7
APS Specific Heat Anomalies in Solids Described by a Multilevel Model Mariano de Souza, 1, * Ricardo Paupitz, 1 Antonio Seridonio, 1 and Roberto E. Lagos 1 1 IGCE, Unesp - Univ Estadual Paulista, Departamento de F´ ısica, 13506-900, Rio Claro, SP, Brazil Specific heat measurements constitute one of the most powerful experimental methods to probe fundamental excitations in solids. After the proposition of Einstein’s model, more than one century ago (Annalen der Physik 22, 180 (1907)), several theoretical models have been proposed to describe experimental results. Here we report on a detailed analysis of the two-peak specific heat anomalies observed in several materials. Employing a simple multilevel model, varying the spacing between the energy levels Δi =(Ei - E0) and the degeneracy of each energy level gi , we derive the required conditions for the appearance of such anomalies. Our findings indicate that a ratio of Δ21 10 between the energy levels and a high degeneracy of one of the energy levels define the two- peaks regime in the specific heat. Our approach accurately matches recent experimental results. Furthermore, using a mean-field approach we calculate the specific heat of a degenerate Schottky-like system undergoing a ferromagnetic (FM) phase transition. Our results reveal that as the degeneracy is increased the Schottky maximum in the specific heat becomes narrow while the peak associated with the FM transition remains unaffected. PACS numbers: 64.60.A-, 65.40.Ba, 65.60.+a, 71.27.+a I. INTRODUCTION In the field of condensed matter Physics, specific heat measurements can be considered as a pivotal experi- mental technique for characterizing the fundamental ex- citations involved in a certain phase transition. In- deed, phase transitions involving spin 1,2 , charge 3 , lattice 4 (phonons) and orbital degrees of freedom, the inter- play between ferromagnetism and superconductivity 5 , Schottky-like anomalies in doped compounds 6 , electronic levels in finite correlated systems 7 , among other features, can be captured by means of high-resolution calorime- try. Furthermore, the entropy change associated with a first-order phase transition, no matter its nature, can be directly obtained upon integrating the specific heat over T , i.e. C(T )/T , in the temperature range of interest. In his seminal paper of 1907 8 , in order to explain the devia- tion of the specific heat of certain materials, like silicon, boron and carbon, from the Dulong-Petit’s law, Einstein proposed a model based on the assumption that all atoms in a solid vibrate independently from each other with the same eigen-frequency. The well-known Einstein’s expres- sion for the phononic specific heat at constant volume, C E ph,v (T ), is given by: C E ph,v (T )=(βhν ) 2 e βhν (e βhν - 1) 2 , (1) hereafter all extensive quantities are defined as per parti- cle, except when otherwise indicated; β = 1/T ,(k B =1 hereafter), ν is the eigen-frequency of the oscillator and h the Planck’s constant. The quantity is the so-called Einstein temperature Θ E . It is worth mentioning that recently, one of us made use of Einstein’s model to de- termine the eigen-energy of the counter-anions libration modes in a molecular conductor 9 . However, thought the success of the model proposed by Einstein, the theoretical description introduced by P.Debye in 1912 10 , constitutes 0 50 100 150 200 0 50 100 150 T (K) C/T (arbitrary units) T (K) C (arbitrary units) FIG. 1. Main panel: Specific heat (C) as a function of tem- perature for an hypothetical Debye’s solid (cf. Eq. 2) with ΘD = 120 K. Inset: C/T versus T showing a maximum centered at T 0.28·ΘD. the hallmark in the description of the phononic contribu- tion to the specific heat in solids, see e.g. 11 . Essentially, Debye’s model is based on the hypothesis of a continuous isotropic solid. The dispersion relation is linear, i.e. the sound velocity is constant and isotropic, being a set of eigen-frequencies allowed to the oscillators. In the frame of Debye’s model, the phononic specific heat at constant volume, C D ph,v (T ), reads: C D ph,v (T )=3 T Θ D 3 Z x 0 x 4 e x (e x - 1) 2 dx, (2) where x D /T and Θ D is the so-called Debye tempera- ture. The behavior of the specific heat as a function of T , according to Eq. 2, well-known from textbooks, is shown arXiv:1504.07525v1 [cond-mat.mtrl-sci] 28 Apr 2015

Speci c Heat Anomalies in Solids Described by a Multilevel ... · Speci c Heat Anomalies in Solids Described by a Multilevel ... Unesp - Univ Estadual Paulista, Departamento de F

Embed Size (px)

Citation preview

Page 1: Speci c Heat Anomalies in Solids Described by a Multilevel ... · Speci c Heat Anomalies in Solids Described by a Multilevel ... Unesp - Univ Estadual Paulista, Departamento de F

APS

Specific Heat Anomalies in Solids Described by a Multilevel Model

Mariano de Souza,1, ∗ Ricardo Paupitz,1 Antonio Seridonio,1 and Roberto E. Lagos1

1IGCE, Unesp - Univ Estadual Paulista, Departamento de Fısica, 13506-900, Rio Claro, SP, Brazil

