D. Van Der Marel, H. J. A. Molegraaf, C. Presura, Alex C. Hewson, Veljko Zlatić - Concepts in Electron Correlation PDF
D. Van Der Marel, H. J. A. Molegraaf, C. Presura, Alex C. Hewson, Veljko Zlatić - Concepts in Electron Correlation PDF
The Series is published by IOS Press, Amsterdam, and Kluwer Academic Publishers in conjunction
with the NATO Scientific Affairs Division
Sub-Series
The NATO Science Series continues the series of books published formerly a s the NATO ASI Series.
The NATO Science Programme offers support for collaboration in civil science between scientists of
countries of the Euro-Atlantic Partnership Council. The types of scientific meeting generally supported
are "Advanced Study Institutes" and "Advanced Research Workshops", although other types of
meeting are supported from time to time. The NATO Science Series collects together the results of
these meetings. The meetings are co-organized bij scientists from NATO countries and scientists from
NATO's Partner countries - countries of the CIS and Central and Eastern Europe.
Advanced Study Institutes are high-level tutorial courses offering in-depth study of latest advances
in a field.
Advanced Research Workshops are expert meetings aimed at critical assessment of a field, and
identification of directions for future action.
A s a consequence of the restructuring of the NATO Science Programme in 1999, the NATO Science
Series has been re-organised and there are currently Five Sub-series as noted above. Please consult
the following web sites for information on previous volumes published in the Series, as well as details of
earlier Sub-series.
http://www.nato.int/science
http://www.wkap.nl
http://www.iospress.nl
http://www.wtv-books.de/nato-pco.htm
edited by
Alex C. Hewson
Department of Mathematics,
Imperial College, London, United Kingdom
and
Veljko Zlatic
Institute of Physics,
Zagreb, Croatia
A C.I.P. Catalogue record for this book is available from the Library of C o n g r e s s .
Preface 5
D. van del' Marel, H. J. A. Molegraaf, C. Presura, 1. Santoso/
Superconductivity by Kinetic Energy Saving? 7
R. Hlubina/ Josephson effect in the cuprates: microscopic
implications 17
M. Sigrist/ Ruthenates: Unconventional superconductivity and
magnetic properties 27
K-D. Schotte, C.T. Liang/ Real structure of perovskites looked at
from the band structure point of view 35
H. Eschrig, K Koepernik, 1. Chaplygin/ Towards models of
magnetic interactions in the cuprates 45
J. Rohler/ c-axis intra-layer couplings in the CU02 planes of
high-T c cuprates 55
P. Prelovsek, A. Ramsak / Spectral functions and pseudogap in a
model of strongly correlated electrons 65
G. Baskaran/ Mott insulator to superconductor via pressure RVB
theory & Prediction of New Systems 75
M. Lang, J. Miiller, F. Steglich, J. Schlueter, T. Sasaki/ Exploring
the phase diagram of the quasi-2D organic superconductors K,
-(BEDT-TTF)2X 85
M. Fabrizio, E. Tosatti, M. Capone, C. Castellani/ Enhancement
of superconductivity by strong correlations: a model study 95
J.K Freericks, T.P. Devereaux, R. Bulla/ Inelastic light scattering
and the correlated metal-insulator transition 115
L.F. Feiner, A.M. Oles/ Orbital physics versus spin physics: the
orbital-Hubbard model 123
J. Bah, A. M. Oles/ Quasiparticles in photoemission spectra of
manganites 133
2
5
SUPERCONDUCTIVITY BY KINETIC ENERGY SAVING?
Key words: Optical Conductivity, Spectral Weight, Specific Heat, Pair Correlation, Kin-
etic Energy, Correlation Energy, Internal Energy
7
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 7–16.
© 2003 Kluwer Academic Publishers
8
200
sz
ro
-2' 100
~
U
~ 20 1.5
OJ
'2 E
,., 15 o
OJ
Q; 1.0 ro::J
C
U
:>
W 10
roc
Q; 5
0.5 E
C
0+-_~~~~~~~~~---10.0
o 50 100 150 200
Temperature (K)
Figure 1. Experimental internal energy and the specific heat obtained by Loram et
al (1). The extrapolations are obtained following the procedure of Ref. (3)
1.2. PAIR-CORRELATIONS
to the finite mean free path of the electrons. As a result the integral over
the center of mass coordinates of the correlation function
(2)
In the weak coupling scenario of BeS theory the electrons have an effective
attractive interaction, as a result of which they tend to form pairs. For the
purpose of the present discussion we will assume that the interaction is of
the form
(3)
where n(r) is the electron density operator. The interaction energy in the
superconducting state becomes lower than in the normal state, due to the
fact that the effective attractive interaction favors a state with enhanced
pair-correlations. The value of the interaction energy of the superconducting
state, relative to the normal state is
where gk and Vk are the Fourier transforms of g(r) and V(r) respectively.
Using the Bogoliubov transformation the correlation function can be ex-
pressed (4) in terms of the gap-function D..k and the single particle energies
Ek = {(Ek - p,)2 + D..k}1/2.
(5)
10
0.2
--...
en
Figure 2. The k-space (top panel) and coordinate space (bottom panel) representation
of the superconductivity induced change of pair-correlation function for d-wave symmetry
(bottom panel). Parameters: 6/W = 0.2, WLJ/W = 0.2. Doping level: x = 0.25
This corresponds to the conversion of a pair (q, -q) to a pair with quantum
numbers (q + k, -q - k). In the expression for the correlation energy, Eq. 4,
the transferred momentum k is carried by interaction kernel Vk .
Starting from a model expression for the single electron energy-momentum
dispersion Ek, and the gap-function D..k, it is a straightforward numer-
ical exercise to calculate the summations in Eq. 5. Adopting the nearest
neighbor tight-binding model with a d-wave gap, and adopting the ratio
D..Clf,O)/W = 0.2, where W is the bandwidth, the correlation function gk
can be easily calculated, and the result is shown in Fig. 2. We see from this
graph that a negative value of (Hi) s - (Hi)n requires either (i) Vk < 0 for
k in the neighborhood of the origin, or (ii) Vk > 0 for k in the vicinity of
Clf, Jr). The corresponding representation in real space, g(r), shown in Fig. 2,
illustrates that the dominant correlation of the d-wave superconducting
state is of pairs where the two electrons occupy a nearest neighboring site,
while the on-site amplitude is zero.
Combining the information of Fig. 2 with Eq. 2, it is clear that the strongest
saving of correlation energy is expected if the electrons interact with an
interaction of the form V(rl' r2) = Va La b(rl - r2 + a) where the vector a
11
15
'>
Q)
.s 10
E
~
<1
.s t = -297.6 (meV)
t' = 81.8 (meV)
i=' -140
'::;:
-141
15
Q'
;>Q) 10
3
i='
U 5
ol:--"o"~--+-+-+--+-+-+-+-+-+--+->--t--..---.--+-+-j
Temperature (K)
Figure 4. Bes prediction of the internal energy and the specific heat.
The BCS prediction for the temperature dependence of the average in-
teraction energy follows from Eq. 4. We then use the BCS variational
wave-function for the statistical average of Eq. 5, resulting in
(7)
(8)
-227.0
:;- Weak coupling BCS-theory
OJ
t' / t = -0.27
-S d-wave pairing, T c = 80 K
~ -227.5
OJ
c
W
.S2 -228.0
Ql
c
S2
-228.5
Correlation Energy
Figure 5. BeS prediction of the kinetic energy and the correlation energy.
-0 dwReo-(w)dw = (9)
/
where the high frequency limit indicates that the integral should include
only the intra-valence band transitions, and the condensate peak at w = 0
if the material is a superconductor. The integral over negative and positive
frequencies (note that cr( w) = cr* ( -w)) avoids ambiguity about the way
the spectral weight in the condensate peak should be counted. If the band
structure is described by a nearest neighbor tight-binding model, Eq. 9
leads to the simple relation
(10)
Hence in the nearest neighbor tight-binding limit the partial f-sum provides
the kinetic energy contribution, which depends both on the number of
particles and the hopping parameter t (11, 12, 13, 14).
However, if the band-structure has both nearest neighbor hopping and next
nearest neighbor hopping, Eq. 10 is not an exact relation, and instead Eq. 9
should be compared directly to the experiments. In Fig. 6 we compare the
spectral weight, calculated directly using Eq. 9 to the result of Eq. 10, using
the same parameters as for Fig. 3. Note that the kinetic energy has to be
14
112.8
112.6
>
OJ 112.4
-S
a: 112.2
112.0
>
OJ
-S
N
:;-
(l)
E
~
--'
Q.
142
:;-
(l)
E
--'
Q.
138
Figure 7. Experimental values of the ab-plane spectral weight function, taken from
Ref. (15)
The trend seen in the experimental data has been predicted by Hirsch in
1992 (17, 18, 19). The model assumption made by Hirsch was, that the
hopping probability of a single hole between two sites becomes larger if one
of the two sites is already occupied by a hole. Although this model provides
good qualitative agreement with the optical experiments, it has one serious
deficiency: It also predicts s-wave symmetry for the order parameter, in
sharp contrast to a large body of experimental data which show that the
superconducting gap in the cuprates has d-wave symmetry.
In a recent set of calculations based on the Hubbard model, Jarrell et al (20)
obtained a similar effect as seen in our experiments, both for underdoped
and optimally doped samples. Crudely speaking the mechanism is believed
due to the frustrated motion of single carriers in a background with short-
range (RVB-type) spin-correlations, which is released once pairs are formed.
6. Conclusions
Acknowledgements
References
R. HLUBINA
Department of Solid State Physics, Comenius University,
Mlynska dolina F2, SK-842 48 Bratislava, Slovakia
Abstract. In the tunnel limit, the current-phase relation of Josephson junctions can be
expanded as I( ¢) = h sin ¢ + h sin 2¢. Standard BCS theory predicts that hRN ~ 61 e
and hih ~ D, where R N is the resistance of the junction in the normal state, 6 is
the superconducting gap, and D « 1 is the junction transparency. In the cuprates, the
experimental value of hR N (hih) is much smaller (larger) than the BCS prediction. We
argue that both peculiarities of the cuprates can be explained by postulating quantum
fluctuations of the pairing symmetry.
1. Introduction
17
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 17–25.
© 2003 Kluwer Academic Publishers
18
detailed analysis of the Josephson effect, new insights into the nature of
quantum fluctuations in the cuprates can be obtained.
Josephson junctions involving the cuprate superconductors have been
studied mainly because they enable phase-sensitive tests of pairing sym-
metry (3). Attention has been paid especially to two particular types of
Josephson junctions: grain boundary (4) and intrinsic Josephson junctions
(5). The best studied type of grain boundary Josephson junctions involves
junctions built on c-axis oriented films, where the weak link forms at the
boundary of grains which are rotated around the [001] axis with respect to
each other. Idealized junctions of this type are characterized by a planar
interface and two angles e1 and e2 between the interface normal and the
crystallographic directions in the grains forming the junction. The proper-
ties of the junctions depend dominantly on the misorientation angle e =
e 2 -e 1 . It is well known (4) that the transparency D of grain bound-
ary junctions decays exponentially with increasing misorientation angle e,
D(e) ex: exp(-e/eo). Thus, for e > eo;::::::; 5°, grain boundary junctions are in
the tunnel limit and (for e not too close to 45°, see Section 3) their current-
phase relation I (¢) = h sin ¢ + h sin 2¢ + ... can be well approximated by
I( ¢) = h sin ¢, neglecting the higher-order harmonics.
(1)
momentum of the Cooper pairs (for a similar calculation, see e.g. (13)).
The conductance per square in the normal state is given by G N
(2e 2 /n)N(0)(Zktkrk1)N, where rk is the inverse lifetime ofthe quasiparticles
and the index N in the Fermi line average means that the quantities are
to be evaluated in the normal state. Therefore standard theory predicts for
the Josephson product of intrinsic Josephson junctions hRN = j1Gil ;:::::
e-l(zktk)/(Zktkrkl)N. In conventional superconductors RN can be meas-
ured at low temperatures in a sufficiently large magnetic field. This is
impossible for the cuprates and thus RN is usually defined as the c-axis
resistivity at T e .
Unfortunately, due to the unknown temperature dependence of Zk, the
theoretical hRN can not be directly tested by experiment. In order to
overcome this problem, in (14) instead of the usual Josephson product
a related characteristic of intrinsic Josephson junctions has been stud-
ied, namely the product of the critical current h and of the resistivity
Rs in the resistive mode of the junction (at low temperatures). Since
the conductance per square in the resistive mode of the superconductor
is (14) Gs ;::::: (8e 2 /h)N(O)(Zktk)node/~, standard BCS-like theory pre-
1
dicts hRs = j1G S ;::::: (7r/2)(~/e)(zktk)/(Zktk)node. In (14), hRs '" ~/e
has been found experimentally and good agreement with theory has been
claimed, since momentum-independent tk and Zk were assumed. However,
according to band structure calculations (15), tk is strongly suppressed in
the nodal directions. If this modulation of the tunnel matrix element tk is
taken into account and the presumably only moderate k-space dependence
of Zk is neglected, the experimental hRs is seen to be drastically reduced
with respect to the theoretical predictions.
Thus we have shown that although the barriers in grain boundaries and
in intrinsic Josephson junctions are of very different nature, both types of
junctions exhibit a suppressed Josephson product. Therefore we believe that
this suppression is not due to specific barrier properties as suggested in (9),
but rather due to some intrinsic property of the high-Te superconductors.
However, the results of (16, 17) are quite mysterious, if we take into
account the actual experimental setup. In fact, standard BCS theory with
ideal featureless barriers implies that in order that hi rv hi, the average
1 1
4. Microscopic implications
where H fluct = -m- 1 Li 8 2 /8<Pr. The second term in Eq. (2) (with (i,j)
denoting a pair of nearest neighbor sites) describes for 0 < W < V a
superconductor with a dominant d-wave pairing V + Wand subdominant
s-wave pairing V - W. Since we concentrate on the q ;:::j (1T,1T) fluctuations
23
5
d+is
4
> 3 disordered
--
"""')
2 d
OIL-----'-----------'---------'----'---------'----------'
o 2 3 4 5 6
WN
Figure 2. Mean field phase diagram of the toy model.
5. Conclusions
Acknowledgements
References
MANFRED SIGRIST
Theoretische Physik, ETH-Honggerberg, 8093 Zurich, Switzer-
land
1. Introduction
For the last two decades transition metal oxides have been among the
leading topics in condensed matter physics. Best known are the cuprates
which are high-temperature superconductors and the manganese oxides dis-
playing the colossal magnetoresistence phenomena. Among these classes of
materials also ruthenates play an important role as systems with intriguing
properties. The interest in this material mainly arose from the discovery of
unconventional superconductivity in Sr2Ru04 with p-wave pairing (1, 2).
Unlike many other stoichiometric compounds Sr2Ru04 is not a Mott insu-
lator, but is a very good metal with clear Fermi liquid properties. This Fermi
liquid shows features of strong correlation and is quasi-two-dimensional due
to the layered crystal structure. It was suggested that this system could be
considered as the electronic analog of 3He with the natural implication that
the superconducting state originates from spin triplet p-wave pairing (3, 4).
The electronic properties are dominated by the nearly degenerate 4d-t2g-
orbitals of the Ru4+-ion, which introduces four electrons per Ru. These or-
27
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 27–34.
© 2003 Kluwer Academic Publishers
28
There are two key experiments determining the symmetry of the pairing
state. The measurement of the spin susceptibility via NMR-Knight shift
addresses the spin part of pair wave function, and indicates that there is
equal-spin-pairing with parallel spins in the Ru02-planes (5). Furthermore,
the zero-field muon spin relaxation shows the appearance of spontaneous
intrinsic magnetism below the superconducting phase transition (6). This
magnetism originates most likely from the net angular momentum of the
Cooper pairs, determining the orbital part of the pair wave function. For a
quasi-two-dimensional system the orbital angular momentum is naturally
oriented perpendicular to the basal plane. The resulting pairing state is
entirely consistent with the symmetry classification including spin-orbit
coupling, and can be represented in the vector notation as
with cLs creating an electron on site i in the orbital f-l with spin s (Ep,vp is
the completely antisymmetric tensor). It is possible to fit the band structure
with a simple tight-binding model. Using a simple phenomenological model
for the pairing interaction Ng and Sigrist concluded that the i-band origin-
ating from 4dxy would have to be dominant for superconductivity in order to
stabilize the chiral p-wave phase (9, 10). This result was recently confirmed
by Yanase and Ogata (17) using a microscopic pairing mechanism based
on a perturbative approach of the extended three-band Hubbard model as
introduced by Nomura and Yamada (19, 18). The latter group had found
that within their theory indeed the i-band would play the leading role for
spin-triplet superconductivity.
The assumption that the pairing symmetry is related with the so-called
orbital-dependent superconductivity (ODS) and pairing symmetry via spin-
orbit coupling is very consistent with several experimental and theoretical
facts. Spin-triplet superconductivity is believed to be supported by fer-
romagnetic spin fluctuations. The analysis of NMR-experiments suggests
that the i-band is closer to ferromagnetism than the a-p-bands which
show strong incommensurate spin correlations due to Fermi surface nesting
features (12, 13).
ODS plays also an important role for the thermodynamics, since mul-
tiple gaps introduce significant modifications in the temperature depend-
ence (8). Power-law behavior of various quantities over a wide temperat-
ure range, which have been attributed to line nodes, can be well under-
stood also within a multi-band picture as shown by Zhitomirsky and Rice
(14), Nomura and Yamada (19) for the specific heat. Kusunose and Sigrist
30
showed that the analysis of the behavior of the London penetration depth
requires one furthermore to take non-local effects into account (15, 16).
Nevertheless, the fitting of the thermodynamic quantities does not lead to
the unique determination of gap shapes and it may remain even difficult to
settle the question whether the gaps in some bands have nodes.
3-Kelvin phase: A surprising feature has been the onset of inhomogen-
eous superconductivity in some samples of Sr2Ru04 at higher temperatures
than the bulk-Teo In accordance with the approximate onset temperature
this phase is called the "3 K-phase" (20). The analysis of the material
revealed the presence of excess-Ru in form of micrometer-sized Ru-metal
inclusion embedded into the pure Sr2Ru04-metal. Since Ru on its own
becomes superconducting below 0.5K only, it has been suggested that the
actual nucleation region of the 3 K phase is the interface between the Ru-
inclusions and Sr2Ru04. It is possible that the transition temperature of
the p-wave superconductivity is slightly enhanced at the interface due to
crystal lattice deformations. Based on this assumption a Ginzburg-Landau
theory of the nucleated phase has been developed (21). Without going into
the calculational details we would like to review here some key results of
this analysis and compare with experimental results as far as available.
1. The interface corresponds to a region of reduced symmetry, so that
the two p-wave components with momentum direction parallel and perpen-
dicular to the interface (kll and kj..) have different onset temperatures. It
is found that p-wave component parallel to the interface is more stable,
because its coherence length along the normal direction of the interface
is shorter, so that the order parameter is better confined to the beneficial
region. Thus the 3 K phase is time reversal symmetry conserving and before
the onset of the bulk phase a second phase transition has to occur breaking
time reversal symmetry.
2. The superconducting interface state corresponds to a gap function
with phase 0 and 7r depending on momentum direction. This is a condi-
tion to generate zero-energy Andreev bound states in the Ru-inclusions
analogous to the zero-energy surface states in d-wave superconductors.
These states are observable in quasiparticle tunneling experiments. Re-
cent data by Mao et al in break junctions, where tunneling is likely to
be dominated by inter-Ru-inclusion tunneling, show the presence of zero-
bias voltage anomalies in the 3 K phase, consistent with the nucleation of
the a single-component p-wave state at the interface (22).
3. The critical magnetic field for the 3 K phase is highest, if the magnetic
field lies parallel to the interface, since in this case orbital depairing is
reduced. Interestingly under this condition the temperature dependence of
H e2 is different from the standard linear bulk behavior. For an infinitely
planar interface there is a square-root behavior, H e2 ex: VT - T* with T*
31
electron in the i-band remains itinerant. This gives rise to a localized spin
S = 1/2 degree of freedom. This would explain the fact that for x close
to 0.5 the spin susceptibility indicates the presence of such a localized spin
per site which is difficult to explain within the picture of entirely itinerant
electrons (25, 26). On the other hand, it is not clear whether this orbital-
selective Mott insulator would give rise to the experimentally observed
degree of metallicity.
The scenario of the orbital-selective Mott insulator introduces with the
three electrons in the d yz / dzx-orbitals beside the spin also an orbital degree
of freedom (isospin 1/2). Thus the effective model of the localized degrees
of freedom corresponds to a Kugel-Khomskii type of model for coupled
spin 1/2 and isospin 1/2 (27). It has been shown that this effective model
in region III would give rise to staggered orbital order, i.e. having a single
hole in the d yz (dzx)-orbital on the sublattice A (B) of the Ru square lattice.
This in turn leads to a ferromagnetic spin interaction via the Goodenough-
Kanamori rule.
Further doping leads into region II where the crystal changes symmetry
from tetragonal to orthorhombic. This Jahn-Teller distortion introduces a
bias on the localized orbital degrees offreedom of the d yz / dzx-orbitals, sup-
pressing the staggered correlation in favor of an alignment of the orbitals.
Eventually the spin-spin interaction turns antiferromagnetic with growing
lattice deformation in the effective model. This behavior is in qualitative
agreement with experiments. The mean field phase diagram of the effective
spin-orbital model including the lattice deformation is shown in Fig.I. For
small deformation the staggered orbital order is prevailing leading to a
ferromagnetic phase at low enough temperature. There is, however, a first
order phase transition to the competing antiferromagnetic order, which
at sufficiently strong deformation truncates abruptly the staggered orbital
(antiferro-orbital) order.
The role of Ca-Sr alloying is to introduce lattice deformations through
octahedra rotation and tilt. In the alloy there is naturally disorder and the
properties show a certain degree of spatial variation. This means that in the
region 0.2 < x < 0.5 the two phases separated by the first order transition,
FM-AFO (ferromagnetic-antiferro-orbital) and AFM-FO (antiferromagnetic-
ferro-orbital), coexist in domains with rather sharp boundaries. This as-
pect allows us to connect the metamagnetic transition and the anomalous
behavior of the magnetoresistance, observed for Ca1.8SrO.2Ru04 (28). At
temperature of about 2 K a smooth metamagnetic transition has been
observed at about B ex = 2.7T and the longitudinal magnetoresistance
passes a pronounced maximum at approximately the same field value. In
a finite magnetic field the domain structure of the material is modified by
increasing the FM-AFO domains. The metamagnetic transition is associ-
33
antiferro-orbital "
I
I
I
I
I antiferromagnetic
I
ferromagnetic
orthorhombic strain
Figure 1. Schematic phase diagram for temperature versus orthorhombic strain for
the Kugel-Khomskii type model with spin-orbital degrees of freedom: The solid (dashed)
lines represent second (first) order phase transitions (29).
Acknowledgements
KK Ng, M. Troyer and, especially, T.M. Rice for the collaboration on the
topics summarized here. I am also very grateful to Y. Maeno, S. Nakatsuji,
H. Yaguchi and K Wada for many helpful discussions. Furthermore I thank
to M. Indergand for careful reading of the manuscript. These works had
financial support from the Swiss Nationalfonds and a Grant-in-Aid by the
Japanese Ministry of Education, Culture, Sports, Science and Technology.
References
Key words; Perovskite, Crystal Structure, Band Structure Calculation, Jahn- Teller Effect
The mineral CaTi0 3 found in the Ural mountains, gave its name perovskite
to a large group of compounds investigated intensively (1, 2). In solid state
physics their magnetic, ferro-electric and superconductivity properties were
the incentive. In mineralogy the high pressure conditions of the earth's
35
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 35–43.
© 2003 Kluwer Academic Publishers
36
, ~ .' -:- .. -. - ~ -.. ---.-
--
,~~-"-:""~-._-_. ~- --~ :,.. - ::;::~
, : ,
,
.. -
k
:0_---
~::::::~ -. - ..
.. - ..... "
:;., "r'
- .. -
x
Figure 1. Structure of Re03 as an ideal network of Re06 octahedra
(1)
is 1 geometrically. Using the ionic radii for the two distances d one obtains
usually smaller values, so that Mg2+ in MgSi0 3 is too small for the space
available. The tolerance ratio t < 0.9 is an indicator for quite a number of
perovskites according to the data collection of Wyckoff (1) deviating from
a simple cubic structure with one formula unit per cell.
The compound ABX 3 considered as a corner sharing network of oc-
tahedral complexes BX6 indicates that the octahedra taken as rigid units
have the freedom to rotate or tilt. Since there are many possibilities for such
displacements a large variety in the structure of perovskites is possible (4).
37
Figure 2. One unit cell of orthorhombic SrRuO;j. The rotations can be classified as
b+ a - a - after (4), that is rotations around the x, y, z directions of an octahedron having
an opposite sense a- or the same b+ in a adjecent octahedra along the rotation axes. Ru
(smallest) and 0 are connected by bonds, the largest spheres are Sr.
10
en
OJ
~
4-<
0
0
q
~
OJ
CI
-10
-20'------'-------'-------'------'------'-------'-------'-------'------'-----"'-----L--'-------'-------'------'------'-------'-------'-------'------'
-10 -8 -6 -4 -2 0 4 6 10
Enerh'Y leV]
structure. However, to fix the X-ion at the high symmetry position is not
possible if the core is too small indicated by the tolerance ratio, see eq(l),
0.75 < t < 0.9. Since further the BX 6 octahedron is a rigid unit only a
collective movement of all the X-ions is possible, as classified by Glazer (4).
A typical example can be seen in fig.2, where the structure of SrRu03
is displayed as determined by neutron scattering (5). The earth's mantle
material MgSi0 3 (3) has the same structure like many others found first in
the ferrite and "parent" substance GdFe03 (6). Four molecules ABX 3 are
contained in a orthorhombic unit cell with the space group Pnma (7).
Provided one has a so called "full-potential" program that avoids spher-
ical averaging of the ionic potentials, one can determine this structure with
the present band structure codes using the local density approximation.
However, to find a minimum of the total energy close to the tilted structure
determined experimentally may be time consuming (see (3) and fig.4).
From band structure point of view a more interesting object than the
insulator MgSi0 3 is a metal like SrRu03' especially since it is a ferromagnet.
Due to the lower symmetry and larger unit cell compared to the cubic one,
diffraction effects could lead to gaps in the density of states. If such changes
are close to the Fermi energy, they influence the electronic transport and
certainly the magneto-resistance.
We repeated the calculation of Singh (8) with the flp02-program (9).
As seen in fig.3 gaps appear, which for spin-up electrons are close to the
39
Fermi energy. The 4 d states of Ru are split by crystal field effect into t g
and ego The e g states are empty and the four d-electrons occupy only the t g
states. In an isolated ion three electrons would fill the up states completely
and one of the down states, which would give spin 1. The band structure
gives a lower value of 0.75. The small gap is due to a small hybridization
of t g and e g , which would be absent for cubic SrRu03 without tilting. We
suspect that correlation effects could pin this feature really to the Fermi
level.
Unlike the tilting of octahedra discussed, the band structure approach has
difficulties to reproduce the J ahn-Teller distortions of the transition metal
perowskites with J ahn-Teller active ions like Cu 2 +, Mn3+ and Cr 2 +. For
these ions the Jahn-Teller effect is equivalent to the orbital ordering (10)
of e g states. These states are doubly degenerate in a perfect octahedral co-
ordination of the transition metal ions. Lifting this degeneracy by lowering
the symmetry is the essence of the Jahn-Teller effect (11).
Spin ordering appears similar to orbital ordering in that the two-fold
degeneracy between up and down direction is lifted. This can be treated by
modifying the local density approximation (LDA) into a local spin density
approximation (LSDA). Otherwise ferromagnetism or anti-ferromagnetism
would not occur in a band structure calculation (12). Similar modifications
seem to be necessary to generate orbital ordering.
This has been done by adding to the standard LSDA an orbital depend-
ent potential (14) which is called shortly LDA + U. That indeed an orbital
ordering and a Jahn-Teller effect in fair agreement with the structural data
occurs had been demonstrated by Liechtenstein (15).
We repeated his calculation for KCuF 3 using the "Wien2k"-code where
a LDA + U version is included (19). In order to keep the computational
effort low we have chosen the ferro-distortive KCuF 3(16) and imposed fer-
romagnetic spin ordering so that only two formula units were in a tetragonal
cell. The type of magnetic ordering has little influence on the structure.
P4/mbm (7) the Wyckoff positions are 4h: (~ - 5, ~ + 5, ~) for F in ab-
plane around the Jahn-Teller active Cu at 2c: (0, ~, ~) as shown in fig.4.
K has the position 2a: (0, 0, 0) and the other F at 2d: (0, ~' 0). We have
taken for U = 0.59, see eq(2), and J = 0.07Ryd. 1 The lattice constants are
a = 5.8543 A and c = 3.9303 A. The ordering of the orbitals can be seen in
fig.5.
1 To run the Wien2k-program the muffin-tin radii were set to Rcu = 2.0, RK = 2.5,
-0.355
, I
" I
", II
, I
TI -0.360
'+, I
I
>. , I
~ \ I
o
t.ri
o
\, I
f
N
o
-{)365
+, I
\ I
I \ I
>.
Dl
, I
<Ii ~ I
c: \ I
w \ experimental /
-0.370 , I I
\
", :)-
I /
/
~ I /
'....... I ",,;'
-0.375 "*-----+'"
o 0.01 0.02 0.03 0.04
Distortion ,j 2S
Figure 4. Total energy as a function of the Jahn-Teller distortion (j as indicated in the
insert showing the tetragonal unit cell of KCuF a seen along the c-axis with (jcxp = 0.022.
Why does "Slater's dream machine" 2 fail so badly with the transition
metal oxides or halides? "Correlation" is the standard answer. However, in
Hartree-Fock without any additions and correlations, the orbital ordering
and generating of gaps seem to be no problem (13). Following ref. (14)
the LDA + U ansatz is constructed by adding a Hubbard-like term and
subtracting its mean value, that is the energy is
where N = 2:i ni is the total number of electrons in the d-shell and ni the
occupation of sublevel i. One assumes here that the standard LDA leads
to a reasonable total energy but the energy of the level should be changed.
By taking the derivative one obtains an orbital dependent potential
(3)
Acknowledgements
We thank Jurgen Kubler, Helmut Eschrig and Peter Blaha for their help.
References
in Paris 2002 referring the significant developments of solid state physics in the past.
42
Figure 5. Density plots: above the sum and below the difference of spin up and down.
The contour lines have values 0.2, 0.4 A-~ .. The horizontal axis is in the ab-plane along
the Jahn-Teller distorted Cu-F bonds, vertical is the c-axis. The e g hole orbitals alternate
as described in ref. (10). Above the Cu ions shrink where the hole orbitals are largest.
2. J.B. Goodenough & J.M. Longo, Crystallograhpic and Magnetic Properties of Per-
ovskites and Perovskite-Related Compounds, Landolt-Bornstein Neue Serie III/4a,
p.126, Springer Verlag (1970)
3. Lars Stixrude and R.E. Cohen, Nature 364, 613 (1993).
4. A.M. Glazer, Acta Cryst. B 28, 3384 (1972), see also Acta Cryst. A31,756 (1975).
5. C.W. Jones, P.T. Battle, P. Lightfoot and W.T.A. Harrison, Acta Cryst. C 45, 365
(1989).
6. S. Geller, J. Chern. Phys. 24, 1236 (1956).
7. International Tables for X-Ray Cr·ystallography, Kynoch Press, Birmingham (1969).
8. D.J. Singh, Electronic and IvIagnetic properties of the 4d itinerant ferromagnet
SrRuO:J, J. Appl. Phys. 79, 4818 (1996).
9. K Koepernik and H. Eschrig, Full-Potential Nonorthogonal Local-Orbital
minimum-basis band structure scheme, Phys. Rev. B 59, 1743 (1999).
10. KL Kugel and D.L Khomskii, The Jalm-Teller Effect and Magnetism: Transition
metal compounds, Sov. Phys. Usp. 25, 231 (1982).
11. M.D. Sturge, Jalm-Teller Effect in Solids, Solid State Physics 20, 91 (1967).
12. J. Kubler, Theory of Itinerant Electron Magnetism, Clarendon Press (2000).
13. M.D. Towler, R. Dovesi and V.R. Saunders, Phys. Rev. B 52, 10150 (1995).
14. V.L Anisimov, LV. Solovyev, M.A. Korotin, M.T.Czyzyk and Sawatzky, Density-
Functional Theory and NiO Photoemission Spectra, Phys. Rev. B48, 16929 (1993).
15. A.L Liechtenstein, V.L Anisimov, J. Zaanen, Phys. Rev. B 52, R5467 (1995).
16. M.T. Hutchings, E.J. Samuelsen, G. Shirane & K Hirakawa, Phys. Rev. BI88, 919
(1969).
17. V. Eyert and K-H. Hock, J. Phys.: Condens. Matter 5, 2987 (1993).
43
Abstract. The electronic structure of undoped cuprates is reanalyzed with the help of the
LDA+U approach of density functional theory. It is found that in-plane oxygen orbitals
forming 'if-bonds with copper are in energetic concurrence to those forming the Zhang-
Rice singlet. The corresponding bands display a surprisingly large c-axis dispersion due
to hybridization with unoccupied cation d-states in the adjacent layers. Hence, they must
have relevance to c-axis magnetic coupling. In view of this situation, the Emery model is
incomplete for the description of magnetism of undoped cuprates.
Key words: Electronic Structure, Undoped Cuprates, Density Functional Theory, LDA
+ U, Strong Correlations, Hubbard Model, Emery Model, Zhang-Rice Singlet, Magnetic
Coupling, t-J Model
The variational principle by Hohenberg and Kohn for the ground state
energy E and spin density n of electrons in a static external field u:
(u I n) = L Jd ru 3
88 m 8 '8 = Jd 3 r(un - B· m), (1 In) = L Jd rn
3
88 ,
88' 8
(2)
holds true in any case of arbitrarily strong correlation. It is based on many-
particle quantum theory by rigorous mathematics. Of course, the density
functional H[n] is unknown.
Since the external field u must be static, it does not cope with non-
adiabatic electron-lattice interaction.
45
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 45–53.
© 2003 Kluwer Academic Publishers
46
The electronic quasiparticle excitations (if they exist) are obtained from
the coherent part (pole term) of the single particle Green's function:
2. Model Functionals
(5)
and 1(n) taken to be the exchange and correlation energy of the homogen-
eous electron liquid.
47
(9)
(10)
The corresponding U-potential is an operator containing a projector:
1 6
(11)
nk 6cP'k eu =
The L(S)DA+U models depend on the basis set of the solver of the KS
equation. We will use for the latter a nonorthogonal local-orbital basis
(1), since a local orbital representation is mandatory for considering strong
correlations. For an LMTO solver see (2), for an LAPW solver (3).
Consider KS orbitals Ik; = IcPk; and orbital occupation numbers nk
as previously; they need not be eigenstates of spin. Let {Il)} 3 Ii) be a
possibly non-orthogonal basis for KS orbitals: Ik; = 2: z ll)Czk' Sll' = (lll').
The occupation matrix ii = ii[cPk' nk] of correlated orbitals Ii) at site R i in
an orthogonal form is introduced as (see Fig. 1)
48
Ii} ~i)
Ik)n1/2 k
Ii} = ~Z Il)(S-l )Zi, (lli} = 6Zi
Il)
Figure 1. Orthogonal projection on the contragredient orbital.
(14)
Averages are:
(15)
This functional was introduced under the name 'around the mean field'
(AMF) in Ref. (4). It is zero if the orbitals of an atomic shell are equally
occupied, hence it depends on orbital polarization. It is given by
} L {(fLafL'-alwlfLafL'-a)nILan/L'-a + (16)
IWILIL~ +[(fLafL~lwlfLafL~) - (fLafL~lwlfL~fLa)] nlLanlLa}
- 2 L { U(N - na) - J ( Na - na) } Na,
IW
Na = L nlLa = (2l + l)na.
/L
The corresponding U-potential is
onlLa
L{ (ItaILalwl/taILa)(fI/L'-a- n-a) +
/L'
+ [(fLafL~ Iw IfLafL~) - (fLafL~ Iw IfL~fLa)] (n/L'a - na)} (17)
49
o·
x
x
~ L {({LU{L'-uliiJl/J,U{L'-u) iiJLuiiJL'-U +
&JLJL'
1 + [( {Lu{Lu/ 1-I / ) - ({Lu{Lu/ 1-I
W {Lu{Lu / )] nJLUnJLU
W {Lu{Lu - - }
-2 ~{UN(N -1) - JJ;Nu(Nu -I)} = (18)
(19)
:> 50 ~ 5.0
~
~
'J 00
>-
~
~
W
-5.0
M A
several orbitals is shown, that is, the square of the coefficient in the expan-
sion of the KS orbital into local basis orbitals is given by the linewidth. It
is clearly seen that both the 0 2pu and 0 2p7r orbitals hybridize equally
strongly with the Cu 3dx L y 2 orbital at the valence band edge. Thereby,
the considerable dispersion in z-direction of the 0 2p7r dominated bands
comes from hybridization with the unoccupied Ca 3dx L y 2 orbitals.
Sr2Cu02Cb is probably the most two-dimensional of all cuprates. The
antiferromagnetic unit cell is shown in Fig. 6 and the same orbital weights
as in the last section are shown on Fig. 7. The situation is quite similar
except that there is practically no dispersion in z-direction.
~ 5.0 r-'----'-::-----4...._Ji""""~-
0.0 ~ 0.0 -
~
~-~---
.-
-50
rXM r
AMF
ZRA
- Z rXM r
AMF
ZRA Z
CaCu021 LSDA+U (U=8 1 eV J=1 eV) MCu=O 71 CaCu021 LSDA+U (U=8.1 eV, J=1 eV) MCu =O.71
~ 50 ~ 5.0_
M M A
Figure 5. Upper panels: orbital weight of the Cu 3d x 2 _y2 orbital and the 0 2p" orbitals,
resp. Lower panels: the same for the 0 2p" and the 0 2pz orbitals.
z
z
x x
--
Sr2Cu02CIZ LSDA+U (U=8.16 eV, J=1 eV) MCu =O.748 Sr2Cu02CIZ LSDA+U (U=8.16 eV, J=1 eV) M Cu =O.748
~ 00
>- I-. ..
~ -2.0
~ ....
W -4.0
-6.0
r X p X, z r r X P Xj Z r
AMF AMF
Sr2Cu02CIZ LSDA+U (U=8.16 eV, J=1 eV) MCu =O.748 Sr2Cu02CIZ LSDA+U (U=8.16 eV, J=1 eV) M Cu =O.748
60 ~ 6.0 -
• 0,
40 4.0 _
~ 2,0 ~ 2.0_
-60 -6.0 =-
Acknowledgements
References
J. ROHLER
UniversitCit zu Koln
Ziilpicher Str. 77, D-50937 Koln, Germany
Abstract. We discuss the static deformations of the doped CU02 lattices in YBa 2Cu 3 0x.
Intrinsic lattice effects driven by the strong correlations of the electron system are separ-
ated from extrinsic chemical and bandstructure effects. c-axis displacements of the planar
copper atoms seem to be a generic property of the metallic CU02 lattices.
Key words: Superconductivity, Cuprates, YBa 2Cu 3 0x, CU02 Lattice, C-Axis, Buckling,
Strongly Correlated Electrons, Doping Mechanism, X-Ray Diffraction, EXAFS
1. Introduction
55
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 55–63.
