8
Dielectric critical behaviour at a metal–insulator transition under lattice compression Shailesh Shukla a , Deepak Kumar a , Nitya Nath Shukla b , Rajendra Prasad b, * a School of Physical Sciences, Jawaharlal Nehru University, New Delhi 110067, India b Department of Physics, Indian Institute of Technology, Kanpur 208016, India Received 28 December 2004; revised 15 March 2005; accepted 15 March 2005 Abstract We study dielectric critical behaviour around a continuous metal–insulator transition in crystalline Cesium Iodide induced by changing the lattice parameter. The ab initio calculations of band structure and various quantities related to the dielectric response are performed in the transition region, within the local density approximation of the density functional theory. These calculations allow us to establish the power- law singularities of various quantities on two sides of the transition. The exponents obtained here are mean-field like due to the approximation in which interactions and disorder are treated. The critical behaviour is discussed by applying the scaling principle to the wavevector and frequency dependent dielectric function. We further investigate the effect of dielectric anomalies on optical properties by calculating the reflectance around transition region taking the ionic contribution to the dielectric function also into account. We find that the reflectance as a function of frequency shows very different kind of behaviour on both sides of the metal–insulator transition. q 2005 Elsevier Ltd. All rights reserved. PACS: 71.30.Ch; 71.20.Mq; 71.20.Nr Keywords: D. Dielectric properties; D. Critical phenomena; A. Electronic materials; D. Optical properties 1. Introduction Metal–insulator transition in solids under variation of lattice parameter has been a subject of great theoretical interest for a long time [1,2]. According to Mott, solids with odd number of valence electrons like H, Na, etc. will undergo transition from metal to insulator on lattice expansion due to electron correlation effects [3]. Solids with even number of electrons per unit cell can also show metal insulator transition due to opening or closing of band gaps at the Fermi level when their lattice parameter is varied [3,4]. In this paper we study one such transition, which occurs through the band overlap process and analyse dielectric anomalies around the transition. Since one is dealing with a quantum phase transition, one expects certain universal features, which arise due to large scale spatial and temporal quantum fluctuations that accompany a continuous zero temperature transition [1,5,6]. There has been a tremendous amount of work toward understanding the critical behaviour around metal–insulator transition (MIT). The first set of ideas were introduced through the work of Wegner [7] and Abrahams et al. [8] in the context of disorder-induced metal insulator transition (MIT) for noninteracting electrons. Since interactions are inevitably present in any physical system, the work following these pioneering papers has naturally focussed on quantitative understanding of MIT in interacting systems [1,2,9,10]. Much of this activity has focussed on Mott– Hubbard transition induced by short ranged interaction as in Hubbard Model. Our aim in this work is to complement such studies by analysing a MIT associated with the closing of a one-particle gap induced by pressure in a realistic system. We should emphasize that this transition is being driven by the one-electron properties of the periodic lattice system. Though both interactions and disorder play a significant role in determining the properties of the transition, neither actually drives the transition. Nevertheless this MIT is associated with many dielectric anomalies, which we Journal of Physics and Chemistry of Solids 66 (2005) 1150–1157 www.elsevier.com/locate/jpcs 0022-3697/$ - see front matter q 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.jpcs.2005.03.005 * Corresponding author. E-mail address: [email protected] (R. Prasad).

Dielectric critical behaviour at a metal–insulator transition under lattice compression

Embed Size (px)

Citation preview

Page 1: Dielectric critical behaviour at a metal–insulator transition under lattice compression

Dielectric critical behaviour at a metal–insulator transition

under lattice compression

Shailesh Shuklaa, Deepak Kumara, Nitya Nath Shuklab, Rajendra Prasadb,*

aSchool of Physical Sciences, Jawaharlal Nehru University, New Delhi 110067, IndiabDepartment of Physics, Indian Institute of Technology, Kanpur 208016, India

Received 28 December 2004; revised 15 March 2005; accepted 15 March 2005

Abstract

We study dielectric critical behaviour around a continuous metal–insulator transition in crystalline Cesium Iodide induced by changing the

lattice parameter. The ab initio calculations of band structure and various quantities related to the dielectric response are performed in the

transition region, within the local density approximation of the density functional theory. These calculations allow us to establish the power-

law singularities of various quantities on two sides of the transition. The exponents obtained here are mean-field like due to the approximation

in which interactions and disorder are treated. The critical behaviour is discussed by applying the scaling principle to the wavevector and

frequency dependent dielectric function. We further investigate the effect of dielectric anomalies on optical properties by calculating the

reflectance around transition region taking the ionic contribution to the dielectric function also into account. We find that the reflectance as a

function of frequency shows very different kind of behaviour on both sides of the metal–insulator transition.

q 2005 Elsevier Ltd. All rights reserved.

