(Methods in Enzymology 463) Richard R. Burgess and Murray P. Deutscher (Eds.) - Guide To Protein Purification, 2nd Edition-Academic Press (2009)
(Methods in Enzymology 463) Richard R. Burgess and Murray P. Deutscher (Eds.) - Guide To Protein Purification, 2nd Edition-Academic Press (2009)
METHODS IN
ENZYMOLOGY
Guide to Protein Purification,
2nd Edition
METHODS IN ENZYMOLOGY
Editors-in-Chief
Founding Editors
METHODS IN
ENZYMOLOGY
Guide to Protein Purification,
2nd Edition
EDITED BY
RICHARD R. BURGESS
McArdle Laboratory for Cancer Research
University of Wisconsin-Madison
Madison, Wisconsin, USA
MURRAY P. DEUTSCHER
Department of Biochemistry and Molecular Biology
University of Miami School of Medicine
Miami, FL, USA
No part of this publication may be reproduced, stored in a retrieval system or transmitted in any
form or by any means electronic, mechanical, photocopying, recording or otherwise without the
prior written permission of the publisher
Permissions may be sought directly from Elsevier’s Science & Technology Rights Department
in Oxford, UK: phone (+44) (0) 1865 843830; fax (+44) (0) 1865 853333; email: permissions@
elsevier.com. Alternatively you can submit your request online by visiting the Elsevier web site at
http://elsevier.com/locate/permissions, and selecting Obtaining permission to use Elsevier material
Notice
No responsibility is assumed by the publisher for any injury and/or damage to persons or
property as a matter of products liability, negligence or otherwise, or from any use or operation
of any methods, products, instructions or ideas contained in the material herein. Because of rapid
advances in the medical sciences, in particular, independent verification of diagnoses and drug
dosages should be made
Contributors xix
Preface xxv
Volumes in Series xxvii
v
vi Contents
5. Setting Up a Laboratory 37
Murray P. Deutscher
1. Supporting Materials 38
2. Detection and Assay Requirements 39
3. Fractionation Requirements 40
8. Quantitation of Protein 73
James E. Noble and Marc J. A. Bailey
1. Introduction 74
2. General Instructions for Reagent Preparation 75
3. Ultraviolet Absorption Spectroscopy 80
4. Dye-Based Protein Assays 83
5. Coomassie Blue (Bradford) Protein Assay (Range: 1–50 mg) 85
6. Lowry (Alkaline Copper Reduction Assays) (Range: 5–100 mg) 86
7. Bicinchoninic Acid (BCA) (Range: 0.2–50 mg) 88
8. Amine Derivatization (Range: 0.05–25 mg) 89
Contents vii
1. Principle 373
2. Practice 374
1. Theory 406
2. Latest Technology in HIC Adsorbents 408
3. Procedures for Use of HIC Adsorbents 409
References 413
1. Introduction 516
2. Materials 527
3. Methods 528
References 538
1. Introduction 604
2. Detergent Structure 604
3. Properties of Detergents in Solution 605
4. Exploiting the Physicochemical Parameters of Detergents
for Membrane Protein Purification 612
5. Detergent Removal and Detergent Exchange 613
6. Choosing the Right Detergent 613
7. Conclusions 615
Acknowledgments 616
References 616
1. Introduction 692
2. Chemical Methods 695
3. Transport Methods 698
4. Scattering Methods 716
5. Presence of Subunits 719
References 721
1. Introduction 726
2. Enrichment Techniques for Identifying PTMs 731
3. Nitrosative Protein Modifications 740
4. Methylation and Acetylation 741
5. Mass Spectrometry Analysis 744
6. CID versus ECD versus ETD 747
7. Quantifying PTMs 750
8. Future Directions 756
Acknowledgments 758
References 758
Contents xvii
1. Introduction 809
Alice Alegria-Schaffer
Thermo Fisher Scientific, Pierce Protein Research, Rockford, Illinois, USA
Marc J. A. Bailey
Nokia Research Centre - Eurolab, University of Cambridge, Cambridge, United
Kingdom
John S. Blanchard
Department of Biochemistry, Albert Einstein College of Medicine, Bronx,
New York
Helena Block
QIAGEN GmbH, Qiagen Strasse 1, Hilden, Germany
Robert E. Bossio
Department of Chemistry, University of Michigan–Dearborn, Dearborn,
Michigan, USA
Nicole Brinker
QIAGEN GmbH, Qiagen Strasse 1, Hilden, Germany
Kimberly Brisack
Bio-Rad Laboratories, Inc., Hercules, California, USA
William H. Brondyk
Genzyme Corporation, Framingham, Massachusetts, USA
Richard R. Burgess
McArdle Laboratory for Cancer Research, University of Wisconsin-Madison,
Madison, Wisconsin, USA
Thomas Chappell
Biogrammatics, Inc., Carlsbad, California, USA
Frank R. Collart
Biosciences Division, Argonne National Laboratory, Lemont, Illinois, USA
James M. Cregg
Keck Graduate Institute of Applied Life Sciences, Claremont, California, USA, and
Biogrammatics, Inc., Carlsbad, California, USA
xix
xx Contributors
Larry J. Cummings
Bio-Rad Laboratories, Inc., Hercules, California, USA
Murray P. Deutscher
Department of Biochemistry and Molecular Biology, University of Miami School
of Medicine, Miami, Florida, USA
David R. H. Evans
Process Biochemistry, Biopharmaceutical Development, Biogen Idec. Inc.,
Cambridge, Massachusetts, USA
Roland Fabis
QIAGEN GmbH, Qiagen Strasse 1, Hilden, Germany
Adam R. Farley
Department of Biochemistry, Vanderbilt University School of Medicine,
Nashville, Tennessee, USA
Katherine M. Foley
McArdle Laboratory for Cancer Research, University of Wisconsin-Madison,
Madison, Wisconsin, USA
Brian G. Fox
Department of Biochemistry, Dr. Brian Fox Lab, University of Wisconsin-
Madison, Madison, Wisconsin, USA
Ashley M. Frank
Biosciences Division, Argonne National Laboratory, Lemont, Illinois, USA
David B. Friedman
Proteomics Laboratory, Mass Spectrometry Research Center, Vanderbilt
University, Nashville, Tennessee, USA
Cornelia Fux
Novartis Pharma AG, Department of NBx/PSD: Scale up, Basel, Switzerland
David E. Garfin
Chemical Division, Research Products Group, Bio-Rad Laboratories,
Incorporated, Richmond, California
Sabine Geisse
Novartis Institutes for BioMedical Research, Department of NBC/PPA, Basel,
Switzerland
Michael A. Goren
Department of Biochemistry, Dr. Brian Fox Lab, University of Wisconsin-
Madison, Madison, Wisconsin, USA
Contributors xxi
Anthony C. Grabski
Department of Research and Development, Semba Biosciences, Inc., Madison,
Wisconsin, USA
Reinhard Grisshammer
Department of Health and Human Services, National Institutes of Health, National
Institute of Neurological Disorders and Stroke, Rockville, Maryland, USA
Guido Guidotti
Department of Molecular and Cellular Biology, Harvard University, Cambridge,
Massachusetts, USA
Rainer Hahn
Department of Biotechnology, University of Natural Resources and Applied Life
Sciences, Vienna, Austria
T. K. Harris
Department of Biochemistry and Molecular Biology, Miller School of Medicine,
University of Miami, Miami, Florida, USA
Sjouke Hoving
Novartis Institutes of BioMedical Research, Basel, Switzerland
Donald L. Jarvis
Department of Molecular Biology, University of Wyoming, Laramie, Wyoming,
USA
Alois Jungbauer
Department of Biotechnology, University of Natural Resources and Applied Life
Sciences, Vienna, Austria
M. M. Keshwani
Department of Biochemistry and Molecular Biology, Miller School of Medicine,
University of Miami, Miami, Florida, USA
Patricia J. Kiley
Department of Biomolecular Chemistry, University of Wisconsin, Madison,
Wisconsin, USA
Jan Kubicek
QIAGEN GmbH, Qiagen Strasse 1, Hilden, Germany
Anasua Kusari
Keck Graduate Institute of Applied Life Sciences, Claremont, California, USA
Jörg Labahn
Institute of Structural Biology (IBI-2), Research Center Jülich, Jülich, Germany
xxii Contributors
Thomas M. Laue
Department of Biochemistry and Molecular Biology, University of New
Hampshire, Durham, New Hampshire, USA
Scott A. Lesley
Genomics Institute of the Novartis Research Foundation, San Diego, California,
USA
Sue-Hwa Lin
Department of Molecular Pathology, University of Texas, Houston, Texas, USA
Andrew J. Link
Department of Microbiology and Immunology, Vanderbilt University School of
Medicine, Nashville, Tennessee, USA
Dirk Linke
Department I, Protein Evolution, Max Planck Institute for Developmental
Biology, Tübingen, Germany
Stuart Linn
Department of Molecular and Cellular Biology, University of California, Berkeley,
California, USA
Andrew Lodge
Thermo Fisher Scientific, Pierce Protein Research, Rockford, Illinois, USA
Knut Madden
Biogrammatics, Inc., Carlsbad, California, USA
Barbara Maertens
QIAGEN GmbH, Qiagen Strasse 1, Hilden, Germany
Shin-ichi Makino
Department of Biochemistry, Dr. Brian Fox Lab, University of Wisconsin-
Madison, Madison, Wisconsin, USA
Arun Malhotra
Department of Biochemistry and Molecular Biology, University of Miami Miller
School of Medicine, Miami, Florida, USA
Justin T. McCue
Biogen Idec Corporation, Cambridge, Massachusetts, USA
Uwe Michelsen
Merck KGaA, Darmstadt, Germany
James E. Noble
Analytical Science, National Physical Laboratory, Teddington, Middlesex, United
Kingdom
Contributors xxiii
Akira Nozawa
Department of Biochemistry, Dr. Brian Fox Lab, University of Wisconsin-
Madison, Madison, Wisconsin, USA
David G. Rhodes
Lipophilia Consulting, Storrs, Connecticut, USA
Jonathan K. Romero
Process Biochemistry, Biopharmaceutical Development, Biogen Idec. Inc.,
Cambridge, Massachusetts, USA
Frank Schäfer
QIAGEN GmbH, Qiagen Strasse 1, Hilden, Germany
Dan Simpson
Promega Corporation, Madison, Wisconsin, USA
Mark A. Snyder
Bio-Rad Laboratories, Inc., Hercules, California, USA
Anne Spriestersbach
QIAGEN GmbH, Qiagen Strasse 1, Hilden, Germany
Elizabeth S. Stalder
McArdle Laboratory for Cancer Research, University of Wisconsin-Madison,
Madison, Wisconsin, USA
Thomas H. Steinberg
McArdle Laboratory for Cancer Research, University of Wisconsin–Madison,
Madison, Wisconsin, USA
Earle Stellwagen
Department of Biochemistry, University of Iowa, Iowa City, Iowa
Vincent S. Stoll
Department of Biochemistry, Albert Einstein College of Medicine, Bronx,
New York
Jay Sunga
Keck Graduate Institute of Applied Life Sciences, Claremont, California, USA
Nancy E. Thompson
McArdle Laboratory for Cancer Research, University of Wisconsin-Madison,
Madison, Wisconsin, USA
Ilya Tolstorukov
Keck Graduate Institute of Applied Life Sciences, Claremont, California, USA, and
Biogrammatics, Inc., Carlsbad, California, USA
Marjeta Urh
Promega Corporation, Madison, Wisconsin, USA
xxiv Contributors
Krishna Vattem
Thermo Fisher Scientific, Pierce Protein Research, Rockford, Illinois, USA
Jörg von Hagen
Merck KGaA, Darmstadt, Germany
Reiner Westermeier
Gelcompany GmbH, Paul-Ehrlich-Strasse, Tübingen, Germany
Matthew Westoby
Process Biochemistry, Biopharmaceutical Development, Biogen Idec. Inc., San
Diego, California, USA
Russell L. Wrobel
Department of Biochemistry, Dr. Brian Fox Lab, University of Wisconsin-
Madison, Madison, Wisconsin, USA
Aixin Yan
School of Biological Sciences, The University of Hong Kong, Hong Kong,
Hong Kong SAR
Sarah Zerbs
Biosciences Division, Argonne National Laboratory, Lemont, Illinois, USA
Kate Zhao
Promega Corporation, Madison, Wisconsin, USA
PREFACE
xxv
xxvi Preface
We hope that the Guide will be an invaluable resource and reference text
for researchers new to protein purification as well as for the more experi-
enced researchers, and that it will find an important place in every protein
biochemistry laboratory.
RICHARD R. BURGESS
MURRAY P. DEUTSCHER
METHODS IN ENZYMOLOGY
xxvii
xxviii Methods in Enzymology
VOLUME 257. Small GTPases and Their Regulators (Part C: Proteins Involved
in Transport)
Edited by W. E. BALCH, CHANNING J. DER, AND ALAN HALL
VOLUME 258. Redox-Active Amino Acids in Biology
Edited by JUDITH P. KLINMAN
VOLUME 259. Energetics of Biological Macromolecules
Edited by MICHAEL L. JOHNSON AND GARY K. ACKERS
VOLUME 260. Mitochondrial Biogenesis and Genetics (Part A)
Edited by GIUSEPPE M. ATTARDI AND ANNE CHOMYN
VOLUME 261. Nuclear Magnetic Resonance and Nucleic Acids
Edited by THOMAS L. JAMES
VOLUME 262. DNA Replication
Edited by JUDITH L. CAMPBELL
VOLUME 263. Plasma Lipoproteins (Part C: Quantitation)
Edited by WILLIAM A. BRADLEY, SANDRA H. GIANTURCO, AND JERE P. SEGREST
VOLUME 264. Mitochondrial Biogenesis and Genetics (Part B)
Edited by GIUSEPPE M. ATTARDI AND ANNE CHOMYN
VOLUME 265. Cumulative Subject Index Volumes 228, 230–262
VOLUME 266. Computer Methods for Macromolecular Sequence Analysis
Edited by RUSSELL F. DOOLITTLE
VOLUME 267. Combinatorial Chemistry
Edited by JOHN N. ABELSON
VOLUME 268. Nitric Oxide (Part A: Sources and Detection of NO; NO Synthase)
Edited by LESTER PACKER
VOLUME 269. Nitric Oxide (Part B: Physiological and Pathological Processes)
Edited by LESTER PACKER
VOLUME 270. High Resolution Separation and Analysis of Biological
Macromolecules (Part A: Fundamentals)
Edited by BARRY L. KARGER AND WILLIAM S. HANCOCK
VOLUME 271. High Resolution Separation and Analysis of Biological
Macromolecules (Part B: Applications)
Edited by BARRY L. KARGER AND WILLIAM S. HANCOCK
VOLUME 272. Cytochrome P450 (Part B)
Edited by ERIC F. JOHNSON AND MICHAEL R. WATERMAN
VOLUME 273. RNA Polymerase and Associated Factors (Part A)
Edited by SANKAR ADHYA
VOLUME 274. RNA Polymerase and Associated Factors (Part B)
Edited by SANKAR ADHYA
Methods in Enzymology xliii
It is with great admiration that we reprint in its entirety Chapter 1 of the first
edition of this volume, ‘‘Why Purify Enzymes’’ by the late Arthur Kornberg.
Arthur was one of the giants of 20th century biochemistry and his forte was
enzymology. Upon reading the chapter, one can appreciate why this was so.
His passion for identifying an enzyme activity and purifying it to homo-
geneity so that many of its properties could be directly studied become clearly
evident from Arthur’s words. Yet, he was clearly mindful of what he called
the enzyme’s ‘‘social face,’’ its interactions with other cellular components,
an attribute that has gained more prominence as we have come to understand
the importance of cell organization.
There have been many advances in the 20 years since the chapter was
written, particularly in the ease with which a protein can be purified if its
gene is known. Nevertheless, the admonition to ‘‘purify, purify, purify’’
remains relevant when one is studying the catalytic properties of an enzyme,
elucidating the structure of a protein, or identifying possible regulatory
factors. Protein purification remains of prime importance if one endeavors
to understand the enzyme responsible for a newly discovered cellular
reaction or the proteins involved in a cellular process. With this in mind,
the lessons to be learned from reading, or re-reading, Arthur’s wonderfully
lucid and pertinent chapter will be extremely rewarding.
MURRAY P. DEUTSCHER
1
C H A P T E R O N E
3
4 Arthur Kornberg
studies, the effects of manipulating the cell’s genome and the actions of
viruses and agents are almost always monitored with intact cells and organ-
isms. Rarely are attempts made to examine a stage in an overall process in a
cell-free system. This reliance in current biological research on intact cells
and organisms to fathom their chemistry is a modern version of the vitalism
that befell Pasteur and that has permeated the attitudes of generations of
biologists before and since.
It baffles me that the utterly simple and proven enzymologic approach to
solving basic problems in metabolism is so commonly ignored. The precept
that discrete substances and their interactions must be understood before
more complex phenomena can be explained is rooted in the history of
biochemistry and should by now be utterly commonsensical. Robert Koch,
in identifying the causative agent of an infectious disease, taught us a century
ago that we must first isolate the responsible microbe from all others.
Organic chemists have known even longer that we must purify and crystal-
lize a substance to prove its identity. More recently in history, the vitamin
hunters found it futile to try to discover the metabolic and nutritional roles
of vitamins without having isolated each in pure form. And so with enzymes
it is only by purifying enzymes that we can clearly identify each of the
molecular machines responsible for a discrete metabolic operation. Con-
vinced of this, one of my graduate students expressed it in a personalized
license plate (Fig. 1.1).
ACKNOWLEDGMENT
This article borrows extensively from ‘‘For the Love of Enzymes: The Odyssey of a
Biochemist,’’ Harvard University Press, 1989.
C H A P T E R T W O
Contents
1. General Considerations 10
1.1. Properties and sensitivities of the protein 10
1.2. For what is the protein to be used 10
1.3. Assays 11
1.4. What should be added to the suspension, storage,
and assay buffers 13
1.5. Storage of protein solutions 14
1.6. Contaminating activities 15
2. Source of the Protein 15
2.1. Preliminary studies to obtain sequence information 15
2.2. Overexpressed protein as the source material 16
3. Preparing Extracts 16
4. Bulk or Batch Procedures for Purification 17
5. Refined Procedures for Purification 18
5.1. High-capacity steps 18
5.2. Intermediate-capacity steps 19
5.3. Low-capacity steps 19
6. Conclusions 19
References 19
Abstract
Prior to embarking upon the purification of a protein, one should begin by
considering what the protein is to be used for. In particular, how much of the
protein is needed, what should be its state of purity, and must it be folded
correctly and associated with various other peptides or cofactors. Using such
criteria, an appropriate assay should be chosen and a procedure be planned
taking into account the source of the protein, how it is to be extracted from the
source, and what agents the protein ought to be exposed to or ultimately be
stored in.
Department of Molecular and Cellular Biology, University of California, Berkeley, California, USA
9
10 Stuart Linn
1. General Considerations
1.1. Properties and sensitivities of the protein
If known, one must consider whether the protein is soluble, and, if not,
what agents might help to solubilize it. Is the protein labile at high or low
concentrations? Is the protein sensitive to high or low salt concentrations,
high or low temperatures, high or low pH, or oxidation, particularly by
oxygen. A few preliminary experiments in this regard might be well worth
the effort so as to learn what reagents ought to be present during the
purification, what agents must be avoided (e.g., oxygen), and under what
conditions the protein ought to be stored.
also for detection or assay of the amount and activity of the protein.
Concentration techniques such as those covered in Chapter 9 often must
be included after an intermediate or the final step of purification.
Other considerations include whether the protein must be active (as an
enzyme, a regulatory protein, or an antibody, for example). If so, must it be
folded correctly into its normal native configuration? (Refolding of dena-
tured- or overexpressed proteins may alter precise kinetic properties, for
example, if subtle, but stable alternative configurations are assumed.) Or,
folding may not be an issue, for example, if the protein is to be utilized for
sequence information or identification. The techniques employed during a
procedure should be as gentle as are necessary. But, whenever possible,
some of the harsher, but often spectacularly successful procedures such as
those that involve extremes of pH, organic solvents, detergents, or hydro-
phobic or strong affinity chromatographic media should also be tried.
How free from contaminants should a protein be and, in particular,
which contaminants? For example, proteins used as pharmaceuticals must
be ultra-pure in all respects, whereas polymerases must be free of enzymes
that degrade the polymer product of the polymerase or proteins that add or
remove protein modifications must be free of their counterparts.
A consideration that is currently receiving a great deal of attention is that
of the appropriate subunit content of a multimeric protein. What are the
subunits of the desired protein? Is an isolated complex identical in content
to that within the starting material? And, how might this content change
with the growth state, growth conditions, or storage of the source material?
These considerations are further exacerbated by the possibility that a protein
in question might exist simultaneously in several or many different com-
plexes or states of aggregation. Likewise, it might exist simultaneously with
alternative posttranslational modifications. In either or both of these
instances, the distribution of the protein among these various alternatives
might be extremely sensitive to stresses imposed upon the cells used as the
source material prior to purification. One should take these possibilities
into account, particularly when the protein seems to partition into several
peaks during purification. These variables may preclude an easy or successful
purification scheme. A case in point is p53 (the product of the human TP53
gene) for which a satisfactory purification of active protein has yet to be
achieved, in spite of its importance as the ‘‘guardian of the human genome,’’
because of its numerous patterns of associated peptides, alternative
modifications, and states of aggregation.
1.3. Assays
Classically, a most important consideration is the development of appropri-
ate assays for the protein. The success of the purification is often dependent
on this. Six considerations are relevant: sensitivity, accuracy, precision,
linearity, substrate availability, and cost as measured in time and money.
12 Stuart Linn
Sensitivity can be the limiting factor as the protein becomes diluted into
column effluents, etc. Before formulating a step, dilution and losses ought to
be estimated so that the ability to detect and utilize the protein will be
possible.
Accuracy and precision are often compromised, but clearly these factors
must be controlled to the extent that the assay is reliable for assuring
reproducibility and the recovery of sufficient material.
The linear range of the assay must be measured both with respect to time
and to protein amount. As a procedure progresses, these ranges ought to be
reevaluated, particularly for activity assays. For example, removal of protein
contaminants might affect the limits if an inhibitor or stimulator were
removed. Or, the limits could be altered if the highly purified protein
becomes less stable during assay incubations.
Substrate availability and cost refer to the practicality of the assay:
Is enough substrate available to perform the entire purification without
interruption? Can the assay be performed simultaneously on a reasonable
number of fractions in a reasonable amount of time? Stopping to prepare
more substrate or skimping on material and/or the number of fractions
assayed usually compromises the purification’s success. On the other hand,
assays at some steps might be compromised, for example, by omitting
specificity controls at later stages of purification or assaying alternate or
combined chromatography fractions.
It is often tempting not to use assays for activity, but to purify a gel band
or an antigen instead. Although this approach is appropriate in some
instances, particularly when an activity is not being sought, it is to be
strongly discouraged when activity is in fact what is desired. It cannot be
emphasized strongly enough that an activity assay is necessary to obtain
optimal yields of unaltered activity, be it one associated with an enzyme, an
antibody, a hormone, a regulatory protein, or a protein that modifies the
secondary or tertiary structure of a polymer.
An inherent part of an assay is the assay conditions. Optimal conditions
must be determined empirically, but, beware: optimal conditions often
change with purification due to removal of interfering inhibitory or stimu-
latory factors, lower overall protein concentrations, etc. Moreover, always
take into account materials added with the substrate or with the protein
storage, suspension or dilution buffers as these may change the pH and ionic
strength or otherwise affect the reliability of the assay.
A final comment pertains to the protein assay. While the goals also
include simplicity, linearity, reproducibility, specificity, and reliability,
accuracy is generally compromised, as no commonly used assay is absolute
with regard to all proteins. With crude fractions, color reactions are proba-
bly best. While the Bradford Method (Bradford, 1976) is the simplest of
these, in our laboratory we find it to be unreliable with crude fractions from
animal cells or when detergents are present. Protein assays can and often
General Strategies and Considerations 13
3. Preparing Extracts
Preparation of extracts is discussed in details in Section IV, so only
general considerations are noted here. In our experience, the manner in
which cells are disrupted has a profound, but unpredictable effect on the
General Strategies and Considerations 17
6. Conclusions
Although the development and execution of protein purification pro-
cedures are often difficult and frustrating processes, the rewards are great.
Moreover, given the continual addition of new technology, high-quality
commercial materials utilized for purification procedures, and genetically
altered sources of proteins, the future bodes well for ever-simpler procedures
accompanied by ever-greater rewards.
REFERENCES
Bradford, M. M. (1976). A rapid and sensitive method for quantitation of microgram
quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. 72,
248–254.
Tombs, M. P., Souter, F., and MacLagan, N. F. (1959). The spectrophotometric determi-
nation of protein at 210 mm. Biochem. J. 73, 167–171.
Waddell, W. J. (1956). A simple ultraviolet spectrophotometric method for the determina-
tion of protein. J. Lab. Clin. Med. 48, 311–314.
C H A P T E R T H R E E
Contents
1. What You Can Learn from an Amino Acid Sequence 22
1.1. Molecular weight of the polypeptide chain 22
1.2. Charge versus pH/titration curve, isoelectric point 22
1.3. Molar extinction coefficient 23
1.4. Cysteine content 23
1.5. Secondary structure 24
1.6. Stability 24
1.7. Hydrophobicity and membrane-spanning regions 24
1.8. Sequence similarity suggests homology
and possible cofactor affinity 24
1.9. Potential modification sites 25
1.10. Solubility on overexpression in E. coli 25
2. What You Cannot yet Predict 25
2.1. Three-dimension structure; shape, surface features 26
2.2. Multisubunit features; homomultimers, heteromultimers? 26
2.3. Precipitation properties 26
3. Conclusion 26
3.1. Protein bioinformatic resources 27
3.2. Purification in the denatured state 27
References 27
Abstract
Now that many hundreds and even thousands of whole genomes have been
sequenced, it is rare to be studying a target protein whose amino acid sequence
is not known. However, it is still often necessary to obtain large amounts of the
target protein for a variety of purposes including structural studies, drug
discovery, enzymology, protein biochemistry, and industrial application. It
would seem that knowing the amino acid sequences would make it much easier
McArdle Laboratory for Cancer Research, University of Wisconsin-Madison, Madison, Wisconsin, USA
21
22 Richard R. Burgess
anion and cation exchange columns under the same conditions. To the
first approximation, the greater the charge on a protein at the pH of
the column buffer, the tighter it will bind and the higher the salt needed
to elute it from the column resin. Of course, if the protein is part of a
multiprotein complex, then you have no idea what its ion exchange column
binding properties will be.
Finally, since a protein is generally least soluble at its pI, one can consider
an isoelectric precipitation step (assuming that the protein is not in a stable
complex with another protein or proteins) (see Chapter 20).
1.6. Stability
It is possible, using ProtParam (see below) to estimate the in vivo half-life and
the instability index of a protein from its sequence. The in vivo half-life is
calculated based on the N-end rule (Varshavsky, 1997) and the N-terminal
amino acid of the protein in question. Approximate half-life estimates are
given for the protein expressed in mammalian cells, yeast and E. coli. The
instability index provides an estimate of the stability of your protein in vitro
and is based on the analysis of dipeptide occurrences in your protein com-
pared to those of a set of test proteins that are known to be unstable or stable
(Guruprasad et al., 1990). This information might be useful since a protein
predicted to be unstable in vitro might warrant special care in maintaining low
temperatures during purification and perhaps in the use of protease inhibitors.
your protein is related to other homologous family members that have been
studied, then perhaps you know, for example, that it usually occurs as a
homodimer. This will help in predicting its behavior on gel filtration
column chromatography. This information might also be used to devise a
suitable assay for your protein.
Again, as above, sequence can identify if it belongs to a protein family all
of which are known to bind to a particular cofactor or substrate. For example,
if it is a member of the AAA ATPase family, it is likely that it will bind to an
affinity column that has an immobilized ATP analog. Such an affinity purifi-
cation step can aid greatly in a purification scheme (see Chapter 26).
3. Conclusion
I think it is very clear that one cannot yet use amino acid sequence to
predict the behavior of a given protein in most of the primary fractionation
methods used by protein purifiers. By far the best approach still is to
Bioinformatics in Protein Purification 27
REFERENCES
Gasteiger, E., Hoogland, C., Gattiker, A., Duvaud, S., Wilkins, M. R., Appel, R. D., and
Bairoch, A. (2005). Protein identification and analysis tools on the ExPASy server.
In ‘‘The Proteomics Protocols Handbook’’, ( J. M. Walker, ed.), pp. 571–607. Humana
Press, Totowa.
Gill, S. C., and von Hippel, P. H. (1989). Calculation of protein extinction coefficients from
amino acid sequence data. Anal. Biochem. 182, 319–326.
28 Richard R. Burgess
Guruprasad, K., Reddy, B., and Pandit, M. W. (1990). Correlation between stability of a
protein and its dipeptide composition: A novel approach for predicting in vivo stability of
a protein from it primary sequence. Protein Eng. 4, 155–161.
Idicula-Thomas, S., and Balaji, P. V. (2005). Understanding the relationship between the
primary structure of proteins and its propensity to be soluble on overexpression in E. coli.
Protein Sci. 14, 582–592.
Knuth, M. W., and Burgess, R. R. (1987). Purification of proteins in the denatured state.
In ‘‘Protein Purification: Micro to Macro’’, (R. R. Burgess, ed.), pp. 279–305. Alan R.
Liss, New York.
Kyte, J., and Doolittle, R. F. (1982). A simple method for displaying the hydrophobic
character of a protein. J. Mol. Biol. 157, 105–132.
Pace, C. N., Vajdos, F., Fee, L., Grimsely, G., and Gray, T. (1995). How to measure and
predict the molar absorption coefficient of a protein. Protein Sci. 11, 2411–2423.
Varshavsky, A. (1997). The N-end rule pathway of protein degradation. Genes Cells 2,
13–28.
Wilkinson, D. L., and Harrison, R. G. (1991). Predicting the solubility of recombinant
proteins in E. coli. Biotechnology 9, 443–448.
C H A P T E R F O U R
Preparing a Purification
Summary Table
Richard R. Burgess
Contents
1. Introduction 29
2. The Importance of Footnotes 32
3. The Value of an SDS–Polyacrylamide Gel Analysis on
Main Protein Fractions 32
4. Some Common Mistakes and Problems 32
Abstract
Once a protein purification scheme has been developed, the purification, char-
acterization, and use/structure of a target protein are usually published. It is
highly desirable to present the major steps in the purification and the
corresponding features of the protein at each step summarized in the form of
a purification summary table. In considering whether to repeat a published
protein purification, a reader needs this information to evaluate the purification,
and to decide if it is worth following or if it needs major modifications. In this
chapter, I discuss the main characteristics of a useful purification summary
table and point out common mistakes and problems I see in many such tables.
1. Introduction
As an executive editor and editor-in-chief of the journal, Protein
Expression and Purification, I have reviewed on the order of 100 protein
purification papers a year for over 18 years. It is remarkable how many
manuscripts I receive where there is either no purification summary or one
that is severely lacking in necessary information and accuracy. The essentials
of a reasonable purification table are illustrated by the example below.
Suppose one set out to purify an enzyme from the bacterium E. coli
starting with 10 g of wet weight cell pellet from a 4-l culture (10 g of wet
weight cells typically would contain about 2 g of dry weight and about
McArdle Laboratory for Cancer Research, University of Wisconsin-Madison, Madison, Wisconsin, USA
29
30 Richard R. Burgess
1200 mg of total protein). The cells are lysed by sonication to give a crude
lysate and the debris is removed by centrifugation to give a crude extract.
A 45–50% saturated ammonium sulfate cut was prepared. The 50%
saturated ammonium sulfate pellet was dissolved in buffer and diluted to
low salt and applied to a DEAE anion exchange column. The column was
washed at low salt and then eluted with a linear salt gradient from 0.1 to
0.6 M NaCl, the peak of activity eluting at about 0.25 M NaCl. The peak
was pooled and applied to a Sephacryl S-300 gel filtration column and
eluted with a buffer at constant salt (isocratically). The fractions of peak
activity were pooled and shown by SDS–PAGE with Coomassie blue
staining to be a single band. The specific activity of the final material is
the same as that of a known pure reference sample. The main fractions were
all assayed for enzyme activity and protein determinations were carried out.
The resulting data are given in Table 4.1.
A purification summary table should allow a reader to evaluate easily the
procedure and readily detect particularly effective and ineffective purifica-
tion steps. It should be easy to see if large losses occurred at a particular step.
A suitable table will contain the following columns:
1. Major steps in the purification. These typically include steps like:
Crude lysate (the result of cell or tissue disruption). This step is often
omitted since assays may be difficult but it is useful and even essential
when much of the expressed recombinant target protein is in an
insoluble inclusion body.
Crude extract (the lysate after any insoluble material has been removed
by centrifugation)
clear that it is accurate to one part in 20 (is it 22, 23, or 24?). No numbers
should be given to more than three significant figures and in general
most percentages can be given to two significant figures. Also, remember
that any number resulting from the division of two numbers each
accurate to 10% will only be accurate to 20%.
2. Calculate values erroneously. Remarkably many tables contain simple
arithmetic errors. All numbers should be checked and rechecked before
submission of a manuscript.
3. Use step yields instead of overall yields. A step yield is the yield from a single
step in the purification procedure, that is, the amount of target protein or
activity after that step compared to that in the previous step of the
procedure. A series of four fractionation steps might all give 60% step
yields, but the overall yield is (0.6)4 ¼ 0.6 0.6 0.6 0.6 ¼ 0.13 or
13%. Overall yield is more useful. A procedure that gives an overall yield
of a few percent may be due to a very lengthy and difficult purification of
a rare or unstable protein, but more likely it indicates that the procedure
has not been optimized very well.
4. Calculate yield as yield of total protein. Often I see a table in which the yield
given is yield of total protein, rather than target protein. This is relatively
useless information. Yield is always recovery of total target protein or
activity.
5. Write up and try to publish the first purification that gives any product. There is
a tendency for readers, especially inexperienced readers, to assume that a
published purification is the result of many cycles of improvement and
optimization, and represents the best way to purify the protein. This is
very often not true. Many times, it is just a series of steps, often chosen
arbitrarily that happens to result in some product. One should look very
carefully at a purification to assess if it is a procedure that is worth trying
to follow if one wants to purify some of the target protein. This is why a
proper purification summary table is so valuable. If huge losses occur at
a particular fractionation step, if the overall yield is very low, if the final
purity is not high, or if similar fractionation steps are used several times,
then perhaps the procedure should merely be used as a beginning point
in designing a better, more effective purification.
6. What to do when purifying a protein where a fusion partner is cleaved off during
the procedure? Very often a recombinant protein is expressed as a fusion
with another protein or tag that aids in its folding or purification
(see Chapter 16). Since most often the desired final product is the target
protein without the fusion partner or tag, especially for structural studies,
the fusion partner must be cleaved off by one of several specific proteases.
Let us say that the target is 20 kDa and the fusion partner is 40 kDa, so the
fusion protein expressed is 60 kDa. At the step where the fusion protein
is cleaved, the yield of target protein seems to decrease by 67% even if
all of it is recovered. How is this indicated in the summary table?
34 Richard R. Burgess
I suggest that the column on target protein amount contain two numbers;
the mg of fusion protein and the calculated mg of the final target protein in
parenthesis. That way at the step where the cleavage has occurred, one can
continue giving just the mg of the cleaved product and the theoretical
amount of cleaved product in the first step can be used to calculate the
overall yield.
C H A P T E R F I V E
Setting Up a Laboratory
Murray P. Deutscher
Contents
1. Supporting Materials 38
1.1. Glassware and plasticware 38
1.2. Chemicals 38
1.3. Disposables 38
1.4. Small equipment and accessories 39
1.5. Equipment and apparatus 39
2. Detection and Assay Requirements 39
3. Fractionation Requirements 40
Department of Biochemistry and Molecular Biology, University of Miami School of Medicine, Miami,
Florida, USA
37
38 Murray P. Deutscher
1. Supporting Materials
1.1. Glassware and plasticware
Tubes, beakers, flasks, bottles, cylinders, funnels, and pipets in a wide
range of sizes (disposable materials are often most useful)
Transfer (Pasteur) pipets
Pipettor tips
Baking dishes
Plastic containers
Large carboys and jars
Ice buckets
1.2. Chemicals
High-grade distilled H2O
Salts (generally chlorides or acetates)
Sodium and potassium phosphates
Enzyme-grade ammonium sulfate
Tris and other organic buffers
EDTA
Acids and bases
Reducing agents (2-mercaptoethanol, dithiothreitol, glutathione)
Protease inhibitors
Detergents
Glycerol
1.3. Disposables
Dialysis tubing
Concentrators/fractionators
Weighing paper and boats
Filter paper
pH paper
Aluminum foil
Glass wool
Syringes and needles
Sterile gloves
Marking tape and pens
Paper wipes and towels
Razor blades
Setting Up a Laboratory 39
3. Fractionation Requirements
Protein purification means protein fractionation. What distinguishes a
protein purification laboratory from the usual biochemistry or molecular
biology laboratory is largely the number and types of fractionation apparatus
and materials available. Subsequent chapters will discuss these items in
detail; they will be mentioned here only briefly.
Probably, the most frequently used piece of equipment in the laboratory
is the centrifuge. The workhorse of the protein purification laboratory is the
refrigerated high-speed centrifuge which attains speeds up to 20,000 rpm.
The usefulness of such a centrifuge is directly related to the presence in the
laboratory of a wide variety of rotors and centrifuge tubes and bottles.
Rotors are available that hold as small a volume as a few milliliters per
tube to ones that hold 500-ml bottles. The large rotors are invaluable for
handling the large volumes of extracts often encountered in early steps of a
protein purification.
In instances in which one wants to remove or prepare subcellular
organelles, access to an ultracentrifuge is desirable. Instruments are available
that can process reasonably large volumes at speeds as high as 80,000 rpm,
and smaller machines on the market can go even faster. The availability of
this instrumentation has greatly reduced the time required to prepare
microsomal or high-speed supernatant fractions. In view of the cost of
Setting Up a Laboratory 41
these machines, and their relatively infrequent use in most cases, they are
often shared among laboratories.
The advent of many microanalytical techniques has also made the
minifuge or microcentrifuge an essential item. Though not really a fraction-
ation apparatus in a protein purification laboratory, it is a necessary addition.
In this regard, the larger centrifuges are also frequently used for assays of
various types, rather than only for fractionation purposes.
In order to isolate proteins, a means of rupturing cells is required.
Various apparatuses are available for this purpose, including hand-held and
motor-driven homogenizers, blenders, sonicators, pressure cells, etc. These
will be discussed in detail in other chapters. In general, it is desirable to have
a variety of the less costly items in individual laboratories, with the remain-
der available as shared equipment.
Column chromatography is a primary protein purification method in
use in most laboratories. Every laboratory involved in protein fractionation
should have available a large supply of columns of various lengths and
diameters in anticipation of every conceivable need, since they will arise
during the course of developing purification schemes. Columns are avail-
able in various degrees of sophistication (and cost). We have found that
simple, gravity-flow, open-top columns fitted with stoppers and syringe
needles, or tubing, for fluid inlet and control, are satisfactory for most
chromatographic procedures. However, for many purposes, prepacked
columns and pressurized FPLC systems may be the preferred route.
A dependable fraction collector is one of the most important pieces of
equipment in the laboratory. Failure of a fraction collector may result in the
loss of several month’s work. In this instance, extra money spent on a good,
versatile machine is a wise investment. Instruments able to handle a large
number of tubes, of various sizes, in different collection modes, are the most
useful. Many different types of fraction collectors are available. Careful analysis
of the various models, and matching to anticipated requirements, is good
practice prior to purchase. A suitable strategy for many laboratories would
be the purchase of one of the more sophisticated instruments for special needs,
and one or more of the less costly, simple machines for routine use.
A number of other accessories to column chromatography are useful, if
not essential. These include a peristaltic pump, various sizes of gradient
makers, and a UV monitor. Gradient makers can be homemade from flasks
or bottles, if necessary. Following the protein elution profile during a
chromatographic run provides important information. This can be done
by determining the absorbance of individual fractions with a spectropho-
tometer, or automatically with an in-stream UV monitor. Dual-wavelength
models with different size flow cells are the most versatile (and also most
costly).
Finally, every laboratory should also have on hand a basic supply of
chromatographic gels and resins. These should include an anion and cation
42 Murray P. Deutscher
Contents
1. Introduction 43
2. Theory 44
3. Buffer Selection 45
4. Buffer Preparation 48
5. Volatile Buffers 49
6. Broad-Range Buffers 50
7. Recipes for Buffer Stock Solutions 50
References 56
1. Introduction
The necessity for maintaining a stable pH when studying enzymes is
well established (Good and Izawa, this series; Johnson and Metzler, this
series). Biochemical processes can be severely affected by minute changes in
hydrogen ion concentrations. At the same time, many protons may be
consumed or released during an enzymatic reaction. It has become increas-
ingly important to find buffers to stabilize hydrogen ion concentrations
while not interfering with the function of the enzyme being studied. The
development of a series of N-substituted taurine and glycine buffers by
Good et al. (1966) has provided buffers in the physiologically relevant range
(6.1–10.4) of most enzymes, which have limited side effects with most
enzymes (Good et al., 1966). It has been found that these buffers are
nontoxic to cells at 50 mM concentrations and in some cases much higher
(Ferguson et al., 1980).
43
44 Vincent S. Stoll and John S. Blanchard
2. Theory
The observation that partially neutralized solutions of weak acids or
weak bases are resistant to pH changes on the addition of small amounts of
strong acid or strong base leads to the concept of ‘‘buffering’’ (Perrin and
Dempsey, 1974). Buffers consist of an acid and its conjugate base, such as
carbonate and bicarbonate, or acetate and acetic acid. The quality of a buffer
is dependent on its buffering capacity (resistance to change in pH by
addition of strong acid or base), and its ability to maintain a stable pH
upon dilution or addition of neutral salts. Because of the following equili-
bria, additions of small amounts of strong acid or strong base result in the
removal of only small amounts of the weakly acidic or basic species;
therefore, there is little change in the pH:
HAðacidÞ Ð Hþ þ A ðconjugate baseÞ ð6:1Þ
BðbaseÞ þ Hþ Ð BHþ ðconjugate acidÞ ð6:2Þ
The pH of a solution of a weak acid or base may be calculated from the
Henderson–Hasselbalch equation:
½basic species
pH ¼ pKa0 þ log ð6:3Þ
½acidic species
The pKa of a buffer is that pH where the concentrations of basic and
acidic species are equal, and in this basic form the equation is accurate
between the pH range of 3–11. Below pH 3 and above pH 11 the
concentrations of the ionic species of water must be included in the
equation (Perrin and Dempsey, 1974). Since the pH range of interest here
is generally in the pH 3–11 range, this will be ignored.
From the Henderson–Hasselbalch equation an expression for buffer
capacity may be deduced. If at some concentration of buffer, c, the sum
[A] þ [HA] is constant, then the amount of strong acid or base needed to
cause a small change in pH is given by the relationship:
( )
d½B ½Ka0 c½Hþ K
þ ½Hþ þ þ
w
¼ 2:303 ð6:4Þ
dpH ðKa0 þ ½Hþ Þ2 H
In this equation, Kw refers to the ionic product of water, and the second
and third terms are only significant below pH 3 or above pH 11. In the pH
range of interest (pH 3–11), this equation yields the following expression:
2:303c
bmax ¼ ¼ 0:576c; ð6:5Þ
4
which represents a maximum value for d[B]/dpH when pH ¼ pKa. The
buffer capacity of any buffer is dependent on the concentration, c, and may
Buffers: Principles and Practice 45
0.025
b
0.015
0.005
−1.0 −0.5 0.0 0.5 1.0
ΔpH
Figure 6.1 Buffer capacity (b) versus DpH over the range 1 pH unit of the pKa for
HEPES (0.05 M). Points calculated using Eq. (6.5), and data from Perrin and Dempsey
(1974).
3. Buffer Selection
There are many factors that must be considered when choosing a buffer.
When studying an enzyme one must consider the pH optimum of the enzyme,
nonspecific buffer effects on the enzyme, and interactions with substrates or
metals. When purifying a protein, cost becomes an important consideration, as
does the compatibility of the buffer with different purification techniques.
Table 6.1 lists a wide variety of buffers covering a broad pH range.
Determining the pH optimum of a protein is a first step in determining the
best buffer to employ (Blanchard, this series). Since the buffering capacity is
maximal at the pK, buffers should be used close to this value. When determin-
ing the pH optimum for an enzyme, it is useful to use a series of related buffers
that span a wide pH range. Once an optimal pH has been approximated,
different buffers within this pH range can be examined for specific buffer effects.
The Good buffers have been shown to be relatively free of side effects.
However, inorganic buffers do have a high potential for specific buffer
effects. Many enzymes are inhibited by phosphate buffer, including car-
boxypeptidase, urease, as well as many kinases and dehydrogenases
(Blanchard, this series). Borate buffers can form covalent complexes with
mono- and oligosaccharides, the ribose moieties of nucleic acids, pyridine
nucleotides, and other gem-diols. Tris and other primary amine buffers may
form Schiff base adducts with aldehydes and ketones.
46 Vincent S. Stoll and John S. Blanchard
4. Buffer Preparation
Once a suitable buffer has been chosen it must be dissolved and titrated to
the desired pH. Before titrating a buffer solution, the pH meter must be
calibrated. Calibration should be done using commercially available pH
Buffers: Principles and Practice 49
5. Volatile Buffers
In certain cases, it is necessary to remove a buffer quickly and
completely. Volatile buffers make it possible to remove components that
may interfere in subsequent procedures. Volatile buffers are useful in elec-
trophoresis, ion-exchange chromatography, and digestion of proteins fol-
lowed by separation of peptides or amino acids. Most of the volatile buffers
(Table 6.2) are transparent in the lower UV range except for the buffers
System pH range
87 ml glacial acetic acid þ 25 ml 88% HCOOH in 11 l 1.9
25 ml 88% HCOOH in 1 l 2.1
Pyridine–formic acid 2.3–3.5
Trimethylamine–formic acid 3.0–5.0
Triethylamine–formic (or acetic) acid 3–6
5 ml pyridine þ 100 ml glacial acetic acid in 1 l 3.1
5 ml pyridine þ 50 ml glacial acetic acid in 1 l 3.5
Trimethylamine–acetic acid 4.0–6.0
25 ml pyridine þ 25 ml glacial acetic acid in 1 l 4.7
Collidine–acetic acid 5.5–7.0
100 ml pyridine þ 4 ml glacial acetic acid in 1 l 6.5
Triethanolamine–HCl 6.8–8.8
Ammonia–formic (or acetic) acid 7.0–10.0
Trimethylamine–CO2 7–12
Triethylamine–CO2 7–12
24 g NH4HCO3 in 1 l 7.9
Ammonium carbonate–ammonia 8.0–10.5
Ethanolamine–HCl 8.5–10.5
20 g (NH4)2CO3 in 1 l 8.9
a
From Perrin and Dempsey (1974).
50 Vincent S. Stoll and John S. Blanchard
6. Broad-Range Buffers
There may be occasions where a single buffer system is desired that can
span a wide pH range of perhaps 5 or more pH units. One method would be
a mixture of buffers that sufficiently covers the pH range of interest. This
may lead to nonspecific buffer interactions for which corrections must be
made. Another common approach is to use a series of structurally related
buffers that have evenly spaced pK values such that each pK is separated by
approximately 1 pH unit (the limit of buffering capacity). The Good
buffers are ideal for this approach since they are structurally related and have
relatively evenly spaced pK values. As the pH passes the pK of one buffer it
becomes nonparticipatory and therefore has no further function. These
nonparticipating buffer components may show nonspecific buffer effects
as well as raising the ionic strength with potential deleterious effects.
A detailed description of buffer mixtures which provide a wide range of
buffering capacity with constant ionic strength is available (Ellis and
Morrison, this series).
x pH x pH
5.0 3.6 16.8 2.8
6.4 3.4 24.2 2.6
8.2 3.2 32.4 –
x y pH
46.5 3.5 3.0
43.7 6.3 3.2
40.0 10.0 3.4
37.0 13.0 3.6
35.0 15.0 3.8
33.0 17.0 4.0
31.5 18.5 4.2
28.0 22.0 4.4
25.5 24.5 4.6
23.0 27.0 4.8
20.5 29.5 5.0
18.0 32.0 5.2
16.0 34.0 5.4
13.7 36.3 5.6
11.8 38.2 5.8
9.5 41.5 6.0
7.2 42.8 6.2
x y pH
46.3 3.7 3.6
44.0 6.0 3.8
41.0 9.0 4.0
36.8 13.2 4.2
30.5 19.5 4.4
25.5 24.5 4.6
14.8 35.2 5.0
10.5 39.5 5.2
8.8 41.2 5.4
4.8 45.2 5.6
x y pH
44.6 5.4 2.6
42.2 7.8 2.8
39.8 10.2 3.0
37.7 12.3 3.2
35.9 14.1 3.4
33.9 16.1 3.6
32.3 17.7 3.8
30.7 19.3 4.0
29.4 20.6 4.2
27.8 22.2 4.4
26.7 23.3 4.6
25.2 24.8 4.8
24.3 25.7 5.0
23.3 26.7 5.2
22.2 27.8 5.4
21.0 29.0 5.6
19.7 30.3 5.8
17.9 32.1 6.0
16.9 33.1 6.2
15.4 34.6 6.4
13.6 36.4 6.6
9.1 40.9 6.8
6.5 43.6 7.0
5. Succinate buffer (G. Gomori, unpublished observation) stock solutions
A: 0.2 M solution of succinic acid (23.6 g in 1000 ml)
B: 0.2 M NaOH
25 ml of A þ x ml of B, diluted to a total of 100 ml:
x pH x pH
7.5 3.8 26.7 5.0
10.0 4.0 30.3 5.2
13.3 4.2 34.2 5.4
16.7 4.4 37.5 5.6
20.0 4.6 40.7 5.8
23.5 4.8 43.5 6.0
6. Cacodylate buffer (Plumel, 1949) stock solutions
A: 0.2 M solution of sodium cacodylate (42.8 g of Na(CH3)2AsO2
3H2O in 1000 ml)
B: 0.2 M NaOH
50 ml of A þ x ml of B, diluted to a total of 200 ml:
Buffers: Principles and Practice 53
x pH x pH
2.7 7.4 29.6 6.0
4.2 7.2 34.8 5.8
6.3 7.0 39.2 5.6
9.3 6.8 43.0 5.4
13.3 6.6 45.0 5.2
18.3 6.4 47.0 5.0
13.8 6.2
7. Phosphate buffer (Sorensen, 1909a,b) stock solutions
A: 0.2 M solution of monobasic sodium phosphate (27.8 g in 1000 ml)
B: 0.2 M solution of dibasic sodium phosphate (53.65 g of Na2HPO4
7H2O or 71.7 g of Na2HPO4 12H2O in 1000 ml)
x ml of A þ y ml of B, diluted to a total of 200 ml:
x y pH x y pH
93.5 6.5 5.7 45.0 55.0 6.9
92.0 8.0 5.8 39.0 61.0 7.0
90.0 10.0 5.9 33.0 67.0 7.1
87.7 12.3 6.0 28.0 72.0 7.2
85.0 15.0 6.1 23.0 77.0 7.3
81.5 18.5 6.2 19.0 81.0 7.4
77.5 22.5 6.3 16.0 84.0 7.5
73.5 26.5 6.4 13.0 87.0 7.6
68.5 31.5 6.5 10.5 90.5 7.7
62.5 37.5 6.6 8.5 91.5 7.8
56.5 43.5 6.7 7.0 93.0 7.9
51.0 49.0 6.8 5.3 94.7 8.0
x pH x pH
1.5 9.2 22.5 7.8
2.5 9.0 27.5 7.6
4.0 8.8 32.5 7.4
6.0 8.6 39.0 7.2
9.0 8.4 43.0 7.0
2.7 8.2 45.0 6.8
17.5 8.0
54 Vincent S. Stoll and John S. Blanchard
x pH
5.0 9.0
8.1 8.8
12.2 8.6
16.5 8.4
21.9 8.4
26.8 8.0
32.5 7.8
38.4 7.6
41.4 7.4
44.2 7.2
10. Boric acid–borax buffer (Holmes, 1943) stock solutions
A: 0.2 M solution of boric acid (12.4 g in 1000 ml)
B: 0.05 M solution of borax (19.05 g in 1000 ml; 0.2 M in terms of
sodium borate)
50 ml of A þ x ml of B, diluted to a total of 200 ml:
x pH x pH
2.0 7.6 22.5 8.7
3.1 7.8 30.0 8.8
4.9 8.0 42.5 8.9
7.3 8.2 59.0 9.0
11.5 8.4 83.0 9.1
17.5 8.6 115.0 9.2
11. 2-Amino-2-methyl-1,3-propanediol (Ammediol) buffer (Gomori,
1946) stock solutions
A: 0.2 M solution of 2-amino-2-methyl-1,3-propanediol (21.03 g in
1000 ml)
B: 0.2 M HCl
50 ml of A þ x ml of B, diluted to a total of 200 ml:
Buffers: Principles and Practice 55
x pH x pH
2.0 10.0 22.0 8.8
3.7 9.8 29.5 8.6
5.7 9.6 34.0 8.4
8.5 9.4 37.7 8.2
12.5 9.2 41.0 8.0
16.7 9.0 43.5 7.8
12. Glycine–NaOH buffer (Sorensen, 1909a,b) stock solutions
A: 0.2 M solution of glycine (15.01 g in 1000 ml)
B: 0.2 M NaOH
50 ml of A þ x ml of B, diluted to a total of 200 ml:
x pH x pH
4.0 8.6 22.4 9.6
6.0 8.8 27.2 9.8
8.8 9.0 32.0 10.0
12.0 9.2 38.6 10.4
16.8 9.4 45.5 10.6
13. Borax–NaOH buffer (Clark and Lubs, 1917) stock solutions
A: 0.05 M solution of borax (19.05 g in 1000 ml; 0.02 M in terms of
sodium borate)
B: 0.2 M NaOH
50 ml of A þ x ml of B, diluted to a total of 200 ml:
x pH
0.0 9.28
7.0 9.35
11.0 9.4
17.6 9.5
23.0 9.6
29.0 9.7
34.0 9.8
38.6 9.9
43.0 10.0
46.0 10.1
14. Carbonate–bicarbonate buffer (Delory and King, 1945) stock solutions
A: 0.2 M solution of anhydrous sodium carbonate (21.2 g in
1000 ml)
B: 0.2 M solution of sodium bicarbonate (16.8 g in 1000 ml)
x ml of A þ y ml of B, diluted to a total of 200 ml:
56 Vincent S. Stoll and John S. Blanchard
x y pH
4.0 46.0 9.2
7.5 42.5 9.3
9.5 40.5 9.4
13.0 37.0 9.5
16.0 34.0 9.6
19.5 30.5 9.7
22.0 28.0 9.8
25.0 25.0 9.9
27.5 22.5 10.0
30.0 20.0 10.1
33.0 17.0 10.2
35.5 14.5 10.3
38.5 11.5 10.4
40.5 9.5 10.5
42.5 7.5 10.6
45.0 5.0 10.7
REFERENCES
Blanchard, J. S. (this series). Vol. 104, p. 404.
Bradford, M. M. (1976). Anal. Biochem. 22, 248.
Clark, W. M., and Lubs, H. A. (1917). J. Bacteriol. 2, 1.
Delory, G. E., and King, E. J. (1945). Biochem. J. 39, 245.
Ellis, K. J., and Morrison, J. F. (this series). Vol. 87, p. 405.
Ferguson, W. J., et al. (1980). Anal. Biochem. 104, 300.
Gomori, G. (1946). Proc. Soc. Exp. Biol. Med. 62, 33.
Gomori, G. Unpublished observations.
Good, N. E., and Izawa, S. (this series). Vol. 24, p. 53.
Good, N. E., Winget, G. D., Winter, W., Connolly, T. N., Izawa, S., and Singh, R. M. M.
(1966). Biochemistry 5, 467.
Hayaishi, O. (this series). Vol. 1, p. 144.
Holmes, W. (1943). Anat. Rec. 86, 163.
Johnson, R. J., and Metzler, D. E. (this series). Vol. 22, p. 3.
Leirmo, S., Harrison, C., Cayley, D. S., Burgess, R. R., and Record, M. T. (1987).
Biochemistry 26, 2095.
Lillie, R. D. (1948). Histopathologic Technique. Blakiston, Philadelphia, PA.
McIlvaine, T. C. (1921). J. Biol. Chem. 49, 183.
Michaelis, L. (1930). J. Biol. Chem. 87, 33.
Perrin, D. D., and Dempsey, B. (1974). Buffers for pH and Metal Ion Control.
Chapman & Hall, London.
Plumel, M. (1949). Bull. Soc. Chim. Biol. 30, 129.
Sorensen, S. P. L. (1909a). Biochem. Z. 21, 131.
Sorensen, S. P. L. (1909b). Biochem. Z. 22, 352.
Walpole, G. S. (1914). J. Chem. Soc. 105, 2501.
C H A P T E R S E V E N
Contents
1. Introduction 58
2. Principles of Catalytic Activity 58
2.1. Chemical kinetics 58
2.2. Basic enzyme kinetics 62
3. Measurement of Enzyme Activity 64
3.1. Continuous assays 65
3.2. Discontinuous assays 67
4. Formulation of Reaction Assay Mixtures 69
5. Discussion 71
Acknowledgments 71
References 71
Abstract
To study and understand the nature of living cells, scientists have continually
employed traditional biochemical techniques aimed to fractionate and charac-
terize a designated network of macromolecular components required to carry
out a particular cellular function. At the most rudimentary level, cellular func-
tions ultimately entail rapid chemical transformations that otherwise would not
occur in the physiological environment of the cell. The term enzyme is used to
singularly designate a macromolecular gene product that specifically and
greatly enhances the rate of a chemical transformation. Purification and char-
acterization of individual and collective groups of enzymes has been and will
remain essential toward advancement of the molecular biological sciences; and
developing and utilizing enzyme reaction assays is central to this mission. First,
basic kinetic principles are described for understanding chemical reaction rates
and the catalytic effects of enzymes on such rates. Then, a number of methods
are described for measuring enzyme-catalyzed reaction rates, which mainly
differ with regard to techniques used to detect and quantify concentration
changes of given reactants or products. Finally, short commentary is given
toward formulation of reaction mixtures used to measure enzyme activity.
Whereas a comprehensive treatment of enzymatic reaction assays is not within
Department of Biochemistry and Molecular Biology, Miller School of Medicine, University of Miami,
Miami, Florida, USA
57
58 T. K. Harris and M. M. Keshwani
the scope of this chapter, the very core principles that are presented should
enable new researchers to better understand the logic and utility of any given
enzymatic assay that becomes of interest.
1. Introduction
One primary goal of biological sciences is to deduce the molecular
bases of all chemical processes that take place in living organisms. It is well
established that the vast majority of biochemical reactions are governed by
protein gene products called enzymes. In this sense, enzymes behave as
finely tuned chemical catalysts, which greatly enhance reaction rates with
both temporal and spatial resolution. Thus, elucidation of biochemical
pathways involving biosynthesis, modification, and degradation of macro-
molecular, metabolic, and signaling molecules remains ultimately tied to
experimental methods for purification, reconstitution, and direct demon-
stration of enzymes to catalyze specific chemical reactions. The focus of this
chapter is to explain the most fundamental principles that must be consid-
ered in developing effective methods for measuring the progress of enzyme-
catalyzed reactions. First, kinetic formulation is described for chemical and
enzymatic reaction processes. Next, instructive commentary is given on
various experimental methods that are commonly used for measuring
enzyme activity; and consideration is given toward numerous procedures
and conditions that must be optimized.
A k+1 k+2
R1 + R2 I P1 + P2
k−1
Reaction coordinate
B k+1 k+2
E+S ES E +P
k−1
Figure 7.1 (A) Basic kinetic mechanism for chemical conversion of two molecular
reactants (R1 and R2) into two molecular products (P1 and P2). In this mechanism, an
intermediate complex (I) is reversibly formed from reactants (kþ 1 and k 1) but is
irreversibly converted to products (kþ 2). (B) Basic kinetic mechanism for enzyme
(E)-catalyzed conversion of a molecular substrate (S) into a molecular product (P). In
this mechanism, an ES intermediate complex is reversibly formed (kþ 1 and k 1) but is
irreversibly converted to product (kþ 2). In this case, the enzyme (E) is regenerated and
can undergo subsequent catalytic cycles. For both mechanisms, a free-energy diagram
depicts typical energy changes that occur along the reaction coordinate leading to
intermediates and products. Ground-state energies for reactants, intermediates, and
products are depicted as valleys, whereas transition-state energies for chemical changes
are depicted by peaks.
Table 7.1 Relationship between rate law, order, and the rate constant, k
d½R d½P
v¼ ¼ ¼ k½R ð7:1Þ
dt dt
3[R]0
v (d[P]/dt)
2[R]0
[P]
[R]0
Time [R]0
Figure 7.2 Determination of a first-order rate constant from measuring the initial
velocity of product formation at differing concentrations of a reactant. (A) Initial
velocities of product formation (slopes ¼ v ¼ d[P]/dt, M s 1) are measured for different
starting concentrations of reactant (i.e., [R]0, 2[R]0, or 3[R]0). (B) The measured initial
velocities are then plotted versus the different starting concentrations of reactant. The
slope of this dependence yields the first-order rate constant for conversion of reactant
into product (slope ¼ k ¼ dv/d[R]0, s 1).
ð ½R ðt
d½R
¼ k dt
½R0 ½R 0
½R
ln ¼ kt
½R0
when the reaction is initiated (i.e., [R] ¼ [R]0 at t ¼ 0) and the concentra-
tion of reactant at a given time after the reaction has started. Figure 7.3A
shows the relationship between measuring instantaneous velocity (i.e.,
v ¼ d[R]/dt, where Dt is very small) according to Eq. (7.1) and measuring
overall time progress (i.e., either [R] or [P] versus t, where Dt is very large)
according to Eq. (7.2) for the simple first-order conversion of reactant to
product. The derivative in Eq. (7.1) is simply the slope of the tangent for the
concentration curve at a specific time.
If no product is present at t ¼ 0, the sum of reactant and product
concentrations at any time must be equal to [R]0 (i.e., [R] þ [P] ¼ [R]0).
Using this relationship, ([R]0 [P]) may be substituted for [R] in Eq. (7.2)
and rearranged to yield the first-order rate Eq. (7.3) that describes the
corresponding increasing product concentration with time (Fig. 7.3B).
A B
[R]
[P]
Time Time
Figure 7.3 (A) Time progress for decreasing reactant concentration [R] in a first-order
reaction. It can be seen that the instantaneous velocity at any given reactant concentration
(lines) decreases with decreasing reactant concentration [R]. (B) Corresponding time
progress curve for increasing product concentration [P] in the same first-order reaction.
d½P
v¼ ¼ kcat ½ES ð7:4Þ
dt
d½ES
¼ kþ1 ½E½S k1 ½ES kcat ½ES ð7:5Þ
dt
Measurement of Enzyme Activity 63
ð½E0 ½ESÞ½S0
½ES ¼ ð7:7Þ
Km
½E0 ½S0
½ES ¼ ð7:8Þ
Km þ ½S0
the substrate (i.e., kcat k1 ), the Km value approximates the true dissoci-
ation constant for the ES complex, Kd ¼ k 1/kþ 1; and the system is
described as being in rapid equilibrium.
If the measured initial velocity is then normalized to amount of enzyme
in the reaction mixture, then Eq. (7.9) takes the form of Eq. (7.10), whereby
the activity of the enzyme is defined
v kcat ½S0
Activity ðs1 Þ ¼ ¼ ð7:10Þ
½E0 Km þ ½S0
v ðdA=dt; s1 Þ
Activity ðs1 Þ ¼ ¼ ð7:11Þ
½E0 ðe; M1 cm1 Þðl; cmÞð½E0 ; MÞ
v
Activity ðmmol min1 mg1 Þ ¼
½E0
ð7:12Þ
ðdA=dt; min1 Þ
¼
ðe; ml mmol 1 cm1 Þðl; cmÞð½E0 ; mg ml1 Þ
d½P cpm
Activityðs1 Þ ¼ ¼
dt½E0 ðSA ; cpm=mmol Þðvol; mlÞðt; sÞð½E; mmol=ml1 Þ
S 1
ð7:13Þ
d½P
Activityðmmolmin1 mg1 Þ ¼
dt½E0
ð7:14Þ
cpm
¼
ðSAS ; cpm=mmol1 Þðvol; mlÞðt; minÞð½E; mg=ml1 Þ
5. Discussion
In this chapter, the most fundamental principles of measuring enzyme
reaction rates were presented, and it is in no way a complete treatment on
the subject. Nevertheless, clear understanding of these principles is abso-
lutely prerequisite toward either (i) effective utilization of well-developed
assays or (ii) efficient development of assays for newly discovered enzymes.
Over the past century, countless research articles, review articles, book
chapters, and even entire books have reported on the development, utiliza-
tion, and interpretation of enzyme activity measurements. The selection of
books and articles cited in this article provide more comprehensive treat-
ments of the various topics related to measuring enzyme activity.
ACKNOWLEDGMENTS
This work was supported by NIGMS, National Institutes of Health Grant GM69868 to
T. K. H. and a Maytag Fellowship to M. M. K.
REFERENCES
Cook, P. F., and Cleland, W. W. (2007). Enzyme Kinetics and Mechanism. Garland
Science, Hamden, CT.
Cornish-Bowden, A. (1995). Fundamentals of Enzyme Kinetics. Portland Press, Ltd.,
London, UK.
Eisenthal, R., and Danson, M. (2002). Enzyme Assays: A Practical Approach. Oxford
University Press, Inc., New York, NY.
Frey, P. A., and Hegeman, A. D. (2007). Enzymatic Reaction Mechanisms. Oxford
University Press, Inc., New York, NY.
House, J. E. (2007). Principles of Chemical Kinetics. 2nd edn. Academic Press, Burlington,
MA.
Laidler, K. J. (1987). Chemical Kinetics. 3rd edn. Prentice Hall, Upper Saddle River, NJ.
Lakowicz, J. R. (2004). Principles of Fluorescence Spectroscopy. 2nd edn. Springer
ScienceþBusiness Media, Inc., New York, NY.
McMaster, M. C. (2007). HPLC, a Practical Users Guide. 2nd edn. John Wiley & Sons, Inc.,
Hoboken, NJ.
Rossomando, E. F. (1990). Measurement of enzyme activity. Methods Enzymol. 182, 38–49.
Segel, I. H. (1975). Enzyme Kinetics. John Wiley & Sons, Inc., New York, NY.
Segal, I. H. (1976). Biochemical Calculations. 2nd edn. John Wiley & Sons, Inc.,
New York, NY.
C H A P T E R E I G H T
Quantitation of Protein
James E. Noble* and Marc J. A. Bailey†
Contents
1. Introduction 74
2. General Instructions for Reagent Preparation 75
3. Ultraviolet Absorption Spectroscopy 80
3.1. Ultraviolet absorbance at 280 nm (Range: 20–3000 mg) 80
3.2. Method 81
3.3. Comments 81
3.4. Ultraviolet absorbance at 205 nm (Range: 1–100 mg) 82
3.5. Calculation of the extinction coefficient 82
4. Dye-Based Protein Assays 83
4.1. Protein concentration standards 83
5. Coomassie Blue (Bradford) Protein Assay (Range: 1–50 mg) 85
5.1. Reagents 85
5.2. Procedure 85
5.3. Comments 86
6. Lowry (Alkaline Copper Reduction Assays) (Range: 5–100 mg) 86
6.1. Reagents 87
6.2. Procedure 87
6.3. Comments 88
7. Bicinchoninic Acid (BCA) (Range: 0.2–50 mg) 88
7.1. Reagents 88
7.2. Procedure 89
7.3. Comments 89
8. Amine Derivatization (Range: 0.05–25 mg) 89
8.1. Reagents 90
8.2. Procedure 90
8.3. Comments 90
9. Detergent-Based Fluorescent Detection (Range: 0.02–2 mg) 91
Methods in Enzymology, Volume 463 Crown Copyright # 2009. Published by Elsevier Inc.
ISSN 0076-6879, DOI: 10.1016/S0076-6879(09)63008-1 All rights reserved.
73
74 James E. Noble and Marc J. A. Bailey
Abstract
The measurement of protein concentration in an aqueous sample is an impor-
tant assay in biochemistry research and development labs for applications
ranging from enzymatic studies to providing data for biopharmaceutical lot
release. Spectrophotometric protein quantitation assays are methods that use
UV and visible spectroscopy to rapidly determine the concentration of protein,
relative to a standard, or using an assigned extinction coefficient. Methods are
described to provide information on how to analyze protein concentration using
UV protein spectroscopy measurements, traditional dye-based absorbance
measurements; BCA, Lowry, and Bradford assays and the fluorescent dye-
based assays; amine derivatization and detergent partition assays. The obser-
vation that no single assay dominates the market is due to specific limitations of
certain methods that investigators need to consider before selecting the most
appropriate assay for their sample. Many of the dye-based assays have unique
chemical mechanisms that are prone to interference from chemicals prevalent
in many biological buffer preparations. A discussion of which assays are prone
to interference and the selection of alternative methods is included.
1. Introduction
The quantity of protein is an important metric to measure during
protein purification, for calculating yields or the mass balance, or determin-
ing the specific activity/potency of the target protein. Various platforms and
methods are available to quantitate proteins and will be described elsewhere
in this volume; however, for this chapter, we will concentrate on spectro-
photometric assays of protein in solution that do not require either enzy-
matic/chemical digestion or separation of the mixture prior to analysis.
The spectrophotometric assays described are UV absorbance methods
and dye-binding assays using colorimetric and fluorescent-based detection.
In comparison to other methods, these assays can be run at a high through-
put, using inexpensive reagents with equipment found in the majority of
biochemical laboratories. These spectrophotometric assays require an
appropriate protein standard or constituent amino acid sequence informa-
tion to make a good estimate of concentration. The choice of method used
to determine the concentration of a protein or peptide in solution is
dependent on many factors that will be discussed. The majority of methods
Quantitation of Protein 75
Is the protein free of post-translational No Avoid using the Bradford, and Lowry
modifications, for example
assay for glycosylated proteins
glycosylation?
Yes
Yes
Avoid BCA,
Are thiol, or reducing agents Yes
Is the protein solution free of interfering No Lowry and
present, e.g. DTT from GST
compounds? CBQCATM
affinity purification?
methods
Yes
Check compatibility
Are high concentrations of Yes with BCA, Lowry
salts, or acids used for
and Bradford
precipitation present?
methods
Figure 8.1 Flow chart for the selection of assays for quantitation or proteins in common protein purification procedures. The chart assumes
that the sample for analysis is relatively pure, that is the analyte for quantitation is the major component, for example fractions from affinity
chromatography, or extraction from inclusion bodies. The ‘‘reference standard protein’’ refers to a standard that is the same protein that is
being quantitated in the same, or similar matrix that is ‘‘matched.’’
Table 8.1 Substance compatibility table
Compatible concentrationa
Amine Fluorescent
Substance BCAb Lowryc Bradfordd derivatizatione detergent f UV Abs280 nmg
Acids/bases
HCl 0.1 M na 0.1 M na 10 mM >1 M
NaOH 0.1 M na 0.1 M na 10 mM >1 M
Perchloric acid <1% >1.25% na na na 10%
Trichloroacetic acid <1% >1.25% na na na 10%
Buffers/salts
Ammonium sulfate 1.5 M >28 mM 1.0 M 10 mM 10–50 mM >50%
Borate 10 mM Undiluted Undiluted Undiluted Undiluted Undiluted
Glycine 1 mM 1 mM 100 mM – na 1M
HEPES 100 mM 1 mM 100 mM na 10–50 mM na
Imidazole 50 mM 25 mM 200 mM na na na
Potassium chloride <10 mM 30 mM 1.0 M na 20–200 mM 100 mM
PBS Undiluted Undiluted Undiluted Undiluted Undiluted Undiluted
Sodium acetate 200 mM 200 mM 180 mM na na na
Sodium azide 0.2% 0.5% 0.5% 0.1% 10 mM na
Sodium chloride 1.0 M 1.0 M 5.0 M na 20–200 mM >1 M
Triethanolamine 25 mM 100 mM na na na na
Tris 250 mM 10 mM 2.0 M 10 mM na 0.5 M
Detergents
Brij 35 5% 0.031% 0.125% na na 1%
CHAPS 5% 0.0625% 5% na na 10%
(continued)
Table 8.1 (continued)
Compatible concentrationa
Amine Fluorescent
Substance BCAb Lowryc Bradfordd derivatizatione detergent f UV Abs280 nmg
Deoxycholic acid 5% 625 mg/ml 0.05% na na 0.3%
Nonidet P-40 5% 0.016% 0.5% na na na
SDS 5% 1% 0.125% – 0.01–0.1% 0.1%
Triton X-100 5% 0.031% 0.125% na 0.001% 0.02%
Tween-20 5% 0.062% 0.062% 0.1% 0.001% 0.3%
Reducing agents
Cysteine na 1 mM 10 mM na na na
DTT 1 mM 0.05 mM 5–1000 mM 0.1 mM 10–100 mM 3 mM
2-Mercaptoethanol 0.01% 1 mM 1.0 M 0.1 mM 10–100 mM 10 mM
Thimerosal 0.01% 0.01% 0.01% na na na
Chelators
EDTA 10 mM 1 mM 100 mM na 5–10 mM 30 mM
EGTA na 1 mM 2 mM na na na
Solvents
DMSO 10% 10% 10% na na 20%
Ethanol 10% 10% 10% na na na
Glycerol 10% 10% 10% 10% 10% 40%
Guanidine–HCl 4.0 M 0.1 M 3.5 M na na na
Methanol 10% 10% 10% na na na
PMSF 1 mM 1 mM 1 mM na na na
Sucrose 40% 7.5% 10% 10% 10–500 mM 2M
Urea 3.0 M 3.0 M 3.0 M na na >1.0 M
Miscellaneous
DNA 0.1 mg 0.2 mg 0.25 mg na 50–100 mg/ml 1 mg
Values relate to the maximum concentration of interfering compound within the protein sample that does not result in significant loss in assay performance. The guide is an
updated version of that prepared by Stoscheck to inform of any issues related to assay interference. Concentrations were obtained from product inserts and references
(Bradford, 1976; Peterson, 1979; Smith et al., 1985; Stoscheck, 1990), where there is not a consensus of values a range is given. Changing the protein-to-dye ratios, or
formulation of many of the dye-based assays can alter the maximum concentration of compound permissible. Interfering compounds have been selected to represent those
commonly encountered in protein purification and enzymology.
a
na indicates the reagent was not tested. A blank indicates that the reagent is not compatible with the assay at the reagent concentrations analyzed. A figure preceded by
(<) or (>) symbols indicates the tolerable limit is unknown but is respectively, less than or greater than the amount shown.
b
Figures indicate the concentration in a 0.1-ml sample using a final reaction volume of 2.1 ml.
c
Figures indicate the concentration in a 0.2-ml sample using a final reaction volume of 1.3 ml.
d
Figures indicate the concentration in a 0.05-ml sample using a final reaction volume of 1.55 ml.
e
Figures indicate interference concentrations with the CBQCATM assay (You et al., 1997) in a 90-ml sample using a final reaction volume of 100 ml.
f
Figures indicate interference concentrations with the NanoOrangeTM and Quant-iTTM assays ((Hammer and Nagel, 1986) and Quant-iTTM product insert) in a 40- and
20-ml sample (respectively) using a final reaction volume of 200 ml.
g
Figures indicate the concentration of the chemical that does not produce an absorbance of 0.5 over water (Stoscheck, 1990).
80 James E. Noble and Marc J. A. Bailey
Quant-IT
Fluorescamine
Lowry
CBQCA
Bradford
BCA
Figure 8.2 The markers designate the upper and lower values for the quantitation
range of dye-based protein assay performed in a microplate format. The quantitation
range was defined as the range of protein amounts (ng) that displayed good precision
and did not show any deviation from the fitted response curve. Figure used with
permission from Noble et al. (2007).
A ¼ am cl ð8:1Þ
3.2. Method
For the measurement of a protein with unknown extinction coefficient,
using a protein standard:
1. Add blank buffer to a clean quartz cuvette and use to zero the
spectrophotometer.
2. Either using a fresh identical cuvette or replace the buffer with the
sample, then measure the absorbance at 280 nm. If the signal is outside
the linear range of the instrument (typically an absorbance greater than
2.0), then dilute the protein in buffer and remeasure.
3. After measurement of the sample remeasure the blank buffer to correct
for any instrument drift.
4. Determine the unknown concentrations from the linear standard
response.
3.3. Comments
The determination of the absorbance coefficient for a protein is discussed
below but if a stock of the protein at known concentration is available then
this can be used as a standard. Very rough estimates can be made from the
relationship that if the cuvette has a path length of 1 cm, and the sample volume
is 1 ml then concentration (mg/ml) ¼ absorbance of protein at 280 nm.
Light scattering from either turbid protein samples or particles suspended
in the sample with a comparable size to the incident wavelength (250–
300 nm) can reduce the amount of light reaching the detector leading to an
increase in apparent absorbance. Filtration using 0.2 mm filter units (that do
not adsorb proteins), or centrifugation can be performed prior to analysis to
reduce light scattering. Corrections for light scattering can be performed by
measuring absorbance at lower energies (320, 325, 330, 335, 340, 345, and
350 nm), assuming the protein does not display significant absorbance at
these wavelengths. A log–log plot of absorbance versus wavelength should
82 James E. Noble and Marc J. A. Bailey
generate a linear response that can be extrapolated back to 280 nm, the
resulting antilog of which will give the scattering contribution at this
wavelength (Leach and Scheraga, 1960).
Nucleic acids absorb strongly at 280 nm and are a common contaminant
of protein preparations. A pure protein preparation is estimated to give a
ratio of A280 to A260 of 2.0 while, if nucleic acid is present, the protein
concentration can be derived by the following formula (Groves et al., 1968).
protein sample. The absorbance spectra from various amino acids are
environmentally sensitive; therefore, e derived for a protein in a set buffer
may not be the same for another buffer system if gross changes in pH (tyrosine
ionization at pH 10.9), or solvent polarity (denaturing agents) occur.
To determine e280, the amino acid composition or sequence of the
protein is required. From the protein sequence, e280 can be calculated from
first principles using a standard formula (Gill and von Hippel, 1989), which
has been refined (Pace et al., 1995). Such models use the absorption coeffi-
cients for specific amino acids (Trp, Tyr, and disulfide bond) to generate a
good estimate of e280 where these amino acids are in abundance. However,
where there is a low abundance of these amino acids (e.g., insulin), the model
can display deviations of up to 15% from that determined by physical methods
(Pace et al., 1995). Physical (empirical) methods to determine extinction
coefficient include amino acid analysis (AAA) via acid hydrolysis and chro-
matographic separation of resulting amino acids (Sittampalam et al., 1988)
and Kjeldahl and gravimetric analysis (Kupke and Dorrier, 1978).
3
IgG
Insulin
Insulin-PEG
2.5 Lysozyme
Ratio of protein concentration estimates:
Lyso-PEG
RNase A
2 RNase B
(BSA standard/AAA)
1.5
0.5
0
BCA Bradford CBQCA Lowry Fluorescamine Quant-iT
Figure 8.3 A comparison of the accuracy of the BSA standard in estimating the
concentration of a protein using dye-based protein quantitation assays. AAA was used
to determine the concentration of the model proteins; from these estimates a calibration
curve for each protein was prepared using the dye-based assays. In the same plate, a
calibration curve using the BSA standard (Pierce; concentration defined by manufac-
turer) was also prepared and the response of this was compared to that of the model
proteins to see how well the BSA standard estimated the true concentration of the
model proteins. The ‘‘ratio of concentration estimations’’ refers to the concentration of
protein derived using the BSA standard when compared to the ‘‘true’’ value using
AAA, where a ratio of 1 indicates the two methods gave the same value. The variation
‘‘% CV’’ associated with the dye-based protein concentration assays ranged from 2% to
8%, dependent on the assay. AAA concentration assignment typically displayed 5% CV
values, dependent on the protein analyzed. Figure adapted with permission from Noble
et al. (2007).
Quantitation of Protein 85
5.1. Reagents
Dissolve 100 mg Coomassie Brilliant Blue G-250 in 50 ml of 95% ethanol
and add 100 ml of 85% phosphoric acid while stirring continuously. When
the dye has dissolved dilute to 1 l in water. The reagent is stable for up to a
month at room temperature; however, for long-term storage keep at 4 C,
if precipitation occurs filter before use.
5.2. Procedure
1. Prepare standards in the range 100–1500 mg/ml in a Bradford-compatible
buffer. For more dilute samples the sensitivity can be extended by increas-
ing the ratio of sample to reagent volumes (Micro Bradford assay: 1–25
mg/ml). If the ratio of the sample to dye is too high, the pH of the reaction
mixture could increase leading to higher background responses.
2. Add the standard and unknown samples to disposable cuvettes (plastic
disposable cuvettes and microplates should be used as the dye sticks to
various surfaces).
3. Allow the Bradford reagent to warm to room temperature. Add 1 ml of
the dye solution to 25 ml of the protein sample, mix and incubate for 10
min at room temperature.
86 James E. Noble and Marc J. A. Bailey
5.3. Comments
The advantages of the Bradford assay include the ease of use, sensitivity and
low cost of the reagents. For microplate-based assays the reagent volumes
can be decreased giving a total volume of 300 ml. Due to the path of the
light source on the majority of microplate spectrophotometers, it is recom-
mended to use commercial sources of Bradford reagent that are less predis-
posed to precipitation during prolonged storage.
We have observed significant variation in response between various
commercial suppliers of Bradford preparations (Noble et al., 2007). This
appears to be most pronounced when analyzing low-molecular-weight
proteins or peptides. Indeed the assay is reported to display a molecular
weight cutoff ‘‘threshold’’; requiring a certain number of residues for full
signal development (de Moreno et al., 1986). Changes in the formulation of
the Bradford reagent are reported to change the response generated from
specific proteins; therefore, care should be taken when comparing Bradford
data from different suppliers or preparations (Chan et al., 1995; Friedenauer
and Berlet, 1989; Lopez et al., 1993; Read and Northcote, 1981).
The Bradford assay is sensitive to interferences from various reagents
detailed in Table 8.1 that include most ionic and nonionic detergents and
glycosylated proteins. If precipitation of the reaction mixture occurs, for
example hydrophobic or membrane proteins, the reaction can be supple-
mented with 1 M NaOH at 5–10% (v/v) to aid solubilization.
6.1. Reagents
6.1.1. Folin and Ciocalteu’s reagent
The preparation of this reagent has been described (Lowry et al., 1951);
however, the solution can be obtained from commercial sources (Sigma).
Mix 10 ml of Folin–Ciocalteu’s Phenol reagent to 50 ml of water.
6.2. Procedure
1. To 1 ml of sample and protein standards 5–100 mg/ml, add 1 ml of the
alkaline copper reagent, mix and allow to stand for 10 min.
2. Add 0.5 ml of Folin–Ciocalteu’s reagent mix, vortex thoroughly and
incubate for 30 min.
3. After incubation vortex again and measure the absorbance at 750 nm.
Absorbance can be read from 650 to 750 nm depending on the avail-
ability of appropriate filters (microplate readers), or if the signal is too
high, without significant loss in assay performance. Lowry is not an
endpoint assay, so samples should be staggered to obtain more accurate
estimates.
4. The response observed will be linear over a limited range of standards.
Polynomial, exponential, and logarithmic models can be used to fit the
data to extend the dynamic range of the response curve.
88 James E. Noble and Marc J. A. Bailey
6.3. Comments
The Lowry method above can be adapted to a microplate format by
reducing the volume of reactants added, resulting in a dynamic range
50–500 mg/ml. The Lowry assay has been largely superseded by the
BCA assay due to sensitivity, linearity, and improved methodology.
The Lowry protein assay is sensitive to many interfering compounds
(Table 8.1), which may not generate a linear response (making extrapola-
tions of interfering data complex). Formation of precipitates can occur with
detergents, lipids, potassium ions, and sodium phosphate.
7.1. Reagents
Reagent A: 1 g sodium bicinchoninate, 2 g Na2CO3, 0.16 g sodium tartrate,
0.4 g NaOH, and 0.95 g NaHCO3, made up to 100 ml and the pH
adjusted to 11.25 with either solid or concentrated NaOH.
Reagent B: 0.4 g CuSO4 5H2O dissolved in 10 ml water. Both reagent
A and B are stable indefinitely at room temperature.
The working solution is prepared by mixing 100 parts of reagent A with
two parts reagent B to form a green solution that is stable for up to a week.
Quantitation of Protein 89
7.2. Procedure
1. Cuvette analysis can be performed with 50–150 ml of protein and 3 ml of
BCA working reagent, whereas microplate assay can use 25 ml of protein
and 200 ml of BCA working reagent, that is a lower reagent to protein
ratio.
2. Incubate the sample and standards 5–250 mg/ml at either 37 or 60 C
for 30 min (longer incubations at 37 C will improve protein-to-protein
variability) and allow the sample to equilibrate to room temperature
before reading. Microplates should be covered during incubation to
avoid evaporation of the sample. For cuvette analysis at 37 C, samples
should be staggered to ensure equal incubation times.
3. Measure absorbance at 562 nm, for filter-based plate readers wavelengths
in the range of 540–590 nm can be used instead without a significant loss
in assay performance.
4. The BCA assay will produce a linear response over a wide concentration
range; however, to extend the dynamic range of the data analysis a
quadratic response can be used to model the data.
7.3. Comments
A microbased BCA assay can be used to improve the sensitivity of the
procedure (1–25 mg/ml). The microbased assay uses a more concentrated
working solution and can be prone to precipitation; again commercial
sources of this modified BCA assay are available (Pierce). The BCA assay
is sensitive to either copper chelators (e.g., EDTA) or reagents that can also
reduce Cu2þ (e.g., DTT), a summary of the maximum tolerances can be
found in Table 8.1.
Fluorescamine reacts directly with the amine functional group, whereas OPA
and CBQCATM require the addition of a thiol (2-mercaptoethanol) or cyanide
(CBQCATM). A cuvette-based format is described for the OPA assay, which
can be converted to a microplate format by adjusting the volume of reactants
and NaOH.
8.1. Reagents
OPA stock: Dissolve 120 mg of o-phthalaldehyde (high purity grade from
Sigma or Invitrogen) in methanol, then dilute to 100 ml in 1 M boric acid,
pH 10.4 (pH adjusted with potassium hydroxide). Add 0.6 ml of polyox-
yethylene (23) lauryl ether and mix. The stock is stable for 3 weeks at room
temperature.
8.2. Procedure
1. At least 30 min before analysis, add 15 ml of 2-mercaptoethanol to 5 ml
of OPA stock, this reagent is stable for a day. Protect all fluorescent
samples and reactions from light at all times.
2. Protein standards (0.2–10 mg/ml) and unknown samples need to be
adjusted to a pH between 8.0 and 10.5 before analysis. Mix 10 ml of test
sample with 100 ml of OPA stock (supplemented with 2-mercaptoethanol)
and incubate at room temperature for 15 min.
3. Add 3 ml of 0.5 N NaOH and mix.
4. Read fluorescence at excitation 340 nm and emission from 440 to
455 nm in a fluorescent cuvette.
5. The relationship between protein concentration and fluorescence
should be linear over the dynamic range of the assay and can be used
to estimate unknown samples.
8.3. Comments
All three dyes offer improved sensitivity and dynamic range when compared
with absorbance-based protein quantitation assays. OPA is generally pre-
ferred over fluorescamine due to its enhanced solubility and stability in
aqueous buffers.
The use of amine-derivatization agents for protein quantitation is limited
as the assay displays a large protein-to-protein variability due to variation in
the number of lysine residues in proteins, requiring the need for a
‘‘matched’’ standard. Assay interference from glycine and amine containing
buffers, ammonium ions, and thiols common in many biological-buffering
systems limit the application of such assays (Table 8.1). The reproducibility
of the assay is dependent on the pH of the reaction, protein samples that
Quantitation of Protein 91
contain residual acids, for example from precipitation steps could reduce the
rate of amine derivatization (You et al., 1997).
A noncovalent amine reactive dye epicocconone can also be used for
total protein assays in solution (Sigma), for which the mechanism has been
reported (Bell and Karuso, 2003; Coghlan et al., 2005).
10.1. Cuvettes
Traditionally, cuvettes have been used for the majority of spectrophoto-
metric protein assays. Quartz cuvettes can be costly, therefore glass cuvettes
are preferred; however, both of these may have to be washed between
measurements to remove dye and adsorbed protein. Disposable plastic
92 James E. Noble and Marc J. A. Bailey
cuvettes are available and can be used to increase the throughput where
many samples have to be measured, or the reagent is prone to sticking to the
cuvette surface, for example Bradford reagent. Staggering of sample analysis
is especially important if the signal is not stable, or does not run to comple-
tion within the time frame of the assay, for example BCA or Lowry assays.
The best precision is obtained from a two-beam instrument incorporating a
reference cell to account for instrument drift. Replacing the cuvette in the
holder between each measurement due to cleaning, or the use of disposable
cuvettes can result in changes in alignment, resulting in significant changes
in amount of light reaching the detector. This is especially important if low-
volume cuvettes are being used where the transmission window is reduced
in size. Care should also be taken with low-volume cuvettes to ensure the
sample covers the entire transmission window.
Care should be taken when handling and cleaning cuvettes. Prevent
fingerprints from contaminating the transmitting surfaces. Cuvettes should
be washed with either water or an appropriate solvent between runs and
dried using a stream of nitrogen gas. If smearing of the transmitting surface is
observed, the cuvette can be rewashed in water, ethanol, and finally ace-
tone, or removed using ethanol and lintless lens tissue. If protein deposition
is a recurring problem, cuvettes can be washed overnight in nitric acid and
thoroughly washed before use.
ACKNOWLEDGMENT
We thank A. Hills for comments and help in preparing this review.
REFERENCES
Alterman, M., Chin, D., Harris, R., Hunziker, P., Le, A., Linskens, S., Packman, L., and
Schaller, J. AAARG2003 study: Quantitation of proteins by amino acid analysis and
colorimetric assays. Poster available for download from www.abrf.org (http://www.abrf.
org/ResearchGroups/AminoAcidAnalysis/EPosters/aaa2003_poster_print.pdf ).
Asermely, K. E., Broomfield, C. A., Nowakowski, J., Courtney, B. C., and Adler, M.
(1997). Identification of a recombinant synaptobrevin-thioredoxin fusion protein by
capillary zone electrophoresis using laser-induced fluorescence detection. J. Chromatogr.
B Biomed. Sci. Appl. 695, 67–75.
Bantan-Polak, T., Kassai, M., and Grant, K. B. (2001). A comparison of fluorescamine and
naphthalene-2, 3-dicarboxaldehyde fluorogenic reagents for microplate-based detection
of amino acids. Anal. Biochem. 297, 128–136.
Bell, P. J., and Karuso, P. (2003). Epicocconone, a novel fluorescent compound from the
fungus epicoccumnigrum. J. Am. Chem. Soc. 125, 9304–9305.
Blakeley, R. L., and Zerner, B. (1975). An accurate gravimetric determination of the
concentration of pure proteins and the specific activity of enzymes. Methods Enzymol.
35, 221–226.
Bradford, M. M. (1976). A rapid and sensitive method for the quantitation of microgram
quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. 72,
248–254.
Burkitt, W. I., Pritchard, C., Arsene, C., Henrion, A., Bunk, D., and O’Connor, G. (2008).
Toward Systeme International d’Unite-traceable protein quantification: From amino
acids to proteins. Anal. Biochem. 376, 242–251.
Chan, J. K., Thompson, J. W., and Gill, T. A. (1995). Quantitative determination of
protamines by coomassie blue G assay. Anal. Biochem. 226, 191–193.
Coghlan, D. R., Mackintosh, J. A., and Karuso, P. (2005). Mechanism of reversible
fluorescent staining of protein with epicocconone. Org. Lett. 7, 2401–2404.
Daban, J. R., Bartolome, S., and Samso, M. (1991). Use of the hydrophobic probe Nile red
for the fluorescent staining of protein bands in sodium dodecyl sulfate-polyacrylamide
gels. Anal. Biochem. 199, 169–174.
de Moreno, M. R., Smith, J. F., and Smith, R. V. (1986). Mechanism studies of coomassie
blue and silver staining of proteins. J. Pharm. Sci. 75, 907–911.
Fountoulakis, M., Juranville, J. F., and Manneberg, M. (1992). Comparison of the
Coomassie brilliant blue, bicinchoninic acid and Lowry quantitation assays, using non-
glycosylated and glycosylated proteins. J. Biochem. Biophys. Methods 24, 265–274.
Friedenauer, S., and Berlet, H. H. (1989). Sensitivity and variability of the Bradford protein
assay in the presence of detergents. Anal. Biochem. 178, 263–268.
Gill, S. C., and von Hippel, P. H. (1989). Calculation of protein extinction coefficients from
amino acid sequence data. Anal. Biochem. 182, 319–326.
Groves, W. E., Davis, F. C. Jr., and Sells, B. H. (1968). Spectrophotometric determination
of microgram quantities of protein without nucleic acid interference. Anal. Biochem. 22,
195–210.
Hammer, K. D., and Nagel, K. (1986). An automated fluorescence assay for subnanogram
quantities of protein in the presence of interfering material. Anal. Biochem. 155, 308–314.
Quantitation of Protein 95
Jones, L. J., Haugland, R. P., and Singer, V. L. (2003). Development and characterization of
the NanoOrange protein quantitation assay: A fluorescence-based assay of proteins in
solution. Biotechniques 34, 850–854, 856, 858 passim.
Kupke, D. W., and Dorrier, T. E. (1978). Protein concentration measurements: The dry
weight. Methods Enzymol. 48, 155–162.
Leach, S. J., and Scheraga, H. A. (1960). Effect of light scattering on ultraviolet difference
spectra. J. Am. Chem. Soc. 82, 4790–4792.
Lopez, J. M., Imperial, S., Valderrama, R., and Navarro, S. (1993). An improved Bradford
protein assay for collagen proteins. Clin. Chim. Acta 220, 91–100.
Lorenzen, A., and Kennedy, S. W. (1993). A fluorescence-based protein assay for use with a
microplate reader. Anal. Biochem. 214, 346–348.
Lowry, O. H., Rosebrough, N. J., Farr, A. L., and Randall, R. J. (1951). Protein measure-
ment with the Folin phenol reagent. J. Biol. Chem. 193, 265–275.
Noble, J. E., Knight, A. E., Reason, A. J., Di Matola, A., and Bailey, M. J. (2007). A
comparison of protein quantitation assays for biopharmaceutical applications. Mol. Bio-
technol. 37, 99–111.
Olson, B. J., and Markwell, J. (2007). Assays for determination of protein concentration.
Curr. Protoc. Protein Sci. Chapter 3, Unit 3.4.
Ozols, J. (1990). Amino acid analysis. Methods Enzymol. 182, 587–601.
Pace, C. N., Vajdos, F., Fee, L., Grimsley, G., and Gray, T. (1995). How to measure and
predict the molar absorption coefficient of a protein. Protein Sci. 4, 2411–2423.
Peterson, G. L. (1977). A simplification of the protein assay method of Lowry et al. which is
more generally applicable. Anal. Biochem. 83, 346–356.
Peterson, G. L. (1979). Review of the Folin phenol protein quantitation method of Lowry,
Rosebrough, Farr and Randall. Anal. Biochem. 100, 201–220.
Read, S. M., and Northcote, D. H. (1981). Minimization of variation in the response to
different proteins of the Coomassie blue G dye-binding assay for protein. Anal. Biochem.
116, 53–64.
Sapan, C. V., Lundblad, R. L., and Price, N. C. (1999). Colorimetric protein assay
techniques. Biotechnol. Appl. Biochem. 29(Pt 2), 99–108.
Scopes, R. K. (1974). Measurement of protein by spectrophotometry at 205 nm. Anal.
Biochem. 59, 277–282.
Sittampalam, G. S., Ellis, R. M., Miner, D. J., Rickard, E. C., and Clodfelter, D. K. (1988).
Evaluation of amino acid analysis as reference method to quantitate highly purified
proteins. J. Assoc. Off. Anal. Chem. 71, 833–838.
Smith, P. K., Krohn, R. I., Hermanson, G. T., Mallia, A. K., Gartner, F. H.,
Provenzano, M. D., Fujimoto, E. K., Goeke, N. M., Olson, B. J., and Klenk, D. C.
(1985). Measurement of protein using bicinchoninic acid. Anal. Biochem. 150, 76–85.
Stoscheck, C. M. (1990). Quantitation of protein. Methods Enzymol. 182, 50–68.
Wiechelman, K. J., Braun, R. D., and Fitzpatrick, J. D. (1988). Investigation of the
bicinchoninic acid protein assay: Identification of the groups responsible for color
formation. Anal. Biochem. 175, 231–237.
Wu, A. M., Wu, J. C., and Herp, A. (1978). Polypeptide linkages and resulting structural
features as powerful chromogenic factors in the Lowry phenol reaction. Studies on a
glycoprotein containing no Lowry phenol-reactive amino acids and on its desialylated
and deglycosylated products. Biochem. J. 175, 47–51.
You, W. W., Haugland, R. P., Ryan, D. K., and Haugland, R. P. (1997). 3-(4-Carbox-
ybenzoyl)quinoline-2-carboxaldehyde, a reagent with broad dynamic range for the assay
of proteins and lipoproteins in solution. Anal. Biochem. 244, 277–282.
C H A P T E R N I N E
Concentration of Proteins
and Removal of Solutes
David R. H. Evans,* Jonathan K. Romero,* and Matthew Westoby†
Contents
1. Chromatography 98
1.1. Gel filtration 99
1.2. Ion-exchange chromatography 99
1.3. Reversed phase chromatography 100
1.4. Hydrophobic and affinity chromatography 100
1.5. Manufacturing-scale chromatographic applications 102
2. Electrophoresis 103
3. Dialysis 104
3.1. Concentration and affinity binding applications 107
4. Ultrafiltration 107
4.1. Ultrafiltration membranes 109
4.2. Ultrafiltration devices 110
4.3. Purification applications 112
5. Lyophilization 113
6. Precipitation 116
7. Crystallization 118
References 118
Abstract
The dramatic advances in recombinant DNA and proteomics technology over the
past decade have supported a tremendous growth in biologics applied to
diagnostics, biomarkers, and commercial therapeutic markets. In particular,
antibodies and fusion proteins have now become a main focus for a broad
number of clinical indications, including neurology, oncology, and infectious
diseases with projected increase in novel first-class molecules and biosimilar
entities over the next several years. In line with these advances are the
improved analytical, development, and small-scale preparative methods
* Process Biochemistry, Biopharmaceutical Development, Biogen Idec. Inc., Cambridge, Massachusetts, USA
{
Process Biochemistry, Biopharmaceutical Development, Biogen Idec. Inc., San Diego, California, USA
97
98 David R. H. Evans et al.
1. Chromatography
Over the past two decades, the innovation of chromatographic tech-
niques for protein concentration and desalting has been driven by the
increasing application of proteomics technologies. Proteomics generally
requires an enrichment of proteins from small amounts of complex mixtures
including the selective concentration of specific species. The high sensitivity
of detection and detailed biochemical characterization provided by mass
spectrometry (MS) and antibody-based assays has found utility in diverse
applications including clinical diagnostics, biologic drug characterization,
and basic research into protein structure–function. These applications
require the analysis of drug targets and biomarkers present at low amounts
in biological fluids, or the acquisition of detailed information on the
structure and posttranslational modification of proteins expressed from
heterologous systems. Samples for proteomics analysis are routinely gener-
ated in aqueous solutions that may contain a range of nonvolatile
salts, solvents, dyes, detergents, chaotropes, DNA/RNA, and lipids.
These contaminants can impair the performance of sensitive analytical
techniques, notably electrospray ionization (ESI) matrix-assisted laser
desorption/ionization (MALDI) MS, where they cause a detrimental effect
on resolution and sensitivity by interfering with sample ionization, adduct
formation, and ion-source fouling. As a result, an excess of microscale solid
phase extraction procedures and devices have been developed by multiple
Concentration of Proteins and Removal of Solutes 99
2. Electrophoresis
Concentration of proteins by gel electrophoresis is generally employed
to enhance the sensitivity of a subsequent analytical technique. In one
example, the sensitivity of Western blot analysis was enhanced by applying
up to 250 ml of a protein sample over five consecutive loadings on a 1–2 cm
stacking gel in a mini-gel system (Sheen and Ali-Khan, 2005). The
approach permitted effective sample concentration with vertical band
broadening as the only artifact. A similar concentration of a dilute sample
prior to western blotting has been achieved via a funnel shaped loading well.
Concentration of protein samples by stacking has been achieved by capillary
electrophoresis (CE) (e.g., see Shihabi, 2002). Moreover, a recent report has
described conditions for CE that permit simultaneous concentration and
desalting of proteins. This is a notable advance, as desalting is normally
required prior to CE because ionic salts cause band broadening through
Joule heating and electrodispersion. In addition, a desirable stacking effect
can be achieved when samples for CE are loaded at a lower ionic strength
than the electrolyte, leading to protein enrichment and improved sensitivity
in analytical applications. Desalting of proteins prior to CE can be achieved
by standard RP extractions. However, while solid phase mini-beds that
permit desalting in line with CE have been described, the approach may be
technically challenging and generally the desalting step is performed as a
separate operation. Thus, to overcome these drawbacks a novel technique
termed capillary isoelectric trapping was developed to simultaneously con-
centrate and desalt protein samples during CE (Booker and Yeung, 2008).
The approach utilizes a discontinuous buffer system to establish a stable pH
boundary within the capillary. The stationary boundary traps zwitterionic
proteins at their pI while removing contaminating ionic salts, which
104 David R. H. Evans et al.
3. Dialysis
Dialysis is the size-based separation of molecules in solution by
selective diffusion through a semipermeable membrane. This technique
is considered the most popular method employed for removal of low-
molecular-weight solutes from large protein molecules. In particular, the
nondenaturing desalting technique permits buffer exchange under benign
or physiological conditions with minimum risk of impacting the target
protein’s properties.
The mechanism of dialysis has not changed, but the techniques and tools
for employing dialysis have improved over the past decade. In particular,
optimized techniques have allowed for higher throughput implementation
(reduced processing times), increased flexibility for sample volume proces-
sing (10 ml ! 100 ml), improved membrane morphology for reduced protein
adsorption and loss, and enhanced convenience for sample handling.
Typically, the volume of the target buffer solution (dialysate) will be several
orders of magnitude greater than the protein sample volume. The sample is
transferred to a sealed compartment and exposed to dialysis membrane which
functions as a semipermeable barrier. The membrane typically possesses a
specific pore size, termed the molecular weight cutoff (MWCO) limit, and
permits the free passage of molecules smaller than this in both directions while
retaining large macromolecules such as proteins. The buffer concentration
gradient across the membrane drives the diffusion of solutes from regions of
high concentration to regions of low concentration (see Fig. 9.1). The diffu-
sional transport or flux is described using Fick’s law for diffusion:
@C
J ¼ D
@x
Concentration
Concentration
CA,t1
CP,o CP,o
CP,o
CA,t = final
0 0 0
Volume of sample (VS) is placed in Since protein P can not permeate Since VS << VD solute A concentration will
a dialysis device containing a membrane or dialysis device, drop to near undetectable levels while
membrane pore size that retains concentration profile remains solute B in sample device will attain
protein of interest (P) and device unchanged. Solute A concentration in concentrations levels near CB,o. Replacing
placed in the dialysate (VD) where sample drops as A diffuses through dialysate with new buffer will ensure
VS<<VD membrane an into dialysate while solute B CB,0 = CB,tfinal. Equilibrium is reached when
diffuses from dialysate into sample solute concentrations are equivalent on
each side of membrane, respectively
Figure 9.1 Schematic of the dialysis process. Top circular diagrams provide graphic representation of solute distributions within ‘‘sample’’
and ‘‘dialysate’’ pools throughout dialysis process. The bottom graphs show the respective concentration profiles from each pool.
