Bio 2
Bio 2
2 Bioenergetics 3
2.1 Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1.1 Forms of Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2 Laws of Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2.1 Conservation of Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2.2 Entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.3 Gibbs Free Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.3.1 Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3 Proteins 8
3.1 Primary Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.1.1 Amino Acids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.2 Secondary Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.2.1 α Helices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.2.2 β Pleated Sheets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.3 Tertiary Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.4 Quaternary Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4 Enzymes 14
4.1 Enzyme Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
4.2 Catalysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.2.1 Energy of Activation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
4.3 Enzymatic Regulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4.3.1 Types of Inhibition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.3.2 Allosteric Regulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.3.3 Cooperativity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.3.4 Feedback Inhibition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.4 Enzyme Kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.4.1 Michaelis-Menten Kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.4.2 Lineweaver-Burk Plot . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.4.3 Effects of Inhibition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
5 Conclusion 27
5.1 Closing Statement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2
USABO Guide Andrew Ung
1 Introduction
1.1 Opening Statement
Biochemistry is a captivating blend of biology and chemistry, and certain principles found in
biochemistry are needed to fully understand the background of life’s most complex biological
processes. The laws of energy which govern our universe, the intricate folding processes of the
proteins which make cellular life possible, the complex ways in which enzymes allow everyday
reactions to be possible, and more essential topics all play their own part in understanding biology
and the phenomenon of life.
Although this paper is structured in a way that allows all readers to grasp the main concepts
and topics, it may be recommended to read this document after finishing the first biochemistry
handout.
2 Bioenergetics
Bioenergetics is defined as the study of energy flow and energy transformations within living
organisms. In order to understand how complex life is able to occur, we must understand the
primary concepts which dictate how energy is and can be manipulated.
2.1 Energy
Energy, at its simplest definition, is the ability or capacity to do work. Energy can have various
forms, all of which have their own definitions and functions in everyday life.
Thermal energy is a type of kinetic energy. However, thermal energy is associated with the
random movement of atoms in a collection of matter, whereas kinetic energy is associated with
a directional movement of atoms in a collection of matter.
3
USABO Guide Andrew Ung
Figure 2.2 A collection of water molecules with thermal energy. Note that, although each water
molecule has kinetic energy, their directions of movement are random, thus there is no net
movement. (Source: Andrew Ung)
Light energy and sound energy are also forms of energy, stemming from electromagnetic
radiation and the vibration of matter, respectively. There are many other types of energy not
outlined in the document due to irrelevance.
Potential energy is a stored form of energy. Potential energy is a form of energy that is
associated with the position, shape, or configuration of an object or system, ready to be converted
into other forms of energy and do work when certain conditions are met.
Imagine lifting a box up from the ground and holding it in place in the air. The box is not
moving, it has no kinetic energy, but it has a substantial amount of potential energy. By dropping
that box, that potential energy is gradually converted into kinetic energy as it speeds up in free
fall.
Figure 2.3 A box with potential energy (left). The potential energy gets converted into kinetic
energy as the box is dropped (right). (Source: Andrew Ung)
Chemical energy is a type of potential energy stored within the chemical bonds of molecules.
Recall that bonds are formed when suitable molecules and atoms approach close enough to share
or give and take electrons, forming an attraction between the two.
• In order to form a bond, energy must be released. Molecules and atoms associated with
large amounts of kinetic or thermal energy move too erratically and frequently to form bonds,
meanwhile molecules and atoms associated with small amounts of kinetic or thermal energy
move slowly enough to form bonds with other molecules and atoms.
4
USABO Guide Andrew Ung
• In order to break chemical bonds, energy must be added. Adding kinetic or thermal
energy results in movement, resulting in the bond breaking as the molecules or atoms move
too far apart to be able to share or give electrons.
Thus, the stronger the bonds within are, the less energy a molecule contains, as it requires a
large energy input to break apart. The weaker the bonds within are, the more energy a molecule
contains, as it requires less energy to break it apart.
Recall glucose, a carbohydrate molecule important to life as a store of energy. Glucose contains
a large amount of chemical potential energy, and when it is broken down, that energy is released.