PACS: 71.30.Ch; 71.20.Mq; 71.20.Nr

Keywords: D. Dielectric properties; D. Critical phenomena; A. Electronic materials; D. Optical properties

1. Introduction

Metal–insulator transition in solids under variation of

lattice parameter has been a subject of great theoretical

interest for a long time [1,2]. According to Mott, solids with

odd number of valence electrons like H, Na, etc. will

undergo transition from metal to insulator on lattice

expansion due to electron correlation effects [3]. Solids

with even number of electrons per unit cell can also show

metal insulator transition due to opening or closing of band

gaps at the Fermi level when their lattice parameter is varied

[3,4]. In this paper we study one such transition, which

occurs through the band overlap process and analyse

dielectric anomalies around the transition. Since one is

dealing with a quantum phase transition, one expects certain

universal features, which arise due to large scale spatial and

0022-3697/$ - see front matter q 2005 Elsevier Ltd. All rights reserved.

doi:10.1016/j.jpcs.2005.03.005

* Corresponding author.

E-mail address: [email protected] (R. Prasad).

temporal quantum fluctuations that accompany a continuous

zero temperature transition [1,5,6].

There has been a tremendous amount of work toward

understanding the critical behaviour around metal–insulator

transition (MIT). The first set of ideas were introduced

through the work of Wegner [7] and Abrahams et al. [8] in

the context of disorder-induced metal insulator transition

(MIT) for noninteracting electrons. Since interactions are

inevitably present in any physical system, the work

following these pioneering papers has naturally focussed

on quantitative understanding of MIT in interacting systems

[1,2,9,10]. Much of this activity has focussed on Mott–

Hubbard transition induced by short ranged interaction as in

Hubbard Model. Our aim in this work is to complement such

studies by analysing a MIT associated with the closing of a

one-particle gap induced by pressure in a realistic system.

We should emphasize that this transition is being driven by

the one-electron properties of the periodic lattice system.

Though both interactions and disorder play a significant role

in determining the properties of the transition, neither

actually drives the transition. Nevertheless this MIT is

associated with many dielectric anomalies, which we

Journal of Physics and Chemistry of Solids 66 (2005) 1150–1157

www.elsevier.com/locate/jpcs

Page 2: Dielectric critical behaviour at a metal–insulator transition under lattice compression

S. Shukla et al. / Journal of Physics and Chemistry of Solids 66 (2005) 1150–1157 1151

explore in the mean field spirit, in this paper. Such critical

properties have only been partially investigated in the

context of metal–insulator transitions induced by disorder or

interactions as in Mott–Hubbard models, due to inherent

limitations of the models studied. Thus our study provides a

broader base of physical properties that can be examined

theoretically and experimentally in the context of MIT.

In our formulation, the fundamental quantities are the

dielectric constant and the conductivity, which are exper-

imentally measurable quantities. This allows us to consider

long range Coulomb interaction, its screening and more

generally the critical behaviour associated with wavevector

and frequency dependent dielectric function. We apply the

principles of spatial and temporal scaling to it and obtain a

set of four exponents, which characterise the transition.

Scaling ideas have by now become a standard tool to

analyse continuous phase transitions [11]. Their application

to continuous quantum phase transitions has been recently

reviewed by Sondhi et al. [5]. Some key physical features of

the transition associated with opening (closing) of the one-

particle gap are different from the disorder-induced

Anderson transition, so there are natural differences in the

scaling relations formulated here from those in the work of

Wegner, Abrahams et al.

We have chosen to study Cesium Iodide solid which is

experimentally well known to show insulator–metal

transition under pressure [12]. Several first-principle

electronic structure calculations for CsI exist, which

support the transition by the band overlap mechanism

under lattice compression. CsI has a direct band gap, and

the calculations show that it undergoes the transition by

closing of the direct gap. At this point it is relevant to

comment on the possibility of the excitonic phase, which

may exist in this situation in which the MIT occurs due to

the closing of a band gap. Mott first pointed out that such a

transition cannot be continuous due to the following

reason. As the MIT is approached from the metallic side

the screening gets weaker due to decreasing density of free

electrons and holes, and at some point this gas becomes

unstable to the formation of bound pairs [3]. For our

purpose an argument due to Knox, who considered the

situation from the insulating side, is of direct relevance.

The binding energy of the exciton is given by e4mr=320 [13],

where mr is the reduced mass and 30 denotes the static

dielectric constant. The exciton level lies within the energy

gap. The key point is how this level changes as the gap

closes. If the binding energy becomes larger than the gap

there is an instability toward exciton formation. This

possibility which has been extensively studied can be

realized when 3 variation is not significant as can happen

when an indirect band gap closes. This leads to interesting

intervening phases with charge and spin density waves. In

our case, the dielectric constant diverges due to the closing

of a direct gap and one needs to consider the relative

variation of the energy gap and the exciton binding energy.

We shall take up this point in a later section.