106 David R. H. Evans et al.
(10 ml ! 100 ml) with negligible product loss. The traditional dialysis
tubing is difficult to prepare before it can be used. Pierce Biotech (Rockford,
IL) offers SnakeSkinÒ dialysis tubing to simplify large sample dialysis. This
technology comes in a pleated regenerated cellulose membrane format (from
3.5 to 10 kDa MWCO) allowing rapid preparation. In the case with small
sample volume dialysis, Pierce offers compact Slide-A-LyzerÒ Mini-dialysis
Units (10–100 ml), and Cassettes (0.1–30 ml) to process limited small biological
samples. These units are disposable cups made of polypropylene and regener-
ated cellulose allowing sample to be easily removed using standard lab pipettes.
The new designs allow for improved recovery of valuable small sample
volumes.
4. Ultrafiltration
Ultrafiltration (UF) exploits a separation mechanism essentially equiv-
alent to that employed for dialysis. Both techniques employ a semiperme-
able barrier or membrane to separate sample (containing desired product or
108 David R. H. Evans et al.
Jv ¼ L½DPTM so ðPW PF Þ
Cwall
Pfiltrate
Cbulk
d
Cfiltrate
PF
Osmotic pressure
driving force
Figure 9.2 Schematic representation of convective and diffusive solute transport for
ultrafiltration processes.
Cross-flow
Dead-end operation (tangential flow)
operation
Feed or retentate
Feed Retentate
Ultrafiltration
membrane
Permeate Permeate
(filtrate) (filtrate)
offers the The AcroPrepTM filter plates (Pall Corporation, Meriden, CT)
in 96- or 384-multiwell Filter Plate format.
5. Stirred-cell device: A flat sheet UF membrane disk is placed on an appro-
priate support plate and sealed in place at the bottom of a cylindrical
housing. Fluid in the stirred device is agitated using a magnetic stirring
bar suspended from the top of the housing with the device designed to
operate with air to provide pressure driving force. Both Millipore and
Sterlitech provide a variety of device sizes with sample volumes ranging
from 3 to 400 ml. These devices are designed to simulate crossflow/
tangential filtration using different ultrafiltration membranes.
Even though the first four devices described above are probably the
easiest to operate and can process very small sample volumes, they are not
designed to provide precise control of concentration or concentration
polarization on the membrane. In contrast, the stirred-cell device supports
improved mass transport and concentration control due to its mixing
capabilities and capability to perform buffer exchange and product concen-
tration simultaneously. The process of buffer exchange using ultrafiltration
is termed ‘‘diafiltration.’’ In diafiltration, the exchange buffer is added to the
retentate or feed during the filtration, with the UF membrane-permeating
buffer species removed from the feed as the excess fluid is filtered through
the membrane. In contrast to dialysis, in which differences in solute con-
centration across the membrane drive buffer exchange, the UF buffer-
exchange process is faster due to convective flow of buffer across the
membrane. This can be beneficial for processing proteins or biologics that
are unstable and that require immediate buffer exchange. Diafiltration can
be performed in one of two modes. In continuous diafiltration, the wash
(exchange) buffer is added continuously throughout the filtration, with the
system often designed so that the total feed volume remains constant during
the process. In discontinuous diafiltration, the feed volume is first reduced
by a predetermined amount using a standard ultrafiltration. The wash buffer
is then added to the feed reservoir and the filtration continued. This process
can be repeated to obtain desired final solute concentration and volume
reduction.
based processes and the approach has been termed high-performance tan-
gential-flow filtration (Sakena and Zydney, 1994; van Reis et al., 1997).
These membranes are designed to separate biological molecules of identical
size but differing in isoelectric points (pI) by exploiting electrostatic inter-
actions between the molecules and the membrane pores. In addition, there
have been significant advances in the area of membrane chromatography (or
affinity membrane chromatography) using membranes with larger pore
distributions (microfiltration membranes) than ultrafiltration membranes.
Membrane adsorbers have become accepted as an alternative to final polish-
ing chromatography steps for removal of trace impurities (Ghosh, 2002).
There are several commercial adsorbers available from Sartorius and Pall
Corporation that serve as anion and cation exchangers which have designed
to overcome the diffusion mass transfer limitations associated with bead-
based chromatography. Typically, the charged ligand is immobilized in the
membrane pores while convective flow delivers the solute molecules in
close proximity to the ligand thus minimizing diffusional limitations. None-
theless, the adoption of this technology has been slow because membrane
chromatography has been limited by a lower binding capacity than that
achieved by conventional chromatography resins even though the high-flux
advantages provided by membrane adsorbers offer the potential for high
productivity.
5. Lyophilization
Concentration of solutes in solution can be achieved thermodynami-
cally by driving solvent into a gaseous head space (evaporation) or
by boiling. Unfortunately, labile solutes such as protein molecules can be
quickly degraded by the heat required to eliminate solvents. Since the
boiling point of a solution is the temperature at which the vapor pressure
of the solution equals the atmospheric pressure (headspace air pressure),
solvents can boil by increasing solution temperature or by lowering the
atmospheric pressure (e.g., applying vacuum). This latter technique is
termed vacuum concentration. For the process of lyophilization the sample
temperature is lowered to the point where the solution freezes and solvents
are removed by sublimation. The freezing and vacuum step can be per-
formed simultaneously when both heat and headspace gases are removed to
concentrate the product of interest. The liquid protein product to be
lyophilized typically contains the active biologic ingredient, a solvent sys-
tem and several bulking and stabilizing agents. The bulking agents are used
to provide support structure for the active molecule while the stabilizing
agent will play a significant role in maintaining the activity of the target
protein in liquid form prior to the lyophilization process and after sample
114 David R. H. Evans et al.
Specialized containers: vials with Vacuum chamber: contains shelves and Heat transfer fluid circulates
stoppers, ampules or syringes that formulation in vials or glass containers to inside shelves to drive heat
contain formulation be lyophilized transfer away from the product
Condenser
Crygenic
refrigerator
Figure 9.4 Schematic of the lyophilization process and key system components.
116 David R. H. Evans et al.
6. Precipitation
Precipitation is a common purification technique utilized to concen-
trate and desalt proteins. Protein precipitation is the process of separating a
protein from a solution as a solid by altering the protein solubility with
addition of a reagent. A detailed discussion of protein precipitation is
presented in Chapter 20. Precipitation is generally inexpensive and scales
easily. In addition, most precipitations can often be performed in crude feed
streams where impurities such as DNA, lipids, and contaminant proteins are
present. There exist several methods or systems by which one can conduct
precipitation. The challenge is to determine which method is most suitable
to accomplish the desired goals. Precipitation may be induced by the
addition of neutral salts, organic solvents, nonionic hydrophilic polymers,
polyvalent metal ions, or by an acid or base to induce isoelectric point
precipitation (Harrison et al., 2003). The most important factors that influ-
ence protein solubility are structure, size, charge, and the solvent (Ladisch,
2007). Once a protein is precipitated, it can be separated from the solution
by subsequent centrifugation or filtration. The precipitate can then be
resolubilized in a smaller volume to concentrate the protein or washed
and resolubilized in a different solution to change the solution conditions.
A number of commercially available filtration plates and tubes are
available in 96-multiwell plate formats (Anachem, UK; Millipore, Bedford,
MA; Pall, Meriden, CT) which can be useful in screening precipitation
conditions with minimal resources. These plates can also enable precipita-
tion techniques to be automated, further reducing resources. Crystallization
screening kits, which contain ready-to-use solutions reflecting potential
precipitation/crystallization conditions, can also be useful for exploring a
Concentration of Proteins and Removal of Solutes 117
7. Crystallization
Crystallization is another purification technique that can be used to
concentrate and desalt proteins similarly to precipitation. Protein crystalli-
zation is a powerful separation technology because it simultaneously con-
centrates, purifies, and stabilizes the product (Etzel, 2006). Crystallization is
a controlled precipitation from an aqueous solution with four main variables
that control crystal morphology and recovery: protein concentration, pre-
cipitant concentration, pH, and temperature. It is similar to precipitation in
that solid particles are formed from solution; however, precipitates have
poorly defined morphology and are characterized by small particle size,
whereas crystals are highly ordered with generally larger particle sizes. High-
throughput screening systems are being developed in multiwell plate and
smaller formats that, combined with robotic liquid handling and novel
analysis methods, are enabling the rapid screening of broad ranges of con-
ditions with nanoliter amounts (Brown et al., 2003). In addition, applica-
tions for microfluidic chips have been developed for biomolecule
crystallization (Teragenics, Watertown, MA). In practice, it is usually
more difficult to crystallize than precipitate a protein and for purposes of
concentrating and desalting a protein, precipitation is generally a preferable
method. In addition, as crystallization processes are scaled up from the lab
bench, mixing dynamics can significantly impact robustness.
REFERENCES
Atha, D. H., and Ingham, K. C. (1981). Mechanism of precipitation of proteins by
polyethylene glycols. J. Biol. Chem. 256, 12108–12117.
Booker, C. J., and Yeung, K. K. C. (2008). In-capillary protein enrichment and removal of
nonbuffering salts using capillary electrophoresis with discontinuous buffers. Anal. Chem.
80, 8598–8604.
Brown, J., Walter, T. S., Carter, L., Abrescia, N. G. A, Aricescu, A. R., Batuwangala, T. D.,
Bird, L. E., Brown, N., Chamberlain, P. P., and Davis, S. (2003). A procedure for setting
Concentration of Proteins and Removal of Solutes 119
Waters, N. J., et al. (2008). Validation of a rapid equilibrium dialysis approach for the
measurement of plasma protein binding. J. Pharm. Sci. 97(10), 4586–4595.
Waziri, S. M., Abu-Sharkh, B. F., and Ali, S. A. (2004). Protein partitioning in aqueous
two-phase systems composed of a pH-responsive copolymer and poly(ethylene glycol).
Biotechnol. Prog. 20, 526–532.
Wessel, D., and Flugge, U. I. (1984). A method for the quantitative recovery of protein in
dilute solution in the presence of detergents and lipids. Anal. Biochem. 138, 141–143.
Wheat, T. E., Rainville, P. R., Gillece-Castro, B. L., Lu, Z., Cravello, L., and Mazzeo, J. R.
(2007). Fast on-line desalting of proteins for determination of structural information
using exact mass spectrometry. Bioprocess. J. 6, 55–59.
Yu, H., Lu, Y., Zhou, Y.-G., Wang, F.-B., He, F.-Y., and Xia, X.-H. (2008). A disposable
microfluidic device for rapid protein concentration and purification via direct-printing.
Lab Chip 8, 1496–1501.
Zeman, L., and Zydney, A. (1996). Microfiltration and Ultrafiltration. Marcel Dekker Inc.
C H A P T E R T E N
Contents
1. Causes of Protein Inactivation 121
2. General Handling Procedures 122
3. Concentration and Solvent Conditions 122
4. Stability Trials and Storage Conditions 123
5. Proteolysis and Protease Inhibitors 124
6. Loss of Activity 125
Abstract
Proteins are fragile molecules that often require great care during purification
to ensure that they remain intact and fully active. Nowadays, many proteins are
also purified in small amounts under denaturing conditions by various gel
electrophoretic techniques, such that inactive proteins are obtained. But even
here, it is usually advantageous to maintain the protein in an intact form. In the
case of enzymes, and other proteins with assayable biological activities, main-
tenance of activity is generally of prime importance, both for following the
protein during purification and for subsequent studies of function. This chapter
will focus on the major points to keep in mind with regard to maintaining the
stability of a protein during purification and storage. Various other chapters
describe in detail stabilization procedures for specific biological systems and
specific classes of proteins.
Department of Biochemistry and Molecular Biology, University of Miami School of Medicine, Miami,
Florida, USA
121
122 Murray P. Deutscher
precautions that could be taken to minimize their effect, will go a long way
toward ensuring a successful purification protocol. If the protein of interest is
lost or inactivated during the course of any procedure, determination of the
reason for this loss will often suggest a simple solution. Thus, if possible,
examine whether the loss of activity is accompanied by loss of the protein or
changes in its structure, or whether the protein remains, but is now inactive.
Distinguishing among these different possibilities might indicate what type
of process is behind the problem and, thus, what an appropriate solution
might be.
proteins with ones that dilute them. For example, elution of proteins from a
column to which they are bound using a batchwise procedure will tend to
concentrate the eluted proteins, whereas gradient elution will tend to dilute
them. Columns to which proteins bind will tend to concentrate, whereas
gel filtration will dilute. By judicious arrangement of purification steps, one
may be able to avoid extensive dilution.
The solution conditions are also extremely important. Although it is not
possible to describe a universal stabilizing solvent applicable to every pro-
tein, the addition of certain components is generally helpful. These include
a buffer, usually around neutrality, to avoid unnecessary pH changes.
Careful attention should be given to the buffer anion since in many cases
Cl may be harmful. EDTA is usually added at about 0.1 mM to chelate
heavy metal ions that could interact with the protein or promote oxidation.
A reducing agent such as 2-mercaptoethanol or dithiothreitol is often
present to counteract oxidative effects, particularly of cysteine residues.
The use of dithiothreitol at about 0.1–1 mM is preferred because it does
not form mixed disulfides with proteins, as 2-mercaptoethanol does. Suffi-
cient reducing agent should be present since it can oxidize relatively rapidly
and lead to protein oxidation. In some cases, salts are also added to maintain
a certain ionic strength, but only if they are compatible with the next
purification or analytical step. Likewise, glycerol at 5–20% often helps to
maintain stability and is compatible with most purification steps at these
concentrations. Occasionally, low levels of a nonionic detergent are added
to prevent aggregation or the adsorption of proteins to surfaces, such
as glassware. Finally, it is good practice to include protease inhibitors,
particularly at early steps.
6. Loss of Activity
The most commonly heard lament during a protein purification is,
‘‘I’ve lost my activity.’’ When this happens, a careful analysis of the situation
is required to determine the cause. Most importantly, one should have a
126 Murray P. Deutscher
careful accounting of enzyme units to evaluate the extent of the activity loss.
For many purification steps, percentage losses of as much as 50% are not
unusual, but of course, these vary with each individual protein. Generally,
purification methods that involve binding of a protein to a matrix, and
which may require conformational changes during binding, have a greater
effect on activity than a procedure such as gel filtration.
If activity is totally lost during a particular purification step, other
possibilities need to be considered. In some cases proteins may bind very
tightly to columns, and require more extreme procedures for elution.
Depending on the type of chromatography (see Sections VI and VII of
this volume), this may require increased ionic strength, use of a chaotropic
salt (e.g., KBr), or inclusion of detergent or ethylene glycol in the elution
buffer.
A second possibility is that more than one component is required for the
activity of the protein, and these components (e.g., another subunit or a
cofactor) are separated during the fractionation step. Thus, either compo-
nent by itself would be inactive or all have to be present to observe activity.
To test for this possibility, all the fractions from the previous step are mixed
back together, and activity measured. In some cases, it may be necessary to
concentrate the mixture back to the original volume in order to observe
activity. If mixing of all the fractions results in the appearance of activity,
one could then test fractions or groups of fractions in a pairwise fashion.
Often it is found that a little activity remains and that the second component
is needed for optimal activity. In this case, one of the required fractions is
already known, and the other fractions can then be tested for their stimulat-
ing activity.
Sometimes activity may be lost between purification steps, such as
during dialysis or concentration, or even during storage. In the former
situations, one should again test for removal of a possible required compo-
nent. The possibility also exists that the protein has stuck to the dialysis
tubing or the concentration membrane. Here, washing the tubing or
membrane with buffer containing some detergent may be helpful. Problems
of stability during storage have been discussed above.
The most frustrating situation is if none of the above possibilities is the
cause of the loss of activity. Under these circumstances the most likely
explanation is actual inactivation of the protein due to denaturation, prote-
olysis, etc. If an independent measure for the protein is available (e.g., a
Western blot), this can be shown directly. If not, an answer to the inactiva-
tion problem may require trial and error experiments to test various condi-
tions. Sometimes, the best solution is simply to avoid that particular
purification step.
Micropurification (purification of very small amounts of protein in the
submilligram range) presents a special set of problems. In particular, one has
to be especially mindful of the possibility of protein loss due to adsorption
Maintaining Protein Stability 127
Contents
1. Introduction 132
2. Escherichia coli 133
2.1. E. coli: Temperature and molecular chaperones 133
2.2. E. coli: Fusion partners 134
2.3. E. coli: Disulfide bond formation 134
2.4. E. coli: Posttranslational modifications 134
3. Pichia pastoris 135
4. Baculovirus/Insect Cells 136
5. Mammalian Cells 138
6. Protein Characteristics 139
6.1. Protein characteristics: E. coli and codon usage 141
6.2. Protein characteristics: Cytoplasmic proteins 141
6.3. Protein characteristics: Secreted proteins 142
6.4. Protein characteristics: Membrane proteins 142
6.5. Protein characteristics: Toxic proteins 142
7. Recombinant Protein Applications 143
8. Conclusion 144
References 144
Abstract
Recombinant proteins are important tools for studying biological processes.
Generating a recombinant protein requires the use of an expression system.
Selection of an appropriate expression system is dependent on the character-
istics and intended application of the recombinant protein and is essential to
produce sufficient quantities of the protein. Over the last 30 years, there have
been considerable advances in the technologies for expressing recombinant
proteins. In this chapter the unique characteristics of four commonly used
expression systems, Escherichia coli, Pichia pastoris, baculovirus/insect cell,
and mammalian cells are described. The E. coli system is a rapid method for
131
132 William H. Brondyk
1. Introduction
Choosing an appropriate method for expressing a recombinant protein
is a critical factor in obtaining the desired yields and quality of a recombinant
protein in a timely fashion. Selecting a wrong expression host can result in
the protein being misfolded or poorly expressed, lacking the necessary
posttranslational modifications or containing inappropriate modifications.
Factors to consider when selecting an expression system include the mass of
the protein and number of disulfide bonds, type of posttranslational mod-
ifications desired on the expressed protein, and the destination of the
expressed protein. The intended application of the purified recombinant
protein is also critical in the decision-making process and the applications can
be categorized into four broad areas: structural studies, in vitro activity assays,
antigens for antibody generation, and in vivo studies. The purpose of this
chapter is to help guide the investigator in the decision-making process for
choosing an appropriate expression system. However, even with the
described guidelines there are many circumstances when it is not obvious a
priori which expression system is the best choice, and the use of multiple
expressions systems must be attempted before an optimal system is identified.
Numerous expression systems are currently being used in academic and
industrial settings. Some of these systems are too new and insufficiently
tested to comment on their utility. In addition, some established systems for
expressing recombinant proteins, such as transgenic animals, are too tech-
nically challenging, time consuming and prohibitively expensive to be a
viable option for the average laboratory. For the purpose of this chapter,
only Escherichia coli, Pichia pastoris, baculovirus/insect cell, and mammalian
expression systems will be considered (for more detailed coverage of these
systems, see Chapters 12–15). These four systems have straightforward
protocols, are readily accessible either from colleagues or from research
product companies (e.g., Invitrogen, EMD-Novagen, Stratagene, and
Promega), and are relatively inexpensive for small-scale production. The
characteristics and available options of these expression systems will be
briefly reviewed with the focus on the differences between the systems.
Strategies will then be presented to help guide the investigator in making
the best choice for an expression system.
Selecting an Expression System 133
2. Escherichia coli
The bacteria E. coli was the first host used to express recombinant
proteins and is still considered to be the workhorse in the field. Using the
E. coli system offers a rapid and simple method for expressing recombinant
proteins due to its short doubling time. Consequently, the assessment of
recombinant gene expression in E. coli can take less than a week. The growth
media for E. coli are inexpensive and there are relatively straightforward
methods to scale-up bioproduction (see Chapter 12).
In E. coli, recombinant proteins are normally either directed to the
cytoplasm or to the periplasm and, to a lesser extent, secreted. Proteins
directed to the cytoplasm are the most efficiently expressed, giving yields of
up to 30% of the biomass ( Jana and Deb, 2005). However, the high
expression of recombinant protein can often lead to the accumulation of
aggregated, insoluble protein that forms inclusions bodies. Inclusion bodies
have been observed not only with eukaryotic proteins but also to a lesser
extent with overexpressed proteins from prokaryotes including E. coli.
The rate of translation and folding in E. coli is almost 10-fold higher than
that observed in eukaryotic cells, and this presumably contributes to the
inclusion body formation of eukaryotic proteins (Andersson et al., 1982;
Goustin and Wilt, 1982). Inclusion bodies can be a significant hindrance in
obtaining soluble, active protein in some situations. However, in some
cases, inclusion bodies are advantageous because they are resistant to prote-
olysis, easy to concentrate by centrifugation, minimally contaminated with
other proteins, and, with some effort, able to be refolded to form active,
soluble proteins (see Chapter 17 for details).
3. Pichia pastoris
Yeast is another traditional, powerful tool for expressing recombinant
proteins and has been used successfully to express a multitude of proteins.
Yeast has many of the advantageous features of E. coli such as a short
doubling time and a readily manipulated genome, but also has the additional
benefits of a eukaryote that includes improved folding and most posttrans-
lational modifications. The first yeast routinely used for recombinant pro-
tein expression was Saccharomyces cerevisiae (Strausberg and Strausberg,
2001). However, in the last 15 years, P. pastoris has become the yeast of
choice because it typically permits higher levels of recombinant protein
expression than does S. cerevisiae (see Chapter 13). P. pastoris is a methyltropic
yeast, and can use methanol as its only carbon source (Cregg et al., 1985). The
growth of P. pastoris in methanol-containing medium results in the dramatic
transcriptional induction of the genes for alcohol oxidase (AOX) and dihy-
droxyacetone synthase (Cregg, 2007). After induction, these proteins com-
prise up to 30% of the P. pastoris biomass. Investigators have exploited this
methanol-dependent gene induction by incorporating the strong, yet tightly
regulated, promoter of the alcohol oxidase I (AOX1) gene into the majority
of vectors for expressing recombinant proteins (Daly and Hearn, 2005). The
P. pastoris expression vectors integrate in the genome whereas by contrast, S.
cerevisiae vectors use the more unstable method of replicating episomally. The
length of time to assess recombinant gene expression with the P. pastoris
method is approximately 3–4 weeks which includes the transformation of
yeast, screening the transformants for integration, and an expression time-
course. An appealing feature of P. pastoris is the extremely high cell densities
achievable under appropriate culture conditions (Cregg, 2007). Using inex-
pensive medium, the P. pastoris culture can reach 120 g/l of dry cell weight
density. An important caveat is that the induction medium requires a low
percentage of methanol. In large-scale cultures, the amount of methanol
becomes a fire hazard requiring a new level of safety conditions.
P. pastoris has been used to obtain both intracellular and secreted recom-
binant proteins. Like other eukaryotes, it efficiently generates disulfide
bonds and has successfully been used to express proteins containing many
disulfide bonds. To facilitate secretion, the recombinant protein must be
engineered to carry a signal sequence. The most commonly used signal
sequence is the pre-pro sequence from S. cerevisiae a-mating factor (Daly
and Hearn, 2005). Because P. pastoris secretes few endogenous proteins,
purification of the recombinant protein from the medium is a relatively
simple task. If proteolysis of the recombinant protein is a concern, expres-
sion can be completed using the pep4 protease-deficient strain of P. pastoris
136 William H. Brondyk
(Gleeson et al., 1998). This strain has reduced vacuole peptidase A activity
which is responsible for activation of carboxypeptidase Y and protease B1.
Yeast has the posttranslational capacity to add glycans at both specific
asparagine residues (N-linked) and serine/threonine residues (O-linked).
These glycan structures are substantially different from the modifications
added by insect and mammalian cells. In P. pastoris the N-linked glycan is a
high mannose type and usually contains 8–17 mannoses, which is quite
different from S. cerevisiae structures that consist of approximately 50–150
mannose residues (Celik et al., 2007; Gemmill and Trimble, 1999). Similar
to insect and mammalian cells, the consensus sequence for N-linked glycans in
yeast is Asn-Xaa-Ser/Thr. Two groups have completed extensive engineering
to create P. pastoris strains that produce complex N-linked glycan structures
comparable to those produced by mammalian cells (Hamilton and Gerngross,
2007). However, only the strains developed by Roland Contreras’ group are
available to investigators and must be licensed through Research Corporation
Technologies. The O-linked structures in P. pastoris have not been studied
comprehensively but are known to be formed by the addition of one to four
mannose residues to serines/threonines (Goto, 2007; Trimble et al., 2004).
Several reports have indicated that expression of certain proteins in P. pastoris
resulted in the addition O-linked glycans not observed when the protein was
expressed endogenously in mammalian cells (Daly and Hearn, 2005).
4. Baculovirus/Insect Cells
Baculovirus-mediated expression in insect cells offers another useful
tool for generating recombinant proteins (see Chapter 14). Baculovirus is a
lytic, large (130 kb), double-stranded DNA virus, and the Autographa
californica virus is the most commonly used baculovirus isolate for recombi-
nant expression. Baculovirus is routinely amplified in insect cell lines
derived from the fall armyworm Spodoptera frugiperda (Sf 9, Sf 21), and
recombinant protein expression is completed either in the aforementioned
lines or in a line derived from the cabbage looper Trichoplusia ni (High-Five)
(Kost et al., 2005). Originally, creating recombinant baculoviruses involved
cotransfecting the gene of interest flanked by baculovirus sequence with
baculovirus DNA into insect cells, and screening for rare homologous
recombination events. Recombinants were identified by screening plaques
with a modified morphology, and often additional rounds of plaque screen-
ing were required to ensure that the recombinant viral preparation was
not contaminated with wild-type virus. This lengthy and laborious process
for generating recombinant viruses has been largely replaced by using
site-specific transposition (Bac-to-Bac or BaculoDirect, Invitrogen) or an
improved homologous recombination method with an engineered
Selecting an Expression System 137
5. Mammalian Cells
Mammalian expression methods have conventionally been considered
to be the least efficient vehicle for expressing recombinant proteins. How-
ever, recent advances have significantly improved the expression levels from
mammalian cell lines (see Chapter 15). For example, stably transfected
Chinese hamster ovary (CHO) cells have been reported to express recom-
binant antibodies up to a level of a few grams per liter (Figueroa et al., 2007;
Wurm, 2004). While many cell lines and expression strategies have been
tested, this chapter will focus on transient transfection in human embryonic
kidney (HEK293) cells and stable transfection with CHO cells.
The HEK293 cell line was derived from human embryonic kidney cells
transformed with adenovirus. HEK293 cells can be transiently transfected
with a high efficiency (>80%) using certain cationic lipids, calcium phos-
phate, or polyethyleneimine as transfection reagents (Durocher et al., 2002;
Jordan et al., 1996). For large-scale transient transfections (>100 ml), cal-
cium phosphate or polyethyleneimine reagents are more cost-effective
options when compared to cationic lipids (Baldi et al., 2007). Transient
transfections have been performed at even the bioreactor level but for most
laboratories this scale is technically challenging (Girard et al., 2002). The
transient transfection method is relatively easy, and the evaluation for a
given recombinant protein can be made in less than 2 weeks.
CHO cells are commonly used for mammalian expression when large
quantities of recombinant protein are needed. For example, most therapeu-
tic antibodies currently on the market are manufactured using this method.
The standard method for stable CHO expression involves transfecting
dihydrofolate reductase (DHFR)-deficient CHO cells with a DHFR selec-
tion cassette along with an expression cassette containing the gene of
interest (Wurm, 2004). Dihydrofolate reductase converts dihydrofolate
into tetrahydrofolate which is required for the de novo synthesis of purines,
certain amino acids, and thymidylic acid. Methotrexate, which binds and
inhibits DHFR, is used as a selection agent and only those cells that have
integrated the DHFR selection cassette will survive. Sequentially increasing
the concentration of methotrexate will result in amplification of the DHFR
gene along with the linked gene of interest. Following at least one round of
selection with the drug methotrexate, the stably transfected pools are sub-
cloned using limiting dilution cloning into multiwell plates. Typically only a
small percentage of the screened subclones will be expressing the recombi-
nant gene at a high level since in the majority of the clones, the expression
cassette has integrated into the heterochromatin region which is transcrip-
tionally inactive. Unfortunately, the entire selection and screening process
takes at least 2–3 months, making this the major drawback of the CHO
Selecting an Expression System 139
6. Protein Characteristics
When choosing an expression system, one can easily survey the
literature to determine if the recombinant protein has previously been
expressed in the past and then assess the success of the published strategy.
140 William H. Brondyk
Expression
systems Advantages Disadvantages
E. coli Rapid expression Limited capacity for
method (days) posttranslational
modifications
Inexpensive Difficult to produce
bioproduction media some proteins in a
and high density soluble, properly
biomass folded state
Simple process scale-up
Well characterized
genetics
P. pastoris Moderately rapid N-linked glycan
expression method structures different
(weeks) from mammalian
forms
Inexpensive Enhanced safety
bioproduction media precautions needed
and high density for large-scale
biomass bioproduction due to
Most posttranslational methanol in induction
modifications and media
high folding capacity
Baculovirus/ Moderately rapid N-linked glycan
insect cell expression method structures different
(weeks) from mammalian
forms
Most posttranslational Low density biomass and
modifications and expensive
high folding capacity bioproduction media
Difficult process scale-up
Mammalian— Moderately rapid Low density biomass and
transient expression method expensive
expression (weeks) bioproduction media
All posttranslational Difficult process scale-up
modifications and
high folding capacity
Mammalian— All posttranslational Lengthy expression
stable modifications and method (months)
expression high folding capacity Low density biomass and
expensive
bioproduction media
Difficult process scale-up
Selecting an Expression System 141
glycoproteins it is not always clear whether the presence of glycans will alter
the immunogenicity of the recombinant antigen (Bhatia and Mukhopadhyay,
1998; Prasad et al., 1995).
Producing recombinant proteins suitable for in vitro activity studies as
well as for in vivo experiments requires appropriate protein folding and
disulfide bond formation. For glycoproteins, the presence of N-linked
glycans along with the glycan structure can have a significant impact on
both applications, and therefore must be considered when selecting a
eukaryotic expression host. N-linked glycosylation has been shown to
positively influence protein structure and increase protein stability (Bhatia
and Mukhopadhyay, 1998). In vitro, the structure of the N-linked glycan on
certain protein ligands has been demonstrated to affect the affinity for
receptor binding and signal transduction. With recombinant immunoglo-
bulins, the conserved N-linked glycan present in the Fc region influences
in vitro effector activity (Presta, 2008). For instance, the presence of fucose
on the N-linked glycan in human IgG1 reduces antibody-dependent cell
cytotoxicity activity (Shinkawa et al., 2003). In vivo, the N-linked glycan
structure on proteins dramatically impacts metabolic clearance and biodis-
tribution. For example, N-linked glycans that are not capped with sialic acid
are cleared by hepatic receptors, including the asialoglycoprotein and man-
nose receptors (Weigel and Yik, 2002). The importance of O-linked
glycosylation has not yet been defined. If the effect of glycosylating the
recombinant protein is unknown then a safer strategy involves choosing an
expression host similar to the source of the recombinant gene.
8. Conclusion
The E. coli, P. pastoris, baculovirus/insect cells and mammalian systems
each have both advantages and disadvantages for expressing recombinant
proteins (Table 11.1). Whether a given system will express a protein at a high
level and generate a quality product is largely protein dependent. A careful
evaluation of the characteristics of the recombinant protein along with the
downstream application must be considered when selecting an expression
method. Unfortunately, there will be circumstances when the expression
choice will not be obvious and several expression hosts must be evaluated.
REFERENCES
Aguiar, R. W., Martins, E. S., Valicente, F. H., Carneiro, N. P., Batista, A. C.,
Melatti, V. M., Monnerat, R. G., and Ribeiro, B. M. (2006). A recombinant truncated
Cry1Ca protein is toxic to lepidopteran insects and forms large cuboidal crystals in insect
cells. Curr. Microbiol. 53, 287–292.
Selecting an Expression System 145
Andersen, C. L., Matthey-Dupraz, A., Missiakas, D., and Raina, S. (1997). A new Escherichia
coli gene, dsbG, encodes a periplasmic protein involved in disulphide bond formation,
required for recycling DsbA/DsbB and DsbC redox proteins. Mol. Microbiol. 26,
121–132.
Andersson, D. I., Boham, K., Isaksson, L. A., and Kurland, C. G. (1982). Translation rates
and misreading characteristics of rpsD mutants in Escherichia coli. Mol. Gen. Genet. 187,
467–472.
Baldi, L., Hacker, D. L., Adam, M., and Wurm, F. M. (2007). Recombinant protein
production by large-scale transient gene expression in mammalian cells: state of the art
and future perspectives. Biotechnol. Lett. 29, 677–684.
Baneyx, F., and Mujacic, M. (2004). Recombinant protein folding and misfolding in
Escherichia coli. Nat. Biotechnol. 22, 1399–1408.
Bardwell, J. C. (1994). Building bridges: Disulphide bond formation in the cell. Mol.
Microbiol. 14, 199–205.
Bessette, P. H., Aslund, F., Beckwith, J., and Georgiou, G. (1999). Efficient folding of
proteins with multiple disulfide bonds in the Escherichia coli cytoplasm. Proc. Natl. Acad.
Sci. USA 96, 13703–13708.
Bhatia, P. K., and Mukhopadhyay, A. (1998). Protein glycosylation: Implications for in vivo
functions and therapeutic applications. Adv. Biochem. Eng. Biotechnol. 64, 155–201.
Browne, S. M., and Al-Rubeai, M. (2007). Selection methods for high-producing mamma-
lian cell lines. Trends Biotechnol. 25, 425–432.
Butler, M. (2006). Optimisation of the cellular metabolism of glycosylation for recombinant
proteins produced by mammalian cell systems. Cytotechnology 50, 57–76.
Celik, E., Calik, P., Halloran, S. M., and Oliver, S. G. (2007). Production of recombinant
human erythropoietin from Pichia pastoris and its structural analysis. J. Appl. Microbiol.
103, 2084–2094.
Chen, W., Shen, Q., and Bahl, O. P. (1991). Carbohydrate variant of the recombinant
b-subunit of human choriogonadotropin expressed in baculovirus expression system.
J. Biol. Chem. 266, 4081–4087.
Cregg, J. M. (2007). Introduction: Distinctions between Pichia pastoris and other expression
systems. Methods Mol. Biol. 389, 1–10.
Cregg, J. M., Barringer, K. J., Hessler, A. Y., and Madden, K. R. (1985). Pichia pastoris as a
host system for transformation. Mol. Cell. Biol. 5, 3376–3385.
Daly, R., and Hearn, M. T. W. (2005). Expression of heterologous proteins in Pichia pastoris:
A useful experimental tool in protein engineering and production. J. Mol. Recognit. 18,
119–138.
Dong, M. S., Bell, L. C., Guo, Z., Phillips, D. R., Blair, I. A., and Guengerich, F. P. (1996).
Identification of retained N-formylmethionine in bacterial recombinant mammalian
cytochrome P450 proteins with the N-terminal sequence MALLLAVFL. . .: Roles of
residues 3–5 in retention and membrane topology. Biochemistry 35, 10031–10041.
Durocher, Y., Perret, S., and Kamen, A. (2002). High-level and high-throughput recombi-
nant protein production by transient transfection of suspension-growing human
293-EBNA1 cells. Nucleic Acids Res. 30, E9.
Dyson, M. R., Shadbolt, S. P., Vincent, K. J., Perera, R. L., and McCafferty, J. (2004).
Production of soluble mammalian proteins in Escherichia coli: Identification of protein
features that correlate with successful expression. BMC Biotechnol. 4, 32–49.
Esposito, D., and Chatterjee, D. K. (2006). Enhancement of soluble protein expression
through the use of fusion tags. Curr. Opin. Biotechnol. 17, 353–358.
Figueroa, B., Ailor, E., Osborne, D., Hardwick, J. M., Reff, M., and Betenbaugh, M. J.
(2007). Enhanced cell culture performance using inducible anti-apoptotic genes E1B–
19 K and Aven in the production of a monoclonal antibody with Chinese Hamster Ovary
cells. Biotechnol. Bioeng. 97, 877–892.
146 William H. Brondyk
Prasad, S. V., Mujtaba, S., Lee, V. H., and Dunbar, B. S. (1995). Immunogenicity enhance-
ment of recombinant rabbit 55-kilodalton zona pellucida protein expressed using the
baculovirus expression system. Biol. Reprod. 52, 1167–1178.
Presta, L. G. (2008). Molecular engineering and design of therapeutic antibodies. Curr. Opin.
Immunol. 20, 460–470.
Rossi, F. M., and Blau, H. M. (1998). Recent advances in inducible gene expression systems.
Curr. Opin. Biotechnol. 9, 451–456.
Sahdev, S., Khattar, S. K., and Saini, K. S. (2008). Production of active eukaryotic proteins
through bacterial expression systems: A review of existing biotechnology strategies. Mol.
Cell Biochem. 307, 249–264.
Saida, F. (2007). Overview on the expression of toxic gene products in Escherichia coli. Curr.
Protoc. Protein Sci. 5.19.1–5.19.13.
Sarramegna, V., Talmont, F., Demange, P., and Milon, A. (2003). Heterologous expression
of G-protein-coupled receptors: Comparison of expression systems from the standpoint
of large-scale production and purification. Cell. Mol. Life Sci. 60, 1529–1546.
Shinkawa, T., Nakamura, K., Yamane, N., Shoji-Hosaka, E., Kanda, Y., Sakurada, M.,
Uchida, K., Anazawa, H., Satoh, M., Yamasaki, M., Hanai, N., and Shitara, K. (2003).
The absence of fucose but not the presence of galactose or bisecting N-acetylglucosamine
of human IgG1 complex-type oligosaccharides shows the critical role of enhancing
antibody-dependent cellular cytotoxicity. J. Biol. Chem. 278, 3466–3473.
Spiess, C., Beil, A., and Ehrmann, M. (1999). A temperature-dependent switch from
chaperone to protease in a widely conserved heat shock protein. Cell 97, 339–347.
Strausberg, R. L., and Strausberg, S. L. (2001). Overview of protein expression in
Saccharomyces cerevisiae. Curr. Protoc. Protein Sci. Chapter 5, Unit 5.6.
Sugyiama, K., Ahorn, H., Maurer-Fogy, I., and Voss, T. (1993). Expression of human
interferon-a2 in Sf 9 cells: Characterization of O-linked glycosylation and protein het-
erogeneities. Eur. J. Biochem. 217, 921–927.
Thomsen, D. R., Post, L. E., and Elhammer, A. P. (2004). Structure of O-glycosidically
linked oligosaccharides synthesized by the insect cell line Sf 9. J. Cell. Biochem. 43, 67–79.
Trimble, R. B., Lubowski, C., Hauer, C. R., Stack, R., McNaughton, L., Gemmill, T. R.,
and Kumar, S. A. (2004). Characterization of N- and O-linked glycosylation of recom-
binant human bile salt-stimulated lipase secreted by Pichia pastoris. Glycobiology 14,
265–274.
Vera, A., Gonzalez-Montalban, N., Aris, A., and Villaverde, A. (2006). The conformational
quality of insoluble recombinant proteins is enhanced at low growth temperatures.
Biotechnol. Bioeng. 96, 1101–1106.
Waugh, D. S. (2005). Making the most of affinity tags. Trends Biotechnol. 23, 316–320.
Weigel, P. H., and Yik, J. H. N. (2002). Glycans as endocytosis signals: The cases of the
asialoglycoprotein and hyaluronan/chondroitin sulfate receptors. Biochim. Biophys. Acta
1372, 341–363.
Wurm, F. M. (2004). Production of recombinant protein therapeutics in cultivated mam-
malian cells. Nature Biotechnol. 22, 1393–1398.
Young, J. C., Agashe, V. R., Siegers, K., and Hartl, P. U. (2004). Pathways of chaperone-
mediated protein folding in the cytosol. Nat. Rev. Mol. Cell Biol. 5, 781–791.
C H A P T E R T W E LV E
Contents
1.Introduction 150
2.Heterologous Protein Production Using Escherichia coli 150
3.Planning a Bacterial Expression Project 151
4.Evaluation of Project Requirements 152
5.Target Analysis 152
6.Cloning 153
7.Preparation of T4 DNA Polymerase-Treated DNA Fragments 155
8.Expression in the E. coli Cytoplasm 156
9.Expression of Cytoplasmic Targets in E. coli 157
9.1. Analysis of protein expression and solubility 157
10. Analysis of Heterologous Protein Expression in E. coli 157
10.1. Analysis of protein solubility 158
10.2. Analysis of protein expression results 159
10.3. Autoinduction method for protein expression 159
11. Small-Scale Expression Cultures in Autoinduction Media Protocol 160
12. Periplasmic Expression of Proteins 160
13. Expression of Periplasmic Targets in E. coli 161
14. Small-Scale Osmotic Shock Protocol 162
15. Alternative Bacterial Systems for Heterologous Protein Production 164
16. Alternative Vector and Induction Conditions 165
17. Production Scale 166
Acknowledgment 166
References 166
Abstract
Proteins are the working molecules of all biological systems and participate in a
majority of cellular chemical reactions and biological processes. Knowledge of the
properties and function of these molecules is central to an understanding of
chemical and biological processes. In this context, purified proteins are a starting
point for biophysical and biochemical characterization methods that can assist in
149
150 Sarah Zerbs et al.
the elucidation of function. The challenge for production of proteins at the scale and
quality required for experimental, therapeutic and commercial applications has led
to the development of a diverse set of methods for heterologous protein produc-
tion. Bacterial expression systems are commonly used for protein production as
these systems provide an economical route for protein production and require
minimal technical expertise to establish a laboratory protein production system.
1. Introduction
Bacterial expression systems and Escherichia coli in particular are frequently
used for production of heterologous proteins at laboratory and industrial
process scales (Baneyx, 1999; Terpe, 2006). The widespread use of bacterial
systems for protein production arises primarily from the nominal cost and
minimal technical requirements for implementation in a laboratory scale
environment. A variety of vector and host options are available and the short
doubling time of most engineered strains enables rapid evaluation of experi-
mental outcome and reduces the stringency of requirements for sterile tech-
nique and facilities. Many of these intrinsic advantages at the laboratory scale
can be easily modified to accommodate automation of the process and imple-
mentation in a high throughput setting (Dieckman et al., 2002). In most cases,
the bacterial expression systems commonly used for benchtop scale processing
can also be adapted to large-scale projects that require production and screen-
ing of hundreds of expression clones or large-scale production of selected
proteins (Baneyx, 1999; Klock et al., 2005; Terpe, 2006; Yokoyama, 2003).
In spite of these advantages, bacterial systems have a number of important
limitations for expression of heterologous proteins that should be considered
in the development of a protein production expression strategy. These limita-
tions are especially apparent for eukaryotic proteins (Dyson et al., 2004) or
proteins that require the coexpression of maturation proteins (Londer et al.,
2008). This is not unexpected since the biological and chemical characteristics
of the bacterial cellular compartments differ from eukaryotic organisms. In
particular, accessory proteins such as chaperones, posttranslational modifica-
tion proteins, or maturation proteins vary widely between eukaryotic and
bacterial organisms. In some cases, these limitations can be circumvented by
the use of genetically modified bacterial expression strains but the selection of
an alternative expression system must often be considered.
and the intended use of the product. E. coli is the most commonly used
bacterial system for production of heterologous proteins. Over a century of
intensive study of E. coli has provided a great deal of information about
regulatory mechanisms and the function of the host accessory proteins that
may impact expression outcome. In addition, there is an extensive resource
of methodological and technological materials in support of protein pro-
duction in a laboratory or commercial setting. For many protein production
projects, the availability of these resources and the minimal technical
requirements make E. coli a host of first choice for preliminary protein
expression screening. Consequently, we use this organism as a model to
illustrate specific approaches and methods for protein production. The
description of specific methods for protein production using E. coli will be
followed by a discussion of alternative bacterial expression systems. Many of
the core concepts and techniques described for protein production using
E. coli are directly applicable to other bacterial systems.
Table 12.1 Scheme for heterologous protein production using a bacterial host
5. Target Analysis
The biochemical and biological attributes of the target are essential
considerations for selection of an expression strategy and are primary pre-
dictors of expression and solubility outcomes. One set of attributes can be
assembled by analysis of the primary sequence for prediction of secondary
structure, biological localization (membrane, cytoplasmic, periplasmic, or
extracellular), classification into families (fold and domains) and inference
of biochemical properties (pI, disordered regions, ligands). Features such as
a high probability for membrane spanning helices may dictate the use of a
system designed for the expression of membrane proteins or the use of
a domain cloning strategy for production of the soluble component of the
protein. Use of the target sequence features to guide the selection of the
expression host and vector construct will contribute to production of a
mature and appropriately localized protein. Sequence information can be
supplemented with experimental or historical data on the target protein or a
homolog to provide further insight into expression system optimization.
For example, proteins known to require a prosthetic group for proper
function may necessitate selection of a specific host to ensure production
of a fully functional protein. A case in point is the production of type c
cytochromes which is facilitated by the use of genetically engineered strains
of E. coli containing accessory proteins for covalent attachment of hemes to
Protein Expression in Bacterial Systems 153
6. Cloning
A variety of options are available for cloning the target sequence
which can be generally categorized into serial and parallel systems. Serial
systems, such as those that utilize restriction enzymes to generate compatible
target and vector termini, are ubiquitous and provide several options to
generate the target/vector compatible ends required to generate an expres-
sion construct. A disadvantage of this approach is the requirement of
validating restriction enzyme cleavage strategies for each target and vector.
There is growing interest in the use of parallel systems or universal cloning
methods that enable easy transfer of the target to multiple vectors and
expression systems regardless of the target sequence (Table 12.2). Examples
of these systems include the Gateway (Esposito et al., 2009), Infusion (Zhu
et al., 2007) or ligation independent cloning (LIC) methods (Aslanidis and
de Jong, 1990; Haun et al., 1992). An advantage of this approach is the
ability to clone a target in multiple vectors and to simultaneously evaluate
multiple expression strategies in a cost effective manner.
LIC is a cost effective method particularly suited for bacterial expression
as the cloning reagents are nonproprietary and available from several com-
mercial suppliers. In the LIC method, specific nucleotide sequences are
appended to the PCR primers and allow any gene to be cloned regardless of
DNA sequence. Compatible ends in both the vector and PCR fragments
are generated by treatment with T4 DNA polymerase in the presence of a
specific nucleoside triphosphate. The procedure generates complementary
10–15 bp overhangs in the vector and PCR fragment that anneal with
sufficient strength to permit transformation without ligation. The process
allows consistent design of PCR primers, is directional, and results in high
cloning efficiency. Although commercial vectors are available, the approach
has been used in several large-scale cloning projects which provide vector
resources to individual investigators. The procedure described below was
developed for use with the suite of vectors designed at the Midwest Center
for Structural Genomics (Eschenfeldt et al., 2009). The method is generally
applicable to other LIC compatible systems provided adjustments are made
for the specific sequences appended to the amplification primers.
154 Sarah Zerbs et al.
3. Keep this mixture on ice and add the T4 DNA polymerase just before
use. Pipette up and down several times to uniformly distribute the
enzyme in the reaction mix.
4. Array 10.4 mL of the LIC reaction mix into a polypropylene 96-well
plate.
5. Add 30 mL (40–100 ng) of purified PCR fragment to the LIC reaction
mix and pipette up and down several times to mix. Incubate at room
temperature for 30 min. Our studies of various fragment-to-vector ratios
(Dieckman et al., 2006) indicate a wide tolerance for variation in the
amount of target DNA fragment on the annealing reaction.
6. Incubate on a heat block at 75 C for 20 min to inactivate the T4 DNA
polymerase.
7. In another 96-well plate, anneal 1–2 mL of the T4-Polymerase-treated
PCR LIC fragments with 4 mL (20–50 ng) T4-Polymerase-treated LIC
vector.
8. Incubate the annealing reaction 5–10 min at room temperature.
9. Use the entire annealing reaction to transform 50 mL competent E. coli
cells and select for transformed colonies (Sambrook and Russell, 2001).
Following the heating process the LIC plates are stored in the refrigera-
tor at 4 C until needed. The preparation of LIC compatible vector is
similar to the above procedure but uses the complementary base dGTP
(Eschenfeldt et al., 2009). The constructs can be validated by sequence
analysis at this stage or the analysis can be performed after analysis for
expression and solubility.
M A B C A B C A B C A B C A B C
Expression analysis
Solubility analysis
Figure 12.1 Expression and solubility for a set of zebrafish proteins. The top figure
represents a Coomassie-stained gel displaying total expression products for proteins
expressed in three different vector systems. The bottom figure is a Coomassie-stained
gel with the soluble fractions of the same constructs. In the second gel, the order of
ZF356 and the molecular weight marker is reversed from the top gel. Vector ‘‘A’’ is the
pMCSG7 vector and produces an N-terminal fusion containing a TEV protease cleav-
able His tag. Vector ‘‘B’’ targets the protein to the periplasm and produces a N-terminal
fusion similar to that described for Vector A. Vector ‘‘C’’ is pMCSG19 which contains a
maltose binding protein (MBP) fusion sequence (Donnelly et al., 2006). Protein acces-
sion numbers are as follows: ZF356; AAH56726.1, ZF384; AAH58296.1, ZF203;
AAH46038.1, ZF254; AAH47843.1, ZF137; AAH67155.1.
3. Remove samples from freezer and thaw slightly. Add 180 mL of lysis buffer
to each sample, cover tightly and vortex to completely resuspend pellet.
4. Return plate to 80 C for 5 min, then remove and incubate at room
temperature until the ice is completely melted. Repeat this freeze–thaw
cycle once more. Samples may turn clear or they may remain cloudy.
5. Pellet cell debris at 3500 rpm for 10–15 min.
6. Remove a 50 mL portion from the top of each sample being careful not
to remove any of the cell debris with the supernatant.
7. Add 60 mL of 2 SDS loading dye to the supernatant and boil for 5 min.
Analyze the samples by denaturing gel electrophoresis. For a 17-well
8 10 cm gel stained with Coomassie, 5 mL of sample is usually sufficient.
8. [Optional] To examine the insoluble fraction of the cell lysate, discard
all of the remaining supernatant from the cell debris in step 7. Be careful
not to disturb or remove the pellet.
9. Add 300 mL of 1 SDS loading dye to the pellet and cover tightly.
Vortex sample until entire pellet is resuspended.
10. Boil sample for 5 min and cool slightly before loading on an SDS–
PAGE gel. For a 17-well 8 10 cm gel stained with Coomassie, 3 mL is
usually enough to visualize expression.
M 1 2 3 4 5 6 7 8 9 10 11 12
Autoinduction
IPTG induction
Figure 12.2 Coomassie-stained gel analysis of target solubility outcome for autoin-
duced and IPTG-induced cultures. Targets were amplified from Shewanella oneidensis
genomic DNA and cloned into the cytoplasmic expression vector pMCSG7 (targets
1–6) or the periplasmic expression vector pBH31 SA (targets 7–12). Soluble fractions
for analysis were prepared as described in the text. Targets are as follows: 1 ¼ SO 3070;
2 ¼ SO 2454; 3 ¼ SO 2444; 4 ¼ SO 1503; 5 ¼ SO 1190; 6 ¼ SO1560; 7 ¼ SO 0809;
8 ¼ SO 0837; 9 ¼ SO 4048; 10 ¼ SO 1503; 11 ¼ SO 1190; 12 ¼ SO 1560.
protein. As the periplasm accounts for 20–40% of the total volume of the
cell, overall expression yields are typically lower when compared to cyto-
plasmic expression. This is apparent from a direct comparison of the expres-
sion levels for targets cloned into a cytoplasmic expression vector (Fig. 12.2,
targets 4–6) and targets cloned into a periplasmic expression vector
(Fig. 12.2, targets 10–12).
SO 1190 SO 0809
H2O
H2O
SET
SET
WC
WC
SP
SP
M
Figure 12.3 An example of osmotic shock fractions for targets from Shewanella
oneidensis cloned into the periplasmic expression vector pBH31 SA. Fractions are
labeled as follows: WC ¼ whole cell expression; SET ¼ SET buffer fraction; H2O ¼
water shock fraction; SP ¼ spheroplast fraction; M ¼ molecular weight markers.
11. [Optional] To examine the spheroplast fractions after the shock proce-
dure remove all remaining liquid from the cell pellet. Add 500 mL
2 SDS loading dye directly to the pellet and resuspend. Boil sample
for 5 min and cool slightly before loading on the gel. For a 17-well,
8 10 cm gel stained with Coomassie, 3 mL was usually sufficient. This
sample can be extremely viscous and may require additional SDS
loading buffer.
ACKNOWLEDGMENT
The submitted manuscript has been created by UChicago Argonne, LLC, Operator of
Argonne National Laboratory (‘‘Argonne’’). Argonne, a U.S. Department of Energy Office
of Science laboratory, is operated under Contract No. DE-AC02-06CH11357. The U.S.
Government retains for itself, and others acting on its behalf, a paid-up nonexclusive,
irrevocable worldwide license in said article to reproduce, prepare derivative works, distrib-
ute copies to the public, and perform publicly and display publicly, by or on behalf of the
Government.
REFERENCES
Ames, G. F., Prody, C., and Kustu, S. (1984). Simple, rapid, and quantitative release of
periplasmic proteins by chloroform. J. Bacteriol. 160, 1181–1183.
Aslanidis, C., and de Jong, P. J. (1990). Ligation-independent cloning of PCR products
(LIC-PCR). Nucleic Acids Res. 18, 6069–6074.
Baneyx, F. (1999). Recombinant protein expression in Escherichia coli. Curr. Opin. Bio-
technol. 10, 411–421.
Baneyx, F., and Palumbo, J. L. (2003). Improving heterologous protein folding via molecu-
lar chaperone and foldase co-expression. Methods Mol. Biol. 205, 171–197.
Birdsell, D. C., and Cota-Robles, E. H. (1967). Production and ultrastructure of lysozyme
and ethylenediaminetetraacetate-lysozyme spheroplasts of Escherichia coli. J. Bacteriol.
93, 427–437.
Boyer, H. W., and Roulland-Dussoix, D. (1969). A complementation analysis of the
restriction and modification of DNA in Escherichia coli. J. Mol. Biol. 41, 459–472.
Carstens, C. P. (2003). Use of tRNA-supplemented host strains for expression of heterolo-
gous genes in E. coli. Methods Mol. Biol. 205, 225–233.
Protein Expression in Bacterial Systems 167
De Marco, V., Stier, G., Blandin, S., and de Marco, A. (2004). The solubility and stability of
recombinant proteins are increased by their fusion to NusA. Biochem. Biophys. Res.
Commun. 322, 766–771.
Derman, A. I., Prinz, W. A., Belin, D., and Beckwith, J. (1993). Mutations that allow
disulfide bond formation in the cytoplasm of Escherichia coli. Science 262, 1744–1747.
Dieckman, L., Gu, M., Stols, L., Donnelly, M. I., and Collart, F. R. (2002). High
throughput methods for gene cloning and expression. Protein Expr. Purif. 25, 1–7.
Dieckman, L. J., Hanly, W. C., and Collart, F. R. (2006). Strategies for high-throughput
gene cloning and expression. Genet. Eng. (NY) 27, 179–190.
Dixon, R. A., and Chopra, I. (1986). Leakage of periplasmic proteins from Escherichia coli
mediated by polymyxin B nonapeptide. Antimicrob. Agents Chemother. 29, 781–788.
Donnelly, M. I., Zhou, M., Millard, C. S., Clancy, S., Stols, L., Eschenfeldt, W. H.,
Collart, F. R., and Joachimiak, A. (2006). An expression vector tailored for large-scale,
high-throughput purification of recombinant proteins. Protein Expr. Purif. 47, 446–454.
Dyson, M. R., Shadbolt, S. P., Vincent, K. J., Perera, R. L., and McCafferty, J. (2004).
Production of soluble mammalian proteins in Escherichia coli: Identification of protein
features that correlate with successful expression. BMC Biotechnol. 4, 32.
Eschenfeldt, W. H., Lucy, S., Millard, C. S., Joachimiak, A., and Mark, I. D. (2009).
A family of LIC vectors for high-throughput cloning and purification of proteins.
Methods Mol. Biol. 498, 105–115.
Esposito, D., Garvey, L. A., and Chakiath, C. S. (2009). Gateway cloning for protein
expression. Methods Mol. Biol. 498, 31–54.
Guzman, L. M., Belin, D., Carson, M. J., and Beckwith, J. (1995). Tight regulation,
modulation, and high-level expression by vectors containing the arabinose PBAD pro-
moter. J. Bacteriol. 177, 4121–4130.
Hartley, J. L., Temple, G. F., and Brasch, M. A. (2000). DNA cloning using in vitro site-
specific recombination. Genome Res. 10, 1788–1795.
Haun, R. S., Serventi, I. M., and Moss, J. (1992). Rapid, reliable ligation-independent
cloning of PCR products using modified plasmid vectors. Biotechniques 13, 515–518.
Hengen, P. (1995). Purification of His-Tag fusion proteins from Escherichia coli. Trends
Biochem Sci. 20, 285–286.
Jiang, W., Fang, L., and Inouye, M. (1996). The role of the 5’-end untranslated region of the
mRNA for CspA, the major cold-shock protein of Escherichia coli, in cold-shock
adaptation. J. Bacteriol. 178, 4919–4925.
Kapust, R. B., and Waugh, D. S. (1999). Escherichia coli maltose-binding protein is
uncommonly effective at promoting the solubility of polypeptides to which it is fused.
Protein Sci. 8, 1668–1674.
Klock, H. E., White, A., Koesema, E., and Lesley, S. A. (2005). Methods and results for
semi-automated cloning using integrated robotics. J. Struct. Funct. Genomics 6, 89–94.
Kunji, E. R., Slotboom, D. J., and Poolman, B. (2003). Lactococcus lactis as host for
overproduction of functional membrane proteins. Biochim. Biophys. Acta 1610, 97–108.
Laible, P. D., Scott, H. N., Henry, L., and Hanson, D. K. (2004). Towards higher-
throughput membrane protein production for structural genomics initiatives. J. Struct.
Funct. Genomics 5, 167–172.
Londer, Y. Y., Giuliani, S. E., Peppler, T., and Collart, F. R. (2008). Addressing Shewanella
oneidensis ‘‘cytochromome’’: The first step towards high-throughput expression of cyto-
chromes c. Protein Expr. Purif. 62, 128–137.
Makrides, S. C. (1996). Strategies for achieving high-level expression of genes in Escherichia
coli. Microbiol. Rev. 60, 512–538.
Malamy, M. H., and Horecker, B. L. (1964). Release of alkaline phosphatase from cells of
Escherichia coli upon lysozyme spheroplast formation. Biochemistry 3, 1889–1893.
168 Sarah Zerbs et al.
Matthey, B., Engert, A., Klimka, A., Diehl, V., and Barth, S. (1999). A new series of pET-
derived vectors for high efficiency expression of Pseudomonas exotoxin-based fusion
proteins. Gene 229, 145–153.
Miroux, B., and Walker, J. E. (1996). Over-production of proteins in Escherichia coli: Mutant
hosts that allow synthesis of some membrane proteins and globular proteins at high levels.
J. Mol. Biol. 260, 289–298.
Moy, S., Dieckman, L., Schiffer, M., Maltsev, N., Yu, G. X., and Collart, A. F. (2004).
Genome-scale expression of proteins from Bacillus subtilis. J. Struct. Funct. Genomics 5,
103–109.
Neophytou, I., Harvey, R., Lawrence, J., Marsh, P., Panaretou, B., and Barlow, D. (2007).
Eukaryotic integral membrane protein expression utilizing the Escherichia coli
glycerol-conducting channel protein (GlpF). Appl. Microbiol. Biotechnol. 77, 375–381.
Neu, H. C., and Chou, J. (1967). Release of surface enzymes in Enterobacteriaceae by
osmotic shock. J. Bacteriol. 94, 1934–1945.
Neu, H. C., and Heppel, L. A. (1965). The release of enzymes from Escherichia coli by
osmotic shock and during the formation of spheroplasts. J. Biol. Chem. 240, 3685–3692.
Nossal, N. G., and Heppel, L. A. (1966). The release of enzymes by osmotic shock from
Escherichia coli in exponential phase. J. Biol. Chem. 241, 3055–3062.
Prinz, W. A., Aslund, F., Holmgren, A., and Beckwith, J. (1997). The role of the thior-
edoxin and glutaredoxin pathways in reducing protein disulfide bonds in the Escherichia
coli cytoplasm. J. Biol. Chem. 272, 15661–15667.
Qing, G., Ma, L. C., Khorchid, A., Swapna, G. V., Mal, T. K., Takayama, M. M., Xia, B.,
Phadtare, S., Ke, H., Acton, T., Montelione, G. T., Ikura, M., et al. (2004). Cold-shock
induced high-yield protein production in Escherichia coli. Nat. Biotechnol. 22, 877–882.
Sambrook, J., and Russell, D. W. (2001). Molecular cloning: A laboratory manual. Cold
Spring Harbor Laboratory Press, Cold Spring Harbor, NY.
Schmidt, F. R. (2004). Recombinant expression systems in the pharmaceutical industry.
Appl. Microbiol. Biotechnol. 65, 363–372.
Shaw, A. Z., and Miroux, B. (2003). A general approach for heterologous membrane protein
expression in Escherichia coli: the uncoupling protein, UCP1, as an example. Methods
Mol. Biol. 228, 23–35.
Smith, D. B., and Johnson, K. S. (1988). Single-step purification of polypeptides expressed in
Escherichia coli as fusions with glutathione S-transferase. Gene. 67, 31–40.
Studier, F. W. (2005). Protein production by auto-induction in high density shaking
cultures. Protein Expr. Purif. 41, 207–234.
Studier, F. W., and Moffatt, B. A. (1986). Use of bacteriophage T7 RNA polymerase to
direct selective high-level expression of cloned genes. J. Mol. Biol. 189, 113–130.
Takayama, Y., and Akutsu, H. (2007). Expression in periplasmic space of Shewanella
oneidensis. Protein Expr. Purif. 56, 80–84.
Terpe, K. (2006). Overview of bacterial expression systems for heterologous protein pro-
duction: from molecular and biochemical fundamentals to commercial systems. Appl.
Microbiol. Biotechnol. 72, 211–222.
Woodcock, D. M., Crowther, P. J., Doherty, J., Jefferson, S., DeCruz, E., Noyer-
Weidner, M., Smith, S. S., Michael, M. Z., and Graham, M. W. (1989). Quantitative
evaluation of Escherichia coli host strains for tolerance to cytosine methylation in plasmid
and phage recombinants. Nucleic Acids Res. 17, 3469–3478.
Yokoyama, S. (2003). Protein expression systems for structural genomics and proteomics.
Curr. Opin. Chem. Biol. 7, 39–43.
Zhu, B., Cai, G., Hall, E. O., and Freeman, G. J. (2007). In-fusion assembly: Seamless
engineering of multidomain fusion proteins, modular vectors, and mutations. Biotechniques
43, 354–359.
C H A P T E R T H I R T E E N
Contents
1.Introduction 170
2.Other Fungal Expression Systems 170
3.Culture Media and Microbial Manipulation Techniques 171
4.Genetic Strain Construction 172
4.1. Mating, creating diploids 172
4.2. Random spore analysis 173
5. Gene Preparation and Vector Selection 174
6. Transformation by Electroporation 176
7. DNA Preparation 176
8. Examination of Strains for Recombinant Protein Production 178
9. Assay Development—The Yeastern Blot 182
10. Posttranslational Modification of the Recombinant Protein
(Proteinases and Glycosylation) 184
11. Selection for Multiple Copies of an Expression Cassette 185
References 187
Abstract
The yeast Pichia pastoris has become the premier example of yeast species
used for the production of recombinant proteins. Advantages of this yeast for
expression include tightly regulated and efficient promoters and a strong
tendency for respiratory growth as opposed to fermentative growth. This chapter
assumes the reader is proficient in molecular biology and details the more
yeast specific procedures involved in utilizing the P. pastoris system for gene
expression. Procedures to be found here include: strain construction by classical
yeast genetics, the logic in selection of a vector and strain, preparation of
electrocompetent yeast cells and transformation by electroporation, and the
yeast colony western blot or Yeastern blot method for visualizing secreted
proteins around yeast colonies.
169
170 James M. Cregg et al.
1. Introduction
Relative to other expression systems, yeast got off to a slow start in the
early 1980s, primarily due to poor results with several proteins in baker’s
yeast Saccharomyces cerevisiae (Romanos et al., 1992). Promising results in
shake flask culture, more often than not, led to disappointing yields when
scaled up in fermentor cultures. In addition, yeast were not likely to be of
value in producing human proteins containing N-linked carbohydrates as
injectable pharmaceutical drugs, a major goal of many biotechnology and
pharmaceutical companies, because their sugars were of a high mannose
type in composition and configuration, quite different from that of humans.
As a result, recombinant glycoproteins made in a yeast system have a typical
fungal-like N-glycosylation pattern, which a human immune system recog-
nizes as foreign and rejects. This results in the rapid clearance of yeast
products from the blood and a strong immune response from a patient
that can possibly result in death. Since the 1980s, these problems have been
addressed and several yeasts have become productive alternative systems for
recombinant protein production. As a eukaryotic microbial expression
system, yeast are a good alternative for proteins for which expression in a
bacterial system leads to the synthesis of improperly folded, and inactive
protein aggregates or inclusion bodies.
This review will focus primarily on the most popular of these new yeast
expression systems, Pichia pastoris. Only details of procedures that are specific
or peculiar to expression in P. pastoris will be covered; for more common
methods (e.g., agarose gel electrophoresis, immunoblotting, general recom-
binant DNA methodology, etc.), readers are referred to the many excellent
books describing these methods including: (Sambrook et al., 1989) or the
series by (Ausubel et al., 2001).
AOX1 promoter
Pme I
ZeoR
pJAZ-aMF
pMB origin
ampR
Sfi I
B GGCT ORF
ATTA
TTGGACAAGAGAGAGGCTGA TAATCAAGAGGATGT
AACCTGTTCTCTCTCCGACTCCGA GTTCTCCTACA
L D K R E A E A
with these Glu–Ala repeats (Fig. 13.1). Another set of Alternative Biogram-
matics vectors such as pJAZ-aMF-KR do not contain the Glu–Ala repeats in
the cloning site. To utilize the aMF signal in a pPICZ alpha vector one must
add sequences to the 50 of the gene such that when ligated to the portion of
aMF present in the vector it results in the reconstruction of a functional aMF
signal sequence (either with or without the sequences encoding the Glu–Ala
repeats in aMF). Typically, XhoI is used to cut the pPICZ vector which cuts it
inside the aMF signal encoding sequences. To restore these sequences, an
oligonucleotide with the sequence:
XhoI
TCT CTC GAG AAA AGA GAG GCT GAA GCT-ABC-DEF. . ..
Ser Leu Glu Lys Arg Glu Ala Glu Ala
is synthesized where ‘‘ABC DEF. . .’’ denotes the nucleotide sequences
encoding the first amino acids of the mature recombinant protein. This
oligonucleotide will hybridize appropriately to the 50 end of the gene
encoding the mature protein and result in the incorporation of the missing
portion of the aMF sequence. Similarly, the ‘‘seamless’’ cloning scheme
used to clone genes into the Biogrammatics vectors utilize type IIS restric-
tion sites to join the ABC DEF nucleotides of a gene of interest to the last
Alanine in the aMF by creating a four base ‘‘sticky’’ end comprising the last
nucleotide in a Glu codon and an entire Ala codon (Fig. 13.1).
6. Transformation by Electroporation
At least four different procedures to introduce foreign plasmid DNA
into P. pastoris have been developed using: spheroplast-generation, LiCl,
polyethelene glycol1000 and electroporation. The electroporation proce-
dure is most commonly used and therefore a modified version of that
described by (Becker and Guarante, 1991) will be outlined in detail. For
the other procedures, readers are referred to either of the volumes of
Methods in Molecular Biology: Pichia Protocols (Cregg, 2008; Higgins
and Cregg, 1998).
7. DNA Preparation
For all transformation methods, linear plasmid DNA is most commonly
transformed into P. pastoris for integration into the yeast genome. The DNA
sequence at the ends of the linear plasmid DNA stimulate integration by a
single crossover recombination event into the locus shared by vector and host
genome. Therefore, linearization of an expression plasmid is performed in
Expression in the Yeast Pichia pastoris 177
Pichia DNA in the plasmid, such as in the promoter (i.e., a PmeI site in the
AOXI promoter). The final vector, prepared in E. coli, is cut with a restriction
enzyme that linearizes the vector, and then the DNA is purified and concen-
trated to at least 100 ng/ul in water prior to transformation. At this point, the
vector is ready for transformation into P. pastoris.
Procedure for preparation of electrocompetent cells (Lin-Cereghino
et al., 2005).
Prepare the following (all solutions should be autoclaved except for the
DTT and HEPES solutions, which should be filter sterilized):
1. 500 ml liquid YPD media in a 2.8 l Fernbach culture shaking flask.
2. H2O (1 l).
3. 1 M sorbitol (100 ml).
4. Appropriate selective agar plates.
5. 1 M DTT (2.5 ml).
6. BEDS solution (9 ml): 10 mM Bicine–NaOH, pH 8.3, 3% ethylene
glycol, 5% DMSO (dimethyl sulfoxide), 1 M sorbitol and 0.1 M DTT.
7. 1 M HEPES buffer, pH 8.0 (50 ml).
8. Sterile 250 ml centrifuge tubes.
9. Sterile electroporation cuvettes.
10. Electroporation instrument: BTX Electro Cell Manipulator 600 (BTX,
San Diego, CA); Bio-Rad Gene Pulser (Bio-Rad, Hercules, CA);
Electroporator II (Invitrogen, San Diego, CA). Parameters for electro-
poration with different instruments vary with the instrument. Be sure
to check instructions for each type of instrument. (Becker and
Guarante, 1991; Grey and Brendel, 1995; Pichia Expression Kit
Instruction Manual; Stowers et al., 1995).
Protocol:
1. Inoculate 10 ml YPD media with a single fresh P. pastoris colony of the
strain to be transformed from an agar plate and grow overnight shaking at
30 C.
2. Use the overnight culture to inoculate a 500 ml YPD culture in a 2.8 l
Fernbach culture flask to a starting OD600 of 0.01 and grow to an OD600 of
1.0 ( 12 h).
3. Harvest the cells by centrifugation at 2000g at 4 C, discard the superna-
tant and resuspend the cells in 100 ml of fresh YPD medium plus HEPES
(pH 8.0, 200 mM ) in a sterile 250 ml centrifuge tube.
4. Add 2.5 ml of 1 M DTT and gently mix.
5. Incubate at 30 C for 15 min with slow rotating.
6. Add 150 ml cold water to the culture and wash by centrifugation at 4 C
with an additional 250 ml of cold water. At this stage and from here on,
keep the cells ice cold and do not vortex the cells to resuspend them
(slow pippeting is best).
178 James M. Cregg et al.
Figure 13.2 Plate assay for detection of phytase constitutively expressed and secreted
by selected P. pastoris strains. Top spot: negative control; remaining spots show five
transformants secreting various amounts of the enzyme.
Expression in the Yeast Pichia pastoris 181
Figure 13.3 Expression of intracellular b-lactamase in P. pastoris. The two spots at the
top of the plate are non-b-lactamase expressing negative control strains.
Figure 13.4 ‘‘Yeastern Blot’’ of P. pastoris colonies secreting both heavy and light
chains of IgG antibodies. Negative controls of P. pastoris strain that does not secrete
antibody shown on second row second streak from the left and at first two positions
from the left on bottom row.
1. Transfer freshly grown yeast colonies from the surface of an agar Petri
plate onto a sterile piece of Whatman No. 1 filter paper by a standard
replica-plating method. The filter paper should be cut to a size that
exactly fits within the plate.
2. Place the filter with yeast cells onto the surface of a fresh plate containing
an appropriate induction medium and incubate the plate for 1–2 days.
3. Prepare a piece of nitrocellulose membrane the same size and dimensions
as the filter paper by soaking the membrane for 5 min or more in 15 ml of
transfer buffer (25 mM Tris base, pH 8.5, 0.2 M glycine, 20% methanol).
Soak two additional pieces of cut filter paper in transfer buffer.
4. Prepare a sandwich of the papers as described below:
a. Place one piece of the filter paper on the anode platform of a western
blotter. (It is essential to remove all bubbles between membrane
layers. This can be done by rolling a pipette over their surface.).
b. Place the soaked nitrocellulose filter on top of the filter paper.
c. Place the filter paper with replica-plated yeast cells on top of the
nitrocellulose paper.
d. Place the second piece of soaked filter paper on top of the filter with
cells.
e. Finally, place the cathode plate on top of the sandwich.
5. Transfer proteins to the nitrocellulose membrane with a constant
current (1–4 mA/cm2) for 1 h.
6. Remove the nitrocellulose membrane and wash it for 5 min in 15 ml
TBS buffer (50 mM Tris–HCl, pH 7.6, 150 mM NaCl). Replace the
TBS buffer with 15 ml of TBST (TBST is prepared by adding Tween-
20 to 0.05% to TBS) containing 1–5% bovine serum albumin (BSA).
184 James M. Cregg et al.
portion of the cells. Due to the extremely high culture densities used with
P. pastoris, the concentration of vacuolar proteases in the medium can be
considerable. To utilize a PEP4 strain, one must transform one of these
strains (i.e., SMD1168 ( pep4 his4) or SMD1168H ( pep4)) with the expres-
sion vector or create deletions in the PEP4 gene in a strain already expres-
sing a protein of interest. To determine if PEP4 deficient strains benefit
the expression of the protein, the recombinant protein should be examined
from wild type and the PEP4 deficient strains after an induction performed
in parallel. Note: PEP4 deficient strains of P. pastoris are not as robust as
wild-type strains. In particular, PEP4 deficient strains die more quickly
when stored on plates, do not transform as efficiently as wild-type strains,
grow more slowly in culture, and are more difficult to induce on methanol.
Furthermore, PEP4 deficient strains are difficult to mate with other non-
PEP4 P. pastoris strains.
If a recombinant protein expressed in P. pastoris is larger than expected
and somewhat heterogenous in size by SDS–PAGE analysis, it may be
glycosylated. One should first examine the amino acid sequence of the
protein for potential glycosylation sites: the signal for addition of N-linked
oligosaccharides is ASN-X-Ser/Thr, and O-linked sugars can be added to
the OH-group of any Ser or Thr residue. Glycosylation can be confirmed by
deglycoslyating suspect protein with PNGase F and examining the product
by SDS–PAGE. A good protocol for deglycoslation of proteins can be found
at: http://www.neb.com/nebecomm/products/productP0704.asp. If this
treatment reduces the apparent molecular size of your protein resulting in a
more homogenous product, the protein is almost certainly glycosylated.
REFERENCES
Ausubel, E. M., Brent, R., Kingston, E., Moore, D. D., Seidman, J. G., Smith, J. A., and
Struhl, K. (eds.) (2001). In ‘‘Current Protocols in Molecular Biology’’ John Wiley and
Sons, Inc., New York.
Becker, D. M., and Guarante, L. (1991). High-efficiency transformation of yeast by electro-
poration. Methods Enzymol. 194, 182–187.
Boer, E., Gellisson, G., and Kunze, G. (2005). Arxula adeninivorans. In ‘‘Production of
Recombinant Proteins’’ (G. Gellisson, ed.), pp. 89–110. Wiley-VCH Verlag GmbH and
Co. KGaA, Weinheim, Germany. Chapter 5.
Brierley, R. A. (1998). Secretion of recombinant human insulin-like growth factor I (IGF-I).
In ‘‘Methods in Molecular Biology Vol. 103 Pichia Protocols’’ (D. R. Higgins and
J. M. Cregg, eds.), pp. 149–178. Humana Press, Totowa, NJ.
Cregg, J. M. (2008). DNA-mediated transformation. In ‘‘Methods in Molecular Biology:
Pichia Protocols’’ ( J. M. Cregg ed.), 2nd ed. pp. 27–42. Humana Press, Totowa, NJ.
Chapter 3.
188 James M. Cregg et al.
(H. Heslot, J. Davies, J. Florent, L. Bobichon, G. Durand, and L. Penasse, eds.) Vol. II,
pp. 477–490. Societe Francaise de Microbiologie, Paris.
Tolstorukov, I., and Cregg, J. M. (2008). Classical Genetics. In ‘‘Methods in Molecular
Biology: Pichia Protocols’’ ( J. M. Cregg, ed.), 2nd ed. pp. 189–202. Humana Press,
Totowa, NJ. Chapter 14.
van Ooyen, A. J., Dekker, P., Huang, M., Olsthoorn, M. M., Jacobs, D. I., Colussi, P. A.,
and Taron, C. H. (2006). Heterologous protein production in the yeast Kluyveromyces
lactis. FEMS Yeast Res. 6, 381–392.
C H A P T E R F O U R T E E N
Baculovirus–Insect Cell
Expression Systems
Donald L. Jarvis
Contents
1. Introduction 192
2. A Brief Overview of Baculovirus Biology and Molecular Biology 193
3. Baculovirus Expression Vectors 195
4. Baculovirus Expression Vector Technology—The Early Years 196
5. Baculovirus Expression Vector Technology—Improved 198
6. Baculovirus Transfer Plasmid Modifications 198
7. Parental Baculovirus Genome Modifications 200
8. The Other Half of the Baculovirus–Insect Cell System 210
9. A New Generation of Insect Cell Hosts for Baculovirus
Expression Vectors 212
10. Basic Baculovirus Protocols 214
10.1. Insect cell maintenance 215
10.2. Isolation of baculovirus genomic DNA 215
10.3. Baculovirus plaque assays 216
References 218
Abstract
In the early 1980s, the first-published reports of baculovirus-mediated foreign
gene expression stimulated great interest in the use of baculovirus–insect cell
systems for recombinant protein production. Initially, this system appeared to
be the first that would be able to provide the high production levels associated
with bacterial systems and the eukaryotic protein processing capabilities asso-
ciated with mammalian systems. Experience and an increased understanding of
basic insect cell biology have shown that these early expectations were not
completely realistic. Nevertheless, baculovirus–insect cell expression systems
have the capacity to produce many recombinant proteins at high levels and they
also provide significant eukaryotic protein processing capabilities. Furthermore,
important technological advances over the past 20 years have improved upon
191
192 Donald L. Jarvis
1. Introduction
It seems that nearly every book chapter or review focusing on the
baculovirus–insect cell expression system begins with a statement emphasiz-
ing the popularity of this system and/or noting its widespread use for
recombinant protein production. While this introduction has become
increasingly redundant, it also has become increasingly accurate over the
past 20 years. An advanced PubMed search using the terms ‘‘baculovirus’’
and ‘‘expression’’ and ‘‘vector’’ at the time of this writing yielded over 2000
hits. While this is a significant number, it vastly underestimates the actual
number of published studies involving recombinant protein production
using the baculovirus–insect cell expression system. Moreover, it does not
include the large number of studies performed behind closed doors in the
biotechnology industry, which are clearly evidenced by the number of oral
presentations given by industrial scientists at various baculovirus–insect cell
technology conferences held over the past two decades.
The previous edition of this Guide to Protein Purification (1990)
included a comprehensive description of the technical details involved in
using the baculovirus–insect cell expression system, as it existed at that time.
This exercise will not be repeated here and the reader should refer to the
previous edition of this book and other sources given below for those
details. In this edition, I will begin by providing the background informa-
tion needed for the reader to understand the basic principles underlying the
original baculovirus–insect cell expression system. I will then focus on the
new tools and approaches developed for the isolation of recombinant
baculoviruses since the publication of the previous edition of this book, as
these have greatly facilitated the ability of virtually any biomedical investi-
gator to utilize the baculovirus–insect cell system, relative to the state of the
art in 1990.
Baculovirus–Insect Cell Expression Systems 193
Plasmid backbone
(e.g. pUC-based)
Transfer plasmid
Transfer plasmid
Polyhedrin gene
Foreign gene
genetically unstable and the newly acquired foreign gene typically would be
lost within a round or two of viral replication. Despite this major limitation,
transfer plasmids that incorporated marker genes into the recombinant
baculovirus genome were widely used and particularly useful in the hands
of investigators aware of the potential single crossover recombination trap.
These investigators would simply use incorporation of the marker gene as a
prescreen and then perform additional screening to map the genomic
position of a foreign gene and confirm that a double crossover recombina-
tion event was responsible for transfer of the gene of interest to the
recombinant baculovirus vector.
The second purpose served by transfer plasmid modifications was to
facilitate recombinant protein expression and purification in the
baculovirus–insect cell system. While the specific, individual modifications
that were undertaken are far too numerous to discuss here, three general
types of transfer plasmid modifications that fell within this category included
the addition of sequences encoding secretory signal peptides, the addition
of sequences encoding amino- or carboxy-terminal purification tags (e.g.,
6 HIS), replacement of the polyhedrin promoter with alternate baculo-
virus promoters, and replacement of the polyhedrin promoter with multiple
promoter elements, which could simultaneously drive the expression of
multiple recombinant proteins during baculovirus infection.
A nearly comprehensive list of baculovirus transfer plasmids with the
two general types of modifications discussed above, together with short
descriptions of their specific, functional features was recently published and
this list is a good source of information on this topic (Possee and King,
2007). However, one type of modification that was not included and
deserves further consideration is the ‘‘immediate-early’’ transfer plasmid,
in which the polyhedrin promoter is replaced by a promoter from a
baculovirus immediate-early gene, such as the AcMNPV ie1 gene ( Jarvis
et al., 1996). The idea of using promoters from earlier classes of baculovirus
genes, such as the ie1 gene, might seem counterintuitive because these
promoters are weaker than the polyhedrin promoter. However, there is
evidence to suggest that higher quality products can be obtained using
recombinant baculoviruses that express foreign genes under the control of
earlier classes of promoters, despite their inability to drive equally high levels
of transcription (Chazenbalk and Rapoport, 1995; Hill-Perkins and Possee,
1990; Jarvis et al., 1996; Murphy et al., 1990; Rankl et al., 1994; Sridhar
et al., 1993). In fact, this approach can be particularly useful for secretory
pathway proteins, which are often produced at relatively low levels when
their genes are expressed under polyhedrin control, as mentioned above.
In this circumstance, an investigator can potentially gain the advantage of
initiating foreign gene expression at earlier times of infection, thereby
avoiding the potentially adverse effects of baculovirus infection on host
protein processing pathways, without sacrificing high production levels.
200 Donald L. Jarvis
Bsu36I
Digest with Bsu36I
Recombine/repair with
transfer plasmid
(co-transfected insect cells)
Bsu36I sites
Digest with Bsu36I
Recombine/repair with
transfer plasmid
(co-transfected insect cells)
Transposition
Recombinant bacmid
Selection
Screening
Bacmid extraction
Recombinant baculoviruses
Recombinant bacmid
ends of Tn7, the chimeric gene could be efficiently transposed from the
transfer plasmid to the polyhedrin locus of the bacmid. This transposition
event would simultaneously delete the lacZ gene in the bacmid, allowing
bacteria containing recombinant bacmids to be detected by standard blue–
white screening on a selective bacterial medium. Thus, one could simply
isolate the recombinant bacmid DNA encoding the gene of interest from an
E. coli clone with a white colony phenotype and use this DNA to transfect
insect cells. Transfection of the DNA would initiate a viral infection and
lead to the production of recombinant baculovirus progeny. This in vivo
transposition approach, which can be used to produce recombinant
baculovirus DNAs at 100% efficiency, was initially commercialized as the
Bac-to-BacTM system by Life Technologies and is now available from
Invitrogen. The bacteriocentric nature of the Bac-to-BacTM system was
important because it allowed investigators who were less familiar with
virological and more familiar with basic bacteriological and molecular
biological methods to work within their comfort zones. One disadvantage
of this system, however, is the fact that it yields recombinant baculoviruses
that retain the bacterial replicon, which appears to be associated with a
higher level of genetically instability upon serial passage of these viruses in
insect cells, relative to baculoviruses lacking this element (Pijlman et al.,
2003; and see below).
Most recently, the Possee and King groups have described a clever cross-
hybridization of the basic ideas underlying the linearizable baculoviral DNA
and bacmid approaches (Possee et al., 2008). In essence, they have created a
new type of bacmid, which consists of a recombinant baculoviral genome
with a bacterial replicon in the polyhedrin locus and a deletion in the
downstream orf 1629 gene (Fig. 14.6). Due to the presence of the replicon
9
orf60 629 orf60 162
3 orf1 3 bac replicon Δorf
Foreign gene
Homologous recombination
bac replicon Δorf1629
(co-transfected insect cells) orf1629
orf603 3
orf60 Foreign gene
Figure 14.6 Producing a baculovirus expression vector using the flashBAC approach.
Baculovirus–Insect Cell Expression Systems 205
and the orf 1629 deletion, this bacmid can replicate in E. coli, but not in
insect cells. Therefore, an E. coli strain carrying this bacmid provides a
convenient source of replication defective parental baculovirus genomic
DNA for recombinant viral vector production. One can simply isolate the
viral DNA from E. coli, mix it with a transfer plasmid, and then use the
mixture to cotransfect insect cells. Homologous recombination between
the viral and transfer plasmid DNAs will simultaneously restore orf 1629,
knockout the bacterial replicon in the polyhedrin locus, and knock-in the
gene of interest at this same site. While this approach does not improve the
efficiency of homologous recombination, it can provide an extremely high
efficiency of recombinant baculovirus production by the cotransfected
insect cells because the parental viral DNA is defective and cannot replicate
on its own. This approach has been commercialized as flashBACTM
by Oxford Expression Technologies. One feature of the flashBACTM sys-
tem that warrants additional emphasis is that the bacterial replicon is deleted
from the bacmid upon homologous recombination with the transfer
plasmid, as mentioned above. This is advantageous because it eliminates
potential problems with the genetic stability of baculovirus expression
vectors produced using conventional bacmids, which retain the bacterial
replicon within their genomes (Pijlman et al., 2003). Another feature that
warrants additional emphasis is that the bacmid used in the flashBACTM
system consists of a defective baculovirus genome that cannot replicate in
insect cells. Clearly, this means that homologous recombination between
the defective parental viral genome and the transfer plasmid is required to
produce helper-independent baculovirus progeny in this system. However,
it is important to recognize that this does not necessarily mean that the
cotransfected insect cells will produce only recombinant baculovirus prog-
eny. These cells also can produce progeny derived from the defective
parental viral genome as a result of genetic complementation. Specifically,
the recombinant virus produced in the cotransfected insect cells could
provide the orf 1629 product as a helper function needed to package defec-
tive viral DNAs in trans. Thus, while Oxford Expression Systems’ commer-
cial literature indicates that it is not necessary, baculovirus vectors produced
using the flashBACTM system should be subjected to at least one round of
plaque purification. Without this step, primary recombinant virus stocks
produced by cotransfected insect cells are likely to be contaminated with
defective, parental viruses that could interfere with downstream vector
replication and foreign gene expression.
Arguably, the simplest way to produce baculovirus vectors is to use a
cross-hybridization of the basic ideas underlying GatewayÒ technology
(Hartley, 2003; Walhout et al., 2000) and the linearized parental viral
DNA approach, which involves in vitro, rather than in vivo recombination
between a GatewayÒ entry plasmid encoding a protein of interest and the
linearized parental viral DNA. In this system, the parental viral DNA is a
206 Donald L. Jarvis
Donor
Site-specific
Polyhedrin
(gateway entry clone)
promoter
recombination
Foreign gene
attR1 IE0-HSV-TK p10-LacZ attR2
(in vitro) Polyhedrin
Bsu36I promoter
obvious. It has been suggested that the viral chitinase gene product, which is
a resident ER protein (Thomas et al., 1998), interferes with the production
of secretory pathway proteins by contributing to the saturation of host
protein translocation machinery (Kaba et al., 2004). If this is true, then
deletion of the viral chitinase gene might be expected to increase secretory
pathway protein production levels by eliminating production of this prod-
uct. However, there are no published studies documenting the impact of
deleting the viral chitinase gene, alone, in an AcMNPV-based vector such
as flashBACTM, and the study documenting the reduced degradation of a
foreign glycoprotein produced by AcBacDCC is complicated by the fact
that this vector lacks both the viral chitinase and cathepsin-like protease
genes. On the other hand, one publication has documented the impact of
deleting the viral chitinase gene, cathepsin-like protease gene, or both, on
the production of a foreign protein by recombinant silkworm baculoviruses
(Bombyx mori nucleopolyhedrovirus; BmNPV; Lee et al., 2006). In this
system, recombinant BmNPV vectors encoding an insect cellulase pro-
duced higher levels of both the foreign protein and its enzymatic activity
in silkworms when either the chitinase or the cathepsin-like protease genes
were deleted. Furthermore, a vector lacking both of these viral genes
produced the highest levels of cellulase protein and enzyme activity. One
could speculate that these results indicate that deleting the viral chitinase
gene from AcMNPV-based vectors would have similar benefits. However,
this speculation would be weakened by the fact that silkworms, not insect
cell lines, were used as the hosts in this study.
DiamondBacTM is another example of a parental baculovirus DNA that has
a deletion in an accessory gene to potentially enhance recombinant protein
production in the baculovirus–insect cell system. As mentioned above, Dia-
mondBacTM is a commercial, prelinearized baculovirus DNA analogous to
BakPAK6, but this viral DNA also lacks a functional p10 gene, which has been
implicated in host cell lysis (Williams et al., 1989). Accordingly, the manufac-
turer’s commercial literature suggests that cells infected with a recombinant
baculovirus vector produced using this parental virus retain higher cell viabil-
ities throughout the course of infection, which contributes to higher recom-
binant protein production levels (Sigma-Aldrich, 2008). Another interesting
and potentially useful feature of DiamondBacTM is that the p10 gene in this
baculoviral genome has been replaced by a protein disulfide isomerase (PDI)
gene, which encodes a chaperone that drives disulfide bonding and contributes
to protein folding. Published evidence suggests that coexpressing PDI in the
baculovirus–insect cell system increases the solubility and secretion of recom-
binant IgG (Hsu et al., 1996). The commercial literature from Sigma-Aldrich
states that the p10 gene deletion and PDI gene insertion provide ‘‘up to a 10-
fold increase in overall protein production for many recombinant proteins’’
(Sigma-Aldrich, 2008). However, there are no published studies comparing
foreign protein production levels by matched pairs of recombinant baculovirus
210 Donald L. Jarvis
vectors with and without the p10 gene deletion and/or the PDI insertion.
Thus, in the final analysis, directed studies with appropriately matched pairs of
AcMNPV-based vectors will be needed to directly determine the general
impact of deleting nonessential viral genes, such as the viral chitinase, cathep-
sin-like protease, and p10 genes, on foreign protein production in the
baculovirus–insect cell system.
In addition to the DiamondBacTM parental baculovirus DNA, other
baculovirus vectors with gene insertions encoding heterologous protein pro-
cessing enzymes designed to enhance foreign protein production have been
described in the literature. One is a recombinant baculovirus that encodes a
polydnavirus vankyrin gene under the control of the p10 promoter (Fath-
Goodin et al., 2006). It has been found that the expression of certain vankyrin
gene products prolongs the viability of baculovirus-infected Sf 9 cells and this,
in turn, can increase the amounts of a foreign protein produced by coexpres-
sion under the control of the polyhedrin promoter. Other types of baculovirus
vectors that fall within this genre include those encoding heterologous glyco-
syltransferases ( Jarvis et al., 2001; Tomiya et al., 2003) or enzymes involved in
CMP-sialic acid biosynthesis (Hill et al., 2006) under the control of baculo-
virus ie1 promoters. The use of the immediate-early promoter in these vectors
allows the newly introduced processing functions to be expressed early in the
infection cycle, well before expression of the foreign gene encoding the
protein of interest. This allows for functional extension of the endogenous
insect cell protein N-glycosylation pathway and the establishment of a path-
way capable of producing glycoproteins with humanized carbohydrate side
chains prior to the onset of production of the glycoprotein of interest.
A transfer plasmid that can be used to simultaneously introduce a poly-
dnavirus vankyrin gene and a foreign gene encoding a protein of interest into
the baculovirus genome is commercially available from Paratechs. Similarly,
parental baculoviral DNAs encoding higher eukaryotic N-glycosylation
functions are likely to be commercially available for recombinant baculovirus
vector isolation in the near future. These tools will facilitate a more global
analysis of the capabilities of these types of vectors in increasing expression
levels, on the one hand, or producing foreign glycoproteins with humanized
carbohydrate side chains, on the other, as neither has been broadly assessed
with a wide array of different foreign protein/glycoprotein products.
components. The first one is, of course, the baculovirus expression vector,
which is an insect virus whose obvious function is to deliver a foreign
gene(s) encoding a protein(s) of interest to the host. In addition, the
baculovirus expression vector typically serves another function, which is
to induce the production of the transcriptional complex needed to tran-
scribe the foreign gene of interest under the control of a late or very late
baculovirus promoter. The second component, which has been given
almost no attention so far, is the host, which is typically a lepidopteran
insect cell line, but alternatively can be a lepidopteran insect.
The Order Lepidoptera includes the moths and butterflies, which are the
hosts for many viruses in the Family Baculoviridae, including AcMNPV.
Grace described the first-established lepidopteran insect cell cultures in 1962
(Grace, 1962) and as of today, over 250 insect cell lines have been described
(Lynn, 2007). However, we can focus on just two lepidopteran insect cell
lines, which are the two most commonly used hosts for baculovirus expres-
sion vectors. One is Sf 9, a cell line described by Smith and Summers in
1987 (Summers and Smith, 1987) as a subclone of IPLB-SF-21, which is a
line that had been isolated from pupal ovarian tissue of Spodoptera frugiperda
(fall armyworm) in 1977 (Vaughn et al., 1977). Sf 9 cells can be obtained
from several different companies, including Invitrogen and Novagen, or the
American Type Culture Collection (ATCC). The other is High FiveTM, a
cell line originally isolated from adult ovarian tissue of Trichoplusia ni
(cabbage looper) and described as BTI Tn 5B-1 by Granados’ and
Wood’s groups in 1992 (Davis et al., 1992; Wickham et al., 1992) and
then commercialized by Invitrogen as High FiveTM.
One interesting and useful feature of Sf 9 and High FiveTM cells is that
they can grow in either adherent or suspension formats. Thus, it is conve-
nient to perform lab-scale protein production experiments by infecting a
few million Sf 9 or High FiveTM cells as adherent cultures in plates or flasks.
Adherent cultures of Sf 21 or Sf 9 cells also are routinely used to plaque
purify and quantify recombinant baculovirus expression vectors. On the
other hand, both Sf 9 and High FiveTM cells can be scaled up to varying
degrees in spinner flasks, shake flasks, or in airlift, stirred tank (Elias et al.,
2007), or Wave (Kadwell and Hardwicke, 2007) bioreactors to produce
larger amounts of recombinant proteins.
The conditions and methods used to culture insect cells are quite
different from those used to culture mammalian cells, which are usually
more familiar to most investigators. For example, the optimal temperature
for insect cell culture is 28 C, rather than 37 C. In addition, both Sf 9 and
High FiveTM cells are loosely adherent and neither EDTA nor trypsin is
required for their subculture. It is not necessary to have a CO2 incubator to
grow insect cells because insect cell culture media are buffered with phos-
phate, rather than carbonate. Sf 9 and High FiveTM cells both can be cul-
tured in either growth media supplemented with serum or in serum-free
212 Donald L. Jarvis
media and both types of growth media are now widely available. In fact,
since the publication of the first edition of this book in 1990, many different
insect cell culture media have become available from many different man-
ufacturers, including Expression Systems, Invitrogen (GIBCO), HyClone,
and Sigma-Aldrich, among others. However, it is important to note that
Sf 9 and High FiveTM cells tend to become adapted to a specific growth
medium. Thus, the abrupt transition of an Sf 9 or High FiveTM culture from
a growth medium supplemented with serum to a serum-free medium can
result in loss of the culture. A much higher rate of success can be achieved
by slowly weaning the cells out of serum containing and into serum-free
medium. For example, one can increase the proportion of serum-free
medium from 0 to 25, 50%, 75, and 100% over four consecutive passages.
Even with this relatively slow type of weaning process, these cells typically
undergo a temporary arrest or grow very slowly when first placed into the
100% serum-free medium. Thus, it is important to be patient, as the cells
will typically recover after a lag of a day or two and begin growing normally
again in the new medium. This type of weaning process is not only useful
for the transition from a serum containing to a serum-free medium, but is
also sometimes needed to move insect cell cultures from one specific type of
serum-free medium into another.
published before and since that time (King and Possee, 1992; Murhammer,
2007; O’Reilly et al., 1992; Richardson, 1995; Summers and Smith, 1987).
Finally, each manufacturer of the new baculovirus transfer plasmids and
genome backbones described in this chapter, which were created to facilitate
the production and isolation of recombinant baculovirus expression vectors,
provides detailed maps, sequences, and stepwise protocols for the use of their
materials, which are freely available at their Web sites. Accordingly, I have
chosen to share a relatively small, but essential set of time-tested protocols
used in my lab, which were originally taken from the original baculovirus
users ‘‘manuals’’ by Summers and Smith (1987) and O’Reilly et al. (1992)
and, in some cases, modified to fit our needs. The reader is referred to the
sources given above for more direct and detailed information on any of the
specific approaches, transfer plasmids, and parental baculovirus vectors
described in this chapter, which need not be reiterated here.
cells in TNM-FH containing 10% fetal bovine serum and 0.1% pluronic
F68, as described above. The requisite number of cells is gently centri-
fuged out of the old growth medium, resuspended in fresh TNM-FH with
no additives, and then seeded into 6-well cell culture plates (Corning) at a
density of 0.75 106 cells/well. The cells are then allowed to settle and
attach to the plastic at 28 C for about an hour. The cells will adhere more
tightly to the plastic when seeded in the absence than when seeded in the
presence of serum. While the cells are attaching, the serial 10-fold dilutions
of the virus stock are performed, working under the assumption that a
typical baculovirus stock will have a plaque assay titer of 1 107 pfu/
mL. We use TNM-FH supplemented with fetal bovine serum and
pluronic F68 as the diluent, with 0.1 mL of virus and 9.9 mL dilution
blanks for 100-fold dilutions and 0.5 mL of virus and 4.5 mL dilution
blanks for 10-fold dilutions. The dilution blanks are set up and the
dilutions performed under aseptic conditions and each dilution is mixed
very well with a vortex mixer between each transfer. After the dilutions
have been performed and the cells have attached to the culture plates, the
cells are inoculated using two wells for each appropriately diluted virus.
The medium is removed from the wells and replaced with 2 mL/well of
the diluted virus. This is a critical step in the assay because the cells dry out
rather quickly and those along the edges will die if you move too slowly.
Thus, it is best to drain the media from only two wells at a time and
immediately inoculate those wells with one virus dilution in duplicate
before moving on to another pair of wells and a new dilution. We
typically plate the virus at dilutions of 10 5, 10 6, and 10 7 to try to
obtain reasonable numbers of plaques for counting. The virus is then
allowed to attach to the cells for 1 h at 28 C and the overlay medium
is prepared during the viral adsorption period. We prepare 125 mL bottles
containing 50 mL of 2% (w/v) SeaplaqueÒ low melting temperature
agarose (Cambrex) in water, then autoclave the bottles and allow the
agarose to solidify. These bottles of solid agarose then can be microwaved
to remelt the agarose, cooled to 60 C, and the liquid agarose can be
mixed with an equal volume of 2 Grace’s medium prewarmed to
30 C (you will need 3 mL per well). If a blue–white screen is being
used to help identify putative recombinant viruses, the chromogenic
substrate should be added to the overlay at this time. The overlay is then
swirled to obtain an even mixture and cooled to about 40–42 C. Finally,
the infected cells are removed from the incubator, the viral inocula
are completely removed with pipettes, and 3 mL of overlay are very gentle
dribbled down the edge of the well. Again, draining two wells at a time
and overlaying those wells with agarose before moving on to another pair
of wells will prevent the cells from drying out and dying. After the overlay
hardens for 15 min, the plates are incubate in sealed plastic baggie,
upside down, for 7–10 days at 28 C. A dissecting microscope is used to
218 Donald L. Jarvis
visualize plaques for careful counting and the plaque counts are used to
calculate the titer of the original virus stock, taking the dilution factor and
plating volume into account.
REFERENCES
Aumiller, J. J., Hollister, J. R., and Jarvis, D. L. (2003). A transgenic lepidopteran insect cell
line engineered to produce CMP-sialic acid and sialoglycoproteins. Glycobiology 13,
497–507.
Azzouz, N., Kedees, M. H., Gerold, P., Becker, S., Dubremetz, J. F., Klenk, H. D., and
Eckert, V. R. T. S. (2000). An early step of glycosylphosphatidyl-inositol anchor
biosynthesis is abolished in lepidopteran insect cells following baculovirus infection.
Glycobiology 10, 177–183.
Berger, I., Fitzgerald, D. J., and Richmond, T. J. (2004). Baculovirus expression system for
heterologous multiprotein complexes. Nat. Biotechnol. 22, 1583–1587.
Bradley, M. K. (1990). Overexpression of proteins in eukaryotes. Methods Enzymol. 182,
112–132.
Breitbach, K., and Jarvis, D. L. (2001). Improved glycosylation of a foreign protein by
Tn-5B1–4 cells engineered to express mammalian glycosyltransferases. Biotechnol. Bioeng.
74, 230–239.
Chazenbalk, G. D., and Rapoport, B. (1995). Expression of the extracellular domain of the
thyrotropin receptor in the baculovirus system using a promoter active earlier than the
polyhedrin promoter Implications for the expression of functional highly glycosylated
proteins. J. Biol. Chem. 270, 1543–1549.
Davis, T. R., Trotter, K. M., Granados, R. R., and Wood, H. A. (1992). Baculovirus
expression of alkaline phosphatase as a reporter gene for evaluation of production,
glycosylation and secretion. Nat. Biotechnol. 10, 1148–1150.
Douris, V., Swevers, L., Labropoulou, V., Andronopoulou, E., Georgoussi, Z., and
Iatrou, K. (2006). Stably transformed insect cell lines: Tools for expression of secreted
and membrane-anchored proteins and high-throughput screening platforms for drug and
insecticide discovery. Adv. Virus Res. 68, 113–156.
Elias, C. B., Jardin, B., and Kamen, A. (2007). Recombinant protein production in large-
scale agitated bioreactors using the baculovirus expression vector system. Methods Mol.
Biol. 388, 225–246.
Ernst, W. J., Grabherr, R. M., and Katinger, H. W. (1994). Direct cloning into the
Autographa californica nuclear polyhedrosis virus for generation of recombinant
baculoviruses. Nucleic Acids Res. 22, 2855–2856.
Fath-Goodin, A., Kroemer, J., Martin, S., Reeves, K., and Webb, B. A. (2006). Polydna-
virus genes that enhance the baculovirus expression vector system. Adv. Virus Res. 68,
75–90.
Friesen, P. D. (1997). Regulation of baculovirus early gene expression. In ‘‘The Baculo-
viruses’’ (L. K. Miller, ed.), pp. 141–170. Plenum Press, New York.