The atoms in glucose are split and rearranged, eventually forming low-energy products such as
carbon dioxide (CO2 ) and water (H2 O). The bonds between atoms of CO2 and H2 O are stronger
than the bonds between the atoms of glucose, meaning they contain less energy, and thus energy
had to have been released when the initial glucose molecule was broken down into those products.
Figure 2.4 A process showing multiple energy transformations and the principle of conservation
of energy. (Source: Andrew Ung)
5
USABO Guide Andrew Ung
Energy can never be lost or gained, simply transformed. This is known as the law of conser-
vation of energy.
2.2.2 Entropy
Entropy is a measure of randomness. Entropy can be found in the arrangements of molecules,
the properties of larger collections of matter, energy dispersion, and more.
The second law of thermodynamics states that the entropy of a system tends to increase
over time. A highly concentrated cloud of gases will eventually disperse over a greater area, a
constructed building will eventually crumble and fall to the ground in small pieces, as ordered ice
crystals melt, their hexagonal structure fades into random chains of water molecules.
Figure 2.5 The inevitability of entropy. An ordered and complex house of cards will tend to fall
into a more stable, simpler arrangement. (Source: Andrew Ung)
Transformations regarding chemical energy also are affected by entropy; our common biological
processes aren’t very energy efficient and lead to the conversion of large amounts of thermal energy.
It is estimated that, when transforming chemical energy into kinetic energy in our muscles, around
60 percent of the original energy stored in chemical bonds is lost as heat.
6
USABO Guide Andrew Ung
This brings us to the topic of what is known as Gibbs free energy, hereby shortened to simply
free energy. Free energy is energy that can be used to perform meaningful work in a system
with equal temperature and pressure. This definition thus does not include thermal energy formed
from entropy, involved with pressure (thermal energy randomly moving air particles leading to
pressure) and temperature. The definition of free energy can thus be simplified to energy present
in a system that is able to perform meaningful work.
Free energy becomes useful when isolating specific chemical reactions—it tells us how much
meaningful energy is gained or lost, and that information allows us to determine other important
properties of the reaction.
∆G = ∆H − T∆S
where ∆G represents change in Gibbs free energy, ∆H represents change in enthalpy, or the
total energy of a system, T represents temperature in Kelvin, and ∆S represents change in entropy.
The change in free energy value allows us to determine whether a chemical reaction will occur
without any energy input, also known as its spontaneity, or how spontaneous a chemical
reaction is.
Remember that entropy is inevitable. The reason that spontaneous reactions occur is because
they result in greater entropy; by releasing energy through a chemical reaction, more stable and
simpler compounds are formed, similarly to how highly concentrated and pressurized molecules
disperse, creating stable formation. The more complex a structure is, the more energy it holds,
the more energy it is able to release, and the more it is able to contribute to entropy.
2.3.1 Equilibrium
Free energy can be thought of as a system’s instability. The more unstable a system is, the more
energy it can release as it heads towards a more stable arrangement. This brings us to the topic
of equilibrium.
Chemical reactions are constantly occurring. Reactants react together to form products in a
forward reaction, and those products then react and reform back into the original reactants in
a reverse reaction. A chemical reaction is at equilibrium when both the forward reactions and
reverse reactions occur at the same rate, or when there is no net change in the concentration of
both products and reactants.
7
USABO Guide Andrew Ung
When a chemical reaction is at equilibrium, it is most stable, and thus is at its lowest possible
free energy value. When is a chemical reaction is not at equilibrium, either the forward reaction rate
or the reverse reaction rate increases in order to reach that equilibrium. This causes a meaningful
shift in the concentrations towards equilibrium, performing work and affecting free energy in the
process.
This is the reason that thermal energy formed from entropy cannot do meaningful work- it is
already at its most stable arrangement. It would be different if that thermal energy were concen-
trated, in which that thermal energy could be harnessed for power, forming the basis of geothermal
reactors. Thermal and chemical energy are analogous, a chemical reaction at equilibrium cannot
do much work, while a more unbalanced chemical reaction can be used for work.
Figure 2.6 Two systems, one that is able to perform work (left), and one that is at equilibrium
(right). This system is analogous to those of chemical reactions. (Source: Andrew Ung)
• The collections of water in the system on the left are not in equilibrium, causing a flow of
water from the left to the right, which is able to generate power, and thus has a negative
∆G value.
