Quantum Critical Points with the Coulomb Interaction and the Dynamical Exponent: When and Why

4
VOLUME 87, NUMBER 13 PHYSICAL REVIEW LETTERS 24 SEPTEMBER 2001 Quantum Critical Points with the Coulomb Interaction and the Dynamical Exponent: When and Why z 5 1 Igor F. Herbut Department of Physics, Simon Fraser University, Burnaby, British Columbia, Canada V5A 1S6 (Received 1 June 2001; published 10 September 2001) A general scenario that leads to Coulomb quantum criticality with the dynamical critical exponent z 1 is proposed. I point out that the long-range Coulomb interaction and quenched disorder have competing effects on z , and that balance between the two may lead to charged quantum critical points at which z 1 exactly. This is illustrated with the calculation for the Josephson junction array Hamiltonian in dimensions D 3 2e. Precisely in D 3, however, the above simple result breaks down, and z . 1. Relation to other studies is discussed. DOI: 10.1103/PhysRevLett.87.137004 PACS numbers: 74.40. +k, 05.30. – d, 71.30. +h, 73.50. –h The crucial difference between the quantum (T 0) and the more familiar finite temperature phase transitions is that while dynamics is irrelevant for the latter, it is es- sential for the former [1]. The link between statics and dy- namics at a continuous quantum phase transition is usually parametrized with the value of dynamical critical exponent z that describes relative scaling of the time and the length scales in the problem [2,3]. Together with the correlation length exponent n, z enters the low-temperature scaling of all physical observables, since in the vicinity of a quantum critical point temperature scales as T jdj zn , where d is the T 0 tuning parameter. The value of z for a given quantum critical point is therefore of great interest, and has often been used to distinguish one universality class from another. In their seminal paper on universal conductivity in two dimensions, Fisher, Grinstein, and Girvin [4] also proposed that when the long-range Coulomb interaction is present, at the criticality energy should scale as the inverse of length, so that z 1 should result. This well-known conjecture has since been used in interpreting some of the most intriguing experiments in modern condensed matter physics, ranging from superconductor-insulator transitions in low-dimensional systems [5], via metal-insulator tran- sitions in Si-metal-oxide-semiconductor field-effect tran- sistors [6], to the universality of the underdoped high- T c cuprates [7]. It has also been utilized in the Monte Carlo simulations where knowing z in advance greatly simplifies the inevitable finite-size scaling analysis [8]. The purpose of this Letter is to provide the theoretical justification for this widely used relation and point to its limitations. I show that in certain dimensions the Coulomb coupling constant (i.e., charge) is protected from re- normalization, and consequently its flow under scaling transformation is directly proportional to its canonical dimension, which is just z 2 1 [4] [see Eq. (2)]. This implies that if a charged critical point exists in the theory, its location is actually determined by the solution of the equation z 1, which then also determines the value of the dynamical critical exponent exactly. I argue that this situation arises when there is an additional coupling in the theory with the competing effect on the dynamical exponent, and which can balance the effect of Coulomb interaction. Such a coupling is shown to be provided by quenched disorder, and a concrete realization of the above scenario is worked out on the example of a disordered Josephson junction array Hamiltonian in D 3 2e dimensions. Finally, the simple relation z 1 is found to break down at special dimensions at which the above renormalization group (RG) protectorate on charge is lifted. Relation to other recent theoretical studies of Coulomb criticality is discussed. To be specific, I will focus on the theory originally con- sidered in [9] which describes an array of coupled Joseph- son junctions, but it will transpire that the underlying physics is more general. Building on an earlier work by Ma [10], Fisher and Grinstein [9] have shown that Coulomb interaction can be represented by a minimal coupling to the soft scalar gauge field. In the long wavelength limit the critical field theory that describes the system of bosons interacting via Coulomb interaction at T 0 and in D dimensions takes the form [9] S Z d D rdt jt 2 iA 0 Cj 2 1 j=Cj 2 1 V r , t 1 m 2 jCj 2 1ljCj 4 1 1 2e A 0 j=j D21 A 0 æ , (1) where by j=j D21 it is meant j kj D21 in the Fourier space. C r , t is the complex superfluid order parameter, and A 0 r , t is the scalar gauge field, which when integrated out introduces the Coulomb interaction into the remain- ing action for C. I have also included a random po- tential V r , t, but more on this in a moment. More generally, the gauge-field propagator in the momentum space is e 2 V c k 2 1 [9], which for the Coulomb inter- action V c r 1r then yields the last term in (1) at small momenta. The presence of a neutralizing background is included by omitting the k 0 components of the gauge field [10]. The above action without randomness V x, t 0 was first studied by Fisher and Grinstein [9], and recently 137004-1 0031-9007 01 87(13) 137004(4)$15.00 © 2001 The American Physical Society 137004-1