Grace, T. D. C. (1962). Establishment of four strains of cells from insect tissues grown in vitro.
Nature 195, 788–789.
Harrison, R. L., and Jarvis, D. L. (2006). Protein N-glycosylation in the baculovirus-insect
cell expression system and efforts to engineer insect cells to produce ‘‘mammalianized’’
recombinant glycoproteins. Adv. Virus Res. 68, 159–191.
Harrison, R. L., and Jarvis, D. L. (2007a). Transforming lepidopteran insect cells for
continuous recombinant protein expression. Methods Mol. Biol. 388, 299–316.
Baculovirus–Insect Cell Expression Systems 219
Harrison, R. L., and Jarvis, D. L. (2007b). Transforming lepidopteran insect cells for
improved protein processing. Methods Mol. Biol. 388, 341–356.
Strausberg, R. L., and Strausberg, S. L. (2003). Overview of protein expression in Saccha-
romyces cerevisiae. Curr. Protoc. Protein Sci. Chapter 5, Unit 5.17.
Hawtin, R. E., Arnold, K., Ayres, M. D., Zanotto, P. M., Howard, S. C., Gooday, G. W.,
Chappell, L. H., Kitts, P. A., King, L. A., and Possee, R. D. (1995). Identification and
preliminary characterization of a chitinase gene in the Autographa californica nuclear
polyhedrosis virus genome. Virology 212, 673–685.
Hill, D. R., Aumiller, J. J., Shi, X., and Jarvis, D. L. (2006). Isolation and analysis of a
baculovirus vector that supports recombinant glycoprotein sialylation by SfSWT-1 cells
cultured in serum-free medium. Biotechnol. Bioeng. 95, 37–47.
Hill-Perkins, M. S., and Possee, R. D. (1990). A baculovirus expression vector derived from
the basic protein promoter of Autographa californica nuclear polyhedrosis virus. J. Gen.
Virol. 71, 971–976.
Hollister, J., and Jarvis, D. L. (2001). Engineering lepidopteran insect cells for sialoglyco-
protein production by genetic transformation with mammalian b1, 4-galactosyltransfer-
ase and alpha 2, 6-sialyltransferase genes. Glycobiology 11, 1–9.
Hollister, J. R., Grabenhorst, E., Nimtz, M., Conradt, H. O., and Jarvis, D. L. (2002).
Engineering the protein N-glycosylation pathway in insect cells for production of
biantennary, complex N-glycans. Biochemistry 41, 15093–15104.
Hollister, J. R., Shaper, J. H., and Jarvis, D. L. (1998). Stable expression of mammalian beta
1, 4-galactosyltransferase extends the N-glycosylation pathway in insect cells. Glycobiology
8, 473–480.
Hsu, T. A., Watson, S., Eiden, J. J., and Betenbaugh, M. J. (1996). Rescue of immunoglo-
bulins from insolubility is facilitated by PDI in the baculovirus expression system. Prot.
Exp. Purif. 7, 281–288.
Jarvis, D. L. (1997). Baculovirus expression vectors. In ‘‘The Baculoviruses’’ (L. K. Miller,
ed.), pp. 389–431. Plenum Press, New York.
Jarvis, D. L., Fleming, J. A., Kovacs, G. R., Summers, M. D., and Guarino, L. A. (1990). Use
of early baculovirus promoters for continuous expression and efficient processing of
foreign gene products in stably transformed lepidopteran cells. Nat. Biotechnol. 8,
950–955.
Jarvis, D. L., and Guarino, L. A. (1995). Continuous foreign gene expression in transformed
lepidopteran insect cells. In ‘‘Baculovirus Expression Protocols’’ (C. D. Richardson, ed.),
Vol. 39, pp. 187–202. Humana Press, Clifton, NJ.
Jarvis, D. L., Howe, D., and Aumiller, J. J. (2001). Novel baculovirus expression vectors that
provide sialylation of recombinant glycoproteins in lepidopteran insect cells. J. Virol. 75,
6223–6227.
Jarvis, D. L., and Summers, M. D. (1989). Glycosylation and secretion of human tissue
plasminogen activator in recombinant baculovirus-infected insect cells. Mol. Cell. Biol. 9,
214–223.
Jarvis, D. L., Weinkauf, C., and Guarino, L. A. (1996). Immediate early baculovirus vectors
for foreign gene expression in transformed or infected insect cells. Prot. Exp. Purif. 8,
191–203.
Kaba, S. A., Salcedo, A. M., Wafula, P. O., Vlak, J. M., and van Oers, M. M. (2004).
Development of a chitinase and v-cathepsin negative bacmid for improved integrity of
secreted recombinant proteins. J. Virol. Methods 122, 113–118.
Kadwell, S. H., and Hardwicke, P. I. (2007). Production of baculovirus-expressed recombi-
nant proteins in wave bioreactors. Methods Mol. Biol. 388, 247–266.
King, L. A., and Possee, R. D. (1992). The baculovirus system: A laboratory guide.
Chapman and Hall, London.
220 Donald L. Jarvis
Kitts, P. A., Ayres, M. D., and Possee, R. D. (1990). Linearization of baculovirus DNA
enhances the recovery of recombinant virus expression vectors. Nucleic Acids Res. 18,
5667–5672.
Kitts, P. A., and Possee, R. D. (1993). A method for producing recombinant baculovirus
expression vectors at high frequency. Biotechniques 14, 810–817.
Kool, M., Voncken, J. W., van Lier, F. L., Tramper, J., and Vlak, J. M. (1991). Detection
and analysis of Autographa californica nuclear polyhedrosis virus mutants with defective
interfering properties. Virology 183, 739–746.
Kost, T. A., Condreay, J. P., and Jarvis, D. L. (2005). Baculovirus as versatile vectors for
protein expression in insect and mammalian cells. Nat. Biotechnol. 23, 567–575.
Lee, K. S., Je, Y. H., Woo, S. D., Sohn, H. D., and Jin, B. R. (2006). Production of a
cellulase in silkworm larvae using a recombinant Bombyx mori nucleopolyhedrovirus
lacking the virus-encoded chitinase and cathepsin genes. Biotechnol. Lett. 28, 645–650.
Legardinier, S., Klett, D., Poirier, J. C., Combarnous, Y., and Cahoreau, C. (2005).
Mammalian-like nonsialyl complex-type N-glycosylation of equine gonadotropins in
Mimic insect cells. Glycobiology 15, 776–790.
Lihoradova, O. A., Ogay, I. D., Abdukarimov, A. A., Azimova Sh, S., Lynn, D. E., and
Slack, J. M. (2007). The Homingbac baculovirus cloning system: An alternative way to
introduce foreign DNA into baculovirus genomes. J. Virol. Methods 140, 59–65.
Lu, A., and Miller, L. K. (1996). Generation of recombinant baculoviruses by direct cloning.
Biotechniques 21, 63–68.
Lu, A., and Miller, L. K. (1997). Regulation of baculovirus late and very late gene
expression. In ‘‘The Baculoviruses’’ (L. K. Miller, ed.), pp. 193–216. Plenum Press,
New York.
Luckow, V. A., Lee, S. C., Barry, G. F., and Olins, P. O. (1993). Efficient generation of
infectious recombinant baculoviruses by site-specific transposon-mediated insertion of
foreign genes into a baculovirus genome propagated in Escherichia coli. J. Virol. 67,
4566–4579.
Lynn, D. E. (2007). Available lepidopteran insect cell lines. Methods Mol. Biol. 388, 117–138.
McLachlin, J. R., and Miller, L. K. (1994). Identification and characterization of vlf-1,
a baculovirus gene involved in very late gene expression. J. Virol. 68, 7746–7756.
Miller, L. K. (1997). The Baculoviruses. Plenum Press, New York.
Mistretta, T. A., and Guarino, L. A. (2005). Transcriptional activity of baculovirus very late
factor 1. J. Virol. 79, 1958–1960.
Murhammer, D. W. (2007). Baculovirus and Insect Cell Expression Protocols. Humana
Press, Totowa, NJ.
Murphy, C. I., Lennick, M., Lehar, S. M., Beltz, G. A., and Young, E. (1990). Temporal
expression of HIV-1 envelope proteins in baculovirus-infected insect cells: Implications
for glycosylation and CD4 binding. Gen. Anal. Tech. Appl. 7, 160–171.
Ooi, B. G., Rankin, C., and Miller, L. K. (1989). Downstream sequences augment
transcription from the essential initiation site of a baculovirus polyhedrin gene. J. Mol.
Biol. 210, 721–736.
O’Reilly, D. R., Miller, L. K., and Luckow, V. A. (1992). Baculovirus Expression Vectors.
W. H. Freeman and Company, New York.
Passarelli, A. L., and Guarino, L. A. (2007). Baculovirus late and very late gene regulation.
Curr. Drug Targets 8, 1103–1115.
Pennock, G. D., Shoemaker, C., and Miller, L. K. (1984). Strong and regulated expression
of Escherichia coli beta-galactosidase in insect cells with a baculovirus vector. Mol. Cell.
Biol. 4, 399–406.
Pijlman, G. P., van Schijndel, J. E., and Vlak, J. M. (2003). Spontaneous excision of BAC
vector sequences from bacmid-derived baculovirus expression vectors upon passage in
insect cells. J. Gen. Virol. 84, 2669–2678.
Baculovirus–Insect Cell Expression Systems 221
Possee, R. D., Hitchman, R. B., Richards, K. S., Mann, S. G., Siaterli, E., Nixon, C. P.,
Irving, H., Assenberg, R., Alderton, D., Owens, R. J., and King, L. A. (2008). Genera-
tion of baculovirus vectors for the high-throughput production of proteins in insect cells.
Biotechnol. Bioeng. 101, 1115–1122.
Possee, R. D., and King, L. A. (2007). Baculovirus transfer vectors. Meth. Mol. Biol. 388,
55–76.
Rankl, N. B., Rice, J. W., Gurganus, T. M., Barbee, J. L., and Burns, D. J. (1994). The
production of an active protein kinase C-delta in insect cells is greatly enhanced by the
use of the basic protein promoter. Prot. Exp. Purif. 5, 346–356.
Richardson, C. D. (1995). Baculovirus Expression Protocols. Humana Press, Totowa, NJ.
Seo, N. S., Hollister, J. R., and Jarvis, D. L. (2001). Mammalian glycosyltransferase expres-
sion allows sialoglycoprotein production by baculovirus-infected insect cells. Prot. Exp.
Purif. 22, 234–241.
Shi, X., and Jarvis, D. L. (2007). Protein N-glycosylation in the baculovirus-insect cell
system. Curr. Drug Targets 8, 1116–1125.
Sigma-Aldrich, D. L. (2008). D6192 DiamondBacTM Baculovirus DNA.
Slack, J. M., Kuzio, J., and Faulkner, P. (1995). Characterization of v-cath, a cathepsin L-like
proteinase expressed by the baculovirus Autographa californica multiple nuclear polyhedro-
sis virus. J. Gen. Virol. 76, 1091–1098.
Smith, P. L., Bousfield, G. R., Kumar, S., Fiete, D., and Baenziger, J. U. (1993). Equine
lutropin and chorionic gonadotropin bear oligosaccharides terminating with SO4–4-
GalNAc and Sia alpha 2, 3Gal, respectively. J. Biol. Chem. 268, 795–802.
Smith, G. E., Fraser, M. J., and Summers, M. D. (1983a). Molecular engineering of the
Autographa californica nuclear polyhedrosis virus genome: Deletion mutations within
the polyhedrin gene. J. Virol. 46, 584–593.
Smith, G. E., Summers, M. D., and Fraser, M. J. (1983b). Production of human beta
interferon in insect cells infected with a baculovirus expression vector. Mol. Cell. Biol.
3, 2156–2165.
Sridhar, P., Panda, A. K., Pal, R., Talwar, G. P., and Hasnain, S. E. (1993). Temporal nature
of the promoter and not relative strength determines the expression of an extensively
processed protein in a baculovirus system. FEBS Lett. 315, 282–286.
Summers, M. D. (2006). Milestones leading to the genetic engineering of baculoviruses as
expression vector systems and viral pesticides. Adv. Virus Res. 68, 3–73.
Summers, M. D., and Smith, G. E. (1987). A manual of methods for baculovirus vectors and
insect cell culture procedures. Tx. Agric. Exp. Stn. Bull. No. 1555.
Thomas, C. J., Brown, H. L., Hawes, C. R., Lee, B. Y., Min, M. K., King, L. A., and
Possee, R. D. (1998). Localization of a baculovirus-induced chitinase in the insect cell
endoplasmic reticulum. J. Virol. 72, 10207–10212.
Tomiya, N., Howe, D., Aumiller, J. J., Pathak, M., Park, J., Palter, K., Jarvis, D. L.,
Betenbaugh, M. J., and Lee, Y. C. (2003). Complex-type biantennary N-glycans of
recombinant human transferrin from Trichoplusia ni insect cells expressing mammalian
b1, 4-galactosyltransferase and b1, 2-N-acetylglucosaminyltransferase II. Glycobiology 13,
23–34.
Vaughn, J. L., Goodwin, R. H., Thompkins, G. J., and McCawley, P. (1977). The
establishment of two insect cell lines from the insect Spodoptera frugiperda (Lepidoptera:
Noctuidae). In Vitro 13, 213–217.
Vialard, J., Lalumiere, M., Vernet, T., Briedis, D., Alkhatib, G., Henning, D., Levin, D., and
Richardson, C. (1990). Synthesis of the membrane fusion and hemagglutinin proteins of
measles virus, using a novel baculovirus vector containing the beta-galactosidase gene.
J. Virol. 64, 37–50.
222 Donald L. Jarvis
Vialard, J. E., and Richardson, C. D. (1993). The 1, 629-nucleotide open reading frame
located downstream of the Autographa californica nuclear polyhedrosis virus polyhedrin
gene encodes a nucleocapsid-associated phosphoprotein. J. Virol. 67, 5859–5866.
Walhout, A. J., Temple, G. F., Brasch, M. A., Hartley, J. L., Lorson, M. A., van den
Heuvel, S., and Vidal, M. (2000). GATEWAY recombinational cloning: Application to
the cloning of large numbers of open reading frames or ORFeomes. Methods Enzymol.
328, 575–592.
Wickham, T. J., Davis, T., Granados, R. R., Shuler, M. L., and Wood, H. A. (1992).
Screening of insect cell lines for the production of recombinant proteins and infectious
virus in the baculovirus expression system. Biotechnol. Progr. 8, 391–396.
Williams, G. V., Rohel, D. Z., Kuzio, J., and Faulkner, P. (1989). A cytopathological
investigation of Autographa californica nuclear polyhedrosis virus p10 gene functions
using insetion/deletion mutants. J. Gen. Virol. 70, 187–202.
C H A P T E R F I F T E E N
Contents
1. Introduction 224
2. HEK293 and CHO Cell Lines Commonly Used in TGE Approaches 224
3. Expression Vectors for HEK293 and CHO Cells 226
4. Cultivation of HEK293 Cells and CHO Cell Lines in Suspension 228
5. Transfection Methods 228
5.1. Small-scale transfection by lipofection in six-well plates 229
5.2. Medium scale transfection in shake flasks with
polyethylenimine as transfer reagent 231
5.3. Large-scale transient transfection in the WaveTM Bioreactor
with polyethylenimine as transfer reagent 232
6. Conclusions 234
Acknowledgments 234
References 235
Abstract
The timely availability of recombinant proteins in sufficient quantity and of
validated quality is of utmost importance in driving drug discovery and the
development of low molecular weight compounds, as well as for biotherapeu-
tics. Transient gene expression (TGE) in mammalian cells has emerged as a
promising technology for protein generation over the past decade as TGE meets
all the prerequisites with respect to quantity and quality of the product as well
as cost-effectiveness and speed of the process. Optimized protocols have been
developed for both HEK293 and CHO cell lines which allow protein production
at any desired scale up to >100 l and in milligram to gram quantities. Along with
an overview on current scientific and technological knowledge, detailed proto-
cols for expression of recombinant proteins on small, medium, and large scale
are discussed in the following chapter.
223
224 Sabine Geisse and Cornelia Fux
1. Introduction
Recombinant proteins including antibodies are essential tools for drug
discovery in the early stage and used in a multitude of applications. While
for production of therapeutic proteins criteria such as growth behavior,
clonality, and stability, as well as the productivity of the producer cell line
are of major importance, recombinant protein production for research
purposes is mainly driven by the cost-effectiveness, simplicity, and speed
of the process in conjunction with adequate yields of the product.
Over the past decade efficient transient transfection protocols have
reportedly been developed, which meet and even exceed the above-men-
tioned prerequisites. Transient production of proteins nowadays is predom-
inantly done in HEK293-derived cell lines, yet strong efforts are ongoing to
develop similarly efficient protocols for CHO cells. This bears the advantage
of maintaining the same host cell background and thus product character-
istics throughout the entire process of therapeutic protein development in
R&D. Detailed summaries of current scientific knowledge with respect to
TGE can be found in some recently published reviews (Baldi et al., 2007;
Geisse, 2009; Geisse et al., 2005; Pham et al., 2006).
Used in conjunction
Expression system Features of cell line Cultivation mediuma with plasmid vectors References Comments
HEK.EBNA EBNA-1 Suspension: for pCEP4 Many, citations, for Most commonly used
(HE, 293- transformed example, ExCell (Invitrogen), example in Baldi cell line in literature,
EBNA) HEK293 cell 293 (SAFC pEAK8, pcDNA et al. (2007) and but has not been
line, originally Biosciences), 3.1, pTT (NRC Pham et al. marketed by
marketed by Freestyle 293 Canada) (2006) Invitrogen for
Invitrogen (Invitrogen) several years
HEK.EBNA EBNA-1 Adherent: DMEM Same as above ATCC/Stanford ATCC CRL-10852
(HE, 293- transformed þ 10% FCS þ University
EBNA) HEK293 cell line 400 mg/ml G418,
suspension not
done
293-SFE (293SF- Suspension adapted Suspension: pTT plasmid series Pham et al. (2005) System available upon
3F6, NRC) EBNA-1 Hybridoma license from NRC
transformed 293 serum-free
cell line medium þ 1%
BCS
HEK293 T SV40 T-antigen Adherent: DMEM pCMV/myc/ER Li et al. (2007) and ATCC CRL-11268,
transformed 293 þ 10% FCS, (Invitrogen) þ Pham et al. also available in
cell line suspension not derivatives (2005) many labs
routinely done
293 Freestyle HEK293 wild type Suspension: Freestyle pcDNA 3.1, pCMV Zhang et al. (2008) Cell line has been
(293-F) medium SPORT Manual on Web carefully selected for
(Invitrogen) site of growth in Freestyle
Invitrogen medium and for 293
fectin reagent
HKB-11 (Hybrid Fusion of 293 cell Suspension: Bayer pTAT/TAR vector Cho et al. (2001) System available upon
of Kidney and with B-Cell proprietary (Bayer) license from Bayer
B-cell) lymphoma cell medium Healthcare
line
a
According to our experience, all of these cell lines can be cultivated adherently in Dulbecco’s Modified Eagle medium (DMEM) or in a 1:2 mixture of DMEM/Ham’s F12 medium, supplemented
with 10% fetal calf serum.
226 Sabine Geisse and Cornelia Fux
(Backliwal et al., 2008a; Pham et al., 2006)) have been described as well. In
brief, a striking advantage of a single promoter/cell line combination could
so far not be demonstrated. Moreover, genetic elements active at the
posttranscriptional level, such as an intron splice element or the Wood-
chuck hepatitis virus regulatory element (WPRE) impact significantly on
protein yields by enhancement of mRNA stability, increased nuclear export
rates of mRNA and increased translation rates (Backliwal et al., 2008a; Klein
et al., 2006; Le Hir et al., 2003; Matsumoto et al., 1998). More details on the
design of various expression plasmids and their performance can be found in
the public domain (Backliwal et al., 2008a; Durocher et al., 2002; Meissner
et al., 2001; Pham et al., 2006; Xia et al., 2006) (Fig. 15.1).
It should, however, be mentioned that despite all benefits to gene
expression by specifically tailored expression plasmids featuring additional
enhancing elements, a large sized expression vector impacts negatively the
stability and yield in plasmid preparation. We have observed a three- to
fivefold reduction in plasmid yields when the size of the empty backbone
vector was doubled or tripled. As quite large quantities of plasmid DNA
need to be prepared for large-scale transient expression approaches and the
size of genes/cDNAs to be expressed appears to be steadily increasing, this
point should be taken into consideration in the design of backbone expres-
sion vectors for TGE approaches.
4.00E + 06 80
Viable cells/mL
3.50E + 06 70
3.00E + 06 60
2.50E + 06 50
2.00E + 06 40
1.50E + 06 30
1.00E + 06 20
5.00E + 05 10
0.00E + 00 0
0 50 100 150 200 250
Time (h)
5. Transfection Methods
For lipofection a wealth of formulations with individual, mostly propri-
etary compositions are commercially available, and doubtlessly many of them
are highly efficacious in introducing the plasmid DNA into the cells. Their
drawback is the cost-of-goods—transient transfection beyond the scale of
several hundred milliliters is economically not feasible when using these
reagents. Large-scale transient transfection approaches call for transfection
reagents readily available in bulk quantities with little batch-to-batch varia-
bility to ensure consistency of the process. Calcium–phosphate-mediated
Transient Transfection, Mammalian Cells, HEK293, CHO 229
6. Conclusions
The generation of recombinant proteins by means of transient trans-
fection technologies has reached an impressive state in terms of method
development and yields during the past several years, as reflected in the
wealth of data published on the subject. As yet, it is probably premature to
envisage the replacement of the tedious process of cell line development for
production of biotherapeutics by large-scale transient approaches, even
though remarkable antibody titers of >1 g/l have been reported already
(Backliwal et al., 2008a). Our own observations based on >100 transient
expression trials indicate a huge heterogeneity in expression yields depend-
ing on the gene(s) expressed, ranging from 1 to >170 mg/l for proteins;
antibody titers amount to 20–40 mg/l on average with frequent outliers to
both sides. There is certainly a need for further technical advancement with
respect to construct design, cultivation conditions, and possibly cell line
engineering, but these efforts are more than justified by the overall impres-
sive success of the approach.
ACKNOWLEDGMENTS
We thank our collaborators Agnès Patoux, Thomas Cremer, Mirjam Buchs, Sibylle Bossart,
Stefan Dalcher, and Rainer Uhrhahn for the development of the protocols and the genera-
tion of the experimental data. Special thanks go to Prof. Bertram Opalka, Uniklinikum
Essen-Duesburg, FRG for critically reviewing the manuscript.
Transient Transfection, Mammalian Cells, HEK293, CHO 235
REFERENCES
Backliwal, G., Hildinger, M., Chenuet, S., Wulhfard, S., De Jesus, M., and Wurm, F. M.
(2008a). Rational vector design and multi-pathway modulation of HEK 293E cells yield
recombinant antibody titers exceeding 1 g/l by transient transfection under serum-free
conditions. Nucleic Acids Res. 36, e96.
Backliwal, G., Hildinger, M., Hasija, V., and Wurm, F. M. (2008b). High-density transfec-
tion with HEK-293 cells allows doubling of transient titers and removes need for a priori
DNA complex formation with PEI. Biotechnol. Bioeng. 99, 721–727.
Backliwal, G., Hildinger, M., Kuettel, I., Delegrange, F., Hacker, D. L., and Wurm, F. M.
(2008c). Valproic acid: A viable alternative to sodium butyrate for enhancing protein
expression in mammalian cell cultures. Biotechnol. Bioeng. 101, 182–189.
Baldi, L., Hacker, D. L., Adam, M., and Wurm, F. M. (2007). Recombinant protein
production by large-scale transient gene expression in mammalian cells: State of the art
and future perspectives. Biotechnol. Lett. 29, 677–684.
Batard, P., Jordan, M., and Wurm, F. (2001). Transfer of high copy number plasmid into
mammalian cells by calcium phosphate transfection. Gene 270, 61–68.
Bertschinger, M., Backliwal, G., Schertenleib, A., Jordan, M., Hacker, D. L., and
Wurm, F. M. (2006). Disassembly of polyethylenimine-DNA particles in vitro: Implica-
tions for polyethylenimine-mediated DNA delivery. J. Control Release 116, 96–104.
Boussif, O., Lezoualc’h, F., Zanta, M. A., Mergny, M. D., Scherman, D., Demeneix, B.,
and Behr, J. P. (1995). A versatile vector for gene and oligonucleotide transfer into cells in
culture and in vivo: Polyethylenimine. Proc. Natl. Acad. Sci. USA 92, 7297–7301.
Chenuet, S., Martinet, D., Besuchet-Schmutz, N., Wicht, M., Jaccard, N., Bon, A. C.,
Derouazi, M., Hacker, D. L., Beckmann, J. S., and Wurm, F. M. (2008). Calcium
phosphate transfection generates mammalian recombinant cell lines with higher specific
productivity than polyfection. Biotechnol. Bioeng. 101, 937–945.
Cho, M.-S., Yee, H., Brown, C., Jeang, K.-T., and Cahn, S. (2001). An oriP expression
vector containing the HIV Tat/TAR transactivation axis produces high levels of protein
expression in mammalian cells. Cytotechnology 37, 23–30.
Cho, M. S., Yee, H., and Chan, S. (2002). Establishment of a human somatic hybrid cell line
for recombinant protein production. J. Biomed. Sci. 9, 631–638.
Dang, J. M., and Leong, K. W. (2006). Natural polymers for gene delivery and tissue
engineering. Adv. Drug Deliv. Rev. 58, 487–499.
Dean, D., Dean, B., Muller, S., and Smith, L. (1999). Sequence requirements for plasmid
nuclear import. Exp. Cell Res. 253, 713–722.
Dean, D., Strong, D., and Zimmer, W. (2005). Nuclear entry of nonviral vectors. Gene Ther.
12, 881–890.
Durocher, Y., Perret, S., and Kamen, A. (2002). High-level and high-throughput recombi-
nant protein production by transient transfection of suspension-growing human 293-
EBNA1 cells. Nucleic Acids Res. 30, E9.
Eliyahu, H., Barenholz, Y., and Domb, A. J. (2005). Polymers for DNA delivery. Molecules
10, 34–64.
Galbraith, D. J., Tait, A. S., Racher, A. J., Birch, J. R., and James, D. C. (2006). Control of
culture environment for improved polyethylenimine-mediated transient production of
recombinant monoclonal antibodies by CHO cells. Biotechnol. Prog. 22, 753–762.
Geisse, S. (2009). Reflections on more than 10 years of TGE approaches. Protein Expr. Purif.
64, 99–107.
Geisse, S., Jordan, M., and Wurm, F. (2005). Large-scale transient expression of therapeutic
proteins in mammalian cells. In ‘‘Therapeutic Proteins’’, (C. Smales and D. James, eds.),
Vol. 308, pp. 87–98. Humana Press, Totowa, NJ.
236 Sabine Geisse and Cornelia Fux
Lungwitz, U., Breunig, M., Blunk, T., and Gopferich, A. (2005). Polyethylenimine-based
non-viral gene delivery systems. Eur. J. Pharm. Biopharm. 60, 247–266.
Majors, B. S., Betenbaugh, M. J., and Chiang, G. G. (2007). Links between metabolism and
apoptosis in mammalian cells: applications for anti-apoptosis engineering. Metab. Eng. 9,
317–326.
Makrides, S. C. (1999). Components of vectors for gene transfer and expression in mamma-
lian cells. Protein Expr. Purif. 17, 183–202.
Matsumoto, K., Wassarman, K. M., and Wolffe, A. P. (1998). Nuclear history of a pre-
mRNA determines the translational activity of cytoplasmic mRNA. EMBO J. 17,
2107–2121.
Meissner, P., Pick, H., Kulangara, A., Chatellard, P., Friedrich, K., and Wurm, F. M.
(2001). Transient gene expression: Recombinant protein production with suspension-
adapted HEK293-EBNA cells. Biotechnol. Bioeng. 75, 197–203.
Muller, N., Girard, P., Hacker, D. L., Jordan, M., and Wurm, F. M. (2005). Orbital shaker
technology for the cultivation of mammalian cells in suspension. Biotechnol. Bioeng. 89,
400–406.
Payne, C. K. (2007). Imaging gene delivery with fluorescence microscopy. Nanomed 2,
847–860.
Pham, P. L., Kamen, A., and Durocher, Y. (2006). Large-scale transfection of mammalian
cells for the fast production of recombinant protein. Mol. Biotechnol. 34, 225–237.
Pham, P. L., Perret, S., Cass, B., Carpentier, E., St-Laurent, G., Bisson, L., Kamen, A., and
Durocher, Y. (2005). Transient gene expression in HEK293 cells: Peptone addition
posttransfection improves recombinant protein synthesis. Biotechnol. Bioeng. 90, 332–344.
Schlaeger, E.-J., and Christensen, K. (1999). Transient gene expression in mammalian cells
grown in serum-free suspension culture. Cytotechnology 30, 71–83.
Shaw, G., Morse, S., Ararat, M., and Graham, F. L. (2002). Preferential transformation
of human neuronal cells by human adenoviruses and the origin of HEK 293 cells.
FASEB J. 16, 869–871.
Silla, T., Hääl, I., Geimanen, J., Janikson, K., Abroi, A., Ustav, E., and Ustav, M. (2005).
Episomal maintenance of plasmids with hybrid origins in mouse cells. J.Virol. 79,
15277–15288.
Stettler, M., Zhang, X., Hacker, D. L., de Jesus, M., and Wurm, F. M. (2007). Novel orbital
shake bioreactors for transient production of CHO derived IgGs. Biotechnol. Prog. 23,
1340–1346.
Sun, X., Goh, P. E., Wong, K. T., Mori, T., and Yap, M. G. (2006). Enhancement of
transient gene expression by fed-batch culture of HEK 293 EBNA1 cells in suspension.
Biotechnol. Lett. 28, 843–848.
Tait, A. S., Brown, C. J., Galbraith, D. J., Hines, M. J., Hoare, M., Birch, J. R., and
James, D. C. (2004). Transient production of recombinant proteins by Chinese hamster
ovary cells using polyethyleneimine/DNA complexes in combination with microtubule
disrupting anti-mitotic agents. Biotechnol. Bioeng. 88, 707–721.
Tigges, M., and Fussenegger, M. (2006). Xbp1-based engineering of secretory capacity
enhances the productivity of Chinese hamster ovary cells. Metab. Eng. 8, 264–272.
Urlaub, G., and Chasin, L. A. (1980). Isolation of Chinese hamster cell mutants deficient in
dihydrofolate reductase activity. Proc. Natl. Acad. Sci. 77, 4216–4220.
Urlaub, G., Käs, E., Carothers, A. M., and Chasin, L. A. (1983). Deletion of the diploid
dihydrofolate reductase locus from cultured mammalian cells. Cell 33, 405–412.
Van Craenenbroeck, K., Vanhoenacker, P., and Haegeman, G. (2000). Episomal vectors for
gene expression in mammalian cells. Eur. J. Biochem. 267, 5665–5678.
Vicennati, P., Giuliano, A., Ortaggi, G., and Masotti, A. (2008). Polyethylenimine in
medicinal chemistry. Curr. Med. Chem. 15, 2826–2839.
238 Sabine Geisse and Cornelia Fux
Xia, W., Bringmann, P., McClary, J., Jones, P. P., Manzana, W., Zhu, Y., Wang, S.,
Liu, Y., Harvey, S., Madlansacay, M. R., McLean, K., Rosser, M. P., et al. (2006). High
levels of protein expression using different mammalian CMV promoters in several cell
lines. Protein Expr. Purif. 45, 115–124.
Zhang, J., Liu, X., Bell, A., To, R., Baral, T. N., Azizi, A., Li, J., Cass, B., and Durocher, Y.
(2009). Transient expression and purification of chimeric heavy chain antibodies. Protein
Expr. Purif. 65, 77–82.
C H A P T E R S I X T E E N
Contents
1. Introduction 240
2. Some Considerations When Designing a Tagged Protein 241
2.1. Affinity and/or solubility? 241
2.2. Which tag(s) to use? 243
2.3. Tandem tags? 244
2.4. N- or C-terminal? 244
2.5. Cleavage sites to remove tags 245
3. Protein Affinity Tags 245
3.1. His-tag 245
3.2. GST tag 246
3.3. Other purification tags 247
4. Solubility Tags 249
4.1. MBP tag 249
4.2. Trx tag 250
4.3. NusA tag 250
4.4. Other solubility tags 250
5. Removal of Tags 251
6. Conclusions 253
Acknowledgment 254
References 254
Abstract
Tags are frequently used in the expression of recombinant proteins to improve
solubility and for affinity purification. A large number of tags have been devel-
oped for protein production and researchers face a profusion of choices when
designing expression constructs. Here, we survey common affinity and solubility
tags, and offer some guidance on their selection and use.
Department of Biochemistry and Molecular Biology, University of Miami Miller School of Medicine, Miami,
Florida, USA
239
240 Arun Malhotra
1. Introduction
Most proteins are made using recombinant techniques in expression
systems. Additional residues or tags can be engineered on either the N- or
C-terminal end of the protein of interest during the cloning step. These
tags, which can range in size from just a few residues to full-length proteins
or domains, can be used to improve protein production or to confer new
properties that can be used for characterization and study of the target
protein. The term ‘‘fusion protein’’ is also often used instead of the term
‘‘tag,’’ and fusion sometimes refers to the simpler end-to-end joining of two
proteins while tags are typically shorter and include linker regions; here, we
will use these terms interchangeably.
The focus of this chapter is to survey some of the common tags that are used
for improving the production of proteins, and highlight the advantages and
pitfalls of their use. Given the wild profusion of tags, often in different
combinations and with different cleavage sites, this review is by no means
exhaustive and the reader is encouraged to look at constructs available from
many commercial sources as well as repository sources such as Addgene
(http://www.addgene.org) or ATCC (http://www.atcc.org). Detailed proto-
cols related to the use of many individual tags have been described previously
in Methods in Enzymology (e.g., Volumes 326 and 327), and in specialized
journals in this field such as Protein Expression and Purification and the newer
Microbial Cell Factories. Handbooks from suppliers of vectors for most expres-
sion systems (Amersham/GE Healthcare Inc., Clontech Inc., Invitrogen Inc.,
New England Biolabs Inc., Novagen/EMD Biosystems Inc., Roche Inc.,
Sigma-Aldrich Inc., and others) are another rich source of protocols.
Tags used to improve the production of recombinant proteins can be
roughly divided into purification and solubility tags. The former are used
along with affinity binding to allow rapid and efficient purification of
proteins, while the latter refer to tags that enhance the proper folding and
solubility of a protein. Tables 16.1 and 16.2 list some of the common
purification and solubility tags that are used for protein expression.
Table 16.3 summarizes common endoproteases used to remove tags and
recover the target protein of interest.
While such tags are quite useful, other tags can be fused to proteins for a
broad range of applications—labeling for imaging and localization studies,
protein detection and quantification, protein–protein interaction studies,
subcellular localization or transduction, and many others. It is important to
keep in mind some of these additional capabilities that can be engineered
into a protein as recombinant constructs are being designed, since multiple
tags can be added together in different combinations (Fig. 16.1).
Tagging for Protein Expression 241
A
TEV cleavage site
N Target protein C
His-tag
B
Thrombin cleavage site
C
TEV cleavage site
Figure 16.1 Example schematics of single and tandem tagged proteins. (A) N-terminally
His-tagged protein; (B) C-terminally GST tagged protein; and (C) target protein with a
removable (using TVMV proteases) C-terminal tandem affinity tag (TAP tag).
and has also the added benefit of increasing the solubility of the target
protein; this is the basis of solubility tags (Table 16.2).
Affinity tags, on the other hand, are crucial during protein purification,
and allow the use of a variety of strategies to bind the target protein on an
affinity matrix (Table 16.1). Some protein tags can function in both affinity
and solubility roles—for example, the glutathione-S-transferase (GST) tag
improves solubility of some proteins, while solubility tags such as maltose-
binding protein (MBP) can be used for affinity purification of the target
protein.
2.4. N- or C-terminal?
Tags can be placed at either the N- or C-terminus of a target protein. One
advantage of placing a tag on the N-terminal end is that the construct can
take advantage of efficient translation initiation sites on the tag. Solubility
tags based on highly expressing proteins such as MBP, Trx, and NusA are
also more efficient at solubilizing target proteins when positioned at the
N-terminal end (Sachdev and Chirgwin, 1998), though recent high-
throughput studies have shown than the MBP tag is still quite effective
when positioned at the C-terminal end (Dyson et al., 2004). Another
advantage of placing a tag on the N-terminal site is that the tag can
be removed more cleanly, since most endoproteases cut at or near the
C-terminus of their recognition sites.
While placing a tag, care should be taken to preserve the positioning of
any signal sequences or modification sites. Sequences at termini of the fusion
protein should be examined for effects on the stability of the final construct,
Tagging for Protein Expression 245
especially at the N-terminal end, which should be inspected for the host
cell’s N-end rule degradation signals (Bachmair et al., 1986; Wang et al.,
2008). It is also useful to examine the sequence of the tagged protein for any
inadvertently created interaction or cleavage sites using motif databases such
as PROSITE (Hulo et al., 2008).
3.3.3. STREP-II-tag
STREP-II-tag (WSHPQFEK) is a tag that takes advantage of the strong and
specific interaction between biotin and streptavidin (Schmidt and Skerra,
1994). This peptide tag binds in the biotin pocket of streptavidin;
Strep-Tactin, a recombinant form of streptavidin optimized to bind the
Strep-II-tag, is used in affinity media. Bound target protein can be eluted
with low levels (2.5 mM ) of desthiobiotin, a biotin analog that competes for
binding to Strep-Tactin in a reversible manner (Skerra and Schmidt, 2000).
The elution conditions are gentle, and buffers can include high levels (up to
50 mM ) of reducing agents such as DTT or b-mercaptoethanol, as well as
chelating agents such as EDTA, which makes this an excellent purification
technique for proteins that are very sensitive to oxidation. Strep-Tactin
chromatographic resins can regenerated and reused a few times. While there
is a low background of naturally biotinylated proteins in most cell extracts,
these usually do not interfere in the purification; if necessary, avidin can be
added to clear biotinylated host proteins (Schmidt and Skerra, 2007). There
has also been some work on using streptavidin/avidin in fusion tags (Sano and
Cantor, 2000), but they are difficult to use for affinity purification because the
very strong interaction between these proteins is difficult to disrupt.
3.3.4. CBP-tag
The calmodulin-binding peptide (CBP) is a short 26-residue sequence
derived from the C-terminus of skeletal muscle myosin light chain kinase
that binds specifically to calmodulin (reviewed in Terpe, 2003). Calmodulin
(CaM) immobilized on chromatographic media (such as CaM-Sepharose)
can be used for affinity purification of target proteins tagged with CBP.
Though calmodulin binds very strongly with CBP (nanomolar affinity), this
interaction is dependent on calcium and the bound protein can be eluted in
a single step using gentle buffers containing a calcium-chelating agent, such
Tagging for Protein Expression 249
4. Solubility Tags
Production of well-folded, soluble proteins is a major bottleneck in
recombinant protein expression, especially when heterologous eukaryotic
proteins are expressed in bacterial cells. Several soluble proteins are used as
tags to improve folding of the target protein. These tags should be used in
conjunction with other approaches to improve protein folding such as
lowering temperature after protein induction or coexpression of chaperones
(Baneyx and Mujacic, 2004; de Marco et al., 2007; Sahdev et al., 2008). It is
also useful to screen multiple solubility tags (Peleg and Unger, 2008). For
recalcitrant proteins, protein purification using denaturing conditions and
refolding can be tried (Cabrita and Bottomley, 2004; Jungbauer and Kaar,
2007; Qoronfleh et al., 2007; see Chapter 17 in this volume).
Zhang et al., 2004). Other small tags are the GB1-tag (56 residues) which is
based on the IgG binding B1 domain of the streptococcal Protein G (Cheng
and Patel, 2004; Zhou et al., 2001), as well as the IgG binding domain of
Protein A (ZZ domain, 116 residues; Inouye and Sahara, 2009; Rondahl et al.,
1992). These small tags are especially useful in protein preparations for NMR
studies (Kato et al., 2007), and some progress has been made in the creation of
NMR-invisible solubility tags using protein ligation methods (Durst et al.,
2008; Kobashigawa et al., 2009).
Small ubiquitin-like modifier (SUMO) protein ( 11 kDa) has
been shown to significantly improve protein stability and solubility as an
N-terminal fusion (Marblestone et al., 2006). The tag can be removed after
purification using SUMO protease (catalytic domain of Ulp1) that recog-
nizes the SUMO structure (Lee et al., 2008; Panavas et al., 2009). The
SUMO-fusion system has also been adapted for use in insect cells and other
eukaryotic expression systems (Liu et al., 2008).
HaloTag is a recently created modular tagging system that uses a 34-kDa
modified haloalkane dehalogenase protein that can bind a variety of syn-
thetic ligands (HaloTag ligands; Promega Inc.). These ligands are comprised
of a constant reactive linker that binds covalently to the HaloTag, and a
variable reporter end that can impart a variety of useful properties to the
fusion protein. Thus, a single tag can be used for subcellular imaging within
live cells, cell labeling and sorting, affinity purification, or even immobili-
zation on solid supports (Los et al., 2008). Ohana et al. (2009) have used a
version of this tag (HaloTag7) for affinity purification, using the chloroalk-
ane linker attached to agarose beads. Since the HaloTag binds covalently to
this linker in a very specific nonreversible manner, even poorly expressed
proteins can be bound efficiently to the chloroalkane resin. The target
protein can then be eluted off the resin using tobacco etch virus (TEV)
protease that cuts at a cleavage site engineered between the HaloTag and the
target protein. Surprisingly, the monomeric compact HaloTag was seen to
dramatically improve fusion protein solubility when tested on a panel of
difficult-to-express recombinant human proteins being expressed in E. coli
(Ohana et al., 2009). In these tests, the HaloTag fared significantly better
than MBP, and appears to function as a bona fide solubility tag.
5. Removal of Tags
After a protein tag has been used for solubility enhancement or affinity
purification, it is often useful to remove it for biological and functional
studies since the tag can potentially interfere with the proper functioning of
the target protein. This is especially true for large tags such GST or MBP,
252 Arun Malhotra
though there are some examples where the fusion protein was more
amenable for crystallization (Smyth et al., 2003).
Most commercial expression vectors that are used to add tags on target
proteins also include cleavage sites with specific sequences that allow the tag
to be removed using recombinant endoproteases. After the initial affinity
purification step, the sample can be treated with the endoprotease to cleave
off the tag, which can subsequently be separated from the target protein by
passing the sample back on the affinity column and collecting the flow-
through. The recombinant endoprotease usually also comes with an affinity
tag, allowing for its easy removal after the cleavage reaction.
Some commonly used endoproteases are listed in Table 16.3. Enteroki-
nase and Factor Xa are useful for removal of N-terminal tags since they cut
at the C-terminal end of their recognition sequence, allowing for the
complete tag and recognition sequence to be removed. However, both
these enzymes are somewhat promiscuous and can cleave at secondary sites,
often at other basic residues. Similar secondary cleavage sites have also been
seen for thrombin, another protease used to remove tags ( Jenny et al., 2003;
Liew et al., 2005). One advantage with proteases such as thrombin, espe-
cially for large-scale protein production, is that it is cheaper and more
efficient than other more specific proteases.
The PreScission protease is a more specific protease with a longer and
stricter recognition sequence. This is the protease 3C from human rhinovi-
rus-14 (3Cpro) with a GST tag, which makes it especially useful for removing
GST tags. Another very specific and popular protease is the TEV protease
(Kapust et al., 2001). The TEV protease prefers a Gly after the cleavage site
(Table 16.3), but can tolerate other residues with only a modest decrease in
activity, which allows it to cleave N-terminal tags with no additional residues
remaining on the target protein in many cases (Kapust et al., 2002). The TEV
protease is easy to produce in-house in large quantities and is usually expressed
with a His-tag for convenient purification and removal after the cleavage
reaction (Tropea et al., 2009). Many forms of TEV proteases, with other
affinity tags or enhanced for higher activity and stability, are sold commercially.
A variant of the TEV protease is the tobacco vein mottling virus
(TVMV) protease that recognizes a different sequence (Nallamsetty et al.,
2004), and can be used to design separate cleavage sites between different
tags (Fig. 16.1C).
Exoproteases can also be used to remove N-terminal tags, such as the
TAGzyme system (Arnau et al., 2006a, available commercially from Qiagen
Inc.). This approach uses dipeptide aminopeptidase I (DAPase) to sequen-
tially chew up the N-terminal tag until a dipeptide ‘‘stop point’’ in the
sequence is reached. Additional variations of this system have been designed
to work with a variety of sequences (Arnau et al., 2006b, 2008).
Complete removal of C-terminal tags is more problematic, since most
endoproteases cut toward the C-terminal end of their recognition sequence.
Tagging for Protein Expression 253
6. Conclusions
A large variety of protein tags are available for facilitating the soluble
expression and purification of recombinant proteins. However, even with
this wide arsenal of tags, structural genomics protein production facilities are
getting soluble purified proteins with success rates of less than 50%
(Structural Genomics Consortium et al., 2008). While many good strategies
254 Arun Malhotra
have emerged from these large-scale studies, given the diversity in how
proteins fold and their distinct biochemical characteristics, there is no
common set of solubility or affinity tags that works for all proteins. Rather,
the choice of tags largely depends on the protein being expressed and the
task at hand. Herein, we have surveyed the most effective protein expres-
sion tags and issues related to their use.
ACKNOWLEDGMENT
The author is supported in part by a grant (R01-GM69972) from the National Institutes
of Health.
REFERENCES
Arnau, J., Lauritzen, C., and Pedersen, J. (2006a). Cloning strategy, production and purifi-
cation of proteins with exopeptidase-cleavable His-tags. Nat. Protoc. 1, 2326–2333.
Arnau, J., Lauritzen, C., Petersen, G. E., and Pedersen, J. (2006b). Current strategies for the
use of affinity tags and tag removal for the purification of recombinant proteins. Protein
Expr. Purif. 48, 1–13.
Arnau, J., Lauritzen, C., Petersen, G. E., and Pedersen, J. (2008). The use of TAGZyme for
the efficient removal of N-terminal His-tags. Methods Mol. Biol. 421, 229–243.
Bachmair, A., Finley, D., and Varshavsky, A. (1986). In vivo half-life of a protein is a function
of its amino-terminal residue. Science 234, 179–186.
Baneyx, F., and Mujacic, M. (2004). Recombinant protein folding and misfolding in
Escherichia coli. Nat. Biotechnol. 22, 1399–1408.
Bauer, A., and Kuster, B. (2003). Affinity purification-mass spectrometry. Powerful tools for
the characterization of protein complexes. Eur. J. Biochem. 270, 570–578.
Bayer, M. E., Bayer, M. H., Lunn, C. A., and Pigiet, V. (1987). Association of thioredoxin
with the inner membrane and adhesion sites in Escherichia coli. J. Bacteriol.
169, 2659–2666.
Bolanos-Garcia, V. M., and Davies, O. R. (2006). Structural analysis and classification of
native proteins from E. coli commonly co-purified by immobilised metal affinity chro-
matography. Biochim. Biophys. Acta 1760, 1304–1313.
Busso, D., Delagoutte-Busso, B., and Moras, D. (2005). Construction of a set Gateway-
based destination vectors for high-throughput cloning and expression screening in
Escherichia coli. Anal. Biochem. 343, 313–321.
Butt, T. R., Edavettal, S. C., Hall, J. P., and Mattern, M. R. (2005). SUMO fusion
technology for difficult-to-express proteins. Protein Expr. Purif. 43, 1–9.
Cabrita, L. D., and Bottomley, S. P. (2004). Protein expression and refolding—A practical
guide to getting the most out of inclusion bodies. Biotechnol. Annu. Rev. 10, 31–50.
Cabrita, L. D., Dai, W., and Bottomley, S. P. (2006). A family of E. coli expression vectors
for laboratory scale and high throughput soluble protein production. BMC Biotechnol.
6, 12.
Carson, M., Johnson, D. H., McDonald, H., Brouillette, C., and Delucas, L. J. (2007).
His-tag impact on structure. Acta Crystallogr. D Biol. Crystallogr. 63, 295–301.
Cheng, Y., and Patel, D. J. (2004). An efficient system for small protein expression and
refolding. Biochem. Biophys. Res. Commun. 317, 401–405.
Tagging for Protein Expression 255
Chong, S., Mersha, F. B., Comb, D. G., Scott, M. E., Landry, D., Vence, L. M.,
Perler, F. B., Benner, J., Kucera, R. B., Hirvonen, C. A., et al. (1997). Single-column
purification of free recombinant proteins using a self-cleavable affinity tag derived from a
protein splicing element. Gene 192, 271–281.
Collins, M. O., and Choudhary, J. S. (2008). Mapping multiprotein complexes by affinity
purification and mass spectrometry. Curr. Opin. Biotechnol. 19, 324–330.
Cunningham, F., and Deber, C. M. (2007). Optimizing synthesis and expression of trans-
membrane peptides and proteins. Methods 41, 370–380.
Davis, G. D., Elisee, C., Newham, D. M., and Harrison, R. G. (1999). New fusion protein
systems designed to give soluble expression in Escherichia coli. Biotechnol. Bioeng. 65, 382–388.
de Marco, A. (2006). Two-step metal affinity purification of double-tagged (NusA-His6)
fusion proteins. Nat. Protoc. 1, 1538–1543.
De Marco, V., Stier, G., Blandin, S., and de Marco, A. (2004). The solubility and stability of
recombinant proteins are increased by their fusion to NusA. Biochem. Biophys. Res.
Commun. 322, 766–771.
de Marco, A., Deuerling, E., Mogk, A., Tomoyasu, T., and Bukau, B. (2007). Chaperone-
based procedure to increase yields of soluble recombinant proteins produced in E. coli.
BMC Biotechnol. 7, 32.
di Guan, C., Li, P., Riggs, P. D., and Inouye, H. (1988). Vectors that facilitate the expression
and purification of foreign peptides in Escherichia coli by fusion to maltose-binding
protein. Gene 67, 21–30.
Döbeli, H., Andres, H., Breyer, N., Draeger, N., Sizmann, D., Zuber, M. T., Weinert, B.,
and Wipf, B. (1998). Recombinant fusion proteins for the industrial production of
disulfide bridge containing peptides: Purification, oxidation without concatemer forma-
tion, and selective cleavage. Protein Expr. Purif. 12, 404–414.
Durst, F. G., Ou, H. D., Löhr, F., Dötsch, V., and Straub, W. E. (2008). The better tag
remains unseen. J. Am. Chem. Soc. 130, 14932–14933.
Dyson, M. R., Shadbolt, S. P., Vincent, K. J., Perera, R. L., and McCafferty, J. (2004).
Production of soluble mammalian proteins in Escherichia coli: Identification of protein
features that correlate with successful expression. BMC Biotechnol. 4, 32.
Einhauer, A., and Jungbauer, A. (2001). The FLAG peptide, a versatile fusion tag for the
purification of recombinant proteins. J. Biochem. Biophys. Methods 49, 455–465.
Esposito, D., and Chatterjee, D. K. (2006). Enhancement of soluble protein expression
through the use of fusion tags. Curr. Opin. Biotechnol. 17, 353–358.
Fairlie, W. D., Uboldi, A. D., De Souza, D. P., Hemmings, G. J., Nicola, N. A., and
Baca, M. (2002). A fusion protein system for the recombinant production of short
disulfide-containing peptides. Protein Expr. Purif. 26, 171–178.
Fritze, C. E., and Anderson, T. R. (2000). Epitope tagging: General method for tracking
recombinant proteins. Methods Enzymol. 327, 3–16.
Gillies, A. R., Hsii, J. F., Oak, S., and Wood, D. W. (2008). Rapid cloning and purification
of proteins: Gateway vectors for protein purification by self-cleaving tags. Biotechnol.
Bioeng. 101, 229–240.
Habig, W. H., Pabst, M. J., and Jakoby, W. B. (1974). Glutathione S-transferases. The first
enzymatic step in mercapturic acid formation. J. Biol. Chem. 249, 7130–7139.
Hernan, R., Heuermann, K., and Brizzard, B. (2000). Multiple epitope tagging of expressed
proteins for enhanced detection. Biotechniques 28, 789–793.
Hu, J., Qin, H., Sharma, M., Cross, T. A., and Gao, F. P. (2008). Chemical cleavage of
fusion proteins for high-level production of transmembrane peptides and protein
domains containing conserved methionines. Biochim. Biophys. Acta 1778, 1060–1066.
Hulo, N., Bairoch, A., Bulliard, V., Cerutti, L., Cuche, B. A., de Castro, E., Lachaize, C.,
Langendijk-Genevaux, P. S., and Sigrist, C. J. (2008). The 20 years of PROSITE. Nucleic
Acids Res. 36, D245–D249.
256 Arun Malhotra
Inouye, S., and Sahara, Y. (2009). Expression and purification of the calcium binding
photoprotein mitrocomin using ZZ-domain as a soluble partner in E. coli cells. Protein
Expr. Purif. 66, 52–57.
Jenny, R. J., Mann, K. G., and Lundblad, R. L. (2003). A critical review of the methods for
cleavage of fusion proteins with thrombin and factor Xa. Protein Expr. Purif. 31, 1–11.
Jungbauer, A., and Kaar, W. (2007). Current status of technical protein refolding.
J. Biotechnol. 128, 587–596.
Kaplan, W., Hüsler, P., Klump, H., Erhardt, J., Sluis-Cremer, N., and Dirr, H. (1997).
Conformational stability of pGEX-expressed Schistosoma japonicum glutathione S-trans-
ferase: A detoxification enzyme and fusion-protein affinity tag. Protein Sci. 6, 399–406.
Kapust, R. B., and Waugh, D. S. (1999). Escherichia coli maltose-binding protein is uncom-
monly effective at promoting the solubility of polypeptides to which it is fused. Protein
Sci. 8, 1668–1674.
Kapust, R. B., Tözsér, J., Fox, J. D., Anderson, D. E., Cherry, S., Copeland, T. D., and
Waugh, D. S. (2001). Tobacco etch virus protease: Mechanism of autolysis and rational
design of stable mutants with wild-type catalytic proficiency. Protein Eng. 14, 993–1000.
Kapust, R. B., Tözsér, J., Copeland, T. D., and Waugh, D. S. (2002). The P10 specificity of
tobacco etch virus protease. Biochem. Biophys. Res. Commun. 294, 949–955.
Kataeva, I., Chang, J., Xu, H., Luan, C. H., Zhou, J., Uversky, V. N., Lin, D., Horanyi, P.,
Liu, Z. J., Ljungdahl, L. G., et al. (2005). Improving solubility of Shewanella oneidensis
MR-1 and Clostridium thermocellum JW-20 proteins expressed into Escherichia coli.
J. Proteome Res. 4, 1942–1951.
Kato, A., Maki, K., Ebina, T., Kuwajima, K., Soda, K., and Kuroda, Y. (2007). Mutational
analysis of protein solubility enhancement using short peptide tags. Biopolymers 85, 12–18.
Kenig, M., Peternel, S., Gaberc-Porekar, V., and Menart, V. (2006). Influence of the protein
oligomericity on final yield after affinity tag removal in purification of recombinant
proteins. J. Chromatogr. A 1101, 293–306.
Kim, S., and Lee, S. B. (2008). Soluble expression of archaeal proteins in Escherichia coli by
using fusion-partners. Protein Expr. Purif. 62, 116–119.
Kobashigawa, Y., Kumeta, H., Ogura, K., and Inagaki, F. (2009). Attachment of an NMR-
invisible solubility enhancement tag using a sortase-mediated protein ligation method.
J. Biomol. NMR 43, 145–150.
LaVallie, E. R., DiBlasio, E. A., Kovacic, S., Grant, K. L., Schendel, P. F., and McCoy, J. M.
(1993). A thioredoxin gene fusion expression system that circumvents inclusion body
formation in the E. coli cytoplasm. Biotechnology (NY) 11, 187–193.
LaVallie, E. R., Lu, Z., Diblasio-Smith, E. A., Collins-Racie, L. A., and McCoy, J. M.
(2000). Thioredoxin as a fusion partner for production of soluble recombinant proteins in
Escherichia coli. Methods Enzymol. 326, 322–340.
Lee, C. D., Sun, H. C., Hu, S. M., Chiu, C. F., Homhuan, A., Liang, S. M., Leng, C. H.,
and Wang, T. F. (2008). An improved SUMO fusion protein system for effective
production of native proteins. Protein Sci. 17, 1241–1248.
Lichty, J. J., Malecki, J. L., Agnew, H. D., Michelson-Horowitz, D. J., and Tan, S. (2005).
Comparison of affinity tags for protein purification. Protein Expr. Purif. 41, 98–105.
Liew, O. W., Ching Chong, J. P., Yandle, T. G., and Brennan, S. O. (2005). Preparation of
recombinant thioredoxin fused N-terminal proCNP: Analysis of enterokinase cleavage
products reveals new enterokinase cleavage sites. Protein Expr. Purif. 41, 332–340.
Listwan, P., Terwilliger, T. C., and Waldo, G. S. (2009). Automated, high-throughput
platform for protein solubility screening using a split-GFP system. J. Struct. Funct.
Genomics 10, 47–55.
Liu, J. W., Boucher, Y., Stokes, H. W., and Ollis, D. L. (2006). Improving protein
solubility: The use of the Escherichia coli dihydrofolate reductase gene as a fusion reporter.
Protein Expr. Purif. 47, 258–263.
Tagging for Protein Expression 257
Liu, L., Spurrier, J., Butt, T. R., and Strickler, J. E. (2008). Enhanced protein expression in
the baculovirus/insect cell system using engineered SUMO fusions. Protein Expr. Purif.
62, 21–28.
Los, G. V., Encell, L. P., McDougall, M. G., Hartzell, D. D., Karassina, N., Zimprich, C.,
Wood, M. G., Learish, R., Ohana, R. F., Urh, M., et al. (2008). HaloTag: A novel protein
labeling technology for cell imaging and protein analysis. ACS Chem. Biol. 3, 373–382.
Malakhov, M. P., Mattern, M. R., Malakhova, O. A., Drinker, M., Weeks, S. D., and
Butt, T. R. (2004). SUMO fusions and SUMO-specific protease for efficient expression
and purification of proteins. J. Struct. Funct. Genomics 5, 75–86.
Marblestone, J. G., Edavettal, S. C., Lim, Y., Lim, P., Zuo, X., and Butt, T. R. (2006).
Comparison of SUMO fusion technology with traditional gene fusion systems: Enhanced
expression and solubility with SUMO. Protein Sci. 15, 182–189.
Mueller, U., Büssow, K., Diehl, A., Bartl, F. J., Niesen, F. H., Nyarsik, L., and
Heinemann, U. (2003). Rapid purification and crystal structure analysis of a small protein
carrying two terminal affinity tags. J. Struct. Funct. Genomics 4, 217–225.
Nallamsetty, S., and Waugh, D. S. (2006). Solubility-enhancing proteins MBP and NusA
play a passive role in the folding of their fusion partners. Protein Expr. Purif. 45, 175–182.
Nallamsetty, S., Kapust, R. B., Tözsér, J., Cherry, S., Tropea, J. E., Copeland, T. D., and
Waugh, D. S. (2004). Efficient site-specific processing of fusion proteins by tobacco vein
mottling virus protease in vivo and in vitro. Protein Expr. Purif. 38, 108–115.
Ohana, R. F., Encell, L. P., Zhao, K., Simpson, D., Slater, M. R., Urh, M., and
Wood, K. V. (2009). HaloTag7: A genetically engineered tag that enhances bacterial
expression of soluble proteins and improves protein purification. Protein Expr. Purif.
68, 110–120.
Panavas, T., Sanders, C., and Butt, T. R. (2009). SUMO fusion technology for enhanced
protein production in prokaryotic and eukaryotic expression systems. Methods Mol. Biol.
497, 303–317.
Pédelacq, J. D., Piltch, E., Liong, E. C., Berendzen, J., Kim, C. Y., Rho, B. S., Park, M. S.,
Terwilliger, T. C., and Waldo, G. S. (2002). Engineering soluble proteins for structural
genomics. Nat. Biotechnol. 20, 927–932.
Peleg, Y., and Unger, T. (2008). Application of high-throughput methodologies to the
expression of recombinant proteins in E. coli. Methods Mol. Biol. 426, 197–208.
Potts, J. T., Young, D. M., and Anfinsen, C. B. (1963). Reconstitution of fully active RNase
S by carboxypeptidase-degraded RNase S-peptide. J. Biol. Chem. 238, 2593–2594.
Prickett, K. S., Amberg, D. C., and Hopp, T. P. (1989). A calcium-dependent antibody for
identification and purification of recombinant proteins. Biotechniques 7, 580–589.
Puig, O., Caspary, F., Rigaut, G., Rutz, B., Bouveret, E., Bragado-Nilsson, E., Wilm, M.,
and Séraphin, B. (2001). The tandem affinity purification (TAP) method: A general
procedure of protein complex purification. Methods 24, 218–229.
Qoronfleh, M. W., Hesterberg, L. K., and Seefeldt, M. B. (2007). Confronting high-
throughput protein refolding using high pressure and solution screens. Protein Expr.
Purif. 55, 209–224.
Raines, R. T., McCormick, M., Van Oosbree, T. R., and Mierendorf, R. C. (2000).
The S.Tag fusion system for protein purification. Methods Enzymol. 326, 362–376.
Richards, F. M., and Vithayarhil, P. J. (1959). The preparation of subtilisin-modified
ribonuclease and the separation of the peptide and protein components. J. Biol. Chem.
234, 1459–1465.
Rondahl, H., Nilsson, B., and Holmgren, E. (1992). Fusions to the 50 end of a gene
encoding a two-domain analogue of staphylococcal protein A. J. Biotechnol. 25, 269–287.
Roodveldt, C., Aharoni, A., and Tawfik, D. S. (2005). Directed evolution of proteins for
heterologous expression and stability. Curr. Opin. Struct. Biol. 15, 50–56.
258 Arun Malhotra
Sachdev, D., and Chirgwin, J. M. (1998). Order of fusions between bacterial and mamma-
lian proteins can determine solubility in Escherichia coli. Biochem. Biophys. Res. Commun.
244, 933–937.
Sahdev, S., Khattar, S. K., and Saini, K. S. (2008). Production of active eukaryotic proteins
through bacterial expression systems: A review of the existing biotechnology strategies.
Mol. Cell. Biochem. 307, 249–264.
Saleh, L., and Perler, F. B. (2006). Protein splicing in cis and in trans. Chem. Rec. 6, 183–193.
Sano, T., and Cantor, C. R. (2000). Streptavidin-containing chimeric proteins: Design and
production. Methods Enzymol. 326, 305–311.
Schmidt, T. G., and Skerra, A. (1994). One-step affinity purification of bacterially produced
proteins by means of the ‘‘Strep tag’’ and immobilized recombinant core streptavidin.
J. Chromatogr. A 676, 337–345.
Schmidt, T. G., and Skerra, A. (2007). The Strep-tag system for one-step purification and
high-affinity detection or capturing of proteins. Nat. Protoc. 2, 1528–1535.
Skerra, A., and Schmidt, T. G. (2000). Use of the Strep-tag and streptavidin for detection
and purification of recombinant proteins. Methods Enzymol. 326, 271–304.
Smith, D. B., and Johnson, K. S. (1988). Single-step purification of polypeptides expressed in
Escherichia coli as fusions with glutathione S-transferase. Gene 67, 31–40.
Smits, S. H., Mueller, A., Grieshaber, M. K., and Schmitt, L. (2008). Coenzyme- and
His-tag-induced crystallization of octopine dehydrogenase. Acta Crystallogr. Sect. F Struct.
Biol. Cryst. Commun. 64, 836–839.
Smyth, D. R., Mrozkiewicz, M. K., McGrath, W. J., Listwan, P., and Kobe, B. (2003).
Crystal structures of fusion proteins with large-affinity tags. Protein Sci. 12, 1313–1322.
Structural Genomics Consortium, China Structural Genomics Consortium, Northeast
Structural Genomics Consortium, Gräslund, S., Nordlund, P., Weigelt, J.,
Hallberg, B. M., Bray, J., Gileadi, O., Knapp, S., et al. (2008). Protein production and
purification. Nat. Methods 5, 135–146.
Terpe, K. (2003). Overview of tag protein fusions: From molecular and biochemical
fundamentals to commercial systems. Appl. Microbiol. Biotechnol. 60, 523–533.
Tolun, A. A., Dickerson, I. M., and Malhotra, A. (2007). Overexpression and purification of
human calcitonin gene-related peptide-receptor component protein in Escherichia coli.
Protein Expr. Purif. 52, 167–174.
Tropea, J. E., Cherry, S., and Waugh, D. S. (2009). Expression and purification of soluble
His(6)-tagged TEV protease. Methods Mol. Biol. 498, 297–307.
Vaillancourt, P., Zheng, C. F., Hoang, D. Q., and Breister, L. (2000). Affinity purification of
recombinant proteins fused to calmodulin or to calmodulin-binding peptides. Methods
Enzymol. 326, 340–362.
Waldo, G. S. (2003). Genetic screens and directed evolution for protein solubility. Curr.
Opin. Chem. Biol. 7, 33–38.
Waldo, G. S., Standish, B. M., Berendzen, J., and Terwilliger, T. C. (1999). Rapid protein-
folding assay using green fluorescent protein. Nat. Biotechnol. 17, 691–695.
Wang, K. H., Oakes, E. S., Sauer, R. T., and Baker, T. A. (2008). Tuning the strength of a
bacterial N-end rule degradation signal. J. Biol. Chem. 283, 24600–24607.
Waugh, D. S. (2005). Making the most of affinity tags. Trends Biotechnol. 23, 316–320.
Yokoyama, S. (2003). Protein expression systems for structural genomics and proteomics.
Curr. Opin. Chem. Biol. 7, 39–43.
Zhang, Y. B., Howitt, J., McCorkle, S., Lawrence, P., Springer, K., and Freimuth, P.
(2004). Protein aggregation during overexpression limited by peptide extensions with
large net negative charge. Protein Expr. Purif. 36, 207–216.
Zhou, P., Lugovskoy, A. A., and Wagner, G. (2001). A solubility-enhancement tag (SET)
for NMR studies of poorly behaving proteins. J. Biomol. NMR 20, 11–14.
C H A P T E R S E V E N T E E N
Contents
1. Introduction 260
2. General Refolding Consideration 262
3. General Procedures 262
4. General Protocol 263
5. Comments on this General Procedure 264
5.1. Overexpression of recombinant proteins in E. coli 264
5.2. Washing IBs 265
5.3. Solubilizing IBs 265
5.4. Refolding 267
5.5. High-resolution ion-exchange chromatography 268
5.6. Reoxidation to form correct disulfide bonds 269
5.7. Characterization 270
6. Performing a Protein Refolding Test Screen 271
6.1. Systematic refolding screens 271
6.2. Variables in refolding 271
6.3. A practical, inexpensive rational approach 274
7. Other Refolding Procedures 275
8. Refolding Database: Refold 277
9. Strategies to Increase Proportion of Soluble Protein 277
10. Conclusion 279
References 279
Abstract
The vast majority of protein purification is now done with cloned, recombinant
proteins expressed in a suitable host. The predominant host is Escherichia coli.
Many, if not most, expressed proteins are found in an insoluble form called an
inclusion body (IB). Since the target protein is often relatively pure in a washed
IB, the challenge is not so much to purify the target, but rather to solubilize an
IB and refold the protein into its native structure, regaining full biological
McArdle Laboratory for Cancer Research, University of Wisconsin-Madison, Madison, Wisconsin, USA
259
260 Richard R. Burgess
activity. While many of the operations of this process are quite general (expres-
sion, cell disruption, IB isolation and washing, and IB solubilization), the precise
conditions that give efficient refolding differ for each protein. This chapter
describes the main techniques and strategies for achieving successful refolding.
1. Introduction
When proteins are expressed at high levels in Escherichia coli and other
expression hosts, it is common to find that most of the protein is not soluble,
but in an insoluble form called an inclusion body (IB). While the exact
mechanism for IB formation is not fully understood and may vary with
different proteins and expression conditions, it is generally thought that
target proteins are being made faster than they can fold into the native
structure. If a protein is partially folded or misfolded, it will generally have
hydrophobic or ‘‘sticky’’ regions exposed that can interact with other
similar proteins and form aggregates. These aggregates form IBs that are
dense, refractile bodies on the order of 0.2–0.5 mm in diameter. Once in an
IB, the protein is usually protected from proteolytic attack and is the
predominant protein. IBs can vary from being mostly native protein,
relatively easily solubilized under mild condition to being misfolded, effec-
tively irreversibly insoluble material that requires high concentrations
of denaturants to be solubilized (see Bowden et al., 1991; Ventura and
Villaverde, 2006). The latter category is by far the most common and
thus, the major barrier to obtaining native, active material is to find condi-
tions under which the denatured protein can be refolded efficiently. This
refolding process and strategies to achieve it are the focus of this chapter.
Anfinsen, in the early 1960s (see Anfinsen, 1973), observed that small
protein molecules could fold spontaneously to the native state. This observa-
tion has led to a thermodynamic hypothesis: the native state has the lowest
free energy. While the thermodynamic hypothesis has become a paradigm of
protein folding, a simple and general model for describing how an unfolded
protein molecule finds its native state is still not available. One cannot yet use
computers to accurately predict protein structure from sequence.
Levinthal (1968) pointed out the paradox that an unfolded protein
molecule does not have the time needed to search all possible conforma-
tions, yet it reaches the single native state rapidly. Consider a (hypothetical)
small protein with 100 residues. Levinthal calculated that, if each residue can
assume only three different conformations (it is probably many more than
this), the total number of structures would be 3100, which is equal to
5 1047. If it takes only 10 13 s to convert one structure into another,
the total search time would be 5 1047 times10 13 s, which is equal to
5 1034 s or 1.6 1027 years! Clearly, it would take much too long for even
Refolding Proteins 261
a small protein to fold properly by randomly trying out all possible con-
formations. This paradox has shaped many studies on the theory of refolding
(e.g., see Clark, 2004; Karplus, 1997).
A great many experimental studies and theoretical analyses have been
carried out to better understand the folding process. The consensus is that
when a denatured protein is no longer in denaturing conditions, some
secondary structure (a-helix and b-sheet) forms, a hydrophobic collapse
occurs to form a ‘‘molten globule’’ state with some secondary structure, but
nonnative tertiary structures, the globular states go through multiple inter-
mediates, and finally reach the lowest energy state, the native conformation.
Solution conditions can have significant effects on the folding pathway of a
given protein. If you think of the refolding energy funnel as a deep, fog-filled
crater in the top of a volcanic mountain, then the rim is the denatured state
and the very bottom of the crater is the native state (the lowest energy state).
If you roll a ball down into the crater from one point on the rim of the crater,
then it may be able to roll all the way to the bottom. From another point on
the rim, the ball may become trapped in a local minimum and not reach the
bottom. In many ways the challenge of finding conditions for efficient
refolding is like trying to guess where on the rim to start rolling the ball.
The rate of protein refolding is rapid, with folding halftimes on the order
of seconds for proteins lacking prolines and on the order of minutes for
proteins with prolines (see Nall, 1994). The amino acid proline can consist
of two isomers, a cis and trans configuration, and the halftime of this
chemical isomerization is on the order of 1 min. In a native protein, each
proline is in a particular configuration, but in a denatured protein an
equilibrium mixture of the two isomers forms consisting of about 70%
trans and 30% cis isomer. Therefore, as a protein refolds, those prolines
that are in the incorrect conformation cannot form the native structure and
must wait for an isomerization event to occur. Thus, proline isomerization
becomes rate limiting for correct protein folding. The enzyme, protein
proline isomerase (PPI), can speed up this isomerization rate.
Even though recombinant proteins were first prepared by refolding
solubilized IBs in the late 1970s and early 1980s, and many improvements
have been made in refolding procedures and techniques, refolding of any
given protein still presents a significant challenge. During the last 10 years
the protein structure initiative (PSI), through its numerous Structural
Genomics Centers, has cloned and expressed over 110,000 proteins (as of
December 2008), but were only able to purify about 29,000 (about 26%).
On the order of 50% of Eubacterial and Archaeal proteins and only about
10% of Eukaryal proteins could be expressed as soluble proteins in E. coli
(Graslund et al., 2008). The proteins that were insoluble were generally
abandoned rather than trying to find effective refolding conditions. I believe
that many of these insoluble proteins could have been salvaged or rescued
by finding suitable refolding conditions as described below.
262 Richard R. Burgess
3. General Procedures
Given that it is very often the case that protein expressed in E. coli and
other hosts is found in IBs, the following steps are almost always needed to
obtain active, native protein:
1. Cell growth and protein overexpression
2. Cell lysis, isolating and washing IBs
3. Solubilizing IBs
4. Identifying suitable refolding conditions
5. Protein refolding
6. Reoxidation to form disulfide bonds where necessary
7. High-resolution ion-exchange chromatography
8. Characterization of final material to determine if refolding has been
successful
Refolding Proteins 263
4. General Protocol
Below is a typical procedure that works well for many proteins. This
protocol is based on and adapted from the protocol first developed
by Nguyen et al. (1993) and used in the Cold Spring Harbor Protein
Purification and Characterization Course section on Purification of Insolu-
ble Recombinant Proteins (Burgess and Knuth, 1996). Other similar
procedures are likely to give similar results, but this is the procedure used
almost exclusively in my laboratory. The critical steps will be discussed in
the following steps.
1. E. coli BL21(DE3)pLysS containing the target gene cloned into a pET
expression vector (Studier et al., 1990) is grown in 1 L of LB in a 2-L
flask at 37 C with shaking until the A600 nm reaches about 0.6.
2. IPTG is added to 0.5–1 mM to induce expression of the T7 RNA
polymerase that then transcribes the target gene.
3. Three to four hours after induction, the cells are harvested by centrifu-
gation at 15,000 rpm for 15 min, the cells are resuspended in a small
volume of the culture supernatant, transferred into a preweighed,
40-ml Oak Ridge tube, and centrifuged to pellet the cells. The wet
weight of the cell pellet is noted and the cells are stored frozen
at 80 C until use. It is common to get 1.5–2.0 g wet weight
E. coli cells per liter using this procedure.
4. Cells are thawed, resuspended in 30 ml of a lysis buffer (50 mM
Tris–HCl, pH 7.9, 0.1 mM EDTA, 5% glycerol, 0.1 mM DTT,
0.1 M NaCl), and then sonicated at 60% power for 3–4 intervals of
20 s each with about 1 min on ice to cool between each interval.
5. Purified Triton X-100 (from a 10%, w/v stock) is added to 1% (w/v) to
dissolve membranes and solubilize membrane proteins. The lysate is
incubated on ice for 10 min and then centrifuged at 15,000 rpm for
15 min to pellet the IB and remove the soluble supernatant.
6. The IB is resuspended in 30 ml of lysis buffer with 1% Triton X-100,
incubated on ice for 10 min, and then centrifuged for 15,000 rpm for
15 min.
7. The drained pellet is resuspended in 30 ml of lysis buffer without Triton
X-100 to remove the Triton X-100 and recentrifuged as above. The pellet
is called the washed IB fraction and is usually over 90% pure.
8. The washed IB pellet is resuspended in a suitable denaturant, incubated,
to allow denaturation and solubilization, and then centrifuged at
15,000 rpm for 15 min to remove any residual insoluble material. We
routinely use the lysis buffer above (lacking the NaCl if diluting from
GuHCl) containing either 6 M guanidine hydrochloride (GuHCl), 8 M
urea, or 0.3% Sarkosyl (n-lauroyl sarcosinate) for IB solubilization.
264 Richard R. Burgess
of the target protein of 10–40% of the total cell protein (Makrides, 1996;
Murby et al., 1996; Sorensen and Mortensen, 2005; Studier et al., 1990)
(see Section 9). Usually one carries out a test to see at what temperature to
grow the cells, how much inducer to use, and how long after induction you
should harvest the cells. After induction of the high-level expression of
the target protein, the cells are usually centrifuged, and frozen at 80 C
until use. There are anecdotal reports that IBs from cells stored for weeks in
the frozen state are harder to refold, but in my experience, we have gotten
excellent refolding results from cells stored frozen for months. Unfortunately,
to my knowledge, no systematic study has addressed this point.
dissociate during the long wash. Experimental analysis of the eluted target
protein indicates essentially no (<1 molecule detergent per 10 molecules of
protein) residual Sarkosyl in the protein (R. Burgess, unpublished). Do not
use MonoS for this purpose since Sarkosyl binds to the column and can
damage the column and contaminate your eluted protein.
SDS can also be used in a somewhat similar fashion, but it is important to
use SDS that is free from higher alkane lengths such as C14, C16, etc.
Refolding experiments using GFP denatured in SDS work quite well
(R. Burgess N. Thompson, and R. Chumanov, unpublished), so SDS
should be considered a viable solubilizing agent.
The successful use of the cationic detergent cetyltrimethylammonium
chloride (CTAC) in solubilization and refolding has also been reported
(Puri et al., 1992). Even distilled water at very low salt has been reported
to achieve solubilization (Song, 2009).
5.4. Refolding
Assuming you have carried out refolding trials as described below, then one
basically dilutes out the denaturant by diluting the solubilized protein into a
suitable refolding buffer. However, there are three ways to dilute during
refolding.
1. Reverse dilution: Add refolding buffer to denatured protein with mixing
between each addition. This method has been successful in several cases
(e.g., Gribskov and Burgess, 1983). However, on reflection, it is clear
that this method of dilution results in the highest concentration of
protein at the critical time when the protein is starting to refold. For
example, if the solubilized protein concentration is 1 mg/ml in 6 M
GuHCl, then if one adds 2–5 volumes of refolding buffer (dilutes to
1–2 M GuHCl), one is likely to go through the refolding transition
between protein concentrations of 330–160 mg/ml.
2. Flash dilution: Add denatured protein to refolding buffer quickly. For
example, add 10 ml of solubilized protein in 6 M GuHCl at one time
with rapid mixing to 590 ml of a suitable refolding buffer to achieve a
60-fold dilution. The protein concentration will be 16 mg/ml during the
refolding, much lower than with reverse dilution and less likely to result
in aggregation.
3. Drip dilution: Add denatured protein to refolding buffer very slowly,
drop-by-drop, over period of 1 h. In theory, this should be the optimal
method, since protein is always refolding at minimal protein concentra-
tion (see Singh and Panda, 2005; Vallejo and Rinas, 2004). For example,
if you add 0.1 ml of solubilized protein as above into 590 ml refolding
buffer, then the protein concentration will only be 0.16 mg/ml. As you
add small incremental amounts of the sample, the protein will always be
268 Richard R. Burgess
low and when the last has been added, the protein concentration will be
16 mg/ml. However, if the denatured protein is added over a period of
1 h, and significant protein refolding to native state is achieved with a
refolding half-life of 1–3 min, then the concentration of partially
refolded sticky protein will always be quite low and the native protein
will not be available to aggregate. While adding solubilized protein over
1 h period is tedious when added drop-by-drop, one can set up a
peristaltic pump at a very slow flow rate to deliver the denatured protein
over 1 h period. This is more reproducible and causes considerably less
graduate student burnout.
Let the protein sit after dilution for 30–60 min and then filter as
described in the generic protocol. If this is not done, then as the protein
solution is loaded on the column (see below) very small particulate material
is likely to clog up the column, causing the pressure to rise, either aborting
the run or requiring one to decrease the flow rate to such low levels that it
takes many hours to load the column.
5.7. Characterization
One of the most common problems I see when reviewing papers involving
refolding is that one has no idea if the material at the end of the procedure is
correctly refolded, monodisperse, or has full biological activity. It is com-
mon to see an enzyme or biological assay, but often without any standard.
You see phrases like ‘‘since the final product has activity it is shown that it is
correctly folded and fully active.’’ Only a comparison with a known, fully
active standard will allow an estimation of the percent of the final product
that is active. Without this comparison you know there is activity, but do
not know if 100%, 10%, 1%, 0.1%, or 0.01% of the refolded material is
active. Such a determination of specific activity of the final product is
essential in a refolding procedure, if possible.
Another important characterization is to determine the size of the
refolded material. Is it monomeric (monodisperse) or does it contain
large amounts of soluble multimer (heterodisperse)? Crystallographers
routinely use dynamic light scattering (DSL) to determine if purified
protein, refolded or not, is monodisperse. Another method, developed
by Mark Knuth at GNF in La Jolla, CA is to analyze the final product by
analytical size exclusion chromatography (ANSEC). We have found this
to be very useful since very small amounts of protein can be analyzed.
Typically, we run 20–50 mg of protein on a 12-ml Shodex KW-802.5
column at 0.5–1.0 ml/min in a buffer containing 0.25 M NaCl, moni-
toring absorbance at 280 and 215 nm. When the column is calibrated
with suitable MW markers, one can quickly determine if most of the
protein is in a single peak (monodisperse) of the size expected for the
monomer (this is good), or if much of the protein elutes earlier indicating
that it is much larger (heterodisperse) due to formation of multimers (this
is bad). Use of this procedure is an important part of a thorough set of
refolding test trials (see below).
One cannot use circular dichroism (CD) or Western blot analysis to
determine activity or monodispersity. There are plenty of examples of
proteins whose CD spectra are very similar or identical to native standard
material and which react just fine in a Western blot analysis that are not
monodisperse or active, respectively.
Refolding Proteins 271
structure will bias refolding toward native and away from aggregated
protein.
Arginine and other amino acids: There is an extensive literature on the use
of arginine to promote refolding of solubilized protein and how it functions
to diminish aggregation (Arakawa et al., 2007; Baynes et al., 2005; Das et al.,
2007; Dong et al., 2004; Reddy et al., 2005). It appears that arginine
decreases aggregation by slowing the rate of protein–protein interactions.
Das et al. argue that this is because arginine forms supramolecular assemblies
in solution. One of its drawbacks is that it is usually used and is most
effective at concentrations of 0.5–1.0 M. At these concentrations, it inter-
feres with subsequent chromatography on Ni2þ-chelate affinity columns
and will prevent the binding of most protein to an ion-exchange column
without further dilution or dialysis. The natural osmoprotectant proline has
also been found to be effective in some cases at increasing solubility both
in vitro and in vivo (Ignatova and Gierasch, 2006).
Glycerol and sugars (sucrose, mannitol, sorbitol, and trehalose): We have found
glycerol to be an excellent refolding additive in many cases, usually used in
the 5–30% range. One extreme method by Shimamoto et al. (1998) involves
solubilization of IB protein to 2–10 mg/ml by 6 M GuHCl, addition of
glycerol to 50% (v/v), dialysis into 75% (v/v) glycerol to remove denaturant,
and then rapid dilution to lower glycerol to 5–10%. Several sugars in the
0.5–1.5 M range have been seen to promote successful refolding (Bowden
and Georgiou, 1988).
Other chemical additives: A large variety of papers have reported the use of
materials such as polyethylene glycol (PEG) (Cleland et al., 1992), cyclo-
dextrin (Rozema and Gellman, 1996), and various nondetergent sulfobe-
taines (NDSBs) (Expert-Bezancon et al., 2003) to increase refolding for
certain proteins. The effects of certain ionic liquids (i.e., organic salts with
melting points below 100 C) such as N 0 -alkyl-N-methylimidazolium
chlorides as refolding additives have been reported (Lange et al., 2005).
The ethyl and propyl derivatives, at concentrations of 0.5–1.0 M, showed
effects in increasing activity and refolding yields comparable to L-arginine.
Detergents: In theory, detergents should be helpful in preventing aggre-
gation during refolding. At low concentrations they bind weakly to exposed
hydrophobic or sticky regions and mask them, preventing aggregation.
As their concentration decreases they dissociate and allow reformation of
native structure. This is thought to be why Sarkosyl works as well as it does.
At high concentration it is a denaturant, but at low concentration it acts as
an artificial chaperone and promotes refolding without aggregation.
Chaotropic agents (denaturants at high concentrations): Since aggregation is
due to interactions among partially folded intermediates, it has sometimes
been useful to have a nondenaturing amount of a chaotrope in the refolding
solution to dissociate aggregates, but not the more stable, native structures.
1 M urea and 0.5 M GuHCl have been used.
274 Richard R. Burgess
Target protein-specific additives: The same argument given above for the
potential benefit of certain divalent cations can be used to suggest that the
presence of appropriate substrates, or cofactors, or essential heme groups, for
example, would stabilize native structure and improve refolding yield.
At 320 nm, the absorbance of the empty well due to the plastic is about
0.2 and can be subtracted from the apparent absorbance readings due to
solution turbidity to give a corrected turbidity.
6. The results of the absorbance readings with the plate absorbance
subtracted can be graphed versus the solution number and one can see
which solutions give the lowest turbidity. Corrected turbidity values
range from 0.0 to about 0.4. We often find several conditions that give
low- or zero-corrected turbidity. One often gets somewhat different
results with protein denatured using different solubilizing agents such as
6 M GuHCl, 8 M urea, or 0.3% Sarkosyl.
7. Choose the best 2–3 conditions that are compatible with loading directly
(after filtration) onto an analytical size exclusion column (ANSEC, see
Section 5.7). To be really useful for large-scale refolding, it must also be
compatible with loading onto an appropriate ion-exchange column at a
salt concentration of about 0.1 m NaCl. The solution (about 200 ml) can
be removed from the corresponding well, filtered through a syringe filter
or centrifuged in a microfuge tube and loaded onto a 12-ml ANSEC
column. A refolding condition that gives low turbidity and mostly monomeric
material on the ANSEC column is the condition that is likely to be useful in the
large-scale refolding of the target protein.
in the middle of the column, and the bottom half of the column is 0 M
urea. One then loads the denatured protein and denaturant is applied at a
slow flow rate to the column. The protein will move down the column
faster than the buffer (because it is partially or completely excluded from
the bead) and move gradually into lower and lower concentrations of
denaturant until it has refolded and elutes from the column. The diffi-
culty here is to decide on the steepness and volume of the reverse
gradient. The various protocols and applications are summarized nicely
in several recent reviews and articles ( Jungbauer and Kaar, 2006;
Jungbauer et al., 2004; Oganesyan et al., 2005; Swietnicki, 2006;
Veldkamp et al., 2007). While this on-column refolding method seems
ideal, it is often the case that as the denaturant concentration is decreased,
the protein precipitates on the column and is lost. This is usually due to
the tendency of many researchers to overload their column or to load
from one end that becomes saturated with protein. In these cases, the fact
that the protein is tethered to the resin does not help because the bound
proteins are close enough together on the column that they can still
aggregate during refolding. One general procedure that sometimes helps
is to bind the protein in denaturant to the column in batch (obviously
not applicable to size exclusion chromatography) at well below saturat-
ing amounts, then load the resin into a column, and finally carry out the
subsequent washing and elution steps in column.
2. Use of folding catalysts such as protein and chemical chaperones, PPI, PDI.
Another nice approach is to covalently attach various enzymes and cha-
perones to a gel filtration column resin and then refold in column. The
enzymes that have been immobilized, either individually or perhaps better
in combinations are: PPI (catalyze proline isomerization), PDI (promote
disulfide shuffling), DsbA and DsbC to catalyze disulfide formation, and
the chaperones GroEL/GroES to aid in refolding (Baneyx and Mujacic,
2004; Jungbauer et al., 2004; Paul et al., 2007; Swietnicki, 2006).
3. High-pressure refolding. A number of papers, summarized by Quronfleh
et al. (2007) report the successful solubilization of IBs by exposure to
high hydrostatic pressure (1.5–2.5 kbar). It is claimed that exposure of
IBs to high pressure within a certain range will disaggregate the aggre-
gated protein but not denature it. The process of pressurizing, holding at
high pressure, and slowly decreasing the pressure in a variety of buffers
has been reported to result in high recovery of soluble and active
enzymes. While this method has been commercialized by Barofold
(Boulder, CO) using a specialized pressure cells (PreEMPTM high-
pressure chamber), the equipment is not widely available. In some
cases, while the IB is solubilized, the protein is largely in the form of
soluble multimers. Like the solubilization with denaturants and refolding
discussed for most of this chapter, one must find the right conditions for
optimal recovery of active protein by screening multiple conditions.
Refolding Proteins 277
4. Alkaline pH shift to solubilize and refold. Singh and Panda (2005) argue that
proteins in IBs are largely folded and can more effectively be refolded if
not subjected to high concentrations of GuHCl or urea. They state that
IB protein (recombinant human growth hormone) was solubilized effi-
ciently at pH 12.5 in 2 M urea and a concentration of 2 mg/ml, but
retained significant secondary structure. The solubilized protein was
then diluted into 2 M urea, pH 8 with a 40% recovery of activity.
6. Fuse to easily refolded protein, for example, a ‘‘solubility tag’’ such as NusA,
Trx, MBP, GST, SUMO, and HaloTag (see Chapter 16 in this volume).
7. Form disulfide bonds in the E. coli cytoplasm. The environment in the E. coli
cytoplasm is reducing, preventing most disulfide bond formation.
Mutations in the glutathione reductase (gor gene) and thioredoxin reduc-
tase (trxB gene) enhance formation of disulfide bonds in the E. coli cyto-
plasm (Bessette et al., 1999). Strains AD494 (DE3) and BL21trxB(DE3) are
trxB deficient. Novagen Origami strains are trxB and gor deficient.
10. Conclusion
Recombinant proteins overexpressed in E. coli often form insoluble
IBs in the cell. The IBs are easy to purify and solubilize, but finding suitable
conditions for efficient refolding of the solubilized protein is sometimes
difficult. The best general strategy seems to be to screen many different
refolding conditions in parallel to find ones that give the highest recovery of
enzyme activity or the lowest turbidity due to aggregation and that show the
least amount of soluble multimers. Many additives may prove useful in
refolding, but the surest way to prevent aggregation and precipitation upon
refolding is to refold at low protein concentration.
REFERENCES
Anfinsen, C. B. (1973). Principles that govern the folding of protein chains. Science
181, 223–230.
Anthony, L. C., Dombkowski, A. A., and Burgess, R. R. (2002). Using disulfide bond
engineering to study conformational changes in the b0 260–309 region of E. coli RNA
polymerase b0 during s70 binding. J. Bacteriol. 184, 2634–2641.
Arakawa, T., Ejima, D., Tsumoto, K., Obeyama, N., Tanaka, Y., Kita, Y., and
Timasheff, S. N. (2007). Suppression of protein interactions by arginine: A proposed
mechanism of the arginine effects. Biophys. Chem. 127, 1–8.
Armstrong, N., De Lencastre, A., and Gouaux, E. (1999). A new protein folding screen:
Application to the ligand binding domains of a glutamate and kainite receptor and to
lysozyme and carbonic anhydrase. Protein Sci. 8, 1475–1483.
Baneyx, F., and Mujacic, M. (2004). Review. Recombinant protein folding and misfolding
in E. coli. Nat. Biotechnol. 22, 1399–1408.
Baynes, B. M., Wang, D. I. C., and Trout, B. L. (2005). Role of arginine in the stabilization
of proteins against aggregation. Biochemistry 44, 4919–4925.
Bessette, P. H., Åslund, F., Beckwith, J., and Georgiou, G. (1999). Efficient folding of
proteins with multiple disulfide bonds in the E. coli cytoplasm. Proc. Natl. Acad. Sci. USA
96, 13703–13708.
Bowden, G. A., and Georgiou, G. (1988). The effect of sugars on b-lactamase aggregation in
E. coli. Biotech. Prog. 4, 97–101.
Bowden, G. A., Paredes, A. M., and Georgiou, G. (1991). Structure and morphology of
protein inclusion bodies in E. coli. Biotechnology (NY) 9, 725–730.
280 Richard R. Burgess
Tsumoto, K., Ejima, D., Kumagai, I., and Arakawa, T. (2003). Practical considerations in
refolding proteins from inclusion bodies. Protein Expr. Purif. 28, 1–8.
Vallejo, L. F., and Rinas, U. (2004). Review. Strategies for the recovery of active proteins
through refolding of bacterial inclusion body proteins. Microb. Cell Fact. 3, 11–22.
Veldkamp, C. T., Person, F. C., Hayes, P. L., Mattmiller, J. E., Haugner, J. C. III, de la
Cruz, N., and Volkman, B. F. (2007). On-column refolding of recombinant chemokines
for NMR studies and biological assays. Protein Expr. Purif. 52, 202–209.
Ventura, S., and Villaverde, A. (2006). Protein quality in bacterial inclusion bodies. Trends
Biotechnol. 24, 179–185.
Vera, A., Gozalez-Montalban, N., Aris, A., and Villaverde, A. (2006). The conformation
quality of insoluble recombinant proteins is enhanced at low growth temperatures.
Biotechnol. Bioeng. 96, 1101–1106.
Vincentelli, R., Canaan, S., Campanacci, V., Valencia, S., Maurin, D., Frassinetti, F., Scap-
pucini-Calvo, L., Bourne, Y., Cambillau, C., and Bignon, C. (2004). High-throughput
automated refolding screening of inclusion bodies. Protein Sci. 13, 2782–2792.
Willis, M. S., Hogan, J. K., Prabhakar, P., Liu, X., Tsai, K., Wei, Y., and Fox, T. (2005).
Investigation of protein refolding using a fractional factorial screen: A study of reagent
effects and interactions. Protein Sci. 14, 1818–1826.
Woycechowsky, K. J., Wittrup, K. D., and Raines, R. T. (1999). A small-molecule catalyst
of protein refolding in vitro and in vivo. Chem. Biol. 6, 871–879.
Xie, Y., and Wetlaufer, D. B. (1996). Control of aggregation in protein refolding: The
temperature-leap tactic. Protein Sci. 5, 517–523.
C H A P T E R E I G H T E E N
Contents
1.Introduction 286
2.Chemical and Enzymatic Cell Disruption 286
3.Mechanical Cell Disruption 290
4.Concluding Remarks 293
5.Procedures, Reagents, and Tips for Cell Disruption 293
5.1. Buffer composition 293
5.2. Cell disruption buffers 296
5.3. Microscale protocols for E. coli cell lysis 296
5.4. Gram-scale mechanical disruption of E. coli 299
References 301
Abstract
There are a variety of reliable methods for cellular disintegration and extraction
of proteins ranging from enzymatic digestion and osmotic shock to ultrasonica-
tion, and pressure disruption. Each method has inherent advantages and
disadvantages. Generally vigorous mechanical treatments reduce extract vis-
cosity but can result in the inactivation of labile proteins by heat or oxidation,
while gentle treatments may not release the target protein from the cells, and
resulting extracts are extremely viscous. Depending on the cell type selected as
the source for target protein expression, cellular extracts contain large amounts
of nucleic acid, ribosomal material, lipids, dispersed cell wall polysaccharide,
carbohydrates, chitin, small molecules, and thousands of unwanted proteins.
Isolation and recovery of a single protein from this complex mixture of macro-
molecules presents considerable challenges. The first and possibly most impor-
tant of these challenges is generation of a cellular extract that can be efficiently
manipulated in downstream processes without inactivation or degradation of
labile protein targets. Cell disruption techniques must rapidly and efficiently
lyse cells to extract proteins with minimal proteolysis or oxidation while reduc-
ing extract viscosity caused by cell debris and genomic DNA contamination.
Department of Research and Development, Semba Biosciences, Inc., Madison, Wisconsin, USA
285
286 Anthony C. Grabski
1. Introduction
Early protein chemists worked primarily with small extracellular
proteins that were stable, plentiful, and easily isolated without cell disrup-
tion. Intracellular proteins expressed at physiological levels might constitute
<0.001% of the total cellular protein and were difficult to extract
and recover without proteolysis or loss of enzymatic activity. Modern
recombinant protein production through genetic engineering of Escherichia
coli can result in expression of the target protein at over 40% of the cellular
total. Recombinant yeasts such as Saccharomyces cerevisiae, Pichia pastoris, and
Kluvyeromyces lactis are capable of expressing complex proteins requiring
posttranslational modifications (Spencer and Spencer, 1997). Recovery of
the intracellular recombinant proteins from these hosts requires effective cell
disruption and extraction techniques. Well established protocols have been
reviewed and detailed procedures described for laboratory and process scale
disruption of various cell types by both mechanical and nonmechanical
methods (Andrews and Asenjo, 1987; Cull and McHenry, 1990; Dignam,
1990; Engler, 1985; Gegenheimer, 1990; Harrison, 1991; Hopkins, 1991;
Jazwinski, 1990; Middelberg, 1995). Radical changes to standard disruption
equipment and procedures had not occurred up to 1995 (Foster, 1995).
However, the evolution of structural and functional proteomics demanded
new reagents and automated methods to streamline the process steps
converting gene sequences to purified proteins (Grabski et al., 2002).
Proteomics and structural genomics efforts require cell extraction meth-
ods that allow screening of multiple host vector combinations for hundreds
of proteins in parallel (Stevens et al., 2001). The best combinations for
correctly folded high-level expression identified in these multiparallel
screens are scaled up in order to produce milligram quantities of purified
protein for functional and structural analysis (Dieckman et al., 2002; Lesley,
2001; Stevens and Wilson, 2001; Zhu et al., 2001). These high-throughput
(HT) protein expression and purification strategies have fostered develop-
ment and optimization of reagents and instrumentation that allow efficient
cell lysis and protein extraction at both micro- and macroscales.
Preparation of Biological Extracts 287
they are gentle and do not generate shear, high temperature or oxidative
damage, and they are simple to use requiring no specialized equipment for
processing. Enzyme treatments of cells and extracts are also combined with
mechanical disruption methods to increase the selectivity of product release,
to increase the rate and yield of extraction, to minimize product damage,
and to reduce viscosity for downstream processes (Andrews and Asenjo,
1987; Grabski et al., 1999). Improved bioprocessing enzymes and enzyme
mixtures (Table 18.1) include high-specific activity phage lysozymes such
as rLysozymeTM, Ready-LyseTM, and Recombinant Lysozyme, nonspecific
nucleases like BenzonaseÒ and OmniCleaveTM, and the lysozyme-nuclease
cocktails LiquisonicTM, LysonaseTM, and EasyLyseTM. Recombinant Lyso-
zyme, Ready-LyseTM and rLysozymeTM phage lysozymes have specific
activities over 200-fold greater than that of chicken egg white lysozyme.
The nuclease of Serratia marcesens, available as a highly purified recombinant
enzyme BenzonaseÒ , is one of the most promiscuous nucleases known. The
enzyme cleaves all forms of DNA and RNA (Meiss et al., 1995; Nestle and
Roberts, 1969). Compared to bovine pancreatic DNase I, Benzonase
attacks substrates more evenly making it an excellent choice for reducing
extract viscosity caused by nucleic acids. ZymolaseÒ and Lyticase glucanases
are enzymes useful for yeast protoplasting or yeast cell lysis.
Potential problems with enzymatic treatment limit its use especially for
industrial cell disruption. These problems include the added lytic enzyme
and impurities in the enzyme preparation that complicate downstream
purification, degradation of the recovered product during lysis, tempera-
ture, and pH optima for the lytic enzyme may be incompatible with the
target protein, and the expense and limited availability of most lytic enzymes
prohibits their use at an industrial scale (Hopkins, 1991).
design with a combined valve seat and impact ring found in the APV
Manton-Gaulin homogenizer. The mechanism for disruption is a combined
pressure drop and shear at the nozzle in pressure extrusion. Numerous
mechanisms for cell disruption with pressure homogenizers have been
proposed including turbulence, cavitation, viscous shear, and impingement
(Kleinig and Middelberg, 1998; Pandolfe, 1999). Universal acceptance for a
single mechanism has not been reached. However, cavitation and impinge-
ment have been reported as the major forces responsible for disruption
(Shirgaonkar et al., 1998; SPX Corp., 2009).
Cell disruption using glass beads to grind the cells in suspension, also
known as bead milling or bead homogenization, is a frequently used
procedure at both laboratory and production scale. Bead milling can be
accomplished with laboratory equipment as simple as a magnetic stirrer,
vortex mixer, or blender and with commercial equipment specialized for
the process in the form of high speed mills, agitators, and mixers. The cell
disintegration is dependent on cell concentration, bead size and composi-
tion, ratio of beads to suspension, processing duration, and the forces
applied (Ramanan et al., 2008). The method is effective with difficult-
to-disrupt cells like yeasts, spores, and microalgae. It is used for bacteria,
plant and animal cells and is a preferred method for large-scale disruption of
fungi (Hopkins, 1991). The mechanism and models of the bead milling
process have been reviewed (Harrison, 1991; Middelberg, 1995), but basi-
cally the cells are crushed, ground, and torn apart by abrasive contact and
shear forces created between the beads, cells, and reaction chamber itself.
The specialized equipment necessary for mechanical cell disruption
techniques has undergone functional innovation and new instruments
have been developed. The majority of mechanical methods are not easily
applied to cell pellets harvested from 5 ml of culture or less, and excess heat
and oxidation are common problems with mechanical disruption. How-
ever, 96-well sonicator heads and microplate horns are available (Misonix,
Inc. Farmingdale, NY; www.misonix.com) and for cases where traces of
detergent in the purified protein preparation might interfere with biochem-
ical characterization or crystallography, this HT physical method is a viable
option. The SonicMan high-throughput sonication system (MatriCal, Inc.
Spokane, WA; www.matrical.com) is a stand alone or integrated unit for
96-, 384-, and 1536-well plate sonications. This touch screen controlled
microplate sonicator uses disposable gasketed pin lids to prevent well-
to-well cross-contamination and has a plate shuttle allowing direct integra-
tion into robotics platforms. Another new sonicator available from BioSpec
Products is the cordless handheld Sonozap Ultrasonic Homogenizer. The
1/8 in. diameter auto-tuned probe is ideal for small samples of 0.3–5.0 ml.
Pressure Biosciences, Inc. (www.pressurebiosciences.com) has developed
the BarocyclerTM, bench top, and PCT Shredder, hand held, sample
preparation systems. These instruments are capable of rapid, high-pressure
292 Anthony C. Grabski
4. Concluding Remarks
The cell disruption methods, reagents, and instrumentation described
in this chapter are useful and effective if appropriately modified for different
cell types and scales, but the question of which method is superior for each
application has certainly been asked. Studies have been conducted compar-
ing methods of microbial disruption and protein quantitation in resulting
extracts (Benov and Al-Ibraheem, 2002; De Mey et al., 2008; Guerlava
et al., 1998; Ho et al., 2008). The mechanical methods of high-pressure
homogenization and bead milling are favorite methods at large scale because
of their effectiveness and ability to rapidly handle various process volumes at
low cost. Sonication, chemical reagents including detergents, enzymatic
treatments, freeze-thaw, and enzymatic plus chemical or physical method
combinations are also very effective and are used more frequently at labora-
tory and especially microscale. The uniqueness of proteins and differences in
host cell structure force a high degree of empiricism in selection and
optimization of cell disruption and extract preparation techniques.
However, the tools and methods for success in biological extract prepara-
tion are available and have kept pace with the demands of modern structural
and functional proteomics.
extraction and purification buffer tailored specifically for its own biochemi-
cal requirements and intended direction through the purification pipeline.
In most cases a more generic extraction buffer can be used with good results
if a few basic criteria are met. These criteria are pH, ionic strength, additives
to prevent degradation and improve stability, and buffer to cell paste ratio.
A very good reference for maintenance of active enzymes through buffer
composition and other means is Scopes (1994; ‘‘Protein Purification’’).
Technical information on pH and buffers including tables for preparation
of pH 1–13 buffers, buffer properties, and the influence of salt, temperature,
and dilution values can be found by Dawson et al. (1986).