• The collections of water in the system on the right are in equilibrium, and there is no net
change occurring, and thus has a ∆G value of zero. Eventually, this is what the system on
the left will tend to become, due to the second law of thermodynamics.
3 Proteins
Proteins are important to life as they are what carry out the various functions of the cell. There
can be thousands of different proteins in a cell, each complete with their own unique function. The
structure and composition of these proteins are what allows them to be so variable, and is what
makes them unique to other molecules. Proteins have multiple levels of structure, each focusing on
different biochemical properties which all combine and interact to form a finalized, folded protein.
8
USABO Guide Andrew Ung
Figure 3.1 The basic structure of an amino acid (left). (Source: Andrew Ung)
The R groups, also known as side chains, are what determine an amino acid’s proper-
ties: buffering capabilities, hydrophobic properties, polarity, compressibility, and more. There are
twenty unique R groups, and thus twenty unique amino acids, each with their own properties.
It is recommended to study and memorize the properties of these amino acids before moving
on.
9
USABO Guide Andrew Ung
Figure 3.2 The twenty amino acids, divided by some of their properties. (Source: Campbell
Biology)
Some other special properties of amino acids relevant to primary structure are as follows:
• Beta-branched amino acids include: leucine, isoleucine, and valine. These beta-branched
amino acids have unique properties in regards to folding freedom, and can destabilize or
stabilize a protein’s structure depending on placement.
• Aromatic amino acids include: phenylalanine, tyrosine, and tryptophan. These aromatic
amino acids act as protein structure stabilizers, and molecular identifier tags.
Polypeptide chains are made of amino acids covalently bounded to each other by their carboxyl
groups and amine groups. This means that one end of the polypeptide chain has an unbounded
amine group, and the other end has an unbounded carboxyl group. These ends are known as the
N-terminus and C-terminus respectively.
Each amino acid has a three-letter code and a one-letter code. Both are commonly used, and
it is recommended to learn these thoroughly using the provided table above. This is useful when
listing the amino acids of a polypeptide chain. One example would be the amino acid structure of
leucine enkephalin, which can be typed out as Tyr-Gly-Gly-Phe-Leu.
10
USABO Guide Andrew Ung
In these groups, the hydrogen atoms have a partial positive charge (δ+), and the oxygen atoms
(from the carboxyl group) have a partial negative charge (δ-). This leads to hydrogen bonding
between different amino acids, which can stabilize the overall protein structure, and even cause new
structures to form. Note that the R group does not play a part in secondary structure formation,
it is simply the backbone constituents, or all the chemical groups excluding the R groups, that
are involved in hydrogen bonding.
3.2.1 α Helices
α helices are such a structure which can arise from this hydrogen bonding. These helices appear
as coils in a polypeptide’s conformation. The helix is held in place by bonding between every
fourth amino acid’s backbone constituents in a chain.
Figure 3.3 Hydrogen bonding between backbone constituents of a polypeptide, causing the
formation of an α helix. (Source: Campbell Biology)
Although the R groups of the amino acids in the polypeptide do not play a role in the formation
of α helices, there are some exceptions which prevent their formation. The following amino acids
are known as helix breakers, or helix disruptors.
• Proline has a unique structure where its R group is incorporated into its backbone chemical
groups (technically making it an imine), causing it to be unable to bend into the correct
conformations to form hydrogen bonds with other amino acids to form an α helix.
• Glycine has less of a pronounced effect on α helices when compared to proline, however
it can still disrupt its overall structure. Glycine, with a very small R group, has a large
amount of conformational freedom and is very flexible. This flexibility causes instability
when incorporated in a polypeptide with an α helix.
11
USABO Guide Andrew Ung
Example 3.1 (Adapted from USABO Open Exam 2018) You discover a protein that exhibits
parts composed of α helices. What is NOT a likely amino acid sequence for some portion
of the α helix containing part of the protein? Select all answers that apply.
(A) ALQQMMDSILDY
(B) ALILALMWWLLF
(C) FPPALGGLPFAMG
(D) LMQKPMDSLPDY
(E) LKMELAKMILLA
Solution: The question is asking which amino acid sequences would most likely not make
up an α helix containing part of a protein. Recall that glycine (G) and proline (P) are
helix breakers. Both sequences FPPALGGLPFAMG and LMQKPMDSLPDY contain either
glycine or proline, thus the answers are C and D.
