Sunteți pe pagina 1din 5

Mott Transition between a Spin-Liquid Insulator and a

Metal in Three Dimensions


Citation
Podolsky, Daniel et al. Mott Transition between a Spin-Liquid
Insulator and a Metal in Three Dimensions. Physical Review
Letters 102.18 (2009): 186401. (C) 2010 The American Physical
Society.
As Published
http://dx.doi.org/10.1103/PhysRevLett.102.186401
Publisher
American Physical Society
Version
Final published version
Accessed
Wed Sep 22 19:22:44 EDT 2010
Citable Link
http://hdl.handle.net/1721.1/51331
Terms of Use
Article is made available in accordance with the publisher's policy
and may be subject to US copyright law. Please refer to the
publisher's site for terms of use.
Detailed Terms
Mott Transition between a Spin-Liquid Insulator and a Metal in Three Dimensions
Daniel Podolsky,
1
Arun Paramekanti,
1
Yong Baek Kim,
1
and T. Senthil
2
1
Department of Physics, University of Toronto, Toronto, Ontario M5S 1A7, Canada
2
Department of Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA
(Received 1 December 2008; published 6 May 2009)
We study a bandwidth controlled Mott metal-insulator transition (MIT) from a Fermi-liquid metal to a
quantum spin-liquid insulator in three dimensions. Using a slave rotor approach including gauge
uctuations, we obtain a continuous MIT and discuss nite temperature crossovers in its vicinity. We
show that the specic heat C $T lnln1,T at the MIT and that the metallic state near the MIT should
exhibit a conductivity minimum as a function of temperature. We suggest Na
4
Ir
3
O
8
as a candidate to
test our predictions and compute its electron spectral function at the MIT.
DOI: 10.1103/PhysRevLett.102.186401 PACS numbers: 71.10.Hf
Introduction.Recent experiments suggest that
Na
4
Ir
3
O
8
at ambient pressure may be the rst example of
a three-dimensional quantum spin-liquid Mott insulator
[1]. This material has Ir local moments residing on the
hyperkagome lattice which is a three-dimensional (3D)
network formed by corner-sharing triangles. Of the various
theoretical proposals for the spin-liquid phase observed in
this material [26], the data may be most consistent with a
particular gapless spin liquid whose low-energy excitations
are charge-neutral spin-1,2 fermions, called spinons,
which live on spinon Fermi surfaces in momentum
space [4,5]. Remarkably, very recent experiments show
that this material undergoes a transition to a metallic state
under pressure [7]. This indicates that Na
4
Ir
3
O
8
may be a
3D analog of the well-studied layered triangular lattice
organic material k-ET
2
Cu
2
CN
3
[8], which is insulating
at ambient pressure but undergoes a Mott transition to a
metal under moderate pressure. Theoretically, the prox-
imity to the Mott transition has been argued to lead to a
2D spin liquid with a spinon Fermi surface in
k-ET
2
Cu
2
CN
3
[9,10].
In this Letter, motivated by the recent experiments on
Na
4
Ir
3
O
8
, we present a theory of a continuous Mott tran-
sition between a spin liquid with a spinon Fermi surface
and a metal in 3D. We use a similar theoretical framework
as developed earlier by one of us [11] for a 2D Mott
transition between a Fermi liquid and a spin-liquid Mott
insulator in relation to k-ET
2
Cu
2
CN
3
. In 2D, several
interesting results such as a universal jump in the residual
resistivity and the emergence of marginal Fermi liquids in
an intermediate crossover scale were found. In this work,
we show that a continuous Mott transition is indeed pos-
sible in 3D, and moreover it is characterized by weaker
singularities in various physical quantities when compared
to the 2D case. In contrast to 2D, right at the Mott critical
point in 3D, the system is insulating. Indeed, we argue that
the electrical conductivity aT in the vicinity of the Mott
quantum critical point (QCP) should exhibit the following
behavior: (i) In the quantum critical (QC) regime T >
T