The pH selected for the extraction buffer should be at least one pH unit
above or below the protein’s isoelectric point. This pH difference from the
pI will prevent isoelectric precipitation by maintaining a positive or nega-
tive charge on the protein and also facilitate ion-exchange purification as a
purification step. The buffer ionic strength should be 20–50 mM with a pKa
within 0.5 unit of the desired pH in order to maintain buffering capacity
with minimal conductivity increase. The ionic strength inside the cytoplasm
of a typical cell is 150–200 mM with very high concentrations of charged
biomolecules for ionic protein interaction. The lysis buffer should contain at
least 50–100 mM NaCl. Increased ionic strength of the extraction buffer
will reduce these ionic interactions and precipitating losses from charged
particulates that would adsorb protein and be removed by centrifugation or
filtration steps. Finally, buffer components to prevent degradation and
improve stability should be added as necessary. These include protease
inhibitors (Table 18.2), reducing agents such as dithiothreitol (DTT), tris
(2-carboxyethyl)phosphine (TCEP), or tris(hydroxypropyl)phosphine
(THP), divalent cations, cofactors, and kosmotropes like glycerol, sorbitol,
or trehalose. The water soluble odorless phosphines TCEP and THP are
more stable and effective for the maintenance of reduced protein disulfide
bonds than the more commonly used thiol reductants 2-mercaptoethanol
and DTT (Cline et al., 2004; Getz et al., 1999; Han and Han, 1994).
Nonionic or zwitterionic detergents may be added to increase solubility
of hydrophobic proteins. Reducing agents, protease inhibitors, and deter-
gents may interfere with some purification and detection methods and
assays. The potential interference of these buffer components must be
considered when selecting the chemicals and concentrations employed.
A very versatile buffer (Buffer B) for bacterial cell disruption is 50 mM
Tris/HCl or sodium phosphate pH 7.5–8.0, 50 mM NaCl, 5% glycerol,
0.5 mM EDTA, and 0.5 mM DTT. Buffer formulations recommended for
yeast disruption (Buffer Y) and insect cell disruption (Buffer I) are given
below.
The volume of buffer used for resuspension of cell pellets must be at least
three times the volume of the original pellet for effective disruption and
good recovery of the liquid fraction after removal of insoluble cell
Table 18.2 Protease inhibitors
in step 8 and the soluble protein in step 9. The samples are typically
compared by electrophoresis, ELISA or enzymatic assay. Direct analysis
of the insoluble fraction can be done electrophoretically by aspiration
of the soluble supernatant, solubilization of the pelleted fraction in
SDS-PAGE sample buffer, and SDS-PAGE.
The larger pore size filter in-line first prevents rapid fouling of the
smaller pore size filter. This final clarified supernatant represents the
soluble cell protein fraction and is ready for chromatography or other
downstream fractionation or analysis procedures.
REFERENCES
Andrews, B. A., and Asenjo, J. A. (1987). Enzymatic lysis and disruption of microbial cells.
TIBTECH 5, 273–277.
Benov, L., and Al-Ibraheem, J. (2002). Disrupting Escherichia coli: A comparison of methods.
J. Biochem. Mol. Biol. 35, 428–431.
Burgess, R. R. (1987). In ‘‘Protein Purification: Micro to Macro’’, (R. R. Burgess, ed.),
Alan R. Liss, New York.
Catley, B. J. (1988). Isolation and analysis of cell walls. In ‘‘Yeast: A Practical Approach’’,
(I. Campbell and J. H. Duffus, eds.), pp. 163–183. IRL Press, Washington.
Chu, R., and Mallia, K. (2001). Method for recovery of proteins prepared by recombinant
DNA procedures. U. S. Patent 6,174,704.
Chu, R., Mallia, K., Brennan, T., and Klenk, D. (1998). Recombinant protein extraction
from E. coli using B-PERÒ Bacterial Protein Extraction Reagent. Previews 2, 12–13.
Cline, D. J., Redding, S. E., Brohawn, S. G., Psathas, J. N., Schneider, J. P., and Thorpe, C.
(2004). New water-soluble phosphines as reductants of peptide and protein disulfide
bonds: Reactivity and membrane permeability. Biochemistry 43, 15195–15203.
Cull, M., and McHenry, C. S. (1990). Preparation of extracts from prokaryotes. In ‘‘Guide
to Protein Purification’’, (M. P. Deutscher, ed.), Meth. Enzymol. 182, pp. 147–153.
Dawson, R. M. C., Elliott, D. C., Elliott, W. H., and Jones, K. M. (1986). pH, buffers,
and physiological media. In ‘‘Data for Biochemical Research’’, pp. 417–448. Oxford
University Press, Inc., New York.
De Mey, M., Lequeux, G. J., Maertens, J., De Muynck, C. I., Soetaert, W. K., and
Vandamme, E. J. (2008). Comparison of protein quantification and extraction methods
suitable for E. coli cultures. Biologicals 36, 198–202.
Guide to protein purification. Deutscher, M. P. (ed.) (1990). Meth. Enzymol.
182, pp. 147–203.
Dieckman, L., Gu, M., Stols, L., Donnelly, M. I., and Collart, F. R. (2002). High-
throughput methods for gene cloning and expression. Protein Expr. Purif. 25, 1–7.
Dignam, J. D. (1990). Preparation of extracts from higher eukaryotes. In ‘‘Guide to Protein
Purification’’, (M. P. Deutscher, ed.), Meth. Enzymol. 182, pp. 194–203.
1
Drott, D., Bahairi, S., and Grabski, A. (2002). Yeast Buster protein extraction reagent for
fast, efficient extraction of proteins from yeast. inNovations 15, 14–16.
Engler, C. R. (1985). Disruption of microbial cells. In ‘‘Comprehensive Biotechnology’’,
(C. L. Cooney and A. E. Humphrey, eds.), Vol. 2, pp. 305–324. Pergamon Press,
Toronto.
Eshaghi, S., Hedren, M., Abdel Nasser, M. I., Hammarberg, T., Thornell, A., and
Norlund, P. (2005). An efficient strategy for high-throughput expression screening of
recombinant integral membrane proteins. Protein Sci. 14, 676–683.
Foster, D. (1995). Optimizing recombinant protein recovery through improvements in cell-
disruption technologies. Curr. Opin. Biotechnol. 6, 523–526.
Garrett, P. E., Tao, F., Lawrence, N., Ji, J., Schumacher, R. T., and Manak, M. M. (2002).
Tired of the same old grind in the new genomics and proteomics era? TARGETS
5, 156–162.
Gegenheimer, P. (1990). Preparation of extracts from plants. In ‘‘Guide to Protein Purifica-
tion’’, (M. P. Deutscher, ed.), Meth. Enzymol. 182, pp. 174–193.
Getz, E. B., Xiao, M., Chakrabarty, T., Cooke, R., and Selvin, P. R. (1999). A comparison
between the sulfhydryl reductants tris(2-carboxyethyl) phosphine and dithiothreitol for
use in protein biochemistry. Anal. Biochem. 273, 73–80.
1
References available online: www.emdbiosciences.com.
302 Anthony C. Grabski
1
Grabski, A., and Burgess, R. R. (2001). Preparation of samples for SDS-polyacrylamide gel
electrophoresis: Procedures and tips. inNovations 13, 10–12.
1
Grabski, A., McCormick, M., and Mierendorf, R. (1999). BugBusterTM and BenzonaseÒ :
The clear solutions to simple efficient extraction of E. coli proteins. inNovations 10, 17–19.
1
Grabski, A., Drott, D., Handley, M., Mehler, M., and Novy, R. (2001). Extraction and
purification of proteins from E. coli without harvesting cells. inNovations 13, 1–4.
1
Grabski, A., Mehler, M., Drott, D., and Van Dinther, J. (2002). Automated purification of
recombinant proteins in a 96-well format with RoboPopTM Kits and robotic sample
processing. inNovations 14, 2–5.
1
Grabski, A., Mehler, M., and Luedke, R. (2003). Automated methods for solubility
screening and recombinant protein purification. Amer. Biotechnol. Lab. 35, 34–38.
Guerlava, P., Izac, V., and Tholozan, J.-L. (1998). Comparison of different methods of cell
lysis and protein measurements in Clostridium perfringens: Application to the cell volume
determination. Curr. Microbiol. 36, 131–135.
Han, J. C., and Han, G. Y. (1994). A procedure for quantitative determination of tris
(2-carboxyethyl)phosphine, an odorless reducing agent more stable and effective than
dithiothreitol. Anal. Biochem. 220, 5–10.
Harris, E. L. V., and Angal, S. (eds.) (1989). In ‘‘Protein Purification Methods, a Practical
Approach’’, IRL Press, Oxford.
Harrison, S. T. L. (1991). Bacterial cell disruption: A key unit operation in the recovery of
intracellular products. Biotechnol. Adv. 9, 217–240.
Ho, C. W., Tan, W. S., Yap, W. B., Ling, T. C., and Tey, B. T. (2008). Comparative
evaluation of different cell disruption methods for the release of recombinant hepatitis B
core antigen from E. coli. Biotechnol. Bioproc. Eng. 13, 577–583.
Hopkins, T. R. (1991). Physical and chemical cell disruption for the recovery of intracellular
proteins. In ‘‘Purification and Analysis of Recombinant Proteins’’, (R. Seetharam and
S. K. Sharma, eds.), pp. 57–83. Marcel Dekker, Inc., New York.
Jazwinski, S. M. (1990). Preparation of extracts from yeast. In ‘‘Guide to Protein Purifica-
tion’’, (M. P. Deutscher, ed.), Meth. Enzymol. 182, pp. 154–174.
Kashino, Y. (2003). Separation methods in the analysis of protein membrane complexes.
J. Chromatgr. B 797, 191–216.
Kleinig, A. R., and Middelberg, A. P. J. (1998). On the mechanism of microbial cell
disruption in high-pressure homogenization. Chem. Eng. Sci. 53, 891–898.
Lesley, S. A. (2001). High-throughput proteomics: Protein expression and purification in
the postgenomic world. Protein Expr. Purif. 22, 159–164.
Lieve, L. (1974). The barrier function of the gram-negative envelope. Ann. N.Y. Acad. Sci.
235, 109–129.
Marshak, D. R., Kadonaga, J. T., Burgess, R. R., Knuth, M. W., Brennan, W. A., and
Lin, S.-H. (1996). Solubilization and purification of the rat liver insulin receptor.
In ‘‘Strategies for Protein Purification and Characterization: A Laboratory Manual’’,
pp. 275–336. Cold Spring Harbor Laboratory Press, New York.
Meiss, G., Friedhoff, P., Hahn, M., Gimadutdinow, O., and Pingoud, A. (1995). Sequence
preferences in cleavage of dsDNA and ssDNA by extracellular Serratia marcescens endo-
nuclease. Biochemistry 34, 11979–11988.
Middelberg, A. P. J. (1995). Process-scale disruption of microorganisms. Biotechnol. Adv.
13, 491–551.
Nestle, M., and Roberts, W. K. (1969). An extracellular nuclease from Serratia marcescens.
J. Biol. Chem. 244, 5219–5225.
Neugebauer, J. M. (1990). Detergents: An overview. In ‘‘Guide to Protein Purification’’,
(M. P. Deutscher, ed.), Meth. Enzymol. 182, pp. 239–253.
Nguyen, H., Martinez, B., Oganesyan, N., and Kim, R. (2004). An automated small-scale
protein expression and purification screening provides beneficial information for protein
production. J. Struct. Funct. Genom. 5, 23–27.
Preparation of Biological Extracts 303
Contents
1. Introduction 306
2. Extraction and Prefractionation of Subproteomes 308
2.1. Density velocity centrifugation 308
2.2. Density gradient centrifugation 311
2.3. Mitochondria enrichment 311
2.4. Endomembrane enrichment (LOPIT) 314
2.5. Differential detergent fractionation 316
2.6. Lipid raft enrichment (detergent-resistant fractionation) 320
2.7. Nuclei and histone enrichment 322
2.8. Purification of core histones 324
2.9. Nuclear extract enrichment 325
2.10. Immunoaffinity reagents for organelle enrichment 326
References 327
Abstract
One of the major challenges in functional proteomics is the separation of
complex protein mixtures to allow detection of low abundance proteins and
provide for reliable quantitative and qualitative analysis of proteins impacted by
environmental parameters. Prerequisites for the success of such analyses are
standardized and reproducible operating procedures for sample preparation
prior to protein separation. Due to the complexity of total proteomes, especially
of eukaryotic proteomes, and the divergence of protein properties, it is often
beneficial to prepare standardized partial proteomes of a given organism to
maximize the coverage of the proteome and to increase the chance to visualize
low abundance proteins and make them accessible for subsequent analysis.
In this chapter we will describe with detailed recipes procedures for the
enrichment and isolation of the currently most investigated organelles and
subcellular compartments in mammalian cells using classical centrifugation
techniques to more sophisticated immunoaffinity-based procedures.
305
306 Uwe Michelsen and Jörg von Hagen
Abbreviations
2D Two-dimensional
DDF Differential detergent fractionation
DRF Detergent-resistant fractions
DRM Detergent-resistant membranes
EMSA electro mobility shift assay
g Gravity unit
GPCR G-Protein coupled receptor
HM Homogenization medium
IEF Isoelectric focussing
iTRAQ Isotope tags for relative and absolute quantification
LC Liquid chromatography
LOPIT Localization of organelle proteins by isotope tagging
M Molar
b-MCD b-Methylcyclodextrin
MS Mass spectrometry
MudPIT Multidimensional protein identification technology
PBS Phosphate buffered saline
PMSF Phenylmethylsulfonyl fluoride
RAM Restricted access material
RP Reversed phase
SCX Strong cation exchange
SDS Sodium dodecyl sulfate
1. Introduction
In proteomics research, one essential step among enrichment
techniques is subcellular fractionation, which is of special importance for
analysis of intracellular organelles and multiprotein complexes. To reduce
sample complexity, subcellular fractionation is a flexible and adjustable
approach and is most efficiently combined with high-resolution 2D gel/
mass spectrometry analysis as well as with gel-independent techniques.
Isolating distinct subcellular compartments by fractionation has been
among the standard strategies established in biochemistry-oriented labora-
tories for decades. Determination of marker enzyme activities was to assay
the efficiency of the subcellular fractionation and a major goal was the
identification of individual new proteins. At that time, the power of protein
identification techniques was very limited, the characterization of the
Isolation of Subcellular Organelles and Structures 307
Sample preparation
Remove high Extraction and Organelle Specific Specific
abundance proteins pre-fractionation isolation enrichment labeling
Separation
1D/2D Liquid Thin layer Free flow
gel electrophoresis chromatography chromatography electrophoresis
Characterization/functional validation
Mass Protein visualization
Crystallography NMR
spectrometry and quantitation
Sample
Mechanically disrupted/precleared
Mitochondria, lysosomes
peroxisomes and 100,000 x g/60 min
chloroplasts
Sample 0.5 ml
15% 1.0 ml
20% 1.0 ml
25% 1.0 ml
27.5% 0.75 ml
30% 0.5 ml
35% 0.5 ml
Figure 19.3 Sucrose gradient setup: 15–35% sucrose density gradient. The gradient is
prepared by layering less dense sucrose solutions upon one another; therefore, the first
solution applied is the 35% sucrose solution. Firstly, a Beckman polyallomer tube is held
upright in a tube stand. Next a tip is placed held steady by a clamp stand and the end of
the tip is allowed to make contact with the inside wall of the tube as shown above.
Alternatively use five and more freeze thaw cycles, five or more soni-
cations (5 pulses of 5 s), French press, high pressure homogenization
(2 cycles of 500–1000 bar) or a bead mill according to the protocol
provided by the various suppliers. For the following application preclear
the lysate by a centrifugation step of 5 min at 500g at 4 C.
Procedure:
1. A mitochondrial membrane suspension at max. 5 mg/mL protein in PBS
is adjusted with lauryl maltoside to a final concentration of 1%.
2. Mix well and incubate on ice for 30 min.
3. Centrifuge at 70,000g for 30 min.
4. The supernatant is collected and the pellet discarded.
5. Add a protease inhibitor cocktail and keep the sample on ice until
centrifugation is performed.
6. Once the sucrose gradient is poured, discrete layers of sucrose should be
visible. Having applied the sample to the top of the gradient the tube
should be handled carefully.
7. The polyallomer tubes should be centrifuged in a swinging bucket type
rotor at 130,000g for 16 h at 4 C with the lowest acceleration and
deceleration profile.
8. Carefully remove 500 mL fractions from top to the bottom and analyze
the fractions. The fractions can now be stored at 80 C.
Table 19.2 Enzyme and substrate markers for different cellular organelles
Digitonin H
O
H O
HO H
H H OH
Xyl-Glc-Gal
Glc-Gal H
In
cr
Triton-X 100 ea
O H sin
O g
n ex
tra
cti
on
eff
ica
cy
Tween 20
HO O O OH
O O
w x
O OH
O y
O
O
O O
z
O
OH
Deoxycholate O
OH
H
HO
Figure 19.4 Commonly used detergents with increased extraction efficacy from
Digitonin to sodium Deoxycholate.
Sample
Adherent or suspension cells
Digitonin extraction
Pe
lle
Cytosolic fraction Triton-X 100 extraction
t
Membrane fraction
(organelles) Tween 20/deoxycholate extraction
Nuclear fraction
Su
SDS extraction
pe
rn
at
an
t
Cytoskelatal fraction
15. Pellets from suspension cultures are washed once with 20 C of 90%
acetone, lyophilized, and weights determined in tared centrifuge tubes.
Samples are stored at 70 C.
16. Monolayers are rinsed in situ with PBS, and the detergent-resistant
residue (Cytoskeleton) is suspended directly into 0.5–1 mL nondenaturing
cytoskeleton solubilization buffer without b-mercaptoethanol. Store at
70 C.
Procedure:
1. Plate 2.0–3.0 107 culture cells (adherent or suspension) in culture
flasks.
2. Wash the cells twice with prewarmed PBS.
3. Where required treat the cells either with b-MCD (10 mM) for 45 min
and/or with cholesterol for 1 h in the serum free medium at 37 C.
4. After treatment or 72 h of infection stimulate the cells where required
5. Lyse the cells in cold 750 mL of TNE-buffer containing 1% Triton
X-100 supplemented with protease and phosphatase inhibitors for
30 min on ice.
6. Mix the lysate with 1500 mL of 70% Nycodenz, dissolved in TNE-
buffer.
7. Load this mix in the bottom of 4 mL of polyallomer ultracentrifuge
tube.
8. Overlay this bottom loaded mix with 25, 21.5, 18, 15, and 8% Nyco-
denz, prepared in TNE-buffer.
9. Spin the tube at 200,000g for 4 h at 4 C using a swing-out rotor.
10. Fractions each of 350 mL from top to the bottom of the tube were
analyzed by western blotting.
11. Probe the blot to confirm caveolin-1 positive (fractions 3–6) or
detergent-resistant membrane fractions.
Note: Unless otherwise indicated, keep all solutions and materials on ice
or at 4 C.
Procedure:
1. 3 109 cells (HeLa) were washed twice in 1 L PBS.
2. Resuspend pellet in 40 mL lysis buffer.
3. Lyse the cells with 15 strokes of a type B pestle Dounce homoge-
nizer. Monitor lysis by light microscopy.
4. Centrifuge 15 min at 3000g, 4 C.
5. Repeat resuspension and pelleting twice with lysis buffer and once with
NPB.
324 Uwe Michelsen and Jörg von Hagen
6. Resuspend nuclei in 2 pellet vol NPB and measure the total volume of
the suspension.
7. While gently stirring, add dropwise 1 total vol of NPB2.
8. Continue gentle stirring for 10 min at 4 C.
9. Pellet nuclei 30 min at 17,500g, 4 C.
10. Nuclear pellets can be frozen in liquid nitrogen or dry ice and kept for
more than a year at 80 C.
Procedure:
Prepare a precleared cell lysate as described above:
1. Incubate the lysate, antibody (according to the suppliers information),
Protein A/G and solid phase (30 mL 1:1 slurry solution) overnight at
4 C with gentle agitation.
Isolation of Subcellular Organelles and Structures 327
Coated vesicle/organelle
Protein A or G
Antibody
ic/ ria
er acte
y m b
ol r
(p c) o
d
a e t i
Be agn
m
REFERENCES
Abdolzade-Bavil, A., Hayes, S., Goretzki, L., Kröger, M., Anders, J., and Hendriks, R.
(2004). Convenient and versatile subcellular extraction procedure, that facilitates classical
protein expression profiling and functional protein analysis. Proteomics 4, 1397–1405.
Billecke, C., Malik, I., Movsisyan, A., Sulghani, S., Sharif, A., Mikkelsen, T., Farrell, N. P.,
and Bögler, O. (2006). Analysis of glioma cell platinum response by metacomparison of
two-dimensional chromatographic proteome profiles. Mol. Cell Proteomics 5, 35–42.
Edelman, M., Epstein, H. T., and Schiff, J. A. (1966). Isolation and characterization of DNA
from the mitochondrial fraction of Euglena. J. Mol. Biol. 17, 463–469.
Fisher, W. D., and Cline, G. B. (1963). A density gradient for the isolation of metabolically
active thymus nuclei. Biochim. Biophys. Acta 68, 640–642.
Ford, T., Graham, J., and Rickwood, D. (1994). Iodixanol: A nonionic iso-osmotic
centrifugation medium for the formation of self-generated gradients. Anal. Biochem.
220, 360–36612.
Goldberg, S. (2008). Mechanical/physical methods of cell disruption and tissue homogeni-
zation. Methods Mol. Biol. 424, 3–22.
Graham, J., Ford, T., and Rickwood, D. (1994). The preparation of subcellular organelles
from mouse liver in self-generated gradients of iodixanol. Anal. Biochem. 220, 367–373.
Kurokawa, M., Kato, M., and Sakamoto, T. (1965). Distribution of sodium-plus-potassium-
stimulated adenosine-triphosphatase activity in isolated nerve-ending particles. Biochem.
J. 97, 833–844.
328 Uwe Michelsen and Jörg von Hagen
Nakamura, E., Kozaki, K., Tsuda, H., Suzuki, E., Pimkhaokham, A., Yamamoto, G.,
Irie, T., Tachikawa, T., Amagasa, T., Inazawa, J., and Imoto, I. (2008). Frequent
silencing of a putative tumor suppressor gene melatonin receptor 1 A (MTNR1A) in
oral squamous-cell carcinoma. Cancer Sci. 99, 1390–1400.
Pertoft, H., Rubin, K., Kjellén, L., Laurent, T. C., and Klingeborn, B. (1977). The viability
of cells grown or centrifuged in a new density gradient medium. Percoll. Exp. Cell Res.
110, 449–457.
Ramsby, M. L., Makowski, G. S., and Khairallah, E. A. (1994). Differential detergent
fractionation of isolated hepatocytes: Biochemical, immunochemical and two-dimensional
gel electrophoresis characterization of cytoskeletal and noncytoskeletal compartments.
Electrophoresis 15, 265–277.
Rickwood, D., Ford, T., and Graham, J. (1982). Nycodenz: A new nonionic iodinated
gradient medium. Anal Biochem. 123, 23–31.
Ross, P. L., Huang, Y. L. N., Marchese, J. N., Williamson, B., Parker, K., Hattan, S.,
Khainovski, N., Pillai, S., Dey, S., Daniels, S., et al. (2004). Multiplexed protein
quantitation in Saccharomyces cerevisiae using amine-reactive isobaric tagging reagents.
Mol. Cell. Proteomics 3, 1154–1169.
Sadowski, P. G., Dunkley, T. P., Shadforth, I. P., Dupree, P., Bessant, C., Griffin, J. L., and
Lilley, K. S. (2006). Quantitative proteomic approach to study subcellular localization of
membrane proteins. Nat. Protoc. 1, 1778–1789.
Schnitzler, G. R. (2001). Isolation of histones and nucleosome cores from mammalian cells.
Curr. Protoc. Mol. Biol. Chapter 21, Unit 21.5.
Shadforth, I., Dunkley, T., Lilley, K., and Bessant, C. (2005). i-Tracker: For quantitative
proteomic using iTRAQ. BMC Genomics 6, 145.
Shah, M. B., and Sehgal, P. B. (2007). Nondetergent isolation of rafts. Methods Mol. Biol.
398, 21–28.
Sherrier, D. J., Prime, T. A., and Dupree, P. (1999). Glycosylphosphatidylinositol-anchored
cell-surface proteins from Arabidopsis. Electrophoresis 20, 2027–2035.
Simon, R. H., and Felsenfeld, G. (1979). A new procedure for purifying histone pairs H2A þ
H2B and H3 þ H4 from chromatin using hydroxylapatite. Nucleic Acids Res. 6, 689–696.
Storrie, B., and Madden, E. A. (1990). Isolation of subcellular organelles. Methods Enzymol.
182, 203–225.
Thompson, E. B., and Miller, J. V. (1975). Enrichment of polysomes synthesizing a specific
protein by use of affinity chromatography. Methods Enzymol. 40, 266–273.
Van Veldhoven, P. P., Baumgart, E., and Mannaerts, G. P. (1996). Iodixanol (Optiprep), an
improved density gradient medium for the iso-osmotic isolation of rat liver peroxisomes.
Anal. Biochem. 237, 17–23.
von Hagen, J. (2008). Proteomics Sample Preparation. New York: Wiley VCH, New York.
Washburn, M. P., Wolters, D., and Yates, J. R. (2001). Large-scale analysis of the yeast
proteome by multidimensional protein identification technology. Nat. Biotechnol. 19,
242–247.
Weber, P. J., Weber, G., and Eckerskorn, C. (2004). Isolation of organelles and prefractio-
nation of protein extracts using free-flow electrophoresis. Curr. Protoc. Protein Sci.
Chapter 22, Unit 22.5.
Willemsen, O., Machtejevas, E., and Unger, K. K. (2004). Enrichment of proteinaceous
materials on a strong cation-exchange diol silica restricted access material: protein-
protein displacement and interaction effects. J. Chromatogr. A. 1025, 209–216.
Zagariya, A., Khrapunov, S., and Zacharias, W. (1993). Rapid method for the fractionation
of nuclear proteins and their complexes by batch elution from hydroxyapatite. J. Chro-
matogr. 648, 275–278.
Zeheb, R., and Orr, G. A. (1986). Use of avidin-iminobiotin complexes for purifying
plasma membrane proteins. Methods Enzymol. 122, 87–94.
C H A P T E R T W E N T Y
Contents
1. Introduction 332
2. Ammonium Sulfate Precipitation 332
2.1. Principles 332
2.2. Basic procedure 334
2.3. Doing an ammonium sulfate precipitation test 336
2.4. Comments/problems/solutions 337
3. Polyethyleneimine Precipitation 337
3.1. Principles 337
3.2. Different modes of use of PEI 338
3.3. Basic procedure for Strategy C 338
3.4. Doing an PEI precipitation test 340
3.5. An example of using PEI to precipitate a basic protein
bound to DNA 340
4. Other Methods 341
4.1. Ethanol and acetone precipitation 341
4.2. Isoelectric precipitation 341
4.3. Thermal precipitation 341
4.4. Polyethylene glycol (nonionic polymer) precipitation 341
5. General Procedures When Fractionating Proteins by Precipitation 341
References 342
Abstract
After cell lysis, the most often used second step in a protein purification
procedure is some sort of a rapid, bulk precipitation step. This is commonly
accomplished by altering the solvent conditions and taking advantage of the
changes in solubility of your protein of interest relative to those of many of
the other proteins and macromolecules in a cell extract. This chapter will focus
on the two most widely used precipitation methods: (1) ammonium sulfate
precipitation and (2) polyethyleneimine (PEI) precipitation. These two methods
work through entirely different principles, but each can achieve significant
enrichment of target protein if optimized and applied carefully.
McArdle Laboratory for Cancer Research, University of Wisconsin–Madison, Madison, Wisconsin, USA
331
332 Richard R. Burgess
1. Introduction
For both laboratory scale and larger scale protein fractionation, there is
a need for a quick, bulk precipitation to remove much of cellular protein
and other components. It is especially important to remove proteases as
early in the procedure as possible to avoid protein degradation. This precip-
itation must be rapid, gentle, scalable, and relatively inexpensive. In addi-
tion to the fractionation, it can also achieve a significant concentration of
the enriched protein. While many different precipitation methods have
been used over the last hundred years, ammonium sulfate (AS) has remained
the most widely used and polyethyleneimine (PEI) has increased in popu-
larity, especially for acidic proteins. These two methods will be discussed in
detail, followed briefly by several other precipitation methods and some
general advice on handling precipitates and obtaining maximal purification
from your precipitation step. An extensive and very useful general overview
of various types of protein precipitation procedures can be found in Scopes
(1994). An excellent review of AS and organic solvent (ethanol and
acetone) precipitation can be found in Englard and Seifter (1990).
0
1 mg/ml
−1
0.1 mg/ml
−2
0.01 mg/ml
−3
26 32 38
0 10 20 30 40 50
Ammonium sulfate (% saturation)
Figure 20.1 Ammonium sulfate solubility curve for a hypothetical protein. This
represents the log solubility of a hypothetical protein as a function of percent saturation
of ammonium sulfate. The ‘‘salting-out’’ line follows the relationship log S ¼ b
Ks(G/2) as described in the text, where G/2 is the ionic strength, which here is given
as ammonium sulfate percent saturation.
extract is 0.1 mg/ml or log S ¼ 1. You can add AS to 32% saturation and
your protein will not precipitate. To achieve 90% precipitation of your
protein, you would have to increase the AS to about 38% saturation
(bottom horizontal dotted line) or carry out a 32–38% saturated AS cut.
You would end up having to use more than 10 times as much AS with the
diluted extract to obtain your protein. This illustrates how important it is to
specify the concentration of your extract.
You do not usually have a curve like that shown in Fig. 20.1 for your
protein of interest so you have to determine the appropriate AS concentra-
tions experimentally as described below.
0 106 134 164 194 226 258 291 326 361 398 436 476 516 559 603 650 697
5 79 108 137 166 197 229 262 296 331 368 405 444 484 526 570 615 662
10 53 81 109 139 169 200 233 266 301 337 374 412 452 493 536 581 627
15 26 54 82 111 141 172 204 237 271 306 343 381 420 460 503 547 592
20 0 27 55 83 113 143 175 207 241 276 312 349 387 427 469 512 557
25 0 27 56 84 115 146 179 211 245 280 317 355 395 436 478 522
30 0 28 56 86 117 148 181 214 249 285 323 362 402 445 488
35 0 28 57 87 118 151 184 218 254 291 329 369 410 453
40 0 29 58 89 120 153 187 222 258 296 335 376 418
45 0 29 59 90 123 156 190 226 263 302 342 383
50 0 30 60 92 125 159 194 230 268 308 348
55 0 30 61 93 127 161 197 235 273 313
60 0 31 62 95 129 164 201 239 279
65 0 31 63 97 132 168 205 244
70 0 32 65 99 134 171 209
75 0 32 66 101 137 174
80 0 33 67 103 139
85 0 34 68 105
90 0 34 70
95 0 35
100 0
a
Reprinted from Englard and Seifter (1990), which was adapted from Dawson et al. (1969).
336 Richard R. Burgess
10 ml
10 ml
supernatant
AS 0.55 g 0.56 g 0.58 g 0.60 g 0.62 g
% saturation 30% 40% 50% 60% 70%
Figure 20.2 How to do an ammonium sulfate precipitation test. This test is carried out
as described in the text and is self-explanatory.
Precipitation Techniques 337
quite an efficient way to determine the optimal conditions that will result in
higher enrichment in this important step.
2.4. Comments/problems/solutions
1. Pellet is not solid. If the AS pellet is not firm after centrifugation it will be
difficult to cleanly pour off the supernatant. One solution is simply to
centrifuge 50% longer such that the precipitate that sedimented to the
bottom of the tube has more time to compact. Another reason for a loose
pellet is the presence of DNA that increases viscosity and slows sedimen-
tation rates. If viscosity is a problem, it can be reduced by sonicating longer
to break cells and shear DNA into shorter pieces. Another approach is to
treat with the recombinant nuclease Benzonase (EMD/Novagen).
2. Pellet floats in high concentrations of AS. Since the density of very high
concentrations of AS approach that of protein aggregates, the AS pre-
cipitate might float rather than sediment to the bottom of the tube
during centrifugation. This can be a problem especially when your
protein contains lipid or if there are nonionic detergents around that
bind to the protein and decrease its density.
3. Published protocols are often hard to follow. Many published proce-
dures fail to indicate the AS saturation convention (saturated at 0, 20, or
25 C) or the protein concentration of the extract. As cautioned above,
the amount of AS needed to achieve a given precipitation is dependent
on the protein concentration.
4. You must interrupt an AS precipitation procedure. If you must stop the
procedure, leave the protein as an AS precipitate. Proteins are remarkably
stable in AS, either as a suspension of precipitated protein or as a pellet.
3. Polyethyleneimine Precipitation
3.1. Principles
PEI, whose trade name is Polymin P, is a basic cationic polymer made by
BASF in large quantities for use in the textile and paper industry. PEI is a
product of polymerization of ethyleneimine to yield a basic polymer with
the structure: CH3CH2N–(–CH2CH2–NH–)n–CH2CH2NH2. Typically,
n equals 700–2000 to give a molecular weight range of 30,000–90,000 Da.
Since the pKa value of the imino group is 10–11, PEI is a positively charged
molecule in solutions of neutral pH. The use of PEI in protein fractionation
originated at Boehringer Mannheim and was published by Zillig et al.
(1970). More extensive examples of its application in protein purification
338 Richard R. Burgess
and several reviews have been published (Burgess, 1991; Burgess and
Jendrisak, 1975; Jendrisak, 1987; Jendrisak and Burgess, 1975).
PEI can be thought of as similar to soluble DEAE cellulose. It binds to
negatively charged macromolecules such as nucleic acid and acidic proteins
and forms a network of PEI and bound acidic molecules that rapidly
precipitates. The binding is stoichiometric. A heavy precipitate rapidly
forms that can be pelleted by centrifugation for 5 min at 5000 rpm. Whether
an acidic protein binds to PEI depends on the salt concentration. At low salt
(0.1 M NaCl), a mildly acidic protein will bind and precipitate, but at
intermediate salt (0.4 M NaCl), it will elute from the polymer and become
soluble. A highly acidic protein will bind at low salt, not be solubilized at
intermediate salt, and will be eluted at high salt (0.9 M NaCl). It should be
noted that when a protein is eluted from a PEI pellet both the protein and
the PEI become soluble. Thus it is necessary to remove the PEI from the
protein before returning to low salt (see below).
EDTA, 0.1 mM DTT, and 0.15 M NaCl. Centrifuge out cell debris at
15,000 rpm for 15 min. All operations are carried out on ice.
3. Based on a PEI precipitation test for your system (see below) add 10%
(v/v) PEI, pH 7.9 to a final concentration of, for example, 0.3% (v/v)
and mix well for 5 min to allow formation of a dense white precipitate.
4. Centrifuge at 5000 rpm for 5 min. Note: Do not centrifuge too hard or the
pellet will be harder to resuspend. Decant the 0.3% PEI supernatant and save
for later analysis. Let the pellet drain for 1–2 min to get rid of as much of
the supernatant as possible.
5. Thoroughly resuspend the 0.3% PEI pellet in 30 ml of the above buffer
containing 0.4 M NaCl. If available, the Tissue Tearor homogenizer
(BioSpec Products, Inc. Cat # 985370-07) works very well to finely
resuspend the pellet and effectively washes out proteins physically
trapped in the pellet, and elutes out mildly acidic protein that are weakly
bound to the PEI in the pellet. Let sit 5 min and then centrifuge at
5000 rpm for 5 min and decant the 0.4 M NaCl wash.
6. Resuspend thoroughly the 0.4 M NaCl pellet in 30 ml of buffer above
but containing 0.9 M NaCl. This elutes more acidic proteins (like E. coli
RNA polymerase), but leaves the nucleic acids in the pellet (it takes
about 1.6 M NaCl to elute the nucleic acids). Let sit about 5 min, mix,
and centrifuge at 15,000 rpm for 10 min.
7. To the 0.9 M NaCl eluate, add solid AS to 60% saturation (add
3.61 g per 10 ml of eluate). Mix well and let precipitation occur for
at least 30 min. Centrifuge at 15,000 rpm for 10 min. Let pellet drain
for 5 min. The pellet contains the AS precipitated protein, but almost
all of the PEI remains in the supernatant. While traces of PEI are
trapped in the pellet, they usually do not interfere with subsequent
operations. If it is necessary to more completely remove these PEI
traces, one can resuspend the pellet in buffer containing 60%
saturated AS and recentrifuge.
This procedure typically gives a sixfold purification of RNA polymerase
from other proteins, greater than 90% recovery, and a nearly complete
removal of nucleic acids in 1–2 h.
you would only have to add 0.03% PEI to achieve the same precipitation
with the 10-fold diluted extract, but of course there would be 10 times
the volume of extract). This reflects the fact that PEI binds tightly to the
acidic components and essentially titrates them.
4. Other Methods
This chapter has focused on AS and PEI precipitation. Other methods
for protein precipitation mentioned very briefly below have been well
described in numerous publications (see, Englard and Seifter, 1990;
Ingham, 1990; Scopes, 1994).
REFERENCES
Burgess, R. R. (1991). The use of polyethyleneimine in the purification of DNA binding
proteins. Meth. Enzymol. 208, 3–10.
Burgess, R. R., and Jendrisak, J. J. (1975). A procedure for the rapid, large-scale purification
of E. coli DNA-dependent RNA polymerase involving Polymin P precipitation and
DNA-cellulose chromatography. Biochemistry 14, 4634–4638.
Burgess, R. R., and Knuth, M. W. (1996). Purification of a recombinant protein over-
produced in E. coli. In ‘‘Strategies for Protein Purification and Characterization:
A Laboratory Manual’’, (D. Marshak, J. Kadonaga, R. Burgess, M. Knuth,
W. Brennan, Jr., and S.-H. Lin, eds.), pp. 219–274. Cold Spring Harbor Press, Cold
Spring Harbor, NY.
Dawson, R. M. C., Elliot, D. C., Elliot, W. H., and Jones, K. M. (1969). In ‘‘Data for
Biochemical Research.’’ 2nd edn., p. 616. Oxford University Press, Oxford.
Duellman, S. J., and Burgess, R. R. (2008). Large-scale Epstein-Barr virus EBNA1 protein
purification. Protein Expr. Purif. 63, 128–133.
Englard, S., and Seifter, S. (1990). Precipitation techniques. Meth. Enzymol. 182, 287–300.
Ingham, K. C. (1990). Precipitation of proteins with polyethylene glycol. Meth. Enzymol.
182, 301–306.
Jendrisak, J. J. (1987). The use of PEI in protein purification. In ‘‘Protein Purification: Micro
to Macro’’, (R. Burgess, ed.), pp. 75–97. A.R. Liss, New York.
Jendrisak, J. J., and Burgess, R. R. (1975). A new method for the large-scale purification of
wheat germ DNA-dependent RNA polymerase II. Biochemistry 14, 4639–4645.
Scopes, R. K. (1994). Protein Purification, Principles and Practice. 3rd edn. pp. 7–101.
Springer-Verlag, New York.
Zillig, W., Zechel, K., and Halbwachs, H. J. (1970). A new method of large scale preparation
of highly purified DNA-dependent RNA-polymerase from E. coli. Hoppe Seylers Z.
Physiol. Chem. 351, 221–224.
C H A P T E R T W E N T Y- O N E
Contents
1. A Representative Protocol 344
Reference 345
Abstract
Passage of an extract or supernatant fraction through a column of Affi-Gel Blue
and batchwise elution can be a rapid and effective early procedure for removal
of nucleic acid, concentration of the sample and purification of nucleotide
binding proteins.
Department of Biochemistry and Molecular Biology, University of Miami School of Medicine, Miami,
Florida, USA
343
344 Murray P. Deutscher
1. A Representative Protocol
The sample to be loaded may be a low-speed (S10) or high-speed
(S100) supernatant fraction. The latter is preferable as there will be no
interference by ribosomes or microsomes. The size of the column to be
used will be determined by the amount of sample to be loaded. A useful rule
of thumb is to use 1 ml of packed gel (50–100 mesh) per 100 mg of protein
in the sample. According to the manufacturer (Bio-Rad, Bulletin 1107),
1 ml of packed Affi-Gel Blue binds 11 mg of albumin. Generally, only
5–10% of total protein binds to the columns under the conditions described,
so the suggested ratio of protein to gel should be sufficient. However, this
should be determined for each extract used by specific assay of the flow-
Early Purification with Affi-Gel Blue 345
REFERENCE
Deutscher, M. P., and Marlor, C. W. (1985). Purification and characterization of Escherichia
coli RNase T. J. Biol. Chem. 260, 7067–7071.
C H A P T E R T W E N T Y- T W O
Ion-Exchange Chromatography
Alois Jungbauer and Rainer Hahn
Contents
1. Introduction 349
2. Principle 351
3. Stationary Phases 353
4. Binding Conditions 355
5. Elution Conditions 361
6. Operation of Ion-Exchange Columns 363
7. Example: Separation of Complex Protein Mixture 366
7.1. Pretreatment of milk 366
7.2. Chromatography conditions 366
8. Example: High-Resolution Separation with a Monolithic Column 367
8.1. Mn peroxidase production 367
8.2. Chromatography conditions 367
References 370
Abstract
Ion-exchange chromatography is the most popular chromatographic method for
separation of proteins. It is a versatile and generic tool and is suited for
discovery of proteins, high-resolution purification, and industrial production
of proteins. Separation conditions are within physiological range of salt and
pH and in the most cases a native protein can be obtained. In this chapter, the
guidance will be provided for binding and elution conditions and selection of
stationary phases.
1. Introduction
Proteins, polynucleotides, and other biomacromolecules expose
charged moieties at the surface and thus they can interact with ion exchan-
gers. Ion-exchange chromatography is a versatile and generic tool for
protein and plasmid separation. It is frequently used for analytical and
Department of Biotechnology, University of Natural Resources and Applied Life Sciences, Vienna, Austria
349
350 Alois Jungbauer and Rainer Hahn
preparative purposes. From the early application to the current state of the
art of ion-exchange chromatography, the method substantially improved
over the years. In the early days of protein separation using ion exchangers,
only extremely low flow rates were applicable due to the soft nature of the
chromatography medium. Often the columns have been operated under
gravity flow and linear flow velocities of only 1–2 cm/h were obtained.
This led to purification cycles lasting several days. As a consequence, protein
purification had to be performed in the cold room to maintain the biological
activity and to prevent microbial growth. Currently, ion-exchange chro-
matography is operated at linear flow velocities of up to 500 cm/h. This is
two orders of magnitude faster than two decades ago. Whenever the
stability of a protein solution allows, ion-exchange chromatography is
performed at room temperature. A substantial step toward improvement
and acceleration of the purification cycles was achieved by introduction of
FPLC in the early 1980s ( Jungbauer, 1993). Since then, ion-exchangers
have been constantly improved ( Jungbauer, 2005). The latest jump in
development was introduction of monoliths ( Jungbauer and Hahn, 2008).
With this chromatography media it is possible to separate proteins in less
than 5 min ( Jungbauer and Hahn, 2004).
The popularity of the methods is based on the high resolution that can be
achieved with ion-exchange chromatography. A dilute protein or polynu-
cleotide solution can be rapidly concentrated and simultaneously purified.
Usually simple salt buffers are sufficient and concentrations are used in a
concentration range, where the so-called salting-in effect on proteins is
observed. This is the range where protein becomes more soluble with
increasing salt concentration. In contrast, this is not always possible with
hydrophobic interaction chromatography. With hydrophobic interaction
chromatography often salt concentrations are required which are in the
salting-out regime.
One drawback is frequently associated with ion-exchange chromatogra-
phy—this is incompatibility with mass spectrometry—especially in case of
ionization mode. The ionization of proteins and peptides is severely per-
turbed by ions.
We can define four general cases, where ion-exchange chromatography
is applied for protein and biomolecular purification and separation.
1. De novo purification: Here only limited information on the molecular
structure of the compounds is available. Often only a biological activity
of a protein solution or a cell extract is known. The task is to identify the
compounds responsible for the biological activity. In this case, intuition
is requested. Only limited rules are available and by observation of how
the biological activity behave on an ion-exchange column will guide to
selection of buffers and materials. Here, the work is really exploratory
and often several methods are interconnected to identify a protein.
Ion-Exchange Chromatography 351
2. Principle
The simplest explanation of ion-exchange chromatography of pro-
teins is based on the attraction of oppositely charged molecules. The protein
carries surface charges depending on pI and pH of the environment. The
ion exchanger is composed of a base matrix usually in the form of porous
beads to provide enough surface area for adsorption. On this base matrix,
352 Alois Jungbauer and Rainer Hahn
A
100
Ke = 0.4, z = 5
Ke = 0.2, z = 3
80
Distribution coefficient, K
Ke = 0.2, z = 5
60
40
20
0
0 100 200 300 400
Salt concentration in mobile phase (Na+), mM
B
100
Ke = 0.4, z = 5
Ke = 0.2, z = 3
Distribution coefficient, K
Ke = 0.2, z = 5
10
0.1
100 200 300
Salt concentration in mobile phase (Na+), mM
Figure 22.1 Relationship between protein retention and salt concentration in ion-
exchange chromatography: (A) linear plot and (B) log–log relationship.
3. Stationary Phases
The evolution of ion-exchange media paralleled the development of
stationary phases, which took place over the last several decades. The
driving force for development of new media has been the need for better
354 Alois Jungbauer and Rainer Hahn
5⬘-6A-3⬘
50 5⬘-6T-3⬘
5⬘-AATATATT -3⬘
5⬘-AATGTATT -3⬘
40 5⬘-GGCACGCC -3⬘
5⬘-GGCGCGCC -3⬘
5⬘-9A-3⬘
30 5⬘-9T-3⬘
5⬘-12A-3⬘
5⬘-12T-3⬘
20 5⬘-20A-3⬘
5⬘-20T-3⬘
5⬘-50A-3⬘
10 5⬘-50T-3⬘
5⬘-9A-3⬘ + 5⬘-9T-3⬘
b -lactoglobulin 5⬘-12A-3⬘ + 5⬘-12T-3⬘
5⬘-20A-3⬘ + 5⬘-20T-3⬘
0 5⬘-50A-3⬘ + 5⬘-50T-3⬘
10,000 20,000 30,000
24mer(P)
Molecular weight, MW 24mer(C)
24mer(P) + 24mer(C)
Figure 22.2 Relationship between the number of binding sites and the molecular mass
or number of charges of polyA and polyT DNAs. As comparison, the number of
binding sites of b-lactoglobulin is shown. Figure from Yamamoto et al. (2007).
4. Binding Conditions
According to the general principles of ion-exchange equilibria, it is
clear that loading of proteins should be preferably at very low salt concen-
tration (Fig. 22.1). Depending of the ligand density of the ion exchanger,
the salt concentration must be adjusted. Usually a concentration of
10–100 mM buffer concentration is recommended. This corresponds to a
conductivity of 1–4 mS/cm. Regular feedstock from bacterial culture or
356
Table 22.1 Selected cation exchangers suited for preparative protein separation
30
25
20
q = [mg/ml]
15
10
8
6 0.0
0.1
C
4
=
0.2
[m
0.3
g/
2
[M ]
m
0.4 I=
l]
0 0.5
0.6
Figure 22.3 Relationship among protein binding capacity, protein concentration, and
salt concentration in a protein solution; I is the ionic strength of the loading buffer, C is
the protein concentration in the mobile phase in equilibrium, and q is the concentration
of protein in the stationary phase.
then the protein solution is added. The supernatant is tested for the protein
and by addition of salt the protein is eluted from the column and it is again
tested in the supernatant. The advantage of microtiter plates are the multiple
experiments which can be run in parallel. Separation of liquid and solid
phase can be easily achieved in a centrifuge. Currently, chromatography
manufacturers also provide ready-to-use microtiter plates.
Several manufacturers have also designed small-scale, prepacked col-
umns, which can be operated by a syringe or a laboratory robot. With
these tiny columns it is easy to optimize binding conditions without wasting
precious proteins. Optimization of ion exchangers is very material consum-
ing, since the capacity is extremely high. These small columns such as
MediaScout MiniColumns from Atoll (Weingarten, Germany) are of
5 mm internal diameter and up to 10 mm bed height (Fig. 22.5). They
contain each up to 0.2 ml of the chromatography adsorbent and are
prepacked to defined density, corrected for each individual resin material,
in order to replicate conditions in larger columns.
360 Alois Jungbauer and Rainer Hahn
12
9
Relative abundance (%)
3 4 5 6 7 8 9 10 11
pl
5. Elution Conditions
Elution of a bound protein can be effected by four modes:
1. Linear salt gradient
2. Step salt gradient
3. pH gradient
4. Displacement development.
Sometime combination between pH gradient and salt is used, but this is
difficult to optimize and robust conditions often are not achieved. Linear
salt and pH gradients are preferred for high-resolution separation, while step
salt gradients are preferred for concentration of proteins. Displacement
chromatography is in principle suited for separation of related species, but
due to the more complicated optimization, is less used in general.
The retention window of proteins is very small; at low salt concentration
proteins are tightly bound and then with increasing salt concentration not
retained at all; compare (Fig. 22.1) the relationship between k0 and salt
concentration in the mobile phase. So in ion-exchange chromatography
linear gradients and even more complex segmented gradients are used.
The resolution Rs
tA tB
Rs ¼ 2 ð22:2Þ
WA þ WB
with the retention time of two different compounds (tA, tB) and the peak
width at the base (WA, WB) depends on the slope of the gradient b. Where
b is defined as
CM CM
0
b¼ ð22:3Þ
tG
with C 0M and CM as the start and final salt concentration of the buffer, and
tG the gradient time. The retention of a compound depends in a complex
manner on the normalized gradient slope (g). A detailed discussion would
exceed the scope of this chapter. A detailed analysis can be found in the
book published by Yamamoto et al. (1988):
bL bV0
g¼ ¼ ; ð22:4Þ
v Q
362 Alois Jungbauer and Rainer Hahn
L is the length of the column and v is the interstitial velocity, V0 the void
volume, and Q is the flow rate.
Resolution is increased with the normalized gradients (g). From
Eq. (22.4) one can infer three possibilities to increase resolution.
1. Increasing the gradient length
2. Increasing the column length/volume
3. Decreasing the interstitial velocity/flow rate.
All three parameters can be varied at once. Equation (22.4) also helps to
transfer a successful elution condition to a larger column. The same resolu-
tion is obtained when g is kept constant.
The resolutions in these cases are 5.00 and 4.97 min, respectively, with a
difference ion scale by a factor of 7 (Fig. 22.6).
As a consequence of the interrelation between protein retention and g, it
has been observed that the peak salt concentration at shallower gradients is
lower compared with steeper gradients. So the protein does not always elute
at the same salt concentration. Resolution depends on g in a reciprocal
manner.
rffiffiffiffiffiffiffiffiffiffiffiffiffi
1 L
Rs / ð22:5Þ
g HETP
0.014
CV = 7.99 ml
0.012 CV = 55.61 ml
0.010
UV 280 nm
0.008
0.006
0.004
0.002
0.000
0 500 1000 1500 2000 2500 3000
Time [s]
counter ion is more retained than the species with the same charge. Thus a
wave with different ions is traveling through the column. To maintain
electroneutrality, the pH changes. A detailed explanation of this phenome-
non can be found at Pabst and Carta (2007). Especially for weak ion-
exchange ligands, a larger volume is required to reach the actual pH.
Loading of the protein solution should be performed at the same pH and
conductivity as the equilibration buffer. Often this is not possible; especially
the composition of the ions in a crude feedstock and the equilibration buffer
may be completely different. Again as a consequence, a pH transition will be
observed during loading. When the protein solution is desalted by size
exclusion chromatography, dialysis, or diafiltration, the same ion composi-
tion can be obtained leading to most robust loading conditions.
After loading the protein solution, the unbound material is washed out.
A first try is always the equilibration buffer. To obtain a higher purity in the
eluate, the wash procedure may consist of several steps; for a tightly bound
protein of interest, washing with equilibration buffer with salt added is often
used. A lot of protocols exist and it seems very protein specific which
compounds are added to a washing buffers. These may be detergents,
urea, sugars, ethylene glycol, etc.
Elution can be effected by a linear or step salt gradient or by displace-
ment development. It is best to start with a linear salt gradient. The gradient
Ion-Exchange Chromatography 365
volume should be about 10 column volumes. The two buffers for the
preparation of the gradient are in the simplest case the equilibration buffer
(Buffer A) and the buffer used for loading with counterions (Buffer B).
In most cases, the second half of the gradient is not required, since most
proteins elute from a conventional ion exchanger in the range between
150 and 500 nM salt. So gradients to maximal 50% B are often sufficient.
The step gradient can also be generated from Buffers A and B. It should be
taken into consideration that an ideal step never can be achieved in a
column. First, due to retention of salt and the dead volume in HPLC
often denoted as dwell time and secondly, due to the washing-out char-
acteristics of a buffer mixer only a sigmoidal-shaped curve will be obtained
(Kaltenbrunner and Jungbauer, 1997). According to the selection of the salt
concentration in the step elution, the protein will migrate with the salt front or
travel behind the salt front. Yamamoto et al. (1988) named this behavior type I
or II elution. When a concentrated eluate is washed then salt concentration
must be selected for type I elution. This is a salt concentration somewhat higher
than the corresponding peak salt concentration in linear gradient elution. For
the type I elution less than one column volume buffer is required.
Elution by pH gradient is more difficult to optimize. So the best start is
to find a condition according to the pI of the protein where the protein does
not bind to the ion exchanger. This is below the pI for an anion exchanger
and above the pI for a cation exchanger. So pH has to be decreased or raised.
Keep in mind that the ion exchanger behaves like an acid or a base and the
pH gradient produced follows a titration curve. It is extremely difficult to
produce a linear pH gradient. This can be done by reacting boronate buffers
with diols such as mannitol (Kaltenbrunner et al., 1993). The fourth possi-
bility is the elution with induced pH gradients. When working with buffer
of low molarity in the range of 10 mM, the application of salt induces a pH
gradient due to the differential migration of the different salt species. Such
an induced pH gradient can be used for high resolution and additional for
focusing of the protein peak (Pabst et al., 2008a,b). In this case also, several
column volumes of buffer are required.
For regeneration of the column, the optimal condition is to use the same
buffer as for loading with salt. This is only possible when a relatively clean
feedstock is processed. Then it is possible to combine two steps into one.
Especially when lipids, endotoxins, and other sticky compounds are loaded
onto an ion-exchange column, then the regeneration with a caustic agent,
preferably up to 1 M NaOH, is recommended. The maximum concentra-
tion of NaOH depends on the resistance of the stationary phase. One
column volume is sufficient but for very tough impurities the column
should be soaked in NaOH. So after pumping one column volume over
the column, flow is stopped and NaOH is left several hours in the column.
The NaOH treatment also has some sanitization effects. It will definitely
reduce the bacterial count in a column, but will not sterilize a column.
366 Alois Jungbauer and Rainer Hahn
OD 280 nm
mS/cm
#1
1.0 40
0.5 20
0.0 0
0 50 100 150 700 750 800 850
B ml
2.0 80
S-sepharose FF #4
#3
1.5 60
#2
OD 280 nm
mS/cm
#1
1.0 40
0.5 20
0.0 00
0 50 100 150 200 650 700 750 800 850
C ml
2.0 80
#3
Fractogel EMD (S)
#4
1.5 #2 60
OD 280 nm
#1
mS/cm
1.0 40
0.5 20
0.0 0
0 50 100 150 200 650 700 750 800 850
D ml
2.0 80
Macro-prep high S #4
#2 # 5 60
1.5 #3
OD 280 nm
mS/cm
#1
1.0 40
0.5 20
0.0 0
0 50 100 150 500 550 600 650
ml
A
400 400
#1 #2
300 la la 300
AU 214 nm
AU 214 nm
200 lgG 200
100 100
0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
B t (min) t (min)
400 400
#1 #2 la
300 300
AU 214 nm
AU 214 nm
la lg
200 200
lgG
100 100
0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
t (min) t (min)
C
400 400
#1 la #2 la
300 300
AU 214 nm
AU 214 nm
lg
200 200
lgG
100 100
0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
t (min) t (min)
D
400 400
lg
#1 #2 la
300 300
AU 214 nm
AU 214 nm
100 100
0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
t (min) t (min)
160 6.5
MnP1 409 nm
140 280 nm
pH gradient
Relative absorbance (409 nm)
6
120
100
pH value (−)
5.5
80
MnP2 5
60
40
MnP4 4.5
3
20 5
0 4
0 0.5 1 1.5 2
Time (min)
This very fast method allows separation of isoforms and two other forms
in less than 2 min. The method is suited for preparative and analytical
applications.
REFERENCES
Afeyan, N. B., Gordon, N. F., Mazsaroff, I., Varady, L., Fulton, S. P., Yang, Y. B., and
Regnier, F. E. (1990). Flow-through particles for the high-performance liquid chro-
matographic separation of biomolecules: Perfusion chromatography. J. Chromatogr. 519,
1–29.
Boardman, N. K., and Partridge, S. M. (1955). Separation of neutral proteins on ion-
exchange resins. Biochem. J. 59, 543–552.
Brooks, C. A., and Cramer, S. M. (1992). Steric mass-action ion exchange: Displacement
profiles and induced salt gradients. AIChE J. 38, 1969–1978.
Hahn, R., Schulz, P. M., Schaupp, C., and Jungbauer, A. (1998). Bovine whey fractionation
based on cation-exchange chromatography. J. Chromatogr. A 795, 277–287.
Janson, J.-C., and Ryden, L. (1998). Protein Purification: Principles, High Resolution
Methods, and Applications 2nd edn. Wiley-VCH, New York.
Jungbauer, A. (1993). Preparative chromatography of biomolecules. J. Chromatogr. 639,
3–16.
Jungbauer, A. (2005). Chromatographic media for bioseparation. J. Chromatogr. A 1065,
3–12.
Ion-Exchange Chromatography 371
Jungbauer, A., and Hahn, R. (2004). Monoliths for fast bioseparation and bioconversion and
their applications in biotechnology. J. Sep.. Sci. 27, 767–778.
Jungbauer, A., and Hahn, R. (2008). Polymethacrylate monoliths for preparative and
industrial separation of biomolecular assemblies. J. Chromatogr. A 1184, 62–79.
Jungbauer, A., and Lettner, H. (1994). Chemical disinfection of chromatographic resins,
Part 1: Preliminary studies and microbial kinetics. BioPharm 7, 46–56.
Jungbauer, A., Lettner, H., Gurrier, L., and Boschetti, E. (1994). Chemical sanitization in
process chromatography, Part 2: In situ treatment of packed columns and long-term
stability of resins. BioPharm 7, 37–42.
Kaltenbrunner, O., and Jungbauer, A. (1997). Simple model for buffer blending in ion-
exchange chromatography. J. Chromatogr. A 769, 37–48.
Kaltenbrunner, O., Tauer, C., Brunner, J., and Jungbauer, A. (1993). Isoprotein analysis by
ion exchange chromatography using a linear pH gradient combined with a salt gradient.
J. Chromatogr. 639, 41–49.
Kopaciewicz, W., Rounds, M. A., Fausnaugh, J., and Regnier, F. E. (1983). Retention
model for high-performance ion-exchange chromatography. J. Chromatogr. A 266, 3–21.
Mueller, E. (2005). Properties and characterization of high capacity resins for biochromato-
graphy. Chem. Eng. Technol. 28, 1295–1305.
Necina, R., Amatschek, K., and Jungbauer, A. (1998). Capture of human monoclonal
antibodies from cell culture supernatant by ion exchange media exhibiting high charge
density. Biotechnol. Bioeng. 60, 689–698.
Pabst, T. M., and Carta, G. (2007). pH transitions in cation exchange chromatographic
columns containing weak acid groups. J. Chromatogr. A 1142, 19–31.
Pabst, T. M., Carta, G., Ramasubramanyan, N., Hunter, A. K., Mensah, P., and
Gustafson, M. E. (2008a). Separation of protein charge variants with induced pH
gradients using anion exchange chromatographic columns. Biotechnol. Prog. 24,
1096–1106.
Pabst, T. M., Antos, D., Carta, G., Ramasubramanyan, N., and Hunter, A. K. (2008b).
Protein separations with induced pH gradients using cation-exchange chromatographic
columns containing weak acid groups. J. Chromatogr. A 1181, 83–94.
Podgornik, H., and Podgornik, A. (2004). Separation of manganese peroxidase isoenzymes
on strong anion-exchange monolithic column using pH-salt gradient. J. Chromatogr.
B Analyt. Technol. Biomed. Life Sci. 799, 343–347.
Tennikova, T. B., Belenkii, B. G., and Svec, F. (1990). High-performance membrane
chromatography. A novel method of protein separation. J. Liq. Chromatogr. 13, 63–70.
Yamamoto, S., Nakanishi, K., and Matsuno, R. (1988). Ion-exchange chromatography of
proteins Marcel Dekker, New York.
Yamamoto, S., Nakamura, M., Tarmann, C., and Jungbauer, A. (2007). Retention studies of
DNA on anion-exchange monolith chromatography: Binding site and elution behavior.
J. Chromatogr. A 1144, 155–169.
C H A P T E R T W E N T Y- T H R E E
Gel Filtration1
Earle Stellwagen
Contents
1. Principle 373
2. Practice 374
2.1. Matrices 376
2.2. Sample preparation 380
2.3. Chromatographic solvents 380
2.4. Preliminary screening 380
2.5. Chromatography using conventional matrix 382
2.6. Scaling upward 383
2.7. Trouble shooting 384
2.8. Further information 385
1. Principle
Gel filtration is performed using porous beads as the chromatographic
support. A column constructed from such beads will have two measurable
liquid volumes, the external volume, consisting of the liquid between the
beads, and the internal volume, consisting of the liquid within the pores of
the beads. Large molecules will equilibrate only with the external volume
while small molecules will equilibrate with both the external and internal
volumes. A mixture of proteins is applied in a discrete volume or zone at the
top of a gel-filtration column and allowed to percolate through the column.
Department of Biochemistry, University of lowa, lowa City, lowa
1
Reprinted from Methods in Enzymology, Volume 182 (Academic Press, 1990)
373
374 Earle Stellwagen
The large protein molecules are excluded from the internal volume and
therefore emerge first from the column while the smaller protein molecules,
which can access the internal volume, emerge later.
The dimensions important to gel filtration are the diameter of the pores
that access the internal volume and the hydrodynamic diameter of the protein
molecules. The latter is defined as the diameter of the spherical volume
created by a protein as it rapidly tumbles in solution. Proteins whose
hydrodynamic diameter is small relative to the average pore diameter of
the beads will access all of the internal volume and are described as being
included in the gel matrix. Proteins whose hydrodynamic diameter is compa-
rable to the average pore diameter will access some but not all of the internal
volume and are described as being fractionally excluded. Proteins whose
hydrodynamic diameter is large relative to the average pore diameter will be
unable to access the internal volume and are described as being excluded.
This conceptualization has led to the gradual renaming of gel filtration as
size-exclusion chromatography. The order of elution of a mixture of
proteins from a size-exclusion column will then be the inverse of their
hydrodynamic diameters. If all the proteins in a mixture are known, or can
be assumed to have the same shape, then the order of elution will be the
inverse of their molecular weights. This discussion will treat protein dimen-
sions in terms of molecular weight since common usage assumes that
protein mixtures contain only globular proteins. However, the reader
should bear in mind that hydrodynamic volume is the operative protein
dimension and that an asymmetrical protein will appear to elute with an
abnormally high-molecular weight compared with globular proteins of
similar molecular weight.
2. Practice
An elution profile obtained by size-exclusion chromatography is illu-
strated in Fig. 23.1A. Zero elution volume is defined as the entry of the
sample into the chromatographic support. The elution volume for the
excluded component is designated V0 for the void volume, which repre-
sents the volume external to the beads. The elution volume for the included
component is designated Vt for the total volume, which represents the sum
of the external volume and the internal volume within the beads. Elution
volumes intermediate between these values are designated Ve. A partition
coefficient, designated Kav, relating these values is given in Eq. (1):
Ve V0
Kav ¼ ð1Þ
Vt V0
Gel Filtration 375
A Volume (ml)
0 20 40 60 80
Vt
Ve
V0
Al
1 2 3
B
1.0
0.8
0.6
Kav
0.4
0.2
0.0
1 10 100 1000
Molecular weight (10−3)
2.1. Matrices
The properties of some conventional and high-performance size-exclusion
matrices are given in Tables 23.1–23.4. It should be noted that suppliers use
a variety of terms and abbreviations to index these products in their catalogs,
including gel-filtration chromatography (GFC), gel-permeation chroma-
tography (GPC), and size-exclusion chromatography (SEC).
The conventional matrices are distinguished by their relative economy
and slow flow rates. These matrices are available in bulk, requiring an
investigator to pour columns of any desired dimensions to accommodate
the volume of the sample to be chromatographed. The flow rates normally
used for chromatography are obtained by multiplying the linear flow rate
listed in Table 23.2 by the cross-sectional area of the column in cm2 to yield
the flow rate in mm/h. A column can be packed with a flow rate approxi-
mately five times that used during chromatography.
The high-performance matrices are distinguished by their convenience,
rapid flow rates, and expense. These matrices are usually purchased as
poured columns which are attached to an existent high-performance chro-
matograph available to the investigator. The smaller analytical columns,
about 8 300 mm, are normally loaded with not more than a few milli-
grams of protein and operated at a flow rate of about 1 ml/min. The larger
preparative columns generally contain beads having a diameter of 30 mm.
The approximately 20 300 mm columns can be loaded with between
10 and 100 mg of protein and can be operated at a flow rate of about
5 ml/min while the very large columns can be loaded with up to 2 g of
protein and be operated at a flow rate of up to 30 ml/min.
Table 23.1 Matrix parameters
Stability
Name Supplier Chemistrya Form supplied pH Temperature ( C) Bead diameter (mm)
Conventional
BioGel A Bio-Rad AG Suspension 4–13 1–30 40–300b
BioGel P Bio-Rad PA Powder 2–10 1–80 40–30b
Sephacryl HR Pharmacia DX Suspension 2–13 1–100 25–75
Sephadex G Pharmacia DX/PA Powder 2–10 1–100 20–300b
Sepharose Pharmacia AG Suspension 4–10 1–40 45–200b
Ultrogel A IBF AG Suspension 3–10 2–36 60–140
Ultrogel AcA IBF AG/PA Suspension 3–10 2–36 60–140
High performance
Protein Pak Waters S Packed column 2–8 1–90 10
Shodex Showa Denko S Packed column 3–7.5 10–45 9
Superose Pharmacia AG Packed column 1–14 4–40 10–13
SynChropak SynChrom S Packed column 2–7 1–60 5–10
TSK-SW Toyo-Soda S Packed column 3–7.5 1–45 10–13
Zorbax DuPont S Packed column 3–8.5 1–100 4–6
a
The following symbols are used to denote the chemical nature of the matrix: AG, cross-linked agarose; PA, cross-linked polyacrylamide; DX, cross-linked dextran;
DX/DA, copolymer of allyl dextran and bisacrylamide; AG/PA, mixture of agarose and polyacrylamide; S, bonded silica.
b
Individual matrices have narrower ranges.
Table 23.2 Powdered matrix parameters
Protein Chromatography
on Hydroxyapatite Columns
Larry J. Cummings, Mark A. Snyder, and Kimberly Brisack
Contents
1. Introduction 388
2. Mechanisms 389
3. Chemical Characteristics 392
3.1. Cationic and anionic modification of HA surfaces 392
3.2. Metal adsorption 393
3.3. HA solubility 394
4. Purification Protocol Development 396
5. Packing Laboratory-Scale Columns 397
6. Process-Scale Column Packing 399
7. Applications 401
References 402
Abstract
The introduction of spherical forms of hydroxyapatite has enabled protein
scientists to separate and purify proteins multiple times with the same packed
column. Biopharmaceutical companies have driven single column applications
of complex samples to simpler samples obtained from upstream column purifi-
cation steps on affinity, ion exchange or hydrophobic interaction columns.
Multiple column purification permits higher protein loads to spherical forms
of hydroxyapatite and improved reduction in host cell protein, aggregates,
endotoxin, and DNA from recombinant proteins. Adsorption and desorption
mechanisms covering the multimodal properties of hydroxyapatite are dis-
cussed. The chemical interactions of hydroxyapatite surface with common
ions, metals, and phosphate species affect column lifetimes. Adsorbed hydrox-
onium ions from low ionic strength buffers are noted by a shift in effluent pH
during column equilibration. Hydroxonium ion desorption is observed by acidic
shifts in the column effluent with the magnitude and duration surprisingly
extreme. Buffering reagents with high buffering capacity reduce both the
magnitude and duration of the acidic shift. Column packing methods for the
387
388 Larry J. Cummings et al.
1. Introduction
The use of hydroxyapatite, a hydroxylated calcium phosphate mate-
rial, in columns for protein chromatography has increased substantially
between 1991 and 2009 primarily due its use for recombinant protein
purification. The methods for using hydroxyapatite (HA) follow those
discussed by Tiselius et al. (1956) and reviewed by Gorbunoff (1985).
Although microcrystalline HA is still manufactured in production-scale
quantities, it is generally used in batch mode operations rather than in
packed columns because the crystals are mechanically unstable in applica-
tions requiring multiple column cycles. Suppliers of HA developed
manufacturing processes to make spherical particles that allow multiple
column cycles. One process forms microcrystalline HA in agarose gel
particles, HA UltrogelÒ sorbent as described in French patent 2231422,
British patent 1468592, Swiss patent 597893, and German patent 2426501.
Another described by Thompson and Miles (1973) forms large spheroids,
approximately 50–600 mm in diameter utilizing microcrystalline HA and
a ceramic hardening process. The spheroidal ceramic particles were com-
mercialized in the early 1970s by BDH Ltd. The adsorption capacity for
protein on agarose–HA is similar if not lower than the microcrystalline HA.
The elution profile of bovine serum albumin is similar between the sphe-
roidal ceramic particles and microcrystalline HA. A granular, irregularly
shaped form of porous ceramic HA is produced by Clarkson Chromatogra-
phy. The most commonly used porous spherical ceramic HA for down-
stream processing of recombinant proteins (CHTTM Ceramic
Hydroxyapatite, Bio-Rad Laboratories). The spherical particles are manu-
factured by spray drying nanoparticles to obtain porous particles in 20, 40,
or 80 mm sizes. The particles are mechanically sized to obtain a narrow size
distribution. Mechanical stability and porosity is imparted to the sized
particles by sintering them at high temperatures. Packed beds of these
particles have 12–20 times the surface area of microcrystalline HA particles
as discussed by Kadoya et al. (1986) and Ichitsuka et al. (1992). The increase
in surface area results in increased protein adsorption to the commercialized
porous CHTTM Types I and II. Type I adsorbs greater than 25 mg of
lysozyme per dry gram of resin from a 10-mM phosphate buffer, pH 6.8
and Type II between 12 and 19 mg/g due to its lower surface area. This
relationship was confirmed by Zhu et al. (2009) studying protein adsorption
to porous and nonporous bioceramics.
Protein Chromatography on Hydroxyapatite Columns 389
2. Mechanisms
The mechanism of protein adsorption and desorption to and from HA
surfaces has been reviewed periodically since 1971 (Bernardi, 1971;
Gorbunoff, 1990). A recent contribution cites (Kandori et al., 2004) earlier
mechanisms in explaining that acidic proteins bind through C (calcium)-
sites while basic proteins bind through P (phosphate)-sites. C- and P-sites
were previously described by Kawasaki et al. (1986) and later further
discussed by Gagnon (1996) with an emphasis on the importance of
C-sites in the purification of monoclonal antibodies; monoclonal antibody
binding occurs in part through calcium coordination complexes with
carboxyl clusters in the antibody. The mechanisms of HA–protein interac-
tions were simplified to the following description. Amino groups are
adsorbed to P-sites but repelled by C-sites. The situation is opposite and
more complex for carboxyl groups. Although amine binding to P-sites and
the initial attraction of carboxyl groups to C-sites are electrostatic, the
binding of carboxyl groups to C-sites involves formation of much stronger
chelation, coordination bonds than anion exchange. Phosphoryl groups on
proteins and other solutes interact more strongly with C-sites than do
carboxyl groups as discussed by Kawasaki (1991).
Further work has involved bone–protein interactions by Dong et al.
(2007) to study the mechanistic role of zero net charge –OH and –NH2
groups in protein for the surface of HA. Steered molecular dynamic simu-
lation demonstrated the strong interaction between the HA surface and
protein –OH and –NH2 groups of bone morphogenetic proteins (growth
factors for bone formation) through a water-bridged H-bond. Their obser-
vations were confirmed by Shen et al. (2008) using a fibronectin fragment
(FN-III10). These and similar investigations support the C- and P-site
model. Hydroxyapatite surfaces have been proposed to interact differentially
with hexahistidine tagged proteins and histidine-rich regions of
390 Larry J. Cummings et al.
3. Chemical Characteristics
3.1. Cationic and anionic modification of HA surfaces
Calcium and magnesium cations change the surface of HA through the
formation of a phosphate–Ca or Mg bridge. Although initially commented
upon over 20 years ago, Gorbunoff (1984a,b) and Gagnon et al. (2009) have
recently used Ca-modified ceramic HA surfaces to improve the purification
of Fab from Mab and Fc antibody components. These workers showed that
Ca-modified ceramic HA is restored to its native form after treatment with
20 column volumes of dilute phosphate buffer.
HA also has a high affinity for pyrophosphate (PPi) and polyphosphates
as explained by Krane and Glimcher (1962). PPi is a contaminant in
anhydrous salts of mono- and disodium phosphate and may be a contami-
nant in its potassium salt counter parts. There are advantages and disadvan-
tages to PPi in phosphate buffers relative to the adsorption and desorption of
proteins to the HA. Figure 24.1 shows the separation of bovine serum
albumin, ovalbumin, a-chymotrypsinogen A, and cytochrome c with
1.0 250
0.5 2 3 100
ODA280
mS/cm
0.0 0
Figure 24.1 Separation of protein mixture using phosphate buffers prepared from
anhydrous phosphate salts: (1) bovine serum albumin, (2) ovalbumin, (3) a-chymo-
trypsinogen, and (4) cytochrome c. Column: 10 mm 40 mm. Flow rate: 2.5 ml/min.
Temperature: 20 C. Program: sample inject, 5-min isocratic A; 20-min 0–80% linear
gradient to B; 10-min isocratic 80% B; 10-min isocratic 100% B; 15-min isocratic
A. Buffer A: 10 mM sodium phosphate, pH 6.8. Buffer B: 400 mM sodium phosphate,
pH 6.8. Reagents: sodium phosphate monobasic anhydrous and sodium phosphate
dibasic anhydrous. Dashed line indicates the relative conductivity of the column effluent.
Protein Chromatography on Hydroxyapatite Columns 393
1.0 250
1
0.5 2 100
ODA280
mS/cm
4
0.0 0
Figure 24.2 Separation of protein mixture using phosphate buffers prepared from
hydrated phosphate salts: (1) bovine serum albumin, (2) ovalbumin, (3) a-chymotryp-
sinogen, and (4) cytochrome c. Column: 10 mm 40 mm. Flow rate: 2.5 ml/min.
Temperature: 20 C. Program: sample inject, 5-min isocratic A; 20-min 0–80% linear
gradient to B; 10-min isocratic 80% B; 10-min isocratic 100% B; 15-min isocratic
A. Buffer A: 10 mM sodium phosphate, pH 6.8. Buffer B: 400 mM sodium phosphate,
pH 6.8. Reagents: sodium phosphate monobasic monohydrate and sodium phosphate
dibasic, heptahydrate. Dashed line indicates the relative conductivity of the column
effluent.
394 Larry J. Cummings et al.
integrity and tissue fluids for artificial bone or cadaver implants, and material
scientists examining soil remediation and metals absorption.
According to Shepard et al. (2000), a column of ceramic HA discolored
after several recombinant protein purification cycles. The column of
ceramic HA was the third column in the process following cation and
anion exchange. They concluded that the metals that discolored the ceramic
HA came from various sources: process equipment, process reagents, pro-
cess water, and fermentation nutrients. However, the discoloration did not
impair the purification of their recombinant protein. Agronomists explain
the preferential absorption of multivalent metals to HA (Ma et al., 1994) and
show that HA has a high capacity for divalent heavy metals. Many of the
described metals were observed in the discolored ceramic HA column,
especially Fe, Al, Zn, and Mn.
3.3. HA solubility
Voids in the column headspace or the formation of channels can occur with
HA resulting from the massive amounts of buffer applied to the column.
The solubility constant for HA is 2.4 10 59 which is equivalent to about
4 ppm. Phosphate ion and basic pH suppress the solubility. However,
adsorbing protein to HA from very dilute phosphate buffer, 1 mM
(pH 6.8), in unbuffered alkali metal solutions, 1 mM NaCl or KCl or in
zwitterionic buffers is common. Some acidic proteins absorb only in water
as determined by Schirch et al. (1985). The low to unbuffered loading
conditions reduce the control over the surface pH of the HA as explained
by Scopes (1993). Some researchers (Schroder et al., 2003) determined that
cobuffers such as 4-morpholinepropanesulfonic acid (MES) could be used
to permit phosphate elution while maintaining the pH of the loading and
elution buffers. However, HA solubility is a risk even with phosphate
suppression. For example, one column was equilibrated with 20 mM
sodium phosphate, pH 6.5, loaded with protein applied in the equilibration
buffer then sequentially eluted with increased sodium chloride buffered
with 20 mM sodium phosphate. In this case, voids formed and the column
displayed increasing pressure from cycle to cycle. The voids and increased
pressure were caused by chemical damage to the HA largely due to the
liberation of hydroxonium ion that accumulated on the surface of the HA.
Figure 24.3 shows the pH profile of the column effluent and the zones
where hydroxonium ion adsorbed (A and B) and desorbed (C and D).
Harding et al. (2005) discuss the adsorption and desorption of hydro-
xonium ion as a function of the zero point charge of HA. The results of
these research (Skartsila and Spanos, 2007) show the relationship of zero
point charge, pH, hydroxonium adsorption, desorption, and calcium
content. The zero point charge determined by Skartsila and Spanos
differs, 7.3 0.1 and 6.5 0.2. These research contributions explain the
Protein Chromatography on Hydroxyapatite Columns 395
8.50
400
8.00
7.50 A B C D 300
7.00
mS/cm
200
pH
6.50
6.00
100
5.50
0
1 2 3 4 5 6 7 8 9 10
0 60 120 180
Min
Fig. 24.3 Effluent pH changes relative to buffer sequence. Column 11 mm 230 mm.
Flow rate: 7.0 ml/min. Temperature: 20 C. Program: (1) 3-min 20 mM sodium
phosphate, pH 6.5 to rinse 0.1 M NaOH from the column; (2) 17-min pH adjustment
with 500 mM sodium phosphate/0.1 M NaCl, pH 6.5; (3) 29-min 20 mM sodium
phosphate/low concentration NaCl, pH 6.5; (4) 29-min 20 mM sodium phosphate, pH
6.5; (5) 34-min 20 mM sodium phosphate/low concentration NaCl, pH 6.5; (6) 18-min
20 mM sodium phosphate/low concentration NaCl, pH 6.5; (7) 3-min 20 mM sodium
phosphate to rinse calcium ion from the column; (8) 29-min regeneration with 500 mM
sodium phosphate/0.1 M NaCl, pH 6.5; (9) 3-min 20 mM sodium phosphate to rinse
concentrated phosphate from the column; (10) 29-min sanitization with 0.5 N NaOH.
Points A and B indicate the change to a more basic effluent than the pH 6.5 input buffer.
Points C and D indicate the change to more acidic effluent than the pH 6.5 input buffer.
Dashed line indicates the relative conductivity of the column effluent. The solid line
indicates the pH of the column effluent.
500
8.50
400
8.00
7.50 A B C D 300
7.00
mS/cm
pH
200
6.50
6.00
100
5.50
0
1 2 3 4 5 6 7 8 9 10
0 60 120 180
Min
Figure 24.4 Effluent pH changes relative to buffer sequence. The buffer sequence is
the same as in Fig. 24.3 except that buffers 1, 3, 4, and 5 contain 75 mM MES. Dashed
line indicates the relative conductivity of the column effluent. The solid line indicates
the pH of the column effluent.
gradient as eluant. Test the protein for purity. Determine the concentration
of the salt or phosphate necessary to elute the protein and apply it as a
desorption step in the purification protocol. Compare the purity of the step
desorbed protein to the gradient desorbed protein. Make adjustment to the
desorption protocol if necessary. Desorb other bound proteins and
biological components from the column by rinsing with 3 column volumes
of phosphate-containing, salt-free, low ionic strength (LIS) buffer followed
by 3–5 column volumes of 0.5 M phosphate buffer, pH 7, 3 column
volumes LIS and 3–5 column volumes of 0.5–1 M NaOH. Prepare the
column for subsequent column cycles by rinsing with 3 column volumes
LIS, 3 column volumes of 0.5 M phosphate buffer at the pH of the
equilibration buffer then with 3 column volumes load buffer.
of the media retention plates (frits or nets). For example, if the distance is
54 cm, the maximum packed height is 27 cm. Calculate the volume of the
packed column. For each liter of packed bed volume, use 630 g of dry
powder and 1.79 l of packing buffer to prepare a 50% (v/v) slurry. With the
column outlet closed, dispense the packing buffer to the column followed
by the dry powder. Agitate the CHT-buffer mixture with a plastic paddle to
hydrate the powder and blend the components to a homogenous slurry.
Reverse the direction of agitation to minimize the motion of the slurry.
Assemble the top adapter following the manufacturer’s instruction and
insert into the column tube. Wait 5 min to allow for a resin free zone
then lower the adapter allowing air to vent through the top process inlet and
to purge the adapter flow distributor and inlet line with packing buffer.
Flow pack at 200–300 cm/h with 2 column volumes of packing buffer.
Once the bed is consolidated, lower the adapter, leaving a headspace of
1–5 mm between the media retention plate and the top of the packed bed.
Do not lower the adapter into the packed bed to avoid irreversible damage
to the CHT particles.
Closed columns such as the InPlaceTM (Bio-Rad Laboratories) and Bio-
ProcessTM LPLC (GE Healthcare) columns are packed with externally
prepared slurries. As with open columns, the maximum packed bed height
cannot be greater than 50% of the height between the surfaces of the media
retention plates (frits). Calculate the volume of the packed column. For each
liter of packed bed volume, use 630 g of dry powder and 1.79 l of packing
buffer to prepare a 50% (v/v) slurry. Dispense the packing buffer to the media
slurry tank followed by the dry powder. Agitate the CHT-buffer mixture to
with a low-shear hydrofoil impeller (Type A3) to hydrate the powder and
blend the components to a homogenous slurry. Transfer all of the slurry to the
column using a media transfer device. Apply up flow at 50 cm/h to expel air
bubbles from the transferred slurry. Stop the flow and wait 5 min to allow a
resin free zone to form. Lower the adapter, allowing air to vent through the
top process inlet and to purge the adapter flow distributor and inlet line with
packing buffer. Axially, flow pack at 200–300 cm/h to consolidate the packed
bed. Continue lowering until the adapter media retention plate is 1–5 mm
above the surface of the packed bed. When packing stainless steel InPlace or
LPLC columns axial flow pack until the adapter reaches 2 cm above the target
height of the column then lower at 10 cm/h until the adapter sensor signals
contact with the top of the bed.
Process-scale columns utilizing microcrystalline HA is not achievable.
HA Ultrogel is utilized in shallow bed columns with large diameters.
Packing this type of column and media requires assistance from the column
and media suppliers.
The evaluation of packed process columns is often required by end users
per the recommendation of regulatory agencies. The United States Food
and Drug Administration and its counterparts in Europe, Canada, Japan,
Protein Chromatography on Hydroxyapatite Columns 401
7. Applications
Phosphate gradient or step elution remains a dominant desorbing
strategy in recombinant purification protocols. Recombinant human cata-
lase expressed in Pichia pastoris was purified to 95% in a three-step process—
ammonium sulfate precipitation, anion-exchange chromatography, and
hydroxyapatite chromatography. The recombinant catalase was desorbed
from the HA column with a linear gradient of 0.05–0.3 M phosphate,
pH 6.8. Shi et al. (2007) obtained a 60% yield of the catalase that was
secreted into the culture medium. Numerous proteins purified by chroma-
tography on HA columns utilize phosphate linear or step gradients as
described by Hsieh et al. (2003), Stránská et al. (2007), Luellau et al.
(1998), and Nuss et al. (2008).
Clones of cDNA for human and two E. coli pyridoxal kinase genes were
used by di Salvo et al. (2004) to express pyridoxal kinases in E. coli. The
ammonium sulfate precipitated cell lysate pellet containing the enzyme was
dissolved with phosphate buffer then purified using three types of chroma-
tography media—hydrophobic interaction gel, anion-exchange resin, and
ceramic HA. The HA purification step for the human pyridoxal kinase
utilized a 20-mM sodium N,N-bis(2-hydroxyethyl)-2-aminoethanesulfonic
acid (BES), pH 7.3 adsorption buffer and a linear gradient to 100 mM
potassium phosphate, pH 7.3 to desorb the enzyme. Methyl jasmonate
hydrolyzing esterase was purified from cell cultures of Lycopersicon esculentum
by Stuhlfelder et al. (2002) using anion-exchange chromatography,
gel-filtration, and chromatography on ceramic HA. The HA step utilized
a 20-mM potassium phosphate, 20-mM b-mercaptoethanol, 0.3-mM
calcium chloride, pH 6.8 adsorption buffer and a linear gradient to
500 mM potassium phosphate, 20 mM b-mercaptoethanol, pH 6.8 to
desorb the esterase. Recombinant erythropoietin (rEPO) is manufactured
with a multistep chromatography process using affinity, hydrophobic inter-
action, HA, and anion-exchange chromatography as explained by these
inventors in a recent application for a patent (Schumann et al., 2007). The
ceramic HA step utilized a 20-mM Tris, 5-mM calcium chloride, 250-mM
sodium chloride, 9% isopropanol, pH 6.9 adsorption buffer. The rEPO is
402 Larry J. Cummings et al.
REFERENCES
Bernardi, G. (1971). Chromatography of proteins on hydroxyapatite. Methods Enzymol. 22,
325–339.
di Salvo, M. L., Hunt, S., and Schirch, V. (2004). Expression, purification, and kinetic
constants for human and Escherichia coli pyridoxal kinases. Protein Expr. Purif. 36, 300–306.
Dong, X., Pan, H.-H., Wang, Q., and Wu, T. (2007). Understanding adsorption–
desorption dynamics of BMP-2 on hydroxyapatite (001) surface. Biophys. J. 93, 750–759.
Protein Chromatography on Hydroxyapatite Columns 403
Ferguson, W. J., Bell, D. H., Braunschweiger, K. I., Braunschweiger, W. R., Good, N. E.,
Jarvis, N. P., McCormick, J. J., Smith, J. R., and Wasmann, C. C. (1980). Hydrogen ion
buffers for biological research. Anal. Biochem. 104, 300–310.
Freitag, R., and Breier, J. (1995). Displacement chromatography in biotechnological down-
stream process. J. Chromatogr. A. 691, 101–112.
French patent 2231422, British patent 1468592, Swiss patent 597893, and German patent
2426501.
Gagnon, P. (1996). Purification tools for monoclonal antibodies. p. 87. Validated Biosystems
Inc., Tucson, AZ.
Gagnon, P. (2008). Improved antibody aggregate removal by hydroxyapatite chromatogra-
phy in the presence of polyethylene glycol. J. Immunol. Methods 336, 222–228.
Gagnon, P., Cheung, C.-W., and Yazaki, P. (2009). Reverse calcium affinity purification of
Fab with calcium derivatized hydroxyapatite. J. Immunol. Methods 342, 115–118.
Gorbunoff, M. J. (1984a). The interaction of proteins with hydroxyapatite: I. Role of
protein charge and structure. Anal. Biochem. 136, 425–432.
Gorbunoff, M. J. (1984b). The interaction of proteins with hydroxyapatite: II. Role of acidic
and basic groups. Anal. Biochem. 136, 433–439.
Gorbunoff, M. J. (1985). Protein chromatography on hydroxyapatite columns. Methods
Enzymol. 117, 370–380.
Gorbunoff, M. J. (1990). Protein chromatography on hydroxyapatite columns. Methods
Enzymol. 182, 329–339.
Gorbunoff, M. J., and Timasheff, S. (1984). The interaction of proteins with hydroxyapatite:
III. Mechanism. Anal. Biochem. 136, 440–445.
Guerrier, L., Boschetti, E., and Flayeux, I. (2001). A dual-mode approach to the selective
separation of antibodies and their fragments. J. Chromatogr. B 755, 37–46.
Harding, I. S., Hing, K. A., and Rashid, N. (2005). Surface charge and the effect of excess
calcium ions on the hydroxyapatite surface. Biomaterials 26, 6818–6826.
Hsieh, H.-Y., Calcutt, M. J., Chapman, L. F., Mitra, M., and Smith, D. S. (2003).
Purification and characterization of a recombinant alpha-N-acetylgalactosaminidase
from Clostridium perfringens. Protein Expr. Purif. 32, 309–316.
Ichitsuka, T., Kawamura, K., Ogawa, T., Sumita, M., and Yokoo, A. (1992). Packing
material for liquid chromatography US Patent 5039408.
Kadoya, T., Ebihara, M., Ishikawa, T., Isobe, T., Kobayashi, A., Kuwahara, H., Ogawa, T.,
Okuyama, T., and Sumita, M. (1986). A new spherical hydroxyapatite for high perfor-
mance liquid chromatography of proteins. J. Liq. Chromatogr. 9, 3543–3557.
Kandori, K., Ishikawa, T., and Miyagawa, K. (2004). Adsorption of immunogamma globulin
onto various synthetic calcium hydroxyapatite particles. J. Colloid Interface Sci. 273, 406–413.
Kawasaki, T. (1991). Hydroxyapatite as a liquid chromatographic packing. J. Chromatogr.
544, 147–184.
Kawasaki, T., Ikeda, K., Kuboki, Y., and Takahashi, S. (1986). Further study of hydroxyap-
atite high-performance liquid chromatography using both proteins and nucleic acids, and
a new technique to increase chromatographic efficiency. Eur. J. Biochem. 155, 249–257.
Kawasaki, A., Kirihara, S., and Mukai, K. (2008). Method for production of erythropoietin
WO/2008/068879.
Krane, S. M., and Glimcher, M. J. (1962). Transphosphorylation from nucleoside di and
triphosphates by apatite crystals. J. Biol. Chem. 237, 2991–2998.
Leibl, H., Eibl, M. M., Mannhalter, J. W., Tomasits, R., and Wolf, H. (1996). Method for
the isolation of biologically active monomeric immunoglobulin A from plasma fraction.
J. Chromatogr. B 678, 173–180.
Luellau, E., Freitag, R., Vogt, S., and von Stockar, U. (1998). Development of a down-
stream process for isolation and separation of immunoglobulin monomers, dimers, and
polymers from cell culture supernatant. J. Chromatogr. A 796, 165–175.
404 Larry J. Cummings et al.
Ma, Q. Y., Logan, T. J., and Traina, S. J. (1994). Effects of aqueous Al, Cd, Cu, Fe(II), Ni,
and Zn on Pb immobilization by hydroxyapatite. Environ. Sci. Tech. 28, 1219–1228.
McCue, J. T., Cecchini, D., Dolinski, E., and Hawkins, K. (2007). Use of an alternative scale-
down approach to predict and extend hydroxyapatite column lifetimes. J. Chromatogr. A
1165, 78–85.
Ng, P. K., Gagnon, P., and He, J. (2007). Mechanistic model for adsorption of immuno-
globulin on hydroxyapatite. J. Chromatogr. A 1142, 13–18.
Nuss, J. E., Choksi, K. B., DeFord, J. H., and Papaconstantinou, J. (2008). Decreased
enzyme activities of chaperones PDI and BiP in aged mouse livers. Biochem. Biophys.
Res. Commun. 365, 355–361.
Ogawa, T., and Hiraide, T. (1996). Effect of pH on gradient elution of different proteins on
two types of Macro-Prep ceramic hydroxyapatite. Am. Lab. 28, 31–34.
Schirch, V., Angelaccio, S., Hopkins, S., and Villar, E. (1985). Serine hydroxymethyltrans-
ferase from Escherichia coli: Purification and properties. J. Bacteriol 163, 1–7.
Schlatterer, J. C., Baeker, R., Kehler, W., Klose, J., Schlatterer, B., and Schlatterer, K.
(2006). Purification of prostaglandin D synthase by ceramic and size exclusion chroma-
tography. Prostaglandins Other Lipid Mediat. 81, 80–89.
Schroder, E., Jonsson, T., and Poole, L. (2003). Hydroxyapatite chromatography: Altering
the phosphate-dependent elution profile of protein as a function of pH. Anal. Biochem.
313, 176–178.
Schumann, C., Hesse, J.-O., and Mack, M. (2007). Method for purifying erythropoietin US
Patent Application Publication, US2007/0293420 A1.
Scopes, R. K. (1993). Protein Purification Principles and Practice. 3 edn. pp. 172–175.
Springer, New York.
Shen, J. W., Pan, H.-H., Wang, Q., and Wu, T. (2008). Molecular simulation of protein
adsorption and desorption on hydroxyapatite surfaces. Biomaterials 29, 513–532.
Shepard, S. R., Brickman-Stone, C., Koch, G., and Schrimsher, J. L. (2000). Discoloration of
ceramic hydroxyapatite used for protein chromatography. J. Chromatogr. A 891, 93–98.
Shi, X.-L., Feng, M.-Q., Shi, J.-H., Shi, J., Zhong, J., and Zhou, P. (2007). High-level
expression and purification of recombinant human catalase in Pichia pastoris. Protein Expr.
Purif. 54, 24–29.
Sinacola, J. R., and Robinson, A. S. (2002). Rapid refolding and polishing of single-chain
antibodies from Escherichia coli inclusion bodies. Protein Expr. Purif. 26, 301–308.
Skartsila, K., and Spanos, N. (2007). Surface characterization of hydroxyapatite: Potentiometric
titrations coupled with solubility measurements. J. Colloid Interface Sci. 308, 405–412.
Stránská, J., Chmelı́k, J., Peč, P., Popa, I., Řehulka, P., Šebela, M., and Tarkowski, P.
(2007). Inhibition of plant amine oxidases by a novel series of diamine derivatives.
Biochimie 89, 135–144.
Stuhlfelder, C., Lottspeich, F., and Mueller, M. J. (2002). Purification and partial amino acid
sequences of an esterase from tomato. Photochemistry 60, 233–240.
Sun, S. (2003). Removal of high molecular weight aggregates from an antibody preparation
using ceramic hydroxyapatite chromatography. In ‘‘Third International Hydroxyapatite
Conference’’, Lisbon, Portugal.
Teeters, M. A., and Quiñones-Garcı́a, I. (2005). Evaluating and monitoring the packing
behavior of process-scale chromatogragphy columns. J. Chromatogr. A 1069, 53–64.
Thompson, A. R., and Miles, B. J. (1973). New materials, especially for chromatography.
Methodol. Dev. Biochem. Prep. Tech. 2, 95–101.
Tiselius, A., Hjertén, S., and Levin, O. (1956). Protein chromatography on calcium phos-
phate columns. Arch. Biochem. Biophys. 65, 132–155.
Zhu, X. D., Fan, H. S., Li, D. X., Luxbacher, T., Xiao, Y. M., Zhang, H. J., and
Zhang, X. D. (2009). Effect of surface structure on protein adsorption to biphasic
calcium-phosphate ceramics in vitro and in vivo. Acta Biomater. 5, 1311–1318.
C H A P T E R T W E N T Y- F I V E
Contents
1. Theory 406
2. Latest Technology in HIC Adsorbents 408
3. Procedures for Use of HIC Adsorbents 409
3.1. Introduction 409
3.2. Choice of adsorbent 409
3.3. Feed/load preparation 410
3.4. Adsorbent preparation 410
3.5. Product elution 410
3.6. Gradient elution 412
3.7. Stepwise (isocratic) elution 412
3.8. Adsorbent regeneration and sanitization 413
References 413
Abstract
Hydrophobic interaction chromatography (HIC) is a valuable tool used in protein
purification applications. HIC is used in the purification of proteins over a broad
range of scales—in both analytical and preparatory scale applications. HIC is
used to remove various impurities that may be present in the solution, including
undesirable product-related impurities. In particular, HIC is often employed to
remove product aggregate species, which possess different hydrophobic proper-
ties than the target monomer species and can often be effectively removed using
HIC. In this chapter, we provide a description of the basic theory of HIC and how it
is used to purify proteins in aqueous-based solutions. Following the theoretical
background, the latest in HIC adsorbent technology is described, including a list
of commonly used and commercially available adsorbents. The basic procedures
for using HIC adsorbents are described next, in order to provide the reader with
useful starting points to apply HIC in protein purification applications.
405
406 Justin T. McCue
1. Theory
Hydrophobic proteins will self-associate, or interact, when dissolved
in an aqueous solution. This self-association forms the basis for a variety of
biological interactions, such as protein folding, protein–substrate interac-
tions, and transport of proteins across cellular membranes ( Janson and
Rydén, 1997). Hydrophobic interaction chromatography (HIC) is used
in both analytical and preparatory scale protein purification applications.
HIC exploits hydrophobic regions present in macromolecules that bind
to hydrophobic ligands on chromatography adsorbents. The interaction
occurs in an environment which favors hydrophobic interactions, such as
an aqueous solution with a high salt concentration.
By itself, water (a polar solvent) is a poor solvent for nonpolar molecules.
Under such an environment, proteins will self-associate, or aggregate, in
order to achieve a state of lowest thermodynamic energy. Prior to self-
association, water molecules form highly ordered structures around each
individual macromolecule (Fig. 25.1A). The self-association of nonpolar
molecules (such as proteins) in the polar solvent is driven by a net increase in
entropy of the environment. During the aggregation process, the overall
surface area of hydrophobic sites of the protein exposed to the polar solvent
Legend
Protein-ligand
binding
Entropy
increase
Base matrix Base matrix
( Jones, 1975; Nozaki and Tanford, 1971; Tanford, 1962; Zimmerman et al.,
1968). Empirical hydrophobic scales for proteins have also been created (Chotia,
1976; Krigbaum and Komoriya, 1979; Manavalan and Ponnuswamy,
1978; Rose et al., 1985; Wertz and Scheraga, 1978) which are based upon the
fraction of amino acids exposed on the protein surface, as well as the degree of
amino acid hydrophobicity. The ability to predict the hydrophobicity of
complex proteins has been only semiquantitative to date, and experiments are
usually required to accurately understand protein hydrophobicity in a given
aqueous solution.
general rule, the strength of hydrophobic binding of the ligand will increase
with the length of the organic chain. Several of the most common ligands
include butyl, octyl, and phenyl, which are linked to the base bead support
through several different coupling approaches (Hjertén et al., 1974; Ulbrich
et al., 1964). Aromatic ligands, such as phenyl, can also interact with the
adsorbed compounds through so-called ‘‘p–p interactions,’’ which can
further strengthen the hydrophobic interaction (Porath and Larsson, 1978).
The hydrophobic interaction strength of the ligand can also be influ-
enced by the ligand loading (ligand density) on the base matrix. The
strength of interaction can increase with higher ligand densities. In order
to have reproducible performance, manufacturers of HIC adsorbents must
often produce adsorbents with narrow ranges of ligand density to ensure
consistent performance from lot to lot.
species. During the elution step, a portion (or fraction) of the eluate may
contain highly purified product, while fractions before and after contain
higher levels of undesirable impurities. A schematic of an elution process
(during a gradient elution) is shown in Fig. 25.2. Figure 25.2 illustrates that
the column effluent collected during the elution step may need to be
fractionated in order to achieve acceptable product purity when using
HIC. The gradient elution process is described in more detail in Section 3.6.
The elution process can be done using either a stepwise (isocratic) or a
gradient approach. The four most common methods (listed from most
common to least common) used to elute the bound protein include the
following:
1. Decrease in the salt concentration (relative to the binding conditions). A decrease
in the salt concentration will decrease the strength of hydrophobic
interaction between the protein and the ligand, and the protein will be
desorbed and eluted from the column.
2. Addition of organic solvents. Addition of an organic solvent (such as
ethylene or propylene glycol) changes the solvent polarity, which dis-
rupts the hydrophobic interaction.
3. Increase in the salt concentration (using a chaotropic salt). Addition of a
chaotropic salt will disrupt the hydrophobic interaction.
4. Detergent addition. Detergents are used as protein displacers, and have
been used mainly for the purification of membrane proteins when using
HIC ( Janson and Rydén, 1997).
Target
product Strongly bound
Weakly impurities
bound (e.g. aggregates)
impurities
High salt
concentration
Protein concentration
Salt concentration
Salt
gradient Low salt
elution concentration
Elution volume
This most common approach used to elute proteins from HIC adsor-
bents is by lowering the salt concentration during the elution step. This
should be the first method that is attempted when using HIC for purifica-
tion of a new protein compound. The other approaches described above
have the disadvantage that an additional component (such as a chaotropic
salt or an organic solvent) needs to be added, which may impact protein
stability. However, such agents may be required in order to effectively elute
a strongly bound protein species from the adsorbent. Each protein must be
evaluated case by case to determine which elution method is appropriate.
The HIC adsorbent used in the purification may also influence which
elution method is effective.
REFERENCES
Ben-Naim, A. (1980). Hydrophobic Interactions. Plenum Press, New York.
Chotia, C. (1976). Surface of monomelic proteins. J. Mol. Biol. 105, 112.
Fausnaugh, J. L., and Regnier, F. E. (1986). Solute and mobile phase contributions to
retention in hydrophobic interaction chromatography of proteins. J. Chromatogr. 359,
131–146.
GE Healthcare (2006). Data File No. 18-1127-63 AC.
Hjertén, S., Rosengren, J., and Påhlman, S. (1974). Hydrophobic interaction chromato-
graphy. J. Chromatogr. 101, 281–288.
Hofmeister, F. (1988). On regularities in the albumin precipitation reactions with salts and
their relationship to physiological behavior. Arch. Exp. Pathol. Pharmakol. 24, 247–260.
Janson, J.-C., and Rydén, L. (eds.) (1997). Protein Purification: Principles, High-Resolution
Methods, and Applications, 2nd edn., p. 284. Wiley-VCH, New York.
Jones, D. D. (1975). Amino acid properties and side-chain orientation in proteins. J. Theor.
Biol. 50, 167–183.
Krigbaum, W. R., and Komoriya, A. (1979). Local interactions as a structure determinant for
protein molecules. Biochim. Biophys. Acta 576, 204–248.
Manavalan, P., and Ponnuswamy, P. K. (1978). Hydrophobic character of amino acid
residues in globular proteins. Nature 275, 673–674.
Melander, W., and Horvath, C. (1977). Salt effects on hydrophobic interactions in precipi-
tation and chromatography of proteins: An interpretation of the lyotropic series. Arch.
Biochem. Biophys. 183, 200–215.
Nozaki, Y., and Tanford, C. (1971). The solubility of amino acids and two glycine peptides
in aqueous ethanol and dioxane solutions. Establishment of a hydrophobicity scale.
J. Biol. Chem. 246, 2211–2217.
414 Justin T. McCue
Påhlman, S., Rosengren, J., and Hjerten, S. (1977). Hydrophobic interaction chromatography
on uncharged Sepharose derivatives. J. Chromatogr. 131, 99–108.
Porath, J., and Larsson, B. (1978). Charge-transfer and water-mediated chromatography.