Transcript of Quantum Critical Points with the Coulomb Interaction and the Dynamical Exponent: When and Why

VOLUME 87, NUMBER 13 P H Y S I C A L R E V I E W L E T T E R S 24 SEPTEMBER 2001

1370

Quantum Critical Points with the Coulomb Interaction and the Dynamical Exponent:When and Why z 5 1

Igor F. HerbutDepartment of Physics, Simon Fraser University, Burnaby, British Columbia, Canada V5A 1S6

(Received 1 June 2001; published 10 September 2001)

A general scenario that leads to Coulomb quantum criticality with the dynamical critical exponentz 1 is proposed. I point out that the long-range Coulomb interaction and quenched disorder havecompeting effects on z, and that balance between the two may lead to charged quantum critical points atwhich z 1 exactly. This is illustrated with the calculation for the Josephson junction array Hamiltonianin dimensions D 3 2 e. Precisely in D 3, however, the above simple result breaks down, andz . 1. Relation to other studies is discussed.

DOI: 10.1103/PhysRevLett.87.137004 PACS numbers: 74.40.+k, 05.30.–d, 71.30.+h, 73.50.–h

The crucial difference between the quantum (T 0)and the more familiar finite temperature phase transitionsis that while dynamics is irrelevant for the latter, it is es-sential for the former [1]. The link between statics and dy-namics at a continuous quantum phase transition is usuallyparametrized with the value of dynamical critical exponentz that describes relative scaling of the time and the lengthscales in the problem [2,3]. Together with the correlationlength exponent n, z enters the low-temperature scaling ofall physical observables, since in the vicinity of a quantumcritical point temperature scales as T jdjzn , where d

is the T 0 tuning parameter. The value of z for a givenquantum critical point is therefore of great interest, and hasoften been used to distinguish one universality class fromanother. In their seminal paper on universal conductivityin two dimensions, Fisher, Grinstein, and Girvin [4] alsoproposed that when the long-range Coulomb interaction ispresent, at the criticality energy should scale as the inverseof length, so that z 1 should result. This well-knownconjecture has since been used in interpreting some of themost intriguing experiments in modern condensed matterphysics, ranging from superconductor-insulator transitionsin low-dimensional systems [5], via metal-insulator tran-sitions in Si-metal-oxide-semiconductor field-effect tran-sistors [6], to the universality of the underdoped high- Tccuprates [7]. It has also been utilized in the Monte Carlosimulations where knowing z in advance greatly simplifiesthe inevitable finite-size scaling analysis [8].

The purpose of this Letter is to provide the theoreticaljustification for this widely used relation and point to itslimitations. I show that in certain dimensions the Coulombcoupling constant (i.e., charge) is protected from re-normalization, and consequently its flow under scalingtransformation is directly proportional to its canonicaldimension, which is just z 2 1 [4] [see Eq. (2)]. Thisimplies that if a charged critical point exists in the theory,its location is actually determined by the solution of theequation z 1, which then also determines the value ofthe dynamical critical exponent exactly. I argue that thissituation arises when there is an additional coupling inthe theory with the competing effect on the dynamical

04-1 0031-90070187(13)137004(4)$15.00

exponent, and which can balance the effect of Coulombinteraction. Such a coupling is shown to be provided byquenched disorder, and a concrete realization of the abovescenario is worked out on the example of a disorderedJosephson junction array Hamiltonian in D 3 2 e

dimensions. Finally, the simple relation z 1 is foundto break down at special dimensions at which the aboverenormalization group (RG) protectorate on charge islifted. Relation to other recent theoretical studies ofCoulomb criticality is discussed.