, we expect aT $T (with log corrections). Here the


crossover scale T

$jg g
c
j
1,2
, where g is the parameter
used to tune the transition (for instance, the pressure or the
strength of the Coulomb repulsion U,i). g
c
is the value of
the parameter at the zero temperature Mott QCP. (ii) On the
insulating side, for T <T

, the conductivity is thermally


activated. (iii) Finally, on the metallic side, for T $T

, we
argue that aT will have a conductivity minimum
before it grows and saturates at low temperature.
Equivalently, there will be a resistivity peak at a nonzero
temperature. This is sketched in Figs. 1(b) and 1(c). This
resistivity peak is the most striking result of this Letter.
As in the 2D case, the nite temperature crossovers near
the Mott transition are characterized by multiple energy
scales T

and T

as depicted schematically in Fig. 1(a). As


noted above, the scale T

determines the crossover in


transport properties associated with the Mott transition.
However, there are only weak changes in physical proper-
T
**
T
*

f,0
T
A
B
C

(c)

f,0
metal
CSL
g (or U/t)
spinliquid
metal
anomalous
T/t
A B C
(a)
T
B
C
A

(b)
T
quantum critical
*
T
**
FIG. 1 (color online). (a) Schematic phase diagram of the 3D
metal-insulator transition. Finite temperature crossovers at the
temperatures T

and T

, shown by dashed lines, separate


regions distinguished by the specic heat and conductivity
behaviors, as discussed in the text (C-SL labels the charge-
uctuation-renormalized spin liquid). (b) Resistivity and
(c) conductivity along the linesA, B, and Cin (a).
PRL 102, 186401 (2009)
P HYS I CAL RE VI E W L E T T E RS
week ending
8 MAY 2009
0031-9007,09,102(18),186401(4) 186401-1 2009 The American Physical Society
ties across the scale T

$jg g
c
j
3,2
. For instance, for
T ) T

, in the regimes labeled QC, anomalous metal,


and charge-uctuation-renormalized spin liquid (C-SL),
the specic heat behaves as C $T lnln1,T. For T ( T

in the Mott insulator C $T ln1,T, while in the same


regime on the metallic side we recover the Fermi-liquid
result C $T. With an eye toward application to Na
4
Ir
3
O
8
,
we also present predictions for the T 0 electron spectral
function and tunneling density of states at the Mott tran-
sition for a tight binding model appropriate to this material.
Microscopic model and eld theory.To understand
the universal features of this kind of 3D Mott transition,
we start with a single band Hubbard model at half lling
on a nonbipartite lattice H i
P
h|]i
c
y
|a
c
]a

U
2
P
|
n
2
|

P
|
n
|
, where c
y
|a
is the creation operator for an electron at
site | with spin a and n
|
c
y
|a
c
|a
is the electron number
operator (repeated spin indices are summed over). This
Hamiltonian is expected to have a metallic Fermi-liquid
ground state for small on-site repulsion U,i ( 1 and to
have a Mott-insulating ground state for U,i ) 1.
(Henceforth, we set i 1.) General arguments [9] suggest
that a spin-liquid state with a spinon Fermi surface is a
likely ground state just on the insulating side of the Mott
transition in this model.
The Mott transition between a Fermi-liquid metal and
such a spin-liquid insulator is most conveniently accessed
in a slave rotor formalism [12], in which the electron
creation operator is written as a product c
y
|a
e
|0
|
]
y
|a
of
a charged spinless rotor eld
|
e
|0
|
and a spin-1,2
charge-neutral spinon ]
|a
. In order to eliminate un-
physical states from the enlarged spinon-rotor Hilbert
space, we must impose the constraint n
0
|
n
]
|
1 at
each site. Upon coarse-graining, this theory can be recast
in the form of a gauge theory which consists of a
dynamical U(1) gauge eld o

o
0
, a coupled to both
the spinon and rotor elds [10]. We begin by writing down
a continuum theory to describe this physics in terms of a
complex bosonic eld $e
|0
, a fermionic eld c
a
, and a
gauge eld o