I. Electron-acceptor ligands on cross-linked dextran. J. Chromatogr. 155, 47–68.
Rose, G. D., Geselowitz, A. R., Lesser, G. J., Lee, R. H., and Zehfus, M. H. (1985).
Hydrophobicity of amino acid residues in globular proteins. Science 229, 834–838.
Tanford, C. (1962). Contribution of hydrophobic interactions to the stability of the globular
conformation of proteins. J. Am. Chem. Soc. 84, 4240–4247.
Tanford, C. (1980). In The Hydrophobic Effect 2nd edn. Wiley, New York.
Ulbrich, V., Makes, J., and Jurecek, M. (1964). Identification of giycidyl ethers. Bis(phenyl-)
and bis(a-naphthylurethans) of glycerol a-alkyl (aryl)ethers. Collect. Czech. Chem.
Commun. 29, 1466–1475.
Wertz, D. H., and Scheraga, H. A. (1978). Influence of water on protein structure.
Macromolecules 11, 9–15.
Yamamoto, S., Nakanishi, K., and Matsuno, R. (1988). Ion-Exchange Chromatography of
Proteins Mercel Dekkar, New York.
Zimmerman, J.M, Eliezer, N., and Simha, R. (1968). The characterization of amino acid
sequences in proteins by statistical methods. J. Theor. Biol. 21(2), 170–201.
C H A P T E R T W E N T Y- S I X
Affinity Chromatography:
General Methods
Marjeta Urh, Dan Simpson, and Kate Zhao
Contents
1. Introduction 418
2. Selection of Affinity Matrix 419
2.1. General features of the support material 419
2.2. Selectivity 420
2.3. Stability 422
2.4. Magnetic affinity beads 422
3. Selection of Ligands 423
3.1. General considerations for ligands design and selection 423
3.2. Characterization of immobilized ligand 424
3.3. Affinity matrices carrying specific ligands 425
3.4. Immunoglobulin binding proteins 425
3.5. Lectins 426
3.6. Biomimetic ligands 427
3.7. Covalent affinity chromatography 428
4. Attachment Chemistry 429
4.1. Activation of surface 430
4.2. Ligand attachment 430
5. Purification Method 433
5.1. Sample preparation 433
5.2. Binding and wash 433
5.3. Elution 434
References 435
Abstract
Affinity chromatography is one of the most diverse and powerful chro-
matographic methods for purification of a specific molecule or a group of
molecules from complex mixtures. It is based on highly specific biological
interactions between two molecules, such as interactions between enzyme
and substrate, receptor and ligand, or antibody and antigen. These interactions,
which are typically reversible, are used for purification by placing one of the
417
418 Marjeta Urh et al.
1. Introduction
Affinity chromatography is a method for selective purification of a
molecule or group of molecules from complex mixtures based on highly
specific biological interaction between the two molecules. The interaction
is typically reversible and purification is achieved through a biphasic inter-
action with one of the molecules (the ligand) immobilized to a surface while
its partner (the target) is in a mobile phase as part of a complex mixture. The
capture step is generally followed by washing and elution, resulting in
recovery of highly purified protein. Highly selective interactions allow for
a fast, often single step, process, with potential for purification in the order
of several hundred to thousand-fold. Additional uses of affinity chromatog-
raphy include the ability to concentrate substances present at low concen-
tration and the ability to separate proteins based on their biological function
where an active form can be separated from the inactive form or a form with
different biological function.
Recent decades have seen tremendous advancements in the utility of
affinity chromatography, with developments in support materials such as
flow-through beads, magnetic beads and monolithic materials as well as new
ligands with a variety of interesting biological properties. In addition, this
approach is no longer used only for purification of specific biomolecules.
It is also quickly becoming a method of choice to study biological interac-
tions and can be used for preparation of samples for mass spectrometry or for
specific removal of contaminants.
Overview of Affinity Chromatography Methods 419
2.2. Selectivity
One key feature of an affinity matrix is its selectivity. It should be specific for
a protein of interest as determined by the specific ligand coupled to the
matrix and inert to all other compounds present in the complex sample.
Since most applications are performed in aqueous solutions, often with low
ionic strength, the support should be hydrophilic and contain limited charge
that may lead to undesirable ionic interaction. Many commercially available
supports fulfill these requirements, either with their native structure or by
coating with suitable materials. Common supports include beaded agarose
and cellulose, available commercially from a number of vendors including
Sepharose from GE Healthcare and Affigel from Bio-Rad (Table 26.1).
Nonspecificity can come from the support itself, such as hydrophobicity
associated with polystyrene beads and negative charge on the surface of
silica. It can also be introduced when modifying a matrix to accept a
particular ligand. In these cases, the attachment chemistry, the ligand, and
the spacer between the ligand and the matrix should be carefully designed,
screened and optimized for selectivity, capacity for target of interest, and
low nonspecific binding.
Table 26.1 Examples of commercially available matrices
Name Vendor Matrix material Particle size (mm) Exclusion limit (Da)
SepharoseTM GE Healthcare Agarose 40–165 10,000–1,000,000
CL6B
SepharoseTM GE Healthcare Agarose 40–165 30,000–5,000,000
CL4B
Bio-gel A-5m Bio-Rad Agarose 75–300 (50–200 mesh) 10,000–5,000,000
medium
Perloza MT100 Iontosorb Cellulose 100–250 2,000,000
medium
Perloza MT50 Iontosorb Cellulose 100–250 100,000
medium
AllTech Grace Silica 7 Pore size
Macrosphere 60–300 Å
Bio-gel P-100 Bio-Rad Polyacrylamide 90–180 5000–100,000
medium
SephacrylTM GE Healthcare Cross-linked ally dextrose 50 1000–100,000
Poros 50 Applied Cross-linked poly(styrene- 50 Pore size 50–100 nm
Biosystems divinylbenzene)
422 Marjeta Urh et al.
2.3. Stability
The affinity matrix must also be chemically and physically stable during the
process, such that the support material itself as well as the attached ligand
should not react to the solvents used in the process, nor should they be
degraded or damaged by enzymes and microbes that might be present in the
sample. The chemical compatibilities of commercially available affinity
matrices are usually supplied by the manufacturer, which should be used
as guidance for developing a successful purification protocol. Cross-linked
agarose can usually withstand a wide pH range (e.g., pH 3–12), most
aqueous solvents (including denaturants), many organic solvents or modi-
fiers, and enzymatic treatments. Materials such as glass and silica are not
stable at alkaline conditions due to hydrolysis; therefore, coating of these
surfaces is often needed before attaching ligands. The matrix should also
withstand physical stress, such as pressure, especially when packed into a
column, and remain intact during the purification process. High pressure
can compress the matrix, causing it to collapse. Agarose beads and other soft
gel matrices are more susceptible to pressure, relative to stronger supports,
such as silica, polystyrene and other highly cross-linked materials.
Monoliths are macroporous, nonbeaded single matrix that can be made
from different materials such as agarose, silica, GMA/EDMA, and cryogel
(Mallik and Hage, 2006; Plievaa et al., 2009). It possesses many of the
desired properties of an affinity support and has become popular due to
the presence of large flow-through pores with no void volume permitting
convective flow instead of diffusion and high flow rate for shortened run
time. In addition, the matrix will not compress and has less pressure drop
during column chromatography and can be made to withstand larger pH
ranges and harsh chemicals. There are several limitations of monoliths as
compared to traditional matrix, such as lower capacity, special processes
required for making each type of monolith affinity supports and the current
limited range of available affinity types (Mallik and Hage, 2006).
3. Selection of Ligands
Selection of the appropriate ligand requires a certain degree of knowl-
edge and understanding of the nature of interactions between the ligand and
the target molecule; where the ligand must specifically bind the target
molecule and should be stable in different binding and elution conditions.
Additionally, when developing affinity purification scheme it is important
to consider whether the ligand is commercially available or de novo devel-
opment of the ligand and affinity matrix will be required. The success of the
second scenario will largely depend on existing knowledge of protein
structure and the nature of interaction, requiring the use of molecular
modeling and combinatorial organic synthesis coupled with immobilization
chemistry and selective binding analysis. The time and effort to design a
novel ligand and to develop appropriate coupling chemistry and matrix may
prove to be too lengthy and costly, whereas the use of nonaffinity-based
purification technique such as ion exchange and hydrophobic interaction
schemes may be a better choice.
Ligand Specificity
Cibacron Blue Albumin, kinases, dehydrogenases, enzymes requiring
F3G-A adenylyl-containing cofactors, NADþ
Blue B Kinases, dehydrogenases, nucleic acid binding proteins
Orange A Lactate dehydrogenase
Green A HAS, dehydrogenases
Polymixin Endoproteins
Benzamidine Serine proteases (thrombin, trypsin, kallikrein)
Biotin Streptavidin, avidin
Gelatin Fibronectin
Heparin DNA binding proteins, serine protease inhibitors
(antithrombin III), growth factors, lipoproteins,
hormone receptors, coagulation factors, DNA, RNA
Lysine Plasminogen, rRNA, dsDNA
Arginine Serine proteases with affinity for arg, fibronectin,
prothrombin
ADT Enzymes with affinity for NADPþ
AMP NAD-dependent dehydrogenases and ATP-dependent
kinases
NAD, NADP Dehydrogenases
Lectins Glycoproteins, polysaccharides, glycolipids
Calmodulin Calmodulin binding proteins, ATPase, adenylate cyclase,
kinases, phosphodiesterase,
Protein A Fc regions of many IgG subtypes, species dependent, weak
interactions with IgA, IgM, IgD
Protein G Fc region of many IgG subtypes, species dependent
Protein L Kappa light chains of antibodies (Fab, single chain variable
region scFv)
and samples are collected into buffers with neutral or slightly basic pH to
avoid denaturation and loss of activity. In cases where biological stability is
lost during elution at low pH, elution with a pH gradient in combination
with salt can be explored.
3.5. Lectins
Lectins are a diverse group of proteins which bind carbohydrates with
high degree of specificity where each lectin has its own specificity profile.
They are often used in affinity purification or enrichment of carbohydrate
moieties of complex glycoconjugates, such as polysaccharides, glycolipids,
Overview of Affinity Chromatography Methods 427
and glycoproteins. They also allow for a specific isolation of different glycol-
forms of a specific protein depending on the nature of glycosylation. Recently,
lectins have been used not only for the purification of sugar-containing
molecules but also for the enrichment of subgroups of glycoproteins for
analysis in mass spectrometry (Hirabayashi, 2008). See Chapter 34.
Most commercially available lectins are of plant origin, and there are
over 100 different lectins available either in conjugated or free form. Lectin
from Canavalia ensiformis, known as Conacanavalin A (ConA) has the
affinity for a-D-mannose, a-D-glucose, and N-acetylglucosamine and is
probably the most frequently used lectin (Hermanson, 1992). Two other
popular lectins are wheat germ agglutinin (WGA) and Jacalin; WGA binds
to sialic acid and molecules containing N-acetyl-D-glucosamine residue and
Jacalin binds to a-D-galactosyl groups.
Coupling of lectin onto resin is often performed at neutral pH and in the
presence of sugar to preserve sugar binding site. Binding to the target
molecule is also usually carried out in neutral pH; note that some lectins
require the presence of divalent metal ions, Ca2þ and Mn2þ in the case of
ConA, for optimal binding. Elution is accomplished by adding an access of
the specific sugar molecule to the elution buffer and this can be done as
stepwise or gradient elution. After elution, the free sugar should be removed
by dialysis or size exclusion chromatography.
A B
N
Solid support
Cl O O N
C H Fluorophore
Chloroalkane
ligand
HaloTag protein
Figure 26.1 HaloTag technology comprises two components: (A) The HaloTag
protein shown on the left with covalently bound HaloTag-TMR ligand. N- and
C-termini are indicated. (B) The chloroalkane ligand. Different functional groups
including, but not limited to, fluorescent dyes or solid support surfaces can be attached
to the chloroalkane ligand, which covalently binds to HaloTag protein and imparts
different functionalities including protein immobilization and fluorescent labeling.
clearly has its advantages, it also creates a challenge in eluting the protein of
interest. Because the covalent bond between HaloTag and chloroalkane
cannot be reversed, traditional approaches to elute proteins from the resin
cannot be utilized. Instead the protein of interest can be released from
HaloTag by specific protease (TEV) as the TEV recognition site is present
between the two moieties (the HaloTag and fusion target protein). Upon
cleavage, HaloTag stays bound to the resin while the fusion partner is
released yielding highly pure protein free of tag (Urh et al., 2008).
Besides using HaloTag for purification of fusion proteins described
earlier, the immobilized HaloTag fusions can also be considered as affinity
ligands in their own right, and can, similarly to protein G, be used to capture
antibodies or other proteins which specifically bind to the proteins fused to
HaloTag. The advantage of this system is that unlike other covalent immo-
bilization techniques, where binding of protein is random and may lead to
multiple attachment sites and improper orientation, immobilization using
HaloTag is a single point attachment through active site and therefore
oriented. Single point, oriented attachment increases capacity, effectiveness
and reproducibility of the system, and HaloTag fusions covalently bound to
the matrix can therefore improve purification of specific antibodies or
isolation of binding partners.
4. Attachment Chemistry
This section will briefly review some of the more common chemistry
for the covalent attachment of affinity ligands to conventional surfaces such
as agarose, cellulose, silica, glass, and synthetic polymeric supports. Strate-
gically, the process can be divided into three components: (1) the surface or
430 Marjeta Urh et al.
Affinity ligand
Bead or surface Activation and/or linkage reactive group
Soft gels: agarose, cellulose Cyanogen bromide Amine
Synthetic supports: Aldehyde (reductive Amine
Polyacrylamide beads, amination)
Trisacryl, Sephacryl, Activated carboxyl ester Amine
Ultragel, Azlactone beads, (succinimidyl ester)
Methacrylate (TSK gel), Carbonyldiimidazole Amine
Eupergit, Polystyrene Carboxyl (activation Amine
(Poros supports) concurrent with
coupling)
FMP activation Amine, thiol
Divinyl sulfone Amine, thiol
Azlactone Amine, thiol
Epoxy (bisoxirane, Amine, thiol
epichlorohydrin)
Tresyl chloride Amine, thiol
Haloacetyl (iodo or Thiol
bromo)
Maleimide Thiol
Pyridyl disulfide Thiol
Amine Carboxyl
(activation
concurrent
with
coupling)
Hydrazide Carbohydrate
(periodate
reduced)
Inorganics: controlled pore 3-(Glycidyloxypropyl) Amine
glass, silica, alumina, trimethoxy-silane
zeolites, etc. 3-(Aminopropyl) Carboxyl (after
trimethoxysilane activation),
aldehyde
5. Purification Method
Purification by affinity starts with proper handling of the sample and
the matrix, followed by selective binding (capture) of the target, washing to
remove nonspecific background, and, finally, elution of the bound target.
Successful affinity purification depends on a number of notable factors
including, the amount and accessibility of the ligand on the resin, the
strength of the interaction, and the integrity of protein to be immobilized.
Usually, conditions are optimized to maximize the interaction between a
target and immobilized ligand during the binding and wash process, then,
switched to substantially weaken the interaction thus allowing for release of
the target. It is recommended to perform small scale trials to select for the
best purification conditions. A short summary of some common practical
issues and considerations affecting affinity purification is given in the
following sections.
the flow rate used for binding. The binding process can be simplified
as Eq. (26.1), assuming a 1:1 molar ratio where Ka is the association
equilibrium constant, [L] is ligand concentration, [T ] is the concentration
of target protein, and [LT ] is the concentration of the complex. Ka equals
[LT ]/([L] [T ]), which can also be expressed as ka/kd, where ka is second
order association rate constant that depends on the concentrations of both
L and T, and kd is the first order dissociation constant that does not depend
on ligand concentration. Higher Ka usually leads to a higher adsorption
ratio, defined as the ratio of bound to total applied target, thus, better
binding. Normally, the ligand concentration on the matrix is around
10 2–10 4 M and to achieve efficient binding, the Ka value should be in
the range of 104–106 M 1.
ka
L þ T ! LT ð26:1Þ
kd
5.3. Elution
Elution of bound target from the resin is essentially the reverse process of
binding, where conditions are optimized to reduce the Ka, that is, weaken-
ing the interaction between target and ligand. The elution condition should
Overview of Affinity Chromatography Methods 435
not denature the target protein, unless such conditions are compatible with
downstream applications.
There are two different types of elution methods, namely, specific and
nonspecific elution. In specific elution, the target protein–ligand complex is
challenged by agents that will compete for either the ligand or the target
thereby releasing the target protein into solution. The concentration and
amount (volume) of competitive reagent used for elution will depend
on their affinity relative to that of the immobilized complex, where weaker
competitive reagents require higher concentration and more volume
as compared to higher affinity additives. A good starting point for weak
competitors is to use concentration 10-fold higher than that of the ligand.
The specific elution is usually milder and proteins are more likely to retain
their activity, but the slow elution, broad elution peaks, and the need to
remove competing agent from the recovered protein are some of the
drawbacks of this approach. For nonspecific elution, solvent conditions
are manipulated to reduce the association rate constant (Eq. (26.1)), which
ideally should approach zero, and to increase the dissociation rate constant,
thus, weakening the overall affinity (Ka) resulting in dissociation of the
complex. Elution conditions can be optimized according to the mechanism
of interaction between the ligand and protein, such as increasing salt
concentration to reduce ionic interactions or by altering pH to change the
protonation/ionization state, thus modulating the strength of hydrogen
bonds, hydrophobic interactions as well electrostatic interactions. An exam-
ple of elution by changing pH is the elution of antibodies from immobilized
protein A or protein G, yet because the affinity of proteinA/G to antibodies
is very strong, with Ka in the 108 M 1 range, a combination of different
elution conditions may be required for maximum antibody release. In many
cases, affinity requires proper three-dimensional folding of a protein so that
chaotropic reagents or reagents that will affect protein folding can be used to
elute target of interest; however, care must be taken to maintain proper
folding of the target after elution by quickly returning to native conditions.
When the affinity is weak, binding is achieved at high concentration of the
target molecule which is then eluted by dissociation of the complex through
dilution. This approach is known as isocratic elution.
REFERENCES
Alexander, C., Andersson, H. S., Andersson, L. I., Ansell, R. J., Kirsch, N., Nicholls, I. A.,
O’Mahony, J., and Whitcombe, M. J. (2006). Molecular imprinting science and tech-
nology: A survey of the literature for the years up to and including 2003. J. Mol. Recognit.
19, 106–180.
Blumberg, P. M., and Strominger, J. L. (1972). Isolation by covalent affinity chromatogra-
phy of the penicillin-binding components from membranes of Bacillus subtilis. Proc. Natl.
Acad. Sci. USA 69, 3751–3755.
436 Marjeta Urh et al.
Brena, B., Ovsejevi, K., Luna, B., and Batista-Viera, F. (1993). Thiolation and reversible
immobilization of sweet potato beta-amylase on thiolsulfonate-agarose. J. Mol. Catal. 84,
381–390.
Campbell, D. H., Luescher, E., and Lerman, L. S. (1951). Immunologic adsorbents: I.
Isolation of antibody by means of a cellulose-protein antigen. Proc. Natl. Acad. Sci.
USA 37, 575–578.
Cecı́lia, A., Roque, A., and Lowe, C. R. (2005). Advances and applications of de novo
designed affinity ligands in proteomics. Biotechnol. Adv. 24, 17–26.
Chandran, S. P., Hotha, S., and Prasad, B. L. V. (2009). Tunable surface modification of
silica nanoparticles through ‘‘click’’ chemistry. Current Science 95, 1327–1333.
Cuatrecasas, P., Wilchek, M., and Anfinsen, C. B. (1968). Selective enzyme purification by
affinity chromatography. Proc. Natl. Acad. Sci. USA 61, 636–643.
Faulmann, E. L., Duvall, J. L., and Boyle, M. D. (1991). Protein B: A versatile bacterial
Fc-binding protein selective for human IgA. Biotechniques 10, 748–755.
Gauchet, C., Labadie, G. R., and Poulter, C. D. (2006). Regio- and chemoselective
covalent immobilization of proteins through unnatural amino acids. J. Am. Chem. Soc.
128, 9274–9275.
GE Healthcare (ed.) (2007). Affinity chromotography, principle and methods. GE
Healthcare.
Guilbault, G. G. (ed.) (1988). In ‘‘Analytical uses of immobilized biological compounds for
detection, medical and industrial uses,’’ Dordrecht, Boston, MA.
Guss, B., Eliasson, M., Olsson, A., Uhlen, M., Frej, A. K., Jornvall, H., Flock, J. I., and
Lindberg, M. (1986). Structure of the IgG-binding regions of streptococcal protein G.
EMBO J. 5, 1567–1575.
Gustavsson, P.-E., and Larsson, P.-O. (2006). Support materials for affinity chromatography.
In ‘‘Handbook of Affinity Immobilization,’’ (D. S. Hage, ed.), pp. 15–34. CRC press/
Taylor and Francis Group, Boca Raton, FL.
Hage, D. S. (ed.) (2006). In ‘‘Handbook of affinity chromatography,’’ CRC press/Taylor
and Francis Group, Boca Raton, FL.
Hage, D. S., Bian, M., Burks, R., Karle, E., Ohnmachi, C., and Wa, C. (2006). Bioaffinity
Chromatography. In ‘‘Handbook of Affinity Immobilization,’’ (D. S. Hage, ed.),
pp. 101–126. CRC press/Taylor and Francis Group, Boca Raton, FL.
Hermanson, G. T. (ed.) (1992). In ‘‘Immobilized affinity ligand techniques,’’ Academic
Press, San Diego, CA.
Hirabayashi, J. (2008). Concept, strategy and realization of lectin-based glycan profiling.
J. Biochem. 144, 139–147.
Koneracka, M., Kopcansky, P., Timbo, M., Ramchand, C. N., Saiyed, Z. M., and Trevan, M.
(2006). Immobilization of enzymes on magnetic particles. In ‘‘Immobilization of Enzymes
and Cells,’’ ( J. M. Guisan, ed.), pp. 217–228. Humana Press.
Korpela, T., and Hinkkanen, A. (1976). A simple method to introduce aldehydic function to
agarose. Anal. Biochem. 71, 322–323.
Kowal, R., and Parsons, R. G. (1980). Stabilization of proteins immobilized on Sepharose
from leakage by glutaraldehyde crosslinking. Anal. Biochem. 102, 72–76.
Labrou, N., and Clonis, Y. D. (1994). The affinity technology in downstream processing.
J. Biotechnol. 36, 95–119.
Labrou, N. E. (2000). Dye-ligand affinity chromatography for protein separation and
purification. Methods Mol. Biol. 147, 129–139.
Labrou, N. E. (2003). Design and selection of ligands for affinity chromatography.
J. Chromatogr. B Analyt. Technol. Biomed. Life Sci. 790, 67–78.
Labrou, N. E., and Clonis, Y. D. (2002). Immobilized synthetic dyes in affinity chromatog-
raphy. In ‘‘Theory and Practice of Biochromatograph,’’ (M. A. Vijayalakshimi, ed.),
pp. 235–251. Taylor and Francis Group, London.
Overview of Affinity Chromatography Methods 437
Labrou, N. E., Karagouni, A., and Clonis, Y. D. (1995). Biomimetic-dye affinity adsorbents
for enzyme purification: Application to the one-step purification of Candida boidinii
formate dehydrogenase. Biotechnol. Bioeng. 48, 278–288.
Langone, J. J. (1982). Applications of immobilized protein A in immunochemical techni-
ques. J. Immunol. Methods 55, 277–296.
Lerman, L. S. (1953a). A biochemically specific method for enzyme isolation. Proc. Natl.
Acad. Sci. USA 39, 232–236.
Lerman, L. S. (1953b). Antibody chromatography on an immunologically specific adsorbent.
Nature 172, 635–636.
Los, G. V., Encell, L. P., McDougall, M. G., Hartzell, D. D., Karassina, N., Zimprich, C.,
Wood, M. G., Learish, R., Ohana, R. F., Urh, M., Simpson, D., Mendez, J., et al.
(2008). HaloTag: A novel protein labeling technology for cell imaging and protein
analysis. ACS Chem. Biol. 3, 373–382.
Lowe, C. R., Lowe, A. R., and Gupta, G. (2001). New developments in affinity chroma-
tography with potential application in the production of biopharmaceuticals. J. Biochem.
Biophys. Methods 49, 561–574.
Madoery, R., and Minchiotti, M. (2006). Cibacron Blue-Eupergit, an affinity matrix for
soybean (Glycine max) phospholipase A2 purification. Enzyme and Microb. Technol. 38,
869–872.
Mallik, R., and Hage, D. S. (2006). Affinity monolith chromatography. J. Sep. Sci. 29,
1686–1704.
Mallik, R., Wa, C., and Hage, D. S. (2007). Development of sulfhydryl-reactive silica for
protein immobilization in high-performance affinity chromatography. Anal. Chem. 79,
1411–1424.
O’Shannessy, D. J., and Wilchek, M. (1990). Immobilization of glycoconjugates by their
oligosaccharides: Use of hydrazido-derivatized matrices. Anal. Biochem. 191, 1–8.
Oates, M. R., Clarke, W., Marsh, E. M., and Hage, D. S. (1998). Kinetic studies on the
immobilization of antibodies to high-performance liquid chromatographic supports.
Bioconjug. Chem. 9, 459–465.
Ohana, R. F., Encell, L. P., Zhao, K., Simpson, D., Slater, M. R., Urh, M., and Wood, K.V
(2009). Halotag7: A genetically engineered tag that enhances bacterial expression of
soluble protein and improves protein purification. Protein Expr. Purif. 68, 110–120.
Ostrove, S. (1990). Affinity chromatography: General methods. Methods Enzymol. 182,
357–371.
Plievaa, F. M., De Seta, E., Galaevb, I. Y., and Mattiasson, B. (2009). Macroporous elastic
polyacrylamide monolith columns: Processing under compression and scale-up. Sep.
Purif. Technol. 65, 110–116.
Porath, J., Axen, R., and Ernback, S. (1967). Chemical coupling of proteins to agarose.
Nature 215, 1491–1492.
Renkin, E. M. (1954). Filtration, diffusion, and molecular sieving through porous cellulose
membranes. J. Gen. Physiol. 38, 225–243.
Saiyed, Z., Telang, S., and Ramchand, C. (2003). Application of magnetic techniques in the
field of drug discovery and biomedicine. Biomagn. Res. Technol. 1, 2.
Starkenstein, E. (1910). Ferment action and the influence upon it of neutral salts. Biochem. Z.
24, 210–218.
Stellwagen, E. (1990). Chromatography on immobilized reactive dyes. Methods Enzymol.
182, 343–357.
Urh, M., Hartzell, D., Mendez, J., Klaubert, D. H., and Wood, K. (2008). Methods for
detection of protein-protein and protein-DNA interactions using HaloTag. Methods Mol.
Biol. 421, 191–209.
Vijayendran, R. A., and Leckband, D. E. (2001). A quantitative assessment of heterogeneity
for surface-immobilized proteins. Anal. Chem. 73, 471–480.
438 Marjeta Urh et al.
Wilchek, M., Knudsen, K. L., and Miron, T. (1994). Improved method for preparing
N-hydroxysuccinimide ester-containing polymers for affinity chromatography.
Bioconjug. Chem. 5, 491–492.
Wilchek, M., Miron, T., and Kohn, J. (1984). Affinity chromatography. Methods Enzymol.
104, 3–55.
Zachariou, M. (ed.) (2007). In ‘‘Affinity Chromatography, Methods and Protocols,’’
Humana Press, Totowa, NJ.
C H A P T E R T W E N T Y- S E V E N
Immobilized-Metal Affinity
Chromatography (IMAC): A Review
Helena Block,* Barbara Maertens,* Anne Spriestersbach,*
Nicole Brinker,* Jan Kubicek,* Roland Fabis,* Jörg Labahn,†
and Frank Schäfer*
Contents
1. Overview on IMAC Ligands and Immobilized Ions 440
2. IMAC Applications 444
2.1. Detection and immobilization 445
2.2. Purification of protein fractions 446
2.3. Purification of His-tagged proteins 448
2.4. General considerations of protein purification by IMAC 451
2.5. Copurifying proteins on IMAC and what to do about it 454
2.6. IMAC for industrial-scale protein production 458
2.7. High-throughput automation of IMAC 460
2.8. Special applications: Purification of membrane proteins 461
2.9. Special applications: Purification of zinc-finger proteins 463
2.10. Protein purification protocols 464
2.11. Cleaning and sanitization 466
2.12. Simplified metal-ion stripping and recharging protocol 467
3. Conclusions 467
Acknowledgments 468
References 468
Abstract
This article reviews the development of immobilized-metal affinity chromatog-
raphy (IMAC) and describes its most important applications. We provide an
overview on the use of IMAC in protein fractionation and proteomics, in protein
immobilization and detection, and on some special applications such as purifi-
cation of immunoglobulins and the Chelex method. The most relevant applica-
tion—purification of histidine-tagged recombinant proteins—will be reviewed
439
440 Helena Block et al.
A
Protein
N N N
N N N
N N N
N N N
Ni-IDA
Ni2+
O
CH2
O
H2O CH2
N
B
Protein
N N N
N N N
N N N
N N N
Ni-NTA
Ni2+
O
CH2
O CH2
N
CH
O
C
Protein
N N N
N N N
N N N
N N N
O
Ni-TED
CH2 Ni2+
CH2
N CH2 N
CH2
CH2
Figure 27.1 Model of the interaction between residues in the His tag and the metal ion
in tri- (IDA), tetra- (NTA), and pentadentate IMAC ligands (TED).
442 Helena Block et al.
This ratio has turned out to be most effective for purification of His-tagged
proteins. Another tetradentate ligand is carboxymethyl aspartate (CM-Asp;
Chaga et al., 1999), commercially available as cobalt-charged Talon resin.
In contrast to tetradentate ligands, IDA coordinates a divalent ion with three
valencies (tridentate, coordination number 3, Fig. 27.1A) leaving three
valencies free for imidazole ring interaction while it is unclear whether
the third is sterically able to participate in the interaction. The coordination
number seems to play an important role regarding the quality of the purified
protein fraction. While protein recovery is usually similar between IDA-
and NTA-based chromatography (Fig. 27.2D), a higher leaching of metal
ions from IDA ligands compared to NTA is observed in general (Hochuli,
1989) and even increased under reducing conditions (Fig. 27.2C). Although
the metal content in the elution fractions (E in Fig. 27.2C) is higher but still
within the same order of magnitude, significantly more Ni2þ ions leach
from the IDA resin in the equilibration and wash steps (W in Fig. 27.2C).
Besides considerable metal leaching, purification of His-tagged proteins
using an IDA matrix frequently results in lower purity compared to
NTA-based purification (Fig. 27.2A and D).
150
[FU]
Ni-IDA
Ni-NTA
100 Ni-TED
50
C 0
10,000
24 26 28 30
1300
1000 [s]
470
log [Ni] (ppb)
100
100
54
10 4
2
1
W E W E W E
Ni-IDA Ni-NTA Ni-TED
D
EMG1 FYN1 JNK1 MAPKAP5 p38a PIM1
NTA IDA NTA IDA NTA IDA NTA IDA NTA IDA NTA IDA
L F WE FWE LFWEFWE LFWEFWE LFWEFWE LFWEFWE LFWEFWE
E a b g d
1 2 3 4 1 2 3 4 1 2 3 4 1 2 3 4 M
200
97
66
55
36
31
21
14
6
Figure 27.2 Purification of His-tagged proteins with NTA, IDA, and TED.
(A) H6-HIV-RT was expressed in E. coli BL21(DE3) and purified via Ni-IMAC in
the presence of 1 mM DTT under standard conditions (IDA, NTA; see Section 2.10 for
standard conditions) or according to the manufacturer’s recommendations (TED).
Corresponding aliquots of the IMAC elution fractions (E1, E2) were analyzed by
SDS–PAGE and Coomassie staining. (B) Bioanalyzer 2100 lab-on-a-chip analysis of
pooled elution fractions. The peaks from the electropherograms corresponding to H6-
HIV-RT were overlayed. Peak areas directly correlate to the protein amount in the
respective pool fraction. (C) Determination of the nickel content in wash (W) and
pooled elution fractions (E) of the H6-HIV-RT purifications described in (A) and (B).
Nickel was measured by ICP-MS (intercoupled plasma mass spectrometry) at Wessling
Laboratories, Bochum, Germany, and values are provided in mg/l (parts per billion,
ppb). (D) QIAgene constructs carrying optimized human genes were used for expres-
sion of the indicated proteins in E. coli BL21(DE3) LB cultures. Cleared lysates were
divided for purification of His-tagged proteins via Ni-NTA (NTA) and Ni-IDA (IDA),
respectively. Fractions were analyzed by SDS–PAGE and Coomassie staining as fol-
lows: L, cleared lysate; F, IMAC flow-through fraction; W, wash fraction; E, peak
elution fraction. (E) Fragments of H6-tagged proteins named a, b, g, and d expressed in
E. coli were purified under standard conditions using NTA and Cm-Asp tetradentate
ligands loaded with Ni2þ or Co2þ as follows: 1, Ni-NTA; 2, Co-NTA; 3, Co-CmAsp;
4, Ni-CmAsp. Aliquots from the peak elution fractions (2 ml each) were analyzed by
SDS–PAGE and Coomassie staining.
444 Helena Block et al.
The reason for the lower purity may be that leaching of metal from the
tridentate ligand generates charged groups which could act as a cation
exchanger and bind positively charged groups on the surface of proteins.
The lowest metal leaching is obtained if a pentadentate ligand is used
(Fig. 27.2C) which coordinates the ions extremely tightly, and such resins
may represent a valid alternative if low metal ion leaching into the protein
preparation is very important. However, in this case only one coordination
site remains for His tag binding and recovery of His-tagged protein is usually
considerably lower than with IDA or NTA (Fig. 27.2B).
The choice of the metal ion immobilized on the IMAC ligand depends
on the application. While trivalent cations such as Al3þ, Ga3þ, and Fe3þ
(Andersson and Porath, 1986; Muszynska et al., 1986; Posewitz and Tempst,
1999) or tetravalent Zr4þ (Zhou et al., 2006) are preferred for capture of
phosphoproteins and phosphopeptides, divalent Cu2þ, Ni2þ, Zn2þ, and
Co2þ ions are used for purification of His-tagged proteins. Combinations of
a tetradentate ligand that ensures strong immobilization, and a metal ion that
leaves two coordination sites free for interaction with biopolymers (Ni2þ,
Co2þ) has gained most acceptance and leads to similar recovery and purity
of eluted protein. As a typical result using such combinations, Fig. 27.2E
shows the purification of several protein fragments derived from different
genes that have been expressed and purified as His-tagged proteins by Ni2þ
and Co2þ immobilized on NTA and Cm-Asp tetradentate ligands.
2. IMAC Applications
Initially developed for purification of native proteins with an intrinsic
affinity to metal ions (Porath et al., 1975), IMAC has turned out to be a
technology with a very broad portfolio of applications. On the chro-
matographic purification side, the range of proteins was expanded from the
primary metalloproteins to antibodies, phosphorylated proteins, and recom-
binant His-tagged proteins. IMAC is being used in proteomics approaches
where fractions of the cellular protein pool are enriched and analyzed
differentially (phosphoproteome, metalloproteome) by mass spectrometrical
techniques; here, IMAC formats can be traditionally bead based or the ligand
can be used on functionalized surfaces such as SELDI (surface-enhanced laser
desorption/ionization) chips. Other chip-based applications include surface
plasmon resonance (SPR) and allow the immobilization of His-tagged
proteins for quantitative functional and kinetic investigations. In addition,
the IMAC principle has been used as an inhibitor depletion step prior to
PCR amplification of nucleic acids from complex samples such as blood in a
technology called Chelex (Walsh et al., 1991). The most relevant of the
IMAC 445
tris-Ni-NTA Ni-NTA
F W1 W2 W3 40 250 F W1 W2 W3 40 250 mM imidazole
97
66 His6-AKT1
36
21
14
Figure 27.3 Purification of His6-tagged AKT1 kinase using tris-Ni-NTA and Ni-NTA
magnetic beads. Human C-terminally His6-tagged AKT1 was expressed cell-free in an
insect cell-based lysate (EasyXpress Insect II system) in a 100 ml reaction volume using
a pIX4.0 vector construct; purification was performed using magnetic agarose beads
functionalized with tris-Ni-NTA (left panel) or Ni-NTA (right panel) under standard
conditions (see Section 2.10) in the presence of 0.05% (v/v) Tween-20 using a
magnet-equipped tube holder. Aliquots of purification fractions as follows were
analyzed by SDS–PAGE and silver staining: F, unbound protein; W1, 2, wash fractions
1 and 2 (10 mM imidazole); W3, wash fraction 3 (20 mM imidazole); 40 ! 250, elution-
fractions (40, 60, 100, and 250 mM imidazole, respectively). Identity of the purified His6-
tagged AKT1 protein was verified by anti-His Western blot analysis (data not shown).
Table 27.1 Reported His tag sequences (single letter amino acid sequence code)
15 TFIIAg
50
TBP
30
75
IRAK4
50
Figure 27.4 Effect of His tag position on protein expression. Proteins were expressed
from PCR product templates generated by two-step PCR using the EasyXpress Linear
Template Kit in E. coli- (A) and insect cell-derived (B) lysates (EasyXpress Protein
Synthesis and EasyXpress Insect II Kits, respectively). Initiator (adapter) primers in the
Linear Template Kit were designed in order to prevent formation of secondary struc-
ture in the translation initiation region on the mRNA and introduce the following tag
sequence(s): N-H6, N-terminal His6 tag; C-H6, C-terminal His6 tag; N-SII, N-terminal
Strep II tag; C-SII, C-terminal Strep II tag; N-H6/C-SII, N-terminal His6 and
C-terminal Strep II tags; N-SII/C-H6, N-terminal Strep II and C-terminal His6 tags.
Corresponding aliquots of total (T) and soluble protein (S, supernatant after centrifu-
gation at 15,000 g for 10 min) were loaded on a SDS gel. Protein bands were
visualized by Western blot analysis using a mixture of Penta anti-His and anti-Strep
tag antibodies. Protein sizes are (kDa) TNFa, 21; TBP, 38; TFIIAab, 55; TFIIAg, 12.5;
MKK3, 39; IRAK4, 55. M, His-tagged protein size markers (kDa).
IMAC 451
His tags. However, in some cases such as the one of IRAK4 the C-terminal
His tag has a more pronounced effect on both expression rate and solubility
(Fig. 27.4). Recently, expression of an insect toxin in E. coli was reported
where the similar observation of higher solubility and thermostability of the
C-terminally His-tagged form was made (Xu et al., 2008). The authors
discussed that the C-terminal tag stabilized the overall protein structure.
Other groups found the His tag to contribute slightly negative to solubility
when compared to the untagged protein but to improve yield when fused to
the C-terminus (Woestenenk et al., 2004). All in all, these data suggest that an
evaluation of at least N- and C-terminally tagged variants of a protein will
increase the chance to obtain reasonable expression and quality of a recom-
binant protein. For expression of proteins to be secreted, tags should be
placed at the C-terminus to prevent interference with membrane trafficking.
A B
0 1 5 10 mM DTT
M L F W E1E2 E3 E4 E2 E3 E4 E2 E3 E4 E2 E3 E4
200
116/97
66 His6- M C NTC
55 HIV-RT 0 1 5 10
36
31
b -actin
21
14
6
Figure 27.5 IMAC under reducing conditions. (A) His6-tagged HIV-1 reverse tran-
scriptase (RT) was purified under standard conditions in the presence of the indicated
DTT concentrations. Aliquots of each chromatographic fraction of the purification
without DTT (M, markers; L, lysate; F, flow-through; W, wash; E, elution fraction)
and of corresponding volumes of elution fractions 2–4 containing DTT were analyzed
by SDS–PAGE and Coomassie staining. Elution fractions were pooled and analyzed for
nickel content by ICP-MS at Wessling Laboratories, Bochum, Germany; nickel
concentrations are given in mg/l (parts per billion, ppb). (B) Equal amounts of HIV-1
RT purified in the presence of the indicated concentrations of DTT were used in
duplicates to reverse transcribe a 1.5 kb b-actin cDNA which was subsequently ampli-
fied by PCR and the PCR products were analyzed by agarose gel electrophoresis and
ethidium bromide staining. Reducing conditions of at least 1 mM DTT were found to
be required for full RT activity (compare weaker amplification with RT purified in the
absence of DTT). C, Omniscript positive control; an equal amount of Omniscript RT
protein was used; NTC, no template negative control.
(Fig. 27.5B) and quantitative real-time RT-PCR (data not shown) show no
inhibitory effect which could origin from high heavy-metal-ion concentra-
tions. Even though nickel ions may become reduced by DTT leading to a
color change of the resin bed they do not increasingly leach from the ligand
(Fig. 27.5A), and resins processed under reducing conditions can be repeat-
edly reused and regenerated (data not shown). These findings suggest that
despite a color change as a consequence of nickel reduction by DTT the
resin is still functional. TCEP, a different, nonthio-based reducing reagent is
more and more replacing DTT and b-ME in protein purification by IMAC
as it seems to be more selective to reduction of disulfide bonds, is odorless
and more stable in aqueous solution. We recommend the use of TCEP with
Ni-NTA chromatography in a concentration of 1–5 mM.
agarose and, in addition, have good pressure stability which makes them
suitable for high-resolution HPLC applications but the silica resins
frequently suffer from low binding capacity and limited resistance to high
pH sanitization procedures. A very recent method that avoids the need to
use solid chromatographic supports completely is called affinity precipita-
tion (Hilbrig and Freitag, 2003) and has been applied to IMAC (Matiasson
et al., 2007). Here, the IMAC ligand is chemically coupled to a responsive
polymer which, after binding to the His-tagged protein, can be aggregated
upon change of environmental conditions such as pH or temperature and
can thus be precipitated by centrifugation. Protocols for its use are still
relatively complicated but as soon as robust and easy to use commercial
materials are available this method may have the potential to play an
important role in IMAC applications. Using a ligand in solution could
overcome steric hindrance of the binding of some His-tagged proteins to
an immobilized ligand as well as mass transport limitations of porous
chromatographic media. Moreover, it is in line with a trend in industrial-
scale chromatography toward single-use disposable materials.
Yet another approach has recently been reported that can be applied for
protein separation from lysates after cell-free expression (Kim et al., 2006).
An E. coli-derived lysate was preincubated with Ni-NTA magnetic agarose
beads to remove proteins with affinity to Ni-NTA prior to template
addition and protein expression; the expression capacity of the S30 extract
was found to remain unaltered and the Ni-NTA purified His-tagged
protein fractions to be of higher purity than without pretreatment.
While the aforementioned strategies to improve the purity of MAC
protein preps have proven useful in many cases they are not generally
applicable and successful. There is a method, however, that almost meets
the criterium of universal applicability regarding improvement of purity,
and it has the additional benefit of resulting in a protein native or near-
native structure: proteolytic His tag cleavage using a His-tagged protease
followed by reverse IMAC (Block et al., 2008). This strategy overcomes the
copurification issue by passing the proteolytically processed protein under
similar or identical conditions over the same column, and the proteins that
bound to the IMAC resin as impurities in the initial purification step will
bind to the same resin again while the cleaved, that is untagged, target
protein is collected in the flow-through fraction (reverse or subtractive
IMAC mode). It can be performed with both exo- and endoproteases that
themselves carry an (uncleavable) His-tag (Nilsson et al., 1997; Polayes et al.,
2008) but the exoproteolytical approach has the advantage that it is faster
and results in a protein with native structure with no vector-derived amino
acids (Arnau et al., 2006; Block et al., 2008; Pedersen et al., 1999). This
approach is especially suitable for demanding downstream applications such
as protein crystallization or biopharmaceutical production. Notably, the
method requires only a single chromatography column to achieve an
458 Helena Block et al.
C His6-TNFa TNFα
B E F
mAU
pH 3 pH 10 pH 3 pH 10 mAU
800 82.36
800 85.12
600
600
400
400
His6-
200 TNFa TNFa 200
0
0
0 20 40 60 80 100 ml
0 20 40 60 80 100 ml
Figure 27.6 Removal of copurifying proteins by His tag cleavage and reverse IMAC. (A) His6-TNFa expressed in E. coli was purified via Ni-NTA
Superflow and processed using the TAGZyme exoproteolytic system as described (Sch€afer et al., 2002a). (B, E) 2D gel electrophoresis and silver
staining of His6-TNFa and TNFa, respectively, was performed as described (Block et al., 2008). The subband pattern in the first dimension between
pI 6.7 and 5.8 for both His6-TNFa and TNFa is in accordance with the report for TNFa produced in yeast (Eck et al., 1988). (C, F) Analytical size
exclusion chromatography (SEC) on HR 10/30 Superdex 200 was run with 1 TAGZyme buffer (Sch€afer et al., 2002a). (D, G) His6-TNFa
tetragonal crystal (D) formed in 2.7 M MgSO4, MES, pH 5.5 and diffracted to 2.5 Å (homelab X-ray source, FR591 Nonius Bruker);
TNFa rhombohedral crystal (G) formed in 1.8 M NH4SO3, 200 mM Tris–HCl, pH 7.8 and diffracted to 2.0 Å (ESRF synchrotron). The size of
typical crystals (mm) is indicated by the bar in (D) and (G).
460 Helena Block et al.
5
SMS1 PK
1 1
p3 EB D
TN g F
IF -6 1A
A
R C
2
N S
D
IL G1
C R
ET PK
Ju C1
Y C2
IF -C
FY Fa
IL B
BI 2
C K1
PI a
EM 1
A
A
C N
N 8a
SM1
JN 1
Fκ
M
IL -4
N
M -7
M
Len
A
D
Y
f
M
PI
200
116/97
66
55
36
31
21
14
A B
M
16
A
M
M L R F W E1 E2
G
D
y6
LD
G
FC
M
N
D
D
O
80 72
70
60 55
50 36
40 28 His6-
30
CAV1
His6-
20 CAV1 17
15 11
A B
40
66 35.6
His6-YY1
55
30
36 25.6
[Me2+] (mM)
31
20
14.2
21
10
14
0.9
0
M L FWE FWE Analyzed Me2+ ion: Ni Zn Ni Zn
Ni-NTA Zn-NTA IMAC resin : Ni-NTA Zn-NTA
3. Conclusions
We have presented the wide variety of applications the IMAC principle
offers for research in general and for production of His-tagged proteins
in particular. Its robustness and versatility are the reasons why IMAC has
468 Helena Block et al.
ACKNOWLEDGMENTS
The authors thank Jacob Piehler for valuable discussions regarding the manuscript and his
support during the transfer of the tris-NTA ligand synthesis procedures. Furthermore, we
thank Annette Zacharias-Koch for excellent technical assistance and her support in prepara-
tion of the manuscript. Part of the work presented here was performed under a grant of the
German Ministry of Education and Research (BMBF, grant no. 0313965B).
REFERENCES
Acton, T. B., Gunsalus, K. C., Xiao, R., Ma, L. C., Aramini, J., Baran, M. C., Chiang, Y.-W.,
Climent, T., Cooper, B., Denissova, N. G., Douglas, S. M., Everett, J. K., et al. (2005).
Robotic cloning and protein production platform of the northeast structural genomics
consortium. Methods Enzymol. 394, 210–243.
Andersson, L., and Porath, J. (1986). Isolation of phosphoproteins by immobilized metal
(Fe3þ)affinity-chromatography. Anal. Biochem. 154, 250–254.
Arnau, J., Lauritzen, C., Petersen, G. E., and Pedersen, J. (2006). Current strategies for the
use of affinity tags and tag removal for the purification of recombinant proteins. Protein
Expr. Purif. 48, 1–13.
Biocompare (2006). Protein chromatography: Tools for protein expression and purification.
Biocompare Surveys and Reports, Biocompare Inc., July 10.
Block, H., Kubicek, J., Labahn, J., Roth, U., and Schäfer, F. (2008). Production and
comprehensive quality control of recombinant human Interleukin-1b: A case study for
a process development strategy. Protein Expr. Purif. 27, 244–254.
Boden, V., Winzerling, J. J., Vijayalakshmi, M., and Porath, J. (1995). Rapid one-step
purification of goat immunoglobulins by immobilized metal ion affinity chromatogra-
phy. J. Immunol. Methods 181, 225–232.
Bolanos-Garcia, V. M., and Davies, O. R. (2006). Structural analysis and classification of
native proteins from E. coli commonly co-purified by immobilized metal affinity
chromatography. Biochim. Biophys. Acta 1760, 1304–1313.
Bornhorst, J. A., and Falke, J. J. (2000). Purification of protein using polyhistidine affinity
tags. Methods Enzymol. 326, 245–254.
Braun, P., Hu, Y., Shen, B., Halleck, A., Koundinya, M., Harlow, E., and LaBaer, J. (2002).
Proteome-scale purification of human proteins from bacteria. Proc. Natl. Acad. Sci. USA
99, 2654–2659.
Busso, D., Kim, R., and Kim, S.-H. (2003). J. Biochem. Biophys. Methods 55, 233–240.
IMAC 469
Büssow, K., Nordhoff, E., Lübbert, C., Lehrach, H., and Walter, G. (2000). A human
cDNA library for high-throughput protein expression screening. Genomics 65, 1–8.
Byrne, B., and Jormakka, M. (2006). Solubilization and purification of membrane proteins.
In ‘‘Structural Genomics on membrane proteins’’ (K. Lundstrom, ed.), pp. 179–198.
CRC Press, Taylor and Francis, Boca Raton.
Cass, B., Pham, O. L., Kamen, A., and Durocher, Y. (2005). Purification of recombinant
proteins from mammalian cell culture using a generic double-affinity chromatography
scheme. Protein Expr. Purif. 40, 77–85.
Cèbe, R., and Geiser, M. (2006). Rapid and easy thermodynamic optimization of the 50 -end
of mRNA dramatically increases the level of wild type protein expression in Escherichia
coli. Protein Expr. Purif. 45, 374–380.
Chaga, G., Hopp, J., and Nelson, P. (1999). Immobilized metal ion affinity chromatography
on Co2þ-carboxymethylaspartate–agarose Superflow, as demonstrated by one-step
purification of lactate dehydrogenase from chicken breast muscle. Biotechnol. Appl.
Biochem. 29, 19–24.
Crowe, J., Döbeli, H., Gentz, R., Hochuli, E., Stüber, D., and Henco, K. (1994). 6xHis-
Ni-NTA chromatography as a superior technique in recombinant protein expression/
purification. Methods Mol. Biol. 31, 371–381.
Dalbge, H., Bayne, S., and Pedersen, J. (1990). In vivo processing of N-terminal methionine
in E. coli. FEBS Lett 266, 1–3.
DeJong, J., and Roeder, R. G. (1993). A single cDNA, hTFIIA/a, encodes both the p35 and
p19 subunits of human TFIIA. Genes Dev. 7, 2220–2234.
Derewenda, Z. S. (2004). The use of recombinant methods and molecular engineering in
protein crystallization. Methods 34, 354–363.
Drews, J. (2000). Drug discovery: A historical perspective. Science 287, 1960–1964.
Eck, M. J., Beutler, B., Kuo, G., Merryweather, J. P., and Sprang, S. R. (1988). Crystalliza-
tion of trimeric recombinant human tumor necrosis factor (catechin). J. Biol. Chem. 263,
12816–12819.
Eshaghi, S., Hedrén, M., Ignatushchenko Abdel Nasser, M., Hammarberg, T., Thomell, A.,
and Nordlund, P. (2005). An efficient strategy for high-throughput expression of recom-
binant integral membrane proteins. Protein Sci. 14, 676–683.
Franken, K. L. M. C., Hiemstra, H. S., van Meijgaarden, K. E., Subronto, Y., den
Hartigh, J., Ottenhoff, T. H. M., and Drijfthout, J. W. (2000). Purification of His-tagged
proteins by immobilized chelate affinity chromatography: The benefits from the use of
organic solvents. Protein Expr. Purif. 18, 95–99.
Garcı́a González, L. A., Rodrigo Tapia, J. P., Sánchez Lazo, P., Ramos, S., and Suárez
Nieto, C. (2004). DNA extraction Using Chelex resin for the oncogenic amplification
analysis in head and neck tumors. Acta Otorrinolaringol. Esp. 55, 139–144.
Garzia, L., André, A., Amoresano, A., D’Angelo, A., Martusciello, R., Cirulli, C.,
Tsurumi, T., Marino, G., and Zollo, M. (2003). Method to express and purify nm23–
H2 protein from baculovirus-infected cells. BioTechniques 35, 384–391.
Gill, P., Kimpton, C. P., and Sullivan, K. (1992). A rapid polymerase chain reaction method
for identifying fixed specimens. Electrophoresis 13, 173–175.
Gräslund, S., Nordlund, P., Weigelt, J., Bray, J., Gileadi, O., Knapp, S., Oppermann, U.,
Arrowsmith, C., Hui, R., Ming, J., dhe-Paganon, S., Park, H.-W., et al. (2008). Protein
production and purification. Nat. Methods 5, 135–146.
Grisshammer, R., and Tucker, J. (1997). Quantitative evaluation of neurotensin receptor
purification by immobilized metal affinity chromatography. Protein Expr. Purif. 11,
53–60.
Hale, J. E., and Beidler, D. E. (1994). Purification of humanized murine and murine
monoclonal antibodies using immobilized metal-affinity chromatography. Anal. Biochem.
222, 29–33.
470 Helena Block et al.
In ‘‘High Throughput Protein Expression and Purification’’ (S. A. Doyle, ed.). Humana
Press, Springer.
Lanio, T., Jeltsch, A., and Pingoud, A. (2000). Automated purification of His6-tagged
proteins allows exhaustive screening of libraries generated by random mutagenesis.
BioTechniques 29, 338–342.
Lata, S., and Piehler, J. (2005). Stable and functional immobilization of histidine-tagged
proteins via multivalent chelator headgroups on a molecular poly(ethylene glycol) brush.
Anal. Chem. 77, 1096–1105.
Lata, S., Gavutis, M., Tampé, R., and Piehler, J. (2006). Specific and stable fluorescence
labeling of histidine-tagged proteins for dissecting multi-protein complex formation.
J. Am. Chem. Soc. 128, 2365–2372.
Lesley, S. A. (2001). High-throughput proteomics: Protein expression and purification in
the postgenomic world. Protein Expr. Purif. 22, 159–164.
Levison, P. R., Badger, S. E., Jones, R. M. H., Toome, D. W., Streater, M.,
Pathirana, N. D., and Wheeler, S. (1995). Validation studies in the regeneration of
ion-exchange celluloses. J. Chromatogr. A 702, 59–68.
Lewinson, O., Lee, A. T., and Douglas, C. R. (2008). The funnel approach to the
precrystallization production of membrane proteins. J. Mol. Biol. 377, 62–73.
Loo, J. A. (2003). The tools of proteomics. Adv. Protein Chem. 65, 353–369.
Lv, G. S., Hua, G.C, and Fu, X. Y. (2003). Expression of milk-derived antihypertensive
peptide in Escherichia coli. J. Dairy Sci. 86, 1927–1931.
Ma, D., Watanabe, H., Mermelstein, F., Admon, A., Oguri, K., Sun, X., Wada, T.,
Imai, T., Shiroya, T., Reinberg, D., and Handa, H. (1993). Isolation of a cDNA
encoding the largest subunit of TFIIA reveals functions important for activated transcrip-
tion. Genes Dev. 7, 2246–2257.
Mateo, C., Fernandez-Lorente, G., Pessela, B. C. C., Vian, A., Carrascosa, A. V.,
Garcia, J. L., Fernandez-Lafuente, R., and Guisan, J. M. (2001). Affinity chromatography
of polyhistidine tagged enzymes: New dextran-coated immobilized metal ion affinity
chromatography matrices for prevention of undesired multipoint adsorptions. J. Chro-
matogr. A 915, 97–106.
Matiasson, B., Kumar, A., Ivanov, A. E., and Galaev, I. Y. (2007). Metal-chelate affinity
precipitation of proteins using responsive polymers. Nat. Protocols 2, 213–220.
Mészárosová, K., Tishchenko, G., Bouchal, K., and Bleha, M. (2003). Immobilized-metal
affinity sorbents based on hydrophilic methacrylate polymers and their interaction with
immunoglobulins. React. Funct. Polym. 56, 27–35.
Mohanty, A. K., and Wiener, M. C. (2004). Membrane protein expression and production:
Effects of polyhistidine tag length and position. Protein Expr. Purif. 33, 311–325.
Muszynska, G., Andersson, L., and Porath, J. (1986). Selective adsorption of phosphopro-
teins on gel-immobilized ferric chelate. Biochemistry 25, 6850–6853.
Nieba, L., Nieba-Axamann, S. E., Persson, A., Hämäläinen, M., Edebratt, F., Hansson, A.,
Lidholm, J., Magnusson, K., Karlsson, A. F., and Plückthun, A. (1997). Biacore analysis
of histidine-tagged proteins using a chelating NTA sensor chip. Anal. Biochem. 252,
217–228.
Nilsson, J., Ståhl, S., Lundeberg, J., Uhlén, M., and Nygren, P.-A. (1997). Affinity fusion
strategies for detection, purification, and immobilization of recombinant proteins. Protein
Expr. Purif. 11, 1–16.
Padan, E., Venturi, M., Michel, H., and Hunte, C. (1998). Production and characterization
of monoclonal antibodies directed against native epitopes of NhaH, the Naþ/Hþ anti-
porter of E. coli. FEBS Lett. 441, 53–58.
Pedersen, J., Lauritzen, C., Madsen, M. T., and Dahl, S. W. (1999). Removal of N-terminal
polyhistidine tags from recombinant proteins using engineered aminopeptidases. Protein
Expr. Purif. 15, 389–400.
472 Helena Block et al.
Polayes, D.A., Parks, T.D., Johnston, S.A., and Dougherty, W.G. (2008). Application of
TEV protease in protein production. Methods Mol. Med. 13, 169–183. In ‘‘Molecular
Diagnosis of Infectious Diseases’’ (U. Reischl. ed.). Humana Press, Springer.
Porath, J., and Olin, B. (1983). Immobilized metal ion affinity adsorption and immobilized
metal ion affinity chromatography of biomaterials. Serum protein affinities for
gel-immobilized iron and nickel ions. Biochemistry 29, 1621–1630.
Porath, J., Carlsson, J., Olsson, I., and Belfrage, G. (1975). Metal chelate affinity chroma-
tography, a new approach to protein fractionation. Nature 258, 598–599.
Posewitz, M. C., and Tempst, P. (1999). Immobilized gallium (III) affinity chromatography
of phosphopeptides. Anal. Chem. 71, 2883–2892.
Prinz, B., Schultchen, J., Rydzewski, R., Holz, C., Boettner, M., Stahl, U., and Lang, C.
(2004). Establishing a versatile fermentation procedure for human proteins expressed in
the yeasts Saccharomyces cerevisiae and Pichia pastoris for structural genomics. J. Struct. Funct.
Genomics 5, 29–44.
QIAGEN (2003). The QIAexpressionist. Handbook for High-Level Expression and Purifi-
cation of 6xHis-Tagged Proteins. 5th edn. Hilden, Germany: QIAGEN, Hilden,
Germany.
Reichel, A., Schaible, D., Al Furoukh, N., Cohen, M., Schreiber, G., and Piehler, J. (2007).
Noncovalent, site-specific biotinylation of histidine-tagged proteins. Anal. Chem. 79,
8590–8600.
Rumbley, J. N., Furlong Nickels, E., and Gennis, R. B. (1997). One-step purification of
cytochrome bo3 from Escherichia coli and demonstration that associated quinone is not
required for the structural integrity of the oxidase. Biochim. Biophys. Acta 1340, 131–142.
Schäfer, F., Blümer, J., Römer, U., and Steinert, K. (2000). Ni-NTA for large-scale
processes—systematic investigation of separation characteristics, storage and CIP condi-
tions, and leaching. QIAGEN. QIAGEN News 4, 11–15. Available from http://www1.
qiagen.com/literature/qiagennews/0400/Ni-NTA%20for%20large-scale.pdf
Schäfer, F., Schäfer, A., and Steinert, K. (2002a). A highly specific system for efficient
enzymatic removal of tags from recombinant proteins. J. Biomol. Tech. 13, 158–171.
Schäfer, F., Römer, U., Emmerlich, M., Blümer, J., Lubenow, H., and Steinert, K. (2002b).
Automated high-throughput purification of 6xHis-tagged proteins. J. Biomol. Tech. 13,
131–142.
Scheich, C., Sievert, V., and Büssow, K. (2003). An automated method for high-throughput
protein purification applied to a comparison of His-tag and GST-tag affinity chromatog-
raphy. BMC Biotechnol. 3, 12. Available from www.biomedcentral.com/1472-6750/3/12
Schmitt, J., Hess, H., and Stunnenberg, H. G. (1993). Affinity purification of histidine-
tagged proteins. Mol. Biol. Rep. 18, 223–230.
Serpa, G., Augusto, E. F. P., Tamashiro, W. M. S. C., Ribeiro, M. B., Miranda, E. A., and
Bueno, S. M. A. (2005). Evaluation of immobilized metal membrane affinity chroma-
tography for purification of an immunoglobulin G1 monoclonal antibody. J. Chromatogr. B
816, 259–268.
Shi, W., and Chance, M. R. (2008). Metallomics and metalloproteomics. Cell. Mol. Life Sci.
65, 3040–3048.
Shi, Y., Seto, E., Chang, L.-S., and Shenk, T. (1991). Transcriptional repression by YY1,
a human GLI-Krüppel-related protein, and relief of repression by adenovirus E1A
protein. Cell 67, 377–388.
Slentz, B. E., Penner, N. A., and Regnier, F. E. (2003). Protein proteolysis and the
multi-dimensional electrochromatographic separation of histidine-containing peptide
fragments. J. Chromatogr. A 984, 97–103.
Stasyk, T., and Huber, L. A. (2004). Zooming in: Fractionation strategies in proteomics.
Proteomics 4, 3704–3716.
IMAC 473
Steen, J., Uhlén, M., Hober, S., and Ottosson, J. (2006). High-throughput protein purifica-
tion using an automated set-up for high-yield affinity chromatography. Protein Expr.
Purif. 46, 173–178.
Stowers, A. W., Zhang, Y., Shimp, R. L., and Kaslow, D. C. (2001). Structural conformers
produced during malaria vaccine production in yeast. Yeast 18, 137–150.
Strömberg, P., Rotticci-Mulder, J., Björnestedt, R., and Schmidt, S. R. (2005). Preparative
parallel protein purification (P4). J. Chromatogr. B 818, 11–18.
Sulkowski, E. (1985). Purification of proteins by IMAC. Trends Biotechnol. 3, 1–7.
Sun, X., Chiu, J.-F., and He, Q.-Y. (2005). Application of immobilized metal affinity
chromatography in proteomics. Expert Rev. Proteomics 2, 649–657.
Svensson, J., Andersson, C., Reseland, J. E., Lyngstadaas, P., and Bülow, L. (2006). Histidine
tag fusions increases expression levels of active recombinant amelogenin in Escherichia coli.
Protein Expr. Purif. 48, 134–141.
Vancan, S., Miranda, E. A., and Bueno, S. M. A. (2002). IMAC of human IgG: Studies with
IDA-immobilized copper, nickel, zinc, and cobalt ions and different buffer systems.
Process Biochem. 37, 573–579.
Walsh, P. S., Metzger, D. A., and Higuchi, R. (1991). Chelex 100 as a medium for simple
extraction of DNA for PCR-based typing from forensic material. BioTechniques 10,
506–513.
Woestenenk, E. A., Hammarström, M., van den Berg, S., Härd, T., and Berglund, H.
(2004). His tag effect on solubility of human proteins produced in Escherichia coli:
Comparison between four expression vectors. J. Struct. Funct. Genomics 5, 217–229.
Xu, C.-G., Fan, X.-J., Fu, Y.-J., and Liang, A.-H. (2008). Effect of the His-tag on the
production of soluble and functional Buthus martensii Karsch insect toxin. Protein Expr.
Purif. 59, 103–109.
Zhaohua, H., Park, J. I., Watson, D. S., Hwang, P., and Szoka, F. C. (2006). Facile synthesis
of multivalent nitrilotriacetic acid (NTA) and NTA conjugates for analytical and drug
delivery applications. Bioconjug. Chem. 17, 1592–1600.
Zhou, H., Xu, S., Ye, M., Feng, S., Pan, C., Jiang, X., Li, X., Han, G., Fu, Y., and Zou, H.
(2006). Zirconium phosphonate-modified porous silicon for highly specific capture of
phosphopeptide and MALDI-TOF MS analysis. J. Proteome Res. 5, 2431–2437.
C H A P T E R T W E N T Y- E I G H T
Contents
1. Introduction 476
2. Polyol-Responsive Monoclonal Antibodies 477
2.1. Properties of a PR-mAb 478
2.2. Source of mAbs 479
2.3. Identification of PR-mAbs by ELISA-elution assay 480
2.4. Producing mAbs in continuous culture 482
2.5. Purification of the antibody 485
2.6. Immobilization of PR-mAbs on a chromatography support 487
2.7. Purification of proteins with PR-mAbs 488
2.8. Purification of proteins using cross-reacting PR-mAbs 490
2.9. Use of epitopes of PR-mAbs as purification tags 491
3. Conclusions 492
Disclosure 493
References 493
Abstract
Immunoaffinity chromatography is a powerful tool for purification of proteins
and protein complexes. The availability of monoclonal antibodies (mAbs) has
revolutionized the field of immunoaffinity chromatography by providing a con-
tinuous supply of highly uniform antibody. Before the availability of mAbs, the
recovery of the target protein from immobilized polyclonal antibodies usually
required very harsh, often denaturing conditions. Although harsh conditions are
often still used to disrupt the antigen–antibody interaction when using a mAb,
various methods have been developed to exploit the uniformity of the antigen–
antibody reaction in order to identify agents or conditions that gently disrupt
McArdle Laboratory for Cancer Research, University of Wisconsin–Madison, Madison, Wisconsin, USA
475
476 Nancy E. Thompson et al.
this interaction and thus result in higher recovery of active protein from immu-
noaffinity chromatography. We discuss here the use of a specific type of
monoclonal antibody that we have designated ‘‘polyol-responsive monoclonal
antibodies’’ (PR-mAbs). These are naturally occurring mAbs that have high
affinity for the antigen under binding conditions, but have low affinity in the
presence of a combination of low molecular weight hydroxylated compounds
(polyols) and nonchaotropic salts. Therefore, these PR-mAbs can be used for
gentle immunoaffinity chromatography. PR-mAbs can be easily identified and
adapted to a powerful protein purification method for a target protein.
1. Introduction
All forms of affinity chromatography are defined by a specific interac-
tion between two components that allows the purification of one of the
components. Immunoaffinity chromatography is a subset of the affinity
chromatography principle where the specific interaction of an antigen
with an antibody is employed (for review, see Subramanian, 2002). Immu-
noaffinity chromatography is really a scaled up extension of an immunopre-
cipitation procedure, except that one of the components (generally the
antigen) is recovered after the chromatography as an active protein.
The ability to produce monoclonal antibodies (mAbs) has revolutio-
nized the field of immunochemistry (for review, see Nelson et al., 2000).
When considering immunoaffinity chromatography, two features give
mAbs an advantage over polyclonal antibodies derived from immune
serum. First, the mAb is a reproducible reagent that can be prepared in
large quantities. Second, a mAb is a homogeneous population that responds
uniformly to an eluting reagent. Generally, the purified antibody is conju-
gated to some type of bead and the antigen-containing solution is applied to
the bead (in batch or in a column). After washing away unbound or loosely
bound material, the antigen is eluted from the bead.
The elution step is usually the most difficult obstacle to overcome in
developing an immunoaffinity chromatography procedure. Antigen–anti-
body interactions are generally a result of a combination of ionic, hydro-
phobic, and hydrogen bonds formed between amino acids in the specific
antigenic determinant of the antigen (epitope) and the protein loops of
the complementarity determining regions (CDRs), which are located in the
variable regions of the heavy and light chain of the antibody molecule.
The ideal way to gently elute the antigen is by competition with a peptide
containing the epitope for the antibody. However, the epitope is not always
known, and a peptide is not always available or is too expensive to synthe-
size in the quantities needed. In addition, sometimes the antibody reacts
with a discontinuous epitope, and the epitope cannot be mimicked by a
Immunoaffinity Chromatography 477
synthetic peptide. In these cases, the antigen is usually eluted with very
harsh conditions (high or low pH values, denaturants such as urea, or an
ionic detergent.) that can inactivate the protein.
Table 28.1 Polyol-responsive monoclonal antibodies that have been used to purify
proteins involved with transcription
2.3.1. Comments
1. Screening for PR-mAbs can be performed at the master-well stage,
immediately after the hybridomas are screened for specific antibody
Immunoaffinity Chromatography 481
A
1
mAb NT63 mAb 8RB13 mAb NT73
0.9
0.8
Absorbance value
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
PG
NaCl
NaCl/PG
TE buffer
NaCl
PG
NaCl/PG
TE buffer
NaCl
PG
NaCl/PG
TE buffer
B
1.4
0% PG
1.2
Absorbance value
1
20% PG
0.8
0.6
0.4
0.2 30% PG
40% PG
0
0 0.2 0.4 0.6 0.8 1
NaCl (M )
Figure 28.1 ELISA-elution assay to identify and characterize PR-mAbs. Each well of
the microtiter plate was coated with 100 ng core RNAP. (A) ELISA-elution assay of
previously identified PR-mAb. mAb NT63 is a control mAb that reacts with E. coli
RNAP (Thompson et al., 1992). mAbs 8RB13 and NT73 are PR-mAbs (Bergendahl
et al., 2003; Thompson et al., 1992). mAb NT63 did not elute from the antigen with
0.75 M NaCl and 40% propylene glycol (NaCl/PG), but mAbs 8RB13 and NT73 did
elute with the salt/polyol combination. None of the mAbs eluted well with 0.75 M
NaCl or 40% propylene glycol (PG) alone. (B) ELISA-elution assay using mAb 8RB13
and varying concentrations of both NaCl (0–1.0 M) and propylene glycol (0–40%).
A
Port to nutrient chamber
Green cap
White cap
B
Cell supernatant
AS supernatant
Markers
F10
F11
F2
F3
F4
F5
F6
F7
F8
F9
KDa
160
110
80
60
HC
50
40
30
LC
20
15
10
3.5
1 2 3 4 5 6 7 8 9 10 11 12 13 14
placed in the cell chamber. Suspend the cell preparation and aspirate into
a 10-ml serological pipette. With the green cap loose, inoculate 5 ml
suspension into cell compartment (white cap) by inserting the pipette
484 Nancy E. Thompson et al.
firmly into the cell compartment port. Remove trapped air bubbles by
pipetting the fluid up and down slowly to allow the bubbles to rise prior
to returning only the fluid to the cell chamber. Replace and tighten the
white cap. Add 350 ml of nutrient medium to the nutrient compartment
and completely tighten the green cap.
4. Cell compartment harvest and culture maintenance: Every 3–7 days remove
nutrient medium from nutrient medium compartment. With green cap
loose, insert 10 ml pipette into the cell compartment and pipette fluid up
and down to thoroughly mix the cells. Then remove entire cell com-
partment liquid volume to a centrifuge tube. Volume may be greater
than 5 ml due to osmotic flux. Take a sample for a cell count and cell
viability using a hemocytometer and trypan blue. Centrifuge the mate-
rial removed from the cell compartment to pellet the cells. Remove and
save the medium, which contains the mAbs; this can be frozen for
purification at a later date. Suspend the cells with fresh complete
medium. Depending upon the initial inoculation density, growth rate,
and maintenance frequency the cells may need to be split back (usually
1:2–1:4) at this point. With the green cap loosened, return 5 ml of cells
back to the cell compartment (white cap). Remove air bubbles and
completely tighten white cap. Add 350 ml of nutrient medium to
nutrient compartment and completely tighten green cap.
5. The Integra CL 350 flask is harvested every 3–7 days after the culture has
been established. Harvest intervals depend on the growth rate of the
hybridoma and the ability of the hybridoma to adapt to the flask
environment. This trait seems to be cell line dependent.
2.4.1. Comments
1. For hybridomas that grow well in the CL 350 flask, approximately 0.5–
1 mg of mAb is present in each ml of harvested cell culture supernatant.
The continuous culture can usually be maintained for about 1 month.
2. Liquid handling: Warm medium to 37 C in a water bath. This helps to
prevent condensation on the flask and temperature shocking the cells.
When adding or removing liquid from the cell compartment (white
cap), always loosen the green cap of the nutrient medium compartment
first to prevent air lock. Always tighten both white and green caps before
placing the flask into the incubator. The use of 10 ml serological pipettes
is recommended for all cell compartment manipulations. The medium
in the nutrient medium chamber can be exchanged by aspirating the
medium out and pouring fresh medium in.
3. The minimum cell concentration is 1.5 106 cells/ml for the inoculum
(Step 2). We have not been successful in reducing the need for fetal
bovine serum in the nutrient medium. Some of the new serum-free
media on the market may be used as the nutrient medium.
Immunoaffinity Chromatography 485
4. It is important to track cell numbers and health (Step 3). This helps to
determine whether to split cells to reduce numbers once the total viable
cell count is greater than 100 106 cells. If the cell viability is greatly
reduced, the maintenance frequency needs to be increased.
5. Because the cells are split back every time the culture is harvested, the
percentage of viable cells (Step 4) will decrease during the continuous
culture due to death of the cells. It is not unusual to end up with only
30–40% viability at the end of the culture period.
6. Some hybridomas are unstable in this long-term, continuous culture and
lose the ability to produce the mAb. Therefore, the antibody production
should be monitored during the culturing. We do this by a standard
ELISA.
7. Information on CELLine flasks can be obtained at: www.integra-
biosciences.com/celline.
2.5.1. Comments
1. The preparation of the DE52 is particularly important. Although the
manufacturer states that precycling is not necessary, we have found that
precycling greatly improves the performance of the chromatography.
Ten grams of resin should be resuspended in 100 ml of water and
washed several times with 100 ml of water, removing the fines (the
small particles that do not settle easily) with each washing. The resin is
then treated with 100 ml of 0.1 M HCl for 30 min. The acid is decanted
and the resin washed at least three times with 100 ml of water for each
wash. After the last wash is decanted, the resin is treated with 0.1 M
NaOH for 30 min. The base is decanted, and the resin is washed at least
three times with 100 ml of water for each wash. The resin is then
washed three times with 100 ml of the antibody buffer and resuspended
in the same buffer. The pH is checked with pH paper, and NaN3 is
added to 0.02%. Resin is distributed to disposable tubes and stored in
the refrigerator.
2. If the mAb is made in ascites fluid, the product is not as clean (approxi-
mately 80–90% pure), but the impurities do not seem to interfere with
the performance of the immunoaffinity resin.
3. Under the conditions described above, most mouse mAbs will flow
through the DE52. However, a few mAbs will bind under these condi-
tions. Therefore, the high salt elution (Step 7) will elute the mAb. It will
Immunoaffinity Chromatography 487
be necessary to use a salt gradient to purify a mAb that binds to the DE52
column.
4. If the mAb is made in the CELLine flask, the DE52 step might not be
necessary.
2.6.1. Comments
1. One gram of this resin yields 3.5 ml of swollen resin. For most purposes,
we find making 0.5–2.0 g at a time to be convenient. We have found
488 Nancy E. Thompson et al.
Elution 3
Elution 4
Markers
Lysate
kDa
260
160 b ′/b
110
80 s 70
60
50
40 a
30
20
15
10
3.5
1 2 3 4 5 6
Figure 28.3 SDS–PAGE (4–12%) of RNAP purified from E. coli by a one-step immu-
noaffinity chromatography procedure, using mAbNT73-Sepharose. The cell lysate was
prepared as described in the text, and the soluble material (Lane 2) was applied to a NT73-
Sepharose column (2 ml). The flow-through material (Lane 3) was collected. After
washing with TE buffer containing 100 mM NaCl, the column was washed briefly with
TE containing 0.5 M NaCl (Lane 4). The column was re-equilibrated in TE containing
100 mM NaCl, and then the RNAP was eluted from the column with TE buffer contain-
ing 0.75 M NaCl and 40% propylene glycol, and 1-ml fractions collected (Lanes 5 and 6).
RNAP subunits are indicated on the right side. Most of the extra bands are RNAP-
binding proteins.
2.7.1. Comments
1. The lysate (Step 3) can be treated with Benzonase (EMD/Novagen
#70746, Madison, WI) to help digest nucleic acids and decrease viscosity
(see Chapter 18 in this volume).
490 Nancy E. Thompson et al.
2. The 500 mM NaCl wash (Step 6) helps to remove nucleic acids and
NusA, a RNA polymerase-binding protein.
3. For reasons that are unknown to us, the elution of the target protein
(Step 8) is significantly more effective if performed at room temperature
than at 4 C.
160 b /b ′
s 70
60
50
40 a
30 GFPep
20
15
10
3.5
1 2 3 4 5
Figure 28.4 SDS–PAGE (4–12%) of RNAP and epitope-tagged GFP purified on PR-
mAbs. Lane 1 contains Novex Sharp Standard markers; Lane 2 contains E. coli RNAP
purified on mAb NT73; Lane 3 contains E. coli core RNAP purified on mAb 8RB13;
Lane 4 contains B. subtilis core RNAP purified on mAb 8RB13; Lane 5 contains
epitope-tagged GFP purified on mAb NT73.
3. Conclusions
We have described the isolation, identification, and use of PR-mAbs
for use in gentle immunoaffinity chromatography. Although the examples
that we have presented here apply to our PR-mAbs and their use to purify
Immunoaffinity Chromatography 493
DISCLOSURE
N. Thompson and R. Burgess are required by the University of Wisconsin–Madison
Conflict of Interest Committee to disclose that they have financial interests in the company
NeoClone which markets many of the mAbs mentioned in this chapter.
REFERENCES
Anthony, J. R., Green, H. A., and Donohue, T. J. (2003). Purification of Rhodobacter
sphaeroides RNA polymerase and its sigma factors. Meth. Enzymol. 370, 54–65.
Bergendahl, V., Thompson, N. E., Foley, K. M., Olson, B. M., and Burgess, R. R. (2003).
A cross-reactive polyol-responsive monoclonal antibody useful for isolation of core
RNA polymerase from many bacterial species. Protein Expr. Purif. 31, 155–160.
Burgess, R. R., and Thompson, N. E. (2002). Advances in gentle immunoaffinity chroma-
tography. Curr. Opin. Biotechnol. 13, 304–308.
Burgess, R. R., Arthur, T. A., and Pietz, B. C. (2000). Mapping protein-protein interaction
domains using ordered fragment ladder far-Western analysis of hexahistidine-tagged
fusion proteins. Meth. Enzymol. 328, 141–157.
Burgess, R. R., Thompson, N. E., and Duellman, S. J. (2007). Immunoaffinity chromatog-
raphy using epitope tags to polyol-responsive monoclonal antibodies. US Patent No.
7,241,580.
Cramer, P., Bushnell, D. A., Fu, J., Gnatt, A. L., Maier-Davis, B., Thompson, N. E.,
Burgess, R. R., Edwards, A. M., David, P. R., and Kornberg, R. D. (2000). Architecture
of RNA polymerase II and implications for the transcription mechanism. Science 288,
640–649.
Duellman, S. J., Thompson, N. E., and Burgess, R. R. (2004). An epitope tag derived from
human transcription factor IIB that reacts with a polyol-responsive monoclonal antibody.
Protein Expr. Purif. 35, 147–155.
Edwards, A. M., Darst, S. A., Feaver, W. J., Thompson, N. E., Burgess, R. R., and
Kornberg, R. D. (1990). Purification and lipid layer crystallization of yeast RNA
polymerase II active in transcription initiation. Proc. Natl. Acad. Sci. USA 87, 2122–2126.
Harlow, E., and Lane, D. (1988). Antibodies: A Laboratory Manual. Cold Spring Harbor
Press, Cold Spring Harbor, NY11724.
Jiang, Y., Zhang, S. J., Wu, S. M., and Lee, M. Y. (1995). Immunoaffinity purification of
DNA polymerase delta. Arch. Biochem. Biophys. 230, 297–304.
Kuznedelov, K., Minakhin, L., Niedziela-Majka, A., Dove, S. L., Rogulja, D.,
Nickels, B. E., Hochschild, A., Heyduk, T., and Severinov, K. (2002). A role for the
interaction of the RNA polymerase flap domain with the sigma subunit in promoter
recognition. Science 295, 855–857.
494 Nancy E. Thompson et al.
Largaespada, D. A., Jackson, M. W., Thompson, N. E., Kaehler, D. A., Byrd, L. G., and
Mushinski, J. F. (1996). The ABL-MYC retrovirus generates antigen-specific plasmacy-
tomas by in vitro infection of activated B lymphocytes from spleen and other murine
lymphoid organs. J. Immunol. Methods 197, 85–95.
Lynch, N. A., Jiang, H., and Gibson, D. T. (1996). Rapid purification of the oxygenase
component of toluene dioxygenase from a polyol-responsive monoclonal antibody.
Appl. Environ. Microbiol. 62, 2133–2137.
Maldonado, E., Drapkin, R., and Reinberg, D. (1996). Purification of human RNA
polymerase II and general transcription factors. Meth. Enzymol. 274B, 72–100.
Nagy, P. L., Griesenbeck, J., Kornberg, R. D., and Cleary, A. (2002). A trithorax-group
complex purified from Saccharomyces cerevisiae is required for methylation of histone H3.
Proc. Natl. Acad. Sci. USA 99, 90–94.
Nelson, P. N., Reynolds, G. M., Waldron, E. E., Ward, E., Giannopoulos, K., and
Murray, P. G. (2000). Demystified: Monoclonal antibodies. J. Clin. Pathol. Mol. Pathol.
53, 111–117.
Probasco, M. D., Thompson, N. E., and Burgess, R. R. (2007). Immunoaffinity purification
and characterization of RNA polymerase from Shewanella oneidensis. Protein Expr. Purif.
55, 23–30.
Rao, L., Jones, D. P., Nguyen, L. H., McMahan, S. A., and Burgess, R. R. (1996). Epitope
mapping using histidine-tagged protein fragments: Application to Escherichia coli RNA
polymerase sigma70. Anal. Biochem. 241, 141–157.
Subramanian, A. (2002). Immunoaffinity chromatography. Mol. Biotechnol. 20, 41–47.
Thompson, N. E., and Burgess, R. R. (1994). Purification of recombinant human transcription
factor IIB (TFIIB) by immunoaffinity chromatography. Protein Expr. Purif. 5, 469–475.
Thompson, N. E., and Burgess, R. R. (1996). Immunoaffinity of RNA polymerase and
transcription factors using polyol-responsive monoclonal antibodies. Meth. Enzymol.
274B, 513–526.
Thompson, N. E., and Burgess, R. R. (1999). Immunoaffinity purification of the RAP30
subunit of the human transcription factor IIF (TFIIF). Protein Expr. Purif. 17, 260–266.
Thompson, N. E., and Burgess, R. R. (2001). Preparation and use of specialized antibodies.
Identification of polyol-responsive monoclonal antibodies for use in immunoaffinity
chromatography. Curr. Protoc. Mol. Biol. Sect. VI (Suppl. 54), 11.18.1–11.18.9.
Thompson, N. E., Aronson, D. B., and Burgess, R. R. (1990). Purification of eukaryotic
RNA polymerase II by immunoaffinity chromatography: Elution of active enzyme with
protein stabilizing agents from a polyol-responsive monoclonal antibody. J. Biol. Chem.
265, 7069–7077.
Thompson, N. E., Hager, D. A., and Burgess, R. R. (1992). Isolation and characterization of
a polyol-responsive monoclonal antibody useful for gentle purification of E. coli RNA
polymerase. Biochemistry 31, 7003–7008.
Thompson, N. E., Arthur, T. M., and Burgess, R. R. (2003). Development of an epitope tag
for the gentle purification of proteins by immunoaffinity chromatography: Application to
epitope-tagged green fluorescent protein. Anal. Biochem. 323, 171–179.
Thompson, N. E., Foley, K. M., and Burgess, R. R. (2004). Antigen-binding properties of
monoclonal antibodies reactive with human TATA-binding protein (TBP) and use in
immunoaffinity chromatography. Protein Expr. Purif. 36, 186–197.
Thompson, N. E., Jensen, D. B., Lamberski, J. A., and Burgess, R. R. (2006). Purification of
protein complexes by immunoaffinity chromatography: Application to transcription
machinery. Genet. Eng. 27, 81–100.
Tsien, R. Y. (1998). The green fluorescent protein. Annu. Rev. Biochem. 67, 509–544.
Woychik, N. A., and Hampsey, M. (2002). The RNA polymerase II machinery: Structure
illuminates function. Cell 108, 453–463.
C H A P T E R T W E N T Y- N I N E
Contents
1. Background 498
2. Polyacrylamide Gels 500
3. Principle of Method 501
4. Procedure 502
4.1. Stock solutions 502
4.2. Catalyst 503
4.3. Electrode buffer 503
4.4. Casting gels 503
4.5. Sample preparation 505
4.6. Electrophoresis 505
4.7. Comments on method 506
4.8. Variations of method 507
5. Detection of Proteins in Gels 508
5.1. Dye staining with Coomassie Brilliant Blue R-250 509
5.2. Silver staining 509
5.3. Copper staining 510
6. Marker Proteins 510
7. Molecular Weight Determination 511
8. Preparative Electrophoresis 511
References 513
Chemical Division, Research Products Group, Bio-Rad Laboratories, Incorporated, Richmond, California
1
Reprinted from Methods in Enzymology, Volume 182 (Academic Press, 1990)
497
498 David E. Garfin
1. Background
Although the detailed theory of gel electrophoresis is complicated and
at present incomplete (Bier et al., 1983; Chrambach and Jovin, 1983; Jovin,
1973), the fundamental concepts are easily understood. Briefly, in an
electrophoretic separation, charged particles are caused to migrate toward
the electrode of opposite sign under the influence of an externally applied
electric field. The movements of the particles are retarded by interactions
with the surrounding gel matrix, which acts as a molecular sieve. The
opposing interactions of the electrical force and molecular sieving result in
differential migration rates for the constituent proteins of a sample.
In general, fractionation by gel electrophoresis is based on the sizes,
shapes, and net charges of the macromolecules. Systems designed to frac-
tionate proteins in their native configurations cannot distinguish between
the effects of size, shape, and charge on electrophoretic mobility. As a
consequence, proteins with differing molecular weights can have the same
mobility in these systems. Thus, while PAGE methods for native proteins
are valuable for separating and categorizing protein mixtures, they should
not be used to assess the purity of a preparation or the molecular weight of
an unknown.
SDS–PAGE overcomes the limitations of native PAGE by imposing
uniform hydrodynamic and charge characteristics on all the proteins in a
One-Dimensional Gel Electrophoresis 499
2. Polyacrylamide Gels
Polyacrylamide gels are formed by copolymerization of acrylamide
monomer, CH2¼CH–CO–NH2, and a cross-linking comonomer, N,N 0 -
methylenebisacrylamide, CH2¼CH–CO–NH–CH2–NH–CO–CH¼CH2,
(bisacrylamide) (Allen et al., 1984; Andrews, 1986; Chrambach, 1985;
Chrambach and Rodbard, 1971; Hames, 1981). The mechanism of gel
formation is vinyl addition polymerization and is catalyzed by a free radical-
generating system composed of ammonium persulfate (the initiator) and an
accelerator, tetramethylethylenediamine (TEMED). TEMED causes the
formation of free radicals from persulfate and these in turn catalyze polymeri-
zation. Oxygen, a radical scavenger, interferes with polymerization, so that
proper degassing to remove dissolved oxygen from acrylamide solutions is
crucial for reproducible gel formation.
The sieving properties of a gel are established by the three-dimensional
network of fibers and pores which is formed as the bifunctional bisacryla-
mide cross-links adjacent polyacrylamide chains (Rodbard and Chrambach,
1970). Within limits, as the acrylamide concentration of a gel increases, its
effective pore size decreases. The effective pore size of a gel is operationally
defined by its sieving properties; that is, by the resistance it imparts to the
migration of protein molecules. By convention, a given gel is physically
characterized by the pair of figures (%T, %C), where %T is the weight
percentage of total monomer (acrylamide þ cross-linker, in g/100 ml), and
%C is the proportion of cross-linker (as a percentage of total monomer) in
the gel. The practical limits for %T lie between 3% and 30%. The factors
governing pore size are complicated, but, in general, the pore size of a gel
decreases as %T increases. For any given fixed %T, pore size is at a minimum
at about 5% C, increasing at both higher and lower cross-linker concentra-
tions (Allen et al., 1984; Andrews, 1986; Chrambach, 1985; Chrambach and
Rodbard, 1971; Hames, 1981).
The use of high-quality reagents is a prerequisite for reproducible, high-
resolution gels. This is particularly true of acrylamide, which constitutes the
most abundant component in the gel–monomer mixture. Residual acrylic
acid, linear polyacrylamide, and ionic impurities are the major contaminants
of acrylamide preparations. Moreover, buffer components should be of
reagent grade and only distilled or deionized water should be used for all
phases of gel electrophoresis.
In SDS–PAGE, the quality of the SDS is of prime importance.
Differential protein-binding properties of impurities such as C10, C14, and
C16 alkyl sulfates can cause single proteins to form multiple bands in gels
(Margulies and Tiffany, 1984). Even with pure SDS, very basic proteins,
very acidic proteins, various glycoproteins, and lipoproteins, because of
their unusual compositions, migrate ‘‘anomalously’’ during electrophoresis
(Allen et al., 1984; Andrews, 1986; Hames, 1981).
One-Dimensional Gel Electrophoresis 501
3. Principle of Method
The most popular electrophoretic method is the SDS–PAGE system
developed by Laemmli (Allen et al., 1984; Andrews, 1986; Blackshear, this
series; Hames, 1981; Laemmli, 1970). This is a discontinuous system con-
sisting of two contiguous, but distinct gels: a resolving or separating (lower)
gel and a stacking (upper) gel. The two gels are cast with different porosities,
pH, and ionic strength. In addition, different mobile ions are used in the gel
and electrode buffers. The buffer discontinuity acts to concentrate large
volume samples in the stacking gel, resulting in better resolution than is
possible using the same sample volumes in gels without stackers. Proteins,
once concentrated in the stacking gel, are separated in the resolving gel.
The Laemmli SDS–PAGE system is made up of four components. From
the top of the cell downward, these are the electrode buffer, the sample, the
stacking gel, and the resolving gel. Samples prepared in low-conductivity
buffer (0.06 M Tris–Cl, pH 6.8) are loaded between the higher conductivity
electrode (0.025 M Tris, 0.192 M glycine, pH 8.3) and stacking gel
(0.125 M Tris–Cl, pH 6.8) buffers. When power is applied, a voltage
drop develops across the sample solution which drives the proteins into
the stacking gel. Glycinate ions from the electrode buffer follow the proteins
into the stacking gel. A moving boundary region is rapidly formed with the
highly mobile chloride ions in the front and the relatively slow glycinate
ions in the rear (Allen et al., 1984; Andrews, 1986; Blackshear, this series;
Bury, 1981; Hames, 1981; Wyckoff et al., 1977). A localized high-voltage
gradient forms between the leading and trailing ion fronts, causing the SDS–
protein complexes to form into a thin zone (stack) and migrate between the
chloride and glycinate phases. Within broad limits, regardless of the height
of the applied sample, all SDS–proteins condense into a very narrow region
and enter the resolving gel as a well-defined, thin zone of high protein
density. (The stacking phenomenon is strikingly demonstrated with
prestained protein standards, which are mixtures of proteins derivatized
with reactive dyes.) The large-pore stacking gel (4% T ) does not retard
the migration of most proteins and serves mainly as an anticonvective
medium. At the interface of the stacking and resolving gels, the proteins
experience a sharp increase in retardation due to the restrictive pore size of
the resolving gel. (Proteins too large to enter the resolving gel will stop
at the interface.) Once in the resolving gel, proteins continue to be slowed
by the sieving of the matrix. The glycinate ions overtake the proteins,
which then move in a space of uniform pH (pH 9.5) formed by the Tris
and glycine. Molecular sieving causes the SDS–polypeptide complexes to
separate on the basis of their molecular weights.
502 David E. Garfin
4. Procedure
Equipment and reagents for SDS–PAGE can be obtained from a
variety of suppliers. Electrophoresis cells vary in design, but their operation
generally follows the steps outlined below. Since the many available cells
differ in size, formulations are presented in conveniently sized units for
simplicity. Required volumes can be prepared using multiples of these unit
sizes. Except where noted, reagents for SDS–PAGE can be prepared as
concentrated stock solutions.
4.2. Catalyst
1. 10% ammonium persulfate (APS): Dissolve 100 mg APS in 1 ml of water.
Make the APS solution fresh daily.
2. TEMED (N,N,N0 ,N0 -tetramethylethylenediamine): Use TEMED undi-
luted from the bottle. Store cool, dry, and protected from light.
4.6. Electrophoresis
Assemble the electrophoresis cell, fill the upper and lower reservoirs with
electrode buffer, and remove the comb from the stacking gel. Load the
prepared samples into the wells in the stacking gel by layering them under
electrode buffer using a microliter syringe or micropipet. The glycerol in
the samples provides the necessary density for them to sink to the bottoms of
the wells and the Bromphenol Blue tracking dye enables the samples to be
seen during loading. Finally, attach leads to the unit and connect them to a
power supply. The lower electrode is the anode and the upper one is the
cathode, in SDS–PAGE.
During an electrophoresis run, electrical energy is converted to heat
which can cause band distortion and diffusion. In general, electrophoresis
should be carried out at power settings at which the run proceeds as rapidly
506 David E. Garfin
as the chamber’s ability to draw off heat will allow. In other words, the run
should be as fast as possible without exceeding desired resolution and
distortion limits.
Many of the power supplies which are available allow control of any
electrical quantity and the choice is almost a matter of preference. Constant
current conditions, as a rule, result in shorter but hotter runs than does
constant voltage (Allen et al., 1984). In the early stages of a run, the resistance
of the gel increases as the chloride ions migrate out of it. Accordingly,
voltage will rise or current will fall, depending on whether constant current
or constant voltage operation is in use.
Small-format minicells, with their thin glass plates, are better able to
efficiently dissipate the heat generated by the initially high currents at the
beginnings of runs than are standard-sized cells. Thus, the recommendation
is that gels should be run under constant current conditions (16–24 mA/mm
of gel thickness) in conventional apparatus and at constant voltage
(20–30 V/cm of gel length) in minicells. The use of recirculated coolant,
where possible, allows higher voltages and currents to be used for shortened
run times. Electrophoresis should be started immediately after the samples
are loaded and is generally continued until the Bromphenol Blue tracking
dye has reached the bottom of the gel.
may not bind the detergent in a constant weight ratio. This makes molecular
weight determinations of these molecules by SDS–PAGE less straightfor-
ward than analyses of fully denatured polypeptides, since it is necessary that
both standards and unknown proteins be in similar configurations for valid
comparisons.
When both SDS and 2-mercaptoethanol are left out of the Laemmli
procedure, what remains is the classical Ornstein–Davis PAGE system
(Davis, 1964; Ornstein, 1964) for native proteins. This is a high-resolution
native PAGE method designed for separation of the full spectrum of serum
proteins. Because the system was meant to separate a wide variety of
proteins, resolution may not be optimal for some restricted ranges of protein
mobilities. Although there are a number of high-resolution native PAGE
systems available to meet differing requirements (Allen et al., 1984;
Andrews, 1986; Blackshear, this series; Chrambach, 1985; Hames, 1981),
the Ornstein–Davis method should perform adequately for the fractionation
of the majority of commonly encountered protein mixtures. Molecular
weights are more difficult to determine by native PAGE than by SDS–
PAGE, since a single native system cannot distinguish the effects of charge
and conformation on protein electrophoretic mobilities (Allen et al., 1984;
Andrews, 1986; Chrambach, 1985; Hames, 1981).
The procedure described here is readily modified for native PAGE.
Merely omit 2-mercaptoethanol from the sample buffer and replace the
10% SDS in the recipes for the gel, sample, and electrode buffers with
equivalent volumes of water. Follow the procedure as otherwise presented,
except for sample treatment. Samples should be diluted in nondenaturing
buffer (0.06 M Tris–Cl, pH 6.8, 10% glycerol, 0.025% Bromphenol Blue)
following the same guidelines as for denaturing gels, but they should not be
heated.
After electrophoresis, remove the gel assembly and separate the glass
plates. The gel will probably stick to one of the two plates. Remove the
spacers and cut off and discard the stacking gel. Place the glass plate holding
the gel into fixative or staining solution and float the gel off of the plate.
All of the steps in gel staining are done at room temperature with gentle
agitation (e.g., on an orbital shaker platform) in any convenient container,
such as a glass casserole or a photography tray. Always wear gloves when
staining gels, since fingerprints will stain. Permanent records of stained gels
can be obtained by photographing them or by drying them on filter paper
using commercially available drying apparatus.
3. Soak the gel for 3–10 min in 200 ml of fresh oxidizer solution
(0.0034 M potassium dichromate, 0.0032 N nitric acid).
4. Wash the gel three or four times for 5–10 min in 400 ml water, until the
yellow color has been washed out.
5. Soak the gel in 200 ml fresh silver reagent (0.012 M silver nitrate) for
15–30 min.
6. Wash the gel with 400 ml water for 1–2 min.
7. Wash the gel for about 1 min in developer (0.28 M sodium carbonate,
1.85% paraformaldehyde).
8. Replace the developer with fresh solution and incubate for 5 min.
9. Replace the developer a second time and allow development to
continue until satisfactory staining has been obtained.
10. Stop development with 5% acetic acid (v/v).
Vertical streaks and sample-independent bands in the 50–70-kDa region
are sometimes seen in silver-stained gels. These artifacts have been attributed
to reduction of contaminants inadvertently introduced into the samples
(Ochs, 1983). They can be eliminated by adding excess iodoacetamide to
sample solutions after treatment with SDS-reducing buffer (Görg et al., 1987).
6. Marker Proteins
Mixtures of marker proteins are available for calibrating gels. PAGE
standards are mixtures of proteins with precisely known molecular weights
blended for uniform staining. They are obtainable in various molecular
One-Dimensional Gel Electrophoresis 511
8. Preparative Electrophoresis
The most satisfactory way to recover proteins separated by SDS–
PAGE for further study is to extract them from bands excised from the
gels. Many attempts have been made to design continuous elution devices
suitable for routine protein purification, in which bands emerging from the
bottoms of electrophoresis gels are swept away to fraction collectors
512 David E. Garfin
cylindrical gel rods. Devices are available for the rapid recovery of proteins
in small volumes with yields of greater than 70% in most cases. Elution takes
about 3 h at 10 mA/tube in 0.025 M Tris, 0.192 M glycine, 0.1% SDS, pH 8.3
(standard SDS–PAGE electrode buffer). SDS can be removed from the
eluted samples by dialysis or ion-exchange chromatography (Furth, 1980).
REFERENCES
Allen, R. C., Saravis, C. A., and Maurer, H. R. (1984). Gel Electrophoresis and Isoelectric
Focusing of Proteins: Selected Techniques. de Gruyter, Berlin.
Andrews, A. T. (1986). Electrophoresis: Theory, Techniques, and Biochemical and Clinical Appli-
cations. 2nd edn. Oxford University Press, New York.
Bier, M., Palusinski, O. A., Mosher, R. A., and Saville, D. A. (1983). Science 219, 1281.
Blackshear, P. J. (this series). Vol. 104, p. 237.
Bury, A. F. (1981). J. Chromatogr. 213, 491.
Chrambach, A. (1985). The Practice of Quantitative Gel Electrophoresis. Weinheim, VCH.
Chrambach, A., and Jovin, T. M. (1983). Electrophoresis 4, 190.
Chrambach, A., and Nguyen, N. Y. (1979). In ‘‘Electrokinetic Separation Methods’’,
(P. G. Righetti, C. J. Van Oss, and J. W. Vanderhoff, eds.), p. 337. Elsevier, Amsterdam.
Chrambach, A., and Rodbard, D. (1971). Science 172, 440.
Davis, B. J. (1964). Ann. NY Acad. Sci. 121, 404.
Dunbar, B. S. (1987). Two-Dimensional Electrophoresis and Immunological Techniques. Plenum,
New York.
Dunbar, B. S., Kimura, H., and Timmons, T. M. (this volume). Chapter 34.
Furth, A. J. (1980). Anal. Biochem. 109, 207.
Görg, A., Postel, W., Weser, J., Günther, S., Strahler, J. R., Hanash, S. M., and Somerlot, L.
(1987). Electrophoresis 8, 122.
Hames, B. D. (1981). In ‘‘Gel Electrophoresis of Proteins: A Practical Approach’’,
(B. D. Hames and D. Rickwood, eds.), p. 1. IRL Press, Oxford.
Harrington, M. (this volume). Chapter 37.
Jovin, T. M. (1973). Biochemistry 12, 871, 879, 890.
Laemmli, U. K. (1970). Nature (London) 227, 680.
Lee, C., Levin, A., and Branton, D. (1987). Anal. Biochem. 166, 308.
Margulies, M. M., and Tiffany, H. L. (1984). Anal. Biochem. 136, 309.
Marshall, T. (1984). Clin. Chem. 30, 475.
Merril, C. R. (this volume). Chapter 36.
Merril, C. R., Goldman, D., Sedman, S. A., and Ebert, M. H. (1981). Science 211, 1437.
Merril, C. R., Goldman, D., and Van Keuren, M. L. (this series). Vol. 104, p. 441.
Neville, D. M. Jr. (1971). J. Biol. Chem. 246, 6328.
Neville, D. M., and Glossmann, H. (this series). Vol. 32, p. 92.
Nielsen, T. B., and Reynolds, J. A. (this series). Vol. 48, p. 3.
Ochs, D. (1983). Anal. Biochem. 135, 470.
Ornstein, L. (1964). Ann. NY Acad. Sci. 121, 321.
Rodbard, D., and Chrambach, A. (1970). Proc. Natl. Acad. Sci. USA 65, 970.
Shapiro, A. L., Vinuela, E., and Maizel, J. V. Jr. (1967). Biochem. Biophys. Res. Commun. 28,
815.
Timmons, T. M., and Dunbar, B. S. (this volume). Chapter 51.
Weber, K., and Osborn, M. (1969). J. Biol. Chem. 244, 4406.
Wilson, C. M. (this series). Vol. 91, p. 236.
Wyckoff, M., Rodbard, D., and Chrambach, A. (1977). Anal. Biochem. 78, 459.
C H A P T E R T H I R T Y
Contents
1. Introduction 516
1.1. Isoelectric focusing—Basic principle 517
1.2. Types of pH gradients for isoelectric focusing 518
1.3. SDS polyacrylamide electrophoresis 521
1.4. Enhancement of resolution for alkaline pH gradients 523
1.5. Difference gel electrophoresis 524
2. Materials 527
2.1. Equipment 527
2.2. Solutions and reagents 527
3. Methods 528
3.1. Protein sample preparation 528
3.2. Sample cleanup and precipitation 530
3.3. Isoelectric focusing using acidic range IPG gels 531
3.4. Isoelectric focusing using alkaline range IPG gels 534
3.5. Equilibration of IPG gels 535
3.6. SDS–PAGE: The second dimension 536
References 538
Abstract
By far the highest resolution of all separation techniques for intact proteins in a
single analytical run continues to be by the combination of isoelectric focusing
(IEF) and SDS–PAGE, originally introduced by O’Farrell [(1975). J. Biol. Chem. 250,
4007–4021]. This analytical platform has seen a number of significant advances
and applications over the past 25 years, including reproducibility using immobi-
lized pH gradient (IPG) strips [Bjellqvist et al. (1982). J. Biochem. Biophys.