To be specific, I will focus on the theory originally con-sidered in [9] which describes an array of coupled Joseph-son junctions, but it will transpire that the underlyingphysics is more general. Building on an earlier work by Ma[10], Fisher and Grinstein [9] have shown that Coulombinteraction can be represented by a minimal coupling tothe soft scalar gauge field. In the long wavelength limitthe critical field theory that describes the system of bosonsinteracting via Coulomb interaction at T 0 and in Ddimensions takes the form [9]

S Z

dD r dt

Ωj≠t 2 iA0Cj2 1 j=Cj2

1 V r , t 1 m2 jCj2 1 ljCj4

112e

A0j=jD21A0

æ, (1)

where by j=jD21 it is meant j kjD21 in the Fourier space.Cr, t is the complex superfluid order parameter, andA0r , t is the scalar gauge field, which when integratedout introduces the Coulomb interaction into the remain-ing action for C. I have also included a random po-tential V r, t, but more on this in a moment. Moregenerally, the gauge-field propagator in the momentumspace is e2Vck 2 1 [9], which for the Coulomb inter-action Vc r 1r then yields the last term in (1) at smallmomenta. The presence of a neutralizing background isincluded by omitting the k 0 components of the gaugefield [10].

The above action without randomness V x, t 0was first studied by Fisher and Grinstein [9], and recently

© 2001 The American Physical Society 137004-1

VOLUME 87, NUMBER 13 P H Y S I C A L R E V I E W L E T T E R S 24 SEPTEMBER 2001

revisited by Ye [11]. In D 3 both the charge (e2) andthe quartic interaction (l) are marginally irrelevant, andthe Coulomb interaction causes the flow to run away tonegative l, which may be interpreted as a sign of a dis-continuous transition. In D 3 2 e the result is moreinteresting: charge is irrelevant at the XY critical point, butif too large may still lead to a runaway flow. Irrelevanceof the charge in the theory (1) arises in a somewhat non-trivial way, and it will prove instructive to understand thisin some detail. In D , 3 the inverse gauge-field propa-gator in (1) is nonanalytic in k and therefore e2 cannotget renormalized by the integration over the high-energymodes. This RG protectorate is reminiscent of the situ-ation in the (2 1 1)-dimensional electrodynamics with theChern Simons term [12], where the statistical angle is ex-actly marginal for a similar reason. Here this means thatthe b function for the charge is determined exclusively byits canonical dimension. Gauge invariance and the accom-panying Ward identity imply that A0 t21, so assumingt Lz yields to

de2

d lnb e2z 2 1 (2)

in D , 3, where b is the standard RG parameter. Theexponent z in the last equation then needs to be deter-mined from the renormalization of the C propagator. Thesimplest one-loop calculation [see the first diagram onFig. 1(a)] in D 3 2 e then gives z 1 2 e23, andthus smaller than one. By Eq. (2) the small charge is thenirrelevant. Two features of this result that are likely to

++ + (a)

++ +

+ + +

(b)

+ + + (c)

(d)+

FIG. 1. One-loop contributions to the renormalization of self-energy, quartic interaction, disorder coupling, and polarization.Wavy lines represent the gauge-field propagator, and dashedlines the disorder vertex.

137004-2

be quite general should be noticed: first, z fi 1 since thescalar gauge field couples only to the time derivative, andthus discriminates between space and time. Second, thenegative sign in Eq. (2) comes from the tendency of thegauge field to soften the C propagator. It is analogous tothe well-known (and much debated) result that the anoma-lous dimension in the scalar (Higgs) electrodynamics isnegative [13–15]. It could therefore be expected that whenhigher order terms in e2 are included z , 1 will remain,and that irrelevancy of the charge in the pure theory (1) ismore general than the simple one-loop calculation wouldsuggest.