© 2003 Kluwer Academic Publishers
56
The pioneering work of Muller and Bednorz (1) has been guided by the
idea of high-Tc superconductivity (HTSC) in transition-metal chalcogen-
ides with strong J ahn-Teller (JT) effects. In fact, the strong JT effect is
confirmed to playa crucial role for the electronic structure of the super-
conducting cuprates selecting from the e g doublet the antibonding 3dx 2_ y 2
as the only relevant d orbital in the plane of the Cu-O polyhedron. It
generates the enormous stability of the square-planar CU02 structure. An
electronically active role of the perpendicular 3d z 2 orbital is however not
confirmed.
But superconductivity is not a purely 2-dimensional effect. T c depends
strongly on interactions of the planes with the doping reservoir, and a
mechanism linking the plane with its perpendicular surroundings is expec-
ted to be involved in the formation of the electronic ground state. Recent
calculations of the one-electron bandstructure of YBa2Cu307 (2) show that
the generic low-energy layer-related features have to be described by the
planar orbitals Cu3d x L y2, 02px,y directing in the plane, and the isotropic
Cu4s orbital. The chemical trend of T cmax is found to be controlled by
the energy of a "perpendicular orbital" , a hybrid between Cu4s, Cu3d z 2,
the apical 02pz, and farther d z 2 or pz orbitals. Interestingly materials with
the perpendicular orbital more localized onto the CU02 layers, i.e. with
pure Cu4s character, exhibit the highest T cmax. It seems that the strong
hybridization between Cu3d x L y 2 and Cu4s allows for atomic displacements
in the rigid covalent CU02 lattice of the parent Mott insulator such that
doped holes are optimally accomodated in the background of the quantum
liquid of spin singlets.
Notably the dependence of T cmax from the structure and the chemistry
perpendicular to the planes enters a single-band many electron Hamiltonian
simply by the ratio of the nnn and nn hopping integrals, t f It. It is worthwile
to stress that the (rare) acknowledgement of the chemistry outside the
planes as an integral part of the HTSC problem does not justify dynamically
independent treatments of the Cu3d and 02p electrons in any kind of multi-
band many electron Hamiltonians.
The CU02 planes of the superconducting cuprates are not perfectly flat
but dimpled. Many diffraction work, e.g. Ref. (3,4,6,8), and some EXAFS
studies (9, 10) yield evidence for dimples in the relevant CU02 planes of the
multi-layer cuprates, i.e. the planes adjacent to the doping reservoirs, cf.
table 1. "Buckled" planes are also observed in single-layer materials, most
57
clearly in the doped La2Cu04 systems, but so far not in the single-layer Hg-
and Tl-cuprates (4). As we shall see, the dimplings in the latter materials
are expected to be small, and since their CU02 planes are crystallographic
inflection planes the usual crystallographic refinements tend to average
them to zero.
We distinguish "chemical" dimples driven by external electrostatic po-
larization along c, mixed valence of the dopants, lattice mismatches and
other bandstructure effects, from "correlation-driven" dimples generic of
the strongly correlated electron system. We show that the former define a
large length-scale of rv 0.3 A, and the latter a small length-scale of rv 0.03
A, corresponding to buckling angles of rv 10°, and rv 10, respectively. The
contribution of the chemical dimples in many of the doped metallic phases
can be traced back to the lattice of their antiferromagnetic parent phase,
and thus be differentiated from the correlation-driven dimples. However in
concentrated alloys, doped by varying ratios of one or two combinations of
heterovalent cations, the correlation-driven dimples may be masked by the
varying external electrostatic polarizations (12). Table I compares the CU02
dimplings in some typical multi-layer cuprates with each other: as a function
of the static charge contrast polarizing the CU02 layer perpendicularly, Pl.,
the number of CU02 layers, N L , and T e . The large length-scale of rv 0.3
A shows clearly up in all compounds with Pl. = 2+ - 3+, but is absent
in compounds with P1.. = 2+ - 2+. Materials alloyed with heterovalent
cations, e.g. Ca2+ / y3+, exhibit intermediate values. Notably CU02 layers
sandwiched between isovalent layers seem to exhibit higher T emax than
those between heterovalent layers, cf. Ref. (7).
58
~ 1.65--------------
co Ba'+·O
N Cu2
o 1.6
>-
8 1.55
YBa Cu 0 Cu2
2 3,
>- 5 K
~ 1.5 (O.20)+O.o.BA
'u 02 0,0211 I --- I
g. 1.45 02,3
03
'-----------------!!-~--~------:--;;;----:-;-;;-----i-
1.4 '--'-~~~~~~~~~
and b between x ':::::' 6.8 and 6.9 connect the two regimes suggesting a thermo-
dynamical instable regime. If the orthorhombic strain originated exclusively
from Cul-04 hybridization in the I-dimensional charge reservoir, we expect
the b-, and the a-axis to exhibit similar orthorhombic deformations in both,
the under- and overdoped regimes. The abrupt change from the quadrupolar
,G-ortho to the monopolar ,G-ortho deformation at Xopt points however to
an additional and different mechanism to be operative.
Figure 2b displays the doping dependence of the basal area, B(x) = a·b.
Evidently accomodation of oxygen atoms in the reservoir layer does not
increase B(x) as might be expected from simple stereochemical grounds,
but reduces the basal area. Hence the doping dependence of B(x) is most
likely controlled by the ground state of the quantum liquids in the metallic
planes, and not by the anisotropic oxygen diffusion processes in the non-
stoichiometric chain layer. To discuss B (x) in more detail we find it useful
to associate "'p ex: -0B /onh with a 2-dimensional "electronic compress-
ibility", eliminating the almost doping independent bandstructure effects.
Here nh ex: x denotes the hole concentration.
"'p ':::::' 0 in the insulating phase. The plane of the spin lattice in the
lightly doped Mott insulator seems to be incompressible as indicated by
the arrow for x = 6.0 - 6.38. The magnetic exchange energy J of order
1500 K determines the electron-electron interactions in the Cu2 planes.
"'p > 0 in the metallic phases. While weakly decreasing in the under-
doped a-ortho regime, B (x) starts to collapse in the overdoped ,G-ortho
regime. The plateau between x = 6.8 and 6.9 points to phase segregation,
possibly into incommensurate stripe phases. In the metallic phase the doped
I
3.88 b ~ I
~ I
I ~ 14.88
1 ro
'3 3.86 (J. I
1
~
D I
ro I ~ 14.86
3.84 <Il
a~' I.
I
I
I
•
\.0
\
YBaCuO I· . 14.84 YBa Cu 0
3.82 2 3 x I 2 3, I "'\
300 K I • ~
300 K : opt. opt.
3.8 L.-~~~~~...L....-~~~-....J 14.82 L.-~~~~~~~~~-...J
6.4 6.5 6.6 6.7 6.8 6.9 6.4 6.5 6.6 6.7 6.8 6.9
I
02 I
YBa Cu 0 YBa CU 0 :
1.435 2 3, 1.435 2 3 x
~ 02_'\".
5K ~ 5 K
I
I
'"c: 1.43
'"o 1.43
I
I
I
>- >- ____ I
• .JI _
N
~ ! ' ~
<02,03"'"- - - - - - - : I"
>- >-
~
:? : ~
...... : ....
1.42
I .1 "-
.~ I I
<f) I I
1.415 I I
I I
~~
I I
03 I I
1.41 L . . . . . ...~~~~~~~~~ 1.41 L........~~~~-L.-..-~~...L.L'--'--"-'
6.5 6.6 6.7 6.8 6.9 6.5 6.6 6.7 6.8 6.9
holes hop with a nn matrix element t ex: d- n , n 2:> 2, and thus shorter
nn distances, d, increase strongly the kinetic energy of the holes. But the
physics of the doped Mott insulator is that of competition between the
exchange energy J and the kinetic energy per hole nht. The doped holes in
the underdoped regime appear only as vacancies in the background of a spin
singlet liquid. The lattice of such a strongly correlated t - J type electron
system is expected to be much harder than that of a nearly noninteracting
electron liquid.
Hence the a- and p-ortho deformations may be identified as charac-
teristic lattice responses to fundamentally different types of electron li-
quids in the metallic phase: a weakly compressible t - J like in the under-
and optimum doped, and a strongly compressible Fermi liquid-like in the
overdoped regimes.
a
a
b
02
a a
b
b
iv. In the overdoped regime the deformations have changed from the
quadrupolar 00- to the monopolar ,6-ortho type stressing the a-, b-axes in
the same direction therewith collapsing the basal Cu2 area. Both oxygens
02,3 shift perpendicularily in the same direction along e.
5. Concluding Remarks
Acknowledgements
References
1. Introduction
65
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 65–73.
© 2003 Kluwer Academic Publishers
66
(2)
(4)
where Tlj = 0: + (So·Sj) /0: and rk = (cos k x + cos k y)/2, r~ = cos k x cos kyo
The central quantity for further consideration is the self energy
(5)
(6)
(9)
"() W (10)
X q,w CX(-2
q + '" 2)(W 2 + w'"2)'
69
3. Pseudogap
The spectral functions show in this approximation two branches E±, sep-
arated by the gap which opens along the AFM zone boundary k = kAFM
and the relevant (pseudo)gap scale is
(12)
,6.CO does not depend on t, but rather on smaller J and in particular t'.
For t' < 0 the gap is largest at (7r, 0), consistent with experiments.
Within the simplified effective band approach, it is not difficult to eval-
uate numerically ~lf beyond the QSA, by taking explicitly into account
X"(q,w), Eq. (10), with"" > 0 and w" 2J"". In Fig. 1 we present results
;v
for A(k, w = 0) at T = 0 for a broad range of "" = 0.01 - 0.6. Curves in fact
display the effective FS determined by the condition G- 1 (k F , 0) = O. At
the same time, intensities A(k, w = 0) correspond to the renormalization
factor ZF. At very small "" = 0.01 we see the hole-pocket FS which follows
from the QSA. Already small "" ;v 0.05 destroys the 'shadow' side of the
pocket. On the other hand, in the gap emerge now QP solutions with very
weak ZF « 1 which reconnect the FS into a large one. We are dealing
nevertheless with effectively truncated FS with well developed arcs. The
effect of larger "" is essentially to increase Z F in the gapped region, in
particular near (7r, 0) .
Note that within present theory the low-energy excitations corresponds
to a Fermi liquid, although a very strange one, where QP exist (at low
T) everywhere along the FS. It is quite remarkable to notice that in spite
κ0.01 κ0.04
κ0.2 κ0.6
71
κ0.1
0_2
wit 0 1---------
(n/8,5n/8) (3n/8,7n/8)
k
Figure 2_ Contour plot of spectral functions A(k, w = 0) across the FS in the pseudogap
regime_
4. Conclusions
We have presented our results for spectral functions and pseudogap within
the t-J model, which is the prototype model for strongly correlated elec-
trons. The physics of the t-J model at lower doping is determined by the
interplay of the magnetic exchange and the itinerant kinetic energy of fer-
mions. At intermediate doping the system is frustrated, the effect showing
up in large entropy, pronounced spin fluctuations, non-Fermi liquid effects
72
etc. Evidently, this is one path towards the metal-insulator transition, but
definitely not the only one possible. In our case, fermionic and spin de-
grees of freedom coexist but are coupled and both active and relevant for
low-energy properties.
Within the present theory (13) the origin of the pseudogap feature is in
the coupling to longitudinal spin fluctuations near the AFM wavevector Q.
It is important to note that apart from extremely small /'l, we are still dealing
with large FS. Still, at /'l, « /'l,* rv 0.5 parts of the FS in the nodal direction
remain well pronounced while the QP weight within the pseudogap (zone
corners) region of the FS are strongly suppressed. QP within the pseudogap
have small weight Z F « 1 but due to nonlocal character of ~ (k, w) not
diminished (or even enhanced) VF. This gives an explanation for a well
known theoretical challenge that approaching the magnetic insulator both
electron and QP DOS decrease and vanish.
We presented results for T = 0, however the extension to T > 0 is
straightforward. Discussing only the effect on the pseudogap, we notice that
it is mainly affected by /'l,. So we can argue that the pseudogap should be
observable for /'l,(Ch' T) < /'l,* rv 0.5. This effectively determines the crossover
temperature T*(Ch) approximately as T* rv To (l- ch/ciJ where To rv 0.6J
and ci, rv 0.15.
References
1. for a review see, e.g., M. Imada, A. Fujimori, and Y. Tokura, Rev. Mod. Phys. 70,
1039 (1998).
2. B. Batlogg et al., Physica C 235 - 240, 130 (1994).
3. J.C. Campuzano and M. Randeria et al., in Proc. of the NATO ARW on Open
Pr'oblems in Strongly Cor'related Electron Systems, Eds. J. Bonca, P. Prelovsek, A.
Ramsak, and S. Sarkar (Kluwer, Dordrecht, 2001), p. 3.
4. D.S. Marshall et al., Phys. Rev. Lett. 76,4841 (1996); H. Ding et al., Nature 382,
51 (1996).
5. M.R. Norman et al., Nature 392, 157 (1998).
6. R. Preuss, W. Hanke, C. Grober, and H.G. Evertz, Phys. Rev. Lett. 79, 1122 (1997).
7. J. Jaklic and P. Prelovsek, Phys. Rev. B 55, R7307 (1997); P. Prelovsek, J. Jaklic,
and K. Bedell, Phys. Rev. B 60, 40 (1999).
8. for a review see J. Jaklic and P. Prelovsek, Adv. Phys. 49, 1 (2000).
9. A.V. Chubukov and D.K. Morr, Phys. Rep. 288, 355 (1997).
10. .J. Schmalian, D. Pines, and B. Stojkovic, Phys. Rev. Lett. 80, 3839 (1998); Phys.
Rev. B 60, 667 (1999).
11. P. Prelovsek, Z. Phys. B 103, 363 (1997).
12. C.L. Kane, P.A. Lee, and N. Read, Phys. Rev. B 39, 6880 (1989).
13. P. Prelovsek and A. Ramsak, Phys. Rev. B 63, 180506 (2001); Phys. Rev. B 65,
174529 (2002).
14. A. Kampf and J.R. Schrieffer, Phys. Rev. B 41, 6399 (1990).
15. A. Ino et al., Phys. Rev. Lett. 81, 2124 (1998).
73
16. Ch. Renner, B. Revaz, K. Kadowaki, 1. Maggio-Aprile, and O. Fischer, Phys. Rev.
Lett. 80, 3606 (1998).
17. J.W. Loram, K.A. Mirza, J.B. Cooper, and W.Y. Liang, Phys. Rev. Lett. 71, 1740
(1993); J.W. Loram, J.L. Luo, J.B. Cooper, W.Y. Liang, and J.L. Tallon, Physica
C 341 - 8, 831 (2000).
MOTT INSULATOR TO SUPERCONDUCTOR VIA PRESSURE
RVB THEORY & PREDICTION OF NEW SYSTEMS
G. BASKARAN
Institute of Mathematical Sciences
Madras 600 113, India
Abstract. This paper summarizes my new work, on the nature of Mott transition in
orbitally non degenerate spin-~ system, motivated by observation of Mott insulator
superconductor transition in real systems. The structure of superexchange perturbation
theory for a repulsive Hubbard model at half filling is analyzed and it suggests, un-
der some conditions, a resonating valence bond (RVB) mechanism of Mott insulator to
superconductor transition; charge 'deconfinement' is accompanied by electron pair delo-
calization aided by a preexisting spin singlet correlations in the insulator. An RVB mean
field theory at half filling illustrating our mechanism of Mott insulator-Superconductor
transition is sketched. We identify some family of compounds as potential candidates for
Mott insulator to superconductor transition under pressure: CuO, BaBi03, thin films of
La 2 CuD4 or CaCu02 (infinite layer compound) under pressure or an effective ab-plane
pressure/epitaxial compressive stress; synthesizing new compounds such as La 2CuS202 ,
La 2CuS4 or CaCuS2 or compounds containing CUS2 or CuSe2 planes to mimic large
ab-plane chemical pressure.
Key words: High Temperature Superconductivity, Mott Insulators, Pressure Induced Su-
perconductivity, Resonating Valence Bond (RVB) Theory, Strongly Correlated Electron
Systems, Organic Superconductors, Slave Boson Theory
1. Introduction
75
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 75–84.
© 2003 Kluwer Academic Publishers
76
followed (5) over nearly 15 years have given a picture that is in qualitative
agreements with a variety of experimental results.
Inspired by cuprates, RVE theory has focussed on metallization of Mott
insulating state by means of hole or electron doping. However, it is also
known that undoped commensurate systems such as the family of quasi-
1 dimensional Eechgaard salts (6, 7), quasi-2 dimensional ET salts (7),
and 3 dimensional fullerites (8, 9) do exhibit unconventional and high
T c superconductivity. For these narrow tight binding systems, a half filled
repulsive Hubbard model provides a correct starting point and various the-
ories (10, 11) show possibility of superconductivity near the Mott transition
point. Recent theoretical work by Capone and collaborators (12) bring this
out in a general context. Experimentally one does see, what may be termed
as Mott insulating spin Peierls or antiferromagnetically ordered states that
undergo a transition to a superconducting state as a function of pressure
or chemical pressure. A striking example (9) is K 3 (N H 3 )C60 , a spin-~ an-
tiferromagnet that becomes a high T c superconductor with a T c (:::::; 32 K)
under pressure.
This has prompted us to reanalyze the issue of Mott insulator metal
transition at commensurate fillings from the point of view of RVE theory.
We focus on orbitally non degenerate spin-~ Mott insulators. We argue
that strong local singlet correlations induced by higher order superexchange
processes in the vicinity of Mott transition play a key role in the Mott
insulator superconductor transition and RVE theory is the most natural
way to understand the transition and the superconducting state itself.
Our paper is organized as follows. In section 2, starting from the large
U repulsive Hubbard model at half filling we look at the higher order terms
of the superexchange perturbation theory. The divergence of the perturba-
tion theory suggests correlated singlet pair delocalization in the conducting
state. In section 3, the above argument is substantiated by an RVE mean
field theory directly applied to the repulsive Hubbard model at half filling,
which shows a Mott insulator superconductor phase transition. In section 4
we suggest that some known Mott insulating systems should become high
Tc superconductors under pressure. We also propose some new systems;
they are predicted to be potential high temperature superconductors, either
on their own or under pressure.
77
The low energy physics of an odd electron (per unit cell) system such as
La2Cu04 is well described by a single band Hubbard model:
• • •
•
• i~1
Figure 1. A higher order superexchange process that signals co-operative delocalization
of charge ±2d singlets.
h
were t = V2
bij 1 (t8 8 t - 8 t 8jT t)·IS a spm .. smgIeteIect ron pair. creat'IOn oper-
ir j1 i1
ator at the bond ij. In the singly occupied subspace the first term becomes
the familiar superexchange term: bIjbij == -(Si . Sj - ininj)' Similarly all
the higher order terms also can be written purely in terms of the spin
operators. The prime in the summation avoids double counting.
I have written the effective Hamiltonian in terms of spinon operators,
rather than the customary Pauli spin operators, as it has some advantage
in proving a theorem and helps in seeing the physics of Mott insulator
to superconductor transition more transparently. Further a dynamically
generated local U(l) gauge symmetry of Heff, in the low energy subspace
of the Mott insulator is also manifest (4): bIj ----t eieibIjeiej.
As far as I know there is no rigorous proof for the convergence of
the above perturbation theory; such a proof should be simple as charge
localization is a rather robust phenomenon in a Mott insulator. I present
a rigorous proof for the convergence by using some of Takahashi's results
(13) for half filled band repulsive Hubbard model.
Theorem: For the repulsive half filled band Hubbard model the hopping
parameter expansion converges for U It >> l.
Proof: From Takahashi's result it is clear that (13) the nth order terms of
the spin Hamiltonian corresponds to closed graphs drawn on the lattice,
connected by n non-zero matrix elements tij; and n has to be even. A
loop may intersect and also may self overlap (figure 1). As the effective
Hamiltonian is also an operator cumulant expansion, there are no diagrams
with disconnected graphs.
It can be shown that the most general term in the expansion has the
. h
lorm, 1·t'1"2
r b b bt b bt b ffi . ti1i2ti2i:, ...tin_1intini1
~ 1·'1'"~2 ~"23
1" 1" ~4'"
3"
~'-'n-l'-'n
~ ~"n1'1' WIt a coe clent rv un 1
They are closed loops on the lattice connected by the hopping parameters
tij'S. It is also easy to see that there are prefactors of individual terms which
79
are bounded by 2n . Since the hopping matrix elements are short ranged the
number of loops of size n is bounded by (ad)n, where a is a number of the
order of unity and d is the spatial dimension.
The matrix elements of the operator bl, b, , bl ,. b, , ... bl"n-lCJn
/Jlu2 "1"2 u2lJ,j "3"4
,b",
"11,"1
in
our single occupancy subspace is bounded by 1. Thus the maximum value
of the sum of n-th order terms is bounded by u n - 1 2n (ad)n. This term
t(max)n
charge degree of freedom. We find that this leads to the possibility of Mott
insulator to superconductor transition (without doping) as a function of
pressure, in a natural way as superconductivity has been argued to arise in
a Mott insulator on doping. From our work, which analysis the Hubbard
model directly, we also reproduce known RVB mean field theory results
(3, 5) for Heisenberg and t-J models as interesting and non-trivial limiting
cases.
Superexchange processes in the Mott insulator is kinetic in origin; second
term of equation 2, a part of the kinetic energy term of the Hubbard model,
generates superexchange process. This enables us to perform a Hartree-
Fock type of factorization of the kinetic energy term of the Hubbard model
(equation 2),
involving singlet spinon pair amplitude, 6.:j == (La asIas}o-) , Xij == (La sIasja) ,
and charge amplitudes rlij == (eidj),eij == (eIe) and dij == (dId j ).
The resulting slave boson mean field Hamiltonian has a charge part:
ij ij
In the slave boson mean field theory we diagonalize the above Hamilto-
nian and determine the Hartree-Fock parameters by minimizing the total
energyjfree energy with the global constraint that total number of spinons
eI
and chargeons Li (dI di + e i + La SIaSia) = N, number of lattice sites.
To establish the principle of an RVB mechanism of superconductivity
without doping at a Mott transition point we quote three general results
from our mean field theory: i) Hch , the boson Hamiltonian has a charge gap
E c which vanishes at aU = Uc , as we decrease U from a large value - this is
the desired Mott transition, ii) the RVB mean field transition temperature
Tc(RV B) starts from a value zero at U = 00 and increases continuously as
we decrease U and iii) the chargeons develop anomalous expectation value
(e i ) = (d i ) i- 0 below Uc with a Tc(BE) i- O.
To see the possibility of superconductivity we express the Cooper pair
order parameter in terms of the slave boson variables in the conducting
phase: (La aCiaCjo-) rv (eI) (e})6.ij + (ei) (d})rJij + otherterms. As (ei), (d i )
81
l~~.?(~~~ t super
o u U C O n ________
Figure 2. (a) T-U Phase diagram (Schematic). Tc(BE) is the chargeon Bose condens-
ation temperature, T c (RVB) is the mean field transition temperature of the RVB order
parameter and E c is the charge gap in the Mott insulating state. (b) Schematic Ground
state phase diagram in the U-n plane
are nonzero in the conducting phase and 6.ij and TJij continue to remain
finite, we have superconductivity within the above mean field theory.
Physically we can understand the above as follows. Coherent electron
singlet pair delocalization naturally leads to a finite density of doubly oc-
cupied and empty sites present in a phase coherent fashion. In terms of
the slave boson picture, presence of real empty and doubly occupied sites
in the ground state in a phase coherent fashion corresponds to the bose
condensation of chargeons.
Figure (2a) illustrates the region where superconductivity is likely to
occur in the T vs U plane. Figure (2b) sketches the ubiquitous supercon-
ducting region in the U vs n plane.
Slave boson mean field as applied directly to the Hubbard model in the
present paper has to be understood carefully. First, as in the RVB mean
field theory applied to Heisenberg Hamiltonian, we get electron pair con-
densation in the Mott insulating state. This has to be properly interpreted,
in view of the presence of charge gap present even in our mean field theory.
Use of gauge fluctuations eliminates this fictitious superconducting order in
the Mott insulating state, consistent with Elitzur's theorem. Secondly, our
mean field theory produces superconductivity even as U ---7 O. This is an
artifact of Hartree-Fock factorization. However, we believe that our results
are qualitatively correct close to the Mott transition point. As pointed
out earlier, nesting instabilities in the conducting phase can be avoided by
appropriate choice of tij'S, the band parameters.
In this paper we do not go into the detailed energetics of mean field
solutions corresponding to different symmetries, for example extended-s
and d symmetry of the spinon pair order parameter 6.ij and the charge
pair order parameter rlij.
82
Acknowledgements
References
MICHAEL LANG
Johann Wolfgang Goethe- University, FOR412
D-60054 Frankfurt am Main, Germany
JENS MULLER and FRANK STEGLICH
Max-Planck Institute for Chemical Physics of Solids
D-01187 Dresden, Germany
JOHN SCHLUETER
Material Science Division, Argonne National Laboratory,
Argonne, Illinois 60439, U.S.A.
TAKAHIKO SASAKI
Institute for Materials Research, Tohoku University,
Sendai, Japan
1. Introduction
85
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 85–93.
© 2003 Kluwer Academic Publishers
86
T (K)
o 20 40 60 80 100 120 140
80 • K-Cl
o K-Ds-Br 20
60
80
60
15 50
~
60
~ 40
10 40
''b
-
'-'
1:t 20 40 5 30
T(K)
26 28 30 32
r- o T 20
·,.r"·
o TN
• K-Hs-Br
-20 - o
-10 I-.t-,----'-----,--"--,---'---r---r---.----r---!
o 40 60 80 100 120
T(K)
Figure 1. Cross-plane thermal expansion coefficient for various K;-(EThX salts; H s (D s )
indicate protonated (deuterated) ethylene endgroups. Inset shows the phase transition
into the AFM ground state for K;-Cl.
tY
65 t
+
r
65
i #'
60
JI
o.j
.~ 55
·~60
j II
50
ij' ~ 20
~
iI
D
I
45
55
18
a 35 b 16 c
55 60 65 70 75 80 85 60 65 70 75 80 85 90 40 45 50 55 60 65 70 75
T~ T~ T~
Figure 2. Blow-up of the cross-plane thermal expansion coefficient around the glass-like
transition for various K;-(EThX salts. Arrows indicate cooling and heating runs.
,...
1'. K-Cu(NCS),
•• c-axis
.- --
.....
b-axis
~\
x
~
c-axis jT,
~-
a -axis U- layers) rP I \
1T .
o 10 20 30 40 50 60 0 10 20 30 40 50 60
T (K) T (K)
Figure 3. Anomalous contributions to the uniaxial thermal expansion coefficients at T c
and T* derived by subtracting a smooth background contribution.
O.72mg
0.8
Tc =9.15K
0.6
Q 00
o 2 4 6 8 10 12
T(K)
Figure 4. Specific heat difference !:::.C = C(OT) - C(8T) = C(OT) - Cn, where C n
denotes the N-state specific heat. Dotted and solid thick lines represent BCS curves
for weak and strong coupling, respectively. Inset shows the electronic contribution in a
BCS-plot.
The present study discloses important new aspects which impose clear
constraints for both experimentalists and theorists attempting to unravel
the nature of the states below and above T c in these K;-(EThX salts. At
elevated temperatures, a glass transition has been identified that defines
the boundary between an ethylene liquid at T > T q and a glassy state
at T < T g where a certain amount of disorder in the terminal [CH 2 b
units becomes frozen. With increasing cooling rate through T g , the level of
disorder (10 - 20 %) increases and may substantially influence the electronic
properties at low temperatures, especially the SC state.
At intermediate temperatures T*, our studies disclose a second-order
phase transition for the metallicjSC salts. We propose that instead of a
pseudogap on the major 2D fractions of the Fermi surface, T* is associated
with the formation of a density wave, i.e. a real gap, on the minor 1D parts
which competes with superconductivity for stability.
92
ethylene liquid
100
g T~ I PMIDW I
E-< 10 1--~
I '
X~
Cu[N(CN),]Cltt Ds Hs
- Cu[N(CN)2]Br CU(NCS)2 I kbar
Figure 5. Compilation of thermal expansion anomalies for the various /'i;-(EThX salts in
a pressure-temperature phase diagram. Arrows indicate the location of the different salts
at ambient pressure. AFI, PM/DW and SC denotes antiferromagnetic insulator, para-
magnetic metal in coexistence with a density-wave state and superconductor, respectively.
Solid lines represent the hydrostatic pressure dependencies of TN and T c taken from the
literature.
References
Abstract.
High-temperature superconductivity in doped Mott insulator cuprates denies the
conventional belief that electron repulsion is detrimental to superconductivity. We discuss
a parallel situation in the alkali fullerides which, even if lacking the cuprate's spectacular
Te's, are no lesser puzzle. The repulsive electron correlations in fullerides are of com-
parable strength to the cuprates: yet trivalent fullerides superconduct, and that with an
s-wave order parameter. We discuss theoretical studies of a model for alkali fullerides
which reveals, in presence of a weak pairing attraction (here of Jahn Teller origin), an
unsuspected superconducting pocket near the Mott insulating state. The peculiar pairing
mechanism in this model shows that close to an ideal, singlet Mott state, superconducting
pairing can manage not just to avoid frustration by strong electron-electron repulsion,
but actually to take vast advantage from it. Although derived by solving a very specific
tetravalent fulleride model, we argue that this mechanism is in fact common to a wider
class of strongly correlated models, including those proposed for cuprates (1).
95
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 95–113.
© 2003 Kluwer Academic Publishers
96
1. Introduction
then
m*
A* ':::::' -A> A,
m
which might indeed produce a T c enhancement.
This hypothetical possibility is based on the assumption that Landau
Fermi liquid theory is valid well below T c . However, when the mass en-
hancement pushes A* rv 1, T c becomes of the order of the quasiparticle
97
After this long introduction, let us describe the model for alkali doped
fullerenes we studied. The Hamiltonian is
(1)
a-ibeing the Pauli matrices. In terms of these operators the Hund term
reads
can and does (10, 21) reverse Hund's rule, generally favoring instead low
spin and angular momentum configurations. We include approximately this
effect by assuming JH < O. This is formally equivalent to treating the
electron-vibron coupling within an antiadiabatic approximation, where it
can be shown to renormalize JH ----+ JH - 3EJT/4, EJT being the Jahn-
Teller energy gain. The antiadiabatic approximation is more justified than
it may seem in strongly correlated fullerides, where the vibron frequencies
of order 0.1 eV, is not in fact larger, but rather comparable, with the
bare electron hopping Itl. Vibrational frequencies are in fact safely larger
than the renormalized hopping Ziti close to the Mott state, where Z « 1.
In any case, the neglect of retardation involved in replacing the true JT
coupling with a negative JH does disfavor superconductivity, by preventing
high energy screening of the Coulomb interaction U. Therefore, adoption
of the JH < 0 approximation in place of the fully retarded JT effect is
conservative, and makes physical sense in this problem.
Let us for a moment consider the isolated molecule with JH < 0 and
with two valence electrons, or, equivalently, four (having in mind K 4 C 60
type compounds). The ground state has quantum numbers 8 = L = O.
The spin gap to the spin triplet state with quantum numbers 8 = L = 1 is
5IJHI. We may therefore fix an approximate JH magnitude by comparing
this value of the spin gap with the exact singlet-triplet gap for the isolated
molecular ion. The n=4 C 60 ion has of course a more favorable Jahn Teller
distortion in the 8=0 ground state than in the 8=1 excited state, and this
gives rise to a spin gap tlspin r-v 0.1 eV. This value is quite comparable with
that observed by magnetic resonance experiments in K 4 C 60 . The compar-
ison leads to a crude estimate JH c:::: -0.02 eV, namely JH/U c:::: -0.02,
the value we adopt hereafter. While this will constitute a very weak pairing
attraction, we shall see that it may still give rise to a remarkably important
consequences.
Since all relevant interactions, namely U and JH, are local, the model
described by the Hamiltonian (1) is suitable to be treated by the so-called
Dynamical Mean Field Theory (DMFT) (22), which is a very efficient tool
to deal with strong local correlations. We studied (1) by DMFT with an
average number of electrons (holes) per site n = 2 at fixed JH /U = -0.02
but at various ratios of U /W. Before presenting the results, let us analyse
the behavior of (1) in the two extreme limits U /W « 1, weak coupling,
and U/W » 1, strong coupling.
In the weak coupling limit, the Hamiltonian (1) describes a metal with
three 1/3-filled degenerate bands. Since JH < 0 is an explicit attraction
present in the model, one needs to check the stability of the metal to-
wards Cooper pairing. In the absence of U, JH would indeed lead to a
102
Mon
METAL SC INSULATOR
I I
I I
o 0.8 0.9 UIW
Piyu.re 1. Pha:;e diagram a:; a funct.ioll of U flV. SC st.ands for superconductor
0.12 around W 1.18U, reaching at this value almost the largest possible
("v
104
0.06
IU "" -50 J1
0.06 t
Ll sPin /20
-<
...........
<I
r:e
0.04
METAL
\
103x~U=O
0.02 MI
Figure 2. Superconducting gap (square points) and spin gap reduced by a factor 20
(circles) as functions of IJH I/W at fixed U = 50lJH I. Also drawn is the superconducting
gap at U = 0, 6u=o, as function of IJHI/W, multiplied by one thousand.
0.2) U.
(Z-T (6)
While at weak coupling, Z "':' 1 the scattering amplitude A* is repulsive,
at sufficiently strong correlation, such that Z <::: Zc "':' 0.067, the repulsion
may turn into attraction and the system becomes unstable towards super-
conductivity. The conjecture (6) for the quasiparticle scattering amplitude
also explains the large values of the gap compared with BCS values. Since
the Mott transition occurs for W "':' U, when Z is further reduced down
to rv Zc/2, the effective attraction A* gets of the same order as the quasi-
particle bandwidth, W* = ZW "':' ZcW/2 "':' ZcU/2 = 10IJHI/6. In the spirit
of Landau Fermi liquid theory, this should correspond to the maximum of
the superconducting gap. The actual gap value, ~ rv W*/10 = IJHI/6
linear in the bare attraction, is in fact compatible with that expected at
the maximum of the ~ vs. >. curve for a generic unretarded attraction at
U = 0 (24). The notable difference is that this value is here obtained for
a bare bandwidth W rv 601JHI, namely for a bare>. rv 0.056, where BCS
theory would rather predict ~BCS rv W exp (-1/ >.) much smaller than the
actual value we obtain ~ "':' W >./20.
lk = L L 4,io- (La)
i,j=l 0-
ij Ck,jo-'
and
3
(la (J"b) k = LL 4,ia (La) ij o-~(jCk,j(j'
i,j=l a{3
w~ = L L 4,io- (Wd)
i,j=l 0-
ij Ck,jo-'
t=LL
3
0.6
•
0.2ITJ
0.1
0.6
•
• 0
0.7 0.8 0.9
0.4
0.2
Figure 3. Quasiparticle residue for the metallic solution. The inset shows the enlarged
region where superconductivity appears, which is signalled by a vertical line at U c::: 0.8W.
The line is the conjectured A* of Eq. (6) which indeed crosses zero at U c::: 0.8W when
Z ~ 0.067.
Here F = FS(A) , HS(A), GS(A) depending on the form of the external field.
By calculating within DMFT the quasiparticle residue Z as well as all the
six susceptibilities, we obtain all F-parameters.
Fig. 3 shows the decrease of Z in the metallic solution on approaching
the Mott transition (MIT). Superconductivity sets in at Z very close to our
conjectured value Zc = 0.67.
The charge compressibility, "', is a decreasing function of U, in agree-
ment with the system approaching the incompressible insulator, see Fig.
4. Through Eq. (8) this implies that F S is always positive and diverges
faster than l/Z at the MIT, actually like 1/Z 2 . All other susceptibilities,
(see for example Fig. 4 where the spin susceptibility is shown), initially
increase due to Stoner enhancement, then turn around at U /W rv 0.7, and
finally vanish at the Mott transition, consistent with the spin and orbital
gaps. This implies that FA, HS(A) and GS(A) all start negative and decrease
roughly like -U/W, as can be perturbatively checked, until U rv ZW, when
they turn upward, cross the zero and finally diverge as + 1/ Z2 at the MIT.
The Landau F-parameters are related to the quasiparticle scattering
amplitudes strictly only at low frequencies and momentum transfers, hence
it is not rigorously justisfied to use them to estimate the scattering amp-
litudes in the Cooper channel, primarily a high momentum transfer process.
Yet the F-parameters have been often used to that purpose, for instance
to justify the triplet p-wave pairing in He 3 . In our case, this approximation
108
6
0.6
~g
.~
"-
~
><
g
0.'
2
0 0
0 0.' 0.6
U/W
Z [ FS pi 11 5 lJA CS CA ]
3
A= 12p 1+FS- 1+pA -31+HS+91+HA +51+Cs-151+GA .
(9)
Fig. 5 shows the scattering I:unplitude A a,s calculated through the above
formula.
At weak coupling, A = U -\- lOJH/3 > 0 is repubive and increases with
UjW until at UIW '" 0.5 it turns downwards and crosses zero signalling
a sllperconducting instability: one which we actually find when allowing
gauge-symmetry breaking. As the l..,lIT is approached, all F-parameters
diverge hence
Z
A-----jo--Z·
p.
saturat.es to a value about equal to half the quasipart.icle bandwidth W.. =
ZlV. For comparison we have also drawn in Fig. 4 our conjectured A.. of
Eq. (6). Up to the metal-to-superconductor transition, A.. seems \.0 agree
quite well wit.h A, alt.hough close to the ~i[1T A.. would remain constant
while A vanishes as Z. If the estimate of A through t.he F-parameters is
at least qualitatively correct, this behavior would imply that the attraction
among the quasiparticles cannot exceed t.heir bandwidth.
109
0.08 .----------r--.--------,,....-----,
0.06
~
<I 0.04
o
Q.
METAL
0.02 MI
O ......c=....---c::F--a____-....----4I
0.012 0.02
Although the expression (9) seem to correctly reproduce the onset of su-
perconductivity, its validity needs further justification. As Nozieres pointed
out to us (25), the same scheme applied to the single band Hubbard model
would lead to an incorrect result. In the single band Hubbard model, the
oS-wave quasiparticle scattering amplitude within the same approximation
is
(10)
Xloeal 1 1
(0)
Xloeal
Z 1 + Floeal
This would still imply Fl~cal rv 1/Z2, but instead Fl~eal = canst. < 0,
namely a repulsive A > 0 at the MIT. This is obvioulsy a more realistic
prediction.
In our model there is in fact no such ambiguity since the uniform and the
local susceptibilities behave exactly in the same way, hence both procedures
lead to the same conclusion A rv -ZW/2 near the MIT. Moreover the use of
local Landau parameters might be justified within DMFT where the lattice
problem is mapped onto an Anderson impurity model where the coupling
to the conduction bath is determined self-consistently by imposing that the
conduction-electron local Green's function coincides with the impurity one.
At self-consistency one can indeed imagine that a Fermi liquid theory for
the effective Anderson impurity model should be valid.