These sheets, when in large concentrations, can be very strong—the fibers found in silk clothing
are made of proteins primarily composed of β pleated sheets.
Figure 3.4 Hydrogen bonding between parallel β strands, forming a β pleated sheet. (Source:
Campbell Biology)
12
USABO Guide Andrew Ung
Hydrophobic interactions are the interactions caused by nonpolar and polar R groups in-
teracting with the environment around them.
• Polar amino acids have R groups with partial charges which can interact and bond with
the water molecules in the aqueous solution surrounding and inside a cell, making them
hydrophilic.
• Nonpolar amino acids have R groups without partial charges and cannot interact or bond
with the water water molecules in the aqueous solution surrounding and inside a cell, making
them hydrophobic.
Nonpolar amino acids in a polypeptide chain thus tend to aggregate in groups, beneath polar
amino acids, similarly to how phospholipids in a cell membrane cluster their hydrophobic tails
inwards and their hydrophilic heads outwards.
Intermolecular forces between R groups of amino acids cause attraction and shape changes
in the resulting protein. Hydrogen bonding and other van der Waals forces keep R groups of
different polarities together.
Ionic bonding and covalent bonding can also occur, leading to shape changes.
• Covalent bonding is present in disulfide bridges, made from two cysteine amino acids
interacting with each other. Cysteine is unique in which it has a sulfhydryl chemical group
making up part of its R group, which can form covalent bonds with other sulfhydryl groups
to create disulfide bridges.
Figure 3.5 The various interactions which make up the tertiary structure of a protein. (Source:
Campbell Biology)
13
USABO Guide Andrew Ung
Amino acids can also be chemically modified. The three amino acids serine, threonine,
and tyrosine can all be modified through the process of phosphorylation, or the adding of the
chemical group phosphate. Phosphate groups are hydrophilic, and thus phosphorylations of these
amino acids can play a role in the tertiary structure of a protein.
One example of a protein with a quaternary structure is collagen. Collagen is made of three
polypeptide chains coiled around each other in a triple helix, creating one strong protein fiber, and
used extensively in animal organisms for its strength in tissue.
Figure 3.6 Collagen, a protein which exhibits a quaternary structure. Each differently-colored
strand is a separate polypeptide chain.
Polypeptides involved in quaternary structure do not necessarily have to be the same polypep-
tide. The protein hemoglobin, which transports oxygen and is found in red blood cells, exhibits
quaternary structure and has two different polypeptide subunits. Hemoglobin is made of two α
and two β subunits, which interact to form a singular hemoglobin protein.
4 Enzymes
Enzymes are a type of protein which allow many chemical reactions to occur in the cell to be
possible. Enzymes catalyze reactions, speeding them up and sometimes being the sole reason they
occur. Without enzymes, many cellular processes simply could not happen.
Each enzyme has a specific region, known as the active site, which chemically binds to and
interacts with the substrate. This active site needs to be shaped specifically for the substrate it’s
binding to in order for it to properly function. The resulting bound enzyme and substrate are
known as an enzyme-substrate complex.
14
USABO Guide Andrew Ung
Figure 4.1 An image depicting the binding of a substrate to an enzyme’s active site, forming an
enzyme-substrate complex. (Source: Campbell Biology)
There are currently two popular models which depict enzyme-substrate binding.
• In the Lock-and-Key hypothesis, an enzyme is structurally rigid and contains the correct
groove that only its substrate can fit in. The enzyme thus acts as a lock, and the substrate
as a key.
• In the Induced Fit hypothesis, an enzyme is not structurally rigid. When a substrate
is bound to the enzyme, interactions between the chemical groups of the enzyme and the
substrate cause the enzyme to undergo changes in its conformation, binding and shifting to
hold on tighter to its bound substrate.
Both models are generally accepted as correct and are complementary to each other.
The polypeptides which make up an enzyme are very complex and specific enough that when
placed in a suitable environment, they would fold into an enzyme protein. This protein has an
almost perfectly-shaped active site which promotes the bonding of a specific substrate. After the
initial binding of a substrate, further interactions result in even stronger binding.