Methods 6, 317–339.], resolution in alkaline IEF using hydroxyethyldisulfide
(HED) [Olsson et al. (2002). Proteomics 2, 1630–1632], and quantification
for differential expression proteomics on intact proteins on a global scale
515
516 David B. Friedman et al.
[DIGE; Unlu et al. (1997). Electrophoresis 18, 2071–2077]. These major improve-
ments will be highlighted in this chapter alongside the principle and theory of 2D
gel electrophoresis, as well as detailed methods for general 2D gel electropho-
resis best use protocols.
1. Introduction
For complex mixtures such as whole cell lysates or enriched subcellu-
lar fractions, two-dimensional gel electrophoresis (2D gel) can typically
resolve hundreds-to-thousands of individual protein species using two
orthogonal separations1 (Fig. 30.1). The first separation is typically based
on charge using denaturing IEF, and the second separation by apparent
molecular mass using denaturing sodium dodecyl sulfate–polyacrylamide
gel electrophoresis (SDS–PAGE).2 2D gel experiments are particularly
powerful for visualizing protein isoforms that result from charged posttrans-
lational modification, such as phosphorylation and sulfation (which add
charge), or acetylation (which neutralize charge), among others. They are
also useful in detecting splice variants and proteolytic cleavages that result in
protein species with altered molecular weight (MW) and isoelectric point
(pI). More recently, 2D gels have been used extensively in differential
expression proteomics experiments, especially since the commercialization
of the difference gel electrophoresis (DIGE) technology in early 2000
(Friedman and Lilley, 2009; Lilley and Friedman, 2004; Unlu et al., 1997).
In modern 2D gel-based proteomics experiments, intact protein resolu-
tion by 2D gel electrophoresis is commonly coupled with protein (‘‘spot’’)
excision, in-gel digestion and mass spectrometry to provide for protein
identification using sophisticated database searching algorithms. As with
all high-end proteomics technologies, this method also requires many
1
Despite this resolution, proteins of extreme MW and pI (10 > kDa < 200; 3 > pI < 11), very hydropho-
bic proteins (e.g., integral membrane proteins with multiple trans-membrane domains), and low-abundance
proteins are typically difficult to resolve or detect on 2D gels. It is in these cases where complementary data
can be obtained from an orthogonal technology, such as liquid chromatography coupled with tandem mass
spectrometry (LC–MS/MS), that is capable of detecting many of these proteins. The added sensitivity of the
LC–MS/MS approach arises from the fact that proteins are detected based on surrogate peptides from an en
masse proteolytic digest that are resolved by liquid chromatography. But, for this very reason, the LC–MS/
MS technique is greatly challenged to identify the multiple protein isoforms that arise from charged or bulky
post-translational modifications that are readily detectable and quantifiable on a 2D gel (especially in a
nontargeted, discovery experiment). Furthermore, the quantification of intact protein expression using the
2D gel platform can now be done using the fluorescent multiplexing technology of DIGE, whereby the
requisite number of individual biological replicates from multiple experimental conditions can be easily
analyzed to provide for statistically powered discovery proteomics.
2
Although denaturing IEF is most commonly used for the highest resolution first-dimensional separations (first
described by O’Farrell, 1975), other techniques such as blue native PAGE (Schagger and von Jagow, 1991)
for analysis of protein complexes, and acidic PAGE in presence of a cationic detergent (Hartinger et al., 1996)
for separating hydrophobic membrane proteins, are often utilized for more specific investigations. However,
SDS–PAGE is almost always being used for the second-dimensional separation.
Isoelectric Focusing and Two-Dimensional Gel Electrophoresis 517
Figure 30.1 High resolution 2D gel (24 cm pH 4–7). 400 mg cellular proteins from a HeLa
cell extract were separated in two orthogonal dimensions, the first based on charge by
isoelectric focusing, and the second based on apparent molecular mass by sodium dodecyl
sulfate polyacrylamide gel electrophoresis. The gel was stained with the fluorescent dye
Sypro Ruby. Typically hundreds-to-thousands of individual protein species, including post-
translationally modified isoforms and proteolytic products, can be resolved and quantified.
The first reagents on the market for this purpose, AmpholinesÒ , were
mixtures of aliphatic oligo-amino/oligo-carbonic acids, of 600–700 differ-
ent homologues exhibiting a spectrum of isoelectric points between pH 3
and 10. These compounds have high buffering capacities at their isoelectric
points and form a pH gradient under the influence of the electric field.
Optimally, they have molecular weights below 1 kDa, and do not bind to
proteins, because they are highly hydrophilic. Mixtures with narrow inter-
vals are available for higher resolution using defined isoelectric point ranges.
More recently came the development of carrier ampholytes, which are
synthesized from different reagents than the AmpholinesÒ and are available
from a variety of commercial vendors. Although the function of forming a
pH gradient under the influence of the electric field is the same for all of
these products, the profiles of the pH gradients, the distribution of buffering
power, and the number of homologues are different (Righetti et al., 2007),
which leads to differences in the focusing pattern.
Carrier ampholytes-based IEF is mostly performed in polyacrylamide
gels, and the original 2D gel electrophoresis procedure by O’Farrell (1975)
accomplishes IEF in carrier ampholyte gradients in thin gel rods (‘‘tube gels’’
for the first dimension). However, these long and soft gels are not easy to
handle, and the pH gradients become unstable with increased running time,
resulting in cathodal drift. Although some laboratories still use the carrier
ampholyte technique, and this method has mostly been supplanted by the
introduction of IPG strips, which are described next.
nonreacted compounds are washed out from the matrix with distilled
water to produce the very low electrical conductivity necessary for IEF
(Westermeier, 2005).
The IPG gels are then dried for long-term storage, and can be rehydrated as
necessary shortly before use. The major benefit of IPGs is the absence of the
cathodal drift: the gradient is fixed to the matrix. The major application of
IPGs is IEF under denaturing conditions, serving as the first dimension of
high-resolution two-dimensional electrophoresis. Carrier ampholytes are still
used, and for improved conductivity and protein solubility they are added to
the sample as well as to the rehydration solution of the gels with IPGs.
Precast IPG strips are now commercially available from a number of
vendors with a wide variety of strip lengths and pH gradients to enable the
optimal separation and display of selected protein groups within a protein
lysate. In particular, pH gradients are available from a very wide range (e.g., pH
3–11), to medium ranges (e.g., pH 4–7, pH 7–11), to narrow ranges (e.g., pH
4–5, pH 5.5–6.5). Strip lengths also vary from as short as 7 cm to as long as
24 cm. Strips typically are 3 mm in width, with an average thickness (after
reswelling) of approximately 0.5 mm (Fig. 30.2). Care should be taken in
selection of the optimal pH range and strip length for the desired experimental
goals (Hoving et al., 2000). For example, 7-cm pH 3–11 IPG strips may seem
best to provide the greatest range in a small gel format, but they provide overall
the lowest resolution and sensitivity for proteome-wide, discovery-based
experiments.
Figure 30.2 Immobilized pH gradient (IPG) strip, shown in the manual process of
preparing for rehydration loading. Sample is first dispensed into a reswelling try, and then
the IPG strip is carefully overlayed on the protein sample solution (face-down, with
plastic backing up) using a forceps to hold one end of the IPG (the plastic backing exceeds
the length of the IPG gel). The strips are then overlaid with 2–3 mL paraffin oil to
prevent urea crystallization during rehydration. IPG strips are provided from commercial
vendors with barcodes to aide sample-tracking.
Isoelectric Focusing and Two-Dimensional Gel Electrophoresis 521
B
Filter paper pad
soaked with water + IPG strip pre-rehydrated −
with sample/rehydration solution
C
Filter paper pad Sample
soaked with water + −
Pre-rehydrated IPG strip
D
Paper bridge
soaked with sample
+ −
Pre-rehydrated IPG strip
Figure 30.3 Different configurations for sample loading into immobilized pH gradient
(IPG) strips for isoelectric focusing. (A) The IPG gel is placed face-down for sample loading
via active rehydration loading, after which isoelectric focusing is also possible. (B) The
sample is rehydrated into the IPG gel, which is then placed “sunny-side up” for isoelectric
focusing. (C) The sample is introduced into a previously rehydrated IPG gel via cup loading
at the anode. This configuration is especially beneficial for isoelectric focusing over alkaline
pH gradients. (D) Modification of cup loading, where the sample is introduced at the anode
via a paper-bridge. Reproduced with permission, Westermeier, Naven and Heopker
(2008), Proteomics in Practice, 2nd edition, Wiley-VCH Verlag GmbH & Co. KGaA.
hydrophobic proteins, all proteins, also very basic ones, migrate into the
same direction, and a separation according to the apparent molecular mass
(commonly referred to as molecular weight) is obtained.
between 1.5 mm and 0.5 mm. Thicker gels have a better mechanical
stability, but are more difficult to stain and provide additional sample loss
and background noise for subsequent mass spectrometry. Thinner gels can
be cooled more efficiently, thus they can be run faster to obtain sharper
resolution.
Of the improvements made to the 2D gel process over the past few
decades, the second-dimension SDS–PAGE separations have remained
largely unchanged. Vertical electrophoresis systems are more commonly
used, and equipment to run individual or 12 or more gels simultaneously is
available. Horizontal flatbed configurations have also been available histori-
cally and are now beginning to see a resurgence due to the low consump-
tion of reagents (e.g., SDS running buffer), sample manipulation, as well as
the potential for increased resolution afforded by short run times. Some
improvements include using flexible plastic backing (including low-fluores-
cent media for DIGE and other fluorescence-based techniques), specialized
equipment for multicasting reproducible acrylamide gradients, and different
acrylamide formulations for increased shelf life and gel durability.
O CH2 S CH2 H
Protein-S− + H CH2 S CH2 O
S CH2 H O CH2
Protein-S CH2 O + H CH2 S−
Figure 30.4 Oxidation of protein thiol groups using hydroxyethyl disulfide (HED)
according to Olsson et al. (2002). The resulting mixed disulfides are stable and prevent
the reformation of disulfide bond during isoelectric focusing over alkaline pH ranges,
whereas the standard reductant DTT becomes charged and migrates away from this
region of the gradient, resulting in streaking and artifacts.
then have to migrate toward the cathode to reach their isoelectric points
(rather having them migrating in both directions).
When the proteins enter the IPG strip, the sulfhydryl groups on cysteine
side chains become immediately oxidized to highly stable mixed disulfides.
This is an equilibrium reaction with high specificity, preventing any
unwanted side reactions (Fig. 30.4). Small amounts of DTT can be tolerated
for the sample preparation when using DeStreak, so that proteins can stay
reduced prior to focusing. However, care must be taken because excessive
DTT will reduce HED to produce 2-mercaptoethanol, which results in
even more horizontal streaking.3
3
When using DeStreak for IEF in alkaline pH gradients (pH 7–11, 3–11), 100 mL sample may contain up to
20 mM reducing agent. DeStreak can also be used for other pH ranges, with different tolerances for
reductants (adapted from the DeStreak rehydration solution manual, GE Healthcare).
Isoelectric Focusing and Two-Dimensional Gel Electrophoresis 525
2. Materials
2.1. Equipment
1. Electrophoresis system for IEF. Several formats are available from a
number of commercial vendors. Most can focus up to 12 IPG strips
simultaneously.
2. Programmable Power Supply capable of producing at least 3500 V but
preferably 8,000–10,000 V.
3. Multigel system for second-dimension SDS–PAGE. Several formats are
available for running 2, 4, 6, 12 or more gels in a single run.
4. Thermostatic circulator (necessary for most second-dimension systems;
check specific vendors).
5. Immobiline DryStrip Reswelling Tray (GE Healthcare).
6. Sample-loading cups (required for IEF in gradients that include
pH > 8).
7. Rotary shaker.
8. Image capture devices (e.g. densitometry, flatbed scanners, fluorescent
imager (must be compatible with Cy dyes for best results with DIGE)).
3. Methods
3.1. Protein sample preparation
Robust sample preparation is vital for any successful analytical measure-
ment. Buffers and materials used should be of the highest quality, and care
should be taken during sample procurement to increase reproducibility and
to minimize unanticipated variation that might be measured in the analysis.
Small molecule protease and phosphatase inhibitors such as aprotinin,
leupeptin, pepstatinA, antipain, AEBSF, sodium orthovanadate, okadaic
acid, and microcystin, among others (or commercial kits that utilize these
4
Mix equal parts butanol and water and shake vigorously. Let the two phases separate overnight, and use the
butanol phase for overlay. Butanol that is not completely water saturated can extract water from the top of the
gel. A more recent modification is to use a 0.1% SDS solution in a conventional spray bottle, used to carefully
spray a fine mist over the top of the gels to thoroughly cover the top of the gel (the gel/overlay interface will
not be as obvious).
Isoelectric Focusing and Two-Dimensional Gel Electrophoresis 529
5
An excellent practical 2D electrophoresis handbook with many tips and tricks with respect to basic
methodology can be downloaded for free at the following site: http://www5.gelifesciences.com/aptrix/
upp01077.nsf/Content/2d_electrophoresis2delectrophoresis_handbook
530 David B. Friedman et al.
6
For protein extraction from cell pellets using 2D gel sample buffer, it is useful for the solution to contain 2%
(v/v) Pharmalytes 3–10 to prevent aggregation of DNA. Cells (e.g., 108 cells) are best disrupted by rapid
addition of 1 mL sample buffer (Rabilloud, 1998) to the cell pellet. Nucleases (DNase, RNase, and
Benzonase) are commonly added as well, but their efficacy is questionable in this highly denaturing buffer.
Isoelectric Focusing and Two-Dimensional Gel Electrophoresis 531
only removes several contaminants at the same time, but also disrupts
complexes efficiently and prevents protease activities irreversibly. The
most effective methods for precipitation are using methanol and chloroform
according to Wessel and Flugge (1984) or TCA, deoxycholate, and acetone
according to Arnold and Ulbrich-Hofmann (1999). Before IEF, the pro-
teins must be redissolved in solubilization solution (see Section 3.1). In
addition, several protein precipitation kits for 2D gel analysis are available
from a number of commercial vendors. An adaption of the method of
Wessel and Flugge (1984) is described as follows:
1. Bring up predetermined amount of protein extract to 100 mL with water.
2. Add 300 mL (3-volumes) water.
3. Add 400 mL (4-volumes) methanol.
4. Add 100 mL (1-volume) chloroform.
5. Vortex vigorously and centrifuge; the protein precipitate should appear
at the interface.
6. Remove the water/MeOH mix on top of the interface, being careful not
to disturb the interface. Often the precipitated proteins do not make a
visibly white interface, and care should be taken not to disturb the interface.
7. Add another 400 mL methanol to wash the precipitate.
8. Vortex vigorously and centrifuge; the protein precipitate should now
pellet to the bottom of the tube.
9. Remove the supernatant and briefly dry the pellets in a vacuum centrifuge.
10. Resuspend the pellets in a suitable amount of 2D gel-compatible buffer
(see Section 3.1).
When using precipitation as a cleanup step, it is recommended to have a
starting protein concentration in the range of 1–10 mg/mL. When protein
samples are too diluted, it will be difficult to quantitatively recover proteins
following precipitation cleanup. Freeze/thawing should also be kept to a
minimum; freezing samples in 1 mL aliquots or less will usually suffice. For
DIGE experiments, this precipitation step greatly facilitates the requirement
for sample labeling to be performed in 2D electrophoresis sample buffer that
is devoid of free amines and pH balanced.
gradient and strip length, using samples derived from the same experiment
(e.g., sample types and composition should be a similar as possible).7
The following method is presented for 24 cm IPG strips pH 4–7:
1. IPG strips are rehydrated in the 2D gel sample buffer (see Section 3.1)
with total protein amounts as high as 0.5–2 mg protein, diluted or
resuspended in a final volume of 450 mL solution per 24 cm strip in a
reswelling tray. The strips are overlaid with 2–3 mL paraffin oil to
prevent urea crystallization.8
2. Rehydration of the strips in the presence of sample should take at least
12 h, but overnight rehydration is recommended, at room temperature.
Some proteins, especially high-molecular-weight proteins, need more
time to enter into the strip during the rehydration step.
3. After complete rehydration, the strips are moved to a flatbed electro-
phoresis system of which the temperature of the cooling block is held
constant at 20 C.
4. Humid paper wicks are applied on the ends of the rehydrated IPG strips,
and electrodes are positioned in place and the strips are covered with
paraffin oil.9
5. See Table 30.1 for typical focusing program.10
Rehydration loading is the easiest method for sample introduction for
2D gel electrophoresis, where protein samples in sample buffer are intro-
duced passively while the IPG strip takes up the sample solution, and the
7
When using the DIGE technique where different combination of samples are multiplexed and coresolved
using a series of gels, the potential of the different strips containing different sample amounts is minimized by
properly randomizing the loading of the samples such that each IPG strip contains the same variety of
samples. For example, in a 3-dye experiment of control versus treated samples using four biological replicates
(using Cy3 and Cy5 to randomly label each of the eight individual samples and Cy2 for the mixed-sample
internal standard), each of the four gels would contain one control and one treated sample along with an
aliquot of the Cy2-labeled mixed-sample internal standard.
8
It is very important to make sure that the rehydration solution is evenly distributed across the entire strip
length. The sample is first distributed along the length of a rehydration well and then the dehydrated IPG
strip is carefully overlayed (with the plastic backing facing up) to ensure that the sample is evenly distributed
beneath the overlayed strip. The IPG strips are covered with paraffin oil to avoid crystallization of the highly
concentrated urea (ca. 10 M urea is saturation at room temperature).
9
Placing humid paper wicks (filter paper wetted with water, but not oversaturated) between the electrodes
and both ends of the IPG strips will provide a sink for excess ions as well as proteins with isoelectric points
outside the pH range of the IPG strip. In extreme cases, it is possible to exchange the paper wicks during the
IEF run. The paraffin oil prevents drying out of the strip and crystallization of the urea/thiourea present, and
also serves to prevent carbon dioxide to enter the system. Carbon dioxide from the air can dissolve into the
IPG gel, acting as a buffer with pK 6.3 and thus changing the pH gradient.
10
IEF is mostly performed in a horizontal flatbed type of equipment. The focusing effect requires high field
strength; therefore, a power supply with high voltage is required. IEF must be performed at an exactly
controlled temperature, because the pK values of the proteins as well as the buffers are highly temperature
dependent. Thus a reliable cooling system is necessary for this method to be reproducible. It is important to
start with a low voltage to avoid aggregation of proteins. It is very useful to apply a voltage/time gradient on
the strips in order to minimize differences in conductivities coming from varying salt loads between the IPG
strips. The best results are obtained with high-end voltages of 8,000–10,000 V, with the overall amount of
total volt-hours being ultimately important for high-resolution focusing. Optimal gradients and total Vh
vary with strip length, pH range, and commercial formulations.
Isoelectric Focusing and Two-Dimensional Gel Electrophoresis 533
contact with the focusing tray (the gel is ‘‘sunny-side up’’) can be advanta-
geous for more difficult samples (Fig. 30.3).
6. Optional: active rehydration (available with some instrument configura-
tions). Sample rehydration occurs with the IPG strip face down and in
contact with the electrodes, which supply a low (typically 30–50 V)
electric field across the strip during rehydration. In some cases, active
rehydration can be found to facilitate the entry of high-molecular-
weight proteins into the IPG strip. IEF units with this configuration
typically commence with the focusing with the IPG strips in the face-
down orientation and do not require user intervention at this point.
11
The protein concentration in the sample cup should not be too high (<10 mg/mL) to avoid protein
precipitation at the entry point. If possible, the sample should be diluted with the sample buffer.
Isoelectric Focusing and Two-Dimensional Gel Electrophoresis 535
carefully positioned just below the anode (on the acidic side of the
gradient) such that the bottom of the cup is in contact with the top of
the IPG gel and makes a seal, but not crushing the gel. The strips are then
covered with paraffin oil.9
5. Ensure that oil does not leak into the cups before proceeding. Then add
approximately 10–20 mL paraffin oil to the cup, and carefully pipet the
sample under the oil (check manufacturer for maximum sample volume
per cup).12
6. See Table 30.1 for typical focusing program.10
12
Paper-bridge loading is a modification of the cup-loading technique that uses a piece of clean filter cardboard
containing the protein sample solution and applied at one of the ends of the IPG strip. Both this as well as the
conventional cup-loading technique needs more handling steps and can lead to losses of proteins with their
isoelectric points close to the pH value of the application point. However, with these techniques all proteins
have the same charge sign and will not form complexes during electrophoretic migration into the IPG strip,
and so in some cases it can deliver better results especially for IEF at pH ranges greater than pH 8.
13
Although suggested in several scientific papers, it is generally not recommended to prealkylate proteins prior
to IEF, because it is often not easy to guarantee 100% alkylation.
536 David B. Friedman et al.
14
Casting a large number of gels requires some practice. The amount of catalyst that is used is minimal to
prevent the gels from polymerizing too rapidly, which leads to excessive heating of the casting chamber.
Ideally, initial polymerization, which can be seen by the development of a distinct gel surface below the
butanol layer, should take 30–60 min. Subsequently, gels can be washed, covered with gel buffer, and stored
at room temperature. The amounts of catalyst can be increased by 10% if necessary. ‘‘Fast’’ polymeration
should occur within an hour or two, after which the isopropanol can be replaced with water or SDS running
buffer. However, gels should ideally be left to polymerize overnight to enable complete polymerization, and
the top of the gels should be lightly sealed with plastic wrap to prevent evaporation of the overlay (especially
if the butanal is left on overnight).
538 David B. Friedman et al.
pressure is applied to the plastic backing, not the IPG gel). Once the
SDS running buffer is overlayed within the running chamber, good
contact and removal of air bubbles can be easily achieved using the
same tapping procedure. Once placed, the IPG strip should not move
during the second-dimension run.
10. Optional: to ensure good contact between the strip and the gel, an
agarose solution can be used to keep the IPG strip in place. The agarose
solution is kept at 65 C and added first on top of the gel. Immediately
after, the equilibrated strip is placed on the gel. This optional procedure
is favored by some, but is cumbersome and can be difficult if the
solution solidifies before the strip is in place and any air bubbles
removed.15
11. Optional: a molecular weight marker can be run at the side (loaded
through a small paper wick).
12. The current is set to 5–10 mA/gel initially to let the proteins enter the
gel and subsequently set to 20 mA/gel over night at 15 C until the blue
front reach the end of the gel. If available, using constant wattage,
setting less than 1 watt per gel for at least the first hour followed by no
more than 15 watt per gel until completion is preferred because the
voltage and amperage will vary during the run as salts and ions leave
the gel.
After completion, gels can be fixed and stained using a variety of visual
and fluorescent staining techniques. DIGE gels can be imaged on a fluores-
cent imager immediately after the second dimension is completed. Gels can
then be stored in the refrigerator or in a cold room at 4 C for months16. It is
no problem to identify proteins using mass spectrometry from 2D gels that
were stored over a long period of time.
REFERENCES
Anderson, N. G., and Anderson, N. L. (1978a). Analytical techniques for cell fractions. XXI.
Two-dimensional analysis of serum and tissue proteins: Multiple isoelectric focusing.
Anal. Biochem. 85, 331–340.
Anderson, N. L., and Anderson, N. G. (1978b). Analytical techniques for cell fractions.
XXII. Two-dimensional analysis of serum and tissue proteins: Multiple gradient-slab gel
electrophoresis. Anal. Biochem. 85, 341–354.
15
The agarose overlay is prepared in such a way that it remains fluid at relative low temperatures. To realize
this a mixture of low-melting agaroses is used (e.g., 0.4% (w/v) standard low Mr from Bio-Rad Laboratories,
and 0.1% (w/v) type VII-A-low gelling temperature from Sigma), dissolved in 1 SDS running buffer.
A trace of bromophenol blue can be added as tracking dye.
16
For long-term storage, 0.02% (w/v) sodium azide can be added (optional) to prevent bacterial and fungal
growth. For DIGE gels and other applications where the gel is affixed to one of the glass plates, the gels can
be stored with the glass cassette intact to minimize dehydration during storage.
Isoelectric Focusing and Two-Dimensional Gel Electrophoresis 539
Sanchez, J. C., Rouge, V., Pisteur, M., Ravier, F., Tonella, L., Moosmayer, M.,
Wilkins, M. R., and Hochstrasser, D. F. (1997). Improved and simplified in-gel sample
application using reswelling of dry immobilized pH gradients. Electrophoresis 18, 324–327.
Schagger, H., and von Jagow, G. (1991). Blue native electrophoresis for isolation of
membrane protein complexes in enzymatically active form. Anal. Biochem. 199,
223–231.
Suehara, Y., Kondo, T., Fujii, K., Hasegawa, T., Kawai, A., Seki, K., Beppu, Y.,
Nishimura, T., Kurosawa, H., and Hirohashi, S. (2006). Proteomic signatures corres-
ponding to histological classification and grading of soft-tissue sarcomas. Proteomics 6,
4402–4409.
Unlu, M., Morgan, M. E., and Minden, J. S. (1997). Difference gel electrophoresis: A single
gel method for detecting changes in protein extracts. Electrophoresis 18, 2071–2077.
Vesterberg, O. (1972). Isoelectric focusing of proteins in polyacrylamide gels. Biochim.
Biophys. Acta 257, 11–19.
Vesterberg, O., and Svensson, H. (1966). Isoelectric fractionation, analysis, and characteri-
zation of ampholytes in natural pH gradients. IV. Further studies on the resolving power
in connection with separation of myoglobins. Acta. Chem. Scand. 20, 820–834.
Wessel, D., and Flugge, U. I. (1984). A method for the quantitative recovery of protein in
dilute solution in the presence of detergents and lipids. Anal. Biochem. 138, 141–143.
Westermeier, R. (2005). Electrophoresis in Practice (4th edn). Method 10 - IEF in immo-
bilized pH gradients, pp. 257–273. Wiley-VCH.
C H A P T E R T H I R T Y- O N E
Contents
1. Introduction 542
2. General Considerations 543
3. Instrumentation: Detection and Documentation 545
4. Total Protein Detection 545
4.1. Colorimetric total protein stains 545
4.2. Fluorescent total protein stains 549
4.3. Preelectrophoresis sample labeling 556
5. Phosphoprotein Detection 556
5.1. Pro-Q Diamond phosphoprotein gel stain 557
5.2. Phos-tag phosphoprotein stains 557
6. Glycoprotein Detection 557
6.1. General glycoprotein detection 557
6.2. O-GlcNAc detection 559
References 559
Abstract
Laboratory scientists who encounter protein biochemistry in many of its myriad
forms must often ask: is my protein pure? The most frequent response: run a
denaturing SDS polyacrylamide gel. Running this gel raises another series of
considerations regarding detection, quantitation, and characterization and so
the next questions invariably center on suitable protein gel staining and detec-
tion methods. A total protein profile can be determined with the colorimetric
methods embodied in Coomassie Blue and silver staining methods, or increas-
ingly, with fluorescent stains. Protein quantitation can be done following stain-
ing, with fluorescence- and instrumentation-based methods offering the
greatest sensitivity and linear dynamic range. Protein posttranslational mod-
ifications such as phosphorylation and glycosylation can be reliably determined
with several fluorescence-based protocols. Staining and detection with two or
more different stains can be done in series to establish relative profiles of
modified versus total protein or to assess purity at two levels of quantitative
McArdle Laboratory for Cancer Research, University of Wisconsin–Madison, Madison, Wisconsin, USA
541
542 Thomas H. Steinberg
sensitivity. The choice of staining method and protocol depends on the required
rigor of detection and quantitation combined with available instrumentation
and documentation capabilities. Other considerations for staining methods
include intended downstream analytical procedures such as mass spectrometry
or peptide sequencing, which preclude some methods. Nonfixative staining
methods allow western blotting after gel staining. Laboratory custom and
budget or intellectual curiosity may be the ultimate determinate of the chosen
gel staining protocol.
1. Introduction
This chapter focuses on electrophoresis-based protein detection with
emphasis on current methods for detection of total protein and detection of
the most frequently encountered protein posttranslational modifications
(PTMs), phosphorylation, and glycosylation. In the 20 years since the
publication of first edition of this volume, denaturing sodium dodecyl
sulfate polyacrylamide gel electrophoresis (SDS–PAGE) followed by pro-
tein staining has remained a fundamental laboratory procedure to determine
the composition and characterization of protein-containing mixtures at all
stages of a purification scheme. It is standard practice to display protein
profiles from a sampling of all purification steps—cell lysate to the final
product—to evaluate polypeptide composition, purity, and quantitation.
Thus a sample from each step should reveal a progressively simpler
polypeptide profile showing a greater proportion of the target protein.
The previously summarized methods and principles of protein electro-
phoresis and detection remain relevant (Dunbar et al., 1990; Garfin, 1990a,b;
Merril, 1990).
The widespread use of SDS–PAGE protein analytical methods has
driven the commercial development and common use of standardized
electrophoresis apparatus and reagents, precast multiwell 1D and 2D slab
gels (minigels; ca. 8 8 cm, 1.0 mm thick), and ready-to-use staining
formulations or kits for simplified, rapid staining. Additionally, proteomics-
oriented electrophoresis techniques have emerged, directed toward analysis
of complex protein mixtures by large format 2-D gel electrophoresis
followed by mass spectrometry (MS) of excised protein spots. The utility
of gel-based methodology in proteome analysis has spurred further com-
mercial development of extant colorimetric stains for total protein and
development of new fluorescent stains and modalities for total protein and
for protein PTMs, driving the coevolution of instrumentation for sensitive
detection, easy documentation, and accurate quantitation of both fluores-
cent and colorimetric signals. A recent compendium of proteomics proto-
cols describes methods and applications for protein detection in 2-D and
Protein Gel Stains 543
2. General Considerations
Common features of all gel staining protocols are grounded in good
laboratory practices: cleanliness, careful manipulations, and attention to
detail. Postelectrophoresis gel staining is generally accomplished in covered
polycarbonate or polypropylene dishes. Accumulation of residual stain may
compromise results, particularly with fluorescent stains. Dishes should be
cleaned with 70–100% ethanol or methanol followed by a water rinse. The
gels are incubated on a rocking platform or an orbital shaker or at a
moderate speed, for example, 60 rpm. The gel should float freely in
solution, neither stuck to the bottom of the dish nor float on top of the
fixing, washing, and staining solutions. If a stacking gel is used, it should be
544 Thomas H. Steinberg
removed; if a commercial, ‘‘prepoured’’ gel is used, the wells and the foot of
the gel should be trimmed away to reduce staining deposition artifacts,
particularly if the SDS bolus from the sample buffer remains in the gel—
this typically travels with the Bromophenol Blue tracking dye. For all
methods and protocols presented below, pure—dionized or double-
distilled—water is used.
Historically, standard gel staining protocols required several sequential
procedures, some requiring more than one step and some requiring
intervening washes with water or buffer; common to all were:
1. The gel is fixed to immobilize protein in the gel and to remove SDS,
buffers, or other interfering components.
2. The gel is stained with organic dyes or silver deposition formulations.
3. Quenching a staining reaction or destaining to reduce background was
required.
Fixation generally has been accomplished with variations of acidic
alcohol mixtures (e.g., 7–10% acetic acid, 30–50% methanol); organic
dyes and destaining solutions were often formulated with the same solvents
as for fixation. Given that literally hundreds of variations of protein stains
and staining protocols have been introduced in the literature over the
years—often with lengthy, painstaking protocols and competing claims of
detection sensitivity (or other features of utility)—the commercialization of
stains, kits, and protocols have been something of a boon to the yeoman
laboratory scientist, although the proprietary considerations that accompany
commercial development may hinder clear description and understanding
of some of these reagents. Trade names that are useful for broad product
recognition can be a bit confusing. For example, ‘‘Gelcode’’ refers to a
product line that encompasses a range of total and PTM-specific stains,
while ‘‘SYPRO’’ refers to a product line of several total protein gel stains
that are chemically diverse. For the curious, material safety data sheets may
give clues to buffer compositions and the patent literature may allow
gleaning of critical physical or chemical information needed to fully under-
stand a staining method or material composition. Many kits have their basis
on the open literature, and vendors that reveal more rather than less about
the intellectual source and composition of the product are preferred.
Gel staining protocols can vary even within the same staining formula-
tions, depending on the time considerations, and desired quantitative rigor of
detection. Not surprisingly, there is a strong market drive for: (1) instant
gratification and reasonable quantitation and (2) staining formulations with
protocols that avoid strong acids, organic solvents, or heavy metals, due to
environmental and disposal concerns. Thus vendors generally offer ‘‘rapid
protocols’’ for total protein detection formulations. These may involve trun-
cation of fixation or staining steps and, frequently, acceleration of staining or
destaining by judicious heating in a microwave oven. In considering the use
Protein Gel Stains 545
3. Instrumentation: Detection
and Documentation
Obviously, detection by visual inspection has been the basis of protein
gel staining; documentation has been accomplished by photography of
stained gels or by drying stained gels and pasting them into a laboratory
notebook. Unless quantitative data are required, most purification detection
and documentation requirements now can be met with relatively low-cost
flatbed scanners or with camera mounted visible or UV light box stations.
In the past two decades, instrumentation for detection and documentation
(often with quantitation software), driven by fluorescence detection
requirements, has become widely available for both fluorescent and colori-
metric protein detection modalities. Fluorescence detection can be as simple
as visual inspection with a handheld light-emitting diode (LED) ‘‘keychain’’
flashlight or as complex as fully automated multiple mode imaging stations
with linear detection capabilities spanning 4 or 5 orders of magnitude and
correspondingly powerful image analysis and quantitation software for
advanced proteomics applications.
The manufactures of advanced instrumentation generally provide good
information regarding the best acquisition modes and data analysis for
protein colorimetric and fluorescent stains, often providing validated pro-
tocols and examples of applications of commercial or open-source staining
formulations. For fluorescent applications, excitation and emission setting
are frequently identified not only by the wavelength settings, but also by the
names of frequently used fluorescent dyes or products for DNA or protein
stains (e.g., Cy dyes, ethidium bromide, SYBR dyes, SYPRO dyes).
In recent years, new varieties of simple transilluminators with LED or
other inexpensive filtered monochromatic light sources, paired with an
orange-filtering emission cover, offer inexpensive, safe modes of detection
of many of the UV-excitable dyes.
ammonium sulfate, 10% (w/v) phosphoric acid, and 20% (v/v) methanol in
a single stable staining solution, prepared by sequentially combining desired
amounts of phosphoric acid, ammonium sulfate, and powdered dye in an
aqueous solution to 80% of the desired volume, then adding methanol to
the final volume (Candiano et al., 2004). Colloidal staining solutions should
not be filtered.
Most of the commercially available Coomassie staining formulations are
proprietary derivatives of the various formulations and protocols referenced
above.
Recommendations regarding fixation vary, but good results can be
obtained with a simple procedure in which following electrophoresis the
gel is washed in water to remove most of the SDS placed in the staining
solution for several hours to overnight. Protein bands are visible within the
first few minutes, and background is initially clear, and eventually weakly
blue. Following staining the gel is washed with water to reduce back-
ground. A useful, recent succinct summary of Coomassie staining protocols
includes a ‘‘hot’’ Coomassie staining protocol utilizing an acetic acid-based
CBB R-250 formulation at 90 C or a ‘‘fast’’ colloidal TCA-based CBB
G-250 formulation at 50 C (Westermeier, 2006). Several of the commer-
cially available formulations include simple microwave-based protocols to
accelerate staining (and destaining) with CBB G-250 colloidal formulations.
2. The gel is incubated with substances that bind proteins and enhance
silver binding or that interfere with background staining by residual
unbound silver. Taken together, these various strategies are referred to
as sensitization.
3. The gel is impregnated with silver ions.
4. The colorimetric silver stain signal is developed by reduction of bound
silver ions to insoluble and visible metallic silver.
5. Development is stopped, to prevent background staining in the gel matrix
that would obscure protein bands. The time dependency of the latter two
steps in particular contribute to the complexity of the technique, as does
the necessity for several time-dependent washes among steps 2–4.
Silver staining protocols are broadly distinguished on the basis of the
silvering agents and corresponding development conditions. Alkaline meth-
ods use a silver diamine complex in an alkaline environment as a silvering
agent with development in an acidic formaldehyde solution, whereas acidic
methods use weakly acidic silver nitrate as a silvering agent with development
in an alkaline formaldehyde solution. Acidic staining methods have recently
become more popular on the basis of reduced cost and the perception that
background staining is more easily controlled, relative to alkaline methods.
There are a wide variety of sensitizing agents. Aromatic sulfonates such
as naphthalene disulfonate or sulfosalicylic acid bind to proteins and also
bind silver; residual protein-bound SDS may also contribute to increased
silver binding. In a similar vein, CBB-stained gels that are poststained with
silver stain also show enhanced silver staining intensity. Glutaraldehyde is
widely used as a sensitizing agent; the basis of utility is binding to free amino
groups in proteins and thus the introduction of protein-bound reducing
aldehyde groups. Sulfiding reagents such as tetrathionates or thiosulfates can
introduce in close proximity to protein a free S2 ion that immediately
reacts with silver ion to form insoluble silver sulfide.
Mass spectrometry of proteins following silver staining may be problem-
atic due to protein modification resulting from use of glutaraldehyde in
fixation and sensitization steps. Protocols that are compatible with mass
spectrometry omit glutaraldehyde, which may reduce relative sensitivity.
The sensitizing steps rely on tetrathionate and thiosulfate, and the develop-
ment step is also correspondingly extended. In that regard, Colloidal
Coomassie Blue staining followed by a rapid glutaraldehyde-free silver stain-
ing protocol may be a convenient two-step colorimetric approach to evaluate
the purity of a protein sample when subsequent mass spectrometry is desired.
Fluorescent protein gel stain Target Excitation peaks (nm) Emission peak (nm) Vendor
Nile red Total protein 270, 550 600 Sigma-Aldrich
SYPRO Orange Total protein 280, 470 569 Life Technologies
SYPRO Red Total protein 280, 547 631 Life Technologies
SYPRO Tangerine Total protein 280, 490 640 Life Technologies
SYPRO Ruby Total protein 280, 450 610 Life Technologies
Deep Purple Total protein 400, 500 610 GE Healthcare
Krypton Total protein 520 580 Thermo Fisher
Krypton Infrared Total protein 690 718 Thermo Fisher
Flamingo Total protein 270, 512 535 Bio-Rad
LUCY 506 Total protein 505 515 Sigma-Aldrich
LUCY 569 Total protein 569 580 Sigma-Aldrich
C16-FL Total protein 470 530 Life Technologies
Pro-Q Diamond Phosphoprotein 555 580 Life Technologies
Phos-tag 300/460 Phosphoprotein 300, 460 630 Perkin Elmer
Phos-tag 540 Phosphoprotein 540 570 Perkin Elmer
Pro-Q Emerald 300 Glycoprotein 288 533 Life Technologies
Pro-Q Emerald 488 Glycoprotein 512 525 Life Technologies
Krypton glycoprotein Glycoprotein 654 673 Thermo Fisher
Glycoprofile III Glycoprotein 430 480 Sigma-Aldrich
TAMRA alkyne O-GlcNAc 545 580 Life Technologies
Dapoxyl alkyne O-GlcNAc 370 580 Life Technologies
Fluorescent protein gel stains discussed in the text are summarized regarding stain specificity, excitation and emission maxima, and vendor. The web sites are as follows, with
parenthetical indications regarding the division specializing in protein stains, if applicable: Bio-Rad Laboratories, http://www.bio-rad.com; GE Healthcare (Amersham), http://
www.gehealthcare.com; Life Technologies (Invitrogen, Molecular Probes), http://www.lifetechnologies.com; Perkin Elmer, http://www.perkinelmer.com; Sigma-Aldrich,
http://www.sigma-aldrich.com; Thermo Fisher Scientific (Pierce), http://www.thermofisher.com.
Protein Gel Stains 551
precludes use of this after dye SDS–PAGE done with running buffer
containing 0.1% SDS. Photostability can also be problematic.
5. Phosphoprotein Detection
The importance of reversible protein phosphorylation at selected
amino acid residues as a fundamental cell signaling mechanism is beyond
dispute. Current phosphoprotein stains are proprietary formulations with
phosphate-binding moieties covalently attached to fluorophores. The mode
of detection is selective binding to phosphorylated amino acids, but with no
fluorescent enhancement. Many soluble fluorescent compounds can be, to
some degree, total protein stains. A selective phosphoprotein staining solu-
tion combines a phosphate-binding moiety, a fluorophore that shows low
intrinsic total protein staining, a formulation that suppresses nonspecific,
total protein staining, and a destaining protocol that favors dissociation of
residual nonspecific stain. There are only a few phosphates per polypeptide
relative to the number of amino acids targeted by total protein stains, so
phosphoprotein detection sensitivity is generally at the same protein mass
sensitivity level as detection by total proteins as colloidal Coomassie staining
or SYPRO Orange staining, and not as sensitive as with silver or SYPRO
Ruby stain. Protein phosphorylation determination in gels must always be
done with appropriate controls, including gel lanes displaying proteins
of known positive or negative phosphorylation status and, if possible,
control samples treated with an effective phosphatase. After the phospho-
protein signal is detected, the gel should be poststained with a total
protein stain. The commercially available phosphoprotein stains are
compatible with subsequent staining with Coomassie, Silver, or total
fluorescent stains, such as SYPRO Ruby, but not with SDS-dependent
fluorogenic stains such as SYPRO Orange. The phosphoprotein stains are
compatible with subsequent analytical procedures such as mass spectrometry
or peptide sequencing.
Protein Gel Stains 557
6. Glycoprotein Detection
6.1. General glycoprotein detection
Glycosylation is the most frequent protein posttranslational modification in
eukaryotes. Oligosaccharides are usually linked to asparagine side chains
(N-linked glycosylation) or to serine and threonine hydroxyl side chains
558 Thomas H. Steinberg
532-nm laser and a 675-nm long-pass filter, with a two- to fourfold increase
in sensitivity over the observable colorimetric limit.
REFERENCES
Agnew, B., Beechem, J., Gee, K., Haugland, R., Liu, J., Martin, V., Patton, W., and
Steinberg, T. (2006). Compositions and methods for detection and isolation of phosphorylated
molecules. United States Patent 7102005.
Agnew, B., Buck, S., Nyberg, T., Bradford, J., Clarke, S., and Gee, K. (2008). Click
chemistry for labeling and detection of biomolecules. Proc. SPIE 6867, 686708.
560 Thomas H. Steinberg
Barbieri, C. M., and Stock, A. M. (2008). Universally applicable methods for monitoring
response regulator aspartate phosphorylation both in vitro and in vivo using Phos-tag base
reagents. Anal. Biochem. 376, 73–82.
Bell, P. J. L., and Karuso, P. (2003). Fluorescent compounds. United States Patent Applica-
tion 20030157518.
Berggren, K., Chernokalskaya, E., Steinberg, T. H., Kemper, C., Lopez, M. F., Diwu, Z.,
Haugland, R. P., and Patton, W. F. (2000). Background-free, high sensitivity staining of
proteins in one- and two-dimensional sodium dodecyl sulfate-polyacrylamide gels using
a luminescent ruthenium complex. Electrophoresis 21, 2509–2521.
Berggren, K. N., Schulenberg, B., Lopez, M. F., Steinberg, T. H., Bogdanova, A.,
Smejkal, G., Wang, A., and Patton, W. F. (2002). An improved formulation of
SYPRO Ruby protein gel stain: Comparison with the original formulation and with a
ruthenium II tris (bathophenanthroline disulfonate) formulation. Proteomics 2, 486–498.
Berkelman, T. R. (2006). Coumarin-based cyanine dyes for non-specific protein binding. United
States Patent Application 20060166368.
Bhalget, M. K., Diwu, Z., Haugland, R. P., and Patton, W. F. (2001). Luminescent protein
stains and their method of use. United States Patent 6316267.
Candiano, G., Bruschi, M., Musante, L., Santucci, L., Ghiggeri, G. M., Carnemolla, B.,
Orecchia, P., Zardi, L., and Righetti, P. G. (2004). Blue silver: A very sensitive colloidal
Coomassie G-250 staining for proteome analysis. Electrophoresis 25, 1327–1333.
Clark, P. M., Dweck, J. F., Mason, D. E., Hart, C. R., Buck, S. B., Peters, E. C.,
Agnew, B. J., and Hsieh-Wilson, L. C. (2008). Direct in-gel fluorescence detection
and cellular imaging of O-GlcNAc-modified proteins. J. Am. Chem. Soc. 130,
11576–11577.
Coghlan, D. R., Mackintosh, J. A., and Karuso, P. (2005). Mechanism of reversible staining
of protein with epicocconone. Org. Lett. 7, 2401–2404.
Czerney, P. T., Desai, S., Lehmann, F. G., Murtaza, Z. S., Schweder, B. G., Wenzel, M. S.,
and Wolf, B. D. (2008). Protein probe compounds, compositions, and methods. United States
Patent Application 20080026478.
Daban, J.-R. (2001). Fluorescent labeling of proteins with Nile red and 2-methoxy-2,
4-diphenyl-3(2H)-furanone: Physicochemical basis and application to the rapid staining
of sodium dodecyl sulfate polyacrylamide gels and Western blots. Electrophoresis 22,
874–880.
Daban, J.-R., Bartolome, S., and Samso, M. (1991). Use of the hydrophobic probe Nile red
for the fluorescent staining of protein bands in sodium dodecyl sulfate-polyacrylamide
gels. Anal. Biochem. 199, 169–174.
Dunbar, B. S., Kimura, H., and Timmons, T. M. (1990). Protein analysis using high-
resolution two-dimensional polyacrylamide gel electrophoresis. In ‘‘Meth. Enzmyol.’’
(M. P. Deutscher, ed.) 182, pp. 441–459. Academic Press, San Diego, CA.
Fernandez-Patron, C. (2005). Zn2þ reverse staining technique. In ‘‘The Proteomics Proto-
col Handbook’’ ( J. M. Walker, ed.), pp. 215–222. Humana Press, Totowa, NJ.
Fernandez-Patron, C., Castellanos-Serra, L., Hardy, E., Guerra, M., Estevez, E., Mehl, E.,
and Frank, R. W. (1998). Understanding the mechanism of the zinc-ion stains of
biomacromolecules in electrophoresis gels: generalization of the reverse-staining tech-
nique. Electrophoresis 19, 2398–2406.
Garfin, D. E. (1990a). One-dimensional gel electrophoresis. In ‘‘Meth. Enzmyol.’’
(M. P. Deutscher, ed.) 182, pp. 425–441. Academic Press, San Diego, CA.
Garfin, D. E. (1990b). Isoelectric focusing. In ‘‘Meth. Enzmyol.’’ (M. P. Deutscher, ed.)
182, pp. 459–478. Academic Press, San Diego, CA.
Golks, A., and Guerini, D. (2008). The O-linked N-acetylglucosamine modification in
cellular signaling and the immune system. EMBO Rep. 9, 748–753.
Protein Gel Stains 561
Wolf, B.D, Desai, S., Czerney, P. T., Lehmann, F. G., Schweder, B. G., and Wenzel, M. S.
(2007). Protein detection and quantitation using hydroxyquinolone dyes. United States Patent
Application 20070281360.
Yamada, S., Nakamura, H., Kinoshita, E., Kinoshita-Kikuta, E., Koike, T., and Shiro, Y.
(2007). Separation of a phosphorylated histidine protein using phosphate affinity
polyacrylamide gel electrophoresis. Anal. Biochem. 360, 160–162.
Yue, S. T., Steinberg, T. H., Patton, W. F., Cheung, C.-Y., and Haugland, R. P. (2003).
Carbazolylvinyl dye protein stains. United States Patent 6579718.
Zacharius, R. M., Zeli, T. E., Morrison, J. H., and Woodlock, J. J. (1969). Glycoprotein
staining following electrophoresis on acrylamide gels. Anal. Biochem. 30, 148–152.
C H A P T E R T H I R T Y- T W O
Contents
1. Introduction 565
2. Elution of Proteins from Gels by Diffusion 566
2.1. Preparing the gel and protecting the protein 567
2.2. Locating protein in the gel 567
2.3. Elution by diffusion 568
2.4. SDS removal and concentration 568
2.5. Renaturation of enzyme activity 569
2.6. Preliminary tests on your enzyme 569
2.7. Limitations of the method 569
2.8. Many applications 569
3. Replacing the SDS Gel with a Reverse Phase HPLC 570
4. Electrophoretic Elution 570
5. Conclusion 571
References 571
Abstract
New protein biochemical technologies have been developed that allow one to
learn more and more about a protein with less and less material (on the order of
mg). Polyacrylamide gel electrophoresis, which was originally strictly an analyt-
ical method, has in some cases become a high-resolution preparative method.
This chapter will focus on the elution of proteins from SDS gels, with an
emphasis on recovering enzyme or biological activity of the resulting protein.
1. Introduction
SDS polyacrylamide gel electrophoresis has proved to be an incredibly
useful analytical method for determining the number and sizes of polypep-
tides in a sample. When performed skillfully it has the ability to resolve
many individual sized proteins. Many of us in the early days of gel
McArdle Laboratory for Cancer Research, University of Wisconsin-Madison, Madison, Wisconsin, USA
565
566 Richard R. Burgess
71%, 90%, and 99% when carrier BSA concentration in the elution
buffer was 0, 15, and 100 mg/ml, respectively (Hager and Burgess, 1980).
3. Tests showed that more than 99.9% of the SDS remained in the acetone
supernatant. The residual SDS can be removed by washing the pellet
briefly with 1 ml of ice-cold 80% acetone and recentrifuging.
rat, human, Drosophila, and many other organisms. It has been used with
enzyme containing multiple identical subunits and with proteins with
multiple different subunits if the appropriate gel slices have been mixed
together. It has been used with proteins that have essential S–S bridges.
4. Electrophoretic Elution
While the above protocols describe in detail the elution of protein
from gels by diffusion, many of the operations would be perfectly applicable
to electrophoretic elution. While these approaches may result in slightly
higher elution efficiency in slightly shorter time than elution by diffusion,
they are more expensive and harder to scale up to many gel sections.
The Schleicher and Schuell ElutrapTM and the Bio-Rad Model 422 Electro-
eluterTM are the two most widely used commercial electrophoretic elution
apparati. Both involve the placing of a section of a gel containing a protein
band of interest into the apparatus and eluting the protein by electrophoresis
of the protein out of the gel piece, through a large-pore membrane or frit
and into a chamber with a small-pore membrane that will only pass small
molecules or proteins smaller than about 5 kDa. The protein of interest is
pipetted out of the chamber and used as desired. The detailed methods for
using these commercial products can be found in excellent reviews by
Harrington (1990) and Seelert and Krause (2008) and in company literature.
Another commercial approach is the ProteoPLUSTM electroeluter. A small
screw capped tube is constructed with upstream and downstream protein
retention membranes that permit the passage of electric current when the
tubes are placed in an electrophoresis tank. After the protein has eluted from
the gel piece, it can be dialyzed in the same tube for subsequent use.
An example of the application of this approach is given by Lei et al. (2007).
The Bio-Rad Whole Gel Eluter allows one to electrophoretically transfer a
whole slab gel into 26 fractions. Typically a protein sample is applied into
Elution of Proteins from Gels 571
one gel-wide lane and subjected to SDS gel electrophoresis. The 26 grooves
of the Eluter are aligned in parallel to the protein bands and the protein
transferred out of the gel, into the grooves and into the harvesting box.
Another recent advance includes the ability to elute a whole slab gel
into numerous individual fractions suitable for proteomic applications in
multiwall plates (Antal et al., 2007).
5. Conclusion
Protein bands separated by high-resolution gel electrophoresis can be
recovered from the gel and used for a multitude of applications. The protein
can be eluted out of the gel section by crushing the gel and allowing the
protein to diffuse out of the gel or it can eluted by applying an electric field
to the gel section and trapping the eluted protein in an appropriate
membrane bounded apparatus. Submicrogram to 100 mg amounts of
protein can be obtained and often renatured to regain enzymatic activity.
REFERENCES
Antal, J., Banyasz, B., and Buzas, Z. (2007). Shotgun electrophoresis: A proteomic tool for
simultaneous sample elution from whole SDS-polyacrylamide gels slabs. Electrophoresis
28, 508–511.
Burgess, R. R., Arthur, T. M., and Pietz, B. C. (2000). Mapping protein–protein interaction
domains by order fragment ladder far-Western analysis using His-tagged proteins. Meth.
Enzymol. 328, 141–157.
Castellanos-Serra, L., and Hardy, E. (2001). Review. Detection of biomolecules in electro-
phoresis gels with salts of imidazole and zinc II: A decade of research. Electrophoresis 22,
864–873.
Chiari, M., Righetti, P. G., Negri, A., Ceciliani, F., and Ronchi, S. (1992). Preincubation
with cysteine prevents modification of sulfhydryl groups in proteins by unreacted
acrylamide in a gel. Electrophoresis 13, 882–884.
Hager, D. A., and Burgess, R. R. (1980). Elution of proteins from sodium dodecyl sulfate-
polyacrylamide gels, removal of SDS, and renaturation of enzymatic activity: Results
with sigma subunit of E. coli RNA polymerase, wheat germ DNA topoisomerase, and
other enzymes. Anal. Biochem. 109, 76–86.
Haldenwang, W. G., Lang, N., and Losick, R. (1981). A sporulation-induced sigma-like
regulatory protein from B. subtilis. Cell 23, 615–624.
Harrington, M. G. (1990). Elution of protein from gels. Meth. Enzymol. 182, 488–495.
Kunitani, M. G., and Kresin, L. M. (1989). Analysis of alkyl sulfates in protein solutions by
isocratic and gradient ion chromatography. Anal. Biochem. 182, 103–108.
Lei, Z., Anand, A., Mysore, K. S., and Sumner, L. W. (2007). Electroelution of intact
proteins from SDS-PAGE gels and their subsequent MALDI-TOF MS analysis. Methods
Mol. Biol. 355, 353–363.
Manchenko, G. P. (2003). Handbook of detection of enzymes on electrophoresis gels. 2nd
edn. CRC Press LLC, Boca Raton.
572 Richard R. Burgess
Prokipcak, R. D., Harris, D. J., and Ross, J. (1994). Purification and properties of a protein
that binds to the C-terminal coding region of human c-Myc mRNA. J. Biol. Chem. 269,
9261–9269.
Seelert, H., and Krause, F. (2008). Review: Preparative isolation of protein complexes and
other bioparticles by elution from polyacrylamide gels. Electrophoresis 29, 2617–2636.
Weber, K., and Kutter, K. (1971). Reversible denaturation of enzymes by sodium dodecyl
sulfate. J. Biol. Chem. 246, 4504–4509.
Wiggs, J., Gilman, M., and Chamberlin, M. (1981). Heterogeneity of RNA polymerase in
Bacillus subtilis: Evidence for an additional sigma factor in vegetative cells. Proc. Natl. Acad.
Sci. USA 78, 2762–2766.
C H A P T E R T H I R T Y- T H R E E
Contents
1. Western Blotting 574
2. Types of Western Blots 575
2.1. Direct and indirect 575
2.2. Far-Western 577
2.3. Semiquantitative Western blotting 577
3. Detection Methods 579
3.1. Enzyme conjugate 579
3.2. Colorimetric detection 581
3.3. Fluorescence detection 581
3.4. Chemiluminescence detection 582
4. The Chemiluminescence Signal 583
4.1. Signal capture 583
4.2. Signal intensity and duration 583
4.3. Optimizing a Western blot 585
5. Common Problems and their Explanations 588
5.1. No signal 588
5.2. Signal fades quickly 588
5.3. High background 589
5.4. New bottle of substrate does not produce a signal 589
5.5. Brown or yellow bands on the membrane 589
5.6. Bands or entire blot glowing in the darkroom 590
5.7. Ghost/hollow bands 590
6. Blotting and Optimization Protocols using
Chemiluminescent Substrates 593
6.1. Western blotting protocol using chemiluminescent substrates 593
6.2. Western blot stripping protocol 595
6.3. Optimizing antigen concentration 596
573
574 Alice Alegria-Schaffer et al.
Abstract
Immunodetection refers to any detection method that exploits the interaction
of an antibody and antigen. The choice of detection method, such as enzyme-
linked immunosorbent assays (ELISAs) or Western blotting, depends on the
researcher’s preferences and requirements. If a researcher wants to quantify a
low-abundance target protein then a chemiluminescent ELISA is used. If a
researcher wants to identify a protein that is in high abundance, a colorimetric
Western blot will suffice. If there are multiple targets within an assay, then
multiplex fluorescence is typically used.
This article focuses on Western blotting. Although colorimetric and fluores-
cent detection methods are discussed, chemiluminescent detection is used
most often and is, therefore, discussed in great detail. Included is specific
information about the chemiluminescent signal and factors that affect its inten-
sity and longevity. We also describe types of blotting and present data and
suggestions for obtaining semiquantitative data. Although classical Western
blotting is typically used for qualitative purposes, we present information about
effective quantitative analysis using specific controls. Common occurrences
within the methodology and their possible explanations are also detailed.
One frequent result is the appearance of ghost bands, which, based on our
research, can be caused by high amounts of target or antibody cross-reactivity.
Also included are the basic Western blot protocol and protocols for trouble-
shooting common problems and optimizing many of the specific factors that
influence results.
1. Western Blotting
Western blotting is a powerful and commonly used tool to identify
and quantify a specific protein in a complex mixture (Towbin et al., 1979).
The technique enables indirect detection of protein samples immobilized
on a nitrocellulose or polyvinylidene fluoride (PVDF) membrane. In a
conventional Western blot, protein samples are first resolved by sodium
dodecyl sulfate polyacrylamide gel electrophoresis (SDS–PAGE) and then
electrophoretically transferred to the membrane. Following a blocking step
using a nonrelevant protein, the membrane is probed with a primary
antibody (poly- or monoclonal) that was raised against the target antigen.
The membrane is then washed and incubated with an enzyme-conjugated
Performing and Optimizing Western Blots 575
A Substrate
Detectable Enzyme
product
Enzyme
Figure 33.1 Schematic of direct and indirect Western blotting methods. In the direct detection method, labeled primary antibody binds to
antigen on the membrane and reacts with the substrate, creating a detectable signal. In the indirect detection method, unlabeled primary
antibody binds to the antigen. Then a labeled secondary antibody binds to the primary antibody and reacts with the substrate.
Performing and Optimizing Western Blots 577
2.2. Far-Western
Occasionally, an antibody to a specific antigen is unsuitable for Western blot
analysis or simply unavailable. Blotting is still possible if a binding partner to
the target protein is available for use as a probe. This type of application is
referred to as a far-Western blot and is routinely used for the discovery or
confirmation of a protein–protein interaction. Variations on this theme are
myriad and involve all the previously mentioned strategies. As with primary
antibodies, labeled binding partners are used, which are frequently labeled
in an in vitro translation reaction with 35S. Biotinylating the probe and
detecting with an avidin or avidin-like conjugate is another possibility and
has the added effect of amplifying the signal. Care must be taken to ensure
the probe is not overlabeled, which can compromise its ability to interact
with the target. Alternatively, expressing a recombinant probe in bacteria
with a tag, such as GST, HA, c-Myc, or FLAG, enables detection via a
labeled antibody to the particular tag (Burgess et al., 2000).
1
DyLightTM, MemCodeÒ , NeutrAvidinÒ , PierceÒ , RestoreÒ and SuperSignalÒ are trademarks of Thermo
Fisher Scientific.
578 Alice Alegria-Schaffer et al.
with an ELISA. Often antibodies directed against one protein may interact
with equal specificity with another closely related protein. In such situa-
tions, an ELISA may generate false positives or overestimation of the target
protein abundance. Because Western blotting involves resolution of pro-
teins by gel electrophoresis, variations in molecular weight can be exploited
to distinguish and quantify the target protein alone (Sato et al., 2002; Xing
and Imagawa, 1999).
To evaluate the effectiveness and accuracy of quantitative Western
blotting, we used a purified target protein as an internal control and to
create a standard curve. The sample must contain sufficient target protein
such that its amount was within the standard curve range. To convert band
intensity into a quantitative measurement, the Western blot was analyzed
densitometrically using a CCD camera and imaging system. Finally, an
ELISA was used to confirm the trends observed by Western blot
(Mathrubutham and Vattem, 2005).
2
SuperSignalÒ Technology is protected by U.S. Patent # 6,432,662.
Performing and Optimizing Western Blots 579
3. Detection Methods
3.1. Enzyme conjugate
Alkaline phosphatase (AP, 140 kDa) used to be the enzyme of choice
and was typically detected with precipitating chromogenic substrates.
Colorimetric reactions proceed at a steady rate, allowing accurate control
of relative sensitivity and reaction development. As protein research pro-
gressed, HRP (40 kDa) became more popular because of its stability and
smaller size, which enables more molecules conjugated per IgG and greater
sensitivity. Furthermore, chemiluminescent substrates for HRP enabled
even higher sensitivity.
C
1,50,000
1,00,000
Net intensity
1 2 3 4 5 6 7 8
A
50,000
B
0
0 1 2 3 4
IkB (ng)
E
0.014
D
0.040 0.012
0.035 0.010
0.030 0.008
0.025 0.006
0.020 0.004
d in in d l l l
te 4h te m
g/ in
m
g/ in
m
g/ 4 h
ul
a 5m 30
m
F
2
ul
a m m m m m
tim
F F EG tim 50 5 50 30 50 2
ns EG EG ns F F F
U U EG EG EG
Figure 33.2 Quantitation of IkBa in EGF-treated cells. Panel A: Known amounts of recombinant IkBa (rIkBa) were Western blotted to
generate a standard curve. Lanes 1–6 contain 0.15–0.012 ng of rIkBa. Panel B: Lysates of EGF-treated A431 cells were Western blotted for
IkBa. Treatments were as follows: nontreated (lanes 1, 2), 50 ng for 5 min (lanes 3, 4); 50 ng for 30 min (lanes 5, 6); and 50 ng for 24 h (lanes
7, 8). Panel C: Recombinant IkBa standard curve (r2 ¼ 0.99). Panel D: Variation in the amount of IkBa with EGF exposure time as
determined by quantitative Western blotting. Panel E: Quantitation by ELISA.
Performing and Optimizing Western Blots 581
3
DyLightTM, MemCodeÒ , NeutrAvidinÒ , PierceÒ , RestoreÒ and SuperSignalÒ are trademarks of Thermo
Fisher Scientific.
Performing and Optimizing Western Blots 583
O O * O
NH O− O−
H2O2 + HRP + Light
NH O− O−
of HRP and a peroxide buffer, luminol oxidizes and forms an excited state
product that emits light as it decays to the ground state. Light emission occurs
only during the enzyme–substrate reaction and, therefore, once the substrate
in proximity to the enzyme is exhausted, signal output ceases. In contrast,
colorimetric substrates, such as TMB, produce precipitate that remains
visible on the membrane even after the reaction has terminated.
120
Too much enzyme
Appropriate enzyme
100
80
Intensity
60
40
20
0
1 2 3 4 5 6
Time (min)
Figure 33.4 Example signal emission curves. When there is too much enzyme present
in a chemiluminescent Western blot system, signal output peaks soon after substrate
application and rapidly exhausts the substrate. In an optimal system, the signal emission
peaks approximately 3 min after applying the substrate and plateaus for several hours.
used and the enzyme-to-substrate ratio present in the system. Although the
amount of substrate on a blot is relatively constant, the amount of enzyme
present depends on how much was added and other factors (Table 33.1).
Too much enzyme conjugate applied to a Western blot system is the single
greatest cause of signal variability, high background, short signal duration,
and low sensitivity.
A signal emission curve that decays slowly (Fig. 33.4) is desirable because
it demonstrates that each component of the system has been optimized and
Performing and Optimizing Western Blots 585
allows reproducible results. A signal that decays too quickly can cause
variability, low sensitivity and lack of signal documentation. A long-lasting
signal minimizes variability with transfer efficiency, different manufacturer
lots of substrate and other factors.
Although HRP continues its activity for as long as substrate is available,
HRP can become inactive with prolonged exposure to substrate. Free
radicals produced during the oxidation reaction can bind to HRP in such
a way that the enzyme can no longer interact with the substrate.
An abundance of HRP in the system in turn produces an abundance of
free radicals that increases the probability of HRP inactivation. Free radicals
also can damage the antigen, antibodies and the membrane, prohibiting
reprobing effectiveness.
Some common protein stains for membranes include: Ponceau S, a red stain
that fades with time; coomassie dye, a sensitive blue stain that tightly binds
to proteins and can interfere with Western blotting; and Thermo Scientific
MemCode4 Stain (Thermo Fisher Scientific), an easily reversible blue stain
(Antharavally et al., 2004).
4.3.4. Antibodies
Not only is the affinity of the primary antibody for the antigen important,
but primary and secondary antibody concentrations also have a profound
effect on signal intensity. Too much HRP on the blot can be caused by
either primary and secondary antibody concentrations or both. Minimal
primary antibody is advantageous, as it promotes target-specific binding and
low background.
If a blot failed to generate an adequate signal, removing all detection
reagents (stripping) from the blot and reprobing with either a different
primary antibody or different concentrations of antibodies often conserves
valuable sample and time; however, insufficient stripping can leave active
HRP on the blot that will produce a signal. Applying substrate on the
stripped blot and subsequent detection will reveal if active HRP remains on
4
DyLightTM, MemCodeÒ , NeutrAvidinÒ , PierceÒ , RestoreÒ and SuperSignalÒ are trademarks of Thermo
Fisher Scientific.
5
TweenÒ is a trademark of ICI Americas.
Performing and Optimizing Western Blots 587
A B
Before After
treatment treatment
Figure 33.5 Reduction of background signal by chemical treatment of the X-ray film.
Recombinant human TNFa was separated by SDS–PAGE and transferred to a nitrocel-
lulose membrane. The blot was probed with mouse antihuman TNFa and goat
antimouse-HRP antibodies and detected with SuperSignal West Dura Substrate.
Panel A: The blot was exposed to film for 30 s, resulting in considerable background
speckling. Panel B: The film was then treated with Pierce Background Eliminator for
2 min to remove speckling.
588 Alice Alegria-Schaffer et al.
6
DyLightTM, MemCodeÒ , NeutrAvidinÒ , PierceÒ , RestoreÒ and SuperSignalÒ are trademarks of Thermo
Fisher Scientific.
Performing and Optimizing Western Blots 591
500
200
100
50
10
5
NFkB (ng)
1:1,000 1:5,000
(10ng/ml) (2ng/ml)
1:20,000 1:50,000
(0.5ng/ml) (0.2ng/ml)
Figure 33.6 Secondary antibody concentration affects the appearance of ghost bands.
Western blots were performed using various concentrations of secondary antibody.
Ghost bands (oval) are present in lanes containing >200 ng of NFkB. Membranes were
incubated in SuperSignal West Pico Substrate for 5 min and exposed to X-ray film for
30 s.
A 1 2 3 4 1 2 3 4 B
p53 Ghost
bands
Milk: 0.5% 2%
1 2 3 4 1 2 3 4
Protein
of interest
p53
Milk: 5% 10%
A431 lysates
Figure 33.7 Low concentration of nonfat milk enhances the detection of low-abun-
dance proteins, but antibodies cross-reacting with milk proteins result in ghost bands.
Panel A: Lanes 1–4 are lysate samples from different cell harvest times (0, 8, 20, and
30 h) after treatment with 50 mM cisplatin. Different concentrations of nonfat milk were
used for blocking and antibody incubation. The low concentration of milk (0.5%)
enhanced detection of p53. Panel B: Membrane was incubated with 5% nonfat
milk blocking buffer for 48 h. The primary antibody bound to the blocked regions,
producing ghost bands.
Gel
Transfer membrane
Filter paper
Pads
Support grid
Gel/membrane/filter
sandwich
Buffer tank
Anode (+)
Electrodes
Cathode (−)
Direction of
transfer
Table 33.3 Primary and secondary antibody concentrations to use with Thermo
Scientific Chemiluminescent Substrates
10. Prepare the substrate. Use a sufficient volume to ensure that the blot is
completely wetted with substrate and the blot does not become dry
(0.1 ml/cm2).
11. Incubate the blot with substrate for 1 min when using ECL or 5 min
when using SuperSignal Substrates.
Performing and Optimizing Western Blots 595
12. Remove the blot from the substrate and place it in a plastic membrane
protector. A plastic sheet protector works well, although plastic wrap
also may be used. Remove all air bubbles between the blot and the
surface of the membrane protector.
13. Image the blot using film or a cooled CCD camera.
3. After washing, cut the blot into strips containing the target.
4. Probe each strip with a different enzyme conjugate concentration. For
example, for SuperSignal West Pico Substrate use 1:40,000, 1:60,000,
and 1:80,000 dilutions (from a 1-mg/ml stock). Incubate strips for 1 h at
room temperature with rocking.
5. Wash the strips and add the substrate. After substrate incubation, image
the strips.
6. Wait 1–2 h and image the strips a second time.
7. Evaluate the results. Use the dilution that produces the strongest signal.
REFERENCES
Antharavally, B. S., Carter, B., Bell, P. A., and Mallia, K. A. (2004). A high-affinity
reversible protein stain for Western blots. Anal. Biochem. 329(2), 276–280.
Bergendahl, V., Glaser, B. T., and Burgess, R. R. (2003). A fast western blot procedure
improved for quantitative analysis by direct fluorescence labeling of primary antibodies.
J. Immunol. Methods 277, 117–125.
Burgess, R. R., Arthur, T. M., and Pietz, B. C. (2000). Mapping protein–protein interaction
domains using ordered fragment ladder far-Western analysis of hexahistidine-tagged
fusion proteins. Methods Enzymol. 328, 141–157.
Choudhary, S., Lu, M., Cui, R, and Brasier, A. R. (2007). Involvement of a novel Rac/
RhoA guanosine triphosphatase-nuclear factor-kappaB inducing kinase signaling path-
way mediating angiotensin II-induced RelA transactivation. Mol. Endocrinol. 21(9),
2203–2217.
Performing and Optimizing Western Blots 599
Desai, S., Dworecki, B., and Cichon, E. (2001). Direct immunodetection of antigens within
the precast polyacrylamide gel. Anal. Biochem. 297, 94–98.
Kaufmann, S. H., Ewing, C. M., and Shaper, J. H. (1987). The erasable Western blot. Anal.
Biochem. 161, 89–95.
Kwok, T. T., Mok, C. H., and Menton-Brennan, L. (1994). Up-regulation of a mutant
form of p53 by doxorubicin in human squamous carcinoma cells. Cancer Res. 54,
2834–2836.
Mathrubutham, M., and Vattem, K. (2005). Methods and considerations for quantitative
Western blotting using SuperSignalÒ Chemiluminescent Substrates. In ‘‘Thermo Fisher
Scientific Application Note # 12’’. www.thermo.com/pierce.
Mattson, D. L., and Bellehumeur, T. G. (1996). Comparison of three chemiluminescent
horseradish peroxidase substrates for immunoblotting. Anal. Biochem. 240, 306–308.
Patonay, G., and Antoine, M. D. (1991). Near-infrared fluorogenic labels: New approach to
an old problem. Anal. Chem. 63, 321A–327A.
Ramlau, J. (1987). Use of secondary antibodies for visualization of bound primary reagents in
blotting procedures. Electrophoresis 8, 398–402.
Sasse, J., and Gallagher, S. R. (2008). Detection of proteins on blot transfer membranes.
Curr. Protoc. Immunol. Chapter 8, Unit 8.10B.
Sato, M., Nishi, N., Shoji, H., Kumagai, M., Imaizumi, T., Hata, Y., Hirashima, M.,
Suzuki, S., and Nakamura, T. (2002). Quantification of Galectin-7 and its localization
in adult mouse tissue. J. Biochem. 131, 255–260.
Sowell, J., Strekowski, L., and Patonay, G. (2002). DNA and protein applications of near-
infrared dyes. J. Biomed. Opt. 7, 571–575.
Spinola, S. M., and Cannon, J. G. (1985). Different blocking agents cause variation in the
immunologic detection of proteins transferred to nitrocellulose membranes. J. Immunol.
Meth. 81, 161.
Sun, L., and Carpenter, G. (1998). Epidermal growth factor activation of NF-kappaB is
mediated through IkappaBalpha degradation and intracellular free calcium. Oncogene 23,
2095–2102.
Tipsmark, C. K., Strom, C. N., Bailey, S. T., and Borski, R. J. (2008). Leptin stimulates
pituitary prolactin release through an extracellular signal-regulated kinase-dependent
pathway. J. Endocrinol. 196(2), 275–281.
Towbin, H., Staehelin, T., and Gordon, J. (1979). Electrophoretic transfer of proteins from
polyacrylamide gels to nitrocellulose sheets: Procedure and some applications. Proc. Natl.
Acad. Sci. USA 76, 4350–4354.
Vattem, K., and Mathrubutham, M. (2005). Factors that cause the appearance of ghost bands
when using chemiluminescent detection systems in a Western blot. In ‘‘Thermo Fisher
Scientific Application Note # 11’’. www.thermo.com/pierce.
Walker, G. R., Feather, K. D., Davis, P. D., and Hines, K. K. (1995). SuperSignalTM
CL-HRP: A new enhanced chemiluminescent substrate for the development of the
horseradish peroxide label in Western blotting applications. J. NIH Res. 7, 76.
Xing, C., and Imagawa, W. (1999). Altered MAP kinase (ERK1, 2) regulation in primary
cultures of mammary tumor cells: Elevated basal activity and sustained response to EGF.
Carcinogenesis 20, 1201–1208.
C H A P T E R T H I R T Y- F O U R
Detergents: An Overview
Dirk Linke
Contents
1. Introduction 604
2. Detergent Structure 604
3. Properties of Detergents in Solution 605
3.1. Phase diagrams and critical micellization concentration 609
3.2. Effects of temperature: Detergent solubility, krafft point,
cloud point, and phase separation 611
4. Exploiting the Physicochemical Parameters of Detergents for
Membrane Protein Purification 612
5. Detergent Removal and Detergent Exchange 613
6. Choosing the Right Detergent 613
6.1. Optical spectroscopy 614
6.2. Mass spectrometry and nuclear magnetic resonance (NMR)
measurements 615
6.3. Protein crystallization 615
7. Conclusions 615
Acknowledgments 616
References 616
Abstract
Detergents are used in molecular biology laboratories every day. They are
present in cell lysis buffers (e.g., in kits for plasmid isolation), in electrophoresis
and blotting buffers, and, most importantly, they are used for cleaning laboratory
glassware and the hands of the laboratory staff. For these routine applications,
a detailed knowledge of detergent properties is not necessary—they just work.
When it comes to the isolation and purification of membrane proteins, one cannot
rely on routine protocols. Many membrane proteins are only stable in a small
number of different detergent buffer systems, and worst of all, different mem-
brane proteins prefer very different detergents. Unfortunately, detergent proper-
ties are considered the domain of colloid science or physical chemistry, and thus,
while the available amount of physico-chemical data on detergents is astound-
ing, this data is rarely compiled in a way that is useful to biochemists. The aim of
Department I, Protein Evolution, Max Planck Institute for Developmental Biology, Tübingen, Germany
603
604 Dirk Linke
1. Introduction
What is a detergent? The term detergent is used in many different
contexts. A household detergent is a complex mixture of surfactants (to dissolve
grease), abrasives (to scour), chelating agents (to counter the effect of ions that
contribute to water hardness), oxidants (for bleaching), enzymes (to degrade
fats, proteins, or complex carbohydrates), colors, perfumes, optical brighteners,
buffer substances (to stabilize the pH), and a number of other stabilizing
ingredients, for example, to modify the foaming properties or to inhibit
bacterial or fungal growth. The term is sometimes also used to distinguish
soap from other cleaning agents (detergents). In molecular biology laboratories,
the term detergent is typically used as a synonym for surfactant. Surfactants
(‘‘surface acting agents’’) are wetting agents that lower the surface tension of a
liquid, and the interfacial tension between two liquids. The chemical and
physical basis of these properties is discussed in detail in this chapter. Obviously,
surfactant is the more precise term, but in scientific literature, it is only used in
the context of physical chemistry and colloid science; in biochemistry, the term
detergent is widespread, and the author of this chapter is not going to try and
change this—from here on, detergent is used as a synonym for surfactant.
Historically, the first detergents used were saponins and soap (or rather,
the fatty acid salts therein). Soap is made from vegetable oils or animal fats
which are hydrolyzed by lye to yield fatty acid salts and glycerol; this process
was already used in ancient Babylon and Egypt, and was brought to
perfection by chemists during the early middle ages in the Middle East.
Saponins are natural detergents that can be extracted from plants, for
example from Soapwort (Saponaria officinalis) or Soapnut (Sapindus). While
soaps are not used in membrane protein science, the saponin Digitonin from
Purple Foxglove (Digitalis purpurea) is. The first synthetic detergents were
developed and used during World War I in Germany, when soap was
scarce. Synthetic detergents have replaced soap in many household applica-
tions since then. Most detergents used in laboratory applications today were
originally developed as constituents of laundry or industry detergents.
2. Detergent Structure
Detergents (or surfactants) are organic compounds of very diverse
structure. Generally speaking, detergent molecules consist of two parts: an
extended, hydrophobic hydrocarbon moiety, and a polar or charged
Detergents: An Overview 605
Nonionic detergents
Big CHAP Gluconamidopropyl Cholesterol 2.9–3.4 10 878.1 9 – SID: 26758300
moieties (2) derivative
Deoxy-Big CHAP Gluconamidopropyl Cholesterol 1.1–1.4 11 862.1 11 – SID: 26758553
moieties (2) derivative
Digitonin Complex Cholesterol – 60–70 1229.3 – – SID: 168187 Natural compound
polysaccharide derivative with high lot-to-lot
variability
Brij-35 (C12E23) Linear PEG (23) Linear hydrocarbon 0.09 40 1199.6 48 >100 SID: 24898176
alcohol (C12)
C12E8 Linear PEG (8) Linear hydrocarbon 0.11 123 538.8 66 74–79 SID: 24898996
alcohol (C12)
C10E4 Linear PEG (4) Linear hydrocarbon 0.64–0.81 54 334.5 18 20 SID: 3729667
alcohol (C10)
C10E6 Linear PEG (6) Linear hydrocarbon – 74 422.6 31 – SID: 24874132
alcohol (C10)
C8E4 Linear PEG (4) Linear hydrocarbon 6.5–8.5 – 306.4 26 35–40 SID: 24900138
alcohol (C8)
C8POE Linear PEG Linear hydrocarbon 6.6 – 330 (average) – 58 Mixture of molecules
alcohol (C8) with different
headgroup lengths
(E ¼ 2–9)
Triton X-100 Linear PEG p-(2,2,4,4- 0.17–0.3 100–150 630 (average) 80 64–65 SID: 7889640 Mixture of molecules
Tetramethylbutyl) with different
phenol headgroup lengths
(average 9.6),
strong UV
absorption
Triton X-114 Linear PEG p-(2,2,4, 0.2–0.35 – 540 (average) – 20–25 SID: 24902129 Mixture of molecules
4-Tetramethylbutyl) with different
phenol headgroup lengths
(average 7–8)
classic detergent for
phase separation,
strong UV
absorption
NP-40 Linear PEG p-(2,2,4, 0.3 100–150 600 60–90 63–67 SID: 813297 Mixture of molecules
4-Tetramethylbutyl) (average) with different
phenol headgroup lengths
(average 9.0),
strong UV
absorption
Tween-20 Polysorbate Linear fatty 0.059 – 1230 – 76 SID: 47207229 Mixture of molecules
acid (C12) (average) with different
headgroup lengths
Tween-80 Polysorbate Linear fatty acid, 0.012 58 1310 76 93 SID: 7848130 Mixture of molecules
unsaturated (C18:1) (average) with different
headgroup lengths
b-Dodecylmaltoside b-Glycosidic maltose Linear hydrocarbon 0.15 98 510.6 70 <0 SID: 691815
alcohol (C12)
b-Decylmaltoside b-Glycosidic maltose Linear hydrocarbon 1.6 – 482.6 – <0 SID: 24894148
alcohol (C10)
b-Octylglucoside b-Glycosidic glucose Linear hydrocarbon 20–25 84 292.4 25 <0 SID: 205023
alcohol (C8)
b-Octylthioglucoside b-Glycosidic glucose Linear hydrocarbon 9 – 308.4 – 7 SID: 698369
thiol (C8)
MEGA-8 N-Methylglucamine Linear fatty acid (C8) 58 – 321.5 – – SID: 737553
MEGA-9 N-Methylglucamine Linear fatty acid (C9) 19–25 – 335.5 – – SID: 737552
MEGA-10 N-Methylglucamine Linear fatty acid (C10) 6–7 – 349.5 – – SID: 751715
LDAO N-Oxide Linear hydrocarbon 2.1–8.3 74 229.4 17 – SID: 158752 Can also be considered
amine as a zwitterionic
detergent
(continued)
Table 34.1 (continued)
Zwitterionic detergents
CHAPS Dimethylammonium- Cholesterol derivative 6.5 4–14 614.9 6 SID: 684915
1-propanesulfonate
CHAPSO Dimethylammonium- Cholesterol derivative 8 11 630.9 7 SID: 698662
1-propanesulfonate
SB-10 Dimethylammonium- Linear hydrocarbon 25–40 41 307.6 12.5 SID: 738617 Zwittergent 3-10
1-propanesulfonate alcohol (C10)
SB-12 Dimethylammonium- Linear hydrocarbon 2–4 55 335.6 18.5 SID: 661588 Zwittergent 3-12
1-propanesulfonate alcohol (C12)
SB-14 Dimethylammonium- Linear hydrocarbon 0.1–0.4 83 363.6 30 SID: 661589 Zwittergent 3-14
1-propanesulfonate alcohol (C14)
Anioinic detergents
Cholate, sodium salt Carboxylic acid Cholesterol derivative 9–15 2 430.6 0.9 SID: 152893 Typically as sodium
salt
Desoxycholate, Carboxylic acid Cholesterol derivative 4–8 4–10 414.6 1.7–4.2 SID: 668198 Typically as sodium
sodium salt salt
Lauroylsarcosine Methylaminoacetate Linear fatty acid (C12) 2 293.4 0.6 SID: 151872 Typically as sodium
(‘‘Sarkosyl’’), salt
sodium salt
Sodium dodecylsulfate Sulfate Linear hydrocarbon 7–10 62 288.4 18 SID: 152210 Also available as
(SDS) alcohol (C12) lithium salt, with
better solubility at
cold temperatures
Cationic detergents
Cetyltrimethyla- Quarternary amine Linear hydrocarbon 0.9 170 364.5 62 SID: 148809 Typically as bromide
mmonium bromide (C16) salt or chloride salt
(CTAB)
Dodecyltrimethy- Quarternary amine Linear hydrocarbon (C12) 15 70 280.3 20 SID: 159639 Typically as bromide
lammonium salt or chloride salt
bromide (DTAB)
a
Note that cloud points are influenced by the buffer composition, see text.
b
Link to PubChem structures and annotations (http://www.ncbi.nlm.nih.gov/sites/entrez?db¼pcsubstance).
Data were compiled from several reviews (Arnold and Linke, 2007, 2008; Helenius et al., 1979; Hinze and Pramauro, 1993; Neugebauer, 1990), and from datasheets of commercial suppliers of detergents.MW, molecular weight;
CMC, critical mizellization concentration.
Detergents: An Overview 609
P > 1/2
P < 1/3
Figure 34.1 The formation of micelles depends on the molecular shape of the deter-
gent. It is best described by the packing parameter P that is calculated by comparing the
headgroup volume with the hydrophobic chain volume and length.
A
Phase separation
Temperature
Micellar
Concentration
B
Temperature
Micellar
Liquid
Monomeric Phase separation crystalline
Concentration
Figure 34.2 Simplified phase diagrams. Panel A shows a detergent with a lower
consolute boundary. Most nonionic detergents fall into this group. Panel B shows a
detergent with an upper consolute boundary. A number of glycosidic and zwitterionic
detergents fall into this group. Note that the liquid-crystalline phase can also consist of
hexagonally packed cylindrical micelles.
y
lit
bi
lu
So
CMC
Krafft point
Temperature
Figure 34.3 Definition of the Krafft point. The gray area describes the possible
concentrations of detergent soluble at a given temperature. Above the Krafft point,
detergent solubility increases dramatically as micelles can be formed.
612 Dirk Linke
7. Conclusions
The most important conclusion, when talking about detergents in
membrane protein research, is that one always has to consider or even test
whether a certain membrane protein–detergent combination is suitable for
the isolation and purification procedure, but also for the downstream
applications planned. To even increase the complexity of the problem,
some membrane proteins are not stable without at least a little amount of
native membrane lipids; this implies that solubilization and purification
should in some cases not be complete. It also implies that sometimes
mixtures of detergents (or lipids) might be more suitable than a pure
substance. This is, in part, realized by commercially available detergent
616 Dirk Linke
ACKNOWLEDGMENTS
The author thanks Andrei Lupas for continuing support. Funding by the Max Planck
Society, the German Science Foundation (SFB766/B4), and the Bill & Melinda Gates
Foundation (Grand Challenges Explorations Program) is gratefully acknowledged.
REFERENCES
Antharavally, B. S., Mallia, K. A., Rangaraj, P., Haney, P., and Bell, P. A. (2009).
Quantitation of proteins using a dye-metal based colorimetric protein assay. Anal.
Biochem. 385, 342–345.
Arnold, T., and Linke, D. (2007). Phase separation in the isolation and purification of
membrane proteins. Biotechniques 43, 427–440.
Arnold, T., and Linke, D. (2008). The use of detergents to purify membrane proteins. Curr.
Protoc. Protein Sci. Chapter 4, 4.8.1–4.8.30.
Ashani, Y., and Catravas, G. N. (1980). Highly reactive impurities in Triton x-100 and
Brij 35: Partial characterization and removal. Anal. Biochem. 109, 55–62.
DeGrip, W. J., van Oostrum, J., and Bovee-Geurts, P. H. M. (1998). Selective detergent-
extraction from mixed detergent/lipid/protein micelles, using cyclodextrin inclusion
compounds: A novel generic approach for the preparation of proteoliposomes. Biochem.
J. 330, 667–674.
Fernandez, C., and Wüthrich, K. (2003). NMR solution structure determination of mem-
brane proteins reconstituted in detergent micelles. FEBS Lett. 555, 144–150.
Fromme, P., and Witt, H. T. (1998). Improved isolation and crystallization of photosystem I
for structural analysis. Biochim. Biophys. Acta Bioenerg. 1365, 175–184.
Furth, A. J., Bolton, H., Potter, J., and Priddle, J. D. (1984). Separating detergent from
proteins. Methods Enzymol. 104, 318–328.
Gu, T. R., and Sjoblom, J. (1992). Surfactant structure and its relation to the Krafft point,
cloud point and micellization—Some empirical relationships. Colloids Surf. 64, 39–46.
Helenius, A., McCaslin, D. R., Fries, E., and Tanford, C. (1979). Properties of detergents.
Methods Enzymol. 56, 734–749.
Hinze, W. L., and Pramauro, E. (1993). A critical review of surfactant-mediated phase
separations (cloud-point extractions)—Theory and applications. Crit. Rev. Anal. Chem.
24, 133–177.
Detergents: An Overview 617
Ishihama, Y., Katayama, H., and Asakawa, N. (2000). Surfactants usable for electrospray
ionization mass spectrometry. Anal. Biochem. 287, 45–54.
Mooney, R. A. (1988). Use of digitonin-permeabilized adipocytes for cAMP studies.
Methods Enzymol. 159, 193–202.
Neugebauer, J. M. (1990). Detergents: An overview. Methods Enzymol. 182, 239–253.
Norris, J. L., Porter, N. A., and Caprioli, R. M. (2003). Mass spectrometry of intracellular
and membrane proteins using cleavable detergents. Anal. Chem. 75, 6642–6647.
Privé, G. G. (2007). Detergents for the stabilization and crystallization of membrane
proteins. Methods 41, 388–397.
Rigaud, J. L., Levy, D., Mosser, G., and Lambert, O. (1998). Detergent removal by
non-polar polystyrene beads—Applications to membrane protein reconstitution and
two-dimensional crystallization. Eur. Biophys. J. Biophys. Lett. 27, 305–319.
C H A P T E R T H I R T Y- F I V E
Contents
1. Introduction 619
2. Preparation of Membranes 620
3. Solubilization of Native Membrane Proteins 622
4. Purification of Membrane Proteins 625
4.1. Lectin-affinity chromatography 626
4.2. Ligand-affinity chromatography 627
4.3. Antibody-affinity chromatography 627
5. Detergent Removal and Detergent Exchange 628
6. Expression and Purification of Recombinant Integral
Membrane Proteins 628
References 629
Abstract
Membrane proteins are pivotal players in biological processes. In order to
understand how a membrane protein works, it is important to purify the protein
to fully characterize it. Membrane proteins are difficult to purify because they
are present in low levels and they require detergents to become soluble in an
aqueous solution. The selection of detergents suitable for the solubilization and
purification of a specific membrane protein is critical in the purification of
membrane proteins. The aim of this chapter is to provide an overview for the
isolation of plasma membranes, selection of detergents for solubilization of
membrane proteins, and how the choice of detergents may affect membrane
protein purification.
1. Introduction
Membrane proteins are pivotal players in biological processes. They
are responsible for connecting cells to each other and to the cell matrix, for
organizing the shape of the organelles and the cells, and for the transport of
619
620 Sue-Hwa Lin and Guido Guidotti
ions, metabolites, and proteins across plasma membranes, and RNA trans-
port across nuclear membrane. Given the importance of membrane proteins
for a plethora of cellular functions, it is not surprising that around 50% of the
current drugs for a variety of diseases target membrane proteins.
In order to understand how a membrane protein works and to generate
drugs that target specific sites within the protein, it is important to purify
the protein to fully characterize it. However, of all known proteins that
were purified and identified biochemically, only very few were membrane
proteins. This is in a large part due to the fact that membrane proteins are
more difficult to purify for the following reasons. First, most membrane
proteins are present at low levels; and second, they are embedded in the
lipid bilayer and require detergents to become soluble in aqueous systems.
The selection of detergents suitable for the solubilization and purification of
the specific membrane protein is critical in the purification of membrane
proteins. In general, the basic principle for the purification of membrane
proteins are the same as those used for soluble proteins, but modifications in
the purification scheme is required in order to accommodate the require-
ment of detergents that are essential in maintaining the solubility and
function of these proteins.
Unlike DNA or RNA, it is impossible to present a single set of methods
for the purification of all proteins, especially membrane proteins. Each
membrane protein possesses a unique set of physical characteristics, prefer-
ence for detergent for solubilization, and conditions for purification. An
overview of the properties of detergents is described in Chapter 41. This
chapter describes methods for the isolation of plasma membranes, and for
solubilization and purification of membrane proteins.
2. Preparation of Membranes
Isolation of plasma membranes from cells or tissues is the first step in
the purification of a plasma membrane protein. Because of the limited
biochemical fractionation methods available for effective separation of
detergent solubilized membrane proteins, the time invested in purifying
plasma membrane fractions will improve the results at the later stages.
Due to the low levels of most membrane proteins, it is important to
select a tissue or cell line that is easy to obtain a large amount of starting
material for purification and that also expresses high specific activity of the
protein of interest. Recently, there has been increased interest in identifying
cell surface proteins as markers for different cell lineages or stem cells. As a
result, obtaining a sufficient amount of plasma membranes from a limited
amount of cells with relatively good enrichment of plasma membrane
marker enzyme activity has become a new focus of membrane isolation.
Purification of Membrane Proteins 621
In the situations, when only a limited amount of tissue culture cells are
used, an increase in the recovery of plasma membrane proteins without
sacrificing the purity of the membrane will be needed. Affinity matrices
provide a simple and quick method to use in the purification of membranes.
The conventional agarose- or acrylamide-affinity matrices cannot be used to
isolate membranes because they sediment with the organelles (e.g., nuclei)
that have relatively high densities. Chemically derived magnetic beads can
be coupled with various proteins, and they have become a new form of
affinity matrix. In contrast to conventional agarose- or acrylamide-based
matrices, magnetic beads can be conveniently separated from the mixture by
using a magnet and thus can be used to separate organelles independent of
their densities. By simply holding the tubes near the magnet, the magnetic
beads can be recovered at the sides of the tubes, allowing easy recovery of
the beads from the mixture. Thus, magnetic beads can be used as a substitute
for centrifugation. This property is likely to have great advantages in
separating plasma membranes from other organelles. We have recently
used magnetic beads with immobilized lectin for the purification of plasma
membrane proteins from cultured epithelial cells (Lee et al., 2008).
This procedure takes advantage of the fact that some of the membrane
proteins are glycosylated and can bind to lectins, the carbohydrate-binding
proteins. In this procedure, the lectin Concanavalin A (ConA) is immobi-
lized onto magnetic beads by binding biotinylated ConA to streptavidin
magnetic beads. The ConA-magnetic beads were mixed with homogenized
cell lysates, and the beads were recovered at the sides of the tubes, allowing
removal of other organelles that are not associated with the magnetic beads.
The membranes were found to be associated with the ConA-magnetic
beads as there was an enrichment in the activity of 50 -nucleotidase, a
membrane protein, from a total cell lysate of prostate PC-3 cells or cervical
HeLa cells. One caveat of this lectin magnetic bead method is that we
could not elute membranes from the ConA-magnetic beads by using the
competitive sugar alpha-methyl mannoside, possibly because the competing
sugar cannot access the binding sites between the plasma membranes and
ConA-magnetic beads. As a result, we used detergents to solubilize the
membrane proteins from the ConA-magnetic beads. The use of detergents
for solubilization of membrane proteins is described in Section 3.
high pH solutions (Schindler et al., 2006), for example 0.5 M NaCl. Because
detergents are not used in the procedure, the peripheral membrane proteins
can be purified by methods similar to those applied to soluble proteins.
Integral membrane proteins need to be solubilized from the lipid bilayer
to become individual proteins before purification. Detergents that possess
amphipathic properties are commonly used to solubilize integral membrane
proteins from membranes. Detergents may be grouped into three classes,
ionic, nonionic, and zwitterionic. A discussion of the different types of
detergents is described in Chapter 41 of this volume. Thus, we will only
discuss their use in terms of purification of membrane proteins.
Proteins are considered ‘‘solubilized’’ from the membrane if the proteins
appear in the supernatant fraction after the detergent treated membranes are
centrifuged at 105,000g for 1 h at 4 C. The process of solubilization of
membrane proteins by detergents can be divided into several phases. In the
first phase, the detergent binds to membranes. As the amount of detergent is
increased, the detergent starts to lyse membranes. Further increase in deter-
gent will lead to the formation of lipid/protein/detergent complexes.
At this stage, the membrane proteins are ‘‘solubilized.’’ Additional amount
of detergent will be needed to ‘‘delipidate’’ the complexes to protein/
detergent and lipid/detergent complexes. Usually, a detergent-to-protein
ratio of around 1–2 is sufficient to solubilize the membrane proteins into
lipid/protein/detergent complex. A detergent-to-protein ratio of around
10 or higher will lead to delipidation (Hjelmeland, 1990). The optimal
detergent to protein ratio that is required to solubilize a specific membrane
protein needs to be empirically determined.
The choice of detergents can be simply stated as the detergent that works
for your protein of interest. Nondenaturing detergents are those detergents
that solubilize the membrane proteins without significantly inactivating the
activity or function of the proteins. Among detergents that are available,
Triton X-100, sodium cholate, CHAPS, octylglucoside are ‘‘nondenatur-
ing’’ most of the time, although a loss of certain fraction of activity during
solubilization is expected.
The presence of detergents will affect several aspects of protein purifica-
tion. The detergents may affect the assay for the activity, for example
transport activity for membrane transporter and ligand-binding assay for a
receptor. As the proteins will no longer be associated with the membranes,
reconstitution of solubilized membrane proteins into phospholipid vesicles
will be needed for measuring transporter activity, and methods that separate
unbound ligand from the ligand–receptor complex will be needed to be
developed for the specific receptor. These requirements may limit the type
of detergents to be used for solubilization. For example, detergents with
high critical micelle concentration (sodium cholate, CHAPS, octylgluco-
side), which are easier to be removed by dialysis, should be used if a
subsequent reconstitution of proteins into phospholipid vesicles is planned.
624 Sue-Hwa Lin and Guido Guidotti
not introduce too much change to the proteins and can increase the
likelihood of expressing the membrane proteins in heterologous cell types.
Metal-affinity chromatography, which has high binding capacity and where
its binding is not affected by the presence of detergents, is a good choice for
expressing and purification of membrane proteins. Other affinity epitope
tags, for example, FLAG, myc, or HA, can also be used. However, purifi-
cation of proteins using immobilized antibody has limited capacity and the
cost is much higher than using metal-affinity chromatography.
Examples of expression and purification of membrane proteins can be
found in several articles. One example is the purification of a cell adhesion
molecule CEACAM1 expressed in insect cells (Phan et al., 2001). In this
study, the integral membrane protein was expressed with a 7-histidine tag at
its C-terminus. The original signal sequence was used. The recombinant
CEACAM1 protein was found to be localized on the plasma membrane.
The cells were lysed and the membranes solubilized by Triton X-100 and
the his-tagged CEACAM1 purified in one-step using metal-affinity
chromatography.
REFERENCES
Bradford, M. M. (1976). A rapid and sensitive method for the quantitation of microgram
quantities of protein utilizing the principle of protein-dye binding. Anal Biochem. 72,
248–254.
Hjelmeland, L. M. (1990). Solubilization of native membrane proteins. Methods Enzymol.
182, 253–264.
Lee, Y. C., Block, G., Chen, H., Folch-Puy, E., Foronjy, R., Jalili, R., Jendresen, C. B.,
Kimura, M., Kraft, E., Lindemose, S., Lu, J., McLain, T., Nutt, L., Ramon-Garcia, S.,
Smith, J., Spivak, A., Wang, M. L., Zanic, M., and Lin, S. H. (2008). One-step isolation
of plasma membrane proteins using magnetic beads with immobilized concanavalin A.
Protein Expr. Purif. 62, 223–229.
Neville, D. M. J. (1968). Isolation of an organ specific protein antigen from cell surface
membrane of rat liver. Biochim. Biophys. Acta 154, 540–552.
Peterson, G. L. (1977). A simplication of the protein assay method of Lowry et al. which is
more generally applicable. Anal. Biochem. 83, 346–356.
Phan, D., Han, E., Birrell, G., Bonnal, S., Duggan, L., Esumi, N., Gutstein, H., Li, R.,
Lopato, S., Manogaran, A., Pollak, E. S., Ray, A., et al. (2001). Purification and
characterization of human cell-cell adhesion molecule 1 (C-CAM1) expressed in insect
cells. Protein Expr. Purif. 21, 343–351.
Schindler, J., Jung, S., Niedner-Schatteburg, G., Friauf, E., and Nothwang, H. G. (2006).
Enrichment of integral membrane proteins from small amounts of brain tissue. J. Neural
Transm. 113, 995–1013.
Smith, P. K., Krohn, R. I., Hermanson, G. T., Mallia, A. K., Gartner, F. H.,
Provenzano, M. D., Fujimoto, E. K., Goeke, N. M., Olson, B. J., and Klenk, D. C.
(1985). Measurement of protein using bicinchoninic acid. Anal. Biochem. 150, 76–85.
C H A P T E R T H I R T Y- S I X
Purification of Recombinant
G-Protein-Coupled Receptors
Reinhard Grisshammer
Contents
1. Introduction 632
2. Solubilization: General Considerations 633
3. Purification: General Considerations 634
3.1. Stability of GPCRs in detergent solution 634
3.2. General affinity purification 634
3.3. Receptor-specific ligand affinity chromatography 636
3.4. Analysis of detergent-solubilized GPCRs 636
4. Solubilization and Purification of a Recombinant Neurotensin
Receptor NTS1 637
4.1. Solubilization of the NTS1 fusion protein 638
4.2. Purification of the NTS1 fusion protein by immobilized metal
affinity chromatography 640
4.3. Purification of the NTS1 fusion protein
by a neurotensin column 640
5. Analysis of Purified NTS1 641
6. Conclusions 642
Acknowledgments 642
References 642
Abstract
Structural and functional analysis of most G-protein-coupled receptors (GPCRs)
requires their expression and purification in functional form. The produced
amount of recombinant membrane-inserted receptors depends on the optimal
combination of a particular GPCR and production host; optimization of expres-
sion is still a matter of trial-and-error. Prior to purification, receptors must be
extracted from the membranes by use of detergent(s). The choice of an
appropriate detergent for solubilization and purification is crucial to maintain
receptors in their functional state. The initial enrichment can be carried out by
affinity chromatography using a general affinity tag (e.g., poly-histidine tag).
Department of Health and Human Services, National Institutes of Health, National Institute of Neurological
Disorders and Stroke, Rockville, Maryland, USA
631
632 Reinhard Grisshammer
If the first purification step does not yield pure receptor protein, purification to
homogeneity can often be achieved by use of a subsequent receptor-specific
ligand column. If suitable immobilized ligands are not available, size exclusion
chromatography or other techniques need to be applied. Many GPCRs become
unstable upon detergent extraction from lipid membranes, and measures for
stabilization are discussed. As an example, the purification of a functional
neurotensin receptor to homogeneity in milligram quantities is given below.
1. Introduction
Structure determination and functional analysis of integral membrane
proteins, which are not naturally abundant, require (i) a recombinant
production system and (ii) a purification strategy to allow the isolation of
functional rather than nonfunctional, incorrectly folded membrane protein.
Expression and purification of prokaryotic and eukaryotic membrane pro-
teins has been covered in the literature. The reader is referred, for example,
to Grisshammer and Tate (1995, 2003) and Grisshammer and Buchanan
(2006). In addition, the reader may consult Chapter 35 of this volume. This
chapter focuses exclusively on G-protein-coupled receptors (GPCRs)
which are eukaryotic integral membrane proteins involved in cell-to-cell
communication and sensory signal transduction (see Gether and Kobilka,
1998).
It is beyond the scope of this chapter to discuss in any great detail all the
possible expression strategies for integral membrane proteins such as GPCRs.
However, a few key points are summarized here. (i) A universal strategy for
the high-level recombinant expression of functional receptors is currently
unavailable. Some GPCRs accumulate in the membrane to high levels,
whereas other often closely related receptors are hardly detected. Despite
their assumed similarities, individual GPCRs behave quite differently in a
given expression host and recombinant production is still a matter of trial-and-
error. Comparative expression studies have, for example, been performed
using the methylotrophic yeast Pichia pastoris (Andre et al., 2006), the baculo-
virus insect cell system (Akermoun et al., 2005), and the Semliki Forest virus
system (Hassaine et al., 2006). A survey of GPCR production in commonly
used expression hosts has been summarized (Sarramegna et al., 2003). (ii) The
use of eukaryotic hosts seems generally better for producing functional,
membrane-inserted GPCRs than prokaryotic hosts (Grisshammer, 2006),
although there are exceptions. For example, the bacterium Escherichia coli
was successfully used for the expression of the neurotensin receptor NTS1
(Grisshammer et al., 1993; White et al., 2004), the M1 muscarinic acetylcho-
line receptor (Hulme and Curtis, 1998), the adenosine A2a receptor
(Weiß and Grisshammer, 2002), and the cannabinoid CB2 receptor
Recombinant GPCR Purification 633
(Calandra et al., 1997; Yeliseev et al., 2005). (iii) Numerous descriptive reports
have been published on the recombinant expression of integral membrane
proteins. However, few publications (see Bonander et al., 2005, 2009; Griffith
et al., 2003; Wagner et al., 2006, 2007) are currently available to understand
the underlying mechanisms of how a given host cell responds to membrane
protein overproduction.