We conclude this overview of the Fermi liquid properties of the Hamilto-
nian (1) close to the Mott transition, by studying another quantity which
sheds some light on the real physical behavior. For our choice of a very small
JH /U = -0.02, superconductivity arises when the Mott-Hubbard bands
are already well separated from the narrow quasiparticle resonance. This
suggests that the effective Anderson impurity model has already flowed in
the Kondo regime, where a two-component description might be used (26).
Let us consider the probability P( n) for a site to be occupied by n electrons
in the ground state. In Fig. 6 we draw P(n) for the superconductor (panels
B to D) and for the nearby Mott insulator (panel A), which are quite
similar. In spite of the large superconducting gap, there is no evidence of
preformed pairs or bipolarons as underlined by the strong steady peaking
of P(n) at n = 2. We assume the two-component form of P(n)
where Pins (n) is the probability distribution in the nearby Mott insulator.
Since Z, P(n) and Pins(n) are all known, this allows to extract the quasi-
paricle probability distribution pq;C)(n), which is drawn in Fig. 6A'-6D'.
It shows oscillations between even and odd n's, as expected for a supercon-
ductor, with no relevant variations as a function of U /W upon approaching
the Mott transition (U/W c:::' 0.9), in agreement with an intermediate
coupling regime as implied by A rv - W*. To check the consistency of the
(METAL)
two-component form, we also extracted the Pqp (n) from the (meta-
stable) metallic solution obtained by preventing gauge-symmetry breaking.
The results are very similar to the probability distribution of the free Fermi
111
u• O.
0.6
SC
0
0 2 . 80 2 .. 80 2 . 8
n n n
Figure 6. Occupation probabilities for the particles, P(n), and for the quasiparticles in
the superconducting and in the metallic phases, n (n) and pJ;;c
I
pJi:
ET AL) (n), respectively.
4. Conclusions
We have shown in a simplified model for alkali doped fullerides that strong
correlations may enhance superconductivity just close to the Mott trans-
ition. The detailed analysis of the results indicates that such effect appears
when the pairing occurs in a channel not involving the charge degrees of
freedom which are most affected by the repulsion. Then superconductivity
is automatically pushed to an intermediate coupling regime where the qua-
siparticle bandwidth, strongly suppressed by correlations, gets of the order
of the attraction, which stays unrenormalized.
112
In our model (1) this newer pairing mechanism, able not only to avoid
U but to take advantage from U, is provided by an inverted Hund's rule
mimicking the J ahn-Teller coupling to molecular vibrations. However the
general mechanism is quite general and seems valid in other cases too, as
for instance in t - J models for cuprates.
Acknowledgements
This work was sponsored through MIUR COFIN, by INFM/G, and by the
EU, through contract ERBFMRXCT970155 (FULPROP). We are grateful
to P. Nozieres and to G. Santoro for illuminating discussions.
References
25. P. Nozieres, private communication and preprint (2002, to appear in Jour. of Stat.
Phys.).
26. G. Kotliar, Em. Phys. J. B 11, 27 (1999), and references therein.
INELASTIC LIGHT SCATTERING AND THE CORRELATED
METAL-INSULATOR TRANSITION
J.K. FREERICKS
Department of Physics, Georgetown University, Washington,
DC 20057, U.S.A.
T.P. DEVEREAUX
Department of Physics, University of Waterloo, Canada
R. BULLA
Theoretische Physik III, Elektronische Korrelationen und Mag-
netismus, Institut fur Physik, Universitiit Augsburg, D-86135,
Augsburg, Germany
1. Introduction
115
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 115–122.
© 2003 Kluwer Academic Publishers
116
2. Formal Development
(1)
Figure 1. Coupled Dyson equations for the inelastic light scattering density-density
correlation functions described by the scattering amplitude fa. Panel (a) depicts the
Dyson equation for the interacting correlation function, while panel (b) is the supple-
mental equation needed to solve for the correlation function (the difference in the two
equations is the number of fa factors). The symbol r stands for the local dynamical
irreducible charge vertex. In situations where there are no charge vertex corrections (like
B lg scattering along the zone-diagonal), the correlation function is simply given by the
first (bare-bubble) diagram on the right hand side of panel (a).
where the 4(k) and ca(k) operators create or destroy an electron with spin
(J" and momentum k and n(v) = 1/[1 - exp(v/T)] is the Bose factor. The
(4)
Here ei,s denote the incident, scattered photon polarization vectors, re-
n
spectively, and we have chosen units kB = c = = t* = 1 and have set the
hypercubic lattice constant equal to 1. We can classify the scattering amp-
litudes by point group symmetry operations. If we choose e i = (1,1,1, ... )
and e S = (1, -1, 1, -1, ... ), then we have the BIg sector, while e i = e S =
(1,1,1, ... ) projects out the A Ig sector since the B 2g component is identically
zero in our model due to the inclusion of only nearest-neighbor hopping (1).
Hence iAl g (q) = -E( q) and iBl g (q) = t* L~I cos qj(-I)j / /d.
The Dyson equation for the density-density correlation function takes
the form given in Figure 1. Note that there are two coupled equations
illustrated in Figures 1 (a) and (b); these equations differ by the number
of ia factors in them. The irreducible vertex function r is the dynam-
ical charge vertex which is known explicitly only for the Falicov-Kimball
118
model (6). If the scattering amplitude r does not have a projection onto
the full symmetry of the lattice, then there are no vertex corrections from
the local dynamical charge vertex (7). This is the only case that can be
analyzed for the Hubbard model.
Let's begin our discussion on the imaginary axis in the BIg sector. If we
restrict ourselves to the zone diagonal, then q = (q, q, q, ... , q). Examining
the diagrams in Figure 1 (a), we see that there is a bare response plus a
vertex correction term. The vertex correction term, however, must vanish,
because the leftmost piece of that diagram is
1
x (5)
iW n + iVI + fL - I;(iw n + iVI) - f(k + ~q)
for "energy" transfer iVI = 2in-Tl and momentum transfer q [in Eq. (5)
iW n = iJrT(2n + 1) is the fermionic Matsubara frequency and fL is the
chemical potential]. It vanishes, because each term indexed by j is equal
in magnitude, but opposite in sign, so the overall summation is equal to
zero. Hence the BIg response on the zone diagonal (including the Raman
response) is given by the bare bubble.
The bare bubble on the zone diagonal is simple to calculate directly (we
shift k ---+ k + q/2):
",",",",'
xo(q, iVI) = -T ~~ 11m ~
"'"' cos(ki
d
+ 2q)cos(k
1
j + 2 q )(-1) J
1 i+"
d---+oo" " 1 d
n k ~,J=
1
x
iW n + fL - I;(iw n ) - f(k)
1
x (6)
iW n + iVI + fL - I;(iw n + iVI) - f(k + q)'
Now, the terms with i # j are all equal in magnitude, but there are as many
positive as negative, so they vanish-only the terms with i = j survive. If
we assume that nand n + l are both larger than 0, and define Zn = iW n +
fL - I; (iw n ), then we can rewrite the fractions as integrals of exponentials
d cos2(k + lq)
XO(q,iVI) = TLL lim L ~ 2
n k d---+oo j=1
X 1 d>.l00 OO
To evaluate this integral, we first expand the functions f(k) and f(k + q)
in terms of the Cartesian momentum components. Then it is obvious that
119
each j term is equal in magnitude, so the sum over j is trivial. The next step
is to expand each exponential factor that has a l/vId prefactor in a Taylor
series expansion, and keep the lowest nonvanishing terms in the multiple
integrals (8). The multiple integral over momentum then becomes
limd---+oo J J J
dk 1 dk2... dk d cos (k 1 + ~)
2
d. . ~2
II (
j=1
1 + _L_~ cos k·
vId J
+ _L_~ cos(k + q) -
vId J
- cos 2 k·
2d J
~X X2
- cos kj cos(kj + q) - - cos 2(k j + q) + ...). (8)
d 2d
Here we have chosen the extra prefactor to lie in the 1 direction. Each
integral over kj , except the first, is equal to each other and equal to 1 -
(~2 + 2~~/X(q) + ~/2)/4d + .... The first integral becomes ~[1 _ (~2 +
2~X X(q)+X2)/4d-(~X +X2 cos q) sin 2 q/8d+ ...]. Here we use the notation
X (q) = limd---+oo "'L1=1 cos qj/ d (= cos q for a zone-diagonal momentum).
The next step is to rewrite each factor as an exponential, and then take the
infinite product. The result for the integral over k is
since the terms proportional to sin 2 q coming from the I-direction are just
a 1/ d correction. The end result for the susceptibility is then
XO(q,iVl) = -1 J 1
dEp(E) Zn _ E vI ~ X2Foo (Z~:l_-:;E), (10)
with F00 the Hilbert transformation of the DOS: F00 (z) = .f dEp( E) / (z - E).
The analytic continuation of this expression is straightforward, and
produces the final result for the BIg response
[f(w) - f(w
(11)
with
j
oo 1 1
Xo(w; X, v) - dEp(E) ( )
. -00 W + It - I: w - E VI - X2
x F (w+V+f.l-I:(W+V)-XE) (12)
00 VI _ X2 '
120
0.3
OJ
(J)
e 0.25
0
0...
(J)
OJ
0.2
l-
e 0.15
0 -- 0.566
E 0.1 -- 0.424
0 -- 0.283
0::: -- 0.141
0.05 -- 0.071
0
0 2 4 6 8 10
Frequency [t*]
Figure 2. Nonresonant BIg Raman response (X = 1) for different temperatures at
U = 4.24 and half filling. The numbers in the legends label the temperature.
and
00 1 1
Xo(w; X, v) - dEp(E)
( )
/ -00 W + JL - I;* W - E VI - X2
x F (W+V+JL-I;(W+V)-XE). (13)
00 VI _ X2
Here f(w) = 1/[1 + exp(wIT)] is the Fermi factor.
A similar, but more complicated analysis can be performed for the A Ig
response, or the BIg response off of the zone diagonal, but we don't have
enough space to report those results here, and they cannot be analyzed
numerically for the Hubbard model.
3. Results
~
X=-1
(f)
-+-'
C
::J
...0
L
0
L........I
Q)
(f)
C
0
0...
(f)
Q)
L
>.
0
L
I
x
o 2 4 6
Frequency [t *]
Figure 3. Nonresonant Big inelastic X-ray response for different temperatures at
U = 3.54 and pc = 0.9. Five values of the transferred photon momentum are plotted,
each shifted by an appropriate amount, and running from the zone center (X = 1) to
the zone boundary (X = -1) along the zone diagonal. The temperature decreases with
decreasing thickness of the lines and ranges from 0.503 to 0.114 to 0.042 to 0.026.
4. Conclusions
Acknowledgements
References
1. J.K Freericks and T.P. Devereaux, J. Condo Phys. (Ukraine) 4, 149 (2001); Phys.
Rev. B 64, 125110 (2001).
2. J.K Freericks, T.P. Devereaux, and R. Bulla, Acta Phys. Pol. B 32, 3219 (2001);
Phys. Rev. B 64, 233114 (2001); unpublished.
3. T.P. Devereaux, G. E. D. McCormack, and J.K. Freericks, submitted to Phys. Rev.
Lett. (cond-matj0208049).
4. J.C. Hubbard, Proc. Royal Soc. London, Ser. A 276, 238 (1963).
5. W. Metzner and D. Vollhardt, Phys. Rev. Lett. 62, 324 (1989).
6. A. M. Shvaika, Physica C 341-348, 177 (2000); J. K Freericks and P. Miller, Phys.
Rev. B 62, 10022 (2000); A. M. Shvaika, J. Phys. Stud. 5, 349 (2002).
7. A. Khurana, Phys. Rev. Lett., 64, 1990 (1990).
8. E. Muller-Hartmann, Int. J. Phys. B 3, 2169 (1989).
9. R. Bulla, Phys. Rev. Lett. 83, 136 (1999); R. Bulla, T. A. Costi, and D. Vollhardt,
Phys. Rev. B 64, 045103 (2001).
10. P. Nyhus, S.L. Cooper, and Z. Fisk, Phys. Rev. B 51, R15626 (1995).
11. P. Nyhus, S.L. Cooper, Z. Fisk, and J. Sarrao, Phys. Rev. B 52, R14308 (1995);
Phys. Rev. B 55, 12488 (1997).
12. X. K Chen, J.G. Naeini, KC. Hewitt, J.C. Irwin, R. Liang, W.N. Hardy, Phys. Rev.
B 56, R513 (1997); J.G. Naeini, X.K Chen, J.C. Irwin, M. Okuya, T. Kimura, and
K Kishio, Phys. Rev. B 59, 9642 (1999); M. Rubhausen, O.A. Hammerstein, A.
Bock, U. Merkt, C.T. Rieck, P. Guptasarma, D.G. Hinks, M.V. Klein, Phys. Rev.
Lett. 82, 5349 (1999); S. Sugai and T. Hosokawa, Phys. Rev. Lett. 95, 1112 (2000);
M. Opel, R. Nemetschek, C. Hoffmann, R. Philipp, P.F. Muller, R. Hackl, 1. Tutto,
A. Erb, B. Revaz, E. Walker, H. Berger, and L. Farro, Phys. Rev. B 61, 9752
(2000); F. Venturini, M. Opel, T. P. Devereaux, J. K Freericks, 1. TuttO, B. Revaz,
E. Walker, H. Berger, L. Forro and R. Hackl, Phys. Rev. Lett. 89, 107003 (2002).
ORBITAL PHYSICS VERSUS SPIN PHYSICS:
THE ORBITAL-HUBBARD MODEL
Abstract. To elucidate the similarities and differences between orbital and spin degrees
of freedom, we analyze an orbital-Hubbard model with two orbital flavors, corresponding
to pseudospin 1/2, and contrast its behavior with that of the familiar (spin-1/2) Hubbard
model. The orbital-Hubbard model describes a partly filled spin-polarized e g band on a
cubic lattice, as occurs in ferromagnetic manganites.
We demonstrate that the absence of SU(2) invariance in orbital space has important
implications - superexchange contributes in all orbital ordered states, the Nagaoka
theorem does not apply, and the kinetic energy is enhanced as compared with the spin
case. As a result orbital-ordered states are destabilized by doping, and instead a strongly
correlated orbital liquid with disordered orbitals is realized.
123
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 123–131.
© 2003 Kluwer Academic Publishers
124
fore strongly frustrated - the quantum fluctuations are then enhanced, and
might even lead to a spin liquid state in LiNi0 2 (4). Another possibility,
that an orbital liquid (OL) is stabilized and coexists with long-range spin
order, was pointed out recently for t2g systems (5). In contrast, the SE in
undoped e g systems, KCuF 3 (2) and LaMn03 (6), favors alternating orbital
(AO) order which coexists with the antiferromagnetic spin (AS) order.
In this paper we study a generic model of correlated electrons in a spin-
polarized e g band, with two orbital flavors described by a pseudospin 1/2
in the orbital Hilbert space, and consider its relation to the (spin) Hubbard
model for electrons with spin s = 1/2. We investigate: (i) in what respect
long-range order in orbital systems is different from that in spin systems,
and (ii) whether the orbitals are ordered or rather form a disordered OL.
These questions are both of fundamental nature and of immediate interest
for understanding the metallic ferromagnetic manganites (3, 7). Our pur-
pose here is to elucidate the physical mechanisms which operate in the eg
band by contrasting them with those known in spin systems.
So we consider spinless eg electrons [in a ferromagnetic spin (FS) state]
on a cubic lattice with kinetic energy
(1)
where hopping t between sites i and j occurs only for the directional orbitals
I(a; oriented along the bond (ij;, i.e., I(a; ex: 3x 2 - r 2 , 3 y 2 - r 2 , and 3z 2 -
r 2 , when (ij; is along the cubic axis Q = a, b, and c, respectively. To
describe the local electron interactions one needs to choose an orthogonal
basis for the two orbital flavors. One possibility is Ix; == x 2 - y2 and Iz; ==
(3z 2 -r 2 )/J3, called real orbitals, whereupon the local Coulomb interaction
becomes Unixniz. The disadvantage is that the expression for the kinetic
energy then takes a different form for each axis (8). We thus prefer to
use instead the basis of complex orbitals 1+; = ~ (Iz; - ilx;) and 1-; =
~(Iz; + ilx;), corresponding to "up"
and "down" pseudospin flavors, and
write the eg orbital-Hubbard model in the form
+e+iXn Ci_Cj+
t )] + U '""
L.J ni+ni-' (2)
i
with Xa,b = ±27f /3, Xc = 0, and r = 1. The phase factors e±ixn are
characteristic of the orbital problem - the orbitals have an actual shape
in real space so that each hopping process depends on the bond direction.
125
which corresponds to cyclic permutation of the cubic axes, leaves the Hamilto-
nian Eq. (2) invariant. (ii) It exhibits clearly the difference between the spin
case and the orbital case, and in fact allowed us to introduce the parameter
, by which one can turn the eg-band orbital-Hubbard model h = 1) into
what looks formally like a spin-Hubbard model h = 0). (iii) It shows
explicitly that rotational SU(2) symmetry for the pseudospins is absent
(2). The total pseudospin operator yz = 2:i Tiz , with Tiz = ~(ni+ - ni-), is
conserved only at , = 0 (i.e., [yz, 'H] = 0), while the terms ex: , commute
instead with the staggered pseudospin operator Yq = 2:i exp(iQ . R i )7iz ,
where Q = (]f,]f,]f).
Because the electrons interact by the local Coulomb interaction U, they
are prone to instabilities towards orbital order, which we compare with
magnetic instabilities in the spin case (9). At half-filling (n = 1) the simplest
possibility to reduce the interaction energy ex: U would be to polarize the
system completely into ferro orbital (FO) states, I<I>FO) = I1i cl ('ljJ, B) 10)
(uniform order (10)). As in the spin case, another possibility is AO order,
I<I>Ao) = I1iEA cl ('ljJA, BA)fl jE B C}('ljJB' BB)IO), i.e., with orbitals alternating
between two sublattices A and B. If the band is partly filled (n < 1), such
states must involve a coherent mixture of occupied and empty sites.
It is an important feature of the I+)-polarized (FO+) and 1+)/1-)-
staggered (AO±) complex states, that they retain cubic symmetry (11).
By contrast, the FO and AO real states, such as Ix)-polarized (FOx), Iz)-
polarized (FOz), or (Ix) + Iz) )/(Ix) - Iz) )-staggered (AOxz), with either
quasi-one-dimensional (FOz and AOxz) or two-dimensional (FOx) disper-
sion, break cubic symmetry (and are thus favored in lower dimensional
systems, e.g. FOx in a 2D square lattice (12)). This nonequivalence between
real and complex states is a manifestation of the broken SU(2) symmetry.
We focus here on the orbital-ordered states with complex orbitals (13),
which occur in Hartree-Fock (HF) approximation for large enough U (11).
In the case of the FO+ state, the electron bands split above a critical
value of U, and lead (14) to a finite order parameter T Z = (Tn of- 0, but the
mechanism of the instability is different from that known in the spin case.
At , = 0 one recovers the Stoner criterion UoN(EF) = 1 for the existence
of FS order, with the FS states becoming saturated only at larger U [see
Fig. l(a)]. By contrast, at, > 0 the FO states appear as a global property
of the band rather than as a Fermi surface instability [Fig. 1(b)]. For large
U (;2: 6t) a gap opens, only the lower band Eko = -tAk + U(~n - T E k )
Z
126
10 , 10
\
,,
\
8 \ 8
,,
-
......
::J
6
4
I
.................. .I\ .... _~;I'
6
4
2 2
(a) (b)
0 0
0.0 0.5 1.0 1.5 2.0 0.0 0.5 1.0 1.5 2.0
n n
Figure 1. HF instabilities towards FO states (full lines) for: (a) 0, and (b) 1 ,= ,=
[dotted line shows N- 1 (w)]. Saturated FS states occur only in (a) above the dashed line.
is occupied, and
It) 2 2] 1/2
E k = [1 + ( UTz B k , (3)
(4)
rather similar to the FO case Eqs. (3), but with the interchange A k +--7
±IBk' The reason is readily recognized from Eq. (2): for FO order, the
diagonal hopping ex cJ±cj ± that gives A k is order-preserving, while the off-
diagonal terms ex I are order-perturbing and reduce T Z • For AO order this
is reversed: the off-diagonal hopping ex cJ±cj:f that gives Bk is compatible
with the order, while the diagonal one disturbs it. The similarity between
the FO and AO states at I '::::::' 1 becomes even more transparent at large U,
where, at x = 1 - n > 0,
0.0
-0.2
-0.4
-0.6
0.0 0.2 0.4 0.6 0.8 1.0
Y
Figure 2. Energies per site at n = 0.75 and U = 00: E in the KR approach for OL
(solid line), FO+ (dashed line), AO± state (long-dashed line), and ED for: the ground
state (filled squares) and excited states (empty symbols) of a single plaquette (P).
+ ry!q+q- (e+
iXa
fl+fj - + e- ixa fLfj+)] , (7)
with niA = flAf iA , yI(i± = [x / (1 - \ ni±) )p /2, and the local constraints
(19) implemented by means of Lagrange multipliers {JLH, JLi-}' The self-
consistent solution corresponds to an OL state with \nH) = \ni-) = ~(1
x) and Gutzwiller renormalization factor q(x) = q±(x) = 2x/(1 + x) (21).
The fermion bands, c~~ = -tq(x) [Ak±rBk]' interpolate correctly between
the uncorrelated (x c:::.' 1) and Mott insulator (x = 0) limits. Since the OL
state is incoherent (it is described by the density matrix Pi = ~ Ii at every
site), it is SU(2) symmetric: random complex or random real orbitals are
equivalent, and indeed the same correlated disordered OL state is obtained
using real orbitals (22).
The ordered states can be obtained too within the present KR slave-
boson formalism by a suitable choice of the Lagrange multipliers, e.g. JL+ =
0, JL- = -00 in Eq. (7) gives the FO+ state. Such states do not experience
any band narrowing, as double occupancy is eliminated at U = 00, and
the correlation energy vanishes (23). As a result, only the cEO
= -tA k
band (ct~ = ±rtBk bands) is (are) partly filled in the FO+ (AO±) state.
Therefor~, the bands in the OL state represent formally a (renormalized)
superposition of the FO and AO bands.
It is instructive to consider the variation with r of the total energy E
of ordered and disordered states at fixed doping (Fig. 2). In the spin model
(r = 0) the FS phase has somewhat lower energy than the disordered state
close to half-filling (24). When r is increased, EFo does not change, whereas
EAO± decreases from zero ex r, but at x = 0.25 still does not surpass the
FO+ state for r = 1. However, in spite of the band narrowing ex q(x),
129
1.0 ......
,,,,
,
...
0.9 "
FO ,,,/
,,
,,
c 0.8 ,,
,,/ OL
,
0.7 ... ,,
0.6
0.0 0.2 0.4 0.6 0.8 1.0
Y
Figure 3. Phase diagram as function of r: OL versus real FOx(z) (full line) and complex
FO+ (dashed line) states at U = 00.
symmetry is absent.
Acknowledgements
References
22. The cubic invariant KR approach for real states consists in substituting bI± f--+
(bIz ± ib!x)/V2, and treating the amplitudes (biz) and (b ix ) in mean field.
23. Equivalent results are therefore obtained for the orbital ordered states by a single
slave-fermion approach.
24. At I = 0 the present method is believed to give an upper bound for the stability of
FS states;(9) they are stable below x c::: 0.33 in the cubic lattice, very close indeed to
x c::: 0.32 found for a single spin-flip in the Gutzwiller wave function [B. S. Shastry,
H. R. Krishnamurthy, and P. W. Anderson, Phys. Rev. B 41, 2375 (1990)].
25. At I = 1 and n = 0.75 the energy gain in the OL state comes close to that in the
exact ground state for a plaquette (Fig. 2).
26. R. Kilian and G. Khaliullin, Phys. Rev. B 58, Rll 841 (1998).
27. A. M. Oles and L. F. Feiner, Phys. Rev. B 65, 052414 (2002).
QUASIPARTICLES IN PHOTOEMISSION SPECTRA OF MANGANITES
Abstract. We compare the spectral functions of a hole moving in the orbital-ordered un-
doped LaMn03, obtained using a self-consistent Born approximation. The quasiparticle
(QP) dispersion and spectral weight depend critically on the type of orbital ordering. If
a hole scatters on orbital excitations, the QP dispersion on the orbiton energy scale is
modified by orbital polarization and by electron-lattice coupling. A lower QP dispersion
and quantum decoherence are obtained due to the inter-planar hole-magnon scattering
in the A-AF phase which disproves the classical concept of a ferromagnetic polaron.
133
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 133–141.
© 2003 Kluwer Academic Publishers
134
(3)
using the notation of Kanamori (10). The last term in Eq. (1) stands for
the elastic energy of the distorted lattice (E z ) (11).
The orbital ordering within the (a, b) planes is stabilized even by the
superexchange interaction alone (3, 12), and one finds alternating occupied
e g orbitals, as observed experimentally (13),
(a) (b)
(n,n) A I
j
I
I
,
I
"
(00)
-
-4 -3 -2 -1 0 1 2 3 4 -4 -3 -2 -1 0 1 2 3 4
mit mit
Figure 1. The hole spectral functions A(k, w) for the orbital ordering given by ¢ = 0
in the absence of the electron-phonon interaction (.\ = 0) for: (a) 6 = 0, and (b) 6 = t,
as obtained along the (0,0) - ('if, 'if) direction in the 2D Brillouin zone.
HJT = -2A [(6x - 6z ) sin 2¢ - 2v3u cos 2¢] (LTiz - LTt), (5)
iEA iEB
with the pseudospin operators rJiz referring to the ground state, Eq. (4).
In the linear orbital-wave (LOW) theory (12) the effective Hamiltonian
represents a coupled hole-orbiton problem in momentum space,
with N being the number of lattice sites and two hole-orbiton vertices,
1.0 ~ 1.0
I:- 0.8 .A> oj(0.8
co O6
. (/v
/V<> S 0.6
~ 0.4 ~~r 0.4
oj( -
, ,
i:>.i:>.AA1:>.A~A+Ai:>.8.AAAAL;.i:>.+1:>.AAAL::..AAA8
-3.0
(0,0) (n,n) (n,O) (0,0) (0,0) (n,n) (n,O) (0,0)
Figure 3, Dispersion of the QP band for: (a) ¢ = 0, and (b) ¢ = -1f/12, with 6 =)., = 0
(0),6 = 0, )., = lOt (e), 6 = t, )., = 0 (D), and 6 = t, )., = lOt (0),
KL L QT'~+2~L L Pi:1;
dyn _
H JT -
i 1;=1,2,3 i 1;=1,2,3
+ x
2V6,\ LTi (sin 2¢ Qi,2 - cos 2¢ Qi,3) , (10)
i
where M stands for the mass of an oxygen ion, with the JT phonon modes:
(11)
where Uiv (with v = x, y, z) is the oxygen vibration along the respective
Mn-O bond and Pi,1; is the conjugate momentum vector corresponding to
the lattice distortion Qi,I;' The breathing mode, Qi,l = (Uix + Uiy + Uiz), Js
does not couple to the orbital excitations. In the LOW order the total
effective Hamiltonian [derived from Eqs. (1) and (10) at ~ = 0] represents
a coupled hole-orbiton-phonon problem in momentum space,
(a) (b)
I I l~JHlulvV' !l! UI
-2.0 -1.0 0.0 1.0 2.0 3.0 -2.0 -1.0 0.0 1.0 2.0 3.0
mit mit
Figure 4. The local hole spectral functions A(w) obtained for the Ix 2 - z2)/l y 2 - Z2)
orbital ordering (¢ = 7[/12) of LaMn03 with: (a) Alt = 0 and (b) Alt = 8.
+ "(
~ (i;)(3t
Mk,q q,~ + Nk,q
(i;)(3t
q+Q,i;
)]
+ H.c. } , (12)
i;=1,2
with the energy of the phonon mode Wo = J2K/M and the hole-phonon
vertex M o = J3/NA/(2KM)1/4. O&i;)(¢) and M~:~ are the energies of
mixed electron-phonon excitations and the respective vertices, which de-
pend on A and ¢ (for more details see Ref. (17)). Moreover, Bq,/L are phonon
operators representing modes which do not couple to orbitons, while (3q,i;
represent mixed orbiton-phonon excitations (18).
In the presence of both orbital ordering and lattice distortions, a hole
can scatter off both orbitons and phonons, already renormalized by each
other. The simplest situation occurs for the Ix 2 - z2) /l y 2 - z2) orbital
ordering with Gk (¢) = O. In this case, similar to a hole in a quantum anti-
ferromagnet (1), a hole can only propagate making mixed orbital-phonon
excitations on its path. Such mixed excitations have already been observed
in FM LaMn03H (19). As the orbitons and phonons having no momentum
dependence in this case, one finds k-independent spectra and the infinite
hole effective mass. In the absence the JT effect, the model reduces to the t-
jZ model (1) with ladder-like spectral functions shown in Fig. 4(a). When
the orbiton-lattice coupling ex: A increases, the QP peak moves to lower
energies and the spectrum looses gradually its coherence; for large values
of A c:::: 8t it separates into a number of subspectra consisting of a series
of lattice vibrations [Fig. 4(b)]. Each hopping of the hole (leading to an
excitation of an orbiton) is accompanied by a few vibrational quanta. This
spectral function well exemplifies the structure of the electron dressing after
a sudden photoemission process. For other orbital orderings the vibrational
peaks merge into a single broad maximum next to the QP peak.
139
(a)
/~
_ (n,O,O)
(b) (n,0,nI2)
(c)
("::1
~.~
~
8'
d
~~
<t:: -~
-~-
~-~
~-
~----
(0,0,0)
=\~-- ~~o,O,nI2) (0,0,0)
-
Figure 5, The hole spectral functions obtained in the hole-magnon 3D model for the
A-AF phase with cP = 1f/24 along high symmetry lines in the 3D Brillouin zone (7),
In this section we consider the limit of rigid orbital ordering (>.. » t) for
~ = O. Then the model given by Eq. (1) leads to free propagation of an eg
hole in the (a, b) plane. However, the A-AF order leads to hole scattering
on spin excitations when it hops out off FM planes. Assuming that the
orbital ordering repeats itself in (a, b) planes, the hopping part of our three-
dimensional (3D) t-J model, H = H t + H.], reads (7),
(13)
where cIa = cIa(l - ni(j). Here (ij)ab and (ij)c represent the nearest-
neighbors in the (a, b) plane and along c direction, and tt = ~t(l + sin 2¢).
The interactions between spins S = 2 of Mn3+ ions are given by:
orbital ordering, where ttlt~b = 0.5 the hole hardly feels the AF ordering
in the c direction and moves almost coherently in the (a, b) plane (7).
Summarizing, we have studied a single hole propagation in undoped
LaMn03 with orbital and magnetic ordering. The hole scatters strongly on
low-energy excitations: phonons, mixed orbiton-phonon, or magnons. The
orbital polarization around a carrier leads to its additional confinement.
Our results predict a large redistribution of spectral weight with respect to
the bands found in local density approximation (LDA) or in LDA+U, and
can have important implications on the angular resolved photoemission
spectroscopy (ARPES). In all cases we found low-energy QP's, but the
minimum of the QP band falls at k = ('if, 'if, kz ) when the hole-orbiton
scattering dominates, while it is at k = (0,0, 'if /2) for the hole-magnon
scattering. Which of the considered processes dominates and which disper-
sion is realized might be answered only by the ARPES experiments. In
fact, the orbitons are gapped, and were found recently at energies larger
than rv 150 meV (20). The magnons are gapless, and have much lower
energies rv 30 meV, so the hole scattering on them might dominate, but
only if the orbital ordering is close to Ix 2 - z2) /l y 2 - z2) (¢ = 'if /12) rather
than to 13x 2 - 7'2) /13 y 2 - 7'2) (¢ = -'if /12). Thus, the results of future
ARPES experiments could also help to establish which type of orbital
ordering is realized in LaMn03, the issue difficult to resolve by resonant
x-ray scattering (13). We also argue that the spectral features measured in
the highly doped La1.2Sr1.sMn207 (21) should change substantially when
magnetic order changes to the A-AF phase under doping.
This work was supported by the Polish State Committee of Scientific
Research (KBN), Project No.5 P03B 055 20.
References
16. W.-G. Yin, H.-Q. Lin, and C.-D. Gong, Phys. Rev. Lett. 87, 047204 (2001).
17. J. Bala, A. M. Oles, and G. A. Sawatzky, Phys. Rev. B 65,184414 (2002).
18. J. van den Brink, Phys. Rev. Lett. 87, 217202 (2001).
19. Y. G. Pashkevich et al., unpublished (2002).
20. E. Saitoh et al., Nature 410, 180 (2001).
21. D. S. Dessau et al., Phys. Rev. Lett. 81, 192 (1998).
METALS NEAR A ZERO-TEMPERATURE MAGNETIC INSTABILITY
H.V.LOHNEYSEN
1 Physikalisches Institut, Universitat Karlsruhe, D-76128 K arls-
ruhe, Germany and
Forschungszentrum Karlsruhe, Institut fur Festkorperphysik,
D-76021 Karlsruhe, Germany
1. Introduction
Metals with strong electronic correlations are often close to a magnetic in-
stability. In a number of these systems, the transition temperature between
a paramagnet and a magnetically ordered phase can be tuned to absolute
zero by some externally controlled parameter such as chemical composi-
tion, pressure, magnetic field, or charge carrier concentration. This offers
the possibility to induce a T = 0 magnetic-nonmagnetic quantum phase
transition (QPT). In the vicinity of this transition, non-Fermi-liquid (NFL)
behavior (1) may occur in thermodynamic and transport properties: the
linear specific-heat coefficient I = C /T acquires an unusual temperature
dependence, often I rv -In(T/To), and the T-dependent part of the elec-
trical resistivity 6.p = P- Po where Po is the residual resistivity, often varies
as 6.p rv T m with Tn < 2, in contrast to the Fermi-liquid (FL) predictions
I = canst and Tn = 2.
143
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 143–153.
© 2003 Kluwer Academic Publishers
144
Pure CeCu6 shows no long-range magnetic order down to very low T due
to the quenching of Ce 4f magnetic moments by the Kondo effect (19, 20),
with some evidence for magnetic ordering (either electronic or nuclear)
occurring at a few mK (21, 22, 23). With r = 1.6 J jmoleK 2 it is one of the
"heaviest" HFS. CeCu6 exhibits a pronounced magnetic anisotropy with
the magnetization ratios along the three axes Me : l'vla : Mb ;:::j 10 : 2 : 1
at low T (20). We use the orthorhombic notation, in spite of the (small)
monoclinic distortion.
Already at relatively high T, i.e. around 1 K, does CeCu6 exhibit in-
tersite antiferromagnetic fluctuations as observed with inelastic neutron
scattering (INS) by peaks in the dynamic structure factor S (q, w) for energy
transfer nw = 0.3 meV at Q = (100) and (0 1±0.15 0) (24,25). The rather
large widths of these peaks correspond to correlation lengths extending
roughly only to the nearest Ce neighbors. Recently, additional features in
the a*c* plane at an energy transfer of 0.1 meV were found (26).
145
Q
2.0 CeCu6 Au -/ '--
-x x / _\
! ~
g 1.5
z
I- 1.0
x p (kbar)
• 0.1 0
,,0.2 0
• • 0.3 0
• o 0.1 6.0
.. 0.2 4.1
o 0.3 8.2
T(K)
Upon alloying with Au, the CeCu6 lattice expands (27), thus weak-
ening the hybridization between conduction electrons and Ce 4f electrons.
Hence the conduction-electron-4f-electron exchange constant J decreases,
leading to a stabilization of localized magnetic moments which can now
interact via the RKKY interaction, with ensuing antiferromagnetic order
(28). Fig. 1 shows the Neel temperature TN of CeCu6-xAux vs x. The
magnetic structure of CeCu6-xAux (0.15 :S x :S 1) as determined with
elastic neutron scattering (7, 29, 30) shows magnetic Bragg peaks in the
a*c* plane (see Fig. 3 below).
The onset of magnetic order in CeCu6-xAux is observed as a kink in
CIT, ef. Fig. 2. For the critical concentration Xc = 0.1 the linear specific-
146
(1)
with 0: = 0.75 (9). The anomalous non-Lorentzian response ( 0: # 1) does
not change for other q away from the critical region (11). For all q the
susceptibility can be expressed as
(2)
In particular, the T dependence of the static uniform susceptibility
X(q = 0, E = 0) for x = 0.1 yields the exponent 0: >:::; 0.8 to a high degree
of accuracy. The simple form of Eq.(2) separates static spatial correlations
from the specific temporal correlations, the latter being independent of q.
These local fluctuations at the quantum critical point have received con-
siderable theoretical attention, although a detailed model is not available
147
I J, d
a
II ~ IJ II
•
.--...
I r ~Ilj
::>
.
"'i::
ffliIII
'-'"
u 0 ••-m---
0 x=O.1 t
C::.Ax=O.2
••
-0.5
x=O.3
x=O.5 CeCu 6 _xAu x
-1
• x=1.0
0 0.5 1 1.5 2
a' (rlu)
Figure 3. Position of the dynamic correlations (x = 0.1, = 0.1 meV, T < 100 mK)
nw
and magnetic Bragg peaks (0.15 x s: s:
1.0) in the a*c* plane in CeCu6- x Au x . Closed
symbols for x = 0.2 represent short-range order peaks. The vertical and horizontal bars
indicate the Lorentzian linewidths for x = 0.1. The four shaded rods are related by the
orthorhombic symmetry (we ignore the small monoclinic distortion). The inset shows a
schematic projection of the CeCu6- x Au x structure onto the ac plane where only the Ce
atoms are shown. The bars in reciprocal space correspond to planes in real space spanned
by b and the lines in the inset.
yet (32, 33). The evolution of the ordered moment with increasing x > Xc
(7), may provide a valuable input to test the different models.
The onset of magnetic order in CeCu6-xAux is attributed to a weak-
ening of J because of the increase of the molar volume upon alloying
with Au as mentioned above. This is confirmed by the observation that
TN of CeCu6-xAux decreases roughly linearly under hydrostatic pressure P
(6,34). At the critical pressure Pc where TN ----t 0, i.e. Pc ~ 4 kbar for x = 0.2
and ~ 8kbar for x = 0.3, CIT exhibits NFL behavior, i.e., CIT rv -lnT,
with the same coefficients a and To as for x = 0.1 at p = 0 (see Fig. 2).
Consequently, p =6 kbar drives the ambient-pressure NFL alloy x = 0.1
into the FL regime.
An inducement of NFL behavior by a magnetic field in the related sys-
tem CeCu6-xAgx was reported previously (35, 36). The low-T properties at
the critical field are compatible with the conventional 3D antiferromagnetic
spin-fluctuation scenario (4). A detailed comparison of the specific heat and
electrical resistivity of CeCu5.sAuo.2 revealed distinctive differences depend-
148
• TN this work
15 - 0 o TN [3]
!1 TN [5]
v TN [12]
g10
~ • Tf this work
z
f-
5 \ 4n
o Tf
D Tf
[5]
[3]
0-
UCu 5 _xPd x IV
I
,~~----
0
0 0.5 1 1.5 2
x
x=1
A = 0.66 0.1
x=1
oCO
000
03
0.03
0.1 ]
0.03 ~
0.1
03 0.Q3
0.1 0.1 5
T(K) T(K)
Figure 5. (a) In CIT vs. In T for UCu5- x Pd x with x = 1,1.25,1.5; solid lines are fits of
the form CIT = T-l+ A . (b) In Xac vs. In T for x = 1, 1.25, 1.5; solid lines are fits of the
form X = T- 1 + A .
imate data of Xne above the spin-glass freezing temperature are consistent
with this prediction (15).