We have done ab initio calculations of various quantities

like dielectric constant, band gap, conductivity, density of

states at Fermi level, etc. for Cesium Iodide in CsCl

structure around the transition region. The calculations are

done in the local density approximation, so that the

exchange and correlation effects have been treated in a

mean field approximation. We have also included

disorder in our calculations of dc conductivity as well as

the frequency dependent dielectric constant. However, the

disorder has also been included in the Drude like

approximation. So any effects like localization due to

disorder or weak-scattering corrections have been ignored.

From the analysis of the data, we have extracted the critical

behaviour of the quantities associated with the dielectric

response. The approximate values of various exponents

characterising the transition have been determined. Since in

our calculations the interactions are treated at a mean field

level, we expect the exponents to be simple fractions. Thus

we believe that the simplest fractions closest to our

numerical results are the correct exponents. We also explore

the effect of dielectric anomalies on infra-red (IR) optical

properties by calculating the reflectance in the transition

region. As the incident photons also interact with the

vibrational modes of the ionic crystal we take ionic

contribution to dielectric function also into account. This

is found to yield interesting features in the spectrum which

should be observed in optical measurements.

The organisation of this paper is as follows. In Section 2,

we deal with anomalies in the dielectric function and its

scaling analysis near the transition. This is followed by

Section 3 containing details of our calculation. In Section 4,

we present our results and its discussion. Finally, we

conclude this work in Section 5.

2. Scaling analysis for the dielectric function

The wavevector-frequency dependent dielectric function

of a solid, 3ððq;u; aÞ provides a unified description of charge

response of both metals and insulators. Here a denotes the

lattice constant, and is the drive parameter for the transition.

We first consider static dielectric function 3ððq; 0; aÞ. In

metals, due to screened coulomb interactions, 3ððq; 0; aÞdiverges as (1/L2q2) in limit ðq/ ð0. Here ‘L’ is the

screening length, which in Thomas Fermi Approximation

is given by LK2 Z4pe2NðEfÞ, where N(Ef) is the density of

states at Fermi level, Ef. In insulators as ðq/ ð0, 3ððq; 0; aÞgoes to a finite limit 30(a), the static dielectric constant. In

approaching the critical point denoted as ac, L and 30 are

expected to diverge according to power laws in

tZ ðaKac=acÞ. To relate power laws on the two sides of

the transition, we follow the scaling arguments due to Imry

et al. [14], which were developed in the context of the

disorder-induced MIT (Anderson transition). Here the basic

assumptions are that (i) there exists a characteristic

correlation length, x in each phase which diverges as

Page 3: Dielectric critical behaviour at a metal–insulator transition under lattice compression

S. Shukla et al. / Journal of Physics and Chemistry of Solids 66 (2005) 1150–11571152

the transition is approached,

x Z x0jtjKn (1)

(ii) the static dielectric function obeys the homogeneity

relation,

3M=Iðq; 0; aÞ Z x2KhfM=IðqxÞ (2)

where the subscripts M and I refer to metallic and insulating

phases, respectively. Behaviour of f(x) for small x in the

metallic and insulating phases can be obtained from the

above considerations; fMðxÞf ð1=x2Þ for a metal and fI(x)fconstant for an insulator. To obtain large x behaviour, we

adopt the physical argument proposed by Imry et al. [14]

that at length scales much smaller than x, the dielectric

behaviour in the two phases should be independent of x.

This requires that at large x, fM=IðxÞf ð1=x2KhÞ. From the two

asymptotic behaviours one obtains the divergence beha-

viours for screening length and static dielectric function in

the critical region as

L2ðxÞfxh f jtjKnh (3)

and

30ðaÞfx2Khf jtjKnð2KhÞ (4)

Next we consider the frequency dependent properties.

The scaling with respect to frequency should be important,

as the transition is associated with an energy scale related to

the band gap. The imaginary part of dielectric function, 3 00 is

related to the real part of conductivity s 0 according to

300ððq;u; aÞ Z4ps0ððq;u; aÞ

u(5)

Let us consider various quantities at zero wavevector and

in small frequency limit. While limu/0s 0(0,u,a)Zsdc, the

dc conductivity in metals, in insulators one finds that

3 00(0,u,a) vanishes as Au in the small frequency limit when

a finite lifetime to the electronic states is allowed. We

assume, following the principles of dynamical scaling, [11]

that (i) there exists a frequency scale, U, in each phase

which vanishes as the critical point is approached according

to a power law in t, and may be written as,

U Z U0xKz (6)

where z is termed as the dynamical exponent, (ii) the

dielectric function satisfies the homogeneity relation with

respect to the frequency scale U,

300M=Ið0;u; aÞ Z UygM=I

u

U

� �(7)