The complexity of protein folding is so vast that, if a given polypeptide were to shift con-
formations every picosecond, it would take longer than the age of the universe to fold in every
possible way. This is the thought experiment known as Levinthal’s paradox. The reason that
polypeptides are able to fold into their biologically correct conformations so quickly, rather than
in the timeline of a universe, is due to various local interactions between the amino acid monomers
of a polypeptide and also help from other proteins. Chaperonins are a group of proteins which
aid in the polypeptide folding process.
4.2 Catalysis
Catalysis is defined as the acceleration of a chemical reaction by a catalyst. Enzymes are the
primary biological catalyst found in cells, and they perform their roles fantastically well.
15
USABO Guide Andrew Ung
In isolation, both of these principles make sense. However, when combining the two, some
inaccuracies appear to arise. For example, why does paper, made of cellulose built from glucose
molecules, not instantly break down into smaller molecules? The breakdown of glucose results
in energy being released, and ∆G is negative, thus the reaction is spontaneous. Why doesn’t
everything simply break down?
This is due to something called activation energy, Ea . Although the breakdown of paper is
energetically favored and spontaneous, it doesn’t simply occur without some form of external input.
Spontaneity doesn’t signify the speed of a reaction, simply whether it is energetically favorable
or not. There first needs to be an input of energy, the Ea , in order for this reaction to proceed.
When put to a flame, the paper burns, releasing energy in the form of heat and light, and breaks
down into its basic constituents. The flame would act as the energy input to fulfill the necessary
Ea , allowing the rest of the paper to burn and release energy. Similarly, a house of cards will not
fall until it is pushed, and the same goes for chemical reactions at the molecular scale.
The reason for the existence of Ea is that there are pre-existing bonds between the atoms of a
molecule. Recall that, in order to break a bond, energy must be inputted to separate individual
atoms until their bonds break. This required input of energy is what forms the basis of Ea . Before
a glucose molecule can break down, energy must be added to separate the bonds. Note that,
although an input of energy is required, the released energy is always greater than the Ea , so the
∆G value is always negative for a spontaneous reaction.
Although theoretically, spontaneous chemical reactions will always proceed, sometimes they
will not occur on its own, or rather will occur too slowly for practical purposes, unless in the
correct environment. Heat is a source of energy to fulfill Ea , however is impractical in many
cases, especially in the cell. Constantly using energy as heat for chemical reactions would be too
inefficient and would lead to many other problems. Heat, in addition to breaking the bonds of
the desired molecule for reacting, will also interact with the bonds of other cellular molecules, like
proteins, leading to denaturing and dysfunction.
Example 4.1 (USABO Open Exam 2010) When the temperature increases, which of the
following statements is not true?
16
USABO Guide Andrew Ung
Solution: Immediately, by looking at the provided answers, the correct choice should be
obvious, even if one doesn’t know whether the other answers are true or not. As temperature
increases, there is more thermal energy free to break and weaken bonds, thus allowing
metabolic reactions to be more likely to reach their activation energy. The answer is thus
C.
Figure 4.2 A graph showing an enzyme’s effect on the Ea of a chemical reaction. Note that ∆G
remains unchanged with or without an enzyme. (Source: Campbell Biology)
Enzymes lower Ea through placing mechanical stress on the its bound substrates. Energy is
only required to excite atoms and separate them enough that bonds break. Enzymes interact with
and bond with the substrate in such a way that it is mechanically pulled, putting strain on the
bonds and pushing atoms away from each other, decreasing the amount of energy needed to fully
separate them.
17
USABO Guide Andrew Ung
Figure 4.3 Models depicting normal enzymatic activity (top), noncompetitive inhibition
(bottom left), and competitive inhibition (bottom right). (Source: Campbell Biology)
18
USABO Guide Andrew Ung
Allosteric regulation is typically always reversible due to the constantly fluctuating needs of
the cell. Although there are some forms of irreversible allosteric regulation, they are few and far
between due to the lack of necessity and danger of a permanent change in enzymatic activity.
Enzymes are not entirely static; they constantly oscillate and dance between multiple possible
configurations. This phenomenon is referred to as conformational dynamics. These configura-
tions have different levels of enzymatic activity.
• The configuration where a substrate is most easily able to bind to an enzyme and catalysis
is most prominent is known as the active form.