The compound ZrZn2 is ferromagnetic (42) despite being made from non-
magnetic and even superconducting elements. The magnetic properties are
believed to derive primarily from the Zr 4d orbitals that have a significant
direct overlap (43). Ferromagnetism develops below the Curie temperature
TFl'vI = 28.5 K with an ordered moment jLs = 0.17jLB per formula unit.
ZrZn2 has a large electronic heat capacity at low temperatures, CiT ;:::::
47 mJ I molK 2, signaling the presence of many low-energy magnetic excita-
tions in addition to spin waves (44). The low TFl'vI and small ordered moment
make ZrZn2 unique among stoichiometric ferromagnetic metals and indicate
that the compound is close to a ferromagnetic QPT. This proximity has
led to numerous proposals that ZrZn2 might be a superconductor (17, 46).
For the highest-quality sample with a low- T residual resistivity of Po
= 0.62 jLOcm, we observe a rapid drop in the electrical resistivity p(T)
below Tsc = 0.29 K at ambient pressure (inset of Fig. 6), suggesting an
incomplete transition to a zero-resistance state. Application of a field of
0.2 T suppresses the drop as would be expected for a superconducting
transition. In addition, a clear diamagnetic signal in the ac susceptibility
associated with superconducting screening is observed below Tsc. For the
lowest excitation amplitudes, Xac approaches -0.65 as T ----t 0, comparable
with the ideal value of -1 (18). This diamagnetic signal rides on top of a
large ferromagnetic background.
It has been known for a long time that ferromagnetism in ZrZn2 is
rapidly suppressed under pressure P (47). Fig. 6 summarizes the effect of
pressure on TFl'vI and Tsc. p suppresses both ferromagnetism and supercon-
ductivity above a critical pressure of Pc = 21 kbar. In view of its sensitivity
to the quality of the sample, the superconductivity in ZrZn2 is likely to be
unconventional. The fact that superconductivity in ZrZn2 only occurs in
the presence of ferromagnetism where Tsc is relative insensitive against
pressure, and is hence promoted by the ferromagnetic state, may arise
naturally in scenarios where the Cooper pairs are in a parallel-spin (triplet)
state, which is already favored in the ferromagnetic state. Such behavior
could well be universal for itinerant ferromagnets in the limit of small Curie
temperature and long electron mean free path. Further work has to establish
the microscopic relation between ferromagnetism and superconductivity
and to investigate if and how the ferromagnet is affected by the onset of
superconductivity.
151
30
_0.6
E
()
22 kbar
1.6 kbar
~0.5 mbient
20
SZ
~
I-
10
ol
D
n
u
TFM
-I'I.. .~
10*Tsc
D
a. 0.4
~O
<x~'"C,~~>------()
, 0.2 0.4
T (K)
i'- .J
0.6
a 5 10 15 20 25
P (kbar)
Acknowledgements
References
47. T. F. Smith, J. A. Mydosh, E. P. Wohlfahrt, Phys. Rev. Lett., 27, 1732, (1971).
48. E. B. Sonin, 1. FeIner, Phys. Rev. B, 57, R14000, (1998).
49. P. Fulde, R. A. Ferrell, Phys. Rev. A, 135, 550, (1964).
50. A.1. Larkin, Y. N. Ovchinnikov, Sov. Phys. JETP, 20, 762, (1975).
HEAVILY DOPED SEMICONDUCTORS: MAGNETIC MOMENTS,
ELECTRON-ELECTRON INTERACTIONS AND THE METAL-
INSULATOR TRANSITION
H.V.LOHNEYSEN
Physikalisches Institut, Universitiit Karlsruhe, D-76128 Karls-
ruhe, Germany and
Forschungszentrum Karlsruhe, Institut fur Festkorperphysik,
D-76021 Karlsruhe, Germany
1. Introduction
Heavily doped semiconductors have become prototype systems for the study
of a metal-insulator transition (MIT) driven by the combined effects of
disorder (Anderson transition) and electron-electron interactions (Mott-
Hubbard transition). In a strict sense, the MIT viewed as a transition from
extended to localized states at the Fermi level, occurs only at T = 0 because
at finite temperature T thermally activated carrier transport may occur.
Hence a metal is defined by a finite dc conductivity O"(T) for T ---+ 0, while
155
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 155–167.
© 2003 Kluwer Academic Publishers
156
It has been known for a long time that a small fraction of localized moments
derived from the spin-degenerate groundstate of the hydrogen-like donor
or acceptor wave function on the insulating side, survive on the metal-
lic side of the MIT. These localized magnetic moments have been mostly
detected with magnetization (2, 3), magnetic resonance (4, 5) and specific-
heat measurements (6, 7). The dependence of the concentration of localized
moments on the P concentration N has been mapped out systematically
from specific-heat measurements for Si:P (8, 9) and more recently also for
Si:(P,B) (10).
The data for Si:P are shown in figure 1 with the overall behavior of
Si:(P,B) being quite similar. The concentration of local moments has been
determined (i) from zero-field excess specific heat 6.C = C -IT - {3T 3 , i.e.
conduction electron and phonon contributions subtracted from the meas-
ured specific heat C, via the entropy S = J(6.C / T) dT = N S ln2, and (ii)
from the Schottky anomaly observed in applied magnetic fields B = 1.5 and
5.7 T, yielding N S eh. In both cases, S = 1/2 is assumed for the spin. N sand
N Seh vary smoothly through the critical concentration N e = 3.52 ·10 18 cm- 3
and localized moments survive in the metallic state due to the disorder and
on-site Coulomb repulsion as will be discussed in detail below.
The random distribution of dopant atoms, which has been demonstrated
convincingly by scanning tunneling microscopy (12), leads to a wide dis-
tribution of nn distances and, hence, nn exchange couplings. This wide
157
, ,
..
<:T
, ,, 00 .0
,,
00
, ,, ~ 1>.1
, ,, M
,, I>.
,,
\
en Q I>.
E , ,,
~
£10 17
,, •
-i;
en
Z
in
Z
• B=0
I>. B= 1.5T
oB=5.7T I>.
10 '6 L..----L---L....................u..LL_....................................LU
0.1 10
(2)
In order to test the validity of the Bhatt-Lee model one can calculate C loc
from Xloc with the above equations and compare it to tlC as measured
directly (8, 9). The overall good agreement found on the insulating side (2,
13) indicates the validity of the Bhatt-Lee model. However, with increasing
N 2 3.3 . 10 18 cm -3, i.e. on the metallic side of the MIT, the agreement
quickly deteriorates (13), indicating that the Bhatt-Lee model is unsuited
for the metallic state.
The origin of localized moments in the metallic state has been investig-
ated in a number of approaches (14, 15, 16, 17,9), most of them starting
with an Anderson-Hubbard model which is the simplest model Hamilto-
nian featuring the essential elements of disorder, through random hopping
integrals tij, and on-site Coulomb interaction U:
(3)
158
S2" -0.5
~
(f) -1
• B=0
-1.5 oB=1.5T
• B = 3T
6 B = 6T
-2 '-_...L..-_-'-_--'-_---'-_----'
o 0.2 0.4 0.6 0.8
T(K)
Cia (ct) is the annihilation (creation) operator for an electron with SpIll
projection (]" in the Is(Ad groundstate of the donor atom at site i, and
nia = ctcia' It is usually assumed that the five excited states generated by
valley-orbit splitting are sufficiently far away (experimentally r-v 10 meV,
see below) that a single-band model is applicable. The host semiconductor
properties enter only through the fact that the positions i are random sites
of the Si lattice and that the effective mass and interaction are renormalized.
The various approaches differ in the type of mean-field model employed.
In the approach taken by Langenfeld and Wolfle (17) an isolated moment
("impurity") forming in an effective homogeneous medium is considered.
The calculated concentration N M of local moments is in good agreement
with the experimentally determined values Ns and NSch on the metallic
side (Fig. 1).
The coupling of the localized moments to the itinerant electrons by
an effective exchange interaction JeI I gives rise to the Kondo effect. Be-
cause of disorder, both the density of states at the Fermi level N(EF)
and the effective exchange interaction Jeff ~ [2/U where t is the average
hopping amplitude from the local-moment site to the neighboring sites,
will fluctuate. This leads to a wide distribution P(TK ) of Kondo temper-
atures T K r-v exp( -(I/N(EF )Jeff). An approximate power-law behavior
P(TK ) r-v Ti/XK is found, with aK ~ 0.9, which may account for flC in the
metallic state of Si:P (9).
" '--'-'-~~~,':-o~-L..J40· 1 10 40
T{K1 T(K)
Figure 3. Negative thermoelectric power times carrier concentration -SN vs. tem-
perature T on a log-log plot for insulating samples of (a) Si:P and (b) Si:(P,B) (24)
Figure 4. Electrical resistivity p plotted vs. inverse temperature for (a) Si:P and (b)
Si:(P,B). Straight lines in (a) indicate fits to obtain the activation energy E 2 . Inset in (b)
shows the Si:(P,B) data plotted as log p vs T- 1 / 2 . Straight lines indicate Efros-Shklovskii
hopping (24).
...............
......
15 .............. (a)
......
" 'e,
5 \
\
O'--_ _----0------01>..
...J.-_ _........_ _----L----"....
150
",,,
I I I
(b)
- -
\
~
\
- \ -
\
o I 0 I
o .~.
o 3
Figure 5. (a) Temperature Ts=o of the thermoelectric power zero and (b) activation en-
ergy E 2 vs. carrier concentration N for uncompensated Si:P (closed circles) and Si:(P,B)
(open circles) (24).
N«N o
.. -
_ _Wll
N~Nc
course, the Hubbard features are absent in Si:(P,B) as they should because
compensated semiconductors are away from half-filling. It is interesting to
note that the density of localized magnetic moments in Si:P is not maximal
at N c but peaks rather precisely at No where the two Hubbard bands in
our scenario are suggested to merge, see Fig. 6.
T (K)
0.01 0.1 0.2 0.5
16 N = 3.2110 18 cm- 3
12
E
C)
----
========
g 8
b
ill
4
oo~-=~=
0.2
.....
0.4
"""==::~=:::r::::'-------J.
0.6 0.8
T1/2 (K1/2)
T (K)
0 0.01 0.05 0.1 0.2
6
4
E
u
S- 2
o
0
0
20
10 0
E
u
glO -1 0 1
0
Figure 8. (a) Low-temperature data of u of Fig. 1 for stress in the immediate vicinity of
the metal-insulator transition plotted against T 1 / a. Dashed line indicates the conductivity
at the critical stress (see text). (b) Extrapolated conductivity u (0) for T ----> 0 versus
uniaxial stress S for two P concentrations N = 3.21 and 3.43.10 18 cm -:3 (open and closed
circles, respectively). The inset shows earlier u(O) versus S - Se data (triangles) together
with the reported I-" = 0.5 (dashed line) from (35) in comparison to our data for sample
1 (circles) (38).
Closer inspection shows that the data near the MIT are actually better
described by a T 1/ 3 dependence for low T, see Fig. 8a. 0"(0) obtained from
the T 1/ 3 extrapolation to T = 0 is shown in Fig. 8b, together with data for
a sample closer to the critical concentration, yielding Be = 1.75 kbar and
1.54 kbar for 3.21 and 3.43 . 1Q18cm -3, respectively. Note that the critical
stress Be is quite well defined, as 0"(0) breaks away roughly linearly from zero
within less than 0.1 kbar, i.e. f-L ;::::; 1. This behavior contrasts with the earlier
stress-tuning data (35) reproduced in the inset of Fig. 8b, where appreciable
rounding close to N e is visible. However, those 0"(0) data between 4 and
16 n-1cm- 1 are compatible with a linear dependence on uniaxial stress.
In order to analyze the scaling behavior of 0" at finite temperatures using
the data of the sample with N = 3.21 . 1Q18cm -3, we employ the scaling
165
SI:P
N ,,3.2110 18 cm-3
metallic
10-2
10-3
10"4 insulating
Figure 9. Scaling plot of u/u, vs. IS - s, I /ScTY for Si:P with N = 3.21·1Q18 cm -3
at different uniaxial pressures S, with Se = 1.75kbar and y = 0.34 (38).
relation (39)
(4)
where CJe(T) CJ(t e,T) is the conductivity at the critical value t e of the
parameter t (in our case S) driving the MIT. If the leading term to CJe(T) is
proportional to T X , one obtains x = Itlvz and y = I/vz from a scaling plot.
Fig. 7 shows that CJ for S close to Se does not exhibit a simple power-law
T dependence over the whole T range investigated. We therefore describe
CJe(T) by the function CJe(T) = aT X (1 + dT W ) with a = 6.01 O-lcm-l, x
= 0.34, d = -0.202, w = 0.863, and T is expressed in K. Here the dT w
term presents a correction to the critical dynamics. All CJ(S, T) curves with
1.00 kbar < S < 2.34 kbar where dCJ I dT ;::: 0 are then used for the scaling
analysis.
5. Acknowledgments
The work reviewed here grew out of a very fruitful collaboration with
students, post-docs and colleagues. Their contributions can be identified
from the references cited. In particular, I would like to acknowledge M.
Lakner, X. Liu, C. Pfleiderer, H. G. Schlager, C. Siirgers, T. Trappmann
and S. Waffenschmidt. I am grateful to P. Wolfle for numerous enlightening
discussions on the theoretical aspects of the metal-insulator transition.
References
Abstract. The large number of low-lying states of d- and i-shells supports a variety of
order parameters. The effective dimensionality of the local Hilbert space depends on the
strength, and kind, of intersite interactions. This gives rise to complicated phase dia-
grams, and an enhanced role of frustration and fluctuation effects. The general principles
are illustrated on the example of the effect of a magnetic field on quadrupolar phase
transitions in the Pr-filled skutterudite PrFe 4P 12 .
Key words: Quadrupolar Order, Multipolar Order, PrFe4P12, Skutterudites, Rare Earth
Compounds, Pr Compounds, Mean Field Theory, Tricritical Point, Metamagnetic Trans-
ition, Quadrupolar Phase Transition in External Magnetic Field
1. Introduction
Transition metal and rare earth compounds show a rich variety of collective
behavior: various kinds of ordered phases as well as strongly fluctuating
states (spin and orbital liquids). The basic reason is that d- and f-shells
have a relatively large number of low-energy states. Crystal field splitting
usually reduces this number (the dimensionality D of the local Hilbert
space) considerably below the free ion value, but complicated physics can
arise even from D = 3 or 4.
Let us briefly consider some examples. D = 3 is, in one interpretation,
the case of S = 1 spin models which turned out to have unanticipated
phases like the spin nematics (1). A different realization is offered by in-
teracting f-electron models based on the low-lying quasi-triplet of Pr ions
in PrBa2Cu307-6 where the nature of Pr ordering is still an open issue
(2, 3). A literal realization of D = 4 is offered by the r s ground state
of Ce ions in CeB 6 which has a rich phase diagram (4). Alternatively, we
169
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 169–177.
© 2003 Kluwer Academic Publishers
170
1 The symmetry group is really not Oh, but the tetrahedral T h . We nevertheless use
the cubic classification, which is an approximation at zero field, but when H # 0, the
symmetry will be in any case substantially lowered.
172
the r l-r 4 scheme is also capable to account for most of the observed static
properties of PrFe4P12.
We now discuss the consequences of assuming a r l-r 4 level scheme.
Since the singlet ground state
(1)
does not carry any kind of moment, the ordered quadrupolar moment has to
be induced by intersite interactions, assuming that the local Hilbert space
contains also the triplet
(5)
Evidently, the system could support either dipolar (r 4), or either of two
kinds of quadrupolar (r 3 or r 5 ) order 2. The quadrupolar order parameters
are the same that appear in the decomposition of a purely r 4 system
(6)
i.e., they are not sustained by inter-level matrix elements 3 . - It may be
of some interest to mention that the r l-r 4 scheme does not offer the
possibility of octupolar order (but the r 3-r 4 scheme would).
0 -2J5/3h x
M(A h ~) = -2J5/3h x ~ + 7A (7)
, x, 0 0
(
o 0
The field couples the singlet ground state Ir 1; to the O§-moment bearing
excited state Irt;. Therefore in the presence of a magnetic field, uniform
quadrupolar moment is no longer" spontaneous". If at H = 0, we had to do
with a transition to a ferroquadrupolar state, it would be smeared out in
H i- 0, and we no longer had a phase boundary to speak about 5 . However,
for antiferroquadrupolar coupling, the appearance of the staggered quad-
rupolar moment is still symmetry breaking, and therefore a sharp phase
transition remains possible also in an external magnetic field. Therefore, if
we had no other evidence than that PrFe4P12 has sharp phase transitions
in external magnetic field, and we adopted the r l-r 4 scheme, we would
have to conclude that the ordered state could not be ferroquadrupolar, but
only antiferroquadrupolar.
Doing the mean field theory (3) for AFQ ordering involves diagonalizing
matrices like (7), and we do not give the details here.
Fig. 1 illustrates that an AFQ transition, though of non-magnetic nature,
may give a susceptibility which looks very much like what you expect from
an antiferromagnet. The cubic (001) and (100) directions are equivalent in
the para phase but the appearance of O§-type AFQ order makes the x-field
susceptibility appear as "transverse", while the z-field susceptibility looks
"longitudinal". Of course, instead of O§ = f; - J;
we might have chosen
J; - fj; or fj; - f;, so in a crystal one would expect an equal mixture of
the corresponding AFQ domains, and the susceptibility suitably averaged.
The experiments may be taken to correspond to this.
Fig. 2 (left) gives the phase diagram. The crystal field splitting ~ and
the quadrupolar coupling A were chosen so as to get at least a rough
numerical agreement with the experimental phase diagram (11). There
is still some freedom in the parameters, but we found that a rather low
~ >:::; 4 - 6K has to be chosen (with z = 8, A >:::; 0.08k B ), if we want to get
both Ttr(H = 0) and the tricritical point right. These estimates are likely
5 Purely in symmetry terms: applying a field in one of the cubic (100) directions, the
symmetry would be lowered to C 4 h, and the decomposition of r 4 of Oh in terms of the
irreps of C 4h would contain the identity which also comes from r 1 ground state.
175
H//<100>
0.61- ~
04
X
0.2
H//<OOI>
H//<100>
0.5
CIT
2 3 4 5 6 7 o 6
TIKI
Figure 2. Left: The boundary of the antiferroquadrupolar phase in the H-T plane.
The curve is drawn in black for first-order transitions, and in grey for continuous phase
transitions. Arrow indicates the tricritical point. Right: The T-dependence of the tem-
perature coefficient of the specific heat for H = 0, 2.5, and 4Tesla (in order of decreasing
transition temperatures).
T~2K T~4.2 K
T~7K
11//<100> IIII<J(X»
11//<100>
II II II
Acknowledgements
References
1. C.D. Batista, G. Ortiz, and J.E. Gubernatis, Phys. Rev. B65, 180402(R), (2002),
and references therein.
2. A.T. Boothroyd, J. Alloys and Compounds 303-304, 489, (2000).
3. A. Kiss and P. Fazekas, cond-mat/0212345, and to be published.
4. R. Shiina, H. Shiba, and P. Thalmeier, J. Phys. Soc. Japan 66, 1741, (1997).
5. A. Takahashi, H. Shiba, J. Phys. Soc. Japan 69, 3328, (2000).
6. G. Mihaly et aI., Phys. Rev. B61, R7381, (2000); G. Mihaly et aI., at this conference.
7. K. Penc et aI., to be published.
8. P. Fazekas, Found. Phys. 30, 1999, (2000).
9. P.W. Anderson, Mat. Res. Bull. 8, 173, (1973); P. Fazekas and P.W. Anderson, Phil.
Mag. 30, 423, (1974).
10. K. Penc, M. Mambrini, P. Fazekas, and F. Mila, cond-mat/0212211.
11. Y. Aoki et aI., Phys. Rev. B 65 064446, (2002); J. Phys. Chern. Sol. 63, 1201, (2002).
12. S.H. Curnoe, K. Ueda, H. Harima, and K. Takegahara, J. Phys. Chern. Sol. 63,
1207, (2002); S.H. Curnoe, H. Harima, K. Takegahara, and K. Ueda ,Physica B
312-313, 837, (2002).
177
13. C. Sekine, T. Uchiumi, 1. Shirotani, and T. Yagi, Phys. Rev. Lett. 79, 3218, (1997).
14. E.D. Bauer et al., Phys. Rev. B 65, 100506(R), (2002); K. Miyake et al.,
unpublished; D.E. MacLaughlin et al., Phys. Rev. Lett. 89, 157001, (2002).
15. K. Iwasa et al., Physica B 312-313, 8:34, (2002).
16. Y. Aoki et al., J. Phys. Soc. Japan 71, 2098, (2002).
LOW DIMENSIONAL SPIN SYSTEMS IN HIGH MAGNETIC
FIELDS: SPIN-PHONON INTERACTION
Abstract. We discuss ESR- and ultrasonic experiments for quasi- one and twodimensional
spin-systems. We found several effects for (VOhP207 and SrCu2(B03h. With ESR we
determined the magnetic excitation spectrum and with ultrasonic measurements we were
able to follow these excitations to lowest frequencies. The spin-phonon interaction is of
the exchange-striction type and with ultrasonic experiments we are able to determine the
strain dependence of the magnetic coupling constant.
Key words: Low Dimensional Spin Systems, Pulse Field Experiments, Spin Phonon
Interaction, Ultrasonic Experiments, ESR Experiments, Magnetization Plateaus
1. Introduction
179
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 179–187.
© 2003 Kluwer Academic Publishers
180
systems. There are several reviews dealing with this topic (2, 3, 4). Here
we analyze in some detail the interaction between magnetic excitations and
sound waves for the different spin systems investigated.
We discuss one- and two-dimensional dimerized spin-l/2 compounds,
such as (VOhP207 and SrCu2(B03h. Here V4+ and Cu2+ have spin-
1/2 and form either chains or two dimensional dimer arrays. The lowest
energy excitations are usually of the singlet-triplet type or modifications
of it. These can be determined either by ESR excitations from the ground
state or within the excited triplet. For B = 0 inelastic neutron scattering
give the excitation spectrum for the whole k-space.
Usually the spin-phonon interaction for transition metal ions is of the single
ion magnetoelastic type. The strain component couples to the electric quad-
rupole moment of the magnetic ion. However for spin 1/2 systems, involving
e.g. Cu2+ or V 4 + and for lower point symmetry than cubic (where the
orbital moment is quenched) this magnetoelastic interaction is negligible.
In this case two ion interactions of the form of exchange striction becomes
important. A soundwave or more generally a phonon can modulate the
exchange interaction thus creating a spin-phonon coupling as given by
(2)
One can describe the effective elastic constant in such systems by consid-
ering the strain susceptibility Xd for an isolated dimer and introducing the
dimer-dimer interaction Kin RPA- approximation. One gets in this way
for a symmetry elastic constant:
(3)
Here N is the number of dimers per unit volume and G is a strain-
dimer coupling constant. For a simple singlet-triplet dimer we have for the
free energy F = -kTlnZ with Z = 1 + 3exp( -fl/kT) and fl the singlet-
triplet splitting. For Xd we get with G = dfl/dcxx: Xd = G- 2d 2F/dc 2 =
3C b./kT /[kTZ 2 ].
In the next section we will give some examples where such effects have
been observed. Note that Xd is the strain susceptibility analogue of the
magnetic susceptibility. Analogously to the case of the temperature de-
pendence we can calculate the field dependence of the symmetry elastic
constant. Instead of the free energy used above we have to take the field
splitting of the excited triplet state into account: F = -kTlnZ with Z =
1 +exp( (-fl + gP,BB)/kT) + exp( -fl/kT) + exp(( -fl- gP,B)/kT). Since in
the systems investigated fl ~ 30-60K and T < 2K, we have only to consider
the lowest triplet level which cuts the singlet branch for gP,BB c = fl. With
this simplification we get for the single dimer strain susceptibility
and the elastic constant c[' is given again by eq.(3) with Xd taken from
eq.(4). The coupling constant G is given again by G = dfl/dc.
There is another way to calculate the field dependent elastic constant us-
ing the Landau theory of phase transitions (6). For a longitudinal wave with
strain cxx one can write for the spin-phonon interaction F me = G' cxxm;
with G' rv dJ/ dcxx. One gets with Cn = d2 F / dc;x with F = FL + Fme the
expression (here F L is Landau-order parameter free energy)
(5)
with X(B) the differential magnetic susceptibility and m/m o the re-
duced magnetization. For (VOhP207 a comparison of the magnetic field
dependence of the elastic constant we are using either the thermodynamic
expression (eq.4) or the Landau formula (eq.6), the latter gives a much
better fit.
182
(VOh P2 0 7
This compound exists in two modifications, an ambient pressure phase
and a high pressure phase (8). ESR and ultrasonics have been performed
only in the former phase, therefore we concentrate on this one. A review on
this compound has been given recently (9). It consists of strongly dimerized
chains along the c-axis in alternate layers along the b-axis. These two chains
have energy gaps at the r point of 35K and 60K respectively. The gaps
have been determined with inelastic neutron scattering (10) and with ESR
(11, 3). From the latter experiment we get by extrapolation a critical field
value of Be = 26T, where the lower triplet branch meets the singlet ground
183
0
0
-2
-2
.".
'0
..--
-4
~
~ L
2. C11 <0
>~ -6 143 MHz -4 >~
...:::] ...:::]
-8
-6
a.) b.)
-10
0 10 20 30 40 0 10 20 30 40 50
B (T)
Figure 1. a: theoretical curves of the sound velocity for VOPO using eq.5 and the
same temperatures as in fig.lb. b: field dependence of the longitudinal sound velocity in
VOPO for temperatures from 1.6K to 11.0K measured in pulsed fields
state. Precisely around this field value we observe a sound velocity anomaly
as shown in fig.1.
The velocity anomaly observed is of the order of t1v/v o ;:::j 5 . 10- 4 .
This rather small value for the Cll mode is due to the small exchange (and
correspondingly small dJ/ de) in the a-direction. For C33 along the c-axis we
would expect a much larger effect, because in this case the propagation is
along the dimerized chain. Nevertheless we have measured a distinct tem-
perature dependence of this elastic anomaly. The effect being largest at low
temperature of 1.6K and disappearing already at 11K. We notice that eq.(5)
describes the experimental data of fig.1 much better than eq.(4). As seen
from a comparison of fig.1a and 1b the shift of the sound velocity minima to
higher fields for higher temperatures is well described by fig.1a and eq.(5).
Preliminary sound attenuation data corroborate our thermodynamic data:
Using eq.(7) we obtain WT ;:::j 0.1 so this approach is justified.
184
o 10 20 30 40 50
B(T)
Figure 2. a: The polarization and propagation directions for the different modes (Cll'
C44, C66) together with the two coupling constants J and J'.b : The field dependence of
the elastic modes Cll, C44 and C66 at 1.5K. c : The magnetization taken from ref(13). The
broken lines indicate the the positions of the different plateaus
This short review shows that the combination of ESR and ultrasonics gives
an efficient spectroscopic tool for soft magnetic modes. This spectroscopy
was applied in field regions where other experimental tools like elastic
and especially inelastic neutron scattering no longer work (> 15T). ESR-
and ultrasonic measurements were also performed on the low dimensional
spin-plateau systems NH4CuCb and CsCuCl 3 (21, 23, 22). But our ex-
periments have a much wider range of applications. We have performed
ultrasonic experiments to determine the B-T -diagram for a mixed valence
compound Ybln1-xAgxCU4(24) and the high field ultrasonic investiga-
tion of the heavy fermion compound URu2Si2 (25). Other pulsed field
experiments with ESR and ultrasonics are possible, e.g. an investigation of
semiconductors, especially heterostructure materials exhibiting quantum
Hall effect.
Acknowledgements
References
Abstract. We report measurements of the real and imaginary parts of the AC conduct-
ivity in the quantum limit, hw > kBT of insulating nominally uncompensated n-type
silicon. The observed frequency dependence shows evidence for a crossover from inter-
acting Coulomb glass-like behavior at lower energies to non-interacting Fermi glass-like
behavior at higher energies across a broad doping range. The crossover is sharper than
predicted and cannot be described by any existing theories. Despite this, the measured
crossover energy can be compared to the theoretically predicted Coulomb interaction
energy and reasonable estimates of the localization length obtained from it. Based on a
comparison with the amorphous semiconductor NbSi, we obtain a general classification
scheme for electrodynamics of electron glasses.
Key words: Long Range Interaction, Electron Glass, Silicon, Si:P, NbSi, Electrodynamics,
Localization, Insulator, Disorder
1. Introduction
189
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 189–197.
© 2003 Kluwer Academic Publishers
190
(1)
where (3 is a constant of order one, go is the non-interacting single particle
density of states (DOS), 10 is the pre-factor of the overlap integral and ~
is the localization length. The concentration dependent localization length
is predicted to diverge as (1 - x/xc)-V as the MIT is approached, where x
is the dopant concentration, Xc is the critical dopant concentration of the
MIT (xc = 3.52 x lO I8 cm- 3 in Si:P (5)) and v is the localization length
exponent.
Neglecting logarithmic factors, Eq. (1) predicts a gradual crossover from
linear to quadratic behavior as the incident photon energy exceeds the inter-
action energy of a typical charge excitation. For the case where the photon
energy, fiuJ > U(r w ), one recovers the quadratic frequency dependence, plus
logarithmic corrections, that Matt originally derived for the non-interacting
Fermi glass case (2). In the opposite limit, fiuJ < U(r w ) the conductivity
shows an approximately linear dependence on frequency, plus logarithmic
corrections, and the material is called a Coulomb glass. We should note
that Eq. (1) was derived for the case where fiuJ > ,6" the Coulomb gap
width. However a quasi-linear dependence (albeit with a different pre-
factor) and an eventual crossover to Matt's non-interacting quadratic law
is still expected even for the case where fiuJ < ,6"
1 WeI
CTI ;::::: 10 In(21o/fiuJ)·
(2)
2. Experiment
3. Results
/
Si:P ..-' /
1
10
? / " /.• / / / /
,x
()
,/
S
b 10'
69%
3
10
10- 1
Si:P
E() 10-'
S
b
10-3
100 1000
Frequency (GHz)
abrupt than the ES function predicts. The dashed line is a fit using the same
method as Ref. (7), namely forcing the linear portion to pass through the
low frequency data, as well as the origin and leaving the pre-factor of the
quadratic term as a free variable. The fit is not satisfactory in either case.
A sharp crossover as such is observed over our entire doping range and has
been observed previously in an analogous system, Si:B, for samples closer
to the MIT (7). Note, that a linear dependence is seen in the imaginary
part of the conductivity (J"2 over the whole measured frequency and doping
range (13). This is consistent with theoretical predictions (14).
Because our data spans a large range of concentrations, the doping
dependence of the crossover energy scale can be analyzed to see whether
its dependence is consistent with other energy scales, e.g. the Coulomb
interaction energy U or the Coulomb gap width 6. as per Ref. (7). Recall
that the Coulomb interaction energy between two sites forming a reson-
ant pair is U(r w ) = e 2 /clr w which is dependent on concentration via the
dielectric constant (measured, but not shown) and the localization length
193
10
2 ! !
II! II! II! ,
\:)
--N
\:)
10'
1;1 2 ~
D i;i
1.0 • • • • • •
~D
ij
0.8
8 ! ! •
0
Si:P
NbSi
\:)
! t
! I
1 - x/x c
Figure 2. The upper panel shows the ratios of the imaginary to the real part of the
complex conductivity for samples of Si:P and amorphous NbSi. The NbSi data is adapted
from Ref 1. The middle panel shows the calculated powers of a as determined from Eq.
3. The dashed line through the NbSi data is a guide to the eye. The bottom panel shows
the divergence of the prefactor of the real part of the conductivity, and the dotted line
is a simple power law fit.
MIT, CT(W) is expected to cross over to the QC dynamics (17, 18), i.e.
CTI ex: w 1 / 2 when ~, the localization length, is of the same scale as £w, the
dephasing length (the characteristic frequency dependent length scale) (19).
This should be a smooth crossover and therefore looking at a fixed window
of frequencies, a continuous change from w ----t w 1 / 2 is expected, similar to
that measured for NbSi shown in the middle panel of Fig 2.
Setting the relations for localization length and dephasing length equal
(19), one finds the crossover condition for the frequency in terms of the
normalized concentration, w ex: £0 (1 - x/xcYv where z is the dynamic
exponent. As the prefactor can vary from system to system, the fact that we
see an 0: >:::; 1 across our entire doping range in Si:P, but an 0: that approaches
1/2 in NbSi indicates that the critical regime in Si:P is much narrower and
out of our experimental window. This is consistent with simple dimensional
arguments (20) that show the crossover should be inversely proportional to
the dopant density of states. The much smaller dopant density in Si:P vs.
195
4. Discussion
Coulomb Glass
a=l
a
a( (i)) ex; (i)
Figure 3. A schematic showing the parameter space for values taken by a for (J" ex: we>.
Note that the boundaries drawn on the plot are smooth crossovers and not sharp onsets.
A classification based on a gives a taxonomy for the electrodynamics of electron glasses.
For dopings not close enough to critical in Si:P, excitation to the conduction
band may be observed before critical dynamics are.
5. Conclusion
Acknowledgements
We wish to thank Phu Tran for assisting with the cavity measurements and
Barakat Alavi for assisting with the sample preparation. We would also
like to thank Steve Kivelson and Boris Shklovskii for helpful conversations.
This research was supported by the National Science Foundation grant
DMR-Ol02405.
References
1. Erik Helgren, George Gruner, Martin R. Ciofalo, David V. Baxter, and John P.
Carini, Phys. Rev. Lett. 87, 116602 (2001).
2. N. F. Mott and E. A. Davis, Electronic Processes in Non-Crystalline Materials,
Second Edition, (Oxford University Press, Oxford, 1979).
3. A. L. Efros and B. 1. Shklovskii, in Electron-electron Interactions in Disordered
Systems, edited by A. L. Efros and M. Pollak (Elsevier New York, 1985), p. 409-482.
4. S. Tanaka and H. Y. Fan, Phys. Rev. 132, 1516 (1963).
5. H. Stupp, M. Hornung, M. Lakner, O. Madel, and H. v. LiShneysen, Phys. Rev.
Lett. 71, 2634 (1993).
6. A. L. Efros and B. 1. Shklovskii, J. Physics C 8, L49 (1975).
7. M. Lee and M. L. Stutzmann, Phys. Rev. Lett. 30, 056402 (2001).
8. M. Lee and J. G. Massey, Phys Rev. B 60,1582(1999).
9. J. G. Massey and M. Lee, Phys. Rev. B 62, R13270 (2000).
10. W. R. Thurber et al., J. Electrochem. Soc. 127,1807 (1980).
11. A. Schwartz et al., Rev. Sci. lnstrum. 66, 2943 (1995).
12. G. Gruner, Millimeter and Submillimeter Wave Spectroscopy of Solids (Springer
Verlag, Berlin, 1998).
13. E. Helgren, N.P. Armitage, and G. Gruner, Phys. Rev. Lett. 89, 246601 (2002).
14. A. L. Efros, Sov. Phys. JETP 62, 1057 (1985). An analagous three-dimensional form
of the two-dimensional theory described in this paper can be derived.
15. B. 1. Shklovskii. pr'ivate comm'Un·ication. The In used is the Bohr energy of
phosphorous.
16. W. L. McMillan, Phys. Rev. B 24, 2739 (1981).
17. H.-L. Lee, J. P. Carini, D. V. Baxter, and G. Gruner, Phys. Rev. Lett. 80, 4261
(1998).
18. H.-L. Lee, John P. Carini, David V. Baxter, W. Henderson, and G. Gruner, Science
287, 633 (2000).
19. S. L. Sondhi, S. M. Girvin, J. P. Carini, and D. Shahar, Rev. Mod. Phys. 69, 315
(1997).
20. S. Kivelson, private communication.
RENORMALIZATION GROUP APPROACHES FOR SYSTEMS
WITH ELECTRON-ELECTRON AND ELECTRON-PHONON
INTERACTIONS
1. Introduction
199
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 199–207.
© 2003 Kluwer Academic Publishers
200
IJ IJ
100
20 - g=O.O 90
---- g=0.025
...... g=0.05 80
._._. g=0.071 70
15
60
1m Xs 50
1m Xc 10
40
30
5: 20
, I
-_.--------. 10
"
O~'c..::.........~;±--~-~===== 00 0.02 0.04
0.02 0.04
0) 0)
Figure 1. Imaginary part of the dynamic charge (Xc) and spin susceptibilities (Xs)
for the symmetric parameters, U = 0.1, Ef = -0.05, 6. = 0.04 and values of
Ucff = 0.1,0.075,0.0, -0.1.
The change in the nature of the charge and spin excitations can be seen
in the spectral densities of the corresponding charge and spin susceptibilities
which are shown in figure 1. For a particle-hole symmetric model with
U = 0.1, Ej = -0.05, and 9 = 0, the spin susceptibility has a low energy
peak, while the peak for the charge susceptibility is suppressed and pushed
to higher energies due to the local repulsive interaction U. As the coupling
strength 9 to the phonons is increased these features are gradually reversed.
For the largest value of the phonon coupling shown, 9 = 0.071, the effective
local interaction is negative U eff = U - 2v = -0.1, and now the spin
excitations are suppressed and the charge excitations develop a low energy
peak. The susceptibilities shown in figure 1 are for Wo = 0.05, and the low
energy peak in the charge susceptibility at Ueff = -0.1 (g = 0.071) can be
seen to be much narrower than that of the corresponding spin susceptibility
at Ueff = 0.1 (g = 0). As WQ ----t 00 the peaks are identical as one would
expect from our earlier discussion.
The narrowing of the low energy peak in the charge susceptibility with
decrease in WQ is reflected in the spectral densities of the local electron
Green's function for the particle-hole symmetric model, which are plotted
in figure 2 for three values of WQ = 0.05,0.1,0.5 with U = 0 and a fixed value
of U eff . The first plot has U eff = -0.05 and so is in the weak to intermediate
coupling regime. The narrowing of the resonance as a function of Wo is
quite pronounced and distinct shoulders on either side of the resonance
can be seen in the plot with the smallest value of W00 The second plot is
for U eff = -0.1 which is in the strong coupling regime. Again the central
resonance can be seen to narrow as Wo is decreased. There are also distinct
202
8 - (00=0.05 8 - (00=0.05
.... (00=0.1 .... (00=0.1
-- (00=0.5 - - (00=0.5
6 6
4 4
2 2
--- ---
0
0 -0.2 0 0.2
Figure 2. In these two figures the spectral density of the local Green's function for the
particle-hole symmetric model with 6 = 0.04, and U = 0, are plotted for three values of
Woo The value of Ucff is maintained as constant in each plot with Ucff = -0.05 for the
upper and lower peaks in all three plots. The lower energy peak is associated
with the adding or removing of an electron from a double or zero occupied
impurity state, which are degenerate in the particle-hole symmetric case.
The corresponding peak above the Fermi-level is associated with the adding
or removing an electron from a singly occupied impurity level. These peaks
are the broadened counterparts of those that exist in the 'atomic limit'
V = O. Further shoulders on these peaks can be seen in the case of the
smallest value of WQ taken, which correspond to additional excitations in
which phonons are emitted or absorbed. In the opposite limit WQ ----t 00 such
excitations are pushed out to very high energies, and in the strong coupling
regime the central peak can be interpreted as a Kondo resonance, with
a Kondo temperature TK given by TK = (IUeffl~/2)1/2exp( -1TIUeffl/8~ +
1T~/IUeffl). The Kondo physics in this case is associated with charge or
pseudo-spin degeneracy and not spin degeneracy.