Equation (2) also implies that if there would exist acritical point in the theory with nonzero charge, in D , 3at that fixed point z 1 exactly. This nonperturbativeresult parallels another exact result hA 1 in the D 3scalar electrodynamics [13,16] where hA is the anomalousdimension of the vector gauge field. The crucial question ishow can such a Coulomb critical point (with e2 fi 0) arise.A clue is provided already in the preceding discussion: oneneeds another coupling which will tend to increase z andbalance the effect of Coulomb interaction in Eq. (2). Aphysically realistic candidate is disorder: in the quantumtheory the random potential V r, t is random in space butstatic and independent of imaginary time. This anisotropywill in general lead to a nontrivial z, and it is well knownthat the effect to the lowest order is always to increase it[17]. In other words, while the weak Coulomb interactionis marginally irrelevant at the XY critical point, I expect itto become relevant at the disordered critical point wherez . 1. The small charge near the random critical pointshould grow until it balances disorder in Eq. (2). As aresult, a new stable (Coulomb) critical point may arise, atwhich z 1 exactly.

Next I demonstrate that the above scenario is indeedborn out in the one-loop calculation in the theory (1),in passing reconciling the apparently different results inRefs. [9] and [11]. I then proceed to show how the equal-ity z 1 is violated in D 3, and comment on relationsto other works.

To exert some control over the fixed points in the the-ory I will assume both a small e 3 2 D and et , wherethe latter is the number of dimensions over which disorderis correlated [18]. The physically interesting case corre-sponds to et 1, but since I am primarily interested in thepoint of principle, convergence properties of the double-eexpansion will not be of much concern here [19]. To aver-age over disorder I utilize the standard replica trick, whichattaches a replica index a 1, . . . , N onto all fields in theaction (1) and introduces another interactionlike term inthe theory

2W2

NXa,b1

ZdD11 R dD11 R0 dD112et R 2 R0

3 jCa Rj2jCb R0j2, (3)

137004-2

VOLUME 87, NUMBER 13 P H Y S I C A L R E V I E W L E T T E R S 24 SEPTEMBER 2001

where the limit N ! 0 is to be taken at the end of thecalculation [18]. Here R r, t, and for et 1 onerecovers the quantum problem where disorder is uncor-related (Gaussian) in space and independent of the imagi-nary time.

To perform the Wilson-Fisher momentum shell renor-malization group one integrates out all the fields withLb , k , L and 2` , v , `, where L is theultraviolet cutoff. The effect of this procedure is to alterthe coefficients in front of v2 Zv, k2 Zk, m2 Zm,l Zl, W ZW , and A2

0-term ZA in the Fourier trans-formed action (1) [20]. One then rescales the momentaand the frequencies as bk ! k and bzv ! v, and thefields as b2DA0 ! A0 and b221D1z2Z

12k C ! C, to

find finally that by defining new coupling constants aslb b42D2zZ22

k Zll, W b b42D2z1et Z22k ZWW ,

and e2b bz21Z21A e2 the action can be restored into its

original form (at the critical point m2 0. Computingnext the Z factors diagrammatically to one-loop order(Fig. 1) gives at the criticality

de2

d lnb e2z 2 1 2 d3,D

112

e4, (4)

dl

d lnb

µe 1

12

e2 1118

W

∂l 2

52

l2 214

e4,

(5)

dW

d lnb

µe 1 et 2 2l 1

12

e2

∂W 1

78

W2. (6)

The exponent z in Eq. (4) is determined by demanding thatb22zZv b22Zk, which gives

z 1 118 W 2

13 e2. (7)