• The configuration where a substrate is least easily able to bind to an enzyme and catalysis
is least prominent is known as the inactive form.
The sites at which allosteric regulator molecules bind to an enzyme are known as allosteric sites.
• Allosteric activators bind to the allosteric site on an enzyme and increase the likelihood
that an enzyme is in its active form. This leads to an increase in overall enzymatic activity.
• Allosteric inhibitors bind to the allosteric site on an enzyme and increase the likelihood
that an enzyme is in its inactive form. this leads to a decrease in overall enzymatic activity.
Figure 4.4 A chart depicting the different types of allosteric regulation. Note the
conformational dynamics of the shown enzyme, and the oscillation between its active and
inactive forms.(Source: Campbell Biology)
19
USABO Guide Andrew Ung
4.3.3 Cooperativity
Many enzymes which undergo allosteric regulation have multiple subunits. Recall the quaternary
structure present in certain proteins and their multiple polypeptide chains. In such enzymes, each
subunit is made of a separate polypeptide chain, and has its own active site and allosteric site.
During allosteric regulation, the inhibition or activation at one allosteric site is reflected on the
rest of the enzyme.
There are two types of cooperativity which can increase or decrease activity throughout an
enzyme.
An example of positive cooperativity can be found in the enzyme ATP synthase. When a H+
ion binds to an active site on a subunit of the enzyme, a conformational change occurs, leading to
greater H+ ion affinity in the other subunits.
An example of negativity cooperativity can be found in the protein hemoglobin, found in our
red blood cells. Although not an enzyme, as it doesn’t catalyze any reactions, hemoglobin exhibits
negative cooperativity in substrate binding. When an O2 molecule binds to the active site of one
of the subunits in hemoglobin, the affinity for CO2 molecules in the rest of the protein is decreased.
Substrates and proteins can have different levels of cooperativity. These degrees of cooper-
ativity are quantified by a value known as the Hill coefficient, or nHill .
nHill can range from 0-∞. Molecules which do not exhibit any binding at all have a nHill value
of 0. In practice, molecules typically have a hill coefficient of 0.5 < nHill < 3. Molecules with nHill
values outside of this range are rarely observed.
20
USABO Guide Andrew Ung
One example of feedback inhibition can be found in the process of cellular respiration. Pyru-
vate dehydrogenase is an enzyme which catalyzes the formation of Acetyl-CoA in the cell.
The molecule Acetyl-CoA then inhibits pyruvate dehydrogenase. This may seem counter-
intuitive at first, however it leads to overall regulation of Acetyl-CoA levels; when too much of
the molecule is present, the enzyme involved in its creation is inhibited, and when too little of the
molecule is present, the enzyme is uninhibited and free to catalyze more reactions.
Feedback inhibition can also reach back many steps in a chemical pathway. ATP, the source of
energy for the cell and one of the final products of cellular respiration, is able to inhibit an enzyme
involved in the very beginning steps of the entire process.
All of these types of enzyme control play a part in maintaining the homeostasis of a cell.
Through allosteric regulation, competitive and noncompetitive inhibitors and activators, feedback
inhibition and activation, cooperativity, and more, the multitudes of processes in a cell are cat-
alyzed and regulated.
21
USABO Guide Andrew Ung
An enzyme catalyzes a reaction very quickly—substrates enter the active site, a reaction is
catalyzed, and a product leaves the enzyme—up to millions of times a second. Most of the time,
an enzyme is idly waiting for a substrate to bind to its active site. By increasing the concentration
of a substrate a in an environment, this wait time is decreased, and the enzyme can thus catalyze
reactions quicker. At some point, increasing a will have no effect on the rate of a reaction, as the
enzyme will have no time between individual catalyses. This is Vmax . KM is a constant which
measures the affinity between the enzyme and its substrate. It is the substrate concentration at
which the reaction rate is half of Vmax .
It is important to note that there are other ways of increasing reaction rate than increasing
a. Simply increasing the amount of enzymes will lead to a greater total reaction velocity. This is
reflected in the Michaelis-Menten equation, where since enzyme concentration is increased, Vmax
is increased. Enzymes also have optimal temperature and pH levels where they work best, also
reflected in the Michaelis-Menten equation by increasing Vmax as conditions become closer to ideal.