The effects of the narrow peak in the charge susceptibility for the
particle-hole symmetric model with U = 0 and Ueff < 0 can be seen the
spectrum for the phonon Green's function D(w) = ((b + bt : b + bt )). A
relation between the two can be derived from the equations of motion (10),
(2)
The low energy peak in Xc( w) induces a low energy peak in the phonon
spectrum, which survives even if we take the limit WQ ----t 00.
203
The NRG calculations for the impurity models can be extended to the
calculation of the response functions for lattice models in the framework
of dynamical mean field theory (DMFT). The DMFT has been applied
widely and is extensively reviewed in reference (6). The lattice problem is
mapped onto an effective impurity problem, which is then subject to a self-
consistency constraint. There are many ways of carrying out calculations for
the effective impurity model, and the Monte Carlo approach is one that has
been used extensively (13). However, the Monte Carlo calculations cannot
be performed for T = 0, or for very low temperatures, so the NRG approach
has a distinct advantage in this regime. The NRG method is also capable
204
of resolving very low energy scales. Most applications so far of the NRG-
DMFT technique have been to the Hubbard and periodic Anderson models.
Using the approach developed in the previous section, we are in a position
to carry out calculations which also include a coupling to phonons, and we
have focussed our attention so far on the Holstein model (9). In this model
the electron occupation number at each site is linearly coupled to a local
boson mode of frequency woo The corresponding Hamiltonian has the form,
where i is a site index. In the limit Wo ----7 00 with Ueff = -2g2/wo finite the
model maps into an effective 'negative-U' Hubbard model. The dynamic
mean field theory approach is strictly speaking only valid for the infinite
dimensional model but should constitute a good starting point for under-
standing the physics of this model, particularly in the intermediate and
strong coupling regime where most methods break down. The effective im-
purity model in this case is the U = 0 Anderson-Holstein model considered
in the previous section. The effective density of states of the host, however,
has to be calculated self-consistently from the DMFT constraint. Results
of applying the DMFT-NRG approach to this model are shown in figure 3.
The first plot gives the spectral density of the local one-electron Green's
function for the particle-hole symmetric model. As the coupling strength 9
is increased from zero, a narrow peak develops at the Fermi-level, which is
quite similar to that seen in the results for the impurity model in figure 2
in the same parameter regime. The results are also quite similar to those
obtained by application of the Migdal-Eliashberg approach (14) in which
all perturbational theory diagrams are summed subject to the neglect of
all vertex corrections.
As the coupling is further increased a pseudo-gap with a very nar-
row central peak begins to develop. At a critical coupling strength gc the
central peak disappears and a gap in the excitation spectrum develops
continously. The quasi-particle spectral weight z decreases monotonically
with increasing coupling strength and goes to zero at 9 = gc. Due to the
attractive effective interaction the electrons pair to form bipolarons as the
interaction becomes stronger, and the lower peak is associated with the
bipolaron formation. The Migdal-Eliashberg (ME) approach breaks down
before a pseudogap develops and the results are limited to the weak coupling
regime. If the ME approach is applied to the Anderson-Holstein model its
limitations can be clearly seen as it does not develop the higher and lower
atomic-like peaks.
It is interesting to contrast these results with those for the positive U-
Hubbard model where a metal-insulator gap develops at a critical coupling
205
- g=0.03
._._. g=0.08
....... g=0.098
---- g=0.12
,, cr(m)
pew) ,,
,,
0.5
, \
\
,,
,,
,,
Figure 3. The spectral density for the local one-electron Green's function p( w) and that
for the phonon Green's function O"(w) for the infinite dimensional Holstein model plotted
for various values of the coupling strength g.
4. Conclusions
Acknowledgements
References
13. J.K. Freericks, M. Jarrell and D.J. Scalapino, Phys. Rev. B 48, 6302 (1993).
14. J.P. Hague and N. d'Ambrumenil, cond-mat/0l06355 (2001).
15. R. Bulla, Phys. Rev. Lett. 83, 136 (1999).
16. S.A. Trugman, J. Bonca, and L.-C. Ku, Int. J. Mod. Phys. B 15, 2707 (2001).
17. O. Gunnarsson, Rev. Mod. Phys. 69, 575 (1997).
18. A.J. Millis, P.B. Littlewood and B.!. Shraiman, Phys. Rev. Lett. 74,5144 (1995).
QUANTUM PHASE TRANSITIONS IN MODELS OF MAGNETIC
IMPURITIES
Abstract. Zero temperature phase transitions not only occur in the bulk of quantum
systems, but also at boundaries or impurities. We review recent work on quantum phase
transitions in impurity models that are generalizations of the standard Kondo model
describing the interaction of a localized magnetic moment with a metallic fermionic host.
Whereas in the standard case the moment is screened for any antiferromagnetic Kondo
coupling as T ----> 0, the common feature of all systems considered here is that Kondo
screening is suppressed due to the competition with other processes. This competition
can generate unstable fixed points associated with phase transitions, where the impurity
properties undergo qualitative changes. In particular, we discuss the coupling to both non-
trivial fermionic and bosonic baths as well as two-impurity models, and make connections
to recent experiments.
Key words: Quantum Phase Transition, Kondo Effect, Local Criticality, Numerical Renor-
malization Group, Two-Impurity Kondo Model, Two-Channel Kondo Model
1. Introduction
209
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 209–217.
© 2003 Kluwer Academic Publishers
210
0.06
~
.3 a b
~
OJ
D-
E 0.04
2
SC
<1
symmetric case
0.02
asymmetric case
fixed point A B LM
\
0.00
0.00 0.25 0.50 0.75
Figure 1. (a) Schematic phase diagram near a quantum phase transition between fixed
points A and B upon variation of a control parameter x; the T = 0 critical point controls
the dynamics in the T > 0 quantum critical region whose crossover boundaries are given
by T ~ Ix - where l/ and z are the correlation length and dynamical critical
exponents, respectively. (b) T = 0 phase diagram for the pseudogap Anderson model in
the p-h symmetric case (solid line, U = 10- 3 , E f = -0.5 . 10- 3 , conduction band cutoff
at -1 and 1) and the p-h asymmetric case (dashed line, Ef = -0.4·10-:1); 6 measures
the hybridization strength, .6.(e) == 7fV2 p(e) = 6lel r .
L1=L1c L1>L1c
local moment quantum critical strong coupling
a b c
2.0
z
W
~
z
<::
1.0
Figure 2. Flow diagrams for the low-energy many-body excitations obtained from
the numerical renormalization group for the three different fixed points of the soft-gap
Anderson model. N is the number of iterations of the NRG procedure, A the NRG
discretization parameter.
so we can easily recognize the three different fixed points for ,6. < ,6.c (Fig.
2a), ,6. = ,6.c (Fig. 2b), and ,6. > ,6.c (Fig. 2c). The structure of the LM
and SC fixed points can be easily understood as that of a free conduction
electron chain (10, 11). The combination of the single-particle states of the
free chain leads to the degeneracies seen in the many-particle states of the
LM and SC fixed points (Fig. 2a and Fig. 2c). In contrast, the structure of
the quantum critical point is unclear. Degeneracies due to the combination
of single-particle levels are missing, probably because the quantum critical
point is not build up of non-interacting single-particle states.
The pseudogap Kondo model has been proposed (8) to describe impurity
moments in d-wave superconductors (r = 1), where signatures of Kondo
physics have been found in NMR experiments (12). Furthermore, a large
peak seen in STM tunneling near Zn impurities (13) can be related to the
impurity spectrum in the asymmetric pseudogap Kondo model.
Models of two impurities offer a new ingredient, namely the exchange in-
teraction, I, between the two impurity spins which competes with Kondo
screening of the individual impurities. This inter-impurity interaction, which
can lead to a magnetic ordering transition in lattice models, arises both from
direct exchange and from the Ruderman-Kittel-Kasuya-Yosida (RKKY)
interaction mediated by the conduction electrons.
In the simplest model of two S = ~ impurities, a ground state sing-
let (Stat = 0) can be realized either by individual Kondo screening (if
I < TK) or by formation of an inter-impurity singlet (if I > TK)' It has
been shown that these two parameter regimes are continuously connected
(without a T = 0 phase transition) as I is varied in the generic situation
without particle-hole symmetry. Notably, in the particle-hole symmetric
case one finds a transition associated with an unstable non-Fermi liquid
fixed point (27, 28, 29).
Quantum phase transitions generically occur in impurity models show-
ing phases with different ground state spin. For two impurities, this can be
realized by coupling to a single conduction band channel only (30). In this
case, a Kosterlitz-Thouless-type transition between a singlet and a doublet
state occurs, associated with a second exponentially small energy scale in
the Kondo regime (30). The physics becomes even richer if multi-channel
physics is combined with multi-impurity physics - here, a variety of fixed
points including such with local non-Fermi liquid behavior can be realized.
Experimentally, quantum dots provide an ideal laboratory to study
systems of two (or more) "impurities" - note that the local "impurity"
states can arise either from charge or from spin degrees of freedom on each
quantum dot. In particular, a number of experiments have been performed
on coupled quantum dot systems which can be directly mapped onto mod-
els of two Kondo or Anderson impurities (31). In addition, experimental
realizations of two-impurity models using magnetic adatoms on metallic
surfaces appear possible.
6. Summary
Acknowledgements
References
28. O. Sakai, Y. Shimizu and T. Kasuya, Solid State Comm. 75, 81 (1990); O. Sakai
and Y. Shimizu, J. Phys. Soc. Jpn 61, 2333 (1992), ibid, 61, 2348 (1992).
29. 1. Affleck, A. W. W. Ludwig and B. A. Jones, Phys. Rev. B 52, 9528 (1995).
:30. M. Vojta, R. Bulla and W. Hofstetter, Phys. Rev. B 65, 140405(R) (2002).
31. W. G. van der Wiel, S. De Franceschi, J. M. Elzerman, S. Tarucha, L. P. Kouwen-
hoven, J. Motohisa, F. Nakajima and T. Fukui, Phys. Rev. Lett. 88, 126803
(2002).
INSTABILITY OF THE FERMI-LIQUID FIXED POINT IN AN
EXTENDED KONDO MODEL
M. LAVAGNA*
Commissariat al'Energie Atomique, DRFMC/SPSMS, 17, rue
des Martyrs, 38054 Grenoble Cedex 9, France
A. JEREZ
European Synchrotron Radiation Facility, 6, rue Jules Horow-
itz, 38043 Grenoble Cedex 9, France
D. BENSIMONt
Department of Appled Physics, Hongo 7-3-1, University of Tokyo,
Tokyo 113-8656, Japan
Key words: Non-Fermi Liquid, Strong versus Intermediate Coupling Fixed Point, Exten-
ded SU(N) Kondo Model
219
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 219–227.
© 2003 Kluwer Academic Publishers
220
1. Introduction
the underscreened regime when K < 28 (where 8 is the value of the spin
in 8U(2)), the SC fixed point is stable. It becomes unstable in the other
regime K > 28, known as the overscreened regime. The underscreened
regime corresponds to the one-stage Kondo effect with the formation of an
effective spin (8 - 1/2) resulting from the screening of the impurity spin
by the conduction electrons located on the same site. The system described
by the SC fixed point behaves as a local Fermi liquid. The instability of
the SC fixed point in the overscreened regime is associated with a multi-
stage Kondo effect in which successively the impurity spin is screened by
conduction electrons on the same site, and then the dressed impurity is
screened by conduction electrons on the neighboring site and so forth. The
overscreened regime leads to the existence of an intermediate coupling fixed
point with non-Fermi-liquid excitation spectrum and an anomalous residual
entropy at zero temperature. It has been recently put forward (3, 4) that
other sources of instability of the SC fixed point may exist other than
the multiplicity of the conduction electron channels. Recent works have
shown that the presence of a more general Kondo impurity where the spin
symmetry is extended from 8U(2) to 8U(N), and the impurity has mixed
symmetry, may also lead to an unstability of the SC fixed point already
in the one-channel case. In the large N limit, Coleman et al. (3) have
found that the SC fixed point becomes unstable as soon as the impurity
parameter q (defined below) is larger than N /2 whatever the value of 28
(defined below) is, giving rise to a two-stage Kondo effect. This result opens
the route for the existence of an intermediate coupling fixed point with
eventually non-Fermi-liquid behavior.
It is worth noting that the supersymmetry theory, or more specific-
ally the taking into account of general Kondo impurities appears to offer
valuable insights into the two issues raised by the breakdown of the Fermi
liquid theory that we have summarized above, i.e. both the existence of
locally critical modes and the question of the instability of the SC fixed
point. Somehow it seems that the consideration of general Kondo impurities
captures the physics present in real systems with the coexistence of the
screening of the spin by conduction electrons responsible for the formation
of quasiparticles , and the formation of localized magnetic moment that
persists and eventually leads to a phase transition as the coupling to other
impurities becomes dominant.
The aim of this work is to study the 8U(N), extended single-impurity
Kondo model in the one-channel case (10). We would like to investigate
how the system behaves when not only the values of the parameters (28, q)
of the representation vary, but also the number nd of conduction electrons
on the neighboring site does. We want to discuss the effect of nd on the
stability of the SC fixed point and to further understand the nature of the
222
We consider the case where J -----+ 00 and we can neglect the kinetic energy
in Eq. 1. In this limit the model can be solved exactly in terms of the
invariants associated to the spin of the electrons and of the impurities (9).
The eigenvalues are of the form
(2)
where I denotes the impurity spin, Y the spin of the n c conduction electrons
coupled to the impurity at the origin, and Rsc the spin of the resulting SC
state at the impurity site. The quantities C2 are the 8U(N) generalization
of the 8U(2) eigenvalues, 8(8+1), and can be readily evaluated (for details,
see Ref. (10)).
223
with, :::::: )(28 + N - l)Cir-_\. Here, cl :::::: cl(O). We denote the ground
state energy by Eo.
The effect of the kinetic term in Eq. 1 is, to lowest order in perturbation
theory, to mix the ground state with excited states where the number of
electrons changes by one. There are three such states, which we denote
by IG8 + l)S, IG8 + l)A and IG8 - I). The labels S(A) indicate that
the additional electron is coupled symmetrically(antisymmetrically) to the
ground state. The states are readily obtained by deriving the relevant SU(N)
Clebsch-Gordan coefficients (10).
Notice that the behavior of both Eq. 4 and Eq. 5 are controlled by the
same factor. This result has the immediate following physical consequence.
The change of sign of Jeff (Fig. 1, Right) -and hence of the stability of
the SC fixed point- is directly connected to the change in the behavior of
b..Etf rv b..E6 with nd. In particular, when b..E6 = b..Etf, b..Etf = -(2t 2 I J)
independently 1 of nd, q or 25. In the regime where the SC fixed point
is stable, qlN < 1/2, Jeff < 0, the lowest energy corresponds to nd =
1, whereas for qlN > 1/2, Jeff> 0, the energy expressed in Eq. (4) is
minimized for nd = (N - 1) (Fig. 1, Left). This is precisely the mechanism
behind the two-stage quenching. The accumulation of electrons on site 1
is not related to Jeff which is independent of nd, but results from the
dependence of b..Etf rv b..E6 with nd.
We finish by making some remarks on the physical properties of the
model in the different regimes. As is common to all models with an antifer-
romagnetic Kondo coupling, there will be a crossover from weak coupling
above a given Kondo scale, T K , to a low-energy regime. When the SC
fixed point is stable, we should expect for T « T K a weak coupling of
the effective impurity at site-O with the rest of the electrons. The physical
properties at low temperature are controlled by the degeneracy of the ef-
fective impurity, d([25 -1]) = C~+iS-2' Thus, we should expect a residual
entropy Si rv In C~+iS-2 and a Curie susceptibility, Xi rv C~+is_2IT, with
logarithmic corrections (9, 16). This is the result that we would expect for
a purely symmetric impurity. Contrary to the purely bosonic case, only
_ -2 ~ . / /
~g ~ I,'
oz °f---==~:;-:f'':'''''''''''==------1
,,/."
,.""
<~
<I -3 -,~ ...
/' .'
-2 /
I
-4 1- - nd~
- n =N-l
d
1 I I
I
-4 I
I
:
0,25 0,5 0,75 ••• (125 0,5 (J,75
gIN q/N
Figure 1. (Left) Leading term in the energy shift, 6.E(i ~ 6.E5, as a function of q/N,
for 1 < nd < N - 1 (shaded region), and in the limiting cases nd/N « 1 (dashed line),
and nd/N ~ 1 (solid line). Notice that the value at q/N = 1/2 is equal to J, for any -2e/
nd. (Right) Energy difference, (6.E[~ - 6.E((), as a function of q/N, for different values
of 28 in the large-N limit.
3. Conclusions
We have studied a Kondo model where the spin of the impurity has mixed
symmetry, as a way of incorporating the phenomena of local moment screen-
ing and magnetic correlation that is observed in some heavy fermion com-
pounds. Such model is naturally realized by extending the spin symmetry
to SU(N), and it displays a rich phase diagram. We find that as long as
the bosonic component of spin is of order N, there is a transition around
226
Acknowledgements
We would like to thank V. Zlatic and A. Hewson for organizing this very
interesting meeting. We are most grateful to N. Andrei for continuous en-
couragement and discussions. We also thank S. Burdin, P. Coleman, Ph.
Nozieres, and C. Pepin for helpful discussions.
References
18. Q. Si, S. Rabello, K. Ingersent and J.L. Smith, Nature 413, 804 (2001).
19. A. Schroder, G. Aeppli, E. Bucher, R. Ramazashvili, and P. Coleman, Phys. Rev.
Lett. 80, 5623 (1998).
20. A. Schroder, G. Aeppli, R. Coldea, M. Adams, O. Stockert, H. v. Lohneysen, E.
Bucher, R. Ramazashvili, and P. Coleman, Nature 407, 351 (2000).
21. Steglich, F., B. Buschinger, P. Gegenwart, M. Lohmann, R. Helfrich, C. Langham-
mer, P. Hellmann, L. Donnevert, S. Thomas, A. Link, C. Geiber, M. Lang, G. Sparn,
and W. Assmus, J. Phys.: Condo Mat. 8, 9909 (1996).
PROJECTION OF THE KONDO EFFECT BY RESONANT
EIGENSTATES INSIDE A CIRCULAR QUANTUM CORRAL
Abstract. We calculate the resonant eigenstates inside a circular quantum corral, mod-
eling the confining potential by a continuous radial 6-function. In order to describe the
corral with a magnetic impurity inside, we use the impurity Anderson model and second
order perturbation theory in the Coulomb repulsion U. The effect of the impurity in
the conduction electron density of states reveals a mirage of the Kondo resonance at a
position different from that of the impurity.
Key words: Quantum Corrals, Mirage, Surface States, Magnetic Impurity, Kondo Effect,
Anderson (Impurity) Model, Perturbation Theory, Differential Conductance, Nanoscale
Structures, Scanning Tunneling Microscopy
1. Introduction
229
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 229–237.
© 2003 Kluwer Academic Publishers
230
was observed not only at that position, but also with reduced intensity at
the other focus (2). This "mirage" is the projection of the Kondo effect,
transmitted from the impurity to distant points by the corral eigenstates.
Similar experiments were done in circular corrals (4). The corral confine-
ment is needed in order to separate the conduction electronic energy levels
(otherwise forming a continuum) and reduce the destructive interference
produced by incoherent superposition of standing waves (5). Among these
eigenstates, the density of the one closest to Ep is clearly displayed in the
change of differential conductance b.dI/dV after adding the impurity (2).
Some theories of the Kondo mirage have attempted to fit or explain the
experimental data assuming the existence of the Kondo resonance, without
actually calculating it (5,6,7). The line shape of b.dI/dV, in the many-
body approaches which can explain it, is very sensitive to the width of
the resonance levels, and this width was taken as a parameter (8, 9, 10).
If this width is reduced to zero, as in ordinary Lanczos calculations (11),
the Kondo resonance splits into two delta functions out of Ep and the
experimental line shape cannot be reproduced. This problem persists if
hopping between impurity and bulk is considered, in spite of the broadening
of the delta functions (12). On the other hand, if the width of the resonance
is very large, the mirage disappears (8).
In this work we calculate the widths of the conduction states 8i for a
circular corral and solve the impurity Anderson model which describes the
mirage experiment using perturbation theory in the Coulomb repulsion U.
Here we consider first the corral without impurity. We model the boundary
of an empty circular corral by a continuous potential W 8(r - ro), where r is
the distance to the center of the circle and ro its radius. The approximation
of a continuous boundary (instead of discrete adatoms forming the corral)
simplifies considerably the mathematics and is justified by the fact that the
Fermi wave length for surface electrons 2Jr/ kp ':::' 30 A is larger than the
average distance between adatoms rv 10 A.
The angular momentum projection perpendicular to the surface lz =
m is a good quantum number. For each m and energy E = (nk)2 /2m~,
where m~ is the effective mass, the eigenstates can be written in the form
!.pkm(r, e) = ?/Jkm(kr) e ime in polar coordinates. The solutions of the 2D
radial Schrodinger equation with this potential can be written in the form
?/Jkm = ?/J::-me(ro - r) + ?/Jkme(r - ro), where e(r) is the step function and
where Jm (Ym ) is the Bessel function of the first (second) kind. We nor-
malize the wave functions CPkm(r, e) inside a hard wall box of large radius
R and take the limit R ---7 00. In this limit, the eigenstates of the problem
form a continuum in the k variable. However, inside the corral, resonances
are formed for finite W (13, 14). The resonant eigenstates of the problem
can be obtained by scattering theory methods, specifically, by finding the
complex poles of the scattering matrix S(k) (13, 14).
The essential information on the spectral properties observed by STM
is contained in the surface conduction electrons Green's function. In the
absence of the impurity, it can be written in the form
J (k m )J (k m ,) im(e-e')
GO(z'r e r' e') =
s , , , ,
Le m
n m n r m
Z -
n r e
E~~ + io~~ ,
(4)
n,m,
where k:';: are the poles in complex k-space of the scattering matrix (m
and n are angular and radial quantum numbers respectively) (15), with
( nknm )2/2m*r 2 = Em - io m and om > 0 Here Em is the energy of the
eo n 71,' n . 'n
resonant eigenstate, 0:';: > 0 is its width, and e~ are the residues at the
poles of the integral in Eq. (3) (12). We have taken an effective mass m~ =
0.38 m e (1, 16). Jm(kr) has no singularities in the complex k-plane. The
poles appearing in Eq. (4) are poles of the normalization factor f (when
continued into the complex k-plane). From the properties of 1/;;;n and 1/;f:rn'
and its derivatives at the boundary ro, one obtains for f:
2 k
(5)
f = 2R(A~(k, W) + B~(k, W))'
where the coefficients Am(k, W) and Bm(k, W) depend on Jm(krO) and
Ym(krO) and the dependence on W is now explicit. In Fig l(a) we show
the adimensional product f2 Rro, proportional to the density of states, as
a function of k (real) for two different values of Wand for states with
232
50
(a) (b)
>' 150
40
.............. W = 1.1geV.A
Q)
.§. 120
,,
\,
30 - - W = 2.38 ev.A
....
0
90
0::: 20
60
~
10
[ ... 30
------.
0
./ ~(,Q
0
0 3 6 9 12 15 18 21 0.0 0.5 1.0 1.5 2.0 2.5 3.0
k [rt W [eV.A]
Figure 1. (a)f2 Rro as a function of radial momentum k, for m=O states. (b) Width of
the resonant state n = 4, m = 1, 5l as a function of the confinement parameter W.
In this section we introduce the magnetic impurity in the form of one non-
degenerate localized d orbital. We consider surface states, described inside
the corral by the Green's function G~, hybridized with the d orbital, for
which there is an important intra-orbital repulsion U (5, 8, 9, 10, 11).
We also include a hybridization of the d orbital with bulk states. The
233
61~-----'--~-----'---~~~~
--pd(m)
5
···········pdo(m)
4
5:'
~3
~2
0:
1
o -_.__._.._...-_...-... ~. ..- _--- _- .
Figure 2. Impurity density of states for u=o (dotted line) and U=l eV (solid line).
Parameters are OF = 20 meV, Ob = 35 meV and Vs = 0.45 eV, and E~ff = EF.
GO(w) = 1 (9)
d w-E~ff +i(5b-(V).)2G~(Ri,Ri,W)·
The term (5b represents here the hybridization between bulk states and the
impurity d level, and is a parameter in our calculations. The function I;(w)
234
/
0.9-r-r~~~---.--.-~~~~---,--"
0.0 +--::,------~~-----,.._~~~_____1
0.6 ••.....
~
>
-0.3
:>
~ 0.3
.....
~ -0.6
Figure 3. Real and imaginary parts of the self-energy for sl = 20 meV (solid line) and
sl = 40 meV (dotted line).
"E(iwz, T) (10)
m
0.03
6.:0.02
....,..0.01 1.0
"'.0.00 0.5
cc
<1- 0.01 o.o~c
• .>;
-1.0 -0.5 -0.5
-1.0
Pigl/,re 4. (a) Space dependence of -6.(:-.(£1") and (b) dell~i1.y of the llurd wall cigcn~tatc
with m = 1 and n = 4,
\Ve have studied the resonances produced inside a circular quantum conal,
assuming a continuous W6(1' - TO) confining potential in a 20 electron
gas. These resonant. eigenstates are very similar to the eigenstal.es of a hard
wall potential and t.hey can be obtained from the poles of the normalization
factor f. This fact. can be easily interpret.ed, since a pole in f indicates that
the state has little weight outside the corral, leading to a "quasi-confined"
state. The introduction of a width iJ.; for each resonanl.:e is the main differ-
ence with a calculation assuming a hard wall corral. Our model leads to a
resonance &., proportional to its energy lSi, in a. first approximation.
236
Acknowledgements
References
1. M.F. Crommie, C.P. Lutz, and D.M. Eigler, Science 262, 218 (1993).
2. H.C. Manoharan, C.P. Lutz, and D.M. Eigler, Nature (London) 403, 512 (2000).
3. S.L. Hulbert, P.D. Johnson, N.G. Stoffel, W.A. Royer, and N.V. Smith, Phys. Rev.
B 31, 6815 (1985).
4. H.C. Manoharan, PASI Conference, Physics and Technology at the Nanometer Scale
(Costa Rica, June 24 - July 3, 2001).
5. D. Porras, .J. Fermindez-Rossier, and C. Tejedor, Phys. Rev. B 63, 155406 (2001).
6. O. Agam and A. Schiller, Phys. Rev. Lett. 86, 484 (2001).
7. G.A. Fiete, J. S. Hersch, E. J. Heller, H.C. Manoharan, C.P. Lutz, and D.M. Eigler,
Phys. Rev. Lett. 86, 2392 (2001) ..
8. A.A. Aligia, Phys. Rev. B 64, 121102(R) (2001).
9. A.A. Aligia, Phys. Status Solidi (b) 230,415 (2002).
10. G. Chiappe and A.A. Aligia, Phys. Rev. B 66, 075421 (2002).
11. K. Hallberg, A.A. Correa, and C.A. Balseiro, Phys. Rev. Lett. 88, 066802 (2002).
237
Abstract. We consider electron transport through a quantum dot in the Coulomb block-
ade regime, when the dot has a total spin ~. The low energy physics of this system is
modeled by a Kondo-type Hamiltonian describing spin-dependent tunneling and exchange
interaction of the electron spins in the leads and the local spin. We show that in the regime
of large transport voltage V and magnetic field B with max(V, B) » T K , where T K is the
Kondo temperature, renormalized perturbation theory is valid. Physical quantities such
as the differential conductance G and the local magnetization 1\;1 may be calculated in a
controlled way by summing the leading logarithmic terms in each order of perturbation
theory. We develop a poor man's scaling renormalization group treatment for frequency
dependent coupling functions, which allows one to calculate G and 1\;1 to leading order
in l/£n[(V, B)/TK].
Key words: Nonequilibrium Transport, Quantum Dot, Kondo Effect, Conductance, Poor
Man's Scaling, Renormalization Group Method
1. Introduction
239
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 239–246.
© 2003 Kluwer Academic Publishers
240
+-1 "
L...J Ja /a S· t
(ca/k/o-,To-'o-Cako-), (1)
~,k' ,0-,0-' ,a,o:' =L,R
where S is the local spin operator on the dot (assumed to be S = ~) and
T is the vector of Pauli matrices. The c~ko- create conduction electrons
with momentum k and spin (]" in lead a. The chemical potential shifts
induced by the bias voltage are given by ML,R = ± V /2 for the left (L) and
right (R) lead. The exchange energies Jaa /, a = L, R, will be assumed
to have the symmetry J RR = J LL and J LR = J RL . We shall use the
dimensionless coupling constants gd = NOJRR and gLR = NOJLR , with
No the local density of states at the Fermi energy, assumed to be flat in
the accessible regime I w I~ V, B. A magnetic field is included, inducing
a Zeeman splitting of the magnetic levels of the local spin = -,B /2, w,
,= ±1. The effect of the magnetic field on the conduction electrons gives
241
only rise to particle-hole asymmetry terms, which are usually small and
may be neglected.
In lowest order perturbation theory the current through the dot is given
by the Golden Rule expression
1= :~ Jdw L n, [gUw-ILL (1 -
,=±l
IW-ILR)
where gl and g2 are the bare dimensionless coupling constants for spin-
nonflip and spin-flip interaction, gl = g2 = NOJLR, IW-lLu is the Fermi
function in the lead 0: = L, R, and IB is the Zeeman energy transfer taking
place in a spin flip process. At low temperatures, T « V, B, the product of
Fermi functions limits the energy integration to the window ILR < w < ILL
and ILR + IB < w < ILL, respectively. The current is seen to depend on
the occupation n, = ~ (1 + 1M) of the local spin states, where M is the
magnetization.
For sufficiently small voltage, V « T, the local spin system is in thermal
equilibrium. In the opposite limit, V » T, the stationary current through
the dot drives the system out of equilibrium. The occupation numbers n,
gt
are then determined by the rate equation n, = C,{n,'} = 0, where C
is the collision integral. This leads to the condition (assuming JLR = hL
here)
nr L
a,a'=L,R
J dwg§Iw-lLu ( 1 - IW-lLu,-B) =
= nl L
a,a'=L,R
J
dwgUw-lLu (1 - IW-lLu,+B) (3)
The spin-flip coupling constant g2 cancels out of this equation, yielding the
nonequilibrium magnetization as given by (5) in the limit J -----+ O.
In the limit V -----+ 0, the equilibrium result M = tanh 2~ is recovered,
whereas in the limit V » B, T, M = 2(/, independent of temperature. This
result has been obtained independently in (18, 20).
In order to calculate higher order contributions in perturbation theory
in J we switch now to a more formal description in terms of nonequilibrium
Green's functions. We find it convenient to represent the local spin operator
in pseudofermion (PF) language (19), S = ~ L", 1," where 1:;
I:;T",
creates a pseudofermion of spin I =1, 1 at the dot. The projection onto the
physical sector of Hilbert space, with pseudofermion occupation number
Q = L, 1:;1, = 1, is done by adding a term AQ to the Hamiltonian and
242
a c
E)+ YX
....
.,. ...... - -... ...,
Figure 1. Feynman diagrams for (a) PF self-energies, (b) current and (c) vertices
entering the I-loop RG equation. PF (electron-) propagators are displayed as dashed
(full) lines.
taking the limit A ----+ 00. This means that the PF system is taken in the
low density limit. The Feynman diagrams of perturbation theory in the
exchange interaction will therefore have one PF loop at most.
The local magnetization can be calculated in terms of the PF Green's
function G~(w) as M = L:f=±1 I J :f:iG~(w), where G< is found by solving
a quantum kinetic equation. In steady state one finds
M = Z (5)
coth 2~ [~+ g~B(l + £(B))] + gIR(c(B) + c(-B))
where
Even for small coupling constants g and for sufficiently large V and B,
such that V, B » TK, but still in the scaling regime V, B « D, such that
gin(D IV) « 1, bare perturbation theory converges slowly. It is necessary
to sum the leading logarithmic contributions in all orders of PT. In the
equilibrium state a powerful method is available to perform this resumma-
tion in a controlled way: the perturbative renormalization group method
(9). It makes use of the fundamental idea that a change of the cut-off D
can be fully absorbed into a redefinition of the coupling constant g. As
long as the running coupling constant g(D) is small, the change of g under
an infinitesimal change of D, 8g/mnD, may be calculated in perturbation
theory. It is well known that in the equilibrium Kondo problem the coup-
ling constant is found to grow to infinity, thus leaving the perturbative
regime beyond g rv 1. A nonperturbative treatment is then necessary, as
shown by Wilson (10) in his pioneering work on the numerical RG. In
the nonequilibrium situation, the RG flow is cut-off by inelastic processes
already within the weak coupling regime, such that perturbation theory is
valid (14). This is true at least for the case hR = JLL, JRR, considered here.
While the discussion of the RG flow in (14) was a qualitative one, motivated
by results obtained in the so-called "Non-crossing approximation" for the
Anderson model, and applied only to the limit B = 0, here we consider the
RG formulation on a more fundamental level. A different and considerably
more involved real-time RG scheme has been developed by Schoeller and
Konig (21).
An analysis of the PT result Eq. (5) suggests that the frequency de-
pendence of the couplings becomes important. In order to understand how
these frequency dependencies are generated on the lowest level, i.e. in one-
loop order, it is sufficient to analyze the behavior of the vertex corrections
shown in Fig. lc under a change of cut-off. Logarithmic corrections in
perturbation theory are generated by PF - conduction electron bubbles
containing the product of a real part of a PF Green's function, 1/(w ±
B /2), and the Keldysh component of the local conduction electron line,
-27fiNo tanh[(w - J1o;)/2T]. If the energy of the PF - conduction electron
intermediate state is within the interval [- D, D] the process will contribute
to the renormalization of the coupling function g (w), otherwise it will not. In
general the coupling functions depend on three frequencies (taking energy
244
conservation into account). Using the fact that the spectral function of the
PFs is sharply peaked at w = ~B /2, we set two of the three frequencies to
w = ~B /2, keeping only one frequency variable. In addition, the frequency
dependence may be neglected within the running band-width, I w 1< D. The
spin structure of the general coupling functions is given by two invariant
amplitudes, fh for spin flip and fh for spin non-flip processes
where 9~.;;' denotes the coupling function for conduction electrons of spin (J"
and ene;gy w in lead 0:( a = (0:, (J", w)) interacting with a pseudofermion in
state b = (" w f) and going into states a', b'. The two frequency-dependent
running coupling functions .ih(w) and fh(w) obey the following flow equa-
tions,
8fh(w)
8lnD
8fh(w)
(8)
8inD
with initial condition .ih(w) = fh(w) = JNo and 8 w = 8(D- 1 w I). In the
limit V, B ----7 0, Eq. (8) reduces to the well-known scaling equations (6).
There is, however, still one additional effect missing: the finite relaxation
of the spins even in the limit T ----7 0 leading to inelastic broadening of the
local spin levels, which we identify with the transverse spin relaxation rate
r = 1/T2 . Within the PF representation this effect shows up as a finite
imaginary part of the PF self-energy.
Assuming the RG flow to be stopped at the scale r, we replace the step
functions 8 w in Eq. (8) by 8(D - vw
2 + r 2 ). The decay rate r has to
r :12 L jdw[91(W)2fw-lLa(1-fw-lLa')
a,a'=L,R,,=i,l
+92(W - I B /2)2 fW-/La (1 - fW-/La,-,B)] (9)
the renormalized r is obtained by replacing 91,2 by 91,2(W) as determined
from the solution of (2).
We are now ready to calculate further physical quantities. The renor-
malized value of the magnetization is obtained by substituting 91,2 (w) in
245
a - RG c 0.15
._. o(i)
0.1 -- o(i) 0.1
o
(9
. I ·~·~_--.j~~r
(5 --------1 i------ __
L-'-'-l
a h-h-+-r-+---,--r---r--h----+--.-+---r-r----r-+-,H-
1
b
0.8
~
0.6
place of 91,2 in the Golden Rule expression (3). In Fig. 2b we show the fully
renormalized result for M as a function of V / B for different values of B.
The charge current I is calculated from Eq. (2) inserting the renor-
malized coupling functions. Fig. 2a shows a comparison of the differential
conductance G(V) = dI/dV obtained in this way, with the bare result (2)
and the PT result including leading logarithmic corrections.
The peak structures in G(V, B) appearing at V = ±B have been detec-
ted in experiment. In order to reach large values of B /TK , it is necessary to
have relatively low TK. This happened to be the case in transport through
metallic point contacts, containing a magnetic impurity (22). In Fig. 2c the
result of our theory using the value of TK ':::::' 30mK given in (22) is compared
to the experimental data, after subtracting a background contribution (see
(17) for details). The agreement is seen to be excellent.
4. Conclusion
5. Acknowledgments
References
T. A. COSTI
Universitat Karlsruhe, Institut fur Theorie der Kondensierten
Materie 16128 Karlsruhe, Germany
1. Introduction
247
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 247–256.
© 2003 Kluwer Academic Publishers
248
and r for the dots in (1, 5) are U rv 0.5 - 2.0 meV and r rv 0.2 - 0.3 meV,
so these dots are strongly correlated.
For temperatures r « T « U quantum fluctuations are small, and
transport is dominated by charging effects. This regime is well understood
(6). The conductance, G, exhibits a series of approximately equidistant
peaks as a function of Vg, with spacing U. The peaks correspond to a
fractional number of electrons on the dot and alternate "Coulomb blockade"
valleys to either an even or an odd number of electrons.
In this paper we shall be interested in the regime T ;S r « U, where
the strong quantum fluctuations can lead to a dramatic modification of the
above picture of Coulomb blockade. In particular, for an odd number of
electrons on the dot, the quantum dot can have a net spin 1/2, and a Kondo
effect can develop. It has been predicted (7, 8), that, at low temperature,
this enhances the conductance in the odd electron valleys, turning them
instead into plateaus of near perfect transmission.
The outline of the paper is as follows. Sec. 2 describes the model and
Sec. 3 the NRC method used to solve it. The results (9, 10, 11) for the
conductance are described in Sec. 4-5 and these are used to compare with
experiment over a range of gate voltages and temperatures in Sec. 6. Ex-
perimental investigation of the regime T ;S r has only recently become
possible as a result of better control of the electrode geometry, allowing
parameters like r to be tuned to values of 1-3K so that T ;S r is accessible
(1, 2, 3, 4, 5). Sec. 7 describes the effect of a magnetic field on the transport
through a quantum dot. Influencing transport via spin effects is of great
current interest (12). The results presented in Sec. 7 should prove use-
ful for interpreting magnetotransport experiments on strongly interacting
quantum dots.