, with action S
R

0
Jr
R
J
3
rL, where
L L
l
L
]
L
g
L
l]
,
L
l
j

|o

j
2
m
2
jj
2
ujj
4
,
L
]
c
y
a

r
|o
0

]
c
a

1
2m
]
jr |ac
a
j
2
,
L
g

1
4g
2

o
:

:
o

2
, L
l]
jc
a
j
2
jj
2
,
(1)
which are all allowed by the symmetries of the system and
thus will arise in the coarse-graining process. The boson
action L
l
is particle-hole symmetric, as appropriate for a
coarse-grained slave rotor theory at half lling.
At mean-eld level, we can ignore uctuations of the
gauge eld o

. In the regime m
2
>0, the charged rotor
eld is gapped, so that h
|
i 0, while the spinons form a
gapless Fermi surface. This is the spin-liquid insulator. On
the other hand, the metallic phase at m
2
<0 corresponds to
a Bose-Einstein condensate of rotors with h
|
i 0. The
spinon Fermi surface now becomes an electron Fermi
surface with a quasiparticle weight proportional to
jh
|
ij
2
. Therefore, the transition to the metal is tuned by
the boson gap parameter m
2
, with m
2
/ U U
c
. The MIT,
in this mean-eld theory, is thus simply a Bose condensa-
tion of rotors with the spinon Fermi surface being contin-
uously connected to the electron Fermi surface, and hence
it is a continuous transition.
We now consider a full theory of this Mott transition
obtained by analyzing uctuation corrections to all low-
energy degrees of freedom (spinons, rotors, and gauge
elds) in a self-consistent fashion. The methods are similar
to those used in 2D [11], so we focus here on stating the
results and discussing experimental implications.
We begin by considering the boson interaction term
ujj
4
together with the fermion-boson density interaction
L
l]
, in the absence of the gauge eld o

. (The gauge eld


is reinstated below.) Upon integrating out fermions [see
Fig. 2(e)], this yields an effective boson-boson interaction
L
0

Z
:
n
:
0
n
w
n
k, k
0
, q
~ uq, |w
n

y
kq,:
n
w
n

k,:
n

y
k
0
q,:
0
n
w
n

k
0
,:
0
n
,
where ~ u u N
0

2
r
]
q, |w
n
, N
0
is the fermionic den-
sity of states, and r
]
q, |w
n
is the density-density polar-
ization function of the fermions. For jqj ( 2k
]
, and in the
limit jw
n
j (
]
q, r
]
! 2
jw
n
j

]
q
. Under a naive rescaling
of space and time with dynamical critical exponent z 1,
L
0
is marginal by power counting at the critical xed point
of the bosonic theory, just like the conventional ujj
4
term. The fate of this term can be determined by a
standard perturbative renormalization group analysis.
Equivalently, we generalize the theory to N
l
boson
avors and consider the N
l
1 limit of the theory.
FIG. 2. Dominant uctuations. The solid and dashed arrows
represent fermions and bosons, respectively. (a) and (b) are the
gauge polarization functions arising from integrating out fermi-
ons and bosons, respectively. (c) Spinon self-energy, where the
double wavy line is the dressed gauge propagator. (d) Boson self-
energy. (e) Residual bosonic interactions.
PRL 102, 186401 (2009)
P HYS I CAL RE VI E W L E T T E RS
week ending
8 MAY 2009
186401-2
In this limit, all diagrams can be summed exactly to
yield a renormalized bosonic interaction u
1
ren
q, w
~ u
1
q, w ln,

w
2
q
2
m
2
p
. Near the QCP, the
second term dominates, and u
ren
! 0. Hence, L
0
is (dan-
gerously) marginally irrelevant.
Now we reinstate the gauge eld and rst consider its
renormalization from integrating out the matter elds. We
work in the Coulomb gauge r a 0, in which o
0
and a
decouple in L
g
. Then the longitudinal eld o
0
is screened
by the gapless fermions and can thus be ignored. On the
other hand, the transverse part of a remains unscreened,
but it is strongly renormalized by the matter elds, whose
contribution dominates over the bare gauge eld action. In
the random phase approximation, the renormalized inverse
transverse gauge eld propagator is D
1
q, |:
n