In order to conform to the low frequency behaviour of the

conductivity and the dielectric constant as discussed above,

as x/0, gI(x)fx for insulators, and gMðxÞf ð1=xÞ for

metals. To determine the large x behaviour of g(x), we again

invoke the argument that for frequencies, u[U the

dielectric behaviour for both the phases is independent of

U. This specifies the form gM/I(x)fxy in large x limit.

It follows from the two asymptotic behaviours that the

divergence of the dielectric function while approaching the

critical region from insulating side is according to,

300ð0;u; aÞfuUyK1 fujtjKnzð1KyÞ (8)

Similarly the conductivity behaviour on approaching

transition from metallic side is obtained as

sdcðaÞfUyC1f jtjnzðyC1Þ (9)

Clearly these conclusions require K1!y!1 to be

satisfied. The four exponents n, h, y, and z fully characterise

the metal–insulator transition driven by lattice parameter.

At this point it is pertinent to mention that the above

discussion is specific to transition driven by the change of

lattice constant and for systems with some disorder so that

the conductivity is finite in the metallic phase. We envisage

a situation of weak disorder so that the transition is not

driven by disorder in three-dimension. Thus our analysis is

different from the scaling analysis for pure systems in which

the transition is driven by interaction parameter or by the

filling fraction [1,2] or the disorder driven-Anderson

transition [7,8].

3. Method of calculation

The macroscopic dielectric function of the crystal is

calculated ab initio using the expression (for uO0) [15]

300ðuÞZ4pe2

m2u2

Xðk ;i;j

jhi;ðkjpxjðk; jij2f

iðk ð1Kfjðk ÞdðEjðk KE

iðk KZuÞ

(10)

It gives the diagonal term of the imaginary part of the

dielectric function for the crystal in RPA, ignoring the

local field effects. Here i and j indicate the band indices, ðkthe wavevector, jikO is the eigenstate with wavevector ðkand band index i, E

iðk , the corresponding eigenvalue, and

fiðk is the occupation number, ‘px’ the x component of

momentum transfer. Through Kramer–Kronig inversion

of 300(u), the calculation of dielectric constant, 30Z30(u)ju/0, where 30(u) is the real part of dielectric

function, becomes possible with an accuracy limited by

the frequency range of our calculation. We allow a finite

life-time ‘GK1’ to the transition between states, by

replacing the delta function in the above expression by

a Lorentzian of width G in the numerical calculation. In

realistic systems, one expects the single-particle states to

have finite lifetimes. The contribution to the life-times

come from disorder as well as electron–electron inter-

actions [16]. The latter are ignored in the RPA

approximation. This procedure was introduced earlier by

Philipp and Ehrenreich [17], and was reasonably success-

ful in achieving agreement with the optical data in some

semiconductors.

Page 4: Dielectric critical behaviour at a metal–insulator transition under lattice compression

S. Shukla et al. / Journal of Physics and Chemistry of Solids 66 (2005) 1150–1157 1153

To calculate the dc conductivity, we use Eq. (5) and take

the limit u/0 in Eq. (10) to obtain [18]

sdc Ze2

3G

Xn;ðk

v2

n;ðkdðE

n;ðk KEfÞ (11)

where ðvn;ðk Z ðVE

n;ðk , is the group velocity for the eigenstate

with quantum numbers n; ðk. Note that in this approximation,

the single-particle relaxation rate instead of the transport

relaxation rate occurs in the expression for the conductivity.

Here this only changes the conductivity by a constant factor.

In addition to these quantities, we calculate the direct

band gap, Eg, and the density of states at the Fermi level,

N(Ef), in insulating and metallic sides, respectively. For all

the numerical calculations the Full Potential Linearised

Augmented Plane Wave method [19] incorporated in

WIEN2k package by Blaha and co-workers [20] is used.

The method is a highly accurate scheme for electronic

structure calculations in solids with the following principal

features (i) The unit cell is divided into two parts: atomic

spheres and interstitial regions, which are described by

atomic like functions and plane waves, respectively. A

suitable linear combination of them is used as the basis set

for expanding Kohn–Sham orbitals. (ii) A modified

tetrahedron method is used for the Brillouin zone inte-

grations (iii) Exchange-correlation effects are treated within

the density functional theory using the Local Spin Density

Approximation (LSDA) [21].

4. Results and discussion

4.1. Electronic properties

Cesium Iodide has a high compressibility and a band gap

(w6 eV), which is the smallest among the alkali halides. CsI

Fig. 1. Band structure of CsI along principal symmetry directions at lattice consta

begins (aZ6.59 a.u.).

has also been subjected to experimental study of metalliza-

tion under pressure. During compression intermediate

phases with different structural symmetries occur at

different pressures. But it is shown experimentally as well

as supported theoretically, that even allowing for structural

transformations, CsI becomes metallic due to overlap of

filled 5p like Iodine band with the empty 5d like Cesium

band at the G—point in the Brillouin zone at a pressure

z100 GPa [12,22,23]. This is equivalent to a volume ratio

of z.50. Such a symmetric and controlled change in

electronic structure of CsI with pressure, in spite of the fact

that ionic character decreases on compression, makes it

ideally suited to our purpose.