2. Model
(2)
3. Method
0.2 0.2
a b
~'.:E:J
0.2
-Pc
----- P, 0.15
-Pc
~ 0.15 0.15 -----
0 P,
II
::r: 0.1
TK £
0"
~- 0.7
0" II ' 0.5
II 0.1 f-< 0.1 0 4 8
f-<
i
0.05 i H
<
<c-
< 0.05 0
-10 10
0.05
0 0
-200 0 200 400 -10 -5 0 5 10 15 20
m/TK m/TK
Figure 1. Comparison of methods for AT (W, 0, H), of the Kondo model, using canonical
(Pc, solid line) (9) and reduced density matrices (Pr, dashed line) (17). TK = 3.97x 10- 5
(in units of half bandwidth D = 1) is the HWHM of the T = H = Kondo resonance: °
(a) at zero field, with the inset showing the Kondo resonance in more detail, and, (b) in
finite field, with the curves from left to right corresponding to H = 0,1,2,5,10 in units
of T K . The inset shows the peak position, Eo (H), as a function of H for the canonical
(circles) and reduced density matrix (squares) approaches. Both Eo and H are measured
in units of T K . The limit (18) limH~o Eo(H)/H = 2/3 is recovered in both approaches.
density matrix of the usual procedure (9). This reduces finite-size effects
in the spectra. The latter are usually small on all energy scales in the case
of zero magnetic field, but can be large in the case of a finite magnetic
field, particularly for the high energy parts of the spectra (since these are
calculated from the shortest chains and are therefore subject to the largest
finite-size effects). Figure: 1a shows that both procedures give equally ac-
curate results for the Kondo model in zero magnetic field on all energy
scales. The same holds for the Anderson model in zero magnetic field. At
finite magnetic field, the reduced density matrix is required for a correct
description of the features at cd and cd+U in the spin-resolved spectra (17).
In contrast, the low energy Kondo resonance in a magnetic field is well
described by either procedure for H;S 10TK, as shown in figure: 1b for the
Kondo model.
The position, Eo(H), of the spin-resolved Kondo resonance in a weak
magnetic field, H « TK, is known from an exact Fermi liquid result due
to Logan (18) (see also (19) showing the low field quasiparticle resonance).
This states that limH-to Eo(H)/H = 2/3. This indicates that correlations
reduce Eo(H)/ H from its expected high field limit of 1 to 2/3. The inset
to figure: 1b shows that both procedures recover this result. The use of the
reduced density matrix has also proven useful for studying magnetic states
in the Hubbard model within the dynamical mean field theory (20).
251
1.5
,,",0.8
o
Me.
Q,l
is
>, 0.6
C ------ Fit
G" c;I 0.4
0.5
0.2
Figure 2. (a) G(T, H = 0) versus 11;1 in the regime, T -s: r, of strong quantum fluctu-
ations, for U /r = 4.712. The symbols are the T = 0 limit, Eq. (4). Temperatures decrease
from bottom to top and correspond to TN = 0.64rA -N, N = 0,1,2, ... ,A = 1.5. The
dashed curve has T = 1.2TK, where TK ~ 0.05r, is the HWHM of the T = 0 Kondo
resonance at mid-valley gate voltage. The regions marked 1,2 and 3 and separated
by vertical dashed lines correspond to the Kondo (Ed :s -0.5r, i.e. nd ~ 1), mixed
valent (IEdl :s 0.5r, i.e. nd ~ 0.5, or by particle-hole symmetry, lEd + UI :s 0.5r, i.e.
nd ~ 1.5), and empty (full) orbital regimes (Ed ~ 0.5r, i.e. nd ~ 0, or, by particle-hole
symmetry, Ed + U :s -0.5r, i.e. nd ~ 2). (b) The universal conductance curve G(T)/G(O)
for a quantum dot in the Kondo regime at mid-valley, Vg/U = 0.5, (circles) (9, 10).
The dashed line is the fit formula G(T)/G(O) = (T'l/(T 2 + T'l))' with s = 0.22 and
T~ = T K /V2 1 / 8 -1 used in (1) to interpolate the NRC results up to lOTK (with T K as
in figure: 2a).
(4)
For a comparison of theory with experiment in all regimes, the complete set
of conductance curves for the Anderson model is required. These are shown
in figure: 3a-c. G(T) in the empty orbital and mixed valent regimes has been
calculated in (23) and used to compare with experimental data in (1). Here
we make a parameter free comparison to similar data from Reference (5)
for all three regimes of interest. The results are shown in figure: 3d. In
making the comparisons, we estimated G(T = 0) from G(T = 30mK),
close to the lowest effective electron temperature T = 40mK (5), and used
it to determine nd from Eq:4 and hence the appropriate Ed to use in the
NRG calculation for G(T) at all T. The calculations also used the values
of r = 231.4JLeV and U ~ 0.5meV from the experiments. It is remarkable
that this zero parameter comparison yields the agreement seen. The Kondo
scale automatically comes out correctly as seen for the Vg = -414mV
253
1.4
- Ed~-O.51 - Ed~1.01
Ed~-O.25r Ed~]·51
1.2 2 d
--- Ed~O.O --- Ed~2.01
1.0 -- Ed~O.251
..
Vg = -420mV
N";g -- Ed~O.51 ~ 1.5 77
0.8 N
rE
~ ~
P 0.6
E 1 Vg = -422mV
C5 0.4 '-'
a 0.5
0.2
0.0 0
10-'10-1 10" 10- 1 10" 10-1 10 100 1000
Ttr T/r Ttr T(mK)
Figure 3. (a-c) G(T) for gate voltages Vg = -Eede in, (a), the Kondo (K) regime, (b),
the mixed valent (MV) regime (Ed = -0.5r, .. , +0.5r), and, (c), the empty orbital (EO)
regime (Ed = +r, .. , +2.0r), and parameters as in figure: 2a. (d) Comparison of theoretical
(lines) to experimental (symbols) conductance curves, G(T), for the quantum dot in (5)
with U = 0.5meV, r = 0.231meV and 13mK :s; T :s; 900mK. The values used for Ed in the
calculations were Ed = -0.85r (Vg = -414mV, K), Ed = -0.25r (Vg = -420mV, MV),
Ed = +0.375r (Vg = -422mV, MV), Ed = +1.00r (Vg = -424mV, EO), Ed = +1.50r
(Vg = -426mV, EO). We find a linear dependence Vg/e = -QEd for Vg :s; -420mV,
which is a consistency check for determining Ed from G(T -+ 0) using Eq. 4.
a
b
1.5
0.8
::c: 53'
N-
~
t t:: 0.6
I.-'
r5 0.4
0.5
0.2
8. Conclusions
Acknowledgements
I would like to thank the authors of Reference (5) for sending me the
experimental data used in figure: 3b and W. van der Wiel for useful discus-
sions. Financial support from the Deutscheforschungsgemeinschaft through
SFB 195, and in part by the National Science Foundation under Grant No.
PHY99-07949 during the writing of this paper at the KITP of UCSB, is
acknowledged.
References
13. A. C. Hewson, The Kondo Problem to Heavy Fermions (Cambridge University Press,
Cambridge, England 1993)
14. K. G. Wilson, Rev. Mod. Phys.47, 773 (1975); H. R. Krishna-murthy, J. W. Wilkins
and K. G. Wilson, Phys. Rev. B21, 100:3 (1980).
15. T. A. Costi, in Density Matrix Renormalization, edited by 1. Peschel, X. Wang, M.
Kaulke and K. Hallberg (Springer, Berlin, Germany 1999).
16. R. Bulla, A. C. Hewson and Th. Pruschke, J. Phys. Condo Matt. 10, 8365 (1998).
17. W. Hofstetter, Phys. Rev. Lett. 85, 1508 (2000).
18. D. E. Logan and N. L. Dickens, J. Phys.: Condens. Matter 13, 9713 (2001).
19. A. C. Hewson, J. Phys.: Condens. Matter, 13, 10011 (2001).
20. R. Zitzler, Th. Pruschke and R. Bulla, European Phys. Journal, B27, 47:3 (2002).
21. J. Nygard, D. H. Cobden and P. E. Lindelof, Nature 408, 342 (2000).
22. D. C. Langreth, Phys. Rev. 150, 516 (1966).
23. J. Konig and H. Schoeller, Phys. Rev. Lett. 84, 3686 (2000).
24. P. Recher, E. V. Sukhorukov and D. Loss, Phys. Rev. Lett. 85, 1962 (2000).
25. J. Nygard, private communication.
26. A. Rosch, J. Paaske, J. Kroha, P. Wolfle, cond-mat/0202404.
PROPERTIES OF CORRELATED NANOSCOPIC SYSTEMS
FROM THE COMBINED EXACT DIAGONALIZATION - AB
INITIO METHOD
Abstract. We review briefly our recent exact results concerning the electronic properties
of atomic, molecular, and nanoscopic systems by combining the exact diagonalization and
ab initio methods (EDABI). Particular emphasis is put on determining the microscopic
parameters. The multiatom system behavior is discussed as a function of interatomic
distance. A transformation from a nanometal (for small lattice constant) to an atomic
system is observed with increasing distance between the atoms.
1. Introduction
257
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 257–268.
© 2003 Kluwer Academic Publishers
258
2. The Method
(2)
(3)
259
and
_ H >-
Ee =< _~' " tij < aicyajCY
t > +-1 '~
" t ajCY2alCY2akCYl
Vijkl < a iCY1 t >,
ijCY 2 ij klCY I CY 2
(5)
where the averaging < ... > is for the ground IWe > or excited states;
we also take into account all accessible occupancies of the states {Iicy >},
{Ijcy' >}, etc. for fixed values of tij and Vijkl regarded as parameters (if
we consider grand canonical ensemble, we diagonalize H - fLN, with N =
Jd3 xW~t (x)w(x)
~
= ~ijCY Jd3 XW i* (X)Wj(X) < aicyajCY
t
> ).
To close the solution we treat the expression (5) as a functional {Wi (X)}.
In the most general case of the grand canonical ensemble and with a
nonorthogonal basis {Wi (X)} this functional can be written as
In effect, the Lagrange-Euler equation, which plays the role of the stationary
self-adjusted wave equation (SWE) for Wi(X), takes the form
(7)
where Aij and fL play the role of Lagrange multipliers (if we use explicitly
orthogonal basis and particle-conserving diagonalization procedure in the
Fock space, then Aij = fL == 0 and Eq.(7) reduces to the usual Euler
equation). Also, when the class of {Wi(X)} is selected variationally (as
in the following), Eq.(6) is then minimized with respect to trial function
parameters (orbital radii ai in the case when the orthogonal basis Wi (x) is
composed of superposed atomic-like wave functions).
260
With the help of the field operator \IJ(x) we can define N-particle wave
function Wa (X1' .. ,XN) for particles distributed among M states as follows
(8)
(9)
where Cj1 ... jN are the expansion coefficients (note that: ]1 == ]10'1). Within
our method we determine C j1 ... jN with the help of Lanczos- or other exact-
diagonalization method (4). Substitution (9) to (8) and using the decom-
position of \IJ(Xi) in terms of {Wi (X)} we obtain that
J
2 2
3 * 1t 2 2e
ta = d r<I> (r)[--\7 - -]<I>(r) (11)
2m r
261
H 1 2 2 0 -1
U = J d3rd3r/I<I>(r)12Ir
2
~ r/I I<I>(r/)1 2 (12)
(13)
This means <I>(r) is determined from the Hartree equation. Assuming that
<I>(r) has Is-like form <I>(r) = (7fa 3)1/2 exp( -ar), and adjusting variationally
the size a = a-I of the renormalized Is orbit we obtain the well known
results (5): Ec = -5.695Ry and a = (16/27)ao, where ao is the Bohr
radius. We can systematically improve on this result by enriching our single-
particle basis. Including Is- and 2s-wave functions, we obtain 6-dimensional
Fock space and by repeating both the diagonalization procedure and the
variational optimization of Ec with respect to Is- and 2s-like state radii
we obtain an improved value of Ec = -5.7549Ry. Finally, taking Is-, 2s-,
and 2p-like states (with the corresponding adjustable radii a1s, a2s, a2p) we
obtain the value Ec = -5.794Ry, which is close to the accepted "exact"
value -5.S074Ry. The values of radii are displayed in Table I for the light
atoms and ions. The ground-state 2-particle wave function has in the Fock
space the form for He atom:
IG >':::' {O.Sa tIs ratls-).I - O.4(a tIs rat2s-).I + at2s ratls-).I) + 0.135a 2s
t rat I
2s-).
262
TABLE I!. Microscopic parameters (in Ry) of the selected atoms and ions
all quantities are calculated for the orthogonalized atomic states.
t U1 U2 U3 Up K 12 K 13 K 23
H- 0.057 1.333 0.369 0.77 0.728 0.519 0.878 0.457
(14)
Thus He atom is indeed the smallest atom in the Universe, and we have
a substantial (0.32) probability of having the two electrons in the singlet
Is-2s configuration. Nota bene, we can determine the first excited states
in this case: the Is-2s triplet with energy E t = - 2.3707 Ry, and the next
singlet of primary Is-2s character with E 2s = -1.3737Ry.
Important are the values of the microscopic parameters, which are
displayed for those systems in Table II. They comprise the Is-2s overlap
integral 5, the Is-2s hopping integral t, the magnitudes of Is (Ud, 2s (U2 ),
2p (U3 for m = 0 and Up for m = ±1), and 2p (Up) Coulomb interaction.
Additionally, K 12 , K 13 , K 23 represent respectively Is-2s, Is-2p, and 2s-2p
Coulomb interactions.
,
\ , '.... AH,- states
\
\, ---->-'--"---'---------------------
~
L........J -2.0
§ Q)
Jj
H 2 ground state
Distance, R/ao
Figure 1. Energies of the ground state for H 2 molecule (solid line), as well as the two
lowest laying states for H:; ions, all quantities as a function of interatomic distance R/ao
(au - Bohr radius).
amplitudes which are listed in Table III. Those values will be compared
with those for the chains containing up to N = 14 atoms to draw some
conclusions concerning a similar behavior of both atomic and nanoscopic
systems.
4. N anochains
\
\
R;j N=8 6 I
I
Figure 2. Schematic representation of the finite chain with periodic boundary conditions
(or planar cluster) used in the calculations.
t2 U Ka Vi V2
I R/ao I
[mRy] [Ry] [Ry] [mRy] [mRy]
3- and 4-site terms in the Gaussian STO-3G basis). The method and the
principal results for the case of Slater orbitals have been published before
[2]. Here we present also the first results obtained for the Gaussian (STO-
3G) basis, which we will try to apply for realistic systems containing Li and
Na atoms in the near future. In Fig.2 we present schematically a ring of
N atoms. We analyze its basic properties in a linear-chain configuration
with periodic boundary conditions (i.e. overlaps, hoppings, interactions
only along the ring), whereas in the next Section we discuss a multidi-
mensional (d = 2,3) configurations. By adopting EDABI to this situation
we can calculate all hopping integrals and therefore, calculate the band
energy Ek, as it evolves as a function of interatomic distance R/ao (2). For
R/ao :2: 4 the tight-binding approximation works very well. The bandwidth
W = 2z Lj(i) Itij I falls with growing R systematically reflecting reduced
overlap. This illustrates the famous Mott remark (6) that by looking at
the single-particle (band) energy one can never observe a transition from
atomic to metallic behavior, only a systematic evolution towards the atomic
265
TABLE V. Microscopic parameters (in Ry) vs. R/ao for N = 4 planar cluster.
The star marks the distance R with the lowest system energy.
N = 6+
N = 8- N = 6-- 1
1. N = 1()D R/ct =~2 5
N = 12 8
N = 14 6
o.
O. O---"=='---~~~~~~----":C= __---l
-1 -0.5 0 0.5
MOMENTUM I kR/n
Figure 3. Momentum distribution nkO" for electrons in linear chain of N = 6-;-12 atoms;
the interatomic distance is specified in units of ao. The solid line represents the parabolic
fit, which is of the same for both k < k F and k > k F . The discontinuity at kFR/Tr = 0.5
is also marked (STO-3G basis is taken).
:>
~
-12.0
-12,5
.
-130
,
~
~
-.
~- N
N~
N-
-
---+-- N =
4, Telrahedron
3, Triangle
4. Square
2, H,
~
z
J' -13,5
,;;
.... _..
11. --------~fiit..··-·...········1II
~
c:
<D
-14,0
. ".:."
~A'
l~
\1 _-:1I.1t.
'"
§ -14,5 i\/It
1
'"
'C \• ;11i"
'\ ~
R •• 2.2 ~-].()jtE"N--]4.40<V I
§ -15.0 I
e , .,'1 ~;j!J.,=2.l u= 1.l2E.:.N=·15.0l cV
CJ
; :
-15,5 1 ].9 OFI.]4 g,.iN: -15.11 ,v
~~.=
Figure 4. Total ground state energy (per atom, including the ionic contribution) of
several cluster systems vs. R/ao. The optimal values of the parameters for the minimal
energy are provided in the insets.
5. Nanoclusters
-0,25
III subband
>:
i.
0::
-0,50
W
ul
<lJ
'OJ
Qj
a3 -0,75
.$
.s
<J)
c
<lJ
Ol
Qj I subband
-1,00 1--\------~~==:"'------'--=='_==_1
E n=16
<lJ
1ii
>-
(f)
Figure 5. Energies of the excited states for the square configuration of H 4 cluster. The
states group into the Hubbard subbands with increasing R.
6. Concluding remarks
Acknowledgements
The work was supported by the Committee for Scientific Research (KBN) of
Poland, Grant No. 2P03B 050 23. The discussions with Profs. J.E. Hirsch
(UCLA) and G. Czycholl (Bremen) at the NATO workshop in Hvar are
appreciated.
References
1. J. Spalek, R. Podsiadly, W. W6jcik, and A. Rycerz, Phys. Rev. B 61, 15676 (2000).
2. A. Rycerz and J. Spalek, Phys. Rev. B63, 073101 (2001); ·ibid. B 65, 035110 (2002).
3. For a brief overview see: J. Spalek et al., Acta Phys. Polonica B 32, 3189 (2001);
ib·id. B31, 2879 (2000).
4. Within the MCI method one determines the Ci., ... i.M coefficients variationally, see
e.g. R.A. Shavitt, in H. Schaeffer (editor), Methods of Electronic StmctuTe Theor'y
(Plenum Press, New York, 1977).
5. See e.g. H.A. Bethe and E.E. Salpeter, Quantum Mechanics of One- and Two-
Electron Atoms (Academic Press, New York, 1957), pp.154-156.
6. N.F. Mott, Metal-InsulatoT Tmnsitions (Taylor & Francis, London, 1990).
7. J. Spalek and A. Rycerz, Phys. Rev. B 64,161105 (2001)(R); d. also: A. Rycerz, J.
Spalek, R. Podsiadly, and W. W6jcik, in Lect'uTes on the Physics of Highly Con'elated
Electron Systems VI: Sixth Training CouTse, edited by F. Mancini, AlP Conf. Proc.
No. 629 (2002).
ON THE MULTICHANNEL-CHANNEL ANDERSON IMPURITY
MODEL OF URANIUM COMPOUNDS
Abstract. In this talk we will present the solution of the two-channel Anderson impurity
model, proposed in the context of the heavy fermion compound UBe13, and discuss
briefly the more general multi-channel case. We will show results for the thermodynamics
in the full range of temperature and fields and make the connection with the current
experimental situation.
Key words: Heavy Fermions, Kondo Model, Anderson (Impurity) Model, Multi-Channel
Models, Non-Fermi Liquid, Uranium Compounds, UBe13, Bethe Ansatz, Integrability, In-
tegrable Models, Mixed Valence, Crystal Field, Thermodynamic Bethe Ansatz, Impurity
Entropy, Impurity Specific Heat
During the last two decades, there has been growing interest in materials
whose desciption falls outside the framework of Landau's Fermi Liquid
Theory. Examples are the high T c superconductors, heavy fermions and
quasi-ID conductors.
Uranium heavy fermion compounds provide very interesting examples,
exhibiting unusual specific heat temperature dependence, I = CV IT rv
in T. This behavior is often related to multichannel Kondo physics, occa-
sioned in the case of uranium by the ground state configuration which is
degenerate and non-magnetic, so that one is led to consider a quadrupolar
Kondo scenario (1, 2).
The starting point is to consider the energies of the uranium ion in
its different states of valence. They fall on a parabola depicted in Fig. 1.
The ground state corresponds to a U4+ ionization state, with a 5f 2 shell-
configuration. Considering the spin-orbit Hund's coupling, the lowest mul-
269
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 269–276.
© 2003 Kluwer Academic Publishers
270
E fl f2 f3 f4 f5
\
\
\
\
\
\
~
/-
Figure 1. Schematics of the energy and lowest angular momentum multiplet for different
states of valence of a uranium ion.
X-field
.",---
J=4
L-S'----;;""""';'--<"
Figure 2. Schematic depiction of the spin-orbit and crystal field splittings for the ground
state valence configuration of uranium in a cubic medium.
---a=t,. Es
E
a=+,- Eq
(8) among others. All these techniques have difficulties accessing the mixed
valence regime.
We showed that this model is integrable and carried out a full analysis
by means of the Bethe Ansatz technique. That allowed us to obtain exact
results characterizing the ground state of the model as a function of its
parameters and very accurate numerical results for the thermodynamics of
the model for all temperatures and fields, as well as all valence regimes,
including the mixed valence. We stress that is in the mixed valence re-
gime where the Bethe Ansatz solution is most crucial, since all the other
techniques have difficulties accessing it.
2
h ~ 0
q
£-J.l= 0
£-J.l= ±2'"
1.5 £-J.l= ±4'"
£-J.l= ±6'"
£-J.l= ±s'"
0.5 r----~=--~~c:::.:...--~-.=-.~--------
T/'"
with 1 < a < 4. For large Ie - p,1 the two peaks are clearly separated and
the areal under the Kondo peak is kB in V2 while that under the Schottky
peak is kB in 2.
E-fl = 0
E-fl = ±2'"
1.5 E-fl = ±4'"
E-fl = ±6'"
E-fl = ±S'"
E-fl<O
T/'"
-"-"-"-"-"-"-"-"-"-"-'" E-f.l= 0
-'-'-'-'-'-'-'-'-'-'-'-'-.-., " E-f.l=±2~
" "
0.8 " '. E-f.l=±M
.......................................
" ., E-f.l=±lO~
".
,", . E-f.l=±20~
.:. ---------------------------,-~~ft"*~,:·j3;-~
0.6 ',',
/ .
/' ,/
....................................... /' /
0.2 /
./
./
_.. _.. _.. _.. _.. _.. _.. _.. _.. _ _.. '
-'-'-'-'-'-'-'-'-'-'-'-'-'-'~
..
/
ot::::::r====:::r=====r=====r:::::::=-L-_----.L__U
10- 3 10- 2 10- 1 10° 10] 10 2 10 3
T/13.
Figure 6. Charge content of the impurity site as a function of temperature for different
values of E - p,. At high temperatures all curves start from n c = 1/2 and flow, as the
temperature is lowered to specific values n~) (E - p,).
if the n-plet is slightly split off the doublet, the area is given by kB In[n~4]
(13). Also the scale of energy involved is then determined by crystal field
splitting and corresponds to the scale observed. The split off non-magnetic
triplet (i.e. n = 3 in the p configuration seems, indeed, to be the most
reasonable scenario for U Be13. We are currently working in this direction.
Acknowledgements
References
0.6
£-J.l= 0
.,I. E-J.l=±2li
0.25
. ,
0.4 I.
£-J.l=±M
I . £-J.l=±lOli
l /Ii."i. ' 0.2
r. ...- 0.2
" .:\. i
.i\....·················.. .· ,. :' \:. i
I . I ..
•••.,l I . . \. 1
1---"=...LU-LllL-----'---'--'-'-'-'-ilL----'---'--'-~~---'--'0
I : I I·.. i 0.15
I . I I·
I :. I:.i i
I : . I: i
I : I I:. 0.1
I : . \: 1
I . I ,..
I .: i i· \
,. .......
... .···t·· i \\ 0.05
/
... .
.
I
I i \.'.,,,
:-: / I \-
.... .....; t'-.-._ ..... . ~
4
0
10
Fig1tre 7. Main plot: impurity contribution to the specific heat as a function of tem-
perature for different values of E - IL. As this parameter approaches zero the Kondo
contribution (left) and the Schottky anomaly (right) collapse in a single peak. Inset:
experimental data for the 5f-derived specific heat of UBela.
B. COQBLIN
Laboratoire de Physique des Solides, Bat. 510, Universite Paris-
Sud, 91405-0rsay, France
Key words: Kondo Effect, Heavy Fermions, Rare-Earth, Uranium, Kondo Lattice, Un-
derscreened Kondo Effect, Spin Glass
1. Introduction
In the series of rare-earth metals, most of them are magnetic with a mag-
netic moment corresponding to the 4fn (with integer n values) configuration
and a valence of 3 corresponding to three electrons in the conduction band;
these metals are called "normal rare-earth metals" (1). There are three ex-
ceptions : cerium which has a peculiar phase diagram with two face-centered
cubic phases r and a which have different valences (respectively almost 3
and clearly intermediate between 3 and 4)(2), europium and ytterbium
which are divalent at normal pressure. On the other hand, many systems
containing rare-earths have been extensively studied and, in addition to
cerium, ytterbium and europium, three other rare-earths have been found
to have "anomalous" behavior in alloys and compounds, namely Pr, 8m
277
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 277–286.
© 2003 Kluwer Academic Publishers
278
and Tm. For example, PrSn3 (3) and TmS (4) present a Kondo effect, while
SmS (5) and TmSe (6) present an intermediate valence behavior. Moreover,
volume-collapse transitions have been observed at extremely high pressures
(7) in praseodymium at 20 CPa (8)and gadolinium at 59 CPa (9); thus, we
can imagine that gadolinium changes valence at 600 kbar !
On the other hand, the 5f electrons are less localized at the begin-
ning of the actinide series than the 4f ones in rare-earths and it results
that uranium, neptunium and plutonium metals are not magnetic, while a
strong magnetism occurs for curium and berkelium. Thus, a strong heavy
fermion character or even magnetism can occur in compounds or alloys
with uranium or also in a few known cases with neptunium and plutonium.
It is well known that different anomalous behaviors have been observed
and theoretically studied in mainly ee, Vb, Eu, U but also in other an-
omalous rare-earth systems (10, 11). In the present paper, we will review
briefly the intermediate valence or the Kondo behavior associated with the
heavy fermion character and we will present some recent features of these
two topics, with a special emphasis on the underscreened Kondo effect and
the effect of disorder in ternary cerium or uranium compounds.
is trivalent in the magnetic phase, but the valence change to the non mag-
netic one is very small, of order 0.1 (14). This transition has been studied
extensively under pressure (15) or in related materials (14, 16).
These isostructural valence transitions have been extensively studied
theoretically. The first theoretical explanation starts from the orbitally de-
generate Anderson Hamiltonian and accounts for the magnetic-nonmagnetic
transition accompanied by a large valence change and the existence of the
critical point (17). Then, the Ramirez-Falicov model starts from an ionic
description of the 4f electrons and can also account for the phase diagram
of cerium (18). The main point of these two first theoretical models is
the explanation of a large valence change. Another theoretical explana-
tion of the i-a transition of cerium was based on the Kondo effect and
assumes a large change of the Kondo temperature when going from the
non magnetic phase to the magnetic one (19, 20). More recently, many
calculations involving detailed band calculations and based in particular on
the dynamical mean-field theory (DMFT) were performed to account for
the phase diagram of cerium and to give in particular a better description
of the photoemission experiments (21, 22, 23). Several theoretical models,
including DMFT calculations within the Falicov-Kimball model (24), have
been developed to account for the experimental phase diagrams of Eu or Yb
systems such as for example YbInCu4 compound; but in addition, there
often exists an antiferromagnetic phase in these phase diagrams and at
present there is not a full theoretical model for explaining such a complete
phase diagram.
In actinide metals at the beginning of the series, the 5f electrons are
less localized than the equivalent 4f electrons in rare-earth metals and the
situation is clearly different : there is a strong d-f hybridization and it
results that U, Np and Pu metals are not magnetic, while these actinides
can be magnetic in alloys or compounds where this hybridization is smaller.
Then the 5f electrons become more localized in the middle of the series and
a strong magnetism occurs for curium and berkelium. The phase diagrams
of the first actinides are complicated and in particular plutonium has five
phases at normal pressure (25, 26). The curve of the atomic volumes lies
between the corresponding one for 5d transition metals up to Pu and that of
lanthanides starting at Am, which is a good indication of the rapid increase
of the 5f localization between Pu and Am. On the other hand, there is a
huge volume expansion (26percent) between the low temperature a phase of
plutonium and the high temperature <5 phase, which is of the same order as
the volume expansion occurring at the i-a transition of cerium. The volume
expansion experiments, as well as the analysis of transport and magnetic
properties suggest a strong localization of 5f electrons and a Kondo effect
in <5 - Pu (27, 28). On the other hand, plutonium is close to be magnetic
280
at low temperatures and the spin fluctuation model has been used a long
time ago in the 70's to account fairly well for the resistivity which is large
and passing through a maximum as a function of temperature (29) and for
the thermoelectric power (30) of 0: - Pu.
The long story of the Kondo effect starts 40 years ago with the third-
order calculation of J. Kondo (31) who explains theoretically the resistivity
minimum observed in dilute alloys such as Cu-based alloys with Fe or Mn
impurities and lanthanum-based alloys with Ce impurities. The properties
of many cerium, ytterbium or also uranium compounds or alloys are well
understood in the theoretical framework of the Kondo effect and we will
describe here some features of this difficult problem. Cerium and ytter-
bium systems are well described by the Kondo effect, when the number
of 4f-electrons is close to 1 in the case of cerium or to 13 (or 1 hole) in
the case of ytterbium. First of all, the single-impurity Kondo effect has
been exactly solved(32): at low temperatures, the system has a "Fermi
Liquid" behavior with a T 2 law for the electrical resistivity and very large
values of both the electronic constant of the specific heat and the magnetic
susceptibility, which were at the origin of the name "heavy fermions" given
to these systems (33, 34, 35). At high temperatures with respect to the
Kondo temperature, the magnetic contribution to the electrical resistivity
passes generally through a maximum corresponding to the overall crystal
field splitting and decreases then as Log T (36). The other transport and
magnetic properties have been extensively studied in the low temperature
and high temperature regimes.
The second extensively studied problem is the Kondo-lattice one. In
the case of a lattice, there exists a strong competition between the Kondo
effect and magnetic ordering, arising from the RKKY (Ruderman-Kittel-
Kasuya-Yosida) interaction between rare-earth atoms at different lattice
sites. This situation is well described by the Doniach diagram(37), which
gives the variation of the Neel temperature and of the Kondo temperature
with increasing antiferromagnetic intrasite exchange interaction JK(> 0)
between localized spins and conduction-electron spins. Usual theories of
the one-impurity Kondo effect and of the RKKY interaction yield a Kondo
temperature TKo, that is proportional to exp( -1 / pJK ), and an ordering
temperature (Neel or, in some cases, Curie), T No , proportional to pJ'k, p
being the density of states for the conduction band at the Fermi energy.
Thus, for small pJK values, T No is larger than TKo and the system tends to
order magnetically, with often a reduction of the magnetic moment due to
the Kondo effect; on the contrary, for large pJK , TKo is larger than T No and
281
the system tends to become non magnetic. The actual ordering temperature
TN, therefore, increases initially with increasing pJK, then passes through
a maximum and tends to zero at a critical value pJ'K corresponding to
a "quantum critical point" (QCP) in the Doniach diagram. Such a be-
havior of TN has been experimentally observed with increasing pressure
in many Kondo compounds, such as for example in CeAb (Ref. (38)) or
in CeRh 2Sid39) Thus, we can conclude that the variation of the Neel
temperature predicted by the Doniach diagram is well observed experiment-
ally in many cerium compounds. We also know that the Neel temperature
starts from zero at a given pressure, and increases rapidly with pressure
in YbCu2Si2 (Ref. (40)) or in related ytterbium compounds, in agreement
with the Doniach diagram.
The one-impurity model predicts an exponential increase of the Kondo
temperature with pJK. This means that the Kondo temperature should
increase with increasing pressure in cerium compounds and with decreasing
pressure in ytterbium compounds, which agrees well with many observa-
tions. However, deviations seem to occur in some cerium compounds, such
as CeRh 2Si 2 (Ref. (39)), CeRu2Ge2(41, 42), or Ce2Rh3Ges (43), where
the actual Kondo temperature observed in a lattice can be significantly
different from the one derived for the single-impurity case. Thus, in order
to account for such an effect, we have studied in detail the Kondo-lattice
model within a mean-field approximation and with both intrasite Kondo
exchange and intersite antiferromagnetic exchange, treating successively
the half-filled case (corresponding to a number of conduction electrons
n = 1) (44) and then the general case n < 1 (45, 46, 47, 48) which
gives a much better description of the metallic cerium systems. We have
studied in detail this problem in Ref. (48) and we have shown that the
Kondo temperature tends to decrease with an increasing intersite exchange
interaction and a decreasing number of conduction electrons corresponding
to the so-called "exhaustion" limit (49). This calculation addresses sev-
eral questions : first, it would be very interesting to understand better
the conditions yielding a Kondo temperature for the lattice much different
than the single-impurity one and further experiments are certainly needed.
Second, the discussion on the Kondo-lattice problem addresses again the
difficult issue of the nature of the ground state and screening in the Kondo-
lattice problem. We have shown here that, as the number of conduction
electrons is reduced, exhaustion may be compensated by formation of in-
tersite singlets of localized spins and exact calculations for small clusters
would be interesting (50, 51). The third interesting question concerns the
derivation of a correlation temperature below which short-range magnetic
correlations appear, in good agreement with neutron scattering experiments
in cerium compounds. Finally, it is interesting to notice that taking into
282
4. Conclusion
References
1. B. Coqblin, The electronic structure of Tare-earth metals and alloys: the magnetic
heavy Tare-earths, Academic Press, (1977).
2. A. Jayaraman, Phys. Rev., 137, A179 (1965).
3. P. Lethuillier and P. Haen, Phys. Rev. Lett., 35, 1391 (1975).
4. F. Lapierre, P. Haen, B. Coqblin, M. Ribault and F. Holtzberg, Proceedings of LT
16, Los Angeles (Aug. 1981).
5. M.B. Maple and D. Wohlleben, Phys. Rev. Lett., 27, 511 (1971).
6. M. Ribault, J. Flouquet, P. Haen, F. Lapierre, J.M. Mignot and F. Holtzberg, Phys.
Rev. Lett., 45, 1295 (1980).
7. A.K. McMahan, C. Huscroft, RT. Scalettar and E.L. Pollock, Journal of Computer-
Aided Materials Design, 5, 131 (1998).
8. G.S. Smith and J. Akella, J. App!. Phys. 53, 9212 (1982).
9. H. Hua, Y.K. Vohra, J. Akella, S. Weir, R Ahuja and B. Johansson, Rev. High
Press. Sci. Techno!., 4, 233 (1998).
10. B. Coqblin, in Magnetism of metals and alloys, ed. by M. Cyrot, (North-Holland,
1982).
11. A. C. Hewson, The Kondo problem to Heavy Fermions, (Cambridge University
Press, Cambridge, 1992).
12. C.U. Segre, M. Croft, J.A. Hodges, V. Murgai, L.C. Gupta and RD. Parks, Phys.
Rev. Lett., 49, 1947 (1982).
13. H. Wada, M.F. Hundley, R Movshovich and J.D. Thompson, Phys. Rev. B, 59,
1141 (1999).
14. J.L. Sarrao, Physica B, 259-261, 128 (1999) and references therein.
285
53. A.J. Arko, J.J. Joyce, A.B. Andrews, J.D. Thompson, J.L. Smith, D. Mandrus,
M.F. Hundley, A.L. Cornelius, E. Moshopoulou, Z. Fisk, P.C. Canfield and Alois
Menovsky, Phys. Rev.B, 56, R7041 (1997).
54. K. Dmeo, T. Takabatake, H. Ohmoto, T. Pietrus, H. von Lohneyssen, K. Koyama,
S. Hane and T. Goto, Phys. Rev. B, 58, 12095 (1998).
55. H. von Lohneyssen, J. Mag. Mag. Mater.,200, 532 (1999).
56. E. Miranda, V. Dobrosavljevic and G. Kotliar, Phys. Rev. Lett., 78, 290 (1997).
57. G.R Stewart, Rev. Modern Phys., 73, 797 (2001).
58. P. Schlottman and P. D. Sacramento, Adv. in Phys., 42, 641 (1993).
59. G.R Stewart, Z. Fisk, J.O. Willis and J.L. Smith, Phys. Rev. Lett., 52, 679 (1984).
60. C. Geibel et al., Z. Phys. B, 83, 305, 1991 and 84, 1 (1991).
61. Karyn Le Hur and B. Coqblin, Phys. Rev. B, 56, 668 (1997).
62. J. Schoenes, private communication and J. Schoenes, Journal Less-Common Met.,
121, 87 (1986).
63. J. C. Gomez Sal, J. Garcia Soldevilla, J.A. Blanco, J.I. Espeso, J. Rodriguez
Fernandez, F. Luis, F. Bartolome and J. Bartolome, Phys. Rev. B, 56, 11741 (1997).
64. J. Garcia Soldevilla, J.C. Gomez Sal, J.A. Blanco, J.I. Espeso and J. Rodriguez
Fernandez, Phys. Rev. B, 61, 6821 (2000).
65. D. Eom, M. Ishikawa, J. Kitagawa and N. Takeda, J. Phys. Soc. Japan, 67, 2495
(1998).
66. S. Majumdar, E.V. Sampathkumaran, St. Berger, JVI. Della Mea, H. Michor, E.
Bauer, JVI. Brando, J. Hemberger and A. Loidl, Solid State Comm., 121,665 (2002).
67. E.V. Sampathkumaran, to be published in the proceedings of SCES02 Conference,
Acta Physica Polonica.
68. M.B. Maple, M.C. de Andrade, J. Herrmann, Y. Dalichaouch, D.A. Gajewski, C.L.
Seaman, R. Chau, R. Movshovich, M.C. Aronson and R. Osborn, J. Low Temp.
Phys., 99, 223 (1995).
69. R. Chau and M.B. Maple, J. Phys. C: Condens. Matter, 8, 9939 (1996).
70. V.S. Zapf, RP. Dickey, E.J. Freeman, C. Sirvent and M.B. Maple, Phys. Rev. B,
65, 024437 (2002).
71. Alba Theumann, B. Coqblin, S.G. Magalhaes and A.A. Schmidt, Phys. Rev. B, 63,
054409 (2001).
72. S.G. Magalhaes, A.A. Schmidt, Alba Theumann and B. Coqblin, to appear in Eur.
Phys. J. B.
73. Alba Theumann and B. Coqblin, to be published.
74. F. Steglich, J. Aarts, C. D. Bredl, W. Lieke, D. Meschede and W. Franz, Phys.
Rev. Lett., 43, 1892 (1979).
75. F. Steglich, J. Mag. Mag. Mater., 226-230, 1 (2001).
76. S.S. Saxena, P. Agarwal, K. Ahilan, F.M. Grosche, RK.W. Haselwimmer, M.J.
Steiner, E. Pugh, I.R Walker, S.R Julian, P. Monthoux, G.G. Lonzarich, A. Huxley,
I. Sheikin, D. Braithwaite and J. Flouquet, Nature, 406, 587 (2000).
DESCRIBING THE VALENCE-CHANGE TRANSITION BY
THE DMFT SOLUTION OF THE FALICOV-KIMBALL MODEL
v. ZLATIC
Institute of Physics, Bijenicka c. 46, 10 001 Zagreb, Croatia
J. K. FREERICKS
Department of Physics, Georgetown University, Washington,
DC 20057, U.S.A.
287
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 287–296.