Note that for D , 3 the flow equation for the charge re-duces to Eq. (2). Precisely in D 3 the inverse propaga-tor for the gauge field becomes analytic, k2. This meansthat in D 3 the charge becomes renormalized by the po-larization diagrams in Fig. 1(d), which contribute the lastterm in Eq. (4). For W 0 the preceding b functions re-duce to those of Ref. [9] when D , 3, and coincide withthose of [11] right at D 3, upon simple redefinitions ofcouplings. They are also equivalent to those of [19] fore2 0. It may be interesting to note that many of the in-dividual diagrams on Fig. 1 are ultraviolet divergent, dueto the independence of the gauge-field propagator on fre-quency. All those divergences exactly cancel out in thefinal result [11]. Finally, the flow of the mass term in (1)yields the correlation length exponent:

n 12

114

µl 1

e2

62

W4

∂. (8)

Let us turn now to the fixed points of the above equa-tions. Besides the Gaussian and the XY fixed pointsat W e2 0, both unstable with respect to disorder,

137004-3

there are two disordered fixed points. First, at e2 0,there is a neutral disordered fixed point [18] at ln 24e 1 11et9 and Wn 8e 1 5et9, which is at-tractive in the l 2 W plane. At the neutral fixed pointzn 1 1 W8 . 1, so a weak Coulomb interaction is arelevant perturbation. With a small charge turned on, theflow is towards a new, stable, Coulomb critical point. InD , 3 this fixed point is located at

lc 100183

e

Ω1 1

389et

100e

1

sµ1 1

389et

100e

∂2

116475000

µ1 1

et

e

∂2æ,

(9)

Wc 1617 2lc 2 e 2 et , (10)

and

e2c 3

8 Wc , (11)

where the last equation ensures that z 1. The readershould note that disorder is necessary for the existence ofthe Coulomb critical point: without it the critical pointwould turn imaginary, and one would find only the stan-dard runaway flow characteristic of the gauge-field fluctua-tions in the e expansion [21]. The Coulomb critical pointtherefore may be considered as an example of a disorderinduced continuous phase transition.

Precisely in D 3 the last term in Eq. (4) becomesfinite. There still exists a stable Coulomb criticalpoint at lc 3.61et, Wc 402lc 2 et41, ande2

c 3Wc10, but with the dynamical critical exponent

z 1 1e2

c

12, (12)

which gives z 1.15, for example, for et 1. One findsz fi 1 in D 3 as a result of the removal of the RG pro-tectorate on charge. The same violation of the simple re-lation z 1 can be expected in other problems in specialdimensions. Also, since at the criticality Vcr 1rz ,one expects that z $ 1 in general [4], since screeningshould certainly not make the interaction longer ranged.The result (12) thus implies the Coulomb interaction hasbeen partially screened at the criticality in D 3, and nowdecays faster (but still as a power law) with distance [22].

The result z 1 was previously also found in the large-N theory of Dirac fermions in D 2 [23], interacting bothvia Coulomb interaction and with the Chern Simons field.This theory may be relevant to the quantum Hall state-insulator transition [2,24]. It was found that for a certainrange of the statistical angle (Chern Simons coupling) thecompetition between the Chern Simons and the Coulombinteractions leads to a nontrivial charged fixed point, atwhich z 1 exactly, just as in Eq. (2). The physical rea-son why the two interactions have competing effects onz only for some values of the Chern Simons coupling

137004-3

VOLUME 87, NUMBER 13 P H Y S I C A L R E V I E W L E T T E R S 24 SEPTEMBER 2001

remained somewhat obscure in this work. Nevertheless,the result in [23] bears a formal resemblance to mine.

Finally, z 1 was also found in the previous work bythe author on the quantum critical behavior of dirty bosonswith Coulomb interaction in D 1 1 e dimensions [25].There it arises as a consequence of a special symmetrythe theory dual to (1) possesses precisely in D 1, andcan therefore be suspected to be an artifact of the specificRG scheme that was employed. This Letter can thus beunderstood as complementing the previous work in that itshows that z 1 is also exact near the D 3, and it maytherefore be expected to hold in the physical case D 2.

It may also be interesting to note that simply settinge et 1 in the one-loop Eqs. (7) and (8), to crudelyestimate the exponents in D 2, besides the exact z 1also yields n 1.46 at the Coulomb criticality. Experi-mentally, n 1 [5], and the result for n is less reliablethan the one obtained in the expansion around D 1 [25].