Our results for the band structure calculation are shown

in Fig. 1. The calculations have been done in a relativis-

tically self-consistent manner by varying the lattice

constant. The two panels on left and right show the energy

band dispersion along the principal symmetry axes in

the Brillouin zone at equilibrium (8.63 a.u.) and at the

point of metal–insulator transition (MIT) (6.59 a.u.).

The equilibrium band gap is underestimated (3.6 eV)

compared to its actual value of 6.3 eV. This is not surprising

since it is well known that the DFT-LDA always under-

estimates the band gap. The MIT occurs at a lattice constant

equal to 6.59 a.u. This is equivalent to a metallization

volume of 0.44 of the uncompressed volume which is in

good agreement with previous calculations [12,22–24]. It is

interesting to note that the transition volume obtained here is

rather close to the Herzfeld’s polarization catastrophe

criterion, which gives a fractional volume of 0.42.

The results of the calculations of 3 00(u) at small

frequencies in lattice parameter range 6.63–6.605 a.u. are

shown in Fig. 2. The linear behaviour with the frequency (in

the limit u/ 0) and the increasing slope with the approach

to the critical lattice parameter is clearly seen in the figure.

The variation of the slope A in the linear region with t, the

nts corresponding to (a) equilibrium (aZ8.63 a.u.) (b) when metallisation

Page 5: Dielectric critical behaviour at a metal–insulator transition under lattice compression

6.63 a.u.

6.61 a.u.

6.62 a.u.

6.615 a.u.

ω (eV)

ε"(ω

)

6.605 a.u.

0

2

4

6

8

10

0 0.2 0.4 0.6 0.8 1

Fig. 2. Imaginary part of dielectric function of CsI versus frequency for

lattice parameters between 6.605 a.u. (top curve) and 6.63 a.u. (last curve at

the bottom) for GZ0.02 eV.

t (10–3 units)

Eg

(10–3

Ryd

)

2

6

10

14

18

22

3 5 7 9 11 1

Fig. 4. Band gap variation as a function of t. The dashed line shows the

power law fit to t.92.

σ dc

(a)

0.2

0.4

0.6

S. Shukla et al. / Journal of Physics and Chemistry of Solids 66 (2005) 1150–11571154

distance from the critical point is shown in Fig. 3a. The

analysis of this data (Fig. 3a) shows a power-law variation,

A(a)fjtjKx. In order to confirm this behaviour and

determine the power law exponent, x, we have done this

calculation by varying the dissipation parameter, G, from

0.1 to 0.002 eV. We find that the exponent increases with

decreasing G, but reaches an asymptotic constant value of

2.40G.03 at GZ0.02 eV. We take this to be the correct

value in consistency with the RPA approximation. The error

limit for the exponents quoted here are the ranges over

which we obtain the best fits of the available data. By using

the dispersion relation, we have also estimated the static

dielectric constant. Its variation with t is shown in Fig. 3b.

The exponent associated with its divergence is .54G.02.

However, this value has a bigger error margin than other

exponents obtained in our computation due to a finite

frequency range of integration in the dispersion relation. In

Fig. 4, we plot the band gap on insulating side around

transition point with respect to critical point distance. It is

seen to vanish as jtj.92 with error limits of the exponents G.02. Note that the value of the exponent for the static

A(a

)

ε 0

t (10–3 units)

(a)

(b)

0

4

8

12

16

4 6 8 10 2 0

4

12

16

8

Fig. 3. (a) Slope of the curves shown in Fig. 2 in region u/0 as a function

of tZ ðjaKacj=acÞ. The dashed line shows the power law fit by tK2.4. Note

the Y-scale for this curve is on left as shown by arrow. (b) Dielectric

constant of CsI as a function of tZ jaKacj=ac. The dashed line shows the

power law fit by tK.54. Y-scale for this curve is on the right.

dielectric constant is inconsistent with the expectation based

on the Penn’s relation 30fEK2g [26].

In Fig. 5a, we present the results for the calculation of the

dc conductivity as a function of lattice constant. One finds

that it also has a power law behaviour sdcfjtj2.1 with G.12

as the error on exponent. The variation of density of states

(DOS) at Fermi level on metallic side is shown in Fig. 5b,

and it varies around the critical point as jtj1.1 within G.14

limits of exponent. It is interesting to note that the exponents

for the conductivity and DOS are different. In this transition,

which is driven by band overlap and hence the DOS at

Fermi level, one expects the conductivity to follow

the exponent for DOS. However, the difference in

t (10-3 units)

(b)

N(E

f) (

10–2

Sta

tes/

eV)

0

1

2

3

4

5

8 12

0

4 0

Fig. 5. (a) Static conductivity as a function of t. The dashed line shows the

power law fit by t2.1. (b) Density of states as a function of t. The dashed line

shows the power law fit by t1.1.