The purification of receptors can be conceptually divided into two steps:
extraction from membranes with a suitable detergent (solubilization), and
subsequent purification by use of general affinity tags, receptor-specific ligand
columns, size exclusion chromatography, and other methods. Of utmost
importance, care must be taken to choose experimental conditions to main-
tain the membrane protein in its active state throughout the purification
procedure. This latter aspect cannot be overemphasized because many
GPCRs become unstable upon detergent extraction from lipid membranes.
comparison of that value with the theoretical value for specific binding of
pure functional receptor. If no test for functionality is available, then good
biochemical behavior such as a symmetric size exclusion chromatography
profile can be used as an indicator for integrity.
The preparation of membranes prior to solubilization constitutes an
initial purification step because soluble proteins are removed before
receptor extraction and the ratio of target receptor to contaminants is there-
fore higher. However, receptors can also be enriched from total cell lysate
rather than from solubilized membranes during the first purification step.
temperature (22 C). This avoids the possible overloading of the E. coli
translocation and membrane insertion machinery by the nascent receptor
chain. Few protein molecules are produced at any given time but accumu-
lation of correctly folded receptors is observed over 2 days. Purification of
10 mg of the NTS1 fusion protein typically requires 250 g of E. coli wet cells
which is equivalent to 50 l of cell culture (White et al., 2004). The growth of
E. coli at this scale is most easily done by fermentation.
The expression levels of the NTS1 fusion protein in E. coli are moderate
(i.e., the ratio of contaminants to target receptor is high). Hence, this requires
an optimized protocol for IMAC to enrich the receptor fusion protein
efficiently. To accomplish this, NTS1 fusion proteins with C-terminal tails
of 10 histidine residues rather than six histidine residues (Grisshammer and
Tucker, 1997) are used in combination with Ni2þ-NTA resin. The tight
binding to Ni2þ-NTA resin of decahistidine-tagged receptors allows strin-
gent washing steps using imidazole at a concentration of 50 mM. Imidazole at
this concentration eliminates binding to the Ni2þ-NTA resin of most of
the E. coli contaminants, but does not cause the elution of the fusion protein.
This strategy not only results in efficient purification of receptors from crude
membranes, but works well for the purification from total cell lysate
(Grisshammer and Tucker, 1997).
The apparent affinity of NT for NTS1 is reduced in the presence of the
high concentrations of sodium ions and imidazole in the NiB buffer
(Grisshammer et al., 1999) precluding the direct binding of functional
receptors in the Ni2þ-NTA column eluate onto the NT column. The
concentration of NaCl and imidazole must therefore be reduced from 200
to 70 mM by dilution with buffer to allow binding of functional NTS1 to
the NT column. Because binding of NTS1 to NT is NaCl sensitive,
receptors can be efficiently eluted from the NT column with NaCl at
high concentration (Fig. 36.1).
M 1 2 3 4 5
100
50
Figure 36.1 Purification of the neurotensin receptor NTS1. The fusion protein
NTS1-624 was purified on a 100-ml Ni2þ-NTA column followed by a 20-ml NT
column, starting from 250 g of E. coli cells. The progress of purification was monitored
by SDS–PAGE (NuPAGE 4–12% Bis–Tris gel, Invitrogen, 1 MES buffer) and
Coomassie R-250 staining. Lane M, Novagen Perfect Protein Marker (15–150 kDa);
lane 1, 10 mg of supernatant; lane 2, 10 mg of Ni2þ-NTA column flow through; lane 3,
5 mg of Ni2þ-NTA column eluate; lane 4, 10 mg of NT column flow through; lane 5,
5 mg of NT column eluate. The incorrectly folded receptors in the NT column flow
through are initially detergent soluble but aggregate over time (R. Grisshammer,
unpublished results). Reprinted from White et al. (2004).
6. Conclusions
The NTS1 receptor is expressed in functional form in E. coli as a fusion
protein. Receptors are solubilized by a detergent mixture from total cell
extract (rather than from a membrane preparation) and are enriched by
immobilized metal affinity chromatography. Contaminants present in the
Ni2þ-NTA column eluate are removed by the use of a subsequent neuro-
tensin column. The above purification protocol is simple and robust and
yields pure, functional NTS1 fusion protein.
ACKNOWLEDGMENTS
The research of R. G. is supported by the Intramural Research Program of the NIH,
National Institute of Neurological Disorders and Stroke.
REFERENCES
Akermoun, M., Koglin, M., Zvalova-Iooss, D., Folschweiller, N., Dowell, S. J., and
Gearing, K. L. (2005). Characterization of 16 human G protein-coupled receptors
expressed in baculovirus-infected insect cells. Protein Expr. Purif. 44, 65–74.
Andre, N., Cherouati, N., Prual, C., Steffan, T., Zeder-Lutz, G., Magnin, T., Pattus, F.,
Michel, H., Wagner, R., and Reinhart, C. (2006). Enhancing functional production of
G protein-coupled receptors in Pichia pastoris to levels required for structural studies via a
single expression screen. Protein Sci. 15, 1115–1126.
Bamber, L., Harding, M., Butler, P. J., and Kunji, E. R. (2006). Yeast mitochondrial ADP/
ATP carriers are monomeric in detergents. Proc. Natl. Acad. Sci. USA 103, 16224–16229.
(Epub ahead of print)Bonander, N., Darby, R. A., Grgic, L., Bora, N., Wen, J., Brogna, S.,
Poyner, D. R., O’Neill, M. A., and Bill, R. M. (2009). Altering the ribosomal subunit
ratio in yeast maximizes recombinant protein yield. Microb. Cell Fact 8, 10.
Bonander, N., Hedfalk, K., Larsson, C., Mostad, P., Chang, C., Gustafsson, L., and
Bill, R. M. (2005). Design of improved membrane protein production experiments:
Quantitation of the host response. Protein Sci. 14, 1729–1740.
Recombinant GPCR Purification 643
Calandra, B., Tucker, J., Shire, D., and Grisshammer, R. (1997). Expression in Escherichia
coli and characterisation of the human central CB1 and peripheral CB2 cannabinoid
receptors. Biotechnol. Lett. 19, 425–428.
Chelikani, P., Reeves, P. J., Rajbhandary, U. L., and Khorana, H. G. (2006). The synthesis
and high-level expression of a beta2-adrenergic receptor gene in a tetracycline-inducible
stable mammalian cell line. Protein Sci. 15, 1433–1440.
Cherezov, V., Rosenbaum, D. M., Hanson, M. A., Rasmussen, S. G., Thian, F. S.,
Kobilka, T. S., Choi, H. J., Kuhn, P., Weis, W. I., Kobilka, B. K., et al. (2007). High-
resolution crystal structure of an engineered human beta2-adrenergic G protein-coupled
receptor. Science 318, 1258–1265.
De Grip, W. J. (1982). Thermal stability of rhodopsin and opsin in some novel detergents.
Methods Enzymol. 81, 256–265.
Gether, U., and Kobilka, B. K. (1998). G Protein-coupled receptors II. Mechanism of
agonist activation. J. Biol. Chem. 273, 17979–17982.
Griffith, D. A., Delipala, C., Leadsham, J., Jarvis, S. M., and Oesterhelt, D. (2003). A novel
yeast expression system for the overproduction of quality-controlled membrane proteins.
FEBS Lett. 553, 45–50.
Grisshammer, R. (2006). Understanding recombinant expression of membrane proteins.
Curr. Opin. Biotechnol. 17, 337–340.
Grisshammer, R., and Buchanan, S. K. (eds.) (2006). ‘‘Structural Biology of Membrane
Proteins’’, The Royal Society of Chemistry, Cambridge, UK.
Grisshammer, R., and Tate, C. G. (1995). Overexpression of integral membrane proteins for
structural studies. Q. Rev. Biophys. 28, 315–422.
Grisshammer, R., and Tate, C. G. (eds.) (2003). ‘‘Special Issue on Overexpression of Integral
Membrane Proteins’’. Biochim. Biophys. Acta 1610, pp. 1–153.
Grisshammer, R., and Tucker, J. (1997). Quantitative evaluation of neurotensin receptor
purification by immobilized metal affinity chromatography. Protein Expr. Purif. 11, 53–60.
Grisshammer, R., Duckworth, R., and Henderson, R. (1993). Expression of a rat neuro-
tensin receptor in Escherichia coli. Biochem. J. 295, 571–576.
Grisshammer, R., Averbeck, P., and Sohal, A. K. (1999). Improved purification of a rat
neurotensin receptor expressed in Escherichia coli. Biochem. Soc. Trans. 27, 899–903.
Hanson, M. A., Cherezov, V., Griffith, M. T., Roth, C. B., Jaakola, V. P., Chien, E. Y.,
Velasquez, J., Kuhn, P., and Stevens, R. C. (2008). A specific cholesterol binding site is
established by the 2.8 Å structure of the human beta2-adrenergic receptor. Structure 16,
897–905.
Hassaine, G., Wagner, R., Kempf, J., Cherouati, N., Hassaine, N., Prual, C., Andre, N.,
Reinhart, C., Pattus, F., and Lundstrom, K. (2006). Semliki Forest virus vectors for
overexpression of 101 G protein-coupled receptors in mammalian host cells. Protein Expr.
Purif. 45, 343–351.
Hulme, E. C. (ed.) (1990). ‘‘Receptor Biochemistry—A Practical Approach’’, IRL Press at
Oxford University Press, Oxford.
Hulme, E. C., and Curtis, C. A. M. (1998). Purification of recombinant M1 muscarinic
acetylcholine receptor. Biochem. Soc. Trans. 26, S361.
Jaakola, V. P., Griffith, M. T., Hanson, M. A., Cherezov, V., Chien, E. Y., Lane, J. R.,
Ijzerman, A. P., and Stevens, R. C. (2008). The 2.6 angstrom crystal structure of a human
A2A adenosine receptor bound to an antagonist. Science 322, 1211–1217.
Klaassen, C. H., Bovee-Geurts, P. H., Decaluwe, G. L., and Degrip, W. J. (1999). Large-
scale production and purification of functional recombinant bovine rhodopsin with the
use of the baculovirus expression system. Biochem. J. 342, 293–300.
Kobilka, B. K. (1995). Amino and carboxyl terminal modifications to facilitate the produc-
tion and purification of a G protein-coupled receptor. Anal. Biochem. 231, 269–271.
644 Reinhard Grisshammer
Weiß, H. M., and Grisshammer, R. (2002). Purification and characterization of the human
adenosine A2a receptor functionally expressed in Escherichia coli. Eur. J. Biochem. 269,
82–92.
White, J. F., and Grisshammer, R. (2007). Automated large-scale purification of a recombi-
nant G-protein-coupled neurotensin receptor. Curr. Protoc. Protein Sci. 6.8.1–6.8.31.
White, J. F., Trinh, L. B., Shiloach, J., and Grisshammer, R. (2004). Automated large-scale
purification of a G-protein-coupled receptor for neurotensin. FEBS Lett. 564, 289–293.
White, J. F., Grodnitzky, J., Louis, J. M., Trinh, L. B., Shiloach, J., Gutierrez, J.,
Northup, J. K., and Grisshammer, R. (2007). Dimerization of the class A G protein-
coupled neurotensin receptor NTS1 alters G protein interaction. Proc. Natl. Acad. Sci.
USA 104, 12199–12204.
Yeliseev, A. A., Wong, K. K., Soubias, O., and Gawrisch, K. (2005). Expression of human
peripheral cannabinoid receptor for structural studies. Protein Sci. 14, 2638–2653.
C H A P T E R T H I R T Y- S E V E N
Contents
1. Introduction 648
2. Overview of Cell-Free Translation 650
3. Expression Vectors 651
4. Gene Cloning 652
4.1. Materials and reagents 654
5. PCR Product Cleanup 655
5.1. Materials and reagents 655
6. Flexi Vector and PCR Product Digestion Reaction 656
6.1. Reagents 656
7. Ligation Reaction 657
7.1. Reagents 657
8. Transformation Reaction 658
8.1. Materials and reagents 658
9. Purification of Plasmid DNA 659
9.1. Reagents 659
10. Preparation of mRNA 660
10.1. Reagents 660
11. Preparation of Liposomes 661
11.1. Materials and reagents 661
12. Wheat Germ Translation Reaction 662
12.1. Materials and reagents 663
13. Purification by Density Gradient Ultracentrifugation 665
13.1. Materials and reagents 665
14. Characterization of Proteoliposomes 667
15. Considerations for Scale-Up 669
16. Isotopic Labeling for Structural Studies 669
647
648 Michael A. Goren et al.
Abstract
Wheat germ cell-free translation is shown to be an effective method to produce
integral membrane proteins in the presence of unilamelar liposomes. In this
chapter, we describe the expression vectors, preparation of mRNA, two types of
cell-free translation reactions performed in the presence of liposomes, a simple
and highly efficient purification of intact proteoliposomes using density gradi-
ent ultracentrifugation, and some of the types of characterization studies that
are facilitated by this facile preparative approach. The in vitro transfer of newly
translated, membrane proteins into liposomes compatible with direct measure-
ments of their catalytic function is contrasted with existing approaches to
extract membrane proteins from biological membranes using detergents and
subsequently transfer them back to liposomes for functional studies.
1. Introduction
Membrane proteins provide the molecular mechanisms through
which useful molecules gain controlled entry into a cell, and likewise
provide the portals through which cellular products are exported from the
cell. Membrane enzymes synthesize the molecules that make up cellular
membranes (e.g., saturated and unsaturated fatty acids, phospholipids, gly-
cerolipids, sphingolipids, sterols, polyisoprenes). Membrane enzyme com-
plexes are responsible for electron transport and generate electrochemical
gradients, and an exquisite membrane motor ATPase uses these gradients to
generate ATP. They harvest light, provide allosteric receptors that trans-
duce the information from binding of external molecules into cellular
responses via signaling cascades, provide essential surface contacts as differ-
entiating embryonic stem cells begin to assemble into more complex tissues,
and help to elicit the antigenic responses observed in response to pathogen
infection. These few examples do not do justice to the incredible diversity
of functions and the essential roles that membrane proteins and enzymes
contribute to cellular function.
Achieving control of integral membrane protein expression, transfer into
the lipid bilayer, and cofactor incorporation are significant experimental
challenges, and the ability to manipulate these events would be of great
scientific utility. Furthermore, identification of novel ways to address
increasing structural complexity, leading to the expression and facile purifi-
cation of fully folded, functional membrane proteins or complexes
Cell-Free Translation 649
automated (Vinarov and Markley, 2005; Vinarov et al., 2006) in 24-, 96-,
and 384-well formats for screening in volumes as low as 50 mL, and 6- and
24-well formats for protein production in volumes up to 6 mL.
Prepare
plasmid DNA, Mix mRNA with
other reagents mRNA synthesis wheat germ extract
Protein
Translation production
Overlay 3h 6h 16 h
Figure 37.1 A schematic representation of the steps used for cell-free translation in a
bilayer mode reaction. A customized plasmid is used to prepare reagent mRNA, which
is added to the translation mixture, overlaid with amino acids, other substrates, and
desired additives, and then the reaction can begin. The photos show an experimental
demonstration of GFP translation over a 16 h period.
Cell-Free Translation 651
3. Expression Vectors
Table 37.1 summarizes specialized vectors used for this work. pEU,
the original cell-free translation vector optimized for wheat germ cell-free
translation (Madin et al., 2000), was modified for high-throughput Flexi
Vector cloning (Promega) to contain 50 -SgfI and 30 -PmeI restriction sites
and a toxic selection cassette in the multicloning site (Blommel et al., 2006).
The vector named pEU-FV produces a protein with no purification tag,
while pEU-His-FV and pEU-HSBC produces a protein with an N-terminal
His6 purification tag. Vectors pEU-NGFP and pEU-GFPC produce
fusions of GFP to either the N- or C-termini of the target protein. These
vectors are useful for control studies and as vehicles for simplified detection
of the translated fusion proteins during purification (Drew et al., 2001).
With proper design of the primer sequences, the GFP tag can be removed
by treatment with tobacco etch virus (TEV) protease (Blommel and Fox,
2007; Sobrado et al., 2008). The vectors pEU-Nb5R and pEU-Cb5 create
fusion proteins with the membrane anchor signals from human cytochrome
b5 reductase and human cytochrome b5, respectively. Fusion proteins con-
taining these tags spontaneously associate with liposomes upon translation
(Nomura et al., 2008), and thus become amenable to purification by density
gradient ultracentrifugation (Goren and Fox, 2008). Other posttranslational
modifications may also be used to target proteins to liposomes (Nosjean and
Roux, 2003). The vectors described herein are available from the NIH
Protein Structure Initiative Material Repository (http://psimr.asu.edu).
4. Gene Cloning
This approach requires a two-step PCR procedure (Blommel et al.,
2006; Thao et al., 2004). Figure 37.2 provides a vector map for pEU-HSCB
and an example of primers designed for cloning of His6-bacteriorhodopsin
(Blommel and Fox, 2007). By using this vector and primer design, the
sequence MGHHHHHHAIAENLYFQ- can be liberated from the trans-
lated protein upon treatment with TEV protease to leave Ser as the first
residue of the mature protein. The tobacco mosaic virus omega sequence is a
translation enhancement sequence (Sawasaki et al., 2000). Incorporation of
the SgfI and PmeI restriction sites in the indicated positions in the respective
primers allows the cloned gene to be transferred by Flexi Vector cloning
(Blommel et al., 2006) into any (or all) of the vectors described in Table 37.1,
along with a number of other commercially available vectors (see http://
www.promega.com for other examples). The sacB-CAT cassette shown in
Fig. 37.2A provides toxic selection when transformants containing this
construct are grown on plates containing 5% sucrose. In this manner, positive
selection for cloning the desired gene can be enforced. The insert also confers
resistance to chloramphenicol. The pF1K homology region enhances the
efficiency of transfer of cloned genes between different Flexi Vectors
(Blommel et al., 2006).
In the example shown in Fig. 37.2B, the first-step PCR uses a 50 forward
primer containing gene-specific nucleotides. An invariant sequence is added
to the 50 end of the first-step PCR forward primer to encode a portion of
the TEV protease site. The first-step PCR also uses a 30 reverse complement
primer as shown in Fig. 37.2C, which contains gene-specific nucleotides
and the PmeI site. Other genes may be cloned by substitution of the
gene-specific sequence in the designated primer sequences.
For the second-step PCR, a universal forward primer (Fig. 37.2B) is
used to add the nucleotides required to complete the TEV site and the SgfI
A
TMV omega Sgfl
200 400 600 800 1000 1200 1400 1600 1800 2000
sacB-CAT cassette
2200 2400 2600 2800 3000 3200 3400 3600 3800 4000 4200
sacB-CAT cassette
4400 4600 4800 5000 5200 5400 5600 5800 6000 6200
Amp resistance
B Sgfl C Pmel
GGTTgcgatgcgCGAAAACCTGTACTTCCAGTCCttggagttattgcc gcgaccagcgactgaTAGTTTAAACGAATTCGAGCTCACAC
CCAAcgctacgcGCTTTTGGACATGAAGGTCAGGaacctcaataacgg cgctggtcgctgactATCAAATTTGCTTAAGCTCGAGTGTG
1st PCR: 3⬘ TEV primer + gene-specific 1st PCR: gene specific + Pmel site
V A I A E N L Y F Q S L E L L P A T S D
Figure 37.2 (A) Linear map of pEU-HSCB showing important features of the vector construction. The TMV omega sequence (blue box)
enhances translation efficiency. The pF1K homology sequence (yellow box) prevents ligation of two plasmid backbones. (B) Example of the
50 primer designed for the cloning of Halobacterium salinarium bacteriorhodopsin (GenBank M11720.1). The first-step PCR forward primer is
654 Michael A. Goren et al.
site. A universal reverse PCR primer is used to duplicate the PmeI site and
add additional nucleotides (Fig. 37.2C). For the second-step PCR, a por-
tion of the first-step PCR reaction is added into a new PCR reaction
containing the universal forward primer and reverse primers to obtain the
PCR product properly functionalized for Flexi Vector cloning.
4.1.1. Protocol
The following steps are used to PCR amplify the desired genes and append
the sequences required for cloning. A sequence-verified gene in a plasmid is
used for the PCR template. These may be obtained from a variety of other
sources or research projects. This PCR protocol can also be used on
genomic DNA, if the gene of interest has no introns, or on reverse
transcribed mRNA (Thao et al., 2004).
If the plasmid being used for the template and the Flexi Vector that the
PCR product will be cloned into have the same antibiotic resistance, DpnI
can be added to the PCR reaction followed by incubation for 1 h at 37 C.
The DpnI will digest the template plasmid while leaving the PCR product
fully intact.
1. Create a PCR Primers plate by combining forward and reverse primers
for each target gene to 10 mM each in a well of an ISC PCR plate.
2. Create a PCR Master Mix consisting of 2.195 mL of water, 250 mL of
10 Pfu Ultra II Buffer, 25 mL of dNTPs (10 mM each), and 50 mL of Pfu
Ultra II Hotstart DNA polymerase This Master Mix will provide up to
96 reactions, and can be scaled as appropriate. Discard the unused mix.
3. Aliquot 23.0 mL of PCR Master Mix to each well of an ISC PCR plate
that will be used. This is the PCR Reaction plate.
4. Add 1 mL of the mixture from each used well of the PCR primers plate
to each used well of the PCR Reaction plate.
5. Add 1 mL ( 100 ng) of purified plasmid DNA for each gene to be
cloned to a separate well of the PCR Reaction plate.
6. Centrifuge the PCR Reaction plate briefly in an Allegra 6R centrifuge
and 6H3.B rotor to get liquid to the bottom of the wells and then cover
the plate with sealing tape.
7. Put the PCR Reaction plate in the thermocycler and cycle using the
following parameters: (1) 95 C for 2.00 min; (2) 94 C for 20 s; (3) 50 C
for 20 s; (4) 72 C, 15 s/kb; (5) repeat steps 1–4 for 19 more times.
8. For the second-step PCR, 20% of the volume of the first-step reaction is
added into a new PCR reaction along with 0.2 mM of each of the
universal forward and reverse primers. The second-step reaction is
then completed using the following cycling parameters: (1) 95 C for
2.00 min; (2) 94 C for 20 s; (3) 50 C for 20 s; (4) 72 C, 15 s/kb;
(5) repeat steps 2–4 for four more times; (6) 94 C for 20 s; (7) 55 C for
20 s; (8) 72 C for 15 s/kb; (9) repeat steps 6–8 for 24 more times; (10) 72
C for 3.00 min; and (11) 4 C and hold.
9. Analyze the completed PCR reactions for the appropriately sized
products using agarose gel electrophoresis.
5.1.1. Protocol
1. Add 4 mL of well-mixed SOPE resin to 20 mL of PCR product from the
FLEXI-TEV PCR protocol.
2. Cover the plate with Secure Seal Sterile tape and vortex. Let the
suspension stand at room temperature while preparing the Performa
Ultra 96-Well Plate.
3. Remove adhesive plate sealers from the top and bottom of the Performa
Ultra 96-Well Plate. Cover with a lid.
4. Stack the Performa Ultra 96-Well Plate on top of the 96-well
flat-bottom microplate.
5. Place the assembly in a cushioned centrifuge plate carrier.
6. Centrifuge for 5 min at 850g.
7. Briefly spin the SOPE/PCR mix to the bottom of the wells. Transfer
the SOPE/PCR reaction mixture by slowly pipetting directly into the
wells of the Performa Ultra 96-Well Plate. Be sure the fluid runs into the
gel matrix. Cover with a lid.
8. Stack the Performa Ultra 96-Well Plate on top of the 96-well V-bottom
microplate.
9. Place the assembly in the centrifuge plate carrier and centrifuge for 5 min
at 850g. Retain the eluates, which contain the purified PCR products.
PCR products can be stored at 20 C until ready for use.
6.1. Reagents
5 Flexi Digest Buffer (Promega)
10 SgfI/PmeI Enzyme Blend (Promega)
pEU vector variant (see Table 37.1)
Purified PCR products from PCR Product Cleanup protocol, step 9
Cell-Free Translation 657
6.1.1. Protocol
1. Create a pEU Vector Digest Master Mix consisting of 158.3 mL of
sterile, deionized water, 44.0 mL of 5 Flexi Digest Buffer, 2.20 mL of
10 SgfI/PmeI Enzyme Blend, and 13.5 mL of the pEU vector variant
desired (e.g., purified pEU-His-FV at a concentration of 150 ng/mL).
The enzyme blend is dense and tends to settle, so must be mixed
thoroughly into the remainder of the mix.
2. Place the pEU Vector Digest Master Mix in the thermocycler and cycle
as follows: (1) 37 C for 40.00 min; (2) 65 C for 20.00 min; and (3) hold
at 4 C until needed.
3. Create the PCR Product Digest Master Mix consisting of 638 mL of
sterile, deionized water, 220 mL 5 of Flexi Digest Buffer, and 22 mL of
10 SgfI/PmeI Enzyme Blend. This Master Mix will provide up to 96
reactions, and can be scaled as appropriate. Discard the unused mix.
4. Add 8.0 mL of the PCR Product Digest Master Mix to each well of an
ISC PCR plate. This is the PCR Digest plate.
5. Add 2.0 mL of the purified PCR products in the PCR Reaction plate from
Gene Cloning protocol, step 8 to each used well of the PCR Digest plate.
6. Place the PCR Digest plate in the thermocycler and cycle as follows:
(1) 37 C for 40.00 min; (2) 65 C for 20.00 min; and (3) hold at 4 C
until needed. If transformation yields no colonies, or colonies containing
vector without the desired insert, decrease the 37 C incubation time to
minimize star activity.
7. Ligation Reaction
Digested and purified PCR products and pEU vector variants are
ligated in this step. For the best efficiency in this ligation reaction, it is
important to use a high concentration of ligase.
7.1. Reagents
10 T4 DNA ligase buffer (Promega)
High concentration T4 DNA ligase (Promega)
pEU vector variant from Flexi Vector Digestion Reaction, step 2
Purified PCR product in PCR Cleanup plate from PCR Product Digestion
Reaction Cleanup, step 6
7.1.1. Protocol
1. Create a Ligation Master Mix containing 225 mL of sterile, deionized
water, 110 mL of 10 T4 ligase buffer, and 50 mL of high-concentration
T4 DNA ligase.
658 Michael A. Goren et al.
2. Add 5.0 mL of purified PCR product from the PCR Cleanup plate,
2.0 mL of the pEU vector digest, and 3.5 mL of Ligation Master Mix to
each well of a new PCR plate. This is the Ligation plate.
3. Incubate the Ligation plate in a thermocycler at 25 C for 3 h.
4. Immediately proceed to the transformation step or store the Ligation
plate overnight at 4 C.
8. Transformation Reaction
The material from the ligation reaction is used to transform competent
cells by the following steps. Competent cells from Invitrogen and Promega
have also been used successfully by following the manufacturers protocols.
8.1.1. Protocol
1. Remove 10G chemically competent cells from 80 C freezer and
thaw on ice.
2. Aliquot 10 mL of cells into a PCR plate or strip tubes that have been
chilled on ice.
3. Add 1.0 mL of the ligation reaction and stir with the pipet tip.
4. Incubate on ice for at least 5 min. Set a thermocycler block to 34 C.
5. Heat shock at 34 C for 30 s (per the manufacturer’s directions for small
reaction volumes).
6. Return the transformation reactions to ice and incubate for 2 min.
7. Remove from ice and add 80 mL of room temp recovery medium.
8. Incubate at 37 C for 1 h without shaking.
9. While transformations are incubating, label the bottoms of 96 YT
plates, containing the appropriate antibiotic, A1–H12 and add 5–10
sterile ColiRoller glass beads to each plate.
10. Apply the entire transformation reaction on each of the correspond-
ingly labeled plates. Shake the plates in a circular motion to spread the
Cell-Free Translation 659
9.1. Reagents
The 10 buffer used with proteinase K is 100 mM Tris–HCl, pH 8.0,
containing 50 mM EDTA and 1% (w/v) SDS.
Proteinase K is from Sigma/Aldrich (St. Louis, MO). Prepare a 10
proteinase K solution by addition of 0.5 mg of proteinase K to 1.0 mL
of 10 proteinase K buffer. Aliquot the proteinase K solution into 10 mL
samples and store them at 80 C.
9.1.1. Protocol
1. Inoculate 150 mL of 2 YT medium with a single colony from a pEU-
His-FV transformation and grow the culture in a shaking incubator
overnight at 37 C. Recover the cells by centrifugation.
2. Purify the plasmid using a Marligen high-purity plasmid Maxiprep kit
(Marligen Biosciences, Ijamsville, MD) and the manufacturer’s instruc-
tions. Resuspend each separate DNA pellet in 500 mL of Milli-Q water
and measure the absorbance at 260 nm to determine the plasmid DNA
concentration. A typical yield is 600–900 mg of plasmid DNA.
3. Plasmid DNA prepared by commercially available kits often contains a
trace contamination of RNase. This contaminant must be removed for
successful transcription and translation. In order to remove trace RNase
contamination, treat the purified plasmid with 50 ng/mL of proteinase K
for at least 60 min at 37 C in 1 proteinase K buffer.
660 Michael A. Goren et al.
10.1. Reagents
Transcription Buffer plus Mg (5) is 400 mM HEPES-KOH, pH 7.8,
containing 100 mM magnesium acetate, 10 mM spermidine hydrochlo-
ride, and 50 mM DTT. Store this buffer at 20 C.
An NTP solution containing 25 mM each of ATP, GTP, CTP, and UTP is
prepared from 0.2 mm filter-sterilized, 100 mM solutions of each NTP
prepared in Milli-Q water. NTP solutions are stored at 80 C.
SP6 RNA polymerase and RNase inhibitor (RNasin) are from Promega.
Cell-Free Translation 661
10.1.1. Protocol
1. Immediately before use, prepare a transcription mixture containing
2 Transcription Buffer plus Mg, 8 mM NTPs, 3.2 unit/mL of SP6
RNA polymerase, and 1.6 unit/mL of RNasin.
2. Dispense 2.5 mL of each separate plasmid DNA from Purification of
Plasmid DNA, step 8, to a separate well of a PCR plate. To the PCR
plate, dispense 2.5 mL of the transcription mixture into each well and
mix. This is called the Transcription plate.
3. Tightly cap the wells of the Transcription plate to avoid concentrating
the samples by evaporation. Incubate the Transcription plate at 37 C for
4 h. If the transcription reaction is proceeding correctly, a white precip-
itate of magnesium pyrophosphate will form and make the transcription
solution turbid.
4. Remove the white precipitate from the transcription reaction by centri-
fugation in the C0650 rotor and Allegra X-22R centrifuge for 5 min at
6230 rpm (4000g) and 26 C. To avoid coprecipitation of mRNA, the
reaction should not be chilled. Transfer the supernatant to a new tube.
This clarified solution will be used in the translation reactions as the
mRNA solution.
11.1.1. Protocol
1. Dissolve 1 g of soybean total extract (20% lecithin) in 3 mL of chloroform.
2. Flush the lipid solution with a stream of N2 gas in order to remove the
bulk of the organic solvent. Dry the remaining lipid further under
vacuum for 30 min.
662 Michael A. Goren et al.
the bilayer method because the constituents of the extract are not diluted by
diffusion, which occurs during the course of the translation in the bilayer
reaction. Thus the dialysis method can yield purified membrane proteins in
the range of 0.2 mg/mL to greater than 2 mg/mL in the standard reactions
(Goren and Fox, 2008). Although not as easily adaptable to robotics, the
dialysis reaction can be performed on the benchtop and is scalable from 50 mL
to 10 mL or greater with no overall changes in volumetric productivity. In
the smaller volumes, this method has utility for simple screening of a few
proteins, while in the larger volumes it also has utility for scale-up of the
production of proteins whose properties have been investigated at small scale
and found to be favorable for more extensive studies. Protocols for cell-free
translation in automated batch, bilayer, and dialysis reactions have been
published (Endo and Sawasaki, 2006; Sawasaki et al., 2002a; Vinarov et al.,
2006). For translation of integral membrane proteins, we have found that
translation reactions are productively modified by the addition of unilamellar
liposomes (Goren and Fox, 2008; Nozawa et al., 2007).
12 kDa molecular weight cutoff (MWCO) dialysis cups are from Cosmo
Bio (Tokyo, Japan). Prior to use, check the integrity of the membrane by
adding 500 mL of Milli-Q water and monitor for any leakage. If there is
no leakage, shake out the water prior to use.
The purified mRNA preparation is from Preparation of mRNA, step 4.
The liposome solution from Preparation of Liposomes, step 6.
24-Deep well pyramid-bottom plate, maximum volume of 10 mL (Artic-
white, Bethlehem, PA).
96-Well U-Bottom Plate (Greiner Bio-One, Monroe, NC).
5. Cover the 24-deep well plate with Saran Wrap to prevent evaporation
of the reservoir dialysis buffer. Incubate the translation reaction at 26 C
for 16 h.
6. Protein levels are determined by denaturing electrophoresis with creatine
kinase serving as an internal intensity standard.
250
1 2 3 4 5 6 7 8 9 10
150
100
75
50
37
25
20
15
10
Figure 37.5 Denaturing PAGE analysis of the density gradient obtained after cell-free
translation of hSCD1. The order of fractions from the density gradient is indicated.
hSCD1 (black star) and a plant Hsp70-like protein (white star) are noted in lane 2,
which represents the interface between the upper buffer (lane 1) and the remainder of
the 30% Accudenz layer. The majority of wheat germ extract proteins remain in the
40% Accudenz layer. The rightmost lane contains molecular weight markers (kDa,
noted alongside image). Adapted from Goren and Fox (2008).
13.1.1. Protocol
1. Carefully mix 75 mL of 80% (w/v) Accudenz solution with up to 75 mL
of the translation reaction, and place the mixed sample in the bottom of
an ultraclear centrifuge tube. This creates the 40% Accudenz layer.
2. Carefully layer 350 mL of 30% (w/v) Accudenz solution on top of the
mixture in the centrifuge tube. Minimize mixing of the gradient by
using a gel loading pipet tip to add the successive buffer layers.
3. Carefully layer 100 mL of lipid rehydration buffer on top of two other
layers in the centrifuge tube. This is the density gradient tube.
4. Spin the density gradient tube in an SW 50.1 rotor and L-60 ultracen-
trifuge for at least 4 h at 45,000 rpm (189,000g) and 4 C. The time
required for floatation is dependent on the properties of the proteolipo-
somes created, so optimization of density gradient and the centrifugation
time may be required to obtain the best separation for other proteins.
5. Carefully remove 60 mL fractions from the top to the bottom of the
gradient in order to fractionate the density gradient. Store the individual
Cell-Free Translation 667
Lane 1 2 3 4 5 6 7 8 9
18:0
18:1
Lane 1 2 3 4 5 6 7 8 9
hSCD1 (mM) 2.9 – 2.9 2.9 2.9 2.9 2.9 2.9 –
cytb5 (mM) – 5.8 5.8 5.8 5.8 5.8 5.8 5.8 –
Fe2+ (mM) 29 – – 29 29 29 29 29 29
Hemin (mM) – 5.8 5.8 – 2.9 5.8 11.6 5.8 5.8
% Conversion 3 – 2 33 54 43 17 42 –
Figure 37.6 Catalytic activity observed from various reconstitutions of the complex of
human stearoyl-CoA desaturase and human cytochrome b5 produced by cell-free transla-
tion and purified by density gradient centrifugation. The soluble domain of human
cytochrome b5 reductase was expressed in E. coli and added to these assays. The influence
of adding Fe2þ (required for hSCD1 activity) and hemin (required for cytb5 activity) is also
demonstrated. Lane 1, hSCD1 plus iron. Lane 2, cytb5 plus hemin. Lane 3, cotranslation of
hSCD1 and cytb5 supplemented with hemin. Lane 4, cotranslation of hSCD1 and cytb5
supplemented with Fe2þ. Lane 5, cotranslation of hSCD1 and cytb5 supplemented with
Fe2þ and 0.5 equiv. of hemin. Lane 6, cotranslation of hSCD1 and cytb5 supplemented
with Fe2þ and 1 equiv. of hemin. Lane 7, cotranslation of hSCD1 and cytb5 supplemented
with Fe2þ and 2 equiv. of hemin. Lane 8, mixture of separately translated and purified
hSCD1 with separately translated and purified cytb5 supplemented with Fe2þ and 1 equiv.
of hemin. Lane 5, cotranslation of hSCD1 and cytb5 supplemented with Fe2þ and 0.5
equiv. of hemin. Adapted from Goren and Fox (2008).
Wu, 2007; Wu et al., 2003). In some cases, the ability to perform protein
synthesis using radiolabeled AAs, unnatural AAs, or to specifically target one
or more AAs with diagnostic mass tags can advance these studies.
Further purification likely requires that the proteoliposomes be dissolved
with detergents, much as would be required for purification of a membrane
protein starting with a microsomal preparation from a living tissue. How-
ever, the combination of cell-free translation, liposomes, and density gradi-
ent centrifugation allows these additional purification efforts to start at a
much higher level of purity. Monitoring the decrease in light scattering
from the proteoliposome as detergent is added provides a simple diagnostic
for detergent optimization (Seddon et al., 2004; Womack et al., 1983). New
approaches using NMR, analytical ultracentrifugation, and gel filtration can
provide insight into the ability of a given detergent to yield monodisperse
protein-detergent micelles (Maslennikov et al., 2007; Slotboom et al.,
2008), which is generally considered to be an indicator of the compatibility
of protein and detergent. It is also possible to perform cell-free translation of
membrane proteins in the presence of detergents (Klammt et al., 2007;
Cell-Free Translation 669
Nozawa et al., 2007), with the constraints that the detergents used are
compatible with the cell-free translation reaction and with solubilization
of the nascent membrane protein.
17. Conclusions
The process simplifications afforded by the cell-free translation
approaches described herein provide significant labor and time advantages,
and this conclusion is emphasized when cell-free translation can produce
integral membrane proteins and enzymes that cannot be reasonably
obtained in a catalytically active state by any other method. By comparison,
other methods for the preparation of membrane proteins beginning with
expression in living hosts are time-, labor-, and material-intensive. Through
application of these methods, uniform samples of membrane proteins can be
easily obtained as the starting point for optimization of catalytic assays,
purification procedures, antibody production, structure determination,
and many other types of studies. This work provides an attractive alternative
to the sequence of expressing in a living host, extracting membrane proteins
with detergents, and then transferring them back into liposomes or other
lipid bilayers for functional studies. It is reasonable to assume that continued
application of this approach could open many new avenues to investigations
of membrane protein structure and function.
ACKNOWLEDGMENTS
This work was supported by NIGMS grant GM50853 to BGF, NIGMS Protein Structure
Initiative grant 1U54 GM074901 ( J. L. Markley, PI, G. N. Phillips, Jr., and B. G. Fox,
Co-Investigators) and NSF East Asia and Pacific Summer Institutes for U.S. Graduate
Students to MAG.
REFERENCES
Abe, M., Hori, H., Nakanishi, T., Arisaka, F., Ogasawara, T., Sawasaki, T., Kitamura, M.,
and Endo, Y. (2004). Application of cell-free translation systems to studies of cofactor
binding proteins. Nucleic Acids Symp. Ser. (Oxf.) 48, 143–144.
Blommel, P. G., and Fox, B. G. (2007). A combined approach to improving large-scale
production of tobacco etch virus protease. Protein Expr. Purif. 55, 53–68.
Blommel, P. G., Martin, P. A., Wrobel, R. L., Steffen, E., and Fox, B. G. (2006). High
efficiency single step production of expression plasmids from cDNA clones using the
Flexi Vector cloning system. Protein Expr. Purif. 47, 562–570.
Boyer, M. E., Stapleton, J. A., Kuchenreuther, J. M., Wang, C. W., and Swartz, J. R. (2008).
Cell-free synthesis and maturation of [FeFe] hydrogenases. Biotechnol. Bioeng. 99, 59–67.
CellFree Sciences, Ltd. Yokohama, Japan. http://www.cfsciences.com.
Chen, Y. J., Pornillos, O., Lieu, S., Ma, C., Chen, A. P., and Chang, G. (2007). X-ray structure
of EmrE supports dual topology model. Proc. Natl. Acad. Sci. USA 104, 18999–19004.
Drew, D. E., von Heijne, G., Nordlund, P., and de Gier, J. W. (2001). Green fluorescent
protein as an indicator to monitor membrane protein overexpression in Escherichia coli.
FEBS Lett. 507, 220–224.
Cell-Free Translation 671
Endo, Y., and Sawasaki, T. (2006). Cell-free expression systems for eukaryotic protein
production. Curr. Opin. Biotechnol. 17, 373–380.
Goerke, A. R., and Swartz, J. R. (2008). Development of cell-free protein synthesis plat-
forms for disulfide bonded proteins. Biotechnol. Bioeng. 99, 351–367.
Goren, M. A., and Fox, B. G. (2008). Wheat germ cell-free translation, purification, and
assembly of a functional human stearoyl-CoA desaturase complex. Protein Expr. Purif. 62,
171–178.
Hall, J. F., Ellis, M. J., Kigawa, T., Yabuki, T., Matsuda, T., Seki, E., Hasnain, S. S., and
Yokoyama, S. (2005). Towards the high-throughput expression of metalloproteins from
the Mycobacterium tuberculosis genome. J. Synchrotron Radiat. 12, 4–7.
Kanno, T., Kitano, M., Kato, R., Omori, A., Endo, Y., and Tozawa, Y. (2007). Sequence
specificity and efficiency of protein N-terminal methionine elimination in wheat-
embryo cell-free system. Protein Expr. Purif. 52, 59–65.
Kiga, D., Sakamoto, K., Kodama, K., Kigawa, T., Matsuda, T., Yabuki, T., Shirouzu, M.,
Harada, Y., Nakayama, H., Takio, K., Hasegawa, Y., Endo, Y., et al. (2002). An
engineered Escherichia coli tyrosyl-tRNA synthetase for site-specific incorporation of an
unnatural amino acid into proteins in eukaryotic translation and its application in a wheat
germ cell-free system. Proc. Natl. Acad. Sci. USA 99, 9715–9720.
Kigawa, T., Yabuki, T., Matsuda, N., Matsuda, T., Nakajima, R., Tanaka, A., and
Yokoyama, S. (2004). Preparation of Escherichia coli cell extract for highly productive
cell-free protein expression. J. Struct. Funct. Genomics 5, 63–68.
Klammt, C., Lohr, F., Schafer, B., Haase, W., Dotsch, V., Ruterjans, H., Glaubitz, C., and
Bernhard, F. (2004). High level cell-free expression and specific labeling of integral
membrane proteins. Eur. J. Biochem. 271, 568–580.
Klammt, C., Schwarz, D., Lohr, F., Schneider, B., Dotsch, V., and Bernhard, F. (2006).
Cell-free expression as an emerging technique for the large scale production of integral
membrane protein. FEBS J. 273, 4141–4153.
Klammt, C., Schwarz, D., Dotsch, V., and Bernhard, F. (2007). Cell-free production of
integral membrane proteins on a preparative scale. Methods Mol. Biol. 375, 57–78.
Madin, K., Sawasaki, T., Ogasawara, T., and Endo, Y. (2000). A highly efficient and robust
cell-free protein synthesis system prepared from wheat embryos: Plants apparently contain
a suicide system directed at ribosomes. Proc. Natl. Acad. Sci. USA 97, 559–564.
Makino, S.-I., Bingman, C. A., Berge, S., Larkin, A., Fox, B. G., Phillips, G. N. Jr, and
Markley, J. L. (2009). Crystal structure of agmatine iminohydrolase produced by wheat germ cell-
free translation (in preparation).
Maslennikov, I., Kefala, G., Johnson, C., Riek, R., Choe, S., and Kwiatkowski, W. (2007).
NMR spectroscopic and analytical ultracentrifuge analysis of membrane protein detergent
complexes. BMC Struct. Biol. 7, 74.
Matsuda, T., Koshiba, S., Tochio, N., Seki, E., Iwasaki, N., Yabuki, T., Inoue, M.,
Yokoyama, S., and Kigawa, T. (2007). Improving cell-free protein synthesis for stable-
isotope labeling. J. Biomol. NMR 37, 225–229.
Matsumoto, K., Tomikawa, C., Toyooka, T., Ochi, A., Takano, Y., Takayanagi, N.,
Abe, M., Endo, Y., and Hori, H. (2008). Production of yeast tRNA (m(7)G46) methyl-
transferase (Trm8-Trm82 complex) in a wheat germ cell-free translation system.
J. Biotechnol. 133, 453–460.
Morita, E. H., Shimizu, M., Ogasawara, T., Endo, Y., Tanaka, R., and Kohno, T. (2004).
A novel way of amino acid-specific assignment in 1H–15N HSQC spectra with a wheat
germ cell-free protein synthesis system. J. Biomol. NMR 30, 37–45.
Nomura, S. M., Kondoh, S., Asayama, W., Asada, A., Nishikawa, S., and Akiyoshi, K.
(2008). Direct preparation of giant proteo-liposomes by in vitro membrane protein
synthesis. J. Biotechnol. 133, 190–195.
672 Michael A. Goren et al.
Contents
1. Composition-Based and Activity-Based Analyses 680
1.1. Problems and pitfalls 681
2. Electrophoretic Methods 681
2.1. Methods 682
2.2. Pitfalls and problems 684
3. Chromatographic Methods 685
3.1. Gel filtration chromatography 685
3.2. Reversed phase HPLC 686
4. Sedimentation Velocity Methods 687
4.1. Method 687
5. Mass Spectrometry Methods 687
6. Light Scattering Methods 688
References 689
Abstract
There is no means for directly quantifying the purity of a protein sample.
Demonstration of purity of a protein preparation always involves an assessment
of the quantity of particular types of impurities, or simply demonstrating their
absence. Whether the goal is interpretation of analytical data, demonstration of
the quality of a process, or assuring safety of a biopharmaceutical product,
determination of sample purity is of central importance. This chapter will focus
primarily on methods for detecting protein impurities in protein samples and
will describe methods, their capabilities, and their limitations.
To assess the purity of a sample, one must first identify the type of
impurity that is to be measured (i.e., nucleic acid, carbohydrate, lipid,
unrelated proteins, isozymes, inactive enzyme), and then identify a
characteristic property (chemical assay or physical feature) which can distin-
guish the putative contaminant(s) in the presence of the protein of interest
in a specified solution condition. Purity is then the demonstration that the
677
678 David G. Rhodes and Thomas M. Laue
the protein or its solvent which might interfere with the use of the
techniques. It is easiest and most common to carry out one or more
fractionation procedures and demonstrate that only one component is
detectable. A wide variety of fractionation procedures may be used to
detect impurities, with the criterion for purity being the presence of a
single detectable component. Orthogonal methods are advisable to guard
against situations in which a chosen detection method is incapable of
resolving an unknown impurity from the primary analyte. Finally, it is
important to handle samples carefully, so that the process of preparing a
sample for analysis does not alter the impurity profile. Cleanliness, proper
temperature, and containers made of appropriate materials will all
contribute to informative analysis.
Many appropriate methods are detailed elsewhere in this volume and
will not be reviewed here. Several analytical methods for determining
molecular weight or molecular size are outlined elsewhere in this volume,
so where appropriate, this discussion will refer to details in that chapter
(Rhodes et al., 2009). In this chapter, emphasis will be on the variations of
those methods which are better suited to detection of contaminant than to
quantitative determination of size, mass, or other molecular parameters.
This chapter will also outline criteria for selecting among these methods,
some of their limitations and advantages, and some description of assays for
purity that do not involve fractionation.
1. Composition-Based and
Activity-Based Analyses
Methods that provide molar quantification of amino acids, specific
prosthetic groups, or of active sites may be used to assess the purity of a
sample. In cases where the specific activity of the pure material is known it
may be appropriate to determine the unit activity as a measure of relative
purity. These methods are secondary in nature, since reference is always
made to the quantity of the analyte that would be expected for a sample that
is free of contaminants. Moreover, these methods are appropriate primarily
with heterogeneous systems in early phases of purification or molecules like
membrane proteins which may require a specific environment to retain
function. Typically, two measurements are required: the mass of the protein
(or material) in the sample that will be used in the analysis, and quantifica-
tion of the known activity or of the particular analyte. A calculation is then
made of the expected quantity of analyte based on the mass of the material
used in the analysis. The purity is then expressed as ratio of the amount
measured to the amount expected. Good quantification has been achieved
using end-group analysis (Chang, 1983), and quantitative analysis for
Determination of Protein Purity 681
specific prosthetic groups, and for enzymatic activity (Biggs, 1976). The
details for this method are the same as those described in this volume
Rhodes et al. (2009).
2. Electrophoretic Methods
Because the electrophoretic methods provide some of the simplest,
least expensive, and often most sensitive approaches to determine the
number of protein components in a sample, they are the most commonly
used methods. Because of the exceptionally low cost and simplicity, these
methods are often the first screen for sample purity, even in early-stage,
highly heterogeneous samples. The methods of SDS gel electrophoresis
(see Friedman, 2009; Garfin, 2009) and electrophoretic determinations of
molecular weight and size are described elsewhere in this volume (Rhodes
et al., 2009). Any of these approaches could be used independently to assess
the purity of a sample, depending on which characteristic of the protein is to
be tested (Table 38.1). If the expected contaminants differ in molecular
weight from that of the desired protein, SDS gel electrophoresis would
resolve the impurity. Molecules of similar molecular weight but different
amino acid composition would generally not be distinguishable using SDS
gel electrophoresis, but could have different electrophoretic mobilities in
nondenaturing gel electrophoresis. Finally, proteins of almost any molecular
weight might be separable using isoelectric focusing techniques. The latter
method is detailed elsewhere in this volume (Friedman, 2009), and will not
be treated here.
Depending on the type of detection method available, nanogram to
microgram quantities of sample are required. Since a contaminant consti-
tutes a fixed weight fraction of a given sample, while the detection of a
contaminant depends on the total amount of contaminant, an overloaded
gel may provide a better chance of detecting a contaminant. However, an
upper limit to the quantity of sample that can be applied to a gel is set by
sample solubility and by considerations of resolution (Lunney et al., 1971).
The latter limit is usually more restrictive, with band broadening and
682 David G. Rhodes and Thomas M. Laue
2.1. Methods
The details of sample preparation depend on the nature of the electropho-
retic method chosen. The reader is referred to the other chapters in this
volume for discussion of native gel electrophoresis and isoelectric focusing
gels. Most frequently, SDS gels provide the ‘‘front-line’’ method for asses-
sing sample purity. Because purity assessment often is not quantitative, one
need not (necessarily) be as concerned about such problems as nonlinearity
in mobility at protein size extremes, aberrations in mobility due to protein
modifications, or nonuniformities in SDS binding that are discussed in
Rhodes et al. (2009). On the other hand, one does need to be concerned
with consistency and uniformity in sample handling and preparation, and
the conditions to which the sample is exposed during fractionation. For
example, a homogeneous protein preparation in which the reduction of
disulfide is not carried out carefully could appear to be heterogeneous.
Likewise, misinterpretation is possible if nonuniformity in a gel cross linking
results in mobility differences across the gel, but suitable replicates and
controls should minimize such issues.
Because one does not normally know the sizes of all protein components
within a sample, it is difficult to predict the gel concentration which will
give the optimal fractionation of a set of protein components. Gradient gels,
or gels of graded porosity, can cover a very wide range of molecular sizes.
Although the conditions of a gradient gel are unlikely to be optimal for a
particular fractionation, they will cover the widest possible range of
conditions, thus maximizing the probability of identifying a contaminant.
Ability to resolve a contaminant will depend largely on the range of gel
concentrations chosen for the gradient. The gradient should be designed
so that the protein of interest bands at an intermediate gel concentration.
The concentration of acrylamide at the top of the gel (low concentration)
Determination of Protein Purity 683
elsewhere in this volume and will not be repeated here. The wide range of
pI can be a considerable advantage, but the sensitivity of the technique to
small differences in isoelectric point can be enhanced by covering a
narrower range of pI values (e.g., 4–5, 5–8). Aside from the obvious
capability of differentiating molecules which differ in pI, this method
may be able to resolve contaminants which differ by properties other
than isoelectric point.
3. Chromatographic Methods
3.1. Gel filtration chromatography
Gel filtration chromatography is one of the simplest methods for the
detection of impurities that differ in size from the molecule of interest.
The method is nondestructive and rapid. Samples are diluted as they pass
through the column because this is a zonal method, so starting concentra-
tions well above the minimum detectable limit must be used. The exact
quantity of material needed will depend on the sensitivity of the assay used
to detect for contaminants.
3.1.1. Method
The method outlined in this volume for sample and column preparation for
the purposes of protein size determination should be used for assessing
impurities (Rhodes et al., 2009). The only difference is that the method of
detection must be sensitive to contaminants as well as to the protein of
interest. Although it is not necessary to calibrate the column to assess the
purity of a sample, it is useful to do so for two reasons. First, by using a
calibrated column, the experimenter gains some information concerning
the size of the impurity. Second, by using a calibrated column, both the test
for impurities and the molecular weight determination can be made simul-
taneously. To do this, one frequently uses two assays. The first is a nonspe-
cific assay for protein (e.g., absorbance at 280 nm and/or 220 nm) and the
second is a specific assay (enzymatic, immunological, mass-spectroscopy,
multiangle light scattering, etc.) to confirm the identity of the analyte
protein. Impurities are detected as either separate peaks in the chromato-
gram or as a broadening (e.g., a shoulder) of the elution profile. In principle,
the elution profile should be nearly Gaussian (with a very slight skewing of
the trailing edge, see Problems and Pitfalls). If the analyte protein peak is
strongly skewed, if the peak has a shoulder, or if the specific and nonspecific
assay results to not superimpose the sample has impurity present.
4.1. Method
Except for band, or zonal, sedimentation, details describing sample prepa-
ration, optical systems available, and the interpretation of the experimental
results are provided in Rhodes et al. (2009) of this volume and reviewed in
Cole et al. (2008). Consult the manufacturer’s instructions for details on
how to set up a preparative centrifuge to conduct a rate-zonal sedimentation
experiment. For zonal methods in the analytical ultracentrifuge, consult
Eason (1984) or Ralston (1993) for details. The analysis of the data is
analogous to that above describing gel-permeation chromatography.
Analysis using the differential sedimentation coefficient distribution, g(s),
(Cole et al., 2008) is a particularly useful approach to detection of impurity.
Because the method is model independent, the presence of multiple com-
ponents may appear as peaks at different values of s or as broadening of the
peak representing the primary analyte. Again, as is the case with gel filtration
chromatography, self-associating proteins may exhibit sedimentation
behavior with broad peaks in g(s) or multiple peaks, even though they are
pure. Analysis using equilibrium ultracentrifugation or another method can
help to resolve the problem.
REFERENCES
Biggs, R. (1976). Tests for fibrinolysis, thrombin clotting time test. In ‘‘Human blood
coagulation, hemostasis and thrombosis’’ (R. Biggs, ed.), 2nd edn. p. 722. Blackwell,
Oxford.
Chang, J. Y. (1983). Amino acid analysis in the picomole range by precolumn derivatization
and high-performance liquid chromatography. Methods Enzymol. 91, 41–48, this series.
Cole, J. L., Lary, J. W., Moody, T. P., and Laue, T. M. (2008). Analytical ultracentrifuga-
tion: Sedimentation velocity and sedimentation equilibrium. Method Cell Biol. 84,
143–179.
Eason, R. (1984). In ‘‘Centrifugation: A practical approach’’. (D. Richard, ed.), 2nd edn.
p. 251. IRL Press, Washington, DC.
Lunney, J., Chrambach, A., and Rodbard, D. (1971). Pore gradient electrophoresis. Anal.
Biochem. 40, 158–173.
Q6B (1998). ICH Q6B Specifications: Test procedures and acceptance criteria for biotech-
nological/biological products.
Ralston, G. (1993). Introduction to analytical ultracentrifugation Beckman Instruments,
Fullerton, CA.
Rhodes, D. G., Bossio, R. E., and Laue, T. M. (2009). Determination of size, molecular
weight, and presence of subunits. This volume.
C H A P T E R T H I R T Y- N I N E
Contents
1. Introduction 692
2. Chemical Methods 695
2.1. Composition 695
2.2. Colligative properties 697
3. Transport Methods 698
3.1. Sedimentation equilibrium 698
3.2. Sedimentation velocity 701
3.3. Gel-filtration (size-exclusion) chromatography 706
3.4. Electrophoresis 708
3.5. Viscosity 711
3.6. Field-flow fractionation 711
3.7. Mass spectrometry 712
4. Scattering Methods 716
4.1. Electron microscopy 718
5. Presence of Subunits 719
References 721
Abstract
The size or apparent molecular weight of a given protein may be the most cited
distinguishing characteristic of the molecule. In addition to being the basis of
many separation methods, the molecular weight, or simply molecular size,
immediately provides the investigator with an idea of the complexity of the
molecule, whether it is likely to be difficult to produce in quantity, and whether
certain analytical methods are likely to be productive. Knowing whether the
polypeptide of interest can self assemble or exists in a heterogeneous complex
with other polypeptides may provide valuable information regarding biosynthesis
691
692 David G. Rhodes et al.
or mechanism. This chapter outlines key methods used to determine the size of
proteins, their molecular weight, and whether subunits are present, with a focus
on the basis of the determinations, their strengths, and their limitations.
1. Introduction
The size of a protein molecule is the property at the basis of many
fractionation methods and is an easy-to-use descriptor for known molecular
entities or unknown impurities (e.g., ‘‘the 20-kDa product’’). Nevertheless,
in spite of significant advances in methodology and data-handling
techniques, care is required if one requires accurate estimates of molecular
size. In this chapter, ‘‘size’’ will refer to the physical dimensions of the
protein, as opposed to the molecular weight of the protein, which is related
to the mass of the protein. This chapter also considers the asymmetry
(or axial ratio) of proteins, as this property normally affects determinations
of the apparent size of the molecule.
Many proteins assemble into larger aggregates, with each constituent
chain being considered a subunit (Timasheff and Fasman, 1971). The
concept of a subunit, however, must be defined in context by the individual
investigator, taking into account the system under consideration and the
objectives of the work with the system. In this discussion, independent
subunits will be defined as those protein moieties which do not have a
contiguous polypeptide backbone. Thus, disulfide-linked peptide chains as
well as segments associated through noncovalent interactions are considered
subunits, even if those subunits may have arisen from proteolytic action on a
parent polypeptide.
The methods used to determine the size or molecular weight of a
protein can be categorized into three broadly defined areas: chemical analysis,
such as composition analysis or the effect of the molecule on the properties
(e.g., vapor pressure, freezing point, or boiling points) of a solvent; transport
methods, based on the movement of the molecule in response to some
applied force (e.g., electrical, centrifugal, mechanical); and scattering methods,
which are based upon the interaction of incident radiation (e.g., light,
X-rays, neutrons) with the molecule. As these methods are sensitive to
different features of the molecule, selection of appropriate methods depends
on what one needs to know about the system and with what accuracy. The
capabilities, advantages, and limitations of a number of techniques are out-
lined in Table 39.1. In addition to these criteria, the protein and the method
must be compatible with regard to the quantity of protein available, the
attainable level of purity (Rhodes and Laue, 2009), solvent requirements,
and any peculiarities of the protein that may interfere with a specific
technique. The various approaches are useful probes of other molecular
Table 39.1 Comparison of techniques for characterization of proteinsa
2. Chemical Methods
2.1. Composition
2.1.1. Overview
Although such methods are used only in unusual circumstances in current
practice, measurements that provide molar quantitation of either amino
acids and/or specific prosthetic groups may be used to estimate the mini-
mum molecular weight of a protein as Mmin ¼ m/n, where Mmin is the
molecular weight estimate (g/mol) of the protein, m is the mass of the
protein used in the analysis, and n is the number of moles of the protein-
related component measured in the analysis. Good quantitation has been
achieved using amino acids analyses, end-group analysis, and quantitative
analysis for specific prosthetic groups (Noble and Bailey, 2009). The quan-
tity of material required for such composition-based molecular weight
analyses is most dependent on the sensitivity of the analytical method
being used. Since many newer analytical techniques are sensitive in the
nanomole to picomole range, only microgram or even smaller quantities of
material may be required.
The error in the minimum molecular weight estimate will depend on the
errors incurred in both the mass estimate and in the analysis for the particular
constituent and, therefore, will depend on the methods chosen for these two
measurements. By combining data from analyses for several different consti-
tuents, the accuracy can be improved. For this reason, independent estimates
based on amino acid analyses, quantitative analysis of the end group, or
quantitation of nonprotein cofactors are recommended.
Because composition analysis provides only a minimum molecular
weight, it is very useful to combine these data with results from other
approaches to obtain the mass of the protein. An accurate composition
analysis can be combined with techniques that provide only low-accuracy
total molecular weights to yield high-accuracy results. Perhaps, the most
often used example of this approach is the combination of amino acid analysis
696 David G. Rhodes et al.
2.1.2. Method
An estimate of protein mass is made most accurately from a measurement of
dry weight when the protein is present in a volatile buffer (e.g., ammonium
bicarbonate). Compensation for nonvolatile buffer components can be
made from the difference in mass between dried samples of the protein
containing solution and the buffer alone, but is difficult to do with great
accuracy and should not be attempted with partially volatile buffers. Alter-
natively, protein concentration measurements are often used in place of dry
weight estimates (the mass value used for analysis simply being the product
of mass concentration and the volume). Concentration estimates can be
made refractometrically, spectrophotometrically, or by chemical assay. The
accuracy of such measurement often is limited, even though the precision
may be reasonable, because standardization is a major problem (Nozaki,
1986). This is especially true for glycoproteins, lipoproteins, or other
proteins with unusual compositions where the method of analysis may be
sensitive only to certain constituents of the protein (e.g., peptide bonds). Of
the methods listed, refractometry (or differential refractometry) is the most
accurate means of estimating concentration. Amino acid analysis also pro-
vides a suitable means of quantitation.
3. Transport Methods
3.1. Sedimentation equilibrium
3.1.1. Overview
Sedimentation equilibrium provides an accurate and powerful method for
the determination of the native molecular weight of a protein (Cole et al.,
2008). It is a simple, nondestructive, relatively rapid method. All parameters
that describe sedimentation equilibrium are either readily measured or easily
estimated. It has several unique capabilities, and can provide quantitative
estimates of molecular weights, stoichiometries, and association constants
for a wide range of chemical systems. For example, by sedimenting in
neutrally buoyant (nonsedimenting) detergents, the native molecular
weight of detergent-solubilized proteins may be measured (Fish, 1978).
Many systems not amenable to analysis by any other technique can be
evaluated successfully by sedimentation equilibrium. However, with the
advent of powerful sedimentation velocity analysis, because sedimentation
equilibrium generally requires sophisticated, expensive equipment, and
because other methods are available which provide data that are adequate
for many purposes, this procedure is used much less frequently than in the
past. Nevertheless, equilibrium sedimentation provides one of the most
powerful methods for the quantitative examination of protein–protein
associations, and is irreplaceable for the study of protein systems with
weak to moderate association constants (Cole et al., 2008).
3.1.2. Method
In an equilibrium sedimentation experiment, the purpose is to produce a
measurable protein concentration gradient along the radial axis of the
centrifuge cell. The length of time needed to achieve sedimentation equi-
librium depends on the length of the solution column, the value of the
reduced molecular weight, s, and the diffusion constant. Because the time
to reach equilibrium depends on the square of the column length, many
cells and techniques have been developed to examine short columns (0.75
or 3 mm long) (Ralston, 1993). Using these cells, equilibrium can be
reached in a matter of minutes to a few hours, depending on the protein.
At each rotor speed, the concentration distribution is measured at intervals
of 30–60 min and invariance of the distribution over time indicates that
equilibrium has been reached. Software (e.g., HeteroAnalysis) helps auto-
mate this process (Cole et al., 2008).
The reduced molecular weight, s, is the measured quantity in a
sedimentation equilibrium experiment (Yphantis, 1964). It is defined as:
Size, Molecular Weight, and Subunits 699
Mð1 vrÞo2
s¼ ; ð39:2Þ
RT
where v (ml/g) is the partial specific volume of the protein, r (g/ml) is the
solution density, and o2 (s 2) is the square of the angular velocity (o ¼
rpm p/30) of the rotor. (Methods for v and r are described below,
following the description of velocity sedimentation). It can be shown that
dln½cr s
¼ ; ð39:3Þ
dr 2 2
where cr is the solute concentration at radial position in the rotor, ‘‘r.’’ Thus,
s can be determined as the slope of a graph of the natural logarithm of the
concentration as a function of r2/2. The slope may be calculated at each of
several radial positions (or, correspondingly, at each of several concentra-
tions). Molecular weights determined in this fashion are weight-average
values.
Alternatively, s may be obtained from nonlinear least-squares fitting to
equations that describe the concentration distribution (i.e., c as a function of
r2/2) that are beyond the scope of this chapter ( Johnson et al., 1981). These
equations have been derived from thermodynamic first principles for a
variety of models that include association and nonideality of the proteins.
A typical equilibrium centrifugation experiment will require approxi-
mately 100–200 ml of solution at each of three or four concentrations,
which typically vary from 0.1 mg/ml to approximately 3 mg/ml. These
solutions must be matched with appropriate buffers of the same composi-
tion. The equilibrium centrifugation experiment is normally run at different
speeds, from a low speed where the concentration ratio from the top of the
cell to the bottom of the cell is about 3–5, to a high speed where the
meniscus is depleted of solute.
The concentration distribution at equilibrium can be determined by any
number of means, without an analytical centrifuge. If a preparative centri-
fuge (Attri and Minton, 1986) or an ‘‘airfuge’’ (Bock and Halvorson, 1983)
is being used for the measurement, any assay that is proportional to the
concentration of protein may be used following fractionation of the content
of the cell. Because any method for concentration determination may be
used (e.g., enzyme activity), this sort of analysis may be used to determine
molecular weights in complex mixtures and concentrated solutions
(Howlett et al., 2006). However, it is then necessary to perform several
experiments for different lengths of time to ensure that equilibrium has been
reached. These techniques suffer a loss of precision due to the collapse of the
concentration gradient that occurs as the rotor decelerates and as the cell
contents are fractionated. The analytical ultracentrifuge alleviates these
problems by permitting the solution contents to be examined optically as
700 David G. Rhodes et al.
6p
f ¼ 6prsphere ¼ ; ð39:7Þ
v3Þ=ð4pN0 Þ1=3
½ðM
where is the viscosity of the solution and rsphere is the radius of the
equivalent sphere. Using Eqs. (39.5)–(39.7), the ratio of f/f may be
calculated and used to estimate the shape of the protein. This ratio, f/f , is
compared to standard and theoretical values to estimate the asymmetry and
overall shape of the protein.
With the advent of powerful computer programs to analyze the data,
sedimentation velocity has become the more widely used than equilibrium
sedimentation (Cole et al., 2008). It is recommended that velocity analysis
be performed on any sample prior to sedimentation equilibrium analysis.
3.2.2. Method
Sedimentation velocity experiments are best performed in an analytical
ultracentrifuge, although methods of reasonable accuracy have been devised
for preparative centrifuges. The advantage of the analytical machine is that
its accuracy (1–3%) is nearly an order of magnitude better than can be
expected from techniques using preparative instruments. Moreover, data
are acquired throughout the experiment so that any unusual behavior can be
observed and better understood. The advantage of the preparative machines
is that sensitive or specific assays, such as ELISA, may be used so that in cases
where the purity of the sample is in doubt, a specific assay can identify the
component of interest. Detailed methods for using the analytical ultracen-
trifuge (Ralston, 1993) and for using preparative centrifuges (Freifelder,
1973; Martin and Ames, 1961) are available.
As noted before, the principal measurement in any sedimentation exper-
iment is the concentration as a function of radial position. The sedimenta-
tion coefficient may be determined from the slope of a graph of the natural
logarithm of the distance that the molecules have sedimented as a function
of time (d ln r/dt):
ðd ln r=dtÞ
s¼ ; ð39:8Þ
o2
where r is the distance of the boundary of the molecules from the center of
the rotor and t is the time from the start of the experiment, usually taken to
be the time at which the rotor reaches two-thirds of the final speed. For
experiments in which the sample being examined is a thin zone of mole-
cules, r is taken to be the point of maximum concentration. For broad zone
or boundary experiments, r is taken to be the midpoint between zero
concentration and the plateau concentration. For proper interpretation
and comparison, the sedimentation coefficient must be corrected to
Size, Molecular Weight, and Subunits 703
Of all the techniques described in this chapter, the problems and pitfalls of
sedimentation analysis are the best documented and most easily overcome.
The most demanding aspect of sedimentation is the availability of
sufficient material for analysis. If the standard optical systems on the current
(Beckman XL-A) analytical ultracentrifuge are used, 0.5 ml of solution with
protein concentrations in the range of 0.1–1 mg/ml is needed to obtain
good data.
For proteins that bind significant levels of buffer components
(e.g., detergent-solubilized proteins), the interpretation of sedimentation
coefficients is made difficult by the fact that the bound components will
contribute to the measured s in four terms: M, v, r, and f. Of these terms,
M, v, and f are usually the most affected by bound components, and unlike
sedimentation equilibrium, there is no way to ‘‘blank out’’ the contribution
of such components. Thus, the measured sedimentation coefficient is for
the complex of the protein with the bound component, making it difficult
to extract any useful information concerning the protein alone.
3.2.4. Variations
The effect of protein concentration on s is assessed by determining s at each
of several protein concentrations. Typically, there is a slight linear decrease
in s as the protein concentration is increased.
Another widely used method of estimating molecular weights of pro-
teins is the sucrose gradient sedimentation method introduced by Martin
and Ames (1961). In this method, one creates a linear gradient of sucrose in
buffer in a swinging bucket centrifuge tube. The sucrose concentration
typically ranges from approximately 5% at the top of the tube to 20% at the
bottom; the actual range is less important than the linearity and
reproducibility of the gradient. The unknown sample is layered onto a
gradient and a set of standard proteins of known molecular weight layered
onto an equivalent gradient. The centrifugation proceeds for an appropriate
interval (typically 12–24 h) and the material on the gradients collected as
fractions. The protein concentration in these fractions is determined by
spectrophotometric, enzymatic, or other assays.
The basis of the method is the fact that in a linear sucrose gradient, the
distance traveled by a molecule should be a linear function of the time of
centrifugation at a specified speed. In addition, the distance will depend
linearly on s. The ratio of the distance traveled by an unknown protein to
that of a standard will be equal to the ratio of their sedimentation
coefficients, which will, in turn be approximated by the ratio of the
molecular weights to the 2/3 power. This method yields only approximate
values of s and M, but is simple, requires no specialized equipment, and can
be used to estimate s and M for very small amounts of material if a suitable
assay is available.
706 David G. Rhodes et al.
3.3.2. Method
The principle of operation and selection of gel media and gel porosity is
described in detail elsewhere (Stellwagen, 2009). SEC stationary phase
materials are available which will fractionate a very wide range of molecular
Size, Molecular Weight, and Subunits 707
sizes; these may not provide adequate resolution for specific needs. In some
cases, a stationary phase with a narrower range may provide superior
resolution, but one must choose a gel in which the protein to be examined
is partially included. Normally, choice of the stationary phase is arbitrary, as
long as the protein does not bind to the matrix. When there is a choice of
bead sizes for a given porosity, the smallest bead size should be used, as this
improves the column resolution. Check the manufacturer’s recommenda-
tions for any limitations on solvents, but in general, nearly any free-flowing
aqueous buffer system may be used. It is recommended that buffers of
moderate ionic strength be used so that electrostatic interactions between
the protein and immobile charges on the matrix are minimized. For best
results and highest resolution, a long, narrow column should be used.
Preparation of the column and equilibration of the column by washing
with buffer should be done according to the manufacturer’s specifications.
Likewise, flow rates should be chosen in accordance with the manufac-
turer’s specifications. In general, lower flow rates afford better resolution
because the solute can fully equilibrate with the gel matrix at all times, but
excessive diffusion can limit resolution if a column is run too slowly. All
samples should be in the same buffer as that used to equilibrate the column.
Sample volumes applied to the column should generally be less than 2% of
the column’s bed volume. In addition, the flow rate of the column should
be kept constant throughout all of the analyses, since flow rate dependence
of the elution volume can be expected (Ackers, 1970).
Elution volume (Ve) is the volume eluted from the column, starting
once one-half of the sample has penetrated the top of the gel bed and
continuing until the maximum (peak) of the protein of interest has eluted.
The void volume (V0) of the column usually can be measured using
commercially available, size-graded Blue Dextran (M ¼ 2,000,000), and
monitoring the effluent spectrophotometrically at 540 or 280 nm. The
included volume can be measured using as a sample some buffer of different
pH or conductivity (avoiding extremes that could affect the column) or one
that contains a small dye (e.g., bromphenol blue). The included volume is
the difference between the elution volume for the small molecule and the
void volume measured with a large, excluded molecule. Care should be
exercised in the choice of a dye, as many aromatic compounds have an
affinity for the stationary phase which results in anomalously high Kav and
Vi values.
Protein standards should be used that span the full range of sizes that can
be analyzed by the stationary phase chosen to assure that the standards
bracket the analyte protein. For best results, a minimum of four different
standards should be used. Kits containing prestained proteins are commer-
cially available. Any convenient assay for detecting the presence of the
protein being analyzed may be used. If, for any reason, the column must
be repacked, the calibration must be performed again.
708 David G. Rhodes et al.
3.4. Electrophoresis
3.4.1. Overview
The most widely used method of evaluating the size of a protein molecule is
electrophoresis. The method is simple, inexpensive, rapid, and for a very
wide range of proteins, is sufficiently accurate for many needs. For these
Size, Molecular Weight, and Subunits 709
reasons, it is the method of choice for most protein systems, and almost
always included in characterization studies. SDS gel electrophoresis is the
most widely used method for determining apparent molecular weights of
denatured proteins (Garfin, 2009), but electrophoretic methods for obtain-
ing size, shape, and molecular weight information are not limited to this
approach. Despite its popularity, it is not necessary to include SDS in the gel
formation; native gel electrophoresis of protein samples may be carried out
under almost any buffer condition required. In addition, the sensitivity of
current staining procedures allows these approaches to be applied to very
small amounts of protein (Garfin, 2009).
3.4.2. Method
The basic procedures are identical to those described earlier (Rhodes and
Laue, 2009; Garfin, 2009) for electrophoresis of denatured proteins, except
that the buffer composition is determined by the needs of the investigator.
There are some restrictions; as with gels under denaturing conditions, the
buffer in which the gel is cast cannot contain reducing agents. In addition,
because the conditions to which the protein is exposed may affect charge,
association, or shape, the composition of the buffer in the gel must be
controlled carefully. One must, therefore, be especially cautious about
relative proportions of catalyst, so that excess oxidant is not left in the gel.
Precast gels are inexpensive and generally have good quality control, but the
probability is low that the buffer in the precast gel will be precisely the
condition that is needed. It is often easiest to prerun gels to remove any
possible by-products of polymerization. Alternatively, if the geometry of
the gel apparatus allows the slab to be exposed (e.g., a horizontal slab or a
vertical slab in which one of two glass plates can be removed and later
replaced), the gel can be dialyzed against the buffer of choice prior to
running. Because slabs are generally quite thin, dialysis for several hours is
generally sufficient. This dialysis procedure can also be used to introduce
reducing reagents postpolymerization, or to use a set of gels cast together
(and thus, presumably, uniform in porosity) with a variety of buffer
conditions.
The basis of electrophoretic protein size analysis is based on a simple
principle: that a charged particle in an electric field is forced through the
surrounding medium by a force proportional to the charge on the particle
and the strength of the field, and is subject to a frictional force proportional
to the velocity, the radius of the particle, and the viscosity of the medium.
As with SDS gel electrophoresis, the investigator may control the frictional
coefficient by controlling the porosity of the gel matrix. In addition, under
nondenaturing conditions, the mobility can be significantly affected by
alterations of the intrinsic charge on the protein due to changes of pH at
which the electrophoresis is carried out. This distinction is important. In
SDS gel electrophoresis, the charge is dominated by the negatively charged
710 David G. Rhodes et al.
SDS associated with the protein so the sample is applied to the cathode end
of the gel and the sample always moves toward the anode. In nondenaturing
electrophoresis, the direction of migration will depend on the buffer pH
relative to the isoelectric point (pI) of the protein. If the pI of the protein is
unknown (and quantities are too limited to allow experimental determina-
tion with isoelectric focusing), a horizontal electrophoresis apparatus may
be used, and the starting wells placed in the center of the gel. Otherwise, a
‘‘best guess’’ might be made by using the pI of related proteins.
The mobility will also be affected by the chosen buffer condition.
Beyond association of the protein or conformational changes associated
with changing buffer conditions, the relative mobilities of various proteins
at a given pH will usually be fairly constant. The absolute mobility, how-
ever, will be strongly affected by the concentrations of counterions.
The analysis of native gel electrophoresis mobility data is analogous to
that of SDS gels. The only significant difference is that, in the native system,
one cannot assume that the proteins have equal charge density and are all in
the shape of long rods. As is necessary in gel-exclusion chromatography,
one must compare the mobility of the sample to that of a set of standards
(Rodbard and Chrambach, 1971). Ferguson analysis can also be used to
identify the feature(s) (size and/or charge) that distinguish two components
and to extrapolate to the mobility expected in the absence of sieving effects
(Andrews, 1986). The slope of the Ferguson plot (log of the relative
mobility, Rf, vs. gel concentration) is proportional to rs, the Stokes radius.
Interpolative estimation of an unknown rs is more reliable using this relation
than using, for example, the slope of the Ferguson plot with molecular
weights of the standards. The analysis simply involves running the unknown
and several standards in a set (at least five) of gels of different total gel
concentrations, and plotting log Rf versus the gel concentration for all
standards and for the unknown. The slopes for the standards are then plotted
as a function of the (known) rs, and the rs of the unknown is derived
from this plot and the measured slope. The range of gel concentrations
used will depend on the size of the protein under study, but at some upper
limit of concentration, the protein will be excluded from the gel. A more
closely spaced group of gel concentrations covering a lower range should
then be used.
3.5. Viscosity
Among the variables upon which the viscosity of a solution depends (T, P,
etc.) are the amount and nature of any solute that is present. The response of
a particle in a fluid under shear will depend on the frictional coefficient of
the particle, which represents the mechanical interaction of the particle with
the moving solution (Yang, 1961). The response of the particle will also
depend on the mass of the particle, which determines the energy required to
attain a given movement ( Johnson et al., 1977). Because the intrinsic
viscosity of a protein depends very strongly on the asymmetry of the
molecule, viscosity measurements are sensitive indicators of protein shape.
In addition, it is possible to combine viscosity data with sedimentation data
to calculate the molecular weight of a protein. Apparatus for rheological
measurements vary widely; their use is generally quite simple, but must be
performed under rigorously controlled conditions.
semipermeable membrane allows solvent, but not protein, to flow out the
bottom of the channel, thus creating a concentration gradient with higher
concentration at the bottom of the channel than at the center or top.
Diffusion works against this concentration gradient and allows the more
rapidly diffusing small molecules to reach the center of the channel where
flow is faster. The larger molecules cannot diffuse as rapidly and move
slowly at the edge of the channel.
To its benefit, this method is nondestructive and can be used with any
buffer system. By appropriate adjustment of the flow rates, this method is
capable of resolving a very wide range of molecular sizes. Current instru-
mentation utilizes many HPLC components and highly automated systems
make the process accessible; a typical experiment runs in approximately
30 min. Nevertheless, the method is technically challenging and will require
time and effort for method development. Finally, the method is very good
for resolving molecules of different sizes (e.g., monomer, dimer, etc.), but
the resulting size distribution is relative and not easily converted to an
absolute size. Consequently, FFF often is used as a fractionation method
for light scattering to provide absolute molecular weights (Liu et al., 2006).
3.7.1. Method
The specific ESI-MS techniques will vary with the individual instrumenta-
tion, but some general guidelines are provided here. As mentioned above,
samples must be in a volatile buffer if a direct presentation to the mass
spectrometer is going to be used. Phosphate- and TRIS-related buffers
cause ion suppression and mass shifts due to noncovalent adduct formation
714 David G. Rhodes et al.
between sodium, potassium, counter ions, and the protein. These salts can
be dialyzed against a volatile buffer or removed chromatographically. If the
sample is analyzed by LC/MS, the small buffer constituents will be fractio-
nated from the larger protein and dialysis is not necessary. In cases where a
protein is added to the column in a salt-based buffer, a diverter valve is used
to divert the salt peak to waste before the salts arrive at the ESI source.
Samples available as a lyophilized powder can be dissolved in a mixture of
methanol or acetonitrile and water with formic or acetic acid, usually at a
ratio of 1:1 aqueous:organic, with the acid present at 5% or less (typically
0.1%). Higher organic concentrations can be used without difficulty for
hydrophobic proteins.
In addition to using the volatile buffers suggested above, it is important
to assure that the protein is fully equilibrated to this condition and that any
titration of the buffer is with suitable acid or base (ammonia, acetic, formic,
or trifluoroacetic). Moreover, for some of these buffers, the concentration
limits are relatively low. Flow rates will be determined by the specific
instrument limits, but in some cases, a splitter will be required so that only
a small portion of the column flow enters the MS. Most LC/MS instru-
ments are capable of operating without a splitter at flow rates as high as
1 ml/min. ESI sources are also available with very small ( 25 micrometer
inner diameter) capillaries, called nano-ESI. These ESI sources operate at
much lower flow rates ( 10–100 nl/min) and are frequently coupled to
capillary LC columns for Capillary LC/nano-ESI-MS. Although the sensi-
tivity is improved, these systems can become clogged without scrupulous
handling of solvents and samples to avoid dust.
The speed of a protein mass spectral analysis is very rapid when analyzing
a single sample (<1 min), and can range up to 1 h for an LC/MS analysis.
New ultrahigh-pressure liquid chromatography analysis (UPLC) is showing
promise to reduce these analysis times and improve the chromatographic
resolution.
Four recently developed ionization modes have allowed for alternate
modes of ionization for large proteins. Desorption electrospray ionization
(DESI), Direct Analysis in Real Time (DART), Laser Ablation Electrospray
Ionization (LAESI) and matrix assisted laser desorption electrospray ioniza-
tion (MALDESI) have shown the ability to ionize proteins, and may prove
useful in cases where ESI and MALDI (matrix-assisted laser desorption
ionization) are ineffective (Faull et al., 2008). MALDESI combines the
advantages of MALDI, such as the ability to handle high-molecular-weight
proteins, with the multiple charging capability of ESI. These ionization
methods may be able to overcome some of the limitations encountered in
the individual methods.
On the mass spectrometer side, an alternate platform is finding new life
in protein analysis, ion mobility spectrometry (IMS). IMS separates ions
according to their mobility through an electric field as they collide with a
Size, Molecular Weight, and Subunits 715
gas (usually He). Protein ions with a large shape/collision cross section
travel slower in the electric field than smaller ions with lower collisional
cross section. From the cross section information, aggregation information
can be determined as well as purity. In some aspects, IMS resembles TOF
but reveals information about protein purity through different measures (the
gas phase ion shape as opposed to its mass). IMS can also be coupled to mass
spectrometers, allowing for interesting multidimensional separations
followed mass analysis. While this technique is still in its starting phases, it
definitely bears watching as the method matures and finds its place in
protein purity analysis.
can result in reduced signals. In such cases, alternative ionization modes can
be used. MALDI is useful in these instances, as this mode can handle very
large (>1 MDa) proteins and has a higher tolerance for salts. Chro-
matographic separation followed by lyophilization of the individual chro-
matographic peaks allows for LC/MALDI analysis. MALDI also tends to
form ions with a single charge state, so that protein ions are largely observed
as their [M þ 1H]þ peaks. This can simplify spectral interpretation, as
dimer peaks will show up at twice the m/z, avoiding the doubly charged
dimer issue altogether.
Mass spectrometry is limited due to the fact that proteins are generally
not isotopically pure. The presence of 13C, 15N, 34S, and other isotopes
occurring in proteins causes broadening of mass spectral peaks in TOF as the
mass of the protein increases. This broadening can eventually make m/z
envelope of a protein ion so broad that the molecular weight becomes
indeterminant. In this instance, peaks due to adduct formation, ligand
binding, and di- or multimer formation may become indistinct. Although
this issue is generally less significant below about 6–7 MDa, it does provide
an upper bound to the method. This can be circumvented in some studies
by producing the protein using amino acids or other starting blocks that are
isotopically depleted, these have a lower than natural abundance amount of
13C, 15N, 34S or similar isotopes.
4. Scattering Methods
Light scattering methods are generally used to obtain molecular
weights and radii of gyration, but specific scattering methods can provide
diffusion coefficients, molecular weight, and thermodynamic parameters
(Chu, 1974; Cottingham, 2005; Timasheff and Townend, 1970).
Instrumentation for key light scattering methods has been developed
which integrates with HPLC and FFF instrumentation and incorporates
nearly all of the data analysis functions. This ease of use and accessibility has
resulted in an resurgence in the application of these powerful methodologies
to protein characterization (Harding et al., 1991).
The basis of light scattering is the interaction of the oscillating electro-
magnetic field of the incident light with the electrons in the protein to
induce oscillating dipoles. These oscillating dipoles radiate light, and the
intensity and directionality of the light depends on many factors, including
the size and structure of the molecule.
Size, Molecular Weight, and Subunits 717
16p2 n0 2 2 2 y
PðyÞ ¼ 1 r sin þ . . . ; ð39:14Þ
3l2 g 2
5. Presence of Subunits
To determine whether subunits are present generally involves charac-
terization of one or more properties of the system under conditions which
favor association and under conditions which are likely to favor dissociation.
Thus, the investigator should have prior knowledge of the conditions under
which association and dissociation should be expected, or will need to
investigate a variety of conditions which have been shown in other cases
to result in dissociation. The investigation to identify these conditions may
require significant effort, but if properly planned, will also yield additional
information about the nature of the system. Before initiating an extensive
search for associating or dissociating conditions, it is worthwhile using
extreme conditions as a preliminary evaluation. One normally carries out
one fractionation procedure under some (often physiological) buffer
condition and a second under strongly denaturing conditions.
One very simple, rapid, and low-cost approach for searching for subunits
is to carry out electrophoresis under nondenaturing conditions, using the
buffer in which the protein was isolated, followed by a second dimension
under dissociating or denaturing conditions. The nature of the denaturing
conditions may vary depending on the type of subunit and the association
one expects to find. For example, one could run the second electrophoresis
at extremes of pH, in the presence of urea or SDS, or in the presence or
absence of other buffered components like calcium. Changes in mobility
due to changes of buffer condition should be accounted for and understood
by analyzing the second dimension. That is, analysis should be based on the
apparent size of the molecule(s) under associating and possibly dissociating
buffer conditions. If one is investigating the possibility that disulfide-linked
subunits are present, SDS gel electrophoresis in the absence of reducing
agents may be carried out as the first dimension, and SDS gel electrophoresis
in the presence of reducing agents may be carried out in the second
dimension. Because in-gel alkylation of disulfides is difficult, it is recom-
mended that mercaptoacetate be included in the second gel running buffer
to avoid reoxidation. Differences in the apparent molecular weight or the
appearance of multiple components in the second dimension will indicate
that disulfide-linked subunits were present. Differences in apparent molec-
ular weight deduced from nondenaturing electrophoresis compared with
the apparent molecular weight based on electrophoresis under completely
denaturing conditions (SDS–PAGE) could indicate the presence of subu-
nits, but one must consider the possibility that the electrophoretic behavior
of the molecule is due to extremes of asymmetry or intrinsic charge.
In the case of subunits which are in equilibrium between associated and
dissociated states, the definitive approach to determining the presence of
720 David G. Rhodes et al.
REFERENCES
Ackers, G. K. (1970). Analytical gel chromatography of proteins. Adv. Protein Chem. 24,
343–446.
Andrews, A. T. (1986). Electrophoresis: Theory, Techniques, and Biochemical and Clinical Appli-
cations. Clarendon Press, Oxford.
Arakawa, T., and Timasheff, S. N. (1985). Calculation of the partial specific volume of
proteins in concentrated salt and amino acid solutions. Methods Enzymol. 117, 60.
Attri, A. K., and Minton, A. P. (1986). Technique and apparatus for automated fractionation
of the contents of small centrifuge tubes: Application to analytical ultracentrifugation.
Anal. Biochem. 152, 319–328.
Baldwin, M. A., Medzihradszky, K. F., Lock, C. M., Fisher, B., Settineri, T. A., and
Burlingame, A. L. (2001). Matrix-assisted laser desorption/ionization coupled with
quadrupole/orthogonal acceleration time-of-flight mass spectrometry for protein discovery,
identification, and structural analysis. Anal. Chem. 73, 1707–1720.
Bartolomeo, M., and Maisano, F. (2006). Validation of a reversed-phase HPLC method for
quantitative amino acid analysis. J. Biomol Tech. 17, 131–137.
Bock, P., and Halvorson, H. (1983). Molecular weight of human high-molecular-weight
kininogen light chain by equilibrium sedimentation in an air-driven ultracentrifuge.
Anal. Biochem. 137, 172–179.
Burgess, R. (2009). Use of bioinformatics in planning a protein purification. Methods
Enzymol. 463, 21–28.
Cantor, C. R., and Schimmel, P. R. (1980). In ‘‘Biophysical Chemistry’’, Part II p. 565.
Freeman, San Francisco, CA.
Chu, B. (1974). Laser Light Scattering. Academic Press, New York.
Cole, J. L., Lary, J. W., Moody, T. P., and Laue, T. M. (2008). Analytical ultracentrifuga-
tion: Sedimentation velocity and sedimentation equilibrium. Methods Cell Biol. 84,
143–179.
Cottingham, K. (2005). Versatile TOFMS. Anal. Chem. 77, 227–231.
Demeler, B., Saber, H., and Hansen, J. C. (1997). Identification and interpretation of
complexity in sedimentation velocity boundaries. Biophys. J. 72, 397–407.
Durschlag, J. (1986). In ‘‘Thermodynamic Data for Biochemistry and Biotechnology,’’
(H. Hinz, ed.), p. 46. Springer-Verlag, Berlin.
Faull, K. F., Dooley, A. N., Halgand, F., Shoemaker, L. D., Norris, A. J., Ryan, C. M.,
Laganowsky, A., Johnson, J. V., and Katz, J. E. (2008). An introduction to the basic
principles and concepts of mass spectrometry. Compr. Anal. Chem. 52, 1–46.
Fish, W. W. (1978). In ‘‘Methods in Membrane Biology,’’ (E. E. Korn, ed.), Vol. 4, p. 289.
Plenum Press, New York.
Folta-Stogniew, E. (2000). Oligomeric states of proteins determined by size-exclusion
chromatography coupled with light scattering, absorbance, and refractive index detec-
tors. Methods Mol. Biol. 328, 97–112.
722 David G. Rhodes et al.
Contents
1. Introduction 726
2. Enrichment Techniques for Identifying PTMs 731
2.1. Phosphorylation 731
2.2. Glycosylation 737
2.3. Ubiquitination and sumoylation 738
3. Nitrosative Protein Modifications 740
4. Methylation and Acetylation 741
5. Mass Spectrometry Analysis 744
6. CID versus ECD versus ETD 747
7. Quantifying PTMs 750
8. Future Directions 756
Acknowledgments 758
References 758
Abstract
Posttranslational modifications (PTMs) of proteins perform crucial roles in
regulating the biology of the cell. PTMs are enzymatic, covalent chemical
modifications of proteins that typically occur after the translation of mRNAs.
These modifications are relevant because they can potentially change a
protein’s physical or chemical properties, activity, localization, or stability.
Some PTMs can be added and removed dynamically as a mechanism for rever-
sibly controlling protein function and cell signaling. Extensive investigations have
aimed to identify PTMs and characterize their biological functions. This chapter
will discuss the existing and emerging techniques in the field of mass spectrom-
etry and proteomics that are available to identify and quantify PTMs. We will
725
726 Adam R. Farley and Andrew J. Link
1. Introduction
Protein posttranslational modifications (PTMs) perform essential roles
in the biological regulation of a cell. PTMs are enzymatic, covalent chemi-
cal modifications of proteins that typically occur after translation from
mRNAs. Chemical modifications of proteins are extremely important
because they potentially change a protein’s physical or chemical properties,
conformation, activity, cellular location, or stability. In fact, most proteins
are altered by the addition or removal of a chemical moiety on either an
amino acid or the protein’s N- or C-terminus. Some PTMs can be added
and removed dynamically as a mechanism for reversibly controlling protein
function. Over 400 specific protein modifications have been identified, and
more are likely to be identified (Creasy and Cottrell, 2004). The most
commonly identified PTMs to date include phosphorylation, sumoylation,
ubiquitination, nitrosylation, methylation, acetylation, sulfation, glycosyla-
tion, and acylation (Table 40.1).
Vast amounts of scientific effort have gone toward identifying PTMs and
elucidating their biological functions. Perhaps the best studied cases involve
eukaryotic histones and the myriad of PTMs that are associated with
transcriptional regulation (Berger, 2007; Fuchs et al., 2009; Issad and Kuo,
2008; Olsson et al., 2007; Reid et al., 2009). Histones associate with
chromosomal DNA to form nucleosomes that bundle together to form
chromatin fibers. The histone code hypothesizes that chromatin–DNA
interactions are regulated by combinations of histone PTMs ( Jenuwein
and Allis, 2001; Strahl and Allis, 2000). The acetylation of lysine was first
discovered in histones and is correlated with actively transcribed genes
(Roth et al., 2001). Combinations of methylation, acetylation, ADP-
ribosylation, ubiquitination, and phosphorylation of histone tails function
to regulate specific gene expression programs (Godde and Ura, 2008).
One of the most ubiquitous PTMs, phosphorylation, is the focus of
many biochemical investigations. It has been estimated that 30% of the
human proteome is phosphorylated (Cohen, 2001; Hubbard and Cohen,
1993). For example, tyrosine kinases and phosphatases transduce signals
from ligand-bound receptors on the cell surface to downstream targets in
the insulin/IGF-1 signaling pathway (Taniguchi et al., 2006).
Classic approaches to detecting PTMs on proteins involved Edman
degradation and thin-layer chromatography (TLC). However, these meth-
ods are hampered by the requirement of significant amounts of starting
Protein Posttranslational Modifications 727
OH
O P OH
N
H
O
OH−
H3PO4 1. b -elimination
N
H
O
SH
2. Michael-like
SH addition
S
HS
N
H
O
S N
S
3. Binding to
affinity resin
H
S N
NH
S
S S +
Figure 40.1 The chemical modification and affinity purification of proteins bearing a
phosphoserine. Reprinted with permission from McLachlin and Chait (2003).
Protein Posttranslational Modifications 735
residues prior to fragmentation along the peptide backbone. The vast majority
of the population undergoes b-elimination reaction, leaving only a fraction
remaining to fragment along the backbone (Bennett et al., 2002; Lee et al.,
2001; Ma et al., 2001). The resulting mass spectrum contains fewer sequenc-
ing ions, and these ions are generally lower in intensity. Therefore, there is a
low signal-to-noise ratio (Fig. 40.2), which hampers interpreting the peptide
sequence. Although capable of losing the phosphate group, phosphotyrosine
seldom generates the intense neutral loss ions that are seen for peptides
containing phosphoserine or phosphothreonine residues. Some mass spectro-
meters that use ETD avoid this neutral loss problem and can yield spectra
suitable for both sequence analysis and phosphoresidue determination. This
alternative fragmentation method will be discussed further when we consider
the various instrument types used in the proteomic identification of PTMs.
On the other hand, neutral loss artifacts can be used to the advantage
of the investigator. The neutral loss signature is a conspicuous indicator
that the peptide contains a site of phosphorylation. The neutral loss peak
observed in the mass spectrum is generally of high intensity due to its
preferential formation. Therefore, if a loss of 98 m/z (H3PO4) or 80 m/z
(HPO3) for a þ1 charged peptide is observed from the precursor ion,
one can infer the presence of a phosphoserine or phosphothreonine
reside within the sequence of the peptide. Modern mass spectrometers
can monitor for this mass loss. An ion trap mass spectrometer can isolate
the neutral loss species and initiate another round of fragmentation (MS3
scan) (Beausoleil et al., 2004; Gruhler et al., 2005; Olsen and Mann,
2004; Ulintz et al., 2009). This additional analysis step can result in
additional sequence coverage to aid in identifying the phosphopeptide
and mapping the PTM to a specific amino acid. An alternative approach
to an MS3 scan is ‘‘Pseudo MSn’’ or multistage activation (MSA) for
phosphopeptide fragmentation in an ion trap mass spectrometer
(Schroeder et al., 2004). The method induces CID of the neutral loss
product from the MS/MS scan of the phosphopeptide. The product ions
from the neutral loss ions along with the initial MS/MS product ions are
trapped together resulting in a composite spectrum. In MSA, the neutral
loss product ions are converted into a variety of structurally informative
fragment ions, which show improved scores in database search algorithms
(Schroeder et al., 2004; Ulintz et al., 2009).
Phosphorylation introduces a negative charge onto the peptide, and this
also affects analysis by mass spectrometry. The negative charge can reduce
the mass spectrometer response when operated in positive-ion mode
(McLachlin and Chait, 2001). This diminished response results in poor
quality spectra and makes identifying the phosphopeptide more difficult.
Furthermore, this increased negative charge can reduce proteolysis during
preparation of protein mixtures prior to mass spectrometry. As a caveat,
Hanno Steen and coworkers found no evidence for decreased ionization/
736 Adam R. Farley and Andrew J. Link
A
100 +2
(M + 2H − H3PO4 )
95
90
85
80
75
70
65
60
55
50
45
40
35
30
25
20
15
10
5
0
400 600 800 1000 1200 1400 1600 1800 2000
m/z
B
10X
9.5
9.0 N M A I N P S* K E N L C S T F C K
8.5
8.0 y8
7.5
7.0 b12
6.5 y12
6.0
5.5
5.0 y6
4.5
4.0 b16
3.5 y10
3.0 y14
2.5 b12
b5
2.0 y5
b4 y4 y9 y13
1.5 b3
b15
1.0
0.5
0.0
400 600 800 1000 1200 1400 1600 1800 2000
M/Z
Figure 40.2 Phosphopeptide MS/MS spectrum from a bovine b-casein tryptic digest.
Panel A demonstrates the intense neutral loss peak observed at 999.2 m/z from the
precursor 1048 at a þ 2 charge state ( 49 Da). This is magnified in Panel B by a factor of
10 to reveal the sequencing ion peaks.
2.2. Glycosylation
Glycosylation is a common PTM and has been shown to affect enzyme
activity, protein localization, stability, signaling, cell adhesion, and protein
interactions (Spiro, 2002). Changes in glycosylation patterns on the surface of
cells and on proteins within biological fluids have been discovered in malig-
nant transformation and tumorogenesis (Durand and Seta, 2000). A variety of
glycosylations can be present within a proteome. O-glycosylated proteins are
modified at serine and threonine, while C-glycosylation targets tryptophan
and N-glycosylation targets asparagine. Proteins with N-glycosylation
share several commonalities such as core structures, prevalence on the
extracellular membrane, and enzymes that remove the N-linked structures
(Medzihradszky, 2005). N-glycosylation sites have has an established
consensus sequence of N-X-S/T where X is any amino acid except proline
(Pless and Lennarz, 1977). Another modification that targets serine and threo-
nine residues is O-linked b-N-acetylglucosamine (O-GlcNAc). Gerald Hart
and coworkers have pioneered the study of this dynamic nucleocytoplasmic
PTM (Holt and Hart, 1986). This modification targets many of the same
proteins selected for phosphorylation. It is postulated that this modification
may serve an antagonistic role to protein phosphorylation (Ahmad et al., 2006).
Some of the same problems to identify phosphorylations with mass
spectrometry also apply to the study of glycosylations. These modifications
tend to be substoichiometric in their abundance relative to unmodified
proteins. Glycosylation and modification by O-GlcNAc are both labile in
the mass spectrometer and often yield poor quality spectra for sequencing.
Even though there is a well characterized consensus sequence known for
the N-glycosylation of proteins, only a subset of the proteins containing
this sequence are glycosylated in vivo.
738 Adam R. Farley and Andrew J. Link
Table 40.2 Ubiquitin-like modifications and their associated nominal mass remnants
from humans and yeast
(Glozak et al., 2005; Grewal and Rice, 2004; Sadoul et al., 2008; Yang and
Seto, 2008). In eukaryotes, N-terminal acetylation is one of the most common
PTMs, occurring on approximately 85% of eukaryotic proteins (Polevoda and
Sherman, 2000). The first nonhistone protein found to be regulated by
acetylation and deacetylation was the oncogene p53 (Gu and Roeder, 1997).
As with many of the other PTMs, the classical approach to identifying
methylation and acetylation was the use of modification-specific antibodies.
This technique is limited by cross-reactivity and specificity problems asso-
ciated with an immunological-based method. In the case of histones,
methods exist for their selective isolation from the nuclei of cells (Garcia
et al., 2007). A more general approach to identifying sites of acetylation and
methylation is the use of an in vitro radiolabeling assay. This technique
utilizes acetyltransferases or methyltransferases to catalyze the addition of
radioactive acetyl or methyl donor group. Thin layer chromatography
(TLC) or HPLC separation followed by microsequencing can reveal the
radiolabeled amino acids (Sobel et al., 1994). The radiolabeling approach has
disadvantages in that it requires significant amounts of pure starting material
and utilizes radioactivity to detect modified amino acids.
There are no efficient methods for the selective enrichment of peptides
or proteins with methylation or acetylation modifications. Additionally,
since acetylation and methylation occur on basic residues, the use of trypsin
as a protease can be less effective in cleaving at its usual sites. This can be
more dramatic in the case of methylation since the basic residues can be
methylated at multiple amine sites. The obvious choice to overcome this
problem is selecting an alternative protease. However, the benefit of the
missed tryptic cleavages can work to the investigator’s advantage, serving as
an indicator that the site may be modified. Tandem mass spectral searching
algorithms can take into account the average mass shift caused by acetylation
and methylation of 42 and 14 Da, respectively, as well as the shift of 28 Da
for dimethylation and 42 Da for trimethylation. Some ambiguity can appear
when considering trimethylation since the mass shift between it and an
acylated peptide is only 0.0363 Da. This mass difference can be resolved
with the use of a mass spectrometer with a high enough degree of mass
accuracy and resolution (Zhang et al., 2004).
Another diagnostic tool to identify sites of acetylation and methylation
is analysis of the immonium ions generated in CID type mass spectro-
meters. Peptides that harbor an unmodified lysine generally produce a
characteristic immonium ion peak at 84 m/z (Trelle and Jensen, 2007).
When the lysine is modified, a different immonium is formed at a unique
m/z value (see Figure 40.3). Di- and trimethylated forms of the immonium
do not form due to constraints of the rearrangement reaction that occurs.
However, the trimethylated form produces a unique neutral loss fragment
59 Da less than the precursor due to loss of trimethylamine (Zhang et al.,
2004). Figure 40.3 shows the rearrangement reaction that takes place as
Protein Posttranslational Modifications 743
O NH
NH2 NH N N+
Immonium ion
Immonium ion
Immonium ion
Immonium ion
Immonium ion
HN HN HN HN HN
m/z 101 m/z 115 m/z 129 m/z 143 m/z 143
Elimination of ammonia
Elimination of ammonia
Elimination of ammonia
Elimination of ammonia
Elimination of ammonia
and rearrangement
and rearrangement
and rearrangement
and rearrangement
and rearrangement
+
NH+ NH+ N+ NH+ NH+ NH+ N
O
Figure 40.3 Proposed structures of the immonium ions generated during CID for
mono-, di-, and trimethylated lysines. Reprinted with permission from Zhang et al. (2004).
well as the associated m/z values of the various immonium ions. These
unique diagnostic characteristics are only generated in mass spectrometers
that are capable of CID.