]
q, |:
n

l
q, |:
n
, where
]
and
l
are fermion
and boson polarization functions, respectively.
The fermionic contribution, shown in Fig. 2(a), is
]

0
j:
n
j
q

0
q
2
. The term proportional to
0
describes the
Landau damping due to the gapless spinons, and
0
is the
diamagnetic susceptibility of the spinons. This form can be
further justied in the large N
]
limit, where N
]
is the
number of avors of the fermions. The corrections appear-
ing in a 1,N
]
expansion do not change this form [13]. The
bosonic contribution shown in Fig. 2(b) for m
2
! 0 (or
U ! U
c
) is
l

q
2
24r
2
ln
m
0
q
2
m
2

1,2
, where m
0
is a nonun-
iversal constant, leading to
l

q
2
24r
2
ln
m
0
q
at the critical
point U U
c
and
l
%
q
2
24r
2
ln
m
0
m
for U >U
c
. On the
other hand, when m
2
<0 or U <U
c
,
l
%
s
, where

s
/ jhij
2
is the superuid stiffness of the bosons.
Thus the bosons determine the different forms of the gauge
eld propagator in various regimes.
It is readily seen that the boson self-energy
l
q, w,
given by Fig. 2(d), acquires only analytic corrections in q
and w, and thus the bosons are not renormalized in an
essential way by the gauge uctuations. The boson sector is
thus not affected by the gauge uctuations in the scaling
limit near the QCP [11].
The fermion self-energy
]
shown in Fig. 2(c) can be
computed by using the gauge eld propagators obtained
above. Near the spinon Fermi surface jkj % k
]
, this leads
to
]
$wlnln1,jwj |
r
2
jwj
ln1,jwj
at the QCP U U
c
and

]
$wln1,jwj |
r
2
jwj in the spin liquid for U >U
c
.
Finally, recall that the electron operator is decomposed
into a boson and a fermion locally (in space and time) as
c
y
a
r, r r, rc
y
a
r, r. This leads to an electron
Greens function that is a convolution (in momentum and
frequency) of the and c
a
Greens functions. Vertex
corrections to this result, arising from interactions with
the gauge eld, are unimportant at low frequencies [11].
In the metallic phase, we thus get the electron Greens
function G
e
jhij
2
G
]

Z
e
we

|u
, with e

$
e

lnln1,
s
and
Z
e
$

s
lnln1,
s
. By using
s
/ U
c
U ln
1
U
c
U
, the effec-
tive mass of the electrons m

e
,m
e
$lnln1,U
c
U di-
verges very weakly. The quasiparticle weight
Z
e
$
U
c
U ln
1
U
c
U
lnln1,U
c
U
diminishes as one approaches the QCP.
Scaling, nite temperature phase diagram, and specic
heat.The nite T crossovers involve multiple energy
scales originating in the different space-time scaling of
the gauge eld and the bosons. Note that the mean-eld
transition of bosons at nite temperature, below which

s
0, becomes a crossover in the full theory and repre-
sents the onset of the charge coherence.
The behavior of the charge excitations is governed by
T $w$q scaling (z 1 theory). In either phase, for T *
T

$jU U
c
j
1,2
, the charge bosons are in their quantum
critical regime. In the insulator, T

may be identied with


the zero temperature charge gap.
The scaling behavior of the coupled spinon-gauge sys-
tem is, however, determined by T $w$q
3
scaling (z 3
theory) up to logarithmic corrections (in principle, one
should use w$q
3
ln1,q scaling). The spinon-gauge sys-
temthus emerges out of the quantumcritical regime only at
a scale T

$jU U
c
j
3,2
. This crossover scale is, in
principle, visible in the specic heat which behaves as
C %
8
>
<
>
:
T lnln1,T T >jU U
c
j
3,2
,