© 2003 Kluwer Academic Publishers
288
evidence for long range order below Tv (9). Y-doping reduces Tv until a crit-
ical concentration of 15 % of Y ions is reached, where the high-temperature
phase extends down to T=O K (5, 6, 10). The experimental results for the
resistivity, susceptibility and thermopower (6) of Yb 1- xYx1nCu4 are shown
in Fig. 1.
,-.....
.:Y
0
0 0.8 (0)
n ~0.15
.
'-..../
()
-0 0.6 E . x=O
"- 0.1 o x=0.3
Q x=O :J
E * x=O.5
"'-...
,-..... 0.4 x=0.1 ~0.05 n x=O.7
.
!"'•••••••• . x~0.9
f-
'-..../ * x=0.5 P
>< 0
()
-0
0.2 x=0.9
0 100 200
Q x x=1 Temperature [K]
0
0 100 200 300
Temperature [K]
6 (b)
4
2
o
-2
• x=0.1
-4
.. x=0.3
fJ) -6
* x=0.5
-8
• x=0.9
-10
o 100 200 300
Temperature [K]
Figure 1. Panel (a) shows the resistivity and the magnetic susceptibility of
Yb 1 - x Y xlnCu4 as function of temperature for various concentrations of Y ions (6).
Note, all the "high-temperature" data can be collapsed onto a single universal curve, by
normalizing the susceptibility with respect to an effective Yb-concentration (not shown).
Panel (b) shows the thermopower of Yb 1 - xY xlnCu4 as a function of temperature for
various concentrations of Y ions (6).
quite suddenly at Tv (3), indicating a mixing of the f-states with the con-
duction band. Neither the susceptibility, nor the resistivity (6), nor the Hall
constant (2) show any temperature dependence below Tv, i.e., the system
behaves as a fermi liquid with a characteristic energy scale TFL » Tv. The
magnetic moment of the rare earth ions is quenched in the ground state
by the f-d hybridization but the onset of the high-entropy phase cannot
be explained by the usual Anderson model in which the low- and high-
temperature scales are the same and the spin degeneracy is not expected
to be recovered below TFL. In these valence-change systems, however, the
f-moment is recovered for Tv « TFL.
The high- T phase that sets in at Tv is also anomalous. In doped systems,
the susceptibility data above Tv can be represented by a single universal
curve, provided one scales the data by an effective concentration of magnetic
f-ions, which is smaller than the nominal concentration of f-ions. The
functional form of the magnetic response agrees well with the "single-ion"
crystal field (CF) theory for all values of the field. The Yb ions seem to
be in the stable 3+ configuration with one f-hole and with the magnetic
moment close to the free ion value nv J(J + l)JLB = 4.53{lB (9L = 8/7 is
the Lande factor and J = 7/2 is the angular momentum of the 4f13 hole).
The dynamical susceptibility obtained from neutron scattering data (12) is
typical of isolated local moments, with well resolved CF excitations (13).
The resistivity of Yb 1 - x YxInCu4 alloys exhibits a weak maximum and
the thermopower has a minimum above 100 K but neither quantity shows
much structure at low temperatures, where the susceptibility drops below
the single-ion CF values. The discontinuity of the thermoelectric power at
the valence transition is a trivial consequence of the different thermoelectric
properties of the two phases: the thermopower of the valence-fluctuating
phase has an enhanced slope and grows rapidly up to Tv, where it suddenly
drops to values characteristic of the high-temperature phase. The resistivity
is not changed much by a magnetic field up to 30 T (14). In typical Kondo
systems, on the other hand, one expects a logarithmic behavior on the scale
T /TK and a large negative magnetoresistance. Here, despite the presence
of the well defined local moments, there are no Kondo-like anomalies. The
Hall constant of the x = 0 compound is large and negative in the high-
temperature phase, typical of a semi-metal (2); the optical conductivity (3)
shows a pronounced maximum of the optical spectral weight at a charge-
transfer peak near 1 eV and a strongly suppressed Drude peak.
The hydrostatic pressure and the magnetic field give rise, like the tem-
perature and the doping, to strong and often surprising effects. The critical
temperature decreases with pressure (14) but the data cannot be explained
with the Kondo volume collapse model (4). We mention also that doping
the Yb sites with Lu3+ ions (5) reduces Tv despite the fact that Lu has a
290
(4)
and
(5)
291
Spin-l/2 conduction holes are created or destroyed at site i by dJ<7 or di<7' the
8-fold degenerate localized f-holes are created or destroyed at site i by f itrt or
firt, and the spinless Y-hole is created or destroyed at site i by cJ or Ci. We
use (J" and TJ labels to denote the angular momentum state of the d- and f-
holes, respectively. The d-, f- and Y-number operators at each site are n~ =
2:<7 nd<7' nj = 2: rt njrt' and ny, respectively, and we have the local constraint
nj +nir ::; 1. The Y-doping reduces the number of f-holes in the conduction
band and provides additional Coulomb scattering for conduction-holes. For
a given concentration x of Y ions, the chemical potential JL is employed to
conserve the total number of remaining d- and f-particles, nd(T) + nf(T) =
ntot - x. In the presence of a magnetic field the magnetic degeneracy of the
f-holes is lifted and the Hamiltonian (1) is supplemented by a Zeeman term.
Using the basis that diagonalizes simultaneously the zero-field Hamiltonian
and the J;/2 component of the angular momentum operator, we can write,
(7)
where the statistical sum runs over all possible quantum states of the system
and depends on ),<7 ( T, T/) for T, T/ E (0, (3). The Hamiltonian of the FK atom,
defines the time evolution of the operators, and the external field defines
the time-evolution operator for the state vectors,
(9)
In the presence of the magnetic field, we add to (8) a Zeeman term that is
obtained from (6) in an obvious way. The Hilbert space can be decomposed
into invariant subspaces with respect to nf and n c , and the matrix elements
in (7) can be calculated within the n f-invariant subspace by replacing
L;TJ fJfTJ in Hat by its eigenvalue (0 or 1) and setting n c = O. Within the
nc-invariant subspace we use nf = 0 and n c = 1. This gives,
(7/2)
where, Zf = L;TJ e-(3(E'1 -/1) and Zy = e-(3(EY-/1Y) are the partition func-
tions of the Yb 3 + and y3+ holes decoupled from the d-states, and ZO(/f-, A)
is the partition function of the Uf = Uy = 0 atomic d-state coupled to the
A-field only. We have,
(11)
where
(12)
and
(13)
(14)
and can be calculated using Eqs.(lO) and (14). In the Matsubara repres-
entation, where GOn = -6InZo/6A~, we obtain
(18)
The DMFT solution for the FK lattice is obtained from the atomic solution
for A-field such that
G IJ (iwn )
at
= J
iWn + /L
p(c)
+ CJ/LBgdH- I;IJ(z) - c
de. (19)
The equations (17)-(19), together with the expressions (10) and (15) for the
partition function, can be solved by iteration. One starts from some trial
self energy and finds G~t using Eq.(19). Then, one finds Go
from Eq.(18),
finds Zat(/L, A) using (10) and (15), calculates G~t by functional derivatives,
recalculates I;IJ using (18), and continues until the fixed point is reached.
Once the numbers No and N f are obtained we can iterate (17), (18) and
(19) on the real axis and find the retarded quantities. In what follows, we
use the DMFT to calculate the thermodynamic and transport properties
of the model corrsponding to the Yb 1 - xYxInCu4 alloys.
3. Numerical results
The calculations are performed assuming that doping by y3+ ions removes
holes from the conduction band. In an undoped sample the total number of
conduction holes and the holes on the Yb 3+ ions is nd+nf = 1.5, while for a
concentration x ofY ions we assume nd+nf = 1.5-x. The parameter space
294
0.4
-- x=0.2
0.2 ---- x=0.15
----- x=O.l
---. x=0.05
----. x=O
(a)
005 01 015 02
Temperature [t"]
10
- - x=0.2
_....-.-- x=0.15
x=O.l
x=0.05
---- x=O
(b)
O'-L-------~------------'
o 0.05 0.1
Teperature [t·]
Figure 2. Panel (a) shows localized electron filling for eightfold degenerate doped
Falicov-Kimball model with UJ = Uy = 2t*, EJ = -O.6t*, and various doping levels x.
Panel (b) shows spin susceptibility normalized to the nominal concentration of Yb ions
in a Yb 1 - xY xlnCu4 alloy.
Panel (a) of Fig.2 shows the effect of temperature and doping on the
average concentration of Yb3+ ions. Panel (b) shows the f-electron contri-
bution to the spin susceptibility, which vanishes below Tv, and is Curie-like
for T » Tv. Above the transition temperature, defined by the inflection
point of the susceptibility curves, the concentration of Yb 3 + ions becomes
significant. The effective Curie constant decreases with doping.
Panel (a) of Fig. 3 shows the result for the DC resistivity, Pdc(T),
normalized by Po = 2fLdad - 2 /e 2 Jr, where a is the lattice spacing and d
is the dimensionality of the system (po;::::; 2.3 X 10- 4 Dcm in d = 3 with
295
2 I
• x=O.2
•
• x=O.15
A x=O.1
•
A
l' x=O.05 • l'
• X=O
• •
l'
0:
:;:;
t 'l
• •
• • • • l'
Q. A
•
•
• •
l'
A
A • (a)
•'"
A
l' l'
0
0 001
Temperature [t·]
1.5 ,-------~-----_=_"
,./'
.//-
0.5 //////
-0.5
i=' -15
Cii
- x=o().2
-2.5 x=o().15
----- x=O.1
x=O.05
-35 ---- x=O (b)
o 0.1 0.2
Temperature [t·)
Acknowledgements
References
A. P. MURANI
Institut Laue Langevin, 6 Rue Jules Horowitz, 38042 Grenoble.
,--+
Measurements of excitations to the upper spin-orbit state in Ce show a dramatic shift
at the a transition from ~ 260 meV in the, phase to around ~ 500 meV in the
a phase (10, 14). A similar enhancement of the spin-orbit excitation energy by about ~
50 ± 30 meV is observed in the compound YbInCu4 as it undergoes a similar transition
from a low-T K (~ 2 meV) heavy fermion state to one with a significantly higher T K (~
32 meV) (15). In CePd a with a ground state TK of ~ 55 meV, the spin-orbit excitation
is observed at ~ 370 meV representing an enhancement of ~ 100 meV (16). Also, a
progressive evolution of the spin-orbit excitation energy from ~ 275 meV to ~ 340 meV
is observed in CeIn:J-xSnx as x is increased form 0 to 3 (2). Remarkably, in all these
systems the enhancement of the spin-orbit energy relative to the free-ion value is about
twice the characteristic energy TK of the ground state.
1. Introduction
297
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 297–305.
© 2003 Kluwer Academic Publishers
298
2. Formalism
(1)
Thus, for r /wo <::::1, (generally, r /wo lies within the range 1/2 to 2/3) we
obtain,
(4)
7f JL2
X(O) = - - (5)
23r
hence, T K = ~r. Note that the relation (5) is implicit via the constant
A in the definition of XI/(w), in relation (1). Furthermore it can be easily
shown that relation (1) satisfies the moment sum rule
where JLo = ---.l!:..-, and g = 2. Unlike the susceptibility integral, the integral
91LI3
of S (w), of course, diverges logarithmically. Hence a cut-off needs to be
applied around We rv 10 - 12wo, or rv lOr for a quasi-elastic distribution.
3. Experimental Results
1
2 Celn 3
·c
::J
...m
.D 0
8 1
-
cD
(/)
0
Celn
1.5
Sn
1.5
4>
1.2
~
c 0.8
::J
.ci
l-
co
r
8 0.4 CePd 3
T = 12 K
~
if)
-0.2
-200 0 200 400 600 80
Energy Transfer (meV)
Figure 2. Paramagnetic spectral response from CePd;J at T = 12 K measured with
high energy neutrons. The continuous curve represents a fit to the data using the Kur-
amoto-Muller Hartmann (KMH) spectral form (20) with a characteristic temperature of
~ 55 meV. The inset shows the additional scattering relative to the fitted curve, on an
expanded vertical scale and re-binned on the energy scale. It represents the spin-orbit
excitation at ~ 370 meV.
shows a fit to the KMH function which is similar to the lorentzian fit over
a very wide energy range. In the region above rv 200 meV one observes
some additional scattering, lying above the fitted curve, which is shown in
the inset on an expanded vertical scale. It corresponds to the spin-orbit
excitation which is centered on rv 370 meV, representing an enhancement
over the free ion value of rv 100 meV.
,,
,
,
,
1:-
'iii
c
4
~ ? o 2 4 6 8 10
Q)
-+-'
c
2
.?,
~
o ~rr2dJ~L...L.......L.-L...J.........L~~~bJ
-50 o 50 100 150
Energy Transfer (meV)
Figure 3. Paramagnetic spectral response from YblnCu4 at 35 K. The continuous
and the dashed curves, which are practically indistinguishable, represent fits to the data
above 10 meV to a lorentzian distribution centered on ~ 32 meV and the KMH spectral
function. The inset shows the variation of the intensity around the peak as a function of
Q. The thick solid line represents a fit to the Yb H form factor.
~
c
600 400
::J
....
.ri
~ 400 200
.c
·m
c
C
Q)
200 o
at ISIS (10). The well defined excitation seen around 260 meV in the I
phase is found to have broadened and shifted to rv 500 meV in the 0:
phase, as found earlier. However, the most remarkable aspect of these later
measurements is that, with the Be-filter technique, at energy transfers of rv
500 meV the momentum transfers Q of rv 15 A-I are attained. Hence, the
presence of measurable intensity in the spin-orbit excitations at such large
Q-vectors further testifies to the single-ion character of the 4f state in the
0: phase. If, the enhancement of the excitation energy by about rv 250 meV
can be interpreted as representing the characteristic energy of the upper
2F7/2 spin-orbit state, then T K 7 /2 is roughly rv 1.5 times the characteristic
energy T K 5 / 2 (rv 170 meV) of the 2F5/2 ground state. As mentioned earlier
the data in figure 2 for CePd 3 indicate T K 7 /2 c:::: 2 x T K 5 / 2 . A very similar
ratio for the excited to ground state characteristic energies is observed for
the series of compounds CeIn3-xSnx over the full range 0 ::; x ::; 3.
4. Conclusions
the localised, single ion, character via the Q dependence of the spectral
intensity as well as its single-ion spectral form.
References
1. A. Severing, A.P. Murani, J.D. Thompson, Z. Fisk, C.-K, Loong, Phys. Rev. B 41,
1739 (1990).
2. A.P. Murani, A.D. Taylor, R.Osborn and Z.A.Bowden, Phys. Rev. B48, 10606
(1993).
3. A.P. Murani, J. Phys. C 33 6359, (1983); Phys. Rev. B 28, 2308 (1983).
4. A.P. Murani, Phys. Rev. B 509882, (1994).
5. A.P. Murani and J. Pierre, Physica B 206-207,329 (1995).
6. R.M. Galera, A.P. Murani, J. Pierre and K.R.A. Ziebeck, J. Magn. Magn. Mater.
63-64, 594 (1987).
7. A.P. Murani, A. Severing and W.G. Marshall, Phys. Rev. B 53, 2641 (1996).
8. A.P. Murani and RS. Eccleston, Physica B 230-232, 126 (1997).
9. A.P. Murani and R.S. Eccleston, Phys. Rev. B 53, 48 (1996).
10. A.P. Murani, Z.A. Bowden, A.D. Taylor, R. Osborn and W.G. Marshall, Phys. Rev.
B 48, 13981 (1993).
11. A.P. Murani, Z.A.Bowden, A.D. Taylor and R Osborn, Phil. Mag. B65, 3092 (1992).
12. A.P. Murani, B. Ouladdif and R.S. Eccleston, Physica B 259-261, 1167 (1999).
13. A.P. Murani and RS. Eccleston, Physica B 241-243, 850 (1998).
14. A.P. Murani, J. Reske, A.S. Ivanov and P. Palleau, Phys. Rev. B 65, 094416 (2002).
15. A.P. Murani D. Richard and R Bewley, Physica B 312-313, 346 (2002).
16. A.P. Murani, R.Raphel, Z.A.Bowden and R.S.Eccleston, Phys.Rev. B53, 8188
(1996).
17. H. Lustfeld and A. Bringer, Solid State Comm. 28, 119 (1978).
18. W. Marshall and RD. Lowde, Rep. Prog. Phys. 31, 705 (1968).
19. G. Zwicknagl, V. Zevin and P.Fulde, Physica B 163, 577 (1990).
20. Y. Kuramoto and E. Muller-Hartmann, J. Magn. Magn. Mater. 52, 122 (1985).
21. A.W. Lawson and T.-Y. Tang, Phys. Rev. 76,301 (1949).
22. 1. FeIner et al. Phys. Rev. B 35, 6956 (1987).
23. G. Polatzek and P. Bonville, Z. Phys. B 88, 189 (1992).
305
307
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 307–315.
© 2003 Kluwer Academic Publishers
308
Hamiltonians which give rise to this behavior is the Anderson lattice model
+L L
00
The prime on the sum indicates, that all sites have to be pairwise different
from each other and (up to y) from the initial site. The scattering matrix
T(z) = V 2 Gj(z) is related to the local f-Green's function to be calculated
from a single impurity theory. Setting y = 0 one obtains a self-avoiding loop
(SAL). Corrections to the SAL especially by multiple visits at a site vanish
for large spatial dimensions (9). At finite dimensions the SAL is known from
the XNCA (10). In (11) some of its mathematical properties were studied.
Here we use a self-avoiding walk to derive DMFT and generalizations of it.
Defining a self-energy via
60 y
GOy(z) = T(z)
v
+ ""'"
~ gox(z)T(z)Gxy(z) + ~ ~ox(z)Gxy(z)
v ""'" v
(4)
x x
In order to find the first contribution to ~ox, start from site 0 and go in
a first step to x. But then in going to site y one must exclude the return
to site O. This correlation in the walk generates the self-energy. For, if one
includes the site 0 into the walk from x to y, one needs a subtraction of the
term which contains site O. In this term we have a walk from x to 0 to y.
Correlations on this walk arise, because in the first part the same site may
309
•o
:- -- - - --- - - -- ~
•o : 2
Figure 1. Principle of exclusion and inclusion for GOy (z). If the correlation line is left
off, one obtains I:~lJ and a subtraction with 4 correlation lines. If these are left off, the
first non-local contribution to I:~2J is obtained and another subtraction etc ..
be visited as in the second. It is easier to see this from the graphics in fig. 1.
Anyway, the first contribution to the self-energy I;W(z) = -6xy T(z)G oo (z),
if it is inserted into the eq. 4 which is multiplied from the right by g gives
the self-consistency equation of DMFT
From the matrix equations of standard many-body theory, involving the full
single p~rticle Green's function G, the unperturbed one g, the scattering
matrix T and the self-energy I;
one obtains
(8)
We next cast the DMFT equation (5) into a similar form by multiplying
eq. (5) with 1 + T(z)Goo(z). This yields
g(z) _ -
Goo(z)(l + T(z)Goo(z)) = (1 _ g(z)T(z) ) = goo(z - I;(z)) = G(z) (9)
l+T(z)Goo(z)
310
0.45 r--------,--------r--------.--------,
-- a=2.0
---- a=2.2
--- a=2.4
0.4 --- a=2.6
o
&. 0.35
0.3
0.25 '-----~-----'--~-----'---~-----'---~------'
-0.04 -0.02 o 0.02 0.04
z
Figure 2. Conduction band density of states with eq. (6). T(z) is approximated by
two Lorentzians T(z) = 1V12C_~~;ibJ + z~bib)' Parameters are 1V1 2 = 0.4, Ej = -2,
aj = b = 0.2, bj = 1.
Given the local quantity T(z) one obtains the self-energy ~(z) via
T(z)
~(z) = 1 + T(z)Goo(z)
----------V~---------'conf
(13)
1 - V(Goo(z) + Goo(z)T(z)Goo(z)) + T(z)Goo(z)
~ -~-conf
Using eq. (10) with G(z) ----+ G(z) we obtain the CPA equation (11). In
view of this, we may interprete the CPA as a special case of the DMFT
for disordered systems. This completes the equivalence of both approaches.
conf
We note in passing, that also the non-local scattering matrix T of the
CPA can be obtained in a closed form from eq. (10)
For the numerical results obtained in section 1 we used either model scatter-
ing matrices T(z) (see fig. 2) or the NCA. The latter, however, is restricted
to temperatures above the single impurity Kondo-temperature. Then the
corrections to the DMFT results by ~1~(z) from eq. (6) are only a few
percent. This is why one has to use another single impurity input. Besides
numerical renormalization group methods or density matrix renormaliza-
tion group, NCA-extensions have been set up. The extension used by the
Wolfie group, the CTMA, can be found in (13), with earlier references
therein. Their derivation via auxiliary bosons can be avoided, if one uses
direct perturbation methods reviewed in (4, 5, 6, 7). In the generating
functionals used before in the direct scheme one can also vary with respect
to the conduction electrons (i.e. the lines crossing the circles in fig. 3). This
can be easily be done, using the formalism in the last appendix of (4). The
first diagram for the generating functional in fig. 3 then reads
(17)
The scattering matrix T I can also be used for the self energy ~~a), as shown
in fig. 6 and for the first part of the conduction band self energy.
~~aJ(
,
J
iw n ) = - V 2 dz. e-(3zjdw2dW3Aa(W2)Aa(W3)f(w2)(1 - f(W3) )Po(z)
27rz
Pa(z + iwn)Po(z + iW n - w3)TI (z + W3, iw n ,W2)PO(Z + W2 - W3)Pa (Z + W2)
(20)
The case with two spin sums is more complicated. In the 3rd diagram
of fig. 4 the curved line entering the lowest vertex carries the energy iW n
and the dashed dotted box is replaced by (L a' T 2,a' (z + iW n + W4, WI, W2)) +
313
Figure 3. Generating functional for improvements of the NCA. The diagrams have the
inverse number of vertices as symmetry factor. All lines are dressed (skeleton diagrams)
and the propagators of the corresponding lines are given in the main text.
"uc,a (.ZW )
n = "NCA(.
uc,a ZW n ) + ,,(a)('
uc,a ZW n ) + ,,(b)('
uc,a ZW n ) (21 )
~c
NCA -
,
a - J
dz -(3z PO(z)Pa(z + zw
-.e
21rZ
.
n) (22)
References
0
Figure 6. Scattering matrix T 1 and self-energies '2:.6 ) and '2:.~0).
Abstract. We investigate the Periodic Anderson model in the regime of itinerant fer-
romagnetism. We compare Quantum Monte Carlo (QMC) results with results obtained
using the Gutzwiller approximation (GA). As expected, the energy of the paramagnetic
state is overestimated by the GA in comparison with QMC results; however, the partially
saturated ferromagnetic (FM) state energies obtained by both methods are surprisingly
similar. While the GA gives the ferromagnetic instability too easily, its results for the
FM ground state, once its existence is confirmed by more dependable methods, can be
used for further examination. This thesis is confirmed by a direct comparison of the mean
values of the occupation numbers nk obtained using both methods.
Introduction
317
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 317–326.
© 2003 Kluwer Academic Publishers
318
Model
The PAM was originally introduced to explain the properties of the the
rare-earth and actinide metallic compounds including the so called heavy
fermion compounds. The basic ingredients of this model are a narrow and
correlated f band hybridized with a dispersive and uncorrelated d band.
The Hamiltonian associated with this model is:
where flu and dbu create an electron with spin (J' in f and d orbitals at lattice
site rand ntu = flu fru. The hopping t is limited to the nearest-neighbor
sites. We will work in units where t = 1.
Gutzwiller method
In the spin-up Fermi sea IFS r }, the d-band is filled up to the Fermi level
EFT, and similarly for down-spins, leaving all f-states empty. The number
of independent variational parameters aa(k) equals the number of k-states
within the Fermi volume.
As in the single-band Gutzwiller method we seek the term dominating
the normalization factor ('If!I'lf!) to set as a replacement for the sum over all
possible configurations. After the average treatment of the determinants
is carried out, Ref.(5), the combinatorics brings us to the self-consistency
condition
(3)
where
1- nf
qa = (4)
1 - nfa'
with nf = nfa + nf(j· With the help of Eq.3 we obtain
where
1 a;(p)
L(1 - nf)2 ~ (q;;l + a;(p))2'
1 L (cf - cp)a;(p) - 2Va a (p)
L(I-nf)2 p (q;;l+a;(p))2
2a a (k)
(8)
onfa C 1 + Aanfa
(11)
oaa(k) a 1 + Aanfa + Aanfij + A a A a (1- nf)'
onfij _ C Aanfij- Aa
(12)
oaa(k) - a 1 + Aanfa + Aanfij + A a A a (1- nf)·
o(H)
oaa(k)
the correct result, Ref.(5), Efu = Efa- = cf - B u/(1 + Au). The Eq.5 is
just a formal relation, since the determination of Efu and nfu still needs
to be done. For a general set of model parameters, this can be achieved
numerically using iterative methods.
Quasiparticle density
Contrary to the Hubbard model, Refs.(7, 8), the number of the strongly
correlated f-electrons in the two-band model is not conserved. In order to
control this number in lower and upper band, we calculate the quasiparticle
density as a function of k. Following the steps in Ref.(5) we find the d-
electron distribution
(16)
for k's contained in the FS u ' The f-band occupation number does not obey
a simple relationship like Eq.16. Instead, just as in the Gutzwiller solution
of the Hubbard model Refs. (7,8), correlation causes the f-state occupation
to spread over the entire Brillouin zone, rather than being confined to the
k-values contained in the FS u ,
(1 )n + qua;(k) . k E FS
/ f )-/ft
\n f)- -qu fu qu-1 + au2(k) , u
. (17)
ku - \ ku ku -
{
(l-qu)nfu ikctFSu
The above relation suggests that the f-electron distribution experiences a
jump at the Fermi wave vector. Analogous procedure, Ref.(5), brings us to
the mixing terms,
(18)
which holds only for k's contained in FS u ' Instead of presenting the actual
dispersion of f - and d-electrons in the k-space, we decide to investigate
the quasiparticle density of hybridized bands defined with the following
operators
t
a ku u(k)flu +v(k)dtu'
13L -v(k)flu + u(k)dtu' (19)
with
Ek - cf
u(k) ,
J (Ek - Cf )2 + V2
-V
v(k) , (20)
J (Ek - Cf )2 + V2
322
-3.08
-1.952
-1.954 V=0.05 (}--{)OMC'Uy
~ Q----O OMC, U=8
~ -1.956 GA -3.10
W -1.958
-1.960 8x8
-3.12 t;=-2.5
z a) V=O.l ~
-1965
~
W -1.97 =--=----- w
-------
~
W -3.14
V=0.5
n=9/32s0.28
-1.975
~
~/_----
-2 -3.16
~
W -2.01 - ,~,
-2.02 -3.18
o 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
M M
Figure 1. a) E/N for 1D PAM vs. AI. Other parameters of the system are N = 50,
n = 0.3, Ef = -1.5. Results are presented for various V = 0.05,0.1,0.2 and U = 4,8.
Gutzwiller results are presented with full lines. b) E/N for 2D PAM vs. lVI. Other
parameters are presented in the inset. QMC results were calculated on a 8 x 8 square
lattice, Gutzwiller results were calculated on a 8 x 8 lattice (full line) and 100 x 100
(dashed lines).
J
and Ek = ~ (ck + Cf + (ck - Cf)2 + 4V2 ). We finally arrive at desired
relations for the upper (0:) and lower ((3) band, respectively,
Results
the magnetization the system becomes less correlated which renders the
GA results more accurate. It is less straightforward to understand why
lowering of V also leads to a similar effect. The main effect of lowering V is
diminishing the crossover region in k-space where the d- and f - bands
hybridize. We would also like to stress that even though GA overestimates
the paramagnetic energy, it predicts reasonably well the value of magnet-
ization of the ground state. This is particularly evident at small V = 0.05.
Comparing systems with substantially different U = 4,8 we notice that at
small V = 0.05 we obtain almost identical results (the two plots for U = 4
and 8 nearly overlap) while at larger values of V = 0.1, and 0.2 differences
are visible nevertheless they remain small. This confirms our hypothesis
from Refs.(l, 2) that the relevant energy scale leading to ferromagnetism
is the hybridization gap ,6,. When U » ,6, the system essentially reaches
U ---t 00 limit.
Results from the 2D system show similar trends as the 1D results, see
Fig.l.b. The main difference is that in the contrast to the 1D case, the
system size used to obtain QMC results in 2D is not large enough to reach
the thermodynamic limit. While finite-size scaling has been performed in
our previous work Refs.(l, 2), we only show QMC results for 8 x 8 system
and GA results for 8 x 8 and 100 x 100 systems. The most prominent
finite-size effect in the GA (comparing 8 x 8 and 100 x 100 systems) is a
slight increase of the ground state energy around M rv 0.42. As in the 1D
case, the GA overestimates the paramagnetic energy in comparison to QMC
calculation. We should nevertheless stress that even in the paramagnetic
case the relative difference between GA and QMC energies is no more than
2%.
Next we compare the mean value of the occupation numbers (n~;(3) for
the 1D PAM as obtained using GA, Eqs.21 and QMC results for the para-
magnetic (Fig.2.a.c.) and ferromagnetic solutions (Fig.2.b.d.) as obtained
using V = 0.05 and V = 0.2. GA calculation captures surprisingly well
many details of (n~~,(3) in both the PM and FM cases. Starting with the
lower band (nk(u,d) for both PM solutions, presented in Fig.2.a.c., we first
see a downturn around k rv 0.27r, which is more pronounced in the V = 0.05
case, followed by the plateau-like k-dependence. Even though the position
of the downturn can be easily predicted by the position of the crossover
from the d-like to the f -like nature of the hybridized 0: band, the actual
hight of the plateau for k > 0.27r is set entirely by correlation effect. Even
more surprising is the finding that the GA nearly perfectly predicts the size
of the jump at the Fermi wave vector k = kp = 0.67r and the magnitude of
(nk(u,d) for k > kp .
Good agreement persists in the FM regime, Fig.2.b.d., where upper
band majority (n ku ) and minority (n kd ) spin occupancies are equally well
324
1.5 1.5
• <n~"(k» 1D U=4 M=0.3
10 U=4 • <n~"(k»
N=70 N=70 ~=-15
~=-15 o <n~,(k» <n~IoI>=O.025(4)
n=O.3 n=0.3 V=0.2
...
V=0.2 0 <n~d{k»
1\
------ GA: <n"U{d)(k» 1\ • <n .(k»
:;z - - GA :<n~"I,){k» :;z • <nud(k»
C a) C
v v o <n'.(k»
,I ~
,I
M=O o <n',(k»
r
0.5 I 0.5 ----- GA:<n".(k»
<n~ ".>=0.044(5) b)
--- GA:<n",(kj>
I -- GA:<n~,,(k»
I
""'-~-'
I
I ---- GA:<n' ,(k» ,
1__ -
I
0 0
0 02 0.4 0.6 0.8 0 02 04 0.6 08
kIn) kIn)
15 15
(k):>
e <n~
10 U=4 M=033
10 U=4 • <n":(k):>
N=70 N=70 £,=-1.5
£,=-1.5 o <n~"{k» <n~",>=0.015(8)
n=O.3 n=0.3 V=0.05
V=0.05 0 <n',(k»
------ GA: <n"U{d)(k):> ,
1\
:;z c) :f - - GA :<n~"I.)(k»
1\
:;z d)
'\\ ~ e<n".(k»
• <nl;\.{k}> ---!
"E'
v M=O
"E'
v tI o <n',(k» i
05 <n~",>=0.050(11)
i
I
I
05 I
\
~ ,
o
-----
---
<n',(k»
GA:<n",(k»
GA:<n",(k»
i
i
- GA:<n~.(k»
~
I I
~!I ---- GA:<n~d(k»
0
0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
kIlt) kIlt)
Figure 2. Mean value of the occupation numbers of the non-interacting band states of
the PAM in the ID lattice for the paramagnetic state .M. = 0 a) and c) and for the ground
state with partially saturated ferromagnetic solution b) and d). Legends: (n~u) repres-
ents lower-band occupation number of up-spins, (n~d) represents upper-band occupation
number of down-spins, etc.
1.5 1.5
12x12 V=-05 M=O 12x12 U=10.0 M=0.42
n=0.29 £,=-3 n=0.29
U=10.0 • <n~"(k»
b) V=-05
£,=-3 • <n~ik»
O<n~"(k»
"
:;<
C
a) o <n~d(k»
- - GA:a
"
:;<
C
• <n"u(k»
• <n"d(k»
v v o <n~,(k»
----- GA:13
0.5 0.5 , 0 <n'd(k»
: ' - - GA:o.
,"6 ------ GA:13
0 0
r M X r r M X r
Figure 3. Mean value of the occupation numbers of the non-interacting band states of
the PAM in the 2D 12 x 12 lattice for the a) paramagnetic state M = 0 and b) for the
ground state with partially saturated ferromagnetic solution with !vI = 0.42.
Conclusions
Acknowledgements
References
1. C. D. Batista, J. Bonca and J. E. Gubernatis, Phys. Rev. Lett. 88, 187203 (2002).
2. C. D. Batista, J. Bonca and J. E. Gubernatis, cond-mat/0208604.
3. T. M. Rice, K. Deda, Phys. Rev. Lett., 55, No.9, 995 (1985); Phys. Rev. B 34,
6420 (1986).
4. C.M. Varma, W. Weber, and L.J. Randall, Phys. Rev. B 33, 1015 (1986).
5. P. Fazekas, B. H. Brandow, Physica Scripta, 36, 809 (1987).
6. P. Fazekas and E. Muller-Hartmann, Z. Phys. B - Condensed Matter 85,285 (1991).
7. M. C. Gutzwiller, Phys. Rev., 137, No. 6A, A1726 (1965).
8. D. Vollhardt, Rev. Mod. Phys., 56, No.1, 99 (1984).
9. S. Zhang, J. Carlson and J. E. Gubernatis, Phys. Rev. Lett., 74, 3652 (1995).
10. S. Zhang, J. Carlson and J. E. Gubernatis, Phys. Rev. B 55, 7464 (1997); J. Carlson,
J. E. Gubernatis, G. Ortiz, and Shiwei Zhang, Phys. Rev. B 59, 12788 (1999).
TRANSPORT PROPERTIES OF HEAVY FERMION SYSTEMS
Abstract. Heavy fermion systems (HFS) are described by the periodic Anderson model
(PAM), which is studied within the dynamical mean-field theory (DMFT), mapping the
PAM on an effective single-impurity Anderson model (SIAM) to be determined selfcon-
sistently, which becomes exact in the limit of high spatial dimensions, d ---+ 00. We use the
modified perturbation theory (MPT) as approximation for the effective SIAM. The MPT
is exact up to second order in the Coulomb correlation U and simultaneously reproduces
the exact results for the atomic limit and the lowest moments. Within this approximation
we have calculated the temperature dependence of the resistivity, the thermopower and
the frequency and temperature dependence of the dynamical conductivity. For all these
quantities the typical HFS-behavior is qualitatively well reproduced within our treatment.
1. Introduction
Besides the specific heat (,-coefficient) and the magnetic susceptibility the
unusual transport properties are most characteristic for heavy fermion sys-
tems (HFS). Pure metallic HFS, for example CePd 3 (1), CeCu6 (2), have
a small zero temperature resistivity R(T = 0), a sharp increase of R(T) for
low temperature T, a maximum at some characteristic temperature T*, and
a R(T) decreasing with increasing T, i.e. a negative temperature coefficient
(NTC), for higher temperatures. There are, however, also some HFS (for
instance U Pt3, YbCuAl) for which no maximum but only a flattening and
a plateau behavior of R(T) is observed for higher T.
Another interesting transport quantity is the thermoelectric power (TEP).
In HFS the TEP S(T) may have a sign change at low T (6, 7, 8, 9, 10) and
it can assume giant values; much of the recent interest in the heavy fermion
327
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 327–334.
© 2003 Kluwer Academic Publishers
328
2. Model
It is well accepted that the basic electronic properties of HFS are described
by the periodic Anderson model (PAM), which contains a conduction band,
strongly correlated f-levels localized at the sites of a lattice simulating
the rare-earth or actinide ions, and a hybridization between f-states and
conduction electron states. For real rare earth or actinide atoms or ions
the f-shells are highly degenerate, namely 14-fold degenerate in the bare
(hydrogen like) atom, which is split into a 6- and an 8-fold degenerate
level due to spin-orbit coupling, and this degeneracy is further reduced in
a crystal due to crystalline electric fields so that the lowest f-level may be
only two-fold degenerate. But the crystal-field (CF) split level is thermally
accessible so that taking into account only a two-fold degenerate f-level is
probably not sufficient to account for the full temperature dependence of
the relevant physical quantities.
Nevertheless for the actual calculations presented in this paper the
standard PAM with only a two-fold degeneracy of the f-levels has been
used, i.e. (with nftlT := f~ITfRIT)
quantities one needs also the current operator, and for the current operator
being consistent with the PAM the component in x-direction is given by
8Vk t
8k UkaCka + c.c.) (2)
, x v
~
= 0, if V k-independent
So for k-independent hybridization V one has in site representation
(3)
«
(«
Cka; cta »z «Cka; ita »z ) _ (z - Ek - V ) -1 (7)
Aa; cta »z «Aa; ita »z - - V Z - E f - ~kf(Z)
3. Approximations
We study the PAM within the dynamical mean-field theory (DMFT) (19),
which becomes correct in the limit of infinite dimension d ---t 00, t ---t 0
330
J dE ( - :~) (E - fL)L(E, E)
S = ----'-------'--;----,--------- (11)
eT J dE (--£) L(E,E)
331
G~SIAM(z)
1
z - Ef - ~f(z) - ~(z)
~L 1 (12)
N k
z - E f - ~f (z) - ~
Z-Ek
Here the "bath" Green function ~(z) (i.e. the effective SIAM conduction
band Green function) has to be determined selfconsistently. But the selfen-
ergy for the effective SIAM is not known rigorously and an additional
suitable approximation for the SIAM has to be used. We have applied the
"modified perturbation theory" (MPT), which starts from the following
ansatz (20, 21, 22)
(13)
where ~1°C (z) is the second order contribution in U to the SIAM selfen-
ergy relative to the Hartree-Fock solution and the parameters Ct, j3 can be
determined by the condition that the atomic limit (of vanishing V) and an
additional criterion (Fermi liquid sum rule, reproduction of the first four
moments) are fulfilled. The MPT has the advantage that it is exact up to
order U 2 and the atomic limit is fulfilled simultaneously.
332
10
--0=-'".2=---=0"'3~~0~.4~~0.5·
00:----=-'0."'"1
m OJ
4. Results
~
1
2:-- = 8z·
N k Z - Ek
(1- )1- (4Z)-2) , (14)
which becomes exact for a Bethe lattice in the limit of infinite coordination
number. Thereby we have chosen the unperturbed conduction band width
as our energy (and temperature) unit. The chemical potential JL has to
be determined selfconsistently for a given total number ntot of electrons
per site. Choosing ntot = 1.5 JL does not fall into a possibly existing
hybridization gap, i.e. the model describes a metallic situation. Typical
results are shown in Figs. 1-3. The resistivity R(T) is small for T ----+ 0,
has a rapid increase with increasing T, a maximum at T*, and an NTC for
higher T > T*. With increasing hybridization V T* also increases and for
larger V one can get a behavior with no maximum but a plateau behavior.