744 Adam R. Farley and Andrew J. Link
+ +
H3N H2N
H
O
H O −O O
H
R C
N
N
OH e− R • H N
OH
H O N
+NH3 +NH 3 O
+NH3 +NH3
NH2
H
O
R H O
NH +
HC N
+NH3 • OH
O
C⬘ Z⬘•
+NH3
Figure 40.4 Fragmentation scheme for production of ions of type c and z by reaction
of a low-energy electron with a multiply protonated peptide. Reprinted with permis-
sion from Udeshi et al. (2007).
A Ac
T A R K S T G G K A P R K Q L
100 [M+4H−2H2O]+4
Normalized intensity: 1.7 x 105
Relative abundance
80
60 y11+2
y13+3 b5
40 b14+3
+2
a5-NH+2
2
y9
20 y10+3
y5 y6
200 400 600 800 1000 Ac 1200 1400 1600
m/z
B T A R K S T G G K A P R K Q L
100
Normalized intensity: 1.0 x 104
Relative abundance
80 .
[M+4H]+3 +2..
[M+4H]
60 [M+4H]+4 C5 C7
C6 C8 Z7
40 C3 C4 Z4 Z6 Z9 C9
Z8 Z10 Z14 C14
20 Z12 ..
Z11
C13 Z13 [M+4H]+1
C2 Z2 Z3 C11 C12
Figure 40.5 A comparison of CID versus ETD spectra for a human histone H3
peptide. Panel A shows the CID spectrum with the corresponding b- and y-ion series.
Panel B depicts the ETD spectrum and its corresponding c- and z-ion series. Reprinted
with permission from Mikesh et al. (2006).
750 Adam R. Farley and Andrew J. Link
phosphopeptides identified with ETD and CID found that only 17.9% of
the sites identified were shared among the two datasets generated (Swaney
et al., 2008). This suggests that the best approach to fragmenting and
identifying sites of PTMs may be a combination of CID and ETD.
Figure 40.5 shows representative spectra generated with both methods
of fragmentation. The characteristic b- and y-ions appear in the CID
spectrum, while c- and z-type ions appear in ETD.
7. Quantifying PTMs
When investigating the biological significance of PTMs, it is often
advantageous to know the relative or absolute abundance of a specific
modification or group of PTMs. This allows for the direct comparison of
the modification of interest across varied biological samples. An example is
comparing the abundance of a PTM in cells or tissues obtained from a normal
versus a disease state. Quantifying these changes can lead to insights into the
role of PTMs in a myriad of processes from cell growth to diseases and
apoptosis. A range of methods are available to quantify PTMs. Traditional
methods utilize 2D-GE and differential staining to identify differences in the
level of protein expression among samples. However, this approach has a low
resolution, and differences in the ability of some proteins to stain can lead to
artifacts (Ong and Mann, 2005). The 2D-GE approach is most amenable to
abundant protein species (Anderson and Anderson, 1998). More recently,
mass spectrometry-based approaches have alleviated some of the problems
posed by gel-based approaches. These methods include protein or peptide
labeling strategies, tagging approaches and differential proteomic methods
using spectral counting (Gygi et al., 1999; Ong et al., 2002; Washburn
et al., 2001). While these methods are more widely applied to quantifying
protein changes among samples, they can also be used to quantify changes in
PTMs from different biological samples.
Quantitative measurements can either be absolute, in terms of concentra-
tion or copy number per cell, or relative, in terms of fold change among
differentially treated samples. Absolute concentrations are more difficult to
obtain, but are more informative and can be used to derive relative measure-
ments. LC-MS/MS experiments allow the plotting of signal intensities as a
peptide elutes from the chromatographic column over time. The area under
this curve for any given species is directly related to the abundance of the
peptide and allows for label-free quantitation. However, the physiochemical
properties of peptides such as hydrophobicity, charge, and size vary widely
and can lead to differences in the mass spectrometer’s detector response. In
addition, co-elution of other peptides and variations in elution conditions can
be problematic for quantitative investigations with this label-free approach.
Protein Posttranslational Modifications 751
The most basic approach for using stable isotopes is termed absolute
quantitation (AQUA) (Gerber et al., 2003). In this approach, isotopically
labeled peptides are generated synthetically and added as internal standards
to the peptide mixture. Steve Gygi and his colleagues used this approach for
determining the change in the abundance of phosphorylations during the
Xenopus cell cycle (Stemmann et al., 2001). This method can become
cumbersome in that it requires that a labeled synthetic peptide be generated
for every PTM to be quantified. Labeled phosphopeptides standards are
readily available, but this is not the case for other modifications, limiting
broad application of this method. The internal standard is typically added
just before or after proteolysis, in which case it is impossible to normalize for
variations in sample preparation upstream of the digestion step. This tech-
nique requires a prior knowledge of the PTM so the synthetic peptide can
be generated. It is also necessary to know the general abundance of the
modified peptide so the labeled standard can be added in a similar amount.
To minimize sample preparation errors, cell populations can be grown
under different conditions with or without an isotopic label, and the
populations pooled prior to sample preparation and analysis. Any errors in
sample preparation affect both populations equally. Mathias Mann’s
752 Adam R. Farley and Andrew J. Link
laboratory has developed the most popular of these approaches, called stable
isotope labeling by amino acids in cell culture (SILAC) (Ong et al., 2002). In
the SILAC approach, the culture media contains 13C6-arginine and 13C6-
lysine, which label the sites at which trypsin cleaves. Therefore, excluding
the C-terminal fragment, each tryptic product includes one labeled amino
acid. The advantage of the SILAC method over total metabolic labeling is
that the number of incorporated labels is defined and independent of the
amino acid sequence. This simplicity makes data interpretation easier with
SILAC than with other methods that label the peptides in more complicated
patterns. SILAC limits comparison to a maximum of three culture condi-
tions, which include either unlabeled, 13C6-labeled, or 13C615N4-labeled
amino acids. Other limitations of SILAC are that it is only useful with
certain model organisms and cell lines that can be grown in isotopically
defined media and can be prohibitively expensive. However, SILAC’s high
degree of accuracy makes this method well suited for the study of PTMs
(Olsen et al., 2006). A variation of SILAC employs labeling with labeled S-
adenosylmethionine for the identification and relative quantification of
protein methylation (Ong et al., 2004).
An alternative to in vivo metabolic labeling is enzymatic labeling in vitro.
This labeling can be performed either during protein digestion or after
proteolysis with an additional reaction step. Using heavy water (H218O),
trypsin or Glu-C will introduce two 18O atoms. The resulting 4 Da shift is
sufficient to resolve isotopes and acquire relative quantification data (Miyagi
and Rao, 2007). However, some proteases such as endoproteinase Lys-N
only incorporate one 18O atom and yield spectra insufficient to resolve
isotope peak overlap (Rao et al., 2005). Full enzymatic labeling is rarely
achieved and different peptides incorporate the label at different rates
thereby complicating data analysis (Ramos-Fernandez et al., 2007). As a
result, this method has not found widespread use in quantitative proteomics
(Ong et al., 2004).
Efficient chemical tagging approaches can alleviate some of the negative
constraints of enzymatic tagging procedures. The first of these was devel-
oped by Ruedi Aebersold’s laboratory and termed isotope-coded affinity
tags (ICAT) (Gygi et al., 1999). The ICAT reagent was initially composed
of a thiol reactive group that targets cysteines, a polyether linker region with
either zero or eight deuterium atoms, and a biotin group for affinity
purification with an avidin column. Since deuterium isotopes can behave
differentially in their chromatographic response relative to the light forms,
13C and 15N tagging reagents have been developed (Bottari et al., 2004).
116.1
117.1
70 70
60 60
116.1
50 50
114.1
40 40
30 30
20 20
10 10
0 0
111.0 112.8 114.6 116.4 118.2 120.0 111.0 112.8 114.6 116.4 118.2 120.0
Mass (m/z) Mass (m/z)
Figure 40.6 iTRAQ reporter ion series for two mixtures of varied ratios of a six
protein digest. Reprinted with permission from Ross et al. (2004).
754 Adam R. Farley and Andrew J. Link
of the iTRAQ tags are lost with the lower 1/3 of the MS/MS data. The use
of pulsed q-dissociation (PQD) was introduced to alleviate this problem but
has found limited use in proteomics (Bantscheff et al., 2008; Cunningham
et al., 2006; Griffin et al., 2007).
Using the iTRAQ approach, it is possible to compare the degree
of phosphorylation of the proteome under different conditions. Studies
performed by Forest White and coworkers used antiphosphotyrosine anti-
bodies followed by IMAC enrichment and iTRAQ labeling to investigate
the effect of epidermal growth factor (EGF) stimulation on the phosphory-
lation state of cells over time (Zhang et al., 2005). This study probed four
time points (0, 5, 10, and 30 min) in a single analysis. They were able to
quantify relative changes in the phosphorylation of 78 sites on 58 proteins
(Zhang et al., 2005). This experimental approach highlights the utility of
combining enrichment for PTMs with quantitation to gain additional
insight into the biology of cells. It is difficult to determine what fraction
of a peptide population is phosphorylated under certain conditions. This is
because the unphosphorylated counterpart will behave differently in its MS-
detector response. So while it is feasible to compare unmodified forms and
other unmodified forms or different phosphorylated forms of the same
peptide under different conditions, it is not reliable to compare modified
versus unmodified forms based on the ratio of reporter ion peaks generated.
A limited number of additional tagging methods aimed specifically at
defined PTMs have been developed. These utilize the selective chemical
reactivity of the chemically modified amino acid side chains. For the
quantitation of phosphorylations, b-elimination of phosphoric acid, and
subsequent Michael addition of ethanedithiol derivatives can be employed
to incorporate isotope tags (Tao et al., 2005). Hydrazine chemistry can be
applied to glycosylated peptides to replace the carbohydrate group with an
isotopically labeled tag (Zhang et al., 2003). These approaches are only
effective for a limited number of PTMs. In addition, if the chemical
reactions involved are not highly efficient, the resulting peptide mixture
can increase in complexity. The increase in complexity can lead to problems
in identifying the peptides and their associated PTMs as well as inaccurate
quantitative information.
Isotope labeling and tagging strategies can be tedious, cumbersome, and
expensive. In addition to the chromatographic peak integration method
discussed previously, other label-free quantitation methods exist that can be
applied to quantifying PTMs. These methods typically involve some form of
spectral counting and normalization for protein length (Washburn et al.,
2001). It is debatable whether these methods are truly quantitative, and as a
result, these methods are more commonly referred to as differential proteo-
mic techniques. The spectral counting methods are based on the principle
that, as the abundance of a protein increases in a given sample, more MS/MS
spectra are isolated for peptides derived from that protein. Relative
Protein Posttranslational Modifications 755
8. Future Directions
Traditional proteomic studies employ a bottom-up approach whereby
whole proteins are digested with a protease and analyzed using MS/MS
strategies to determine the peptide sequences. From these results, the
identities of the proteins can be inferred. However, complete coverage of
a protein is rarely accomplished. As a result, if peptides harboring the
modification are not detected, the PTMs cannot be identified. Top-down
investigations aim to determine the entire primary structure of a single
protein by performing tandem mass spectrometry on the intact protein to
produce fragmentation information. The combination of precursor mass
measurement and fragmentation sequencing data allows the identity of the
protein to be determined. Top-down methods can be useful in determining
specific PTMs and identifying splice variants and degradation products not
typically identified in bottom-up approaches (Waanders et al., 2007). Top-
down mass spectrometry also can discern gene products with a high degree
of sequence identity (Parks et al., 2007; Siuti et al., 2006). Top-down
approaches are readily applicable to the identification of PTMs. The initial
MS event on the whole protein determines if it has been modified from its
native form based upon any mass shift that may be present. The change in
the observed mass can be accounted for by only a limited number of
modifications. Additional fragmentation data generated by MS/MS can
help to further characterize the modification based on the observed mass
shift(s) for the peptides.
Top-down proteomics requires the use of instruments with high reso-
lution and high mass accuracy. This has relegated top-down studies to
FT-ICR and Orbitrap mass spectrometers with typical resolving powers
> 500,000 and 100,000, respectively (Macek et al., 2006). The high resolv-
ing power allows for the accurate determination of charge states, which is
essential for measuring the intact protein’s mass. The sub- to low-ppm mass
accuracy enables the discrimination between PTMs with the same nominal
mass, such as acetylation and trimethylation. Recently, studies by Andrea
Armirotti and coworkers and Scott McLuckey and his group have demon-
strated the ability to perform top-down proteomics on a q-TOF mass
spectrometer (Armirotti et al., 2009; Liu et al., 2009). This extending the
utility of this approach beyond the generally more expensive FT-ICR and
Orbitrap mass spectrometers and opens this methodology to a broader
section of investigators. Top-down approaches possess the ability to handle
a dynamic range of protein sizes. Parks et al. (2007) used such an approach to
identify 22 proteins ranging from 14 to 35 kDa in S. cerevisiae. One group
extended this range to characterize the PTMs of a 115-kDa cardiac myosin-
binding protein (Ge et al., 2009).
Protein Posttranslational Modifications 757
tools are all crucial in determining the successfulness of the approach. This
chapter has presented many of the techniques and technologies that exist today
in the field of proteomics that can be tailored toward the goal of determining
PTMs. The chemical modifications of proteins are extremely important
because they potentially change many properties of a protein. Therefore, it is
important to take advantage of the approaches offered in mass spectrometry to
determine sites of PTMs and elucidate their biological function.
ACKNOWLEDGMENTS
We acknowledge Elizabeth M. Link and Kevin Schey for helpful comments and suggestions
in the preparation of this manuscript. A. R. F. was supported by NIH training grant T32
CA009385 and GM64779. A. J. L. was supported by NIH grants GM64779 and AR055231.
REFERENCES
Ahmad, I., Hoessli, D. C., Walker-Nasir, E., Choudhary, M. I., Rafik, S. M., and
Shakoori, A. R. (2006). J. Cell. Biochem. 99, 706–718.
Anderson, N. L., and Anderson, N. G. (1998). Electrophoresis 19, 1853–1861.
Armirotti, A., Benatti, U., and Damonte, G. (2009). Rapid Commun. Mass Spectrom. 23,
661–666.
Bantscheff, M., Schirle, M., Sweetman, G., Rick, J., and Kuster, B. (2007). Anal. Bioanal.
Chem. 389, 1017–1031.
Bantscheff, M., Boesche, M., Eberhard, D., Matthieson, T., Sweetman, G., and Kuster, B.
(2008). Mol. Cell. Proteomics 7, 1702–1713.
Barnes, P. J., Shapiro, S. D., and Pauwels, R. A. (2003). Eur. Respir. J. 22, 672–688.
Bayer, P., Arndt, A., Metzger, S., Mahajan, R., Melchior, F., Jaenicke, R., and Becker, J.
(1998). J. Mol. Biol. 280, 275–286.
Beausoleil, S. A., Jedrychowski, M., Schwartz, D., Elias, J. E., Villen, J., Li, J., Cohn, M. A.,
Cantley, L. C., and Gygi, S. P. (2004). Proc. Natl. Acad. Sci. USA 101, 12130–12135.
Beckman, J. S., Beckman, T. W., Chen, J., Marshall, P. A., and Freeman, B. A. (1990). Proc.
Natl. Acad. Sci. USA 87, 1620–1624.
Bennett, K. L., Stensballe, A., Podtelejnikov, A. V., Moniatte, M., and Jensen, O. N. (2002).
J. Mass Spectrom. 37, 179–190.
Berger, S. L. (2007). Nature 447, 407–412.
Bernstein, E., and Allis, C. D. (2005). Genes Dev. 19, 1635–1655.
Blagoev, B., Ong, S. E., Kratchmarova, I., and Mann, M. (2004). Nat. Biotechnol. 22,
1139–1145.
Blume-Jensen, P., and Hunter, T. (2001). Nature 411, 355–365.
Bodenmiller, B., Mueller, L. N., et al. (2007). Nat. Methods 4, 231–237.
Bottari, P., Aebersold, R., et al. (2004). Bioconjug. Chem. 15, 380–388.
Bunger, M. K., Cargile, B. J., et al. (2008). Anal. Chem. 80, 1459–1467.
Cantin, G. T., Shock, T. R., Park, S. K., Madhani, H. D., and Yates, J. R. 3rd (2007). Anal.
Chem. 79, 4666–4673.
Carr, S. A., Huddleston, M. J., and Annan, R. S. (1996). Anal. Biochem. 239, 180–192.
Chalkley, R. J., Baker, P. R., Medzihradszky, K. F., Lynn, A. J., and Burlingame, A. L.
(2008). Mol. Cell. Proteomics 7, 2386–2398.
Chang, I. F. (2006). Proteomics 6, 6158–6166.
Protein Posttranslational Modifications 759
Chen, Y. R., Chen, C. L., Chen, W., Zweier, J. L., Augusto, O., Radi, R., and
Mason, R. P. (2004). J. Biol. Chem. 279, 18054–18062.
Cirulli, C., Chiappetta, G., Marino, G., Mauri, P., and Amoresano, A. (2008). Anal. Bioanal.
Chem. 392, 147–159.
Clauser, K. R., Baker, P., and Burlingame, A. L. (1999). Anal. Chem. 71, 2871–2882.
Cohen, P. (2001). Eur. J. Biochem. 268, 5001–5010.
Collier, T. S., Hawkridge, A. M., Georgianna, D. R., Payne, G. A., and Muddiman, D. C.
(2008). Anal. Chem. 80, 4994–5001.
Coon, J. J., Syka, J. E. P., Schwartz, J. C., Shabanowitz, J., and Hunt, D. F. (2004). Int. J.
Mass Spectrom. 236, 33–42.
Coon, J. J., Ueberheide, B., et al. (2005). Proc. Natl. Acad. Sci. USA 102, 9463–9468.
Covey, T. R., Huang, E. C., and Henion, J. D. (1991). Anal. Chem. 63, 1193–1200.
Creasy, D. M., and Cottrell, J. S. (2004). Proteomics 4, 1534–1536.
Cunningham, C. Jr., Glish, G. L., and Burinsky, D. J. (2006). J. Am. Soc. Mass Spectrom. 17,
81–84.
Denison, C., Kirkpatrick, D. S., and Gygi, S. P. (2005). Curr. Opin. Chem. Biol. 9, 69–75.
Dongre, A. R., Jones, J. L., Somogyi, A., and Wysocki, V. H. (1996). J. Am. Chem. Soc. 118,
8365–8374.
Du, Y., Parks, B. A., Sohn, S., Kwast, K. E., and Kelleher, N. L. (2006). Anal. Chem. 78,
686–694.
Durand, G., and Seta, N. (2000). Clin. Chem. 46, 795–805.
Fang, S., and Weissman, A. M. (2004). Cell. Mol. Life Sci. 61, 1546–1561.
Ferguson, J. T., Wenger, C. D., Metcalf, W. W., and Kelleher, N. L. (2009). J. Am. Soc.
Mass Spectrom. 20, 1743–1750.
Ficarro, S. B., McCleland, M. L., Stukenberg, P. T., Burke, D. J., Ross, M. M.,
Shabanowitz, J., Hunt, D. F., and White, F. M. (2002). Nat. Biotechnol. 20, 301–305.
Fuchs, S. M., Laribee, R. N., and Strahl, B. D. (2009). Biochim. Biophys. Acta 1789, 26–36.
Garcia, B. A., Shabanowitz, J., and Hunt, D. F. (2005). Methods 35, 256–264.
Garcia, B. A., Shabanowitz, J., and Hunt, D. F. (2007). Curr. Opin. Chem. Biol. 11, 66–73.
Gatzka, M., and Walsh, C. M. (2007). Autoimmunity 40, 442–452.
Ge, Y., Lawhorn, B. G., et al. (2002). J. Am. Chem. Soc. 124, 672–678.
Ge, Y., Rajkumar, L., Guzman, R. C., Nandi, S., Patton, W. F., and Agnew, B. J. (2004).
Proteomics 4, 3464–3467.
Ge, Y., Rybakova, I. N., Xu, Q., and Moss, R. L. (2009). Proc. Natl. Acad. Sci. USA 106,
12658–12663.
Gerber, S. A., Rush, J., Stemman, O., Kirschner, M. W., and Gygi, S. P. (2003). Proc. Natl.
Acad. Sci. USA 100, 6940–6945.
Gingras, A. C., Raught, B., Gygi, S. P., Niedzwiecka, A., Miron, M., Burley, S. K.,
Polakiewicz, R. D., Wyslouch-Cieszynska, A., Aebersold, R., and Sonenberg, N.
(2001). Genes Dev. 15, 2852–2864.
Glozak, M. A., Sengupta, N., Zhang, X., and Seto, E. (2005). Gene 363, 15–23.
Godde, J. S., and Ura, K. (2008). J. Biochem. 143, 287–293.
Good, D. M., Wirtala, M., McAlister, G. C., and Coon, J. J. (2007). Mol. Cell. Proteomics 6,
1942–1951.
Goodlett, D. R., Keller, A., Watts, J. D., Newitt, R., Yi, E. C., Purvine, S., Eng, J. K., von
Haller, P., Aebersold, R., and Kolker, E. (2001). Rapid Commun. Mass Spectrom. 15,
1214–1221.
Goshe, M. B., Conrads, T. P., Panisko, E. A., Angell, N. H., Veenstra, T. D., and
Smith, R. D. (2001). Anal. Chem. 73, 2578–2586.
Grewal, S. I., and Rice, J. C. (2004). Curr. Opin. Cell Biol. 16, 230–238.
Griffin, T. J., Xie, H., Bandhakavi, S., Popko, J., Mohan, A., Carlis, J. V., and Higgins, L.
(2007). J. Proteome Res. 6, 4200–4209.
760 Adam R. Farley and Andrew J. Link
Gruhler, A., Olsen, J. V., Mohammed, S., Mortensen, P., Faergeman, N. J., Mann, M., and
Jensen, O. N. (2005). Mol. Cell. Proteomics 4, 310–327.
Gu, W., and Roeder, R. G. (1997). Cell 90, 595–606.
Gygi, S. P., Rist, B., Gerber, S. A., Turecek, F., Gelb, M. H., and Aebersold, R. (1999).
Nat. Biotechnol. 17, 994–999.
Haas, W., Faherty, B. K., Gerber, S. A., Elias, J. E., Beausoleil, S. A., Bakalarski, C. E.,
Li, X., Villen, J., and Gygi, S. P. (2006). Mol. Cell. Proteomics 5, 1326–1337.
Hansen, B. T., Jones, J. A., Mason, D. E., and Liebler, D. C. (2001). Anal. Chem. 73,
1676–1683.
Hansen, B. T., Davey, S. W., Ham, A. J., and Liebler, D. C. (2005). J. Proteome Res. 4,
358–368.
Hayduk, E. J., Choe, L. H., and Lee, K. H. (2004). Electrophoresis 25, 2545–2556.
Holt, G. D., and Hart, G. W. (1986). J. Biol. Chem. 261, 8049–8057.
Huang, Y., Triscari, J. M., Tseng, G. C., Pasa-Tolic, L., Lipton, M. S., Smith, R. D., and
Wysocki, V. H. (2005). Anal. Chem. 77, 5800–5813.
Hubbard, M. J., and Cohen, P. (1993). Trends Biochem. Sci. 18, 172–177.
Ikeda, F., and Dikic, I. (2008). EMBO Rep. 9, 536–542.
Issad, T., and Kuo, M. (2008). Trends Endocrinol. Metab. 19, 380–389.
Ito, S., Hayama, K., and Hirabayashi, J. (2009). Methods Mol. Biol. 534, 195–203.
Jenuwein, T., and Allis, C. D. (2001). Science 293, 1074–1080.
Kelleher, N. L., Lin, H. Y., Valaskovic, G. A., Aaserud, D. J., Fridriksson, E. K., and
McLafferty, F. W. (1999). J. Am. Chem. Soc. 121, 806–812.
Kirkpatrick, D. S., Denison, C., and Gygi, S. P. (2005). Nat. Cell Biol. 7, 750–757.
Lange, V., Picotti, P., Domon, B., and Aebersold, R. (2008). Mol. Syst. Biol. 4, 222.
Larsen, M. R., Sorensen, G. L., Fey, S. J., Larsen, P. M., and Roepstorff, P. (2001).
Proteomics 1, 223–238.
Larsen, M. R., Thingholm, T. E., Jensen, O. N., Roepstorff, P., and Jorgensen, T. J. (2005).
Mol. Cell. Proteomics 4, 873–886.
Lee, C. H., McComb, M. E., Bromirski, M., Jilkine, A., Ens, W., Standing, K. G., and
Perreault, H. (2001). Rapid Commun. Mass Spectrom. 15, 191–202.
Lim, K. B., and Kassel, D. B. (2006). Anal. Biochem. 354, 213–219.
Lochnit, G., and Geyer, R. (2004). Biomed. Chromatogr. 18, 841–848.
Lu, B., Motoyama, A., Ruse, C., Venable, J., and Yates, J. R. 3rd (2008). Anal. Chem. 80,
2018–2025.
Ma, Y., Lu, Y., Zeng, H., Ron, D., Mo, W., and Neubert, T. A. (2001). Rapid Commun.
Mass Spectrom. 15, 1693–1700.
MacCoss, M. J., McDonald, W. H., Saraf, A., Sadygov, R., Clark, J. M., Tasto, J. J.,
Gould, K. L., Wolters, D., Washburn, M., Weiss, A., Clark, J. I., and Yates, J. R. 3rd
(2002). Proc. Natl. Acad. Sci. USA 99, 7900–7905.
Macek, B., Waanders, L. F., Olsen, J. V., and Mann, M. (2006). Mol. Cell. Proteomics 5,
949–958.
Makarov, A. (2000). Anal. Chem. 72, 1156–1162.
Makarov, A., Denisov, E., Kholomeev, A., Balschun, W., Lange, O., Strupat, K., and
Horning, S. (2006). Anal. Chem. 78, 2113–2120.
Mann, M., and Kelleher, N. L. (2008). Proc. Natl. Acad. Sci. USA 105, 18132–18138.
Manza, L. L., Codreanu, S. G., Stamer, S. L., Smith, D. L., Wells, K. S., Roberts, R. L., and
Liebler, D. C. (2004). Chem. Res. Toxicol. 17, 1706–1715.
McAlister, G. C., Berggren, W. T., et al. (2008). J. Proteome. Res. 7, 3127–3136.
McAlister, G. C., Phanstiel, D., et al. (2007). Anal. Chem. 79, 3525–3534.
McLachlin, D. T., and Chait, B. T. (2001). Curr. Opin. Chem. Biol. 5, 591–602.
McLachlin, D. T., and Chait, B. T. (2003). Anal. Chem. 75, 6826–6836.
Mechref, Y., Madera, M., and Novotny, M. V. (2008). Methods Mol. Biol. 424, 373–396.
Protein Posttranslational Modifications 761
Tsaprailis, G., Nair, H., Somogyi, A., Wysocki, V. H., Zhong, W., Futrell, J. H.,
Summerfield, S. G., and Gaskell, S. J. (1999). J. Am. Chem. Soc. 121, 5142–5154.
Udeshi, N. D., Shabanowitz, J., Hunt, D. F., and Rose, K. L. (2007). FEBS J. 274,
6269–6276.
Ulintz, P. J., Yocum, A. K., Bodenmiller, B., Aebersold, R., Andrews, P. C., and
Nesvizhskii, A. I. (2009). J. Proteome Res. 8, 887–899.
Unwin, R. D., Griffiths, J. R., Leverentz, M. K., Grallert, A., Hagan, I. M., and
Whetton, A. D. (2005). Mol. Cell. Proteomics 4, 1134–1144.
Unwin, R. D., Griffiths, J. R., and Whetton, A. D. (2009). Nat. Protoc. 4, 870–877.
Vestal, M. L., and Campbell, J. M. (2005). Methods Enzymol. 402, 79–108.
Waanders, L. F., Hanke, S., and Mann, M. (2007). J. Am. Soc. Mass Spectrom. 18,
2058–2064.
Washburn, M. P., Wolters, D., and Yates, J. R. 3rd. (2001). Nat. Biotechnol. 19, 242–247.
Wells, L., Vosseller, K., et al. (2002). Mol. Cell Proteomics 1, 791–804.
White, F. M. (2008). Curr. Opin. Biotechnol. 19, 404–409.
Williamson, B. L., Marchese, J., and Morrice, N. A. (2006). Mol. Cell Proteomics 5, 337–346.
Witze, E. S., Old, W. M., Resing, K. A., and Ahn, N. G. (2007). Nat. Methods 4, 798–806.
Wolf-Yadlin, A., Hautaniemi, S., Lauffenburger, D. A., and White, F. M. (2007). Proc. Natl.
Acad. Sci. USA 104, 5860–5865.
Wu, S. L., Jardine, I., et al. (2004). Rapid Commun. Mass Spectrom. 18, 2201–2207.
Wu, C. C., MacCoss, M. J., Howell, K. E., and Yates, J. R. 3rd (2003). Nat. Biotechnol. 21,
532–538.
Yang, Z., and Hancock, W. S. (2005). J. Chromatogr. A 1070, 57–64.
Yang, X. J., and Seto, E. (2008). Mol. Cell. 31, 449–461.
Yeo, W. S., Lee, S. J., Lee, J. R., and Kim, K. P. (2008). BMB Rep. 41, 194–203.
Zabrouskov, V., Senko, M. W., et al. (2005). J. Am. Soc Mass Spectrom. 16, 2027–2038.
Zappacosta, F., Collingwood, T. S., Huddleston, M. J., and Annan, R. S. (2006). Mol. Cell.
Proteomics 5, 2019–2030.
Zhan, X., and Desiderio, D. M. (2004). Biochem. Biophys. Res. Commun. 325, 1180–1186.
Zhang, R., and Regnier, F. E. (2002). J. Proteome Res. 1, 139–147.
Zhang, X., Herring, C. J., Romano, P. R., Szczepanowska, J., Brzeska, H.,
Hinnebusch, A. G., and Qin, J. (1998). Anal. Chem. 70, 2050–2059.
Zhang, K., Williams, K. E., Huang, L., Yau, P., Siino, J. S., Bradbury, E. M., Jones, P. R.,
Minch, M. J., and Burlingame, A. L. (2002). Mol. Cell. Proteomics 1, 500–508.
Zhang, H., Li, X. J., Martin, D. B., and Aebersold, R. (2003). Nat. Biotechnol. 21, 660–666.
Zhang, K., Yau, P. M., Chandrasekhar, B., New, R., Kondrat, R., Imai, B. S., and
Bradbury, M. E. (2004). Proteomics 4, 1–10.
Zhang, Y., Wolf-Yadlin, A., Ross, P. L., Pappin, D. J., Rush, J., Lauffenburger, D. A., and
White, F. M. (2005). Mol. Cell. Proteomics 4, 1240–1250.
Zhen, Y., Xu, N., Richardson, B., Becklin, R., Savage, J. R., Blake, K., and Peltier, J. M.
(2004). J. Am. Soc. Mass Spectrom. 15, 803–822.
C H A P T E R F O R T Y- O N E
Contents
1. Introduction 767
2. Strategies Based on End-Use 768
3. Parallel Cloning Strategies for Creating Expression Constructs 771
4. Small-Scale Expression Screening to Identify Suitable Constructs 774
5. Analytical Testing of Proteins for Selection 778
6. Large-Scale Parallel Expression 780
7. Conclusion 783
Acknowledgments 783
References 784
Abstract
Protein properties are highly diverse, making parallel expression and purifica-
tion a particular challenge. Parallel methods are typically used when a number
derivatives of a target protein are desired or when multiple homologs are
needed. A typical scenario involves target evaluation, cloning and mutagenesis
of the target, expression screening, large-scale expression and purification, and
analytical and biophysical testing of the resulting protein. This chapter
describes some of the strategies and methods employed for parallel protein
expression and purification.
1. Introduction
Parallel protein purification is mainly a challenge of logistics. There are
many purification approaches, which can be employed to isolate proteins.
Some of these are inherently more amenable to parallel processing than
others. (Graslund et al., 2008a; Kim et al., 2004; Lesley et al., 2002b) Most
parallel purification strategies employ the use of purification tags to facilitate
simple affinity purification with standardization of methods and protein
Genomics Institute of the Novartis Research Foundation, San Diego, California, USA
767
768 Scott A. Lesley
1 170 + +
1 166 −
1 162 + +
1 158 + +
1 154 + +
1 150 −
1 146 −
1 142 −
1 138 −
1 134 −
1 130 − 2HR2
1 126 −
6 174 −
20 174 + −
6 162 + +
20 162 + +
Figure 41.1 Identifying suitable protein expression constructs through parallel protein expression, purification, and characterization. Nested
N- and C-terminal truncations were evaluated in parallel for suitability for crystallization trials. Proteins showing both soluble expression with
sufficient protein yield and monodisperse behavior by ANSEC are indicated in gray. One such construct was entered into crystallization trials
and yielded a high-resolution structure.
770 Scott A. Lesley
related proteins and align them to identify conserved regions. These regions
typically constitute the core structural domains which need to be retained
within expression constructs and can be used to identify domain boundaries.
The basic local alignment search tool (BLAST) (Altschul et al., 1990) is the most
common tool for simple retrieval of related sequences. The NCBI (http://blast.
ncbi.nlm.nih.gov/Blast.cgi?CMD¼Web&PAGE_TYPE¼BlastHome) pro-
vides many versions of this essential tool along with access to the majority of
the genome sequences available to query. A hidden Markov model (HMM)
can be developed for a more sophisticated search but a simple PSI-BLAST
query is usually sufficient to find the most highly related sequences which
are the most useful.
Most proteins have already been aligned and placed into families.
Pfam (http://pfam.sanger.ac.uk/) (Finn et al., 2008) is a database of these
alignments and provides much useful information on comparative studies of
related proteins including domain boundaries and sequence conservation.
UniProt (http://www.uniprot.org/) (Consortium, 2008) also provides a
useful source of summary data including secondary structure predictions
for most proteins. Ligands can also be a useful tool for stabilizing proteins
and activity testing. Ligand predictions may be obvious from annotations,
but several databases such as KEGG (http://www.genome.jp/kegg/)
(Kanehisa et al., 2008) and BioCyc (http://www.biocyc.org) (Karp et al.,
2005) can provide additional ligand suggestions. Finally, a simple search of
the literature can often identify related proteins which have been expressed
previously to give guidance for expression. While seemingly an obvious
task, the rush to get an expression construct into whatever system is readily
available often leads to cursory or no review of past publications. There are
many other useful and easily accessible bioinformatic tools available. The
basic advice is to know what is known about your protein and those which
are closely related and to use that information for your experimental design.
Having gathered the available information on the target of interest,
expression constructs can now be designed. The most successful approach
is to evaluate not only the full-length open reading frame (ORF), but also to
create and test multiple truncations of this ORF in parallel. Defining the
N- and C-terminal boundaries is critical. We have observed many cases of
one or a few amino acid differences at the ends completely altering protein
behavior. Even with information on sequence conservation, multiple
N- and C-termini should be tried.
1. Query existing genome sequences for related proteins using BLAST and
perform sequence alignments. Tools such as Pfam can provide a conve-
nient forum for these queries and provides prevetted alignments and
domain boundaries. Incorporate information from literature searches.
2. Identify locations of N- and C-terminal sequence conservation as initial
truncation boundaries. In some cases, internal domain boundaries may
Parallel Methods for Expression and Purification 771
B X X X X X
C X X X X X
D X X X X X X
E X X X X
F X X X X X X
G X X X X X X
H X X X X X X
Figure 41.2 Parallel cloning process. The PIPE cloning method allows for simple
parallel cloning of many expression constructs, including deletions and point mutations
for screening. Appropriately designed primers are used for vector amplification and
insert amplification. Resulting amplifiers are simply mixed and transformed. Isolated
colonies are screened by a dPCR amplification and detection of PCR product by
fluorescent dye. Typically, two colonies are screened for each attempted construct
which is sufficient to identify insert-containing clones (shaded X) for most targets.
Insert-containing clones are then sequenced for confirmation.
aa 1–195
aa 1–191
aa 1–187
aa 1–183
aa 1–195
aa 1–191
aa 1–187
0 2 4 6 8 10 12 min
C
100,000
90,000
80,000
70,000
Fluorescence units
60,000
50,000
40,000
30,000
20,000
10,000
0
0 20 40 60 80 100 120
Temperature (C)
D
1 MRGMMLGMLAETHIHSGAGRSEGFVDLPVA 30
31 REAVTSYPVIAGSSLKGALRDAARERGMDE 60
61 SIFGDQDRAGDVLVSDARLLLLPVRSLTGS 90
Uncleaved
91 YRWVTCPHILERLSRDMRLCGISDGFEGAS 120
Cleaved
31790.801
1 MRGMMLGMLAETHIHSGAGRSEGFVDLPVA 30
31 REAVTSYPVIAGSSLKGALRDAARERGMDE 60
18035.201 61 SIFGDQDRAGDVLVSDARLLLLPVRSLTGS 90
91 YRWVTCPHILERLSRDMRLCGISDGFEGAS 120
121 VERGKACCTDDLNQIFLEEREFQRSNGIDG 150
9017.601 151 ALIDALKKMVPHKQTASRLERQLVIISDDD 180
181 FGWFASYGLPVIARNKLDDNKKSKNLWYEE 210
211 ALAPDTLMYAMVFERKDGALGKVQSMFETK 240
241 PYLQLGGNETVGMGWFAVKILEQGEGR 267
Figure 41.3 Analytical profiling of proteins derived from parallel expression and purifica-
tion. (A) Capillary electrophoresis results of a truncation series of YP_958225.1. (B) ANSEC
profile of a truncation series of YP_958225.1 showing variation in the amount of soluble
monomer (white box) and soluble aggregate (gray box) between samples. (C) Thermofluor
ligand binding assay showing increased thermostability YP_164977.1 (threonine aldolase)
in the presence of pyridoxal phosphate (filled symbols) versus the apo enzyme. (D) Partial
trypsin proteolysis of YP_843077.1 shows stable fragments visible by SDS–PAGE (cleaved
lane). Mass spectrometry can be used to define stable fragments of the original protein. Two
masses (18032.201 and 9017.601 Da) correspond to peptide fragments (underlined) which
can be targeted for subsequent subcloning and further expression studies.
Parallel Methods for Expression and Purification 777
Transfer pump T1
Buffer pump B1
Buffer
Air regulator
Sample
tube Cleaning
station
Waste Solenoid
value
Toe fitting
Heat Microfluidizer
exchanger
resuspend the pellet and effectively lyses the cells. If insufficient lysis is
observed, passage through a microfluidizer (Microfluidics) or French
press can be used.
3. Centrifuge the lysates for 30 min at 32,000g. The clarified lysates are
ready for affinity purification.
In parallel purification, conventional purification approaches can be a
significant bottleneck. Instrumentation and protocols for multisample pro-
cessing using HPLC instrumentation has been described (McMullan et al.,
2005). For most applications, however, an optimized affinity purification
will provide sufficient purity along with the requisite throughput. We have
found that a two-step nickel purification combined with a proteolytic
removal of the purification tag provides excellent purification and can be
performed in parallel for hundreds of samples at a diversity of scale. This
approach provides material that has proved suitable for crystallization trials
and enzymatic screening studies. Our vector contains a his-tag followed by a
TEV protease cleavage site. This protocol can be adapted for other tags and
cleavage enzymes.
1. Decant each clarified lysates onto two gravity-fed 1.5-ml nickel-chelat-
ing resin columns pre-equilibrated with Lysis Buffer. Collect each flow-
through into a waste tray.
2. Add 7.5 ml of Wash Buffer (50 mM HEPES pH 8.0, 300 mM NaCl,
40 mM imidazole, 10% glycerol, 1 mM TCEP). Add 0.5 ml of Elution
Buffer (20 mM HEPES pH 8.0, 300 mM imidazole, 10% glycerol, 1 mM
TCEP). This volume of elution buffer should be no more than one third
of the bed resin. The purpose is to push the elution front of the buffer to
the bottom of the resin to minimize the elution volume for the next step.
3. Place each column over a PD10 desalting column which has been pre-
equilibrated with Digestion Buffer (20 mM HEPES pH 8.0, 200 mM
NaCl, 40 mM imidazole, 1 mM TCEP). Add 2.5 ml of Elution Buffer to
the Ni-chelate resin and collect the elution directly onto the PD-10
column.
4. Elute the protein from the PD-10 column by further addition of 3.5 ml
of Digestion Buffer, collecting the elution in a 15 ml disposable tube.
Remove a small sample for protein assay and SDS–PAGE analysis.
5. The next step involves removal of the his-tag through proteolytic
cleavage. The protease itself is his-tagged. Therefore, subsequent passage
of the protein over a nickel resin will remove the protease, as well as
uncleaved proteins and those which stick nonspecifically to the nickel
resin while the protein of interest passes in the column flow-through.
For each 15 mg of protein obtained from the PD-10 elution, add 1 mg of
his-tagged TEV protease (Tropea et al., 2009). Bring the total volume of
the digestion to 9 ml using Digestion Buffer. Mix by inversion at room
temperature for 2 h or overnight at 4 C.
Parallel Methods for Expression and Purification 783
7. Conclusion
Parallel protein expression and purification is necessary to adequately
evaluate multiple targets or multiple derivatives of single targets. The
methods described here utilize bacterial expression systems. These systems
are by far the most convenient and robust platforms for expression, but often
suffer problems when applied to mammalian proteins. In many cases,
focusing on domains of interest and systematically optimizing expression
constructs to identify the proper domain boundaries can result in excellent
expression levels of active proteins (Graslund et al., 2008b; Klock et al.,
2008). Identifying suitable domain boundaries is one of the best applications
of parallel expression and purification. For example, kinase catalytic
domains have been defined and expressed in bacteria with great success
and have contributed significantly to drug design (Marsden and Knapp,
2008). Despite the difficulties and tedium of mammalian and baculovirus
expression systems, these platforms can also be implemented in a parallel
processing way (Peng et al., 1993). Platforms such as BacMam (Kost and
Condreay, 2002) offer additional flexibility in exploring multiple expression
systems. Regardless of the expression platform, the key to successful protein
expression is to understand the ultimate requirements of the protein in
terms of its physical properties and activities. Too often, simple solubility
and total expression levels are used as the sole selection criteria when
evaluating expression options. Even very small changes to the domain
boundaries of the expression construct can cause drastic changes to protein
behavior without visibly changing apparent soluble yield (Klock et al.,
2008). Combining parallel expression and purification with parallel analyti-
cal techniques and data analysis provides the best opportunity to identify a
successful expression construct and purification combination.
ACKNOWLEDGMENTS
The protocols described here were developed through the contributions of many people.
The author would particularly like to acknowledge Heath Klock, Mark Knuth, Carol Farr,
Anna Grzechnik, Julie Feuerhelm, Daniel McMullan, and Dennis Carlton for their
contributions to this chapter.
784 Scott A. Lesley
REFERENCES
Altschul, S. F., Gish, W., Miller, W., Myers, E. W., and Lipman, D. J. (1990). Basic local
alignment search tool. J. Mol. Biol. 215, 403–410.
Aslanidis, C., and de Jong, P. J. (1990). Ligation-independent cloning of PCR products
(LIC-PCR). Nucleic Acids Res. 18, 6069–6074.
Consortium, U. (2008). The universal protein resource (UniProt). Nucleic Acids Res. 36,
D190–D195.
Finn, R. D., Tate, J., Mistry, J., Coggill, P. C., Sammut, S. J., Hotz, H. R., Ceric, G.,
Forslund, K., Eddy, S. R., Sonnhammer, E. L., and Bateman, A. (2008). The Pfam
protein families database. Nucleic Acids Res. 36, D281–D288.
Graslund, S., Nordlund, P., Weigelt, J., Hallberg, B. M., Bray, J., Gileadi, O., Knapp, S.,
Oppermann, U., Arrowsmith, C., Hui, R., Ming, J., dhe-Paganon, S., et al. (2008a).
Protein production and purification. Nat. Methods 5, 135–146.
Graslund, S., Sagemark, J., Berglund, H., Dahlgren, L. G., Flores, A., Hammarstrom, M.,
Johansson, I., Kotenyova, T., Nilsson, M., Nordlund, P., and Weigelt, J. (2008b). The
use of systematic N- and C-terminal deletions to promote production and structural
studies of recombinant proteins. Protein Expr. Purif. 58, 210–221.
Harrison, J., Molloy, P. L., and Clark, S. J. (1994). Direct cloning of polymerase chain
reaction products in an XcmI T-vector. Anal. Biochem. 216, 235–236.
Hatley, J. L., Temple, G. F., and Brasch, M. A. (2000). DNA cloning using in vivo site-
specific recombination. Genome Res. 10, 1788–1795.
Kanehisa, M., Araki, M., Goto, S., Hattori, M., Hirakawa, M., Itoh, M., Katayama, T.,
Kawashima, S., Okuda, S., Tokimatsu, T., and Yamanishi, Y. (2008). KEGG for linking
genomes to life and the environment. Nucleic Acids Res. 36, D480–D484.
Kapust, R. B., and Waugh, D. S. (1999). Escherichia coli maltose-binding protein is uncom-
monly effective at promoting the solubility of polypeptides to which it is fused. Protein
Sci. 8, 1668–1674.
Karp, P. D., Ouzounis, C. A., Moore-Kochlacs, C., Goldovsky, L., Kaipa, P., Ahren, D.,
Tsoka, S., Darzentas, N., Kunin, V., and Lopez-Bigas, N. (2005). Expansion of the
BioCyc collection of pathway/genome databases to 160 genomes. Nucleic Acids Res. 33,
6083–6089.
Kim, Y., Dementieva, I., Zhou, M., Wu, R., Lezondra, L., Quartey, P., Joachimiak, G.,
Korolev, O., Li, H., and Joachimiak, A. (2004). Automation of protein purification for
structural genomics. J. Struct. Funct. Genomics 5, 111–118.
Klock, H. E., Koesema, E. J., Knuth, M. W., and Lesley, S. A. (2008). Combining the
polymerase incomplete primer extension method for cloning and mutagenesis with
microscreening to accelerate structural genomics efforts. Proteins 71, 982–994.
Kost, T. A., and Condreay, J. P. (2002). Recombinant baculoviruses as mammalian cell
gene-delivery vectors. Trends Biotechnol. 20, 173–180.
Lesley, S. A. (2001). High-throughput proteomics: Protein expression and purification in
the postgenomic world. Protein Expr. Purif. 22, 159–164.
Lesley, S. A., Graziano, J., Cho, C. Y., Knuth, M. W., and Klock, H. E. (2002a). Gene
expression response to misfolded protein as a screen for soluble recombinant protein.
Protein Eng. 15, 153–160.
Lesley, S. A., Kuhn, P., Godzik, A., Deacon, A. M., Mathews, I., Kreusch, A., Spraggon, G.,
Klock, H. E., McMullan, D., Shin, T., Vincent, J., Robb, A., et al. (2002b). Structural
genomics of the Thermotoga maritima proteome implemented in a high-throughput
structure determination pipeline. Proc. Natl. Acad. Sci. USA 99, 11664–11669.
Liu, Q., Li, M. Z., Leibham, D., Cortez, D., and Elledge, S. J. (1998). The univector
plasmid-fusion system, a method for rapid construction of recombinant DNA without
restriction enzymes. Curr. Biol. 8, 1300–1309.
Parallel Methods for Expression and Purification 785
Marsden, B. D., and Knapp, S. (2008). Doing more than just the structure-structural
genomics in kinase drug discovery. Curr. Opin. Chem. Biol. 12, 40–45.
Martinez Molina, D., Cornvik, T., Eshaghi, S., Haeggstrom, J. Z., Nordlund, P., and
Sabet, M. I. (2008). Engineering membrane protein overproduction in Escherichia coli.
Protein Sci. 17, 673–680.
McMullan, D., Canaves, J. M., Quijano, K., Abdubek, P., Nigoghossian, E., Haugen, J.,
Klock, H. E., Vincent, J., Hale, J., Paulsen, J., and Lesley, S. A. (2005). High-throughput
protein production for X-ray crystallography and use of size exclusion chromatography
to validate or refute computational biological unit predictions. J. Struct. Funct. Genomics 6,
135–141.
Pantoliano, M. W., Petrella, E. C., Kwasnoski, J. D., Lobanove, V. S., Myslik, J., Graf, E.,
Carver, T., Asel, E., Springer, B. A., Lane, P., and Salemme, F. R. (2001). High-density
miniaturized thermal shift assays as a general strategy for drug discovery. J. Biomol. Screen.
6, 429–440.
Peng, S., Sommerfelt, M., Logan, J., Huang, Z., Jilling, T., Kirk, K., Hunter, E., and
Sorscher, E. (1993). One-step affinity isolation of recombinant protein using the bacu-
lovirus/insect cell expression system. Protein Expr. Purif. 4, 95–100.
Shuman, S. (1994). Novel approach to molecular cloning and polynucleotide synthesis using
vaccinia DNA topoisomerase. J. Biol. Chem. 269, 32678–32684.
Studier, F. W. (2005). Protein production by auto-induction in high density shaking
cultures. Protein Expr. Purif. 41, 207–234.
Tropea, J. E., Cherry, S., and Waugh, D. S. (2009). Expression and purification of soluble
His(6)-tagged TEV protease. Methods Mol. Biol. 498, 297–307.
Waldo, G. S., Standish, B. M., Berendzen, J., and Terwilliger, T. C. (1999). Rapid protein-
folding assay using green fluorescent protein. Nat. Biotechnol. 17, 691–695.
C H A P T E R F O R T Y- T W O
Contents
1. Introduction 788
2. Anaerobic Isolation of 4Fe-FNR 790
2.1. Principles and procedures for purifying O2-labile 4Fe-FNR 790
2.2. Protocols 794
2.3. Isolation of protein for special applications 797
3. Characterization of [4Fe–4S]2þ Cluster Containing FNR 799
3.1. Fe and sulfide analysis 800
3.2. ICP-MS analysis of the 57Fe content in the 57Fe
labeled 4Fe-FNR 801
3.3. UV–visible spectrophotometric analysis of the [4Fe–4S]2þ
cluster 802
3.4. Mössbauer spectroscopic analysis 802
3.5. Low-temperature EPR 802
4. Summary 803
References 803
Abstract
Many key enzymes in biological redox reactions require metal centers or
cofactors for optimum activity and function. While the metal centers provide
unique properties for protein structure and function, some also render protein
activity sensitive to environmental O2 and cause experimental challenges to
isolation and biochemical analysis. Iron–sulfur (Fe–S) clusters represent an
important class of such metal centers and Fe–S proteins are widely distributed
in nature. Here, we utilize FNR, a regulatory Fe–S protein from Escherichia coli,
as an example to describe the techniques essential to purifying O2-labile
proteins and summarize various approaches for their biochemical analysis.
These methods can be readily adapted to purify other O2-labile proteins and
advance our understanding of this interesting class of proteins.
* School of Biological Sciences, The University of Hong Kong, Hong Kong, Hong Kong SAR
{
Department of Biomolecular Chemistry, University of Wisconsin, Madison, Wisconsin, USA
787
788 Aixin Yan and Patricia J. Kiley
1. Introduction
Many key proteins in biology contain O2-labile metal centers or
cofactors. Such proteins are found widely distributed in various physiologi-
cal processes in nature, such as oxidoreductases in the life-sustaining process
of respiration or photosynthesis, nitrogenases for N2 fixation, dehydratases
in metabolic pathways, and DNA helicases in DNA replication and repair,
etc. (Beinert et al., 1997; Brzoska et al., 2006; Kiley and Beinert, 2003;
Meyer, 2008). While the presence of such metal centers or cofactors
provides extraordinary diversity in protein structure and function, it may
also render the activity of a protein sensitive to oxidants, such as atmo-
spheric oxygen (O2) or its reactive oxygen species. These properties present
technical challenges in isolating proteins in their active state and in
performing subsequent in vitro analyses.
Creating and maintaining an anaerobic environment throughout the
entire process of protein purification is the key to obtain active O2-labile
proteins. Successful methods for isolating O2-sensitive proteins in their
active state were first reported in the 1960s and employed sealed benchtop
glove boxes in which O2 was removed by evacuating and flushing with an
inert gas (Chase and Rabinowitz, 1968; Kajiyama et al., 1969; Vandecasteele
and Burris, 1970). This type of device, while significantly advancing the
study of O2-labile metalloproteins, was limited in its ability to maintain an
O2-free environment in procedures requiring sequential transfers of mate-
rials into the chamber. These early glove boxes were also limited in the type
of equipment that could be enclosed in a chamber, such as those required
for centrifugation, chromatography, spectroscopy, etc. As a result, the yield
of protein isolated was often low and the purified proteins often exhibited
heterogeneity, resulting in some ambiguity in the subsequent biophysical
and biochemical characterizations (Gotto and Yoch, 1982; Saari et al.,
1984). The invention and construction of the anaerobic laboratory chamber
in the late 1970s solved many of these problems by providing greater
flexibility in adapting many standard laboratory techniques for anaerobic
use (Gunsalus et al., 1980; Poston et al., 1971). Introduction of this device
has greatly advanced the isolation and characterization of O2-labile proteins
in the last several decades and has led to an exponential increase in our
knowledge of this class of proteins.
In this chapter, we describe the methods used in our laboratory to isolate
the O2-sensitive iron–sulfur (Fe–S) protein, FNR, emphasizing current
practices for creating and maintaining an O2-free environment for the
purification and study of O2-labile proteins. Fe–S proteins are an ancient
class of proteins that exist across all biological kingdoms (Beinert et al., 1997;
Brzoska et al., 2006; Johnson et al., 2005; Kiley and Beinert, 2003;
Purification of O2-Labile Proteins 789
Meyer, 2008) and many contain O2-sensitive Fe–S clusters. Fe–S proteins
contain tetrahedral iron usually coordinated with sulfur, provided either by
inorganic sulfide or the thiolates of cysteine residues in the polypeptide
backbone. The basic types of these metal centers include the mononuclear
[Fe(Cys)4] complex and the bi-, tri-, and tetranucleated [2Fe–2S], [3Fe–4S],
and [4Fe–4S] clusters (Meyer, 2008), although more complex cluster types
can be found in nitrogenases and hydrogenases. While the cluster types are
relatively simple in structure, the protein environment surrounding these Fe–S
clusters can be remarkably different, resulting in great variation in cluster
stability or the redox potential of Fe–S centers (Dey et al., 2007). In addition
to their well-known roles in catalyzing electron transfer reactions, the known
functions of Fe–S proteins have expanded to also include regulatory factors that
mediate transcriptional and translational control of gene expression (Cabiscol
et al., 2000; Crack et al., 2008; Kiley and Beinert, 2003). Examples include the
well-characterized c-aconitase in eukaryotes, which binds specific RNA
sequences (iron regulatory elements) when it is in its apoform (the cytosolic
apoform of c-aconitase is also called iron regulatory protein, IRP1) and
controls protein translation in response to iron status (Kennedy et al., 1992),
and several bacterial transcription factors, such as FNR, IscR, and SoxR. FNR
and perhaps IscR appear to exploit the stability of their Fe–S clusters to sense
environmental signals and regulate the expression of a large number of genes in
E. coli ( Johnson et al., 2005; Kiley and Beinert, 2003).
In E. coli, FNR is a global transcription factor that mediates the transition
from an aerobic to anaerobic lifestyle by altering the expression profile of
hundreds of genes under anaerobic growth conditions (Constantinidou
et al., 2006; Grainger et al., 2007; Kang et al., 2005; Salmon et al., 2003).
The ability of FNR to directly and rapidly sense O2 levels is attributed to its
[4Fe–4S]2þ cluster that is ligated to FNR through four essential cysteine
residues: Cys20, Cys23, Cys29, and Cys120 (Khoroshilova et al., 1995).
This form of the protein is referred to as 4Fe-FNR. Under anaerobic
conditions, the [4Fe–4S]2þ cluster promotes FNR dimerization and
subsequent site-specific DNA binding and transcription regulation
(Lazazzera et al., 1993, 1996). Upon a switch to aerobic conditions, O2
rapidly reacts with the [4Fe–4S]2þ cluster and converts it to a [2Fe–2S]2þ
cluster. The resulting protein is referred to as 2Fe-FNR. The 2Fe-FNR is
unable to dimerize, and thus is inactive in site-specific DNA binding and
transcription regulation (Khoroshilova et al., 1997; Sutton et al., 2004).
Since the conversion from a [4Fe–4S]2þ cluster to a [2Fe–2S]2þ cluster by
O2 is sufficient to inactivate FNR in vivo and in vitro, isolation of the natively
active FNR must be carried out exclusively under anaerobic conditions.
Here, we describe the techniques and protocols necessary to isolate
native FNR protein in high yield and cluster occupancy that is suitable
for kinetic and various spectroscopic studies. We also summarize various
790 Aixin Yan and Patricia J. Kiley
Figure 42.1 Diagram of a sparging station connected to an argon tank. A, argon tank;
B, copper-furnace apparatus; C, manifold attached with gas dispersion tubes or needles.
Purification of O2-Labile Proteins 791
(Upchurch Scientific) or needles (Fig. 42.1). Standard needles are also used
to vent gas during the sparging. The glove-box chamber provides sufficient
space for storage and operation during the process of protein purification.
The vinyl chamber is filled with a gas mixture composed of 80% N2, 10%
CO2, and 10% H2 such that any O2 that enters the chamber will be reduced
by H2 using a palladium catalyst and the generated H2O will be absorbed by
a desiccant. To purify proteins, FPLC or HPLC purification systems are
advantageous because the seals are resistant to O2 diffusion and the newer
designs are sufficiently compact to operate within a chamber (assuming
minimal heat production). In addition, their largely automated design
makes it easy to manipulate within the chamber. However, the design of
our older model FPLC requires that most parts including the pump,
column, and detection system are placed outside of the chamber and
connected to the buffer and fraction collector inside the chamber through
PEEK tubing (Upchurch Scientific) (Fig. 42.2), which is impermeable to
O2, so that the entire contents of the system remain anaerobic. The path of
solutions through the FPLC is from liquid containers located inside of the
chamber, to the pumps and/or column, located outside of the chamber, and
then back into the chamber to the fraction collector (Fig. 42.2). Most
solutions such as buffers are first sparged with argon gas for a minimum of
20 min before introduction into the anaerobic chamber via the airlock and
then equilibrated for at least 12 h in the chamber before use. All glassware
and plastic items are equilibrated in the anaerobic chamber for a minimum
of 24 h before being used to ensure any O2 inadvertently introduced into
the chamber is reduced.
E B
C A
anaerobic chamber through two cycles in the airlock. The French press cell
is immediately rinsed twice with anaerobic buffer. Typically, the cell pellet
is resuspended (to 0.5% of the original volume) in anaerobic buffer A
supplemented with dithionite (DTH) and dithiothreitol (DTT) (see
Table 42.1) and then pipetted immediately into the rinsed cold French
press cell. The French press cell containing the cell suspension is sealed and
the sample outlet tubing is fitted with a capped 18-gauge needle before the
entire cell is removed from the anaerobic chamber. Once the cell is
assembled onto the French press, the needle is inserted into a butyl rub-
ber-stoppered glass vial, previously sparged with argon gas, for collection of
the cell lysate. After a single pass through the French press at 20,000 psi, the
anaerobic lysate is collected into the sealed glass vial and is immediately
brought back into the anaerobic chamber, transferred into ultracentrifuge
tubes and sealed with a gas-tight lid before being removed from the
chamber. Cell debris is removed by ultracentrifugation (Beckman Optima
LE-80K, 70.l Ti rotor) at 4 C for 60 min at 45,000 rpm. The samples are
then returned to the chamber, where the supernatant is used as the source of
4Fe-FNR for protein isolation.
Buffer Composition
A (low salt buffer) 50 mM phosphate (pH 6.8), 10% glycerol,
0.1 M KCl
B (high salt buffer) 50 mM phosphate (pH 6.8), 10% glycerol,
1 M KCl
C (no salt buffer) 50 mM phosphate (pH 6.8), 10% glycerol
D (elution buffer) 50 mM phosphate (pH 6.8), 10% glycerol,
0.4 M KCl
E (buffer to dissolve DTH) 0.1 M Tris–HCl (pH 7.9)
794 Aixin Yan and Patricia J. Kiley
2.2. Protocols
A step-by-step protocol for isolating 4Fe-FNR is summarized as follows1
Day 1
1. Inoculate 100 ml of Luria Broth containing 1 ml of 20% glucose and
50 ml of 200 mg/ml Ampicillin (Ap) with strain PK872 in the early
afternoon and grow overnight at 37 C.
2. Prepare and autoclave four 2-l flasks containing 1 l of 1 M9 (Miller,
1972) minimal media.
Day 2
1. To each 1-l flask of M9-minimal medium, add:
10 ml 20% glucose
10 ml 20% casamino acids
2 ml vitamin B1 (thiamine) (2 mg/ml stock)
1 ml MgSO4 (1 M stock)
100 ml CaCl2 (1 M stock)
1 ml ferric ammonium citrate (10 mg/ml stock, filter sterilized)
250 ml Ap (200 mg/ml stock, filter sterilized)
2. For each of the four flasks, inoculate 10 ml of the overnight culture into
1 l of the complete M9 media. Grow at 37 C with shaking at 200 rpm
to an OD600 of 0.3. Induce FNR synthesis by adding 1 ml of 0.4 M
IPTG to each flask (final concentration is 0.4 mM ) and shaking for 1
additional hour.
1
This protocol was created and amended by several members in the Kiley group: Lazzaera, B., Khoroshilova,
K., Voet, K., Sutton, V., Moore, L., Mettert, E., Yan, A.
Purification of O2-Labile Proteins 795
3. Pool cells into a 9-l flask and sparge cells with argon gas at 4 C for 14–16 h.
4. Prepare solutions and glassware that will be used next day and bring
them into the anaerobic chamber: (the chamber is normally equipped
with standard pipetting devices and tips, a microcentrifuge, a vortex
machine and eppendorf tubes.)
(a) Solutions:
Buffer A: prepare 600 ml, divide into a 500- and 100-ml screw-
capped glass bottles (recipe as shown in Table 42.1)
Buffer B: prepare 600 ml, divide into a 500- and 100-ml screw-
capped glass bottles (recipe as shown in Table 42.1)
Buffer C: 100 ml in a screw-capped glass bottle (recipe as shown in
Table 42.1)
Buffer E: 100 ml in a screw-capped glass bottle (recipe as shown in
Table 42.1)
Water: 100 ml bottle
Sparge solutions with argon gas for 20 min by adjusting the argon flow to
produce steady gas-flow through the custom-built sparging system as
described in 2.1 (Fig. 42.1) with the gas dispersion tubes (Upchurch Scientific)
placed at the bottom of the buffer solutions.
(b) Glassware and plastic items to be brought into anaerobic chamber:
Two ultracentrifuge tubes and caps
Fraction collector tubes
Beaker for waste
Superloop with piston and cap
Glass vials (5–6 small, 2 medium) and butyl rubber caps
Glass Pasteur pipettes
Four to six centrifuge bottles and caps with O rings
500 ml graduated cylinder
2 and 4-l plastic beakers
25 ml glass pipettes
Pasteur pipette bulb
Needle and tubing for French Press
Bring these items into anaerobic chamber through air lock, leave the
buffers uncapped and equilibrate for 12 h.
Day 3
1. Pour sparged cells from the 9-l flask to a plastic 4-l beaker and immedi-
ately cover it with Saran wrap and tape to seal. Bring the beaker into
the anaerobic chamber and aliquot 450 ml into each of the 500-ml
centrifuge bottles and seal the bottles with the lids.
2. Bring the bottles out of the anaerobic chamber and measure the weight
to balance. Unbalanced bottles are brought back to the anaerobic
chamber to adjust the weight. Spin at 8000 rpm for 15 min at 4 C.
796 Aixin Yan and Patricia J. Kiley
3. After centrifugation, bring the bottles back to the chamber and remove
the supernatant. Repeat the centrifugation step until all cells are cen-
trifuged. The cell pellet should have an army green color.
4. During the centrifugation steps, weigh 154 mg dithiothreitol (DTT) and
150 mg dithionite (DTH) into individual 1.5 ml eppendorf tubes and
seal tightly. Bring them into the chamber together with the bottles after
centrifugation.
5. In the chamber, add 1.5 ml buffer E to DTH and 1.0 ml H2O to DTT.
Add 300 ml of DTH (final concentration 1.7 mM) and 100 ml of DTT
(final concentration 1.0 mM) to the 100 ml bottle of buffer A. Resus-
pend the cells in 20 ml of buffer A containing 1.7 mM DTH and 1.0 mM
DTT. The cell suspension can either be used immediately for cell lysis or
transferred into a medium glass vial, sealed with a butyl rubber stopper,
frozen on dry ice and then removed from the chamber and stored in a
80 C freezer until ready for purification.
Day 4
1. To the 500 ml of buffers A and B, add 1.5 ml of DTH and 0.5 ml of
DTT (both prepared freshly as described above).
2. Place pump leads into the buffers such that buffer A is connected to
pump A and buffer B is connected to pump B.
3. Connect BioRex-70 column to FPLC system, making sure that there
are no air bubbles in the column.
4. Run FPLC system program to wash the column with 100% buffer B
(contained within the chamber) for 1 h at a flow rate of 0.17 ml/min
followed by 100% buffer A for another 1 h.
5. Bring the frozen cells into the chamber and thaw the cell suspension
under anaerobic conditions (do not thaw cells outside the chamber as O2
will be drawn into vial).
6. Bring French press cell into the chamber. Rinse the sample cell twice
with anaerobic buffer A containing DTT and DTH. Fill cell with the
resuspended cells, seal, and attach sample outlet with the tubing fitted
with a capped 18-gauge needle, and remove from chamber along with a
butyl rubber-capped glass vial for collection of cell lysate.
7. Assemble cell onto French Press located preferably in a cold room. To
exclude O2 from the lysed extract, insert the needle from the cell into
the collection vial and insert two additional needles into the vial, using
one to gently flow argon gas and the second to vent the gas. Lyse cells at
20,000 psi at 4 C.
8. Remove needles and bring the cell lysate into anaerobic chamber and
divide the lysate into the two ultracentrifuge tubes. Bring them out of
the chamber and weigh the tubes to balance before ultracentrifugation.
Unbalanced tubes are brought back into the chamber to adjust the weight.
9. Centrifuge the lysate with 70.1 Ti rotor for 60 min at 45,000 rpm, 4 C.
Purification of O2-Labile Proteins 797
10. After centrifuging, bring tubes into the chamber, pour supernatant into
superloop and seal. Fill the other end of the superloop with buffer A
and seal. Remove the superloop from the chamber and connect super-
loop to the FPLC system.
11. Run FPLC program to load the sample and initiate the appropriate
sequence of buffer solutions to allow FNR to elute from the BioRex-
70 column.
12. Program the fraction collector located in the anaerobic chamber to
collect the column eluate.
Day 5
Concentration and storage
1. Fractions containing 4Fe-FNR are easily recognized by their green
color. Pool these fractions into a medium sized flask and dilute 1:4
with buffer C.
2. Construct a 1-ml BioRex-70 gravity column and wash with 2 column
volumes of buffer B followed by 6 column volumes of buffer C.
3. Load protein solution onto column, wash with 2 column volumes of
buffer A followed by elution with 0.8 M KCl buffer (made from buffers
B and C). Collect only colored drops.
4. Aliquot the concentrated 4Fe-FNR into 0.5 ml eppendorf tubes, which
have had their caps removed, and place tubes into a small glass vial. The
glass vial is topped with butyl rubber stoppers and crimp sealed with
aluminum caps (Bellco). The protein samples stored in the glass vial are
immediately frozen in a dry ice-ethanol slurry, which is introduced into
the anaerobic chamber though the air lock. Samples can be safely stored
in a 80 C freezer for many months using this method. Before thaw-
ing, it is critical that the sealed glass vials be placed in dry ice-ethanol
slurry during transfer to the anaerobic chamber to prevent O2 from
being drawn into the vial while thawing.
copy and because of the vital roles of Fe–S containing proteins in biological
systems, Mössbauer spectroscopy has been widely used to characterize this
class of proteins. In order to prepare 57Fe enriched 4Fe-FNR, we remove the
naturally abundant 56Fe from the growth media by passing the 1 M9 media
through a Chelex 100 (200–400 mesh, BioRad) column (40 g of resin per 1 l
of medium) and then supplement the growth media with 57Fe in the form of
ferrous ethylenediammonium sulfate. The protocol for preparing Fe-free
growth media and supplementation with 57Fe is as follows:
1. Prepare Chelex 100 column by placing 200 g of Chelex resin and
200 ml H2O into a beaker. After incubating at room temperature
for 30 min, pour resin into a 250-ml gravity column. Wash the column
with 5 column volumes of H2O before use. To regenerate column, wash
with 2 column volumes of 1 N HCl followed by 5 column volumes of
H2O; then wash with 2 column volumes of 1 N NaOH and another
5 column volumes of H2O till the pH of the column eluate is 7.0.
2. Deferrate the growth media: Pass 5 l of 1 M9 media through the newly
prepared or regenerated Chelex 100 column. Collect the flow-through
and autoclave. After use, the column is stored in 0.5 M ammonium
sulfate.
3. Cell growth and isolation of FNR containing 57Fe labeled [4Fe–4S]2þ
cluster:
Inoculate 100 ml of Luria Broth with strain PK872. Add 1 ml of 20%
glucose and 50 ml of 200 mg/ml Ampicillin to the medium and grow
overnight at 37 C.
Centrifuge the overnight cell culture to remove the LB media. Add
100 ml of deferrated 1 M9 media to resuspend the cell pellet,
centrifuge at 6000 rpm at 4 C and decant the supernatant.
Resuspend the pellet in 100 ml of deferrated 1 M9 media, and
inoculate 10 ml of cell suspension to 1 l of deferrated M9-minimal
medium containing the same supplements as described earlier, except
Purification of O2-Labile Proteins 799
tuberculosis (Carroll et al., 2005), although it is not yet clear how generally
applicable this method will be.
Reagent Composition
a 1.35 g of sodium dodecyl sulfate dissolved in 30 ml of H2O and
mixed with 0.45 ml of saturated Fe-free sodium acetate
ba 270 mg of ascorbic acid and 9 mg of sodium metabisulfate
(Na2S2O5) dissolved in 5.6 ml of H2O and mixed with 0.4 ml of
saturated acetate
c 18 mg of Ferene in 1 ml of H2O
a
Reagent b is not stable and has to be remade about every 2 weeks or kept frozen.
2
The Fe and S analysis of our protein preparations were performed by our long-time collaborator and friend
Dr. Helmut Beinert who passed away on December 21, 2007.
Purification of O2-Labile Proteins 801
strong acid and/or with the presence of an oxidizing agent (such as FeCl3).
The preventative measures include: (1) perform the entire procedure in a
small glass tube with a 10-mm o.d. and a total volume of 1.7 ml that has a
conical bottom and is fitted with a stir bar (10 3 3 mm) and a glass
stopper; (2) use gentle mixing in steps where the solution is acidic or when an
oxidizing agent is added. This approach minimizes the agitation of the liquid
in the tube and the formation of new air–water interfaces which facilitates
the escape of H2S if it is formed; (3) add a few drops of phenolphthalein in the
solution to follow the efficiency of mixing by gentle mixing; (4) strongly acid
DMPD solution is added to the zinc hydroxide suspension using a micropi-
pet to underlay this heavy solution under the lighter one. The typical
procedure is as follows: prior to the measurement, the glass tubes and H2O
are brought into the anaerobic chamber to equilibrate for 24 h. Then 10 ml of
protein sample and 90 ml of H2O are added to each of the glass tubes, the cap
is added and the samples are removed from the anaerobic chamber. With the
tubes secured above a stir plate, 300 ml of freshly prepared 1% zinc acetate
[Zn(C2H3O2)2] is added and gently stirred. Then 15 ml of 12% NaOH is
added immediately and stirred vigorously until the schlieren of NaOH
disappears or the phenolphthalein color is homogeneous. After incubation
at room temperature for 5 min, 75 ml of 0.1% DMPD solution (in 5 N HCl)
is layered under the protein solution. The solution is gently stirred until only
the top 2-mm layer of liquid has undissolved zinc hydroxide or pink color.
Then 10 ml of 23 mM FeCl3 (in 1.2 N HCl) is added directly to the solution
and quickly stirred to achieve a homogeneous (colorless) solution. After
incubation for 30 min at 20–25 C, the tube is centrifuged for 15–20 min
until protein is packed. The supernatant is transferred to a cuvette and the
absorption values at 670, 710, and 750 nm are measured. The ratio of the
absorbance should be approximately A670:A750:A710 ¼ 3:2:1. The sulfide
content is determined from the extinction coefficient of 34,500 M 1 cm 1
at 670 nm (Beinert, 1983).
4. Summary
In summary, by adapting common procedures to those that can be
carried out in the absence of O2, we routinely isolate 4Fe-FNR from E. coli
with high purity and cluster occupancy. It is noted that special precautions
must be taken during the entire process of the protein purification, includ-
ing cell growth, protein isolation, and protein storage, in order to achieve a
satisfactory yield of this class of O2-labile proteins. The procedures we
routinely use are based upon well-established approaches for eliminating
O2 exposure and maintaining a reduced environment and thus should be
applicable for isolation and/or characterization of other O2-labile proteins.
However, modification of this approach may be required in order to
optimize expression of a different O2-labile protein, particularly, if it has a
different cofactor or if the cluster cannot be inserted posttranslationally as it
is with FNR (Sutton and Kiley, 2003). Other alternatives to studies of O2-
sensitive metal center proteins include purifying the apoform of the protein
under aerobic conditions and reconstituting the O2-labile centers through
appropriate in vitro systems (Crack et al., 2008) or use of O2-stable protein
variants (Bates et al., 2000; Kiley and Reznikoff, 1991). However, it is an
open question whether reconstituted Fe-S protein exhibits all of the same
properties of natively isolated protein (Saunders et al., 2008). Nonetheless,
there is no doubt that purification and characterization of more O2-labile
proteins and enzymes will help to further broaden our knowledge of this
interesting class of proteins.
REFERENCES
Bates, D. M., Popescu, C. V., Khoroshilova, N., Vogt, K., Beinert, H., Munck, E., and
Kiley, P. J. (2000). Substitution of leucine 28 with histidine in the Escherichia coli
transcription factor FNR results in increased stability of the [4Fe–4S]2þ cluster to oxygen.
J. Biol. Chem. 275, 6234–6240.
Beinert, H. (1983). Semi-micro methods for analysis of labile sulfide and of labile sulfide plus
sulfane sulfur in unusually stable iron–sulfur proteins. Anal. Biochem. 131, 373–378.
804 Aixin Yan and Patricia J. Kiley
Beinert, H., Holm, R. H., and Münck, E. (1997). Iron–sulfur clusters: Nature’s modular,
multipurpose structures. Science 277, 653–659.
Brzoska, K., Meczynska, S., and Kruszewski, M. (2006). Iron–sulfur cluster proteins:
Electron transfer and beyond. Acta Biochim. Pol. 53, 685–691.
Cabiscol, E., Tamarit, J., and Ros, J. (2000). Oxidative stress in bacteria and protein damage
by reactive oxygen species. Int. Microbiol. 3, 3–8.
Cammack, R., and Cooper, C. E. (1993). Electron paramagnetic resonance spectroscopy of
iron complexes and iron-containing proteins. Methods Enzymol. 227, 353–384.
Carroll, K. S., Gao, H., Chen, H., Leary, J. A., and Bertozzi, C. R. (2005). Investigation of
the iron–sulfur cluster in Mycobacterium tuberculosis APS reductase: Implications for sub-
strate binding and catalysis. Biochemistry 44, 14647–14657.
Chase, T. Jr., and Rabinowitz, J. C. (1968). Role of pyruvate and S-adenosylmethioine in
activating the pyruvate formate-lyase of Escherichia coli. J. Bacteriol. 96, 1065–1078.
Constantinidou, C., Hobman, J. L., Griffiths, L., Patel, M. D., Penn, C. W., Cole, J. A., and
Overton, T. W. (2006). A reassessment of the FNR regulon and transcriptomic analysis
of the effects of nitrate, nitrite, NarXL, and NarQP as Escherichia coli K12 adapts from
aerobic to anaerobic growth. J. Biol. Chem. 281, 4802–4815.
Crack, J. C., Le Brun, N. E., Thomson, A. J., Green, J., and Jervis, A. J. (2008). Reactions of
nitric oxide and oxygen with the regulator of fumarate and nitrate reduction, a global
transcriptional regulator, during anaerobic growth of Escherichia coli. Methods Enzymol.
437, 191–209.
Dey, A., Jenney, F. E. , Adams, M. W., Babini, E., Takahashi, Y., Fukuyama, K.,
Hodgson, K. O., Hedman, B., and Solomon, E. I. (2007). Solvent tuning of electro-
chemical potentials in the active sites of HiPIP versus ferredoxin. Science 318, 1464–1468.
Gotto, J. W., and Yoch, D. C. (1982). Purification and Mn2þ activation of Rhodospirillum
rubrum nitrogenase activating enzyme. J. Bacteriol. 152, 714–721.
Grainger, D. C., Aiba, H., Hurd, D., Browning, D. F., and Busby, S. J. (2007). Transcrip-
tion factor distribution in Escherichia coli: Studies with FNR protein. Nucleic Acids Res. 35,
269–278.
Gunsalus, R. P., Tandon, S. M., and Wolfe, R. S. (1980). A procedure for anaerobic column
chromatography employing an anaerobic Freter-type chamber. Anal. Biochem. 101,
327–331.
Johnson, D. C., Dean, D. R., Smith, A. D., and Johnson, M. K. (2005). Structure, function,
and formation of biological iron–sulfur clusters. Annu. Rev. Biochem. 74, 247–281.
Kajiyama, S., Matsuki, T., and Nosoh, Y. (1969). Separation of the nitrogenase system of
azotobacter into three components and purification of one of the components. Biochem.
Biophys. Res. Commun. 37, 711–717.
Kang, Y., Weber, K. D., Qiu, Y., Kiley, P. J., and Blattner, F. R. (2005). Genome-wide
expression analysis indicates that FNR of Escherichia coli K-12 regulates a large number of
genes of unknown function. J. Bacteriol. 187, 1135–1160.
Kennedy, M. C., Kent, T. A., Emptage, M., Merkle, H., Beinert, H., and Münck, E. (1984).
Evidence for the formation of a linear [3Fe-4S] cluster in partially unfolded aconitase.
J. Biol. Chem. 259, 14463–14471.
Kennedy, M. C., Mende-Mueller, L., Blondin, G. A., and Beinert, H. (1992). Purification
and characterization of cytosolic aconitase from beef liver and its relationship to the
iron-responsive element binding protein. Proc. Natl. Acad. Sci. USA 89, 11730–11734.
Khoroshilova, N., Beinert, H., and Kiley, P. J. (1995). Association of a polynuclear iron–
sulfur center with a mutant FNR protein enhances DNA binding. Proc. Natl. Acad. Sci.
USA 92, 2499–2503.
Khoroshilova, N., Popescu, C., Münck, E., Beinert, H., and Kiley, P. J. (1997). Iron–sulfur
cluster disassembly in the FNR protein of Escherichia coli by O2: [4Fe–4S] to [2Fe–2S]
conversion with loss of biological activity. Proc. Natl. Acad. Sci. USA 94, 6087–6092.
Purification of O2-Labile Proteins 805
Kiley, P. J., and Beinert, H. (2003). The role of Fe–S proteins in sensing and regulation in
bacteria. Curr. Opin. Microbiol. 6, 181–185.
Kiley, P. J., and Reznikoff, W. S. (1991). Fnr mutants that activate gene expression in the
presence of oxygen. J. Bacteriol. 173, 16–22.
Lazazzera, B. A., Bates, D. M., and Kiley, P. J. (1993). The activity of the Escherichia coli
transcription factor FNR is regulated by a change in oligomeric state. Genes Dev. 7,
1993–2005.
Lazazzera, B. A., Beinert, H., Khoroshilova, N., Kennedy, M. C., and Kiley, P. J. (1996).
DNA binding and dimerization of the Fe–S-containing FNR protein from Escherichia coli
are regulated by oxygen. J. Biol. Chem. 271, 2762–2768.
Mettert, E. L., and Kiley, P. J. (2007). Contributions of [4Fe–4S]-FNR and integration host
factor to FNR transcriptional regulation. J. Bacteriol. 189, 3036–3043.
Mettert, E. L., Outten, F. W., Wanta, B., and Kiley, P. J. (2008). The impact of O2 on the
Fe–S cluster biogenesis requirements of Escherichia coli FNR. J. Mol. Biol. 384, 798–811.
Meyer, J. (2008). Iron–sulfur protein folds, iron–sulfur chemistry, and evolution. J. Biol.
Inorg. Chem. 13, 157–170.
Miller, J. H. (ed.) (1972). In ‘‘Experiments in Molecular Genetics’’, Cold Spring Harbor
Laboratory Press, Plainview, NY.
Moore, L. J., and Kiley, P. J. (2001). Characterization of the dimerization domain in the
FNR transcription factor. J. Biol. Chem. 276, 45744–45750.
Popescu, C. V., Bates, D. M., Beinert, H., Münck, E., and Kiley, P. J. (1998). Mössbauer
spectroscopy as a tool for the study of activation/inactivation of the transcription regula-
tor FNR in whole cells of Escherichia coli. Proc. Natl. Acad. Sci. USA 95, 13431–13435.
Poston, J. M.,, Stadtman, T. C., and Stadtman, E. R. (eds.) (1971). In An Anaerobic
Laboratory, Academic Press, New York.
Raulfs, E. C., O’Carroll, I. P., Dos Santos, P. C., Unciuleac, M. C., and Dean, D. R. (2008).
In vivo iron–sulfur cluster formation. Proc. Natl. Acad. Sci. USA 105, 8591–8596.
Saari, L. L., Triplett, E. W., and Ludden, P. W. (1984). Purification and properties of the
activating enzyme for iron protein of nitrogenase from the photosynthetic bacterium
Rhodospirillum rubrum. J. Biol. Chem. 259, 15502–15508.
Salmon, K., Hung, S. P., Mekjian, K., Baldi, P., Hatfield, G. W., and Gunsalus, R. P.
(2003). Global gene expression profiling in Escherichia coli K12. The effects of oxygen
availability and FNR. J. Biol. Chem. 278, 29837–29855.
Saunders, A. H., Griffiths, A. E., Lee, K. H., Cicchillo, R. M., Tu, L., Stromberg, J. A.,
Krebs, C., and Booker, S. J. (2008). Characterization of quinolinate synthases from
Escherichia coli, Mycobacterium tuberculosis, and Pyrococcus horikoshii indicates that [4Fe–4S]
clusters are common cofactors throughout this class of enzymes. Biochemistry 47,
10999–11012.
Sutton, V. R., and Kiley, P. J. (2003). Techniques for studying the oxygen-sensitive
transcription factor FNR from Escherichia coli. Methods Enzymol. 370, 300–312.
Sutton, V. R., Mettert, E. L., Beinert, H., and Kiley, P. J. (2004). Kinetic analysis of the
oxidative conversion of the [4Fe–4S]2þ cluster of FNR to a [2Fe–2S]2þ cluster.
J. Bacteriol. 186, 8018–8025.
Vandecasteele, J. P., and Burris, R. H. (1970). Purification and properties of the constituents
of the nitrogenase complex from Clostridium pasteurianum. J. Bacteriol. 101, 794–801.
C H A P T E R F O R T Y- T H R E E
Contents
1. Introduction 809
1. Introduction
Every protein purification that you undertake should provide you not
only with purified material but also with considerable information about the
protein. Thus, during the course of purification you most likely will learn
about the stability of the protein under a variety of conditions, as well as about
its size, charge, and, perhaps, its affinity properties. You will have learned
whether the protein can be concentrated, diluted, dialyzed, or exposed to a
variety of agents. In addition, you may have prepared an antibody against the
protein, subjected it to a limited sequence analysis or mass spec analysis or
determined whether it has any covalent modifications. Finally, the sequence
information obtained from the purified protein may allow you to identify
the gene encoding it, opening up the possibility of overexpression and the
generation of mutants for physiological and mechanistic studies.
All of this information can be of great help in deciding whether you have
developed an optimal purification scheme. Obviously, in some cases you
may not care. However, if this protein is one you plan on studying in some
detail, and you can foresee many purifications ahead, a rapid and efficient
(meaning high purity and high yield) purification scheme can save you an
enormous amount of work in the long run. By taking advantage of what
you have learned about the protein, it generally should be possible to
streamline and optimize the procedure a great deal. There is a natural
tendency, especially after having spent many mouths learning to purify a
Department of Biochemistry and Molecular Biology, University of Miami School of Medicine, Miami,
Florida, USA
809
810 Murray P. Deutscher
Contents
1. Introduction 814
2. SDS Gel Electrophoresis Sample Preparation 814
2.1. Proteases at room temperature 815
2.2. Asp–Pro bond cleavage at 100 C 815
2.3. Keratin contamination in sample or SDS sample buffer 816
3. Buffers 816
3.1. Reducing agent can become an oxidizing agent 816
3.2. Contaminant in sulfonylethyl buffers 817
4. Chromatography 817
4.1. EDTA binding to anion-exchange columns 817
4.2. EDTA in sample can strip nickel from a Ni-chelate
affinity column 818
5. Protein Absorption During Filtration 818
6. Chemical Leaching from Plasticware 819
7. Cyanate in Urea 819
References 820
Abstract
Many subtle and sometimes obscure artifacts exist that can have major effects
on the outcomes of otherwise carefully performed experiments. This chapter
focuses on a few of these artifacts involved in processes such as SDS gel
sample preparation, buffers, Ni-chelate affinity and ion-exchange chromatogra-
phy, urea preparation, use of plasticware, and filtration of protein samples.
McArdle Laboratory for Cancer Research, University of Wisconsin–Madison, Madison, Wisconsin, USA
813
814 Richard R. Burgess
1. Introduction
Many of us who are experienced in protein biochemistry and protein
purification are called on regularly to troubleshoot the problems of our
colleagues. I often find that the answer to their problem is a little known bit
of knowledge that they have never heard of, but which I have encountered
many times. Some of us teach courses in protein purification and add little
tidbits of information to our lectures that we feel will help our students
avoid mysterious problems with their experiments. Many of us review
multiple manuscripts on protein purification and occasionally find that
certain artifacts reappear again and again, even when the effect has been
seen and reported many years earlier. One of the problems is that while
these sorts of artifact are many times reported, they are seldom the subject of
the paper, but merely a sentence or paragraph buried deep in the paper or
figure legend where they are rarely seen by a wider audience.
Other times, the importance has not been sufficiently emphasized in the
original paper, and the novice scientist may have heard of the problem, but
has chosen to ignore it or has assumed it really is not a problem most of the
time. Finally, young scientists often assume that there is nothing of impor-
tance in literature older than a few years old and simply are unaware of issues
that were well known 10–20 years ago.
Here, I present a few such potential artifacts and how to avoid them.
I hope that some of this obscure and often forgotten information might help
readers. These are just a few examples that come to mind, or have made a
particularly strong impression on me, but I am sure that there are many,
many more that are equally if not more important in various subareas of
protein biochemistry.
If you are aware of other such artifacts or have tips, or cautions that you
want to share, I would love to hear from you about them (burgess@oncology.
wisc.edu).
3. Buffers
3.1. Reducing agent can become an oxidizing agent
One very commonly adds a reducing agent to buffers to help prevent air
oxidation of enzymes and other proteins. It is typical to add 0.1–1 mM 1,4-
dithiothreitol (DTT) (HSCH2CH2OHCH2OHCH2SH) (Cleland, 1964).
Usually such buffers are stored in the cold and used within a few days, but
often buffers are used that are months old. The problem is that slowly over
time, the reducing agent becomes oxidized by oxygen in the air and becomes an
oxidizing agent; effectively, it is now much worse that having no thiol agent
in the buffer. How do you know whether a buffer is still good and has
effective reducing capability? A simple test is to treat a small portion of the
buffer with Ellman’s reagent (5, 50 -dithiobis-(2-nitrobenzoic acid) or
DTNB) (Ellman, 1959). This chemical reacts with free thiol groups, such
as the sulfhydryl of the DTT and releases 2-nitro-5-thiobenzoic acid, which
is yellow colored and absorbs light with a molar extinction coefficient at
Important Artifacts in Protein Chemistry 817
412 nm of 14,150 M 1 cm 1 (Riddles et al., 1983). Read the A412 nm of the
solution and a control buffer carefully prepared with fresh DTT. If all the
DTT is active then there should be two sulfhydryls per mole of DTT. Buffers
should be made fresh if more than 10–20% of the DTT is oxidized. While this
effect has been known for decades it is still surprising that this simple test is
not more commonly used. The problem is even more pronounced with
2-mercaptoethanol which is more prone than DTT to air oxidation and is
usually used at 5–50 times higher concentrations than DTT.
4. Chromatography
4.1. EDTA binding to anion-exchange columns
Recently, Robert Chumanov in my lab observed a sharp peak of
215 nm-absorbing material eluting during a salt gradient elution of a
MonoQ anion exchange resin (GE Healthcare). It turned out that it was
EDTA, bound quite tightly to MonoQ and eluting at a NaCl concentration of about
250 mM (R. Chumanov and R. Burgess, unpublished). We routinely
monitor column outputs at 215 nm when our protein/peptide lacks absorp-
tion at 280 nm or if we are loading low amounts of protein and want more
sensitivity (most proteins absorb at least 15 times better at 215 nm than
280 nm). It is not surprising that the di- or trivalent anion EDTA binds to
MonoQ column resin. We were surprised that so much bound that it
completely depleted EDTA from our low salt equilibration buffer originally
containing 0.1 mM EDTA. When the salt increased, the EDTA eluted with
a peak EDTA concentration of up to 40 mM. This could have unexpected
effects on many experiments. The fractions where the EDTA elutes will, for
example, bind Mg2þ needed for many in vitro enzyme assays (such as
transcription and replication assays) and appear to contain a mysterious
818 Richard R. Burgess
materials, such as cellulose nitrate and polystyrene are able to bind much
more material and should be avoided for filtration or storage of protein.
7. Cyanate in Urea
Urea is widely used as a protein-denaturing agent, but it is not widely
appreciated that urea solutions contain substantial amounts of ammonium
cyanate, which is in chemical equilibrium with urea (H2N)2C¼O in equilib-
rium with NH4þ þ NCO). Isocyanic acid H–N¼C¼O can react with
amino groups on proteins (e-amino group of lysine and the amino terminus,
and to a lesser extent with arginine and cysteine) to form a carbamylated
protein (Stark, 1965). This carbamylation can alter charge and in some cases
interfere with enzyme function, block certain protease cleavage reactions, and
add 43 Da per carbamylation event to the mass as measured by mass spec-
trometry. How can one diminish or prevent this from happening? One can
treat a urea solution with a mixed bed resin such as Bio-Rad AG 501-X8 to
remove these contaminant ions. The progress of deionization is easily moni-
tored by measuring conductivity of the solution. Unfortunately, this is a
chemical equilibrium and relatively soon (within a few days) the ammonium
cyanate builds back up again to levels in the 0.5–3 mM range in an 8 M urea
solution and can reach values of 20 mM (Lin et al., 2004). Certain chemical
scavengers such as ethylenediamine, glycylglycine, or glycinamide in the
5–25 mM range can also reduce cyanate to less than 0.1 mM in 8 M urea,
Tris pH 8 (Lin et al., 2004). Perhaps the best general practice if you want to
use urea is to replace some of the NaCl in the buffer with some ammonium
salt (such as 25–50 mM ammonium chloride) to push the equilibrium back by
the common ion effect toward less cyanate. Reaction of isocyanic acid with
amino groups in proteins is slowed by lower temperature and by acidic
conditions and can be minimized by restricting the exposure of the protein
to the urea solution to the shortest time possible.
820 Richard R. Burgess
REFERENCES
Cleland, W. W. (1964). Dithiothreitol, a new protective reagent for SH groups. Biochemistry
3, 480–482.
Deutsch, D. G. (1976). Effect of prolonged 100 C heat treatment in SDS upon peptide
bond cleavage. Anal. Biochem. 71, 300–303.
Ellman, G. L. (1959). Tissue sulfhydryl groups. Arch. Biochem. Biophys. 82, 70–77.
Grabski, A., and Burgess, R. R. (2001). Preparation of protein samples for SDS-polyacryl-
amide gel electrophoresis: Procedures and tips. inNovations 13, 10–12.
Lin, M.-F., Williams, C., Murray, M. V., Conn, G., and Ropp, P. A. (2004). Ion chro-
matographic quantification of cyanate in urea solutions: Estimation of the efficiency of
cyanate scavengers for use in recombinant protein manufacturing. J. Chromatogr. B 803,
353–362.
McDonald, G., Hudson, A., Dunn, S., You, H., Baker, G., Whittal, R., Martin, J., Jha, A.,
Edmondson, D., and Holt, A. (2008). Bioactive contaminants leach from disposable
laboratory plasticware. Science 322, 917.
Ochs, D. (1983). Protein contaminants of sodium dodecyl sulfate-polyacrylamide gels. Anal.
Biochem. 135, 470–474.
Pringle, J. R. (1975). Methods for avoiding proteolytic artefacts in studies of enzymes and
other proteins from yeasts. Methods Cell Biol. 12, 149–184.
Riddles, P. W., Blakeley, R. L., and Zerner, B. (1983). Reassessment of Ellman’s reagent.
Meth. Enzymol. 91, 49–60.
Sharpe, J. C., and London, E. (1997). Inadvertent concentration of EDTA by ion exchange
chromatography: Avoiding artifacts that can interfere with protein purification. Anal.
Biochem. 250, 124–125.
Smith, B. D., Soellner, M. B., and Raines, R. T. (2003). Potent inhibition of ribonuclease A
by oligo(vinylsulfonic acid). J. Biol. Chem. 278, 20934–20938.
Stark, G. R. (1965). Reactions of cyanate with functional groups of proteins: Reaction with
amino and carboxyl groups. Biochemistry 4, 1030–1036.
Volkin, D. B., Mach, H., and Middaugh, C. R. (1995). Degradative covalent reactions
important to protein stability. (B. A. Shirley, ed.)Protein stability and folding: Theory and
practiceMethods Mol. Biol. 40, 35–63.
Author Index
821
822 Author Index
Mann, M., 735, 744, 750–752, 757 Mohanty, A. K., 449, 463
Manza, L. L., 739 Molday, R. S., 635
Marblestone, J. G., 251 Monsma, S. A., 137
Margulies, M. M., 500 Mooney, R. A., 612
Marin, M., 140 Moore, L. J., 798
Markley, J. L., 649–650 Morita, E. H., 669
Markwell, J., 75, 93 Mortensen, K. K., 265
Marlor, C. W., 344 Mueller, E., 354–355
Marshak, D. R., 289, 293 Mueller, U., 244
Marshall, K., 703 Mujacic, M., 133, 249, 262, 276
Marshall, T., 507 Mukhopadhyay, A., 139, 144
Marston, F. A. O., 265 Muller, N., 232
Martin, R., 702, 705 Murby, M., 265
Martinez Molina, D., 774 Murhammer, D. W., 215
Maslennikov, I., 668 Muszynska, G., 444
Mateo, C., 456
Mathrubutham, M., 578, 590 N
Matiasson, B., 457
Matsuda, T., 649 Nagel, K., 79, 89
Matsumoto, K., 227, 649 Nagy, P. L., 477
Matthey, B., 155 Nakamura, E., 320
Mattson, D. L., 575 Nall, B., 261
Mc Donald, G., 819 Nallamsetty, S., 249, 252
M. C.slin, D. R., 703 Necina, R., 358
M. C.e, J. T., 395–396, 405–413 Nelson, P. N., 476
M. H.nry, C. S., 286 Neophytou, I., 164
M. I.vaine, T. C., 51 Nestle, M., 290
M. L.chlin, D. T., 731–735 Neu, H. C., 162
M. L.chlin, J. R., 195 Neugebauer, J. M., 289, 605, 608–609
M. M.ster, M. C., 67 Neuhoff, V., 546
M. M.ekin, T. L., 703 Neville, D. M., Jr., 498–499, 621
M. M.llan, D., 782 Nguyen, H., 287
Mechref, Y., 738 Nguyen, L., 263, 265
Medzihradszky, K. F., 737, 745 Nguyen, N. Y., 512
Meiss, G., 290 Nieba, L., 445, 448
Meissner, P., 226–227, 229 Nielsen, T. B., 499
Melander, W., 407 Nilsson, J., 457
Melchior, F., 739 Noble, J. E., 73–93
Merril, C. R., 508–509, 542, 547 Nomura, S. M., 652
Mészárosová, K., 447 Norris, J. L., 615
Mettert, E. L., 792, 797 Northcote, D. H., 86
Meyer, J., 788–789 Nosjean, O., 652
Michael, M. Z., 154 Nossal, N. G., 162
Michaelis, L., 53 Nozaki, Y., 408, 696
Michelsen, U., 305–327 Nozawa, A., 647–670
Middelberg, A. P. J., 262, 272, 286, 290–291 Nuss, J. E., 401
Mikesh, L. M., 728, 747, 749
Miles, B. J., 388 O
Miller, I., 543 O’Farrell, P. H., 515–516, 519
Miller, J. V., 326 O’Farrell, P. Z., 523
Miller, L. K., 193–195, 207 O’Reilly, D. R., 196, 198, 215–216
Minchiotti, M., 427 O’Shannessy, D. J., 432
Minden, J., 556 Oates, M. R., 432
Minton, A. P., 699–700 Ochs, D., 510, 816
Miroux, B., 154, 164 Oda, Y., 729, 733
Mistretta, T. A., 195 Oganesyan, N., 276
Miyagi, M., 751–752 Ogawa, T., 390
Moffatt, B. A., 154, 156 Ohana, R. F., 428
Author Index 829
A AmpholinesÒ , 519
Analytical size exclusion chromatography
Absolute quantitation (AQUA), 751 (ANSEC), 777
A431 cells (ATCC), 590 Antibody-affinity chromatography, 627
Acetate buffer, 51 Arginine, 273
Affinity chromatography Assay selection quantitation, 75–76
amine reactive linkages, 432 Autoinduction media protocol, 160
covalent attachment, 430
ligand selection B
characterization, 424–425
covalent affinity chromatography, 428–429 Bacterial systems, heterologous proteins
de novo development, 423 production
group specific ligands, 425 analysis
lectins, 426–427 autoinduction method, 159–160
specificity, 426 expression, 159
matrix selection solubility, 158–159
examples of, 421 autoinduction media protocol, 160
magnetic affinity beads, 422–423 cloning
pore size, 419 expression hosts and vectors, 153–155
Renkin equation, 420 ligation independent cloning (LIC)
specific ligand, 420 methods, 153
stability, 422 cytoplasmic expression, 156
purification method cytoplasmic targets expression and solubility,
elution, 434–435 157
GPCRs, 634–636 Escherichia coli, 150–151
sample preparation, 433 gram-negative and positive bacterium,
surface activation, 430 164–165
Affinity tags osmotic shock protocol, 162–164
basis, 243 periplasmic targets expression, 160–162
calmodulin-binding peptide (CBP)-tag, 248–249 planning and sequential steps, 151–152
epitope, 247–248 production scale, 166
GST, 246–247 project requirements evaluation, 152
His-tag, 245–246 target analysis, 152–153
S-tag, 248 T4 DNA polymerase-treated DNA fragments,
STREP-II, 248 155–156
Alkaline copper reduction assays. See Lowry assay vector and induction conditions, 165
Amine derivatization assay, 89–91 Bacteria lysis, 289
Amino acid sequence. See Bioinformatics, protein BaculoDirect approach, 205–206
purification Baculovirus–insect cell expression systems
2-Amino-2-methyl-1,3-propanediol (ammediol) baculovirus biology and molecular biology,
buffer, 54–55 193–195
Amino-terminal epitope tags, 739 basic protocols
Ammonium sulfate (AS) precipitation genomic DNA isolation, 215–216
concentration, 335 insect cell maintenance, 215
principles, 332–334 plaque assays, 216–218
problems/solutions, 337 conditions and methods, 211–212
procedure, 334 expression vector, 195–196
solubility curve, 333 foreign gene deliver and protein
test, 336 encoding, 211
835
836 Subject Index
C protocol, 659–660
reagents, 659
Cacodylate buffer, 52–53 proteoliposomes characterization, 667–668
Calcium-phosphate-mediated transfection, 229 scale-up considerations, 669
Calmodulin-binding peptide (CBP)-tag, 248–249 steps, 650–651
Carbonate–bicarbonate buffer, 55–56 transformation reaction
Carboxymethyl aspartate (CM-Asp), 442 materials and reagents, 658
Catalytic activity, principles protocol, 658–659
cell-free translation, 667–668 wheat germ translation reaction
chemical kinetics bilayer reaction protocol, 664
basic kinetic mechanism, 58–59 bilayer reaction vs. dialysis reaction,
first-order rate constant, 60–61 662–663
rate law, order, and rate constant, k, dialysis reaction protocol, 664–665
59–60
materials and reagents, 663–664
reactant concentration, 61–62
Ceramic fluoroapatite (CFT), 391
reaction velocity, 58, 60
Ceramic hydroxyapatite type I (CHT I),
enzyme kinetics
391, 397
initial velocity, 64
CFT. See Ceramic fluoroapatite
Michaelis constant, 63–64
Chaotropic agents, 273
steady-state approximation, 63
Charged-coupled device (CCD) cameras, 575
turnover number, 62
Chemical additives, 273
Cell disruption
Chemical and enzymatic cell disruption
buffer composition
bacteria lysis, 289
basic criteria, 294
high-throughput (HT) extraction, 287
protease inhibitors, 295
lytic enzyme treatments, 290
volume, 294, 296
reagents and enzymes, 288
chemical and enzymes, 287–290
yeast lysis, 289
bacteria lysis, 289
Chemiluminescence signal
high-throughput (HT) extraction, 287
signal capture, 583–584
lytic enzyme treatments, 290
signal intensity and duration, 584–585
reagents, 288
Western blot optimizing
yeast lysis, 289
antibodies, 586–587
E. coli cell lysis, 296–300
blotting membrane, 585
high-pressure homogenizers mechanism,
detection method, 587–588
290–292
Cell-free translation, integral membrane proteins signal reduction, chemical treatment, 587
density gradient ultracentrifugation, 665–667 target protein, 585–586
expression vectors, 651–652 CHO cell lines. See HEK293 cell lines
flexivector and PCR product digestion CHT I. See Ceramic hydroxyapatite type I
reaction Citrate-phosphate buffer, 50–52
protocol, 657 Cleavage sites, 245
reagents, 656 Click chemistry reagents, 433, 559
gene cloning Cloning, bacterial systems
materials and reagents, 654 expression hosts and vectors, 153–155
protocol, 654–655 ligation independent cloning (LIC) methods,
sacB-CAT cassette, 652–653 153
isotopic labeling, 669 Cloud point and phase separation, 611–612
ligation reaction, 657–658 CMC. See Critical micellization concentration
liposomes preparation CNBr. See Cyanogen bromide
materials and reagents, 661 Collision induced dissociation (CID), 747
protocol, 661–662 Column chromatography, 708
mRNA preparation Conacanavalin A (ConA), lectins, 427
protocol, 661 Continuous assays, 65–67
reagents, 660 Coomassie blue assay
PCR product cleanup observation, 86
materials and reagents, 655–656 procedure, 85–86
protocol, 656 reagents, 85
plasmid DNA purification Coomassie brilliant blue (CBB), 546–547
838 Subject Index