Page 6: Dielectric critical behaviour at a metal–insulator transition under lattice compression

S. Shukla et al. / Journal of Physics and Chemistry of Solids 66 (2005) 1150–1157 1155

the behaviours is considerable. According to Eq. (11)

sdc f hv2n;kiNðEfÞ, N(Ef) being density of states at Fermi

level. If group velocity hv2n;ki did not vary as Fermi surface

vanishes, conductivity is indeed proportional to DOS. But

when the bands just overlap and give rise to small Fermi

surfaces, hv2n;ki does seem to depend on the Fermi surface

size as our computation in the transition region indicates.

The results therefore indicate that the mobility variation in

the transition region is also crucial even for the band overlap

transition.

Let us recall that the calculations have been performed in

the LDA which may not precisely predict the position of

metal–insulator transition and the exact behaviour of

various quantities around it. The many body correlation

effects [16] especially in the low energy range may strongly

affect the electronic band structure in the critical region.

Thus keeping in mind the mean-field character of the LDA

with regard to electron–electron interactions, one expects

integer or simple fractional values for these exponents. Our

numerical accuracy in the determination of exponents is

obviously limited by the following factors. On the metallic

side, in the transition region the Fermi surface becomes very

small and we expect errors in the number of k-points used in

sampling it. Similarly on the insulating side the energies in

the denominator become very small, and errors due to finite

grid of k-points are quite likely. Furthermore, the critical

behaviour is limited to a rather small range of lattice

constants and thus the number of points available to

determine power law exponents is small. In view of these

possible errors, it would be well within the accuracy of our

calculations to take the nearest rational exponents to be the

correct values. The values for the various exponents

determined in the calculation along with their closest

rational approximations are summarized in Table 1. From

this data we have also obtained the exponents for the

divergence of the length and frequency scales introduced

above. These are also listed in Table 1. Note that the

analysis of the Drude weight and the compressibility in case

of transitions controlled by filling yields the correlation

length exponent n!(1/2) and the dynamical exponent zO2

in correlated systems [1].

We can return now to the question of the existence of

intervening excitonic phase in the present situation.

Spontaneous formation of excitons is favoured if the

binding energy of the electron–hole pair exceeds the band

gap as the band gap closes. The exciton binding energy

goes as the inverse square of 30 [13]. Given that the

exponent of divergence of 30 is .54 while band gap

vanishes as t.92, it is seen that the binding energy decreases

Table 1

Power law dependence of A, 30, s, N(Ef) and Eg on t and the values of critical ex

A 30 s N(Ef) Eg

K2(K2.4) K(1/2)(K.54) 2(2.1) 1(1.1) 1(.92)

Values obtained from the actual calculations are given in the brackets while thos

a little faster than the gap. Thus the excitonic phase and the

associated discontinuous transition is marginally excluded

in this case. However, since the exponents have been

obtained from a mean field type calculation and the

excitonic binding energy scales approximately same as the

band gap under lattice compression, resolving the issue of

excitonic instability requires a more subtle and rigorous

study.

4.2. Reflectance studies

We explore optical properties around metal–insulator

transition by calculating reflectivity and examine its relation

with divergence of dielectric function. We restrict ourselves

to IR region of spectrum in which total dielectric function

has contribution from both electrons and ions (eZelectronic, iZionic):

3ðuÞ Z 3eðuÞC3iðuÞ (12)

The reflectance for the normal incidence is given by the

well-known Fresnel’s formula [27],

RðuÞ ZðnðuÞK1Þ2 Ck2ðuÞ

ðnðuÞC1Þ2 Ck2ðuÞ(13)

where nðuÞC ikðuÞZffiffi3

pðuÞ is the complex refractive index

function.

Let us take insulating and metallic cases separately.

4.2.1. Insulating phase

The total dielectric function in insulating phase at small

frequencies can be given as

3ðuÞ Z 30 C iAu CU2

p

ðU2T Ku2Þ

(14)

The first two terms constitute the electronic part in limit

u/0 while the third term is the ionic contribution ignoring

its absorptive part. Here Up is the ionic plasma frequency,

Up Z4pn

Me*2

T (15)

in which n is the number of dipoles per unit volume, M the

reduced mass, and e*T the effective charge including the

deformation from the deformable shells of the electrons. UT

is the transverse harmonic frequency of the phonon, which

couples to the light. To find Up at lattice parameters away

from the equilibrium value we use the relation [28]

e*T Z

1

3ð30 C2Þe*

s (16)

ponents

n h y z

3/4(.82) 4/3(1.34) 0(K.06) 8/3(2.74)

e outside it are the rationalised ones.