The Michaelis-Menten equation for a chemical reaction can be graphed, creating a Michaelis-
Menten curve, allowing biologists to visualize the change in v against a.
Example 4.2 (USABO Semifinal Exam 2015) Hypothetical data for an enzyme on its
reaction kinetics were collected in the presence and absence of an inhibitor. Assuming that
the enzyme is properly modeled using Michaelis-Menten kinetics, Use the two plots below
22
USABO Guide Andrew Ung
Note that the plots show the same data, however, the horizontal axes are scaled differently.
What is the maximum reaction rate possible for the uninhibited enzyme under these condi-
tions?
(A) 10 uM/min
(B) 40 uM/min
(C) 80 uM /min
(D) 100 uM /min
(E) 120 uM /min
Solution: In this question, we are being asked to find the Vmax of the uninhibited enzyme.
We can thus ignore the red line representing the kinetics of the inhibited enzyme, and focus
solely on the blue one. Recall that the Michaelis-Menten equation is:
Vmax a
v=
KM + a
In the case of this graph, only substrate concentration is being increased. As a approaches
infinity, v eventually approaches Vmax , forming a horizontal asymptote at v = Vmax . When
looking at the graph, there is a clear horizontal asymptote at v = 100 uM /min. The answer
is thus D.
23
USABO Guide Andrew Ung
Figure 4.7 A Lineweaver-Burk plot. Note that substrate concentration is represented by [S] in
this depiction. (Source: PhD Nest)
Example 4.3 (IMDO Exam 2022) In the Lineweaver Burk plot, what is the x-intercept?
(A) Vmax
(B) KM
(C) 1/Vmax
(D) -1/KM
24
USABO Guide Andrew Ung
x-intercept of a graph is the point at which y = 0, so setting y = 0 and solving for x will
lead to the x-intercept.
KM 1
0= x+
Vmax Vmax
KM 1
=−
Vmax Vmax
1
x=−
KM
The x-intercept of a Lineweaver-Burk plot is − K1M , and thus the answer is D.
• In competitive inhibition, an inhibitor molecule binds to the active site of an enzyme, effec-
tively reducing the total amount of active enzymes available for binding by the substrate. A
higher substrate concentration is required to achieve half of the maximum reaction rate, so
KM is increased.
• Since the competitive inhibitor does not affect the catalytic properties of an enzyme, Vmax
is unchanged.
• Since the noncompetitive inhibitor affects the catalytic properties of an enzyme, Vmax is
decreased.
It is also important to note the existence of uncompetitive inhibition, which leads to both
decreased KM and decreased Vmax .
• When inhibiting an enzyme-substrate complex, the substrate is trapped and cannot release
as easily. This decreases the catalytic efficiency of the enzyme, and thus Vmax is decreased.
25
USABO Guide Andrew Ung
Figure 4.9 Multiple Lineweaver-Burk plots depicting the effects of competitive, uncompetitive,
and noncompetitive inhibition on an enzyme-catalyzed reaction. (Source: Jack Westin)
Example 4.4 (Adapted from IMDO Exam 2022) Which of the following is found in non-
competitive inhibition of an enzyme?
26
USABO Guide Andrew Ung
Solution: It may be helpful to think about how each type of inhibition affects an enzyme and
deduce its effects on KM and Vmax from there. For example, remember that noncompetitive
inhibition is a form of allosteric regulation where the inhibitor molecule does not bind to
the active site of an enzyme. Because of this, noncompetitive inhibition has no effects on
KM , as it doesn’t affect the affinity of the enzyme to its substrate. The answer is thus D.
5 Conclusion
5.1 Closing Statement
The concepts which have been outlined in this paper have hopefully been enlightening to a new
reader who has not yet learned the laws of thermodynamics and the universe, and opened their
eyes to the depth of biology and how far life has come. As we study physics and make observations
of the world around us, we realize the pure, unbreakable laws which govern our lives, of entropy
and energy. As we study biology, we see the miraculous ways our cells have evolved in order to
overcome and bypass such laws. We exist in a world constantly breaking apart, and eventually the
entropic heat death of the universe will occur, but our cells and life as we know it have seemed to
temporarily overcome entropy and have built our house of cards.
27