Changing U for fixed V only the absolute value of R(T) varies whereas
T* remains fixed. In the thermopower S(T) we observe the large absolute
value (of the magnitude 100JLV / K), a low T extremum and a sign change
at a temperature rv T*. In the frequency dependent (optical) conductivity
O"(w, T) (Fig.3) we observe a Drude peak and an additional (mid infrared)
peak at finite frequency for low T. With increasing T we obtain a crossover
to a behavior with only one rather broad peak around zero frequency.
The position of the mid infrared peak obviously increases with increasing
hybridization V.
Choosing ntot = 2, E f = -0.5, U = 1 we have the "symmetric PAM"
and the chemical potential JL falls into a hybridization gap at zero T; this
333
fO.15
, ~-4
~
'iii , --
, :: ...... --
if)
~ -6
& 0.1 ' .....
~
0--8
o
0.05 E ~ V=0.20
",-10 --+-V = 0.30
..c
f- - - V =0040
-12
0.1 0.2 0.3 0 0.05 0.1
Temperature T Temperature T
Figure 4. T-dependence of R(T) (left panel) and S(T) (right panel) for ntot = 2,
Ef = -0.5, U = 1, V = 0.2
situation may, therefore, model a Kondo insulator. Results for R(T), S(T)
and O'(w, T) are shown in Figs. 4, 5. The hybridization gap (of width 6.)
manifests itself in an activation behavior, i.e. an exponential increase of
R(T) for very low T and in a vanishing of O'(w, T = 0) for w < 6.. With
increasing T the hybridization gap disappears and we get a crossover to a
behavior similar as in metallic HFS for sufficiently high T. The thermopower
is also extremely large, even one magnitude larger than in the metallic HFS.
5. Conclusion
disordered systems, similarly as it has been done for R(T) in (23) using
the SOPT. The treatment can also be applied to a more realistic version
of the PAM with a more realistic (higher) orbital degeneracy of the f-levels
and more realistic assumptions for the hybridization (dispersion); then also
crystal field effects can be studied and included. Furthermore, using a two-
channel Anderson model it should be possible to study also non-Fermi liquid
systems within this approach.
References
Abstract. We develop a new general algorithm for finding a regular tight-binding lattice
Hamiltonian in infinite dimensions for an arbitrary given shape of the density of states
(DOS). The availability of such an algorithm is essential for the investigation of broken-
symmetry phases of interacting electron systems and for the computation of transport
properties within the dynamical mean-field theory (DMFT). The algorithm enables us
to calculate the optical conductivity fully consistently on a regular lattice, e.g., for the
semi-elliptical (Bethe) DOS. We discuss the relevant f-sum rule and present numerical
results obtained using quantum Monte Carlo techniques.
Key words: High Dimensions, Infinite Dimensions, Correlated Electrons, Optical Con-
ductivity, Quantum Monte Carlo, Transport, Density of States, General Dispersion Method,
F-Sum Rule
1. Introduction
where the operators cka and cR;a create and destroy electrons of spin (]"
on site R i , respectively; nRw measures the corresponding occupancy. A
general nonperturbative treatment of this model is only possible in the
limit of infinite dimensionality where the dynamical mean-field theory be-
comes exact: due to a local self-energy the model reduces for d -----+ 00 to a
335
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 335–343.
© 2003 Kluwer Academic Publishers
336
(3)
1
00 (TO
dw (T(w) --(Hal' (4)
A
=
a 4
However, the hc DOS is unbounded which is hardly compatible with the
single-band assumption. In fact, no regular lattice model with sharp band
edges in d ----7 00 could be constructed so far. In this situation, many DMFT
studies have focussed on the so-called Bethe lattice which is not a regular
lattice, but a tree in the sense of graph theory as shown in Fig. la. The
semi-elliptic DOS p( E) = vi 4 - E2 I (27f) of this model (for Z ----7 (0) fixes
the local properties of the model; transport, however, is a priori undefined.
A derivation of (T (w) directly for the Bethe tree (using the level-picture
Fig. la) by Chung and Freericks (4) is still incomplete (5); up to a factor
of 3, the same expression p(E) ex (4 - E2 ) p( E) was obtained (6) in a heuristic
scheme by enforcing the hc f-sum rule (4). An alternative direct approach
(7) fails to describe the coherent transport expected in the metallic regime.
The local DMFT problem is unchanged when a finite number of hopping
bonds per site are added. Therefore, the periodically stacked Bethe lattice
(Fig. 1b) is still a Bethe lattice in the DMFT sense; potentially coherent
337
a) c)
/ ./ /
,/ ,,/ ../
... ...
.. /
"
... "
./'
.. /
"
// ./' //
Using the orthogonality of the Hermite polynomials, one may express the
hopping matrix elements in terms of the transformation function F(x):
tt = ~joo
2Jr D!
dEF(E)HeD(E)e-E2/2.
-00
(8)
(9)
which leads to
(10)
0.3
A
~1
V
0.2
0.5
0.1
Figure 2. a) The average squared Fermi velocity (Ivk 2)(E) is constant for the hypercubic
1
lattice (or for the x-component of a stacked lattice); in contrast it vanishes for the isotropic
lattice defined in this work. For comparison, the form suggested by Millis is also shown.
b) Resulting function jJ( E) of the full isotropic model (solid line) in comparison with
truncated models (D max = 3 or D max = 5), evaluated in finite dimensions 5 :S d :S 100.
The asymptotic exponent -3/4 is only slightly smaller than the threshold
value -1/2 required for a finite variance JC"X)CXJ du 2 p( f) = 2:;D tj}. For a
rectangular model DOS, t'D rv 2- n n- 3 / 4 already decays exponentially fast.
In this section, we will apply the new formalism to the Bethe semi-elliptic
DOS in order to determine a corresponding tight-binding Hamiltonian
defined on the hypercubic lattice with the same local properties as the
Bethe lattice (with NN hopping) in the limit d ---+ 00. From (11), we derive
the average squared Fermi velocity defined in (3) in closed form:
(I Vk
2)() 27f exp [
f = ---2 -2(er f_l(fV1-f2/4+2arCSin(f/2))~2]. (14)
1
4-f 7f
Here, we have used the fact that (IVkI2)(f) is effectively constant (and equals
1 for unit variance and lattice spacing) in the hypercubic case. The result
(solid line in Fig. 2a) has all the qualitative features expected for this
observable in any finite dimension: (IVkI2)(f) is maximal near the band
center, strongly reduced for large (absolute) energies and vanishes at the
band edges: states at a (noninteracting) band edge do not contribute to
transport. The violation of this principle in the stacked case (dashed lines
in Fig. 2a), which corresponds to an application of the hc formalism to
the Bethe DOS with (IVkI2) constant up to the band edges, is clearly
340
U=4.0 - U=4.0 -
U=4.6 ----- U=4.6
U=50 - U=5.0
U=5.5 U=55
0.2 0.2
'8
b
0.1
Figure 3. Numerical results for the half-filled Hubbard model with semi-elliptic DOS for
00 in the paramagnetic phase at T = 0.05. a) Local spectral function A(w) obtained
d ---+
from QMC (using a discretization 6T = 0.1) and MEM. b) Optical conductivity u(w) for
w
the isotropic "redefined Bethe lattice". The inset shows the partial f-sum Jo dw' u(w').
pathological. Therefore, our method has not only the merit of yielding
isotropic transport, but also of avoiding unphysical behavior.
In order to determine the microscopic model, we have to apply (8) to
the numerically evaluated transformation function F. Again, the scaled
hopping matrix elements fall off exponentially fast: only a fraction 10- 3
of the total energy variance arises from hopping amplitudes beyond third
nearest neighbors and only a fraction 10- 6 results from hopping beyond gth_
nearest neighbors. This result suggests that properties of the model should
be robust with respect to truncation. In fact, p( E) (and consequently the
definition of CT( w)) hardly changes when hopping is cut off beyond 3rd or
5th nearest neighbors, even when evaluated in finite dimensions as seen
in Fig. 2b. This behavior is very general so that results for CT(W) of a
local theory in finite dimensions will depend on d predominantly via the
interacting DOS A(w) and only very little via p( E).
The local spectral functions for T = 0.05, i.e., slightly below the critical
temperature T* ~ 0.055, are shown in Fig. 3a as obtained from QMC /MEM
(5). In the metallic phase, the spectral density at the Fermi level (w = 0)
is approximately pinned at the noninteracting value p(O) = II-IT ~ 0.32
for U ;oS 4.4. The quasiparticle weight decreases drastically and a shoulder
develops for U 2: 4.6 before a gap opens for U 2: 4.8. An application of
(2) to these spectra for the isotropic model characterized by (14) yields
the estimates for the optical conductivity CT( w) shown in Fig. 3b. A low-
frequency Drude peak (of Lorentzian form) and a mid-infrared peak at
w ~ U /2 are present in the metallic phase and decay towards the metal-
insulator transition at U ~ 4.7. For large U, the optical spectral weight
341
a) b) 0.5 ,..,------,_.------,------,_-,------,--------,------,
isotropic --e-
U=4.0 disordered - -A--
1.5
stacked ··0··
0.4 """
'-.....
G>·····B.~"'''',
8' 0.3
8'
>....,
. '"0-.>~.,.,
b
b 8
"0
.~ 0.2
-1[/2 Ekin
0.5 isotropic -
disordered - 0.1
stacked -
Figure 4. a) Optical conductivity a(w) for T = 0.05 and U = 4.0 for the new iso-
tropic model with semi-elliptic DOS in comparison with disordered and stacked models
consistent with the same DOS. b) Optical i-sum for T = 0.05.
4. Conclusion
Acknow ledgements
References
1. J. Hubbard, Proc. Roy. Soc. London A276, 238 (1963); M. C. Gutzwiller, Phys.
Rev. Lett. 10, 59 (1963); J. Kanamori, Prog. Theor. Phys. 30, 275 (1963).
2. A. Khurana, Phys. Rev. Lett. 64, 1990 (1990).
3. T. Pruschke, D. L. Cox and M. Jarrell, Phys. Rev. B 47, 3553 (1993).
4. W. Chung and J. K. Freericks, Phys. Rev. B 57, 11955 (1998). J. K. Freericks,
private communication (2000, 2002).
5. N. Blumer, Ph.D. Thesis, Universitiit Augsburg (2002).
343
6. A. Chattopadhyay, A. J. Millis and S. Das Sarma, Phys. Rev. B 61, 10738 (2000);
A. J. Millis, private communication (2002).
7. M. P. H. Stumpf, Ph.D. Thesis, University of Oxford (1999).
8. G. S. Uhrig and R Vlaming, J. Phys. Condo Matter 5, 2561 (1993).
9. V. Dobrosavljevic and G. Kotliar, Phys. Rev. Lett. 71, 3218 (1993).
10. M. Jarrell, J. K. Freericks and T. Pruschke, Phys. Rev. B 51,11704 (1995).
11. M. J. Rozenberg, G. Kotliar, H. Kajiiter, G. A. Thomas, D. H. Rapkine, J. M.
Honig, and P. Metcalf, Phys. Rev. Lett. 75, 105 (1995).
FROM CEIN 3 TO PUCOGA 5 : TRENDS IN HEAVY FERMION
SUPERCONDUCTIVITY
1. Crystal Chemistry
345
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 345–351.
© 2003 Kluwer Academic Publishers
346
Figure 1. Generalized doping phase diagram for CeMIn5. The left-hatched regions
indicate antiferromagnetic order (AFM), while the right-hatched regions indicated super-
conductivity (SC). Note the substantial regions of coexistence of both superconductivity
and antiferromagnetism. After (14, 19).
12 r----r---..,.-----,r------,
1(y~
o
o
8
o
6 o
o
4
0L...-_ _
0'
TN
o
~ ,&l,aMDA I".
.- TI".l".
...I'_-_ _.. . . _ _
c 1".4,
....L._ _......I
o 20 40 60 80
P (kbar)
Figure 2. Generalized pressure phase diagram for CeMIn5. The open(closed) circles
indicate the evolution of antiferromagnetism(superconductivity) for CeIn3. The open
and closed squares have the same meaning for CeRhIn51 with the pressure offset by
13 kbar from the measured data to account for chemical pressure effects. Similarly, the
triangles represent the evolution of superconductivity with pressure in CeCoIn5 including
an overall shift of 28 kbar relative to CeIn3. See (16, 23) for details.
enormous.
The extent to which proximity to a quantum critical point influences the
properties of CeMln5 has also been the topic of much discussion. Non-Fermi
liquid behavior has been observed in both CeColn5 and CelrIn5 (6, 2, 24),
and the extent to which magnetism and superconductivity are proximate
to each other would suggest the the presence of one or more quantum
critical points. Furthermore, it can be argued that CeColn5 is equivalent to
CeRhln5 under 16 kbar of pressure (16, 23). This supposition yields a phase
diagram (see Figure 2) that not only produces a smooth evolution of su-
perconducting transition temperature with pressure, Tc(P), but also yields
a normal state phase diagram quite reminiscent of the high-T c cuprates.
To take an alternative view of these data, one could interpret the gen-
eralized phase diagram for Ce(Rh,Ir,Co)In5 of Figure 1 as revealing a more
typical competition between magnetism and superconductivity. The ob-
served ground states can be rationalized in terms of the deleterious effects
of magnetism on T c convolved with the beneficial effects of increasing aniso-
tropy on T c' Further measurements of the spin fluctuation spectra of these
materials, as well as an understanding of their combined pressure-doping
phase diagram, are needed to clarify this issue.
Finally, the extent to which the f-electrons in CeMln5 are localized or
itinerant remains an issue. Reasonable agreement between Fermi surfaces
349
3. Superconductivity in PuCoGa5
Acknowledgements
References
Key words: Gd Compounds, Electron Correlations, Kondo Effect, Heavy Fermions, Spin
Disorder, Electrical Resistivity, Kondo Lattice, Magnetic Polaron, Heat Capacity, Mag-
netoresistance
1. Introduction
353
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 353–361.
© 2003 Kluwer Academic Publishers
354
09
sz
0
0
C'0
'-"
0.8
3.3
3.0 -
~
if]
'§
--
E
Q.
Q.
0.7
0.6
2.7
..0
@
'-"
2.4 Q.
0.5 '--~--'-~_'--~--'-~----' '--~----L_~----'-_~--'-----'
0.040 50 100 150 0 20 40 60
+
E r,. F/'\\T 20 g
u 0.03 I GdldSi3 K
16
I
~ 02'
,~"
_ 0.02 I-"~ " . .;d) 12 ]
~.,.,
j 0.01
N -......
............-.
(b)
... ....
>.. . . . .
8 0
Q.
.................. 0.8
-:---. .... ...• ".
-..;...:..,.-A•. 4
~ 0.00
~ ;.
" o
20 40 60 0 10 20 30 40 50
Temperature (K)
Figure 1. (a) The electrical resistivity (p) of the alloys, (Gd 1 - xY x,hPdSi3 , normalised to
respective 300 K values, as a function of temperature (T). (b) An estimate of "excess p"
as a function of T for Gd 2 PdSi 3 , obtained as described in the text. (c) The p as a function
of T for Gd 2 PdSi;J in the absence of a magnetic field and in the presence of 50 kOe. (d)
The 4f-contribution to heat-capacity as a function of T for the alloys, (Gd 1 - xY xhPdSi 3 .
one gets an idea about T -dependence of" excess p" (P4f= p(T)-Plattice) from
the derived data (shown in Fig. 1b). Irrespective of the nature of reference
for lattice contribution, the temperature coefficient of P4f is found to be
distinctly negative in the paramagnetic state.
We have also measured P in the presence of an externally applied H.
For x= 0.0, it is found that MR, defined as [p(H)-p(O)]/ p(O), is about -
10% at 40 K in the presence of, say, H= 70 kOe, which is very large for a
normal intermetallic compound in this temperature range (5). Interestingly,
the minimum in p(T) vanishes in the presence of H (say at 50 kOe) and p
exhibits a positive T coefficient in the entire range (Fig. 1b). The magnitude
of MR keeps increasing with decreasing T for a given H. These features are
characteristics of Kondo alloys as well. It is to be noted (5) that a large MR
was not observed even for the analogous Kondo compound, Ce2PdSi3, which
emphasizes the importance of the finding in the Gd compound. Another
noteworthy finding is that the T at which MR starts taking noticeably large
values (with a negative sign) decreases with decreasing Gd concentration
(5), following the trend in TN. This implies that the observed anomalies
are magnetic in origin.
With respect to the C data (Fig. 1d), apart from the peaks attributable
to magnetic ordering at TN, the 4f contribution (C rn ) to C exhibits a tail
extending over a wide T range above TN. For instance, for x= 0.8, for
which TN is lower than 2 K, the tail in Crn(T) extends to T as high as 20
K, with the value of CIT as large as 1 J/mol K 2 at 2 K, comparable to the
behaviour of prototype heavy-fermion system, CeCu2Si2.
It is clear from the above results that the compound, Gd 2PdSh, exhibits
all the characteristics of (antiferromagnetically) ordering Kondo lattice
systems in p, MR and C data, including a resistivity minimum. These
features could be confirmed even in single crystals of this compound (8).
In addition, the thermopower also exhibits a large value at 300 K, typical
of Kondo systems (6). In order to explore whether these anomalies are
due to an artifact of spin-glass freezing setting in around 45 K, we have
probed the frequency-dependence of the ac susceptibility as well as pos-
sible irreversibilities in zero-field-cooled and field-cooled dc susceptibility
behaviour carefully (6, 9). The results obtained, apart from establishing
that the magnetic ordering below 21 K is of a long-range type, rule out
any type of magnetic ordering above 21 K. Therefore, a question arises
whether this compound could be the first Gd-based Kondo-lattice. Our
photoemission data on the other hand revealed (14) that there is no 4f-
derived feature at the Fermi level (Ep) and that 4f-level is placed about
8 eV well below Ep, thereby conclusively establishing that this compound
can not be categorised as a Kondo alloy. We will return to further discussion
on this" magnetic precursor effect" (T-dependent" excess" p and Crn above
357
To) in Section 4.
Before closing this section, it is worth stating that two more isostruc-
tural Gd compounds, viz., Gd 2PtSi 3 (13) and Gd2CuGe3 (12), have been
subsequently identified to exhibit properties resembling those of Gd 2PdSi 3.
This suggests that the anomalies discussed here must be more common
among Gd intermetallics.
4. Discussion
5. Conclusion
Acknowledgements
References
L. DECIORCI
Laboratorium fur Festkorperphysik, ETH Zurich, CH-8093 Zurich,
Switzerland
Abstract. We review selected optical results on correlated systems. This review mainly
focuses the attention on some prototype Kondo materials, characterized by the heavy
electron behaviour, and on the linear chain Bechgaard salts, as representative systems
of "real" one-dimensional materials. We will address a variety of relevant problems and
concepts associated with the physics of an interacting electron gas in three and low
dimensions, as Kondo behaviour, dimensionality crossover, and Luttinger liquid state.
Key words: Optical Properties, Correlated Systems, Heavy Electron Materials, Low
Dimensional Systems, Kondo Problem, Mott-Hubbard
1. Introduction
363
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 363–369.
© 2003 Kluwer Academic Publishers
364
(a) (a)
Figure 1. (left) (a) Real part O"l(W) of the optical conductivity obtained from the
Kramers-Kronig analysis of the absorptivity of CeAI 3 , and directly from the surface im-
pedance measurements (Z,,). (b) An enlargement of the dynamical conductivity showing
the development of the low-frequency resonance at low temperatures (10, 11). (right)
Sketch of the calculated frequency dependent conductivity O"I(w) after Millis and Lee
(12).
Figure 1 (10) shows the real part of the optical conductivity 0"1 (w) of CeAl 3 ,
a prototype HE compound. The general trend of the optical conductivity
is rather common in all HE materials (11). In fact, at high temperatures
we can easily recognize a rather broad metallic-like behaviour (i.e., Drude),
indicative of a scattering rate r rv liT (i.e., the width of the Drude term)
of the order of 0.1 eV. Moreover, the conductivity drops below its dc value
at frequencies of the order of 1 IT and is starting to flatten again above
3000 em-I. In contrast, at relatively low temperatures we observe the
gradual emergence of a narrow mode centered at zero frequency in O"l(W),
which merges into the high temperature 0"1 (w) at FIR frequencies. This
dramatic temperature dependence does, nevertheless, preserve the good
agreement between the O"dc values and the zero energy limit of 0"1 (w).
366
Another major issue and current topic of debate in the field of highly-
correlated electron systems is the fundamental question whether these sys-
tems, in their normal state, may be described as simple Fermi liquids. For
some HE compounds (e.g., U Pt3) this seems indeed to be the appropri-
ate picture. More recent experimental work (11) has indicated, however,
that several HE compounds and related alloys display quite remarkable
properties, which may be much less well related with" conventional" Fermi
liquid behaviour. We refer to our review (11) for a broader experimental
and theoretical perspective on these issues.
Figure 3 displays the temperature dependence of the real part 0"1 (w) of the
optical conductivity in different Bechgaard salts for Ella, the chain direction
(14, 15, 16, 17, 18, 19). The optical conductivity of the TMTTF salts
displays several absorption features mostly ascribed to lattice vibrations
(phonon modes), mainly due to the inter and intramolecular vibrations
of the TMTTF unit. The optical properties of the (TMTSF)2PF6 ana-
logs (Fig. 3b), for which the dc conductivity gives evidence for metallic
behaviour down to low temperatures, are markedly different from those
of a simple metal. A well-defined absorption feature around 25 me V and
a zero-frequency mode (15, 16, 17, 18) are observed at low temperatures.
367
10
'
1000 (TMITF),Br -10K
-··90K
Ella - - 300K
Mott insulator
} 500 ~
~
"
i"j---
o ,,-;-,,-~-"-;,...-..-~'-;:-"-~-"-;,...-..-=-~
10-4 10- 3 10- 2 10- 1
co
Photon Energy (eV)
This latter mode is responsible for the large metallic conductivity. Figure
3 also displays the temperature evolution of O"l(W) (14).
There is an ample theoretical literature, which is hopeless to review here
in great detail (14, 15, 17). As introduction to the theoretical expectation for
the electrodynamic response in one dimension, we briefly sketch the optical
response as suggested by Ref. (20,21), which catches the essential features.
Due to the commensurate filling, a strictly one-dimensional Luttinger liquid
with Umklapp scattering effects transforms into a Mott insulating state,
which is dominated by the charge correlation gap excitation. For a strict
one-dimensional Mott insulator, the charge correlation gap E gap , corres-
ponding to the excitation between the lower and upper Hubbard bands,
appears as a singularity in the real part 0"1 (w) of the optical conductivity
(Fig. 4). If the interchain coupling t.d rv tb) between chains is relevant (i.e.,
t.1. > E gap ), small deviations from commensurate filling due to the warping
of the Fermi surface exist, and should lead to effects equivalent to real
doping on a single chain. Then, it is energetically favorable for the charge
carriers to hop between parallel chains (20, 21). The scenario calculated for
a doped one-dimensional Mott semiconductor consists of a Mott (pseudo)
gap (as reminder of the originallD limit with Umklapp scattering process)
and a zero-energy mode (i.e., theoretically a D!5(w) function at w = 0,
representing the Drude resonance of the effective metallic contribution with
368
scattering rate r = 0) for small doping levels (Fig. 4). The spectral weight D
encountered in the" Drude resonance" is proportional to the effective charge
doping, induced by the interchain coupling t~, in the upper Hubbard band.
Therefore, by increasing t~ there is a dimensionality crossover, which is
accompanied by the evolution from a one-dimensional Mott insulator to a
"doped" semiconductor. Indeed, the (Tl\IITTFhX salts, with X = P F 6 or
BT, are insulators at low temperatures (8) with a substantial (Mott) charge
gap (see arrows for different experimental estimations of E gap in Fig. 3a).
For these compounds the correlation gap E gap is so large that the interchain
hopping (t~) is not relevant. Charge carrier hopping on parallel chains is
here strongly suppressed leading to a truly ID insulating phase (20, 21). In
analogy to the TMTTF salts, the strong FIR excitation of the TMTSF
salts (Fig. 3b) is ascribed, within the scenario depicted in Fig. 4, to the
so-called pseudo charge correlation gap. Additionally, there is the narrow
Drude peak, originating from deviations of commensurability due to t~.
The existence of a gap or pseudogap in the charge excitations (Fig. 3) with
the absence of a gap for spin excitations indicates, moreover, spin-charge
separation in the metallic state (14, 18).
Furthermore, at frequencies greater than t ~, the interchain electron
transfer is irrelevant and calculations based on the ID Hubbard model
should be appropriate (20, 21). The theoretical expectation (Fig. 4) consists
in a powerlaw of the frequency-dependent optical conductivity (71 (w)rvw-'
for frequencies greater than t~ and E gap but less than the on-chain band-
width 4t a . The theory also predicts that the exponent r = 5 - 4n 2 K p , K p
being the so-called Luttinger liquid parameter and n the degree of commen-
surability (20, 21, 22). Our results on the powerlaw behaviour (15, 18, 23)
are very robust and allow us to discriminate among different regimes and
type of correlations. This issue is thoroughly reviewed elsewhere (14).
4. Conclusion
This short review only touched a few issues and problems, which recently
attracted a lot of interest in strongly correlated systems. Signatures of
heavy quasiparticles in Kondo materials and of the Luttinger phenomeno-
logy in quasi-ID systems have been observed. However, the interpretation
of the spectra remains problematic on various issues and await further
experimental and theoretical work.
5. Acknowledgements
References
1. H.R. Ott and Z. Fisk, in Handbook on the Physics and Chemistry of the Actinides,
Eds. A.J. Freeman and G.H. Lander (Elsevier, Amsterdam, 1987), p.85.
2. Z. Fisk, D.W. Hess, C.J. Pethick, D. Pines, J.L. Smith, J.D. Thompson and J.O.
Willis, Science 239, 33 (1988).
3. N.B. Brandt and V.V. Moshchalkov, Adv. Phys. 33, 373 (1984).
4. P. Fulde, J. Phys. F.: Met. Phys. 18, 601 (1988).
5. J. Luttinger, J. Math. Phys. 4, 1154 (1963).
6. S. Tomonaga, Prog. Theor. Phys. 5, 554 (1950).
7. for a review, see H.J. Schulz, Int. J. Mod. Phys. B 5, 57 (1991).
8. D. Jerome and H.J. Schulz, Adv. Phys. 31, 299 (1982).
9. D. Jerome, in Strongly interacting fermions and high Tc superconductivity, Eds. B.
Doucot and J. Zinn-Justin (Elsevier Science, 1995), p. 249.
10. A.M. Awasthi et al., Phys. Rev. B 48, 10692 (1993).
11. for a review, see L. Degiorgi, Rev. Mod. Phys. 71, 687 (1999).
12. A.J. Millis and P. A. Lee, Phys. Rev. B 35, 3394 (1987).
13. F. Wooten, in Optical Properties of Solids, (Academic Press, New York, 1972).
14. for a review, see L. Degiorgi, in Strong interactions in low dimensions, Eds. D.
Baeriswyl and L. Degiorgi (Kluwer, 2003).
15. V. Vescoli et al., Em. Phys. J. B 13, 503 (2000).
16. M. Dressel et al., Phys. Rev. Lett. 77, 398 (1996).
17. A. Schwartz et al., Phys. Rev. B 58, 1261 (1998).
18. V. Vescoli et al., Science 281, 1181 (1998).
19. V. Vescoli et al., Solid State Commun. 111,507 (1999).
20. T. Giamarchi, Physica B 230-232, 975 (1997).
21. S. Biermann et al., cond-mat/0201542.
22. J. Voit, Em. Phys. J. B 5, 505 (1998) and Proceedings of the NATO Advanced
Research Workshop on The Physics and Mathematical Physics of the Hubbard Model,
San Sebastian, October 3-8, 1993, Eds. D. Baeriswyl, D. K. Campbell, J. M. P.
Carmelo, F. Guinea, and E. Louis (Plenum Press, New York, 1995).
23. F. Zwick et al., Solid State Commun. 113,179 (2000).
QUASIPARTICLE UNDRESSING: A NEW ROUTE TO COLLECTIVE
EFFECTS IN SOLIDS
J.E. HIRSCH
Department of Physics, University of California, San Diego
La Jolla, CA 92093-0319
Abstract. The carriers of electric current in a metal are quasiparticles dressed by electron-
electron interactions, which have a larger effective mass m* and a smaller quasiparticle
weight z than non-interacting carriers. If the momentum dependence of the self-energy
can be neglected, the effective mass enhancement and quasiparticle weight of quasi-
particles at the Fermi energy are simply related by z = m/m* (m=bare mass). We
propose that both superconductivity and ferromagnetism in metals are driven by qua-
siparticle 'undressing', i.e., that the correlations between quasiparticles that give rise to
the collective state are associated with an increase in z and a corresponding decrease in
m* of the carriers. Undressing gives rise to lowering of kinetic energy, which provides the
condensation energy for the collective state. In contrast, in conventional descriptions of
superconductivity and ferromagnetism the transitions to these collective states result in
increase in kinetic energy of the carriers and are driven by lowering of potential energy
and exchange energy respectively.
Quasiparticles are 'dressed' bare particles, and they have a smaller quasi-
particle weight and a larger effective mass than bare particles. In several
superconductors and ferromagnets of current interest there is experimental
evidence that quasiparticles 'undress', and resemble more free particles,
when correlations build up and the system orders. Associated with this,
that the kinetic energy, that is supposed to be optimal in the Fermi liquid
normal state, decreases rather than increases in the ordered state. This
behavior is counterintuitive, since in a normal Fermi liquid description it
is expected that quasiparticles should become further dressed and less like
371
A.C. Hewson and V. Zlatić (eds.), Concepts in Electron Correlation, 371–380.
© 2003 Kluwer Academic Publishers
372
free particles when they develop the correlations leading to the collective
state, and that they should pay, rather than gain, kinetic energy. Several
'unconventional' theories have been proposed to explain these phenomena.
Instead we propose here that in fact quasiparticle undressing is a unifying
concept that can describe these collective effect in both new and conven-
tional materials. The only difference is that only in the newer materials
is the 'undressing phenomenology' strong enough that it is easily seen in
experiments.
0'L,
'L,(k,w) = 'L,(k, 0) + wOw (2)
where the first term is the quasiparticle part and the second term is the
incoherent part, and the 'quasiparticle weight' Zk is given by
(4)
then we have simply Zk = Z and Ek = ZEk which implies that the quasi-
particle weight and effective mass m* are simply related by
m
-=Z (6)
m*
hence a highly dressed particle will have a small quasiparticle weight and
a large effective mass. For the models of interest in this paper the 'local
approximation' Eq. (5) is reasonable. Finally, the 'kinetic energy' of the
system is defined by
(7)
373
where the occupation number nk is obtained from the single particle Green's
function
00 1
nk = -00 dwf(w)(-;ImG(k,w)) (8)
/
The low energy effective Hamiltonians in our theory can be derived from
the single band generalized Hubbard model
(11)
374
On the other hand, the quasiparticle weights for the single electron and the
single hole in this single band model are simply Zel = Zhole = 1, hence the
expected relation between quasiparticle weight and effective mass Eq. (6)
is strongly violated.
In addition, the optical conductivity sum rule is violated. The integral
of the optical conductivity in a tight binding model is given by
1 00
(proportionality factors are omitted). For the Hamiltonian Eq. (10), the
average kinetic energy for holes from the ij bond is
< Tkin >ij= -(t~ + nh(~t)ij) < c!a.Cja > + < T kin >ij (17a)
< T kin >ij= -(~t)ij[< cl,-acla >< CjaCi,-a > + < c},_acla >< CjaCj,-a >]
(17b)
and it decreases below T c as the anomalous expectation values in Eq. 17(b)
become nonzero. Hence the integrated optical spectral weight (left side of
Eq. (16)) increases. A similar situation occurs for ferromagnetism. However,
in a real system the total optical spectral weight is conserved (optical sum
rule), hence the optical sum rule is 'violated' if kinetic energy lowering
occurs. The resolution of both this violation and the unphysical relation
between m* and z arises from consideration of other degrees of freedom
not contained in the effective Hamiltonian Eq. (9).
(21 )
dressed _----'::....::...ce...::..-_[>
s. t. undressed
carrIers carrIers
,.. ~ ~ ~
f-
H
undressed
paIrs
n
Figure 1. Undressing phenomenology in the cuprates. Both when hole carriers pair
and when the hole concentration increases by doping, there is an increase in the local
hole density around a given carrier that gives rise to undressing. This is accompanied by
spectral weight transfer from high to low frequencies
This relation implies that the doubly occupied state 1> is a single Slater 1 r
determinant. The very fact that electrons interact makes this an incor-
rect assumption. Hence the conventional Hubbard model fails to describe
the most basic aspect of the electronic correlation problem it purports to
embody, namely correlation of electrons in the same Wannier orbital.
Recognition of the fact that the doubly occupied orbital I 1> is a r
correlated state rather than a single Slater determinant leads to dynamic
Hubbard models (9). The correct form of Eq. (23) is
where 1- iJ >n are excited state of the singly occupied orbital, and 1- iJ >::::::
1 - iJ (24) is
>n=O the ground state. Because electrons interact, S in Eq.
never unity, and the second term on the right side of Eq. (24) is never zero.
This leads to the relation between bare particle and quasiparticle operators
Eq. (18).
We have discussed various realizations of dynamic Hubbard models,
involving either an auxiliary boson degree of freedom at each site or more
than one orbital per site (10). A new energy scale enters, given by the
excitation energies of the states iJ >n. This is the energy range from
1 -
which the high frequency spectral weight gets transfered from. In dynamic
Hubbard models the Hubbard U becomes a dynamical variable, which can
take more than one value depending on the relative state of the two elec-
r
trons in the correlated state I 1>, and destruction of an electron in that
correlated state never yields the singly occupied state liJ > with its full
amplitude. The study of dynamic Hubbard models is only in its beginning
stages but it is clear already that they exhibit very rich physics absent in
the conventional Hubbard model.
We are proposing that there is a single unifying concept behind the two most
common collective effects in metals, superconductivity and ferromagnetism:
quasiparticle undressing. Our proposal rests on four pillars, namely: (1)
378
References
1. M.R. Norman, M. Randeria, B. Janko and J.C. Campuzano, Phys. Rev. B 61,14742
(2000).
2. J. E. Hirsch and F. Marsiglio, Phys. Rev. B 62, 15131 (2000) and references therein.
3. J. E. Hirsch, Phys. Rev. B 43, 705 (1991) and references therein.
4. J.E. Hirsch, Phys.Rev. B48, 3327 (1993).
5. J.E. Hirsch, Phys. Rev. B 62, 14487 (2000), 14498 (2000).
6. J. E. Hirsch and F. Marsiglio, Phys. Rev. B 45, 4807 (1992) .
7. J.E. Hirsch, in Folarons and Bipolarons in high-Tc Superconductors and Related Ma-
terials, ed. by E.K.H. Salje, A.S. Alexandrov and W.Y. Liang (Cambridge University
Press, Cambridge, 1995) p. 234 ; Physica C 201, 347 (1992).
8. J.E. Hirsch, Phys. Rev. B 62, 14131 (2000).
9. J.E. Hirsch, Phys. Rev. Lett. 87, 206402 (2001); Phys.Rev.B 65, 184502 (2002).
10. J.E. Hirsch, Phys.Rev. B 65, 214510 (2002); Phys.Rev. B 66, 064507 (2002); cond-
mat/0207369 (2002).
11. Y Okimoto et aI, Phys.Rev. B57, 9377 (1998).
12. L. Degiorgi et aI, Phys. Rev. Lett. 79, 5134 (1997); Phys.Rev. B65, 121102 (2002).
13. H. Okamura et aI, Phys.Rev. B64, 180409 (2001).
14. E.J. Singley et aI, Phys.Rev.Lett. 89, 097203 (2002) .
15. J.F. Janak and A.R. Williams, Phys.Rev. B14, 4199 (1976).
16. H. Ding et aI, Phys. Rev. Lett. 87, 227001 (2001).
17. H. J. A. Molegraaf et aI, Science 295, 2239 (2002).
18. A.F. Santander-Syro et aI, cond-mat/0111539 (2001).
19. S.Uchida et aI, Phys. Rev. B 43, 7942 (1991).
20. Z.M. Yusof et aI, Phys. Rev. Lett. 88, 167006 (2002).
21. YAndo et aI, Phys. Rev. Lett. 87, 017001 (2001).
22. LK. Kikoin and S.W. Gubar, J.Phys. USSR 3,333 (1940).
23. A.F. Hildebrand and M.M. Saffren, Fmc. 9th Int. Conf. on Low Ternp.Fhys., ed. by
J.G. Daunt et al (Plenum, New York, 1965) p.459.
24. A.A. Verheijen et aI, Nature 345, 418 (1990).
25. LM. Chapnik, Sov.Phys. Dokl. 6, 988 (1962).
381
List of Contributors
A. Aligia [email protected]
N. Andrei [email protected]
N. P. Armitage [email protected]
J. BaLa [email protected]
G. Baskaran [email protected]
C. D. Batista [email protected]
K. Baumgartner [email protected]
D. Bensimon [email protected]
N. Blümer [email protected]
C. J. Bolech [email protected]
J. Bonča [email protected]
R. Bulla [email protected]
M. Capone [email protected]
C. Castellani [email protected]
I. Chaplygin [email protected]
B. Coqblin [email protected]
T. A. Costi [email protected]
C. Czycholl [email protected]
L. Degiorgi [email protected]
T. P. Devereaux [email protected]
S. El Shawish [email protected]
H. Eschrig [email protected]
M. Fabrizio [email protected]
P. Fazekas [email protected]
L.-F. Feiner [email protected]
J. K. Freericks [email protected]
E. M. Görlich [email protected]
C. Grenzebach [email protected]
G. Grüner [email protected]
J. E. Gubernatis [email protected]
E. Helgren [email protected]
A. C. Hewson [email protected]
J. E. Hirsch [email protected]
R. Hlubina [email protected]
A. Jerez [email protected]
H. Keiter [email protected]
A. Kiss [email protected]
K. Koepernik [email protected]
J. Kroha [email protected]
M. Lang [email protected]
.
382
M. Lavagna [email protected]
C.T. Liang [email protected]
A. Lobos [email protected]
H. v. Löhneysen [email protected]
B. Lüthi [email protected]
R. Mallik [email protected]
D. Meyer [email protected]
H. J. A. Molegraaf [email protected]
J. Müller [email protected]
A. P. Murani [email protected]
A. M. Oleś [email protected]
D. Otto [email protected]
J. Paaske [email protected]
R. PodsiadLy [email protected]
P. Prelovšek [email protected]
C. Presura [email protected]
A. Ramšak [email protected]
J. Röhler [email protected]
A. Rosch [email protected]
A. Rycerz [email protected]
E. V. Sampathkumaran [email protected]
I. Santoso [email protected]
J. L. Sarrao [email protected]
T. Sasaki [email protected]
J. Schlueter [email protected]
K.–D. Schotte [email protected]
M. Sigrist [email protected]
J. Spalek [email protected]
F. Steglich [email protected]
J. D. Thompson [email protected]
E. Tosatti [email protected]
D. van der Marel [email protected]
P. G. J. van Dongen [email protected]
M. Vojta [email protected]
W. Wójcik [email protected]
B. Wolf [email protected]
P. Wölfle woelfl[email protected]
R. Zahorbeński [email protected]
S. Zherlitsyn [email protected]
V. Zlatić [email protected]
383
Index