Page 7: Dielectric critical behaviour at a metal–insulator transition under lattice compression

6.54 a.u.

6.56 a.u.

6.58 a.u.

ω (eV)

R (ω

)

0.4

0.6

0.8

1

0 0.02 0.04 0.06 0.08 0.1

Fig. 7. Reflectivity spectra for CsI at different lattice constants on the

metallic side of the transition.

6.66 a.u.

6.62 a.u.

6.605 a.u.

ω (eV)

R (ω

)

0

0.2

0.4

0.6

0.8

0 0.02 0.04 0.06

Fig. 6. Reflectivity spectra for CsI at different lattice constants on the

insulating side of the transition.

S. Shukla et al. / Journal of Physics and Chemistry of Solids 66 (2005) 1150–11571156

where e*s is the Szigetti charge whose variation with volume

V is given as

d log e*s

d log Vx:6 (17)

Similarly the variation of phonon frequency UT with

volume can be obtained from the relation[28]

d log UT

d log VZKg (18)

where the Gruneisen parameter gZ2 for cesium halides at

TZ0. Knowing the variations of 30, A, Up, UT as a function

of volume, we calculate reflectance using Eq. (14) for the

dielectric function. Fig. 6 shows the infra-red reflectivity

spectra at lattice parameters aZ6.66, 6.62, and 6.605 a.u.,

respectively. It is seen to have a strong peak at uZUT,

where 3(u) possesses a real pole. This is the well known

Reststrahalen region in which there occurs strong inter-

action of the photon with the transverse optic (TO) phonon

as they are in resonance. In this case, the major contribution

to dielectric function comes from the lattice part and it

follows from Eq. (13) that reflectance R/1.

The behaviour in the region u!UT can be inferred from

the following analysis. We can write

3ðuÞx30 C iAu CU2

p

U2T

(19)

In the small u regime, where Au/ 30CU2

p

U2T

� �, the

refractive index n[k, and the reflectivity R may be

approximated as

Rx1 K4n

ðn C1Þ2(20)

For example at aZ6.605 a.u., 30Z16.07, U2p=U

2T x:5

yields value of R(u)Z.37, which agrees well with the

calculation. R(u) is nearly independent of u over a broad

range below UT as seen in the calculations. The reflectance

increases towards unity as the MIT boundary is approached.

4.2.2. Metallic phase

In metallic phase, small frequency behaviour of elec-

tronic dielectric function is dominated by intraband

contribution, which is basically conductance. Including

the lattice part, it is given as

3ðuÞ Z 30 C i4psdc

uC

U2p

ðU2T Ku2Þ

(21)

and sdc is the static conductivity. Here we also assume that

even in the metallic phase of this ionic solid, the optical

phonon exists with the volume variation of the parameters

as described above. Taking the values of sdc from the ab

initio calculations, we obtain reflectivity using above

expression (21) for dielectric function. The results for the

metallic phase at lattice parameters aZ6.58, 6.56, and

6.54 a.u. of CsI are displayed in Fig. 7. Note that the

behaviour of R(u) for metallic phase is very different from

Fig. 6 for insulating phase particularly near uZUT.

The reflectance is seen to decrease as the MIT is

approached from the metallic side. This is in accord with the

fact that higher conductivity implies good reflectivity. In the

region u!UT!sdc, the behaviour of R is of Hagen–Ruben

kind in respect to frequency as well as the static

conductivity:

R Z 1 K2u

psdc

� �1=2

(22)

This is well obeyed by the detailed computation. Below

the electronic plasma frequency, the TO phonon frequency

again gives rise to a feature in the reflectance spectrum. At

uZUT, this spectrum also shows a pronounced peak.

Interestingly the peak gets widened with the decrease in

lattice parameter and finally appears to coalesce with the

Hagen–Ruben behaviour. This is understandable, since at

higher compression free electron like character supercedes

the role of lattice vibration. In this region, with loss of polar

character of CsI, the free electron Drude absorption

becomes as important as the absorption due to resonant

optical mode.

Page 8: Dielectric critical behaviour at a metal–insulator transition under lattice compression

S. Shukla et al. / Journal of Physics and Chemistry of Solids 66 (2005) 1150–1157 1157

R(u) for uOUT but !up (upZelectronic plasma

frequency) saturates to a constant value. This can be

understood by noting that the dielectric function in this

region can be written in the form,

3ðuÞ Z 1 Ku2

p

uðu C iGÞC

U2p

ðU2T Ku2Þ

(23)

where we have used the Drude form for the conductivity.

Then for large frequencies, the reflectivity R is approxi-

mately given as

Rx1 K2G

up

(24)

which matches the calculation. It becomes larger with

increase in up i.e. enhanced metallicity.