1
T ln1,T T <U U
c

3,2
,

2
T T <U
c
U
3,2
,
(2)
where
1
/ 1, ln
1
UU
c

and
2
$lnln
1
U
c
U
. Note, how-
ever, that there are only rather weak changes in properties
across this second crossover temperature T

.
The crossover scale T

is, on the other hand, visible in


the electrical conductivity which is readily measured in
experiments. It is obtained from the Ioffe-Larkin rule,
which states that, in a slave-particle formalism, the elec-
trical resistivity of the system is the sum of the resistivities
of the fermions and the bosons:
]

l
. For the
fermions, we assume that the conductivity is dominated
by elastic scattering from impurities: a
]
a
],0
. The bo-
sonic conductivity, on the other hand, is strongly tempera-
ture- and pressure-dependent. In the Mott insulator, when
the bosons are gapped (T <T

), the bosonic conductivity


is activated, and a
l
T $e
T

,T
. On the other hand, in the
metal, the bosons are condensed, and a
l
T ! 0 ! 1. In
the quantum critical regime, standard scaling arguments
give a
l
T $Tln
2 1
T
(the log corrections are due to the
marginally irrelevant boson-boson interactions). As this
goes to zero for small T, the boson contribution dominates
over the fermions due to the Ioffe-Larkin rule. Hence, the
QCP itself is insulating.
These results for the conductivity are summarized in
Figs. 1(b) and 1(c). Note that, in the metallic phase near
the QCP, a is not monotonic with temperature. At high T,
the system is in the (insulating) quantum critical fan, and
conductivity decreases with decreasing T. Only at low T
does the system realize that it is inside the metal, and aT
begins to rise. Hence, it is possible for a system to be
PRL 102, 186401 (2009)
P HYS I CAL RE VI E W L E T T E RS
week ending
8 MAY 2009
186401-3
metallic and yet display insulating behavior at high T.
The nonmonotonicity becomes more marked for cleaner
samples, i.e., when a
],0
is large. Experimental observation
of such a conductivity would provide dramatic evidence
for the 3D MIT discussed here.
Mean-eld theory for hyperkagome lattice.We have
shown that the bosons and spinons are only weakly af-
fected by the gauge uctuations beyond mean-eld theory
near the QCP. Encouraged by this observation, we closely
investigate the mean-eld theory for the Hubbard model on
the hyperkagome lattice (of Na
4
Ir
3
O
8
) by approximating
the ground and excited states as direct products of spinon
and rotor wave functions ji j
0
ij
]
i [14]. We nd
[15] a continuous quantum phase transition at U
c
% 6 from
a spin-liquid insulator to a metal, with the electron spectral
function given by
A
e
k, w
X
q,n
M
n
k q
2
q
f1 n
q
](
n
kq