To summarize, it was shown that the competition be-tween the correlated (quantum) disorder and Coulomb in-teraction may lead to a new charged critical point at whichthe dynamical critical exponent z 1 exactly. This simpleresult breaks down in special dimensions, and an exampleof the Josephson junction array when this happens in D 3 was provided. I argued that similar results should be ex-pected whenever there are two couplings in the theory withcompeting effects on z.

The author is grateful to Matthew Fisher and ShivajiSondhi for their useful comments and interest in this work.This research was supported by NSERC of Canada and theResearch Corporation.

[1] John A. Hertz, Phys. Rev. B 14, 1165 (1976).[2] S. L. Sondhi, S. M. Girvin, J. P. Carini, and D. Shahar, Rev.

Mod. Phys. 69, 315 (1997).[3] S. Sachdev, Quantum Phase Transitions (Cambridge Uni-

versity Press, Cambridge, U.K., 1999).

137004-4

[4] M. P. A. Fisher, G. Grinstein, and S. M. Girvin, Phys. Rev.Lett. 64, 587 (1990).

[5] For a review, see A. Goldman and N. Markovic, Phys.Today 51, No. 11, 39 (1998).

[6] E. Abrahams, S. V. Kravchenko, and M. P. Sarachik, Rev.Mod. Phys. 73, 251 (2001).

[7] T. Schneider and J. M. Singer, Phase Transition Approachto High-Temperature Superconductivity (Imperial CollegePress, London, 2000); T. Schneider, e-print cond-mat/0104053 (unpublished).

[8] M. Wallin, E. Sorensen, S. M. Girvin, and A. P. Young,Phys. Rev. B 49, 12 115 (1994).

[9] M. P. A. Fisher and G. Grinstein, Phys. Rev. Lett. 60, 208(1988).

[10] S.-K. Ma, Phys. Rev. Lett. 29, 1311 (1972).[11] J. Ye, Phys. Rev. B 58, 9450 (1998).[12] G. W. Semenoff, P. Sodano, and Y. S. Wu, Phys. Rev. Lett.

62, 715 (1989).[13] I. F. Herbut and Z. Tesanovic, Phys. Rev. Lett. 76, 4588

(1996).[14] J. Hove, S. Mo, and A. Sudbo, Phys. Rev. Lett. 85, 2368

(2000).[15] H. Kleinert and F. S. Nogueira, e-print cond-mat/0104573,

and references therein.[16] P. Olsson and S. Teitel, Phys. Rev. Lett. 80, 1964 (1998).[17] I. F. Herbut, Phys. Rev. Lett. 79, 3502 (1997); Phys. Rev.

B 57, 13 729 (1998), and references therein.[18] S. N. Dorogovtsev, Phys. Lett. 76A, 169 (1980); D. Boy-

anovsky and J. L. Cardy, Phys. Rev. B 26, 154 (1982); I. D.Lawrie and V. V. Prudnikov, J. Phys. C 17, 1655 (1984).

[19] See R. Mukhopadhyay and P. B. Weichman, Phys. Rev.Lett. 76, 2977 (1996).

[20] Gauge invariance of (1) guarantees preservation of theminimal coupling form in the renormalized theory. See,for example, Ref. [11].

[21] B. I. Halperin, T. C. Lubensky, and S.-K. Ma, Phys. Rev.Lett. 32, 292 (1974).

[22] Gauge invariance prohibits a mass term for the gauge fieldin Eq. (1) that would yield the Yukawa form of the screenedinteraction in real space.

[23] J. Ye and S. Sachdev, Phys. Rev. Lett. 80, 5409 (1998).[24] S. L. Sondhi and S. A. Kivelson (unpublished).[25] I. F. Herbut, Phys. Rev. Lett. 81, 3916 (1998); Phys. Rev.

B 58, 971 (1998); 61, 14 723 (2000).

137004-4