5. Summary and concluding remarks

We have examined the critical behaviour of the

wavevector and frequency dependent dielectric function

around the lattice-driven metal–insulator transition in

Cesium Iodide. We have obtained exponents for the

critical behaviour of conductivity, imaginary part of

dielectric function in the small frequency limit, screening

length and band gap. We have discussed these results in

terms of the ideas of dynamical scaling appropriate to this

quantum phase transition. The exponents obtained here

are mean field like as (a) the exchange-correlation effects

are treated in the local density approximation; (b) the

effect of disorder has been included in the conductivity in

the Drude approximation, and in the frequency dependent

calculation by allowing for a level-independent broad-

ening. We have also examined the possibility of the

existence of an excitonic phase. Our present estimate of

exponents barely exclude the exciton phase, however,

given the inaccuracy of these exponents, this question

needs further analysis.

We have also studied IR reflectivity in the transition

region as an observable property, which is strongly affected

by the dielectric anomalies. In this analysis we include the

effects due to ionic part coming from the optical phonon and

find an interesting feature at the resonance whose obser-

vation should provide interesting knowledge about the

critical dielectric behaviour.

Acknowledgements

We are indebted to Prof. Deepak Dhar for some useful

remarks on this work. We also gratefully acknowledge

financial support from the Department of Science and

Technology, New Delhi under project Nos. SP/S2/M-46/97

and SP/S2/M-51/96.

References

[1] For a review see, for example M. Imada, A. Fujimori, Y. Tokura,

Rev. Mod. Phys. 70 (1998) 1039.

[2] M. Imada, J. Phys. Soc. Jpn 64 (1995) 2954.

[3] N.F. Mott, Phil. Mag. 6 (1961) 287.

[4] G.J. Ackland, Rep. Prog. Phys. 64 (2001) 483.

[5] S.L. Sondhi, S.M. Girvin, J.P. Carini, D. Sahar, Rev. Mod. Phys. 69

(1997) 315.

[6] M.A. Continentino, Phys. Rep. 239 (1994) 179.

[7] F.J. Wegner, Z. Phys, Z. Phys. 25 (1976) 327.

[8] E. Abrahams, P.W. Anderson, D.C. Licciardello, T.V. Ramakrishnan,

Phy. Rev. Lett. 42 (1979) 673.

[9] W.L. McMillan, Phys. Rev. B 24 (1981) 2739.

[10] D. Belitz, T.R. Kirkpatrick, Rev. Mod. Phys. 66 (1994) 261.

[11] P.C. Hohenberg, B.I. Halperin, Rev. Mod. Phys. 49 (1977) 435.

[12] Q. Williams, R. Jeanloz, Phys. Rev. Lett. 56 (1986) 163.

[13] B.I. Halperin, T.M. Rice, Rev. Mod. Phys. 40 (1968) 755.

14 Y. Imry, Y. Gefen, D.J. Bergmann, Phys. Rev. B 26 (1982) 3436;

Y. Gefen, Y. Imry, Phys. Rev. B 28 (1983) 3569.

[15] S.L. Adler, Phys. Rev. 126 (1962) 413.

[16] A. Bardasis, D. Hone, Phys. Rev. 153 (1967) 849.

[17] H.R. Philipp, H. Ehrenreich, Phys. Rev. 129 (1963) 1550.

[18] S. Doniach, E.H. Sondheimer, Green’s Functions for Solid State

Physicists, Imperial College Press, London, 1998.

[19] D. Singh:, Plane Waves, Pseudopotentials and LAPW method,

Kluwer, Boston, 1994.

[20] P. Blaha, K. Schwarz, G.K.H Madsen, D. Kvasnicka, J. Luitz,

WIEN2k, An Augmented Plane WaveCLocal Orbitals Program for

Calculating Crystal Properties (Wien, Austria: Karlheinz Schwarz,

Techn. Universitat Wien, 2001) ISBN 3-9501031-1-2.

[21] Fora review see, for example W. Kohn, P. Vashishta, in: S. Lundqvist,

N.H. March (Eds.), Theory of the Inhomogeneous Electron Gas,

Plenum, New York, 1983.

[22] J. Aidun, M.S.T. Bukownsky, M. Ross, Phys. Rev. B 29 (1984) 2611.

[23] S. Satpathy, N.E. Christensen, O. Jepsen, Phys. Rev. B 32 (1985)

6793.

[24] Although density functional eigenvalues do not represent actual

excitation energies, the estimation of metallization is expected to be

reasonable [25].

[25] S. Wei, H. Krakauer, Phys. Rev. Lett. 55 (1985) 1200.

[26] D.R. Penn, Phys. Rev. 128 (1962) 2093.

[27] C. Kittel, Introduction to Solid State Physics, fifth ed., Wiley,

New York. (1976).

[28] G. Martinez, T.S. Moss (Eds.), Handbook on Semiconductors, North-

Holland Publishing Company, 1980.