uw(
n
kq

q

n
q
](
n
kq
uw(
n
kq

q
g,
(7)
where (
n
k
is the dispersion of the spinon band n 1 . . . 12,

q
is the boson dispersion, and n and ] are the Bose and
Fermi distribution functions, respectively. Here, the form
factor M
n
k q arises fromthe spinon wave function in
band n. At the QCP, we assume a boson dispersion of the
form
2
q
4
2
l
sin
2
q
x
,2 sin
2
q
y
,2 sin
2
q
z
,2, so
that
q
%
l
jqj for small momenta. For illustrative pur-
poses, we choose
l
0.5. Deep in the metallic phase at
T 0, where the bosons are fully condensed,
A
met
e
k, w
P
n
M
n
kuw(
n
k
.
Figure 3 shows the T 0 spectral function along some
high symmetry directions of the Brillouin zone (BZ). Deep
in the metallic phase, the spectral function clearly shows
12 bands (after accounting for the fourfold degeneracy of
the at band at high energy). Momentum variations of the
form factor M
n
k lead to changes in the intensity for
the various bands as one traverses the BZ. As one ap-
proaches the QCP, the overall bandwidth gets suppressed
due to the Hubbard U. In contrast to the metallic phase, the
spectra at the QCP are very broad due to rotor uctuations.
Finally, as evident from Fig. 3, there is strong suppression
of spectral weight around the chemical potential. This
pseudogap is also reected in the tunneling density of
states. Integrating A
crit
k, w over all k, we nd
N
crit
tunn
w $w
2
at low frequencies.
Discussion.Recently, it was shown that the specic
heat of the insulating Na
4
Ir
3
O
8
at not-too-lowtemperatures
is well described by a renormalized mean-eld theory of a
gapless spin liquid with a spinon Fermi surface [4]. Such
success, however, may seem at odds with the fact that the
gauge eld uctuations would give rise to C,T $ln1,T
deep inside the spin-liquid regime. In light of the present
work and the recent discovery of the insulator-metal tran-
sition in Na
4
Ir
3
O
8
upon increasing pressure [7], we may
expect that the spin-liquid phase of Na
4
Ir
3
O
8
is already
near the QCP. In this case, the nite temperature behavior
of the specic heat would be dominated by the C-SL
regime with C,T $lnln1,T, while the bona de spin-
liquid insulator appears only at very low T. This weak
singularity would indeed be consistent with the apparent
success of the renormalized mean-eld theory. Another
important point is that, while gauge uctuations deep in
the spin liquid are strong and could potentially destabilize
the spinon Fermi surface, these uctuations are very weak
in the C-SL regime, making it a robust feature of such a
MIT. Further progress on understanding the MIT in
Na
4
Ir
3
O
8
needs a careful comparison between future ex-
periments and our detailed predictions for thermodynamics
and transport.
We thank Professor H. Takagi for helpful discussions
and for sharing unpublished data with us. This work was
supported by CIFAR, CRC (D. P. and Y. B. K.), the NSERC
of Canada (D. P., A. P., and Y. B. K.), Sloan Foundation and
Ontario ERA (A. P.), and NSF Grant No. DMR-0705255
(T. S.).
[1] Y. Okamoto et al., Phys. Rev. Lett. 99, 137207 (2007).
[2] J. M. Hopkinson, S. V. Isakov, H-Y. Kee, and Y. B. Kim,
Phys. Rev. Lett. 99, 037201 (2007).
[3] M. J. Lawler et al., Phys. Rev. Lett. 100, 227201 (2008).
[4] M. J. Lawler et al., Phys. Rev. Lett. 101, 197202 (2008).
[5] Y. Zhou et al., Phys. Rev. Lett. 101, 197201 (2008).
[6] G. Chen and L. Balents, Phys. Rev. B 78, 094403 (2008).
[7] H. Takagi (unpublished).
[8] Y. Shimizu et al., Phys. Rev. Lett. 91, 107001 (2003).
[9] O. Motrunich, Phys. Rev. B 72, 045105 (2005).
[10] S-S. Lee and P. A. Lee, Phys. Rev. Lett. 95, 036403 (2005).
[11] T. Senthil, Phys. Rev. B 78, 045109 (2008).
[12] S. Florens and A. Georges, Phys. Rev. B 70, 035114 (2004).
[13] Y. B. Kim et al., Phys. Rev. B 50, 17 917 (1994); B. L.
Altshuler, L. B. Ioffe, and A. J. Millis, Phys. Rev. B 50,
14 048 (1994).
[14] E. Zhao and A. Paramekanti, Phys. Rev. B 76, 195101
(2007).
[15] D. Podolsky, A. Paramekanti, and Y. B. Kim (unpub-
lished).
FIG. 3 (color online). Intensity plot of the electron spectral
function along high symmetry directions in the metal (left) and
at the QCP (right). 0, 0, 0, R r, r, r, and M
r, r, 0, and w 0 is the chemical potential.
PRL 102, 186401 (2009)
P HYS I CAL RE VI E W L E T T E RS
week ending
8 MAY 2009
186401-4

S-ar putea să vă placă și