Specific heat measurements constitute one of the most powerful experimental methods to probefundamental excitations in solids. After the proposition of Einstein’s model, more than one centuryago (Annalen der Physik 22, 180 (1907)), several theoretical models have been proposed to describeexperimental results. Here we report on a detailed analysis of the two-peak specific heat anomaliesobserved in several materials. Employing a simple multilevel model, varying the spacing betweenthe energy levels ∆i = (Ei − E0) and the degeneracy of each energy level gi, we derive the requiredconditions for the appearance of such anomalies. Our findings indicate that a ratio of ∆2/∆1 ≈10 between the energy levels and a high degeneracy of one of the energy levels define the two-peaks regime in the specific heat. Our approach accurately matches recent experimental results.Furthermore, using a mean-field approach we calculate the specific heat of a degenerate Schottky-likesystem undergoing a ferromagnetic (FM) phase transition. Our results reveal that as the degeneracyis increased the Schottky maximum in the specific heat becomes narrow while the peak associatedwith the FM transition remains unaffected.

PACS numbers: 64.60.A-, 65.40.Ba, 65.60.+a, 71.27.+a

I. INTRODUCTION

In the field of condensed matter Physics, specific heatmeasurements can be considered as a pivotal experi-mental technique for characterizing the fundamental ex-citations involved in a certain phase transition. In-deed, phase transitions involving spin1,2, charge3, lattice4

(phonons) and orbital degrees of freedom, the inter-play between ferromagnetism and superconductivity5,Schottky-like anomalies in doped compounds6, electroniclevels in finite correlated systems7, among other features,can be captured by means of high-resolution calorime-try. Furthermore, the entropy change associated with afirst-order phase transition, no matter its nature, can bedirectly obtained upon integrating the specific heat overT , i.e.C(T )/T , in the temperature range of interest. Inhis seminal paper of 19078, in order to explain the devia-tion of the specific heat of certain materials, like silicon,boron and carbon, from the Dulong-Petit’s law, Einsteinproposed a model based on the assumption that all atomsin a solid vibrate independently from each other with thesame eigen-frequency. The well-known Einstein’s expres-sion for the phononic specific heat at constant volume,CEph,v(T ), is given by:

CEph,v(T ) = (βhν)2 eβhν

(eβhν − 1)2, (1)

hereafter all extensive quantities are defined as per parti-cle, except when otherwise indicated; β = 1/T , (kB = 1hereafter), ν is the eigen-frequency of the oscillator andh the Planck’s constant. The quantity hν is the so-calledEinstein temperature ΘE . It is worth mentioning thatrecently, one of us made use of Einstein’s model to de-termine the eigen-energy of the counter-anions librationmodes in a molecular conductor9. However, thought thesuccess of the model proposed by Einstein, the theoreticaldescription introduced by P. Debye in 191210, constitutes

0 5 0 1 0 0 1 5 0 2 0 00 5 0 1 0 0 1 5 0

T ( K )

C/T (a

rbitra

ry un

its)

T ( K )

C (arb

itrary

units)

FIG. 1. Main panel: Specific heat (C) as a function of tem-perature for an hypothetical Debye’s solid (cf. Eq. 2) with ΘD

= 120 K. Inset: C/T versus T showing a maximum centeredat T ' 0.28·ΘD.

the hallmark in the description of the phononic contribu-tion to the specific heat in solids, see e.g. 11. Essentially,Debye’s model is based on the hypothesis of a continuousisotropic solid. The dispersion relation is linear, i.e. thesound velocity is constant and isotropic, being a set ofeigen-frequencies allowed to the oscillators. In the frameof Debye’s model, the phononic specific heat at constantvolume, CDph,v(T ), reads:

CDph,v(T ) = 3

(T

ΘD

)3 ∫ x

0

x4ex

(ex − 1)2dx, (2)

where x = ΘD/T and ΘD is the so-called Debye tempera-ture. The behavior of the specific heat as a function of T ,according to Eq. 2, well-known from textbooks, is shown

arX

iv:1

504.

0752

5v1

[co

nd-m

at.m

trl-

sci]

28

Apr

201

5

Page 2: Speci c Heat Anomalies in Solids Described by a Multilevel ... · Speci c Heat Anomalies in Solids Described by a Multilevel ... Unesp - Univ Estadual Paulista, Departamento de F

2

in the main panel of Fig. 1. A particular behavior, usu-ally not explicitly shown and discussed in textbooks12,is observed by plotting C/T versus T , cf. inset of Fig. 1.The maximum centered at 0.28·ΘD corresponds to theinflexion point of Eq. 2 and manifests itself as a directeffect of the accessible energy levels upon increasing thetemperature. In other words, as the temperature of acertain Debye’s solid is increased the number of accessi-ble energy levels is reduced and, as a consequence, theentropy variation rate is reduced. The latter indicatesthat for any system which obeys a Debye-like behavior,ΘD can be directly estimated by plotting C/T versus T .Similarly in Einstein’s model (Eq. 1), C/T versus T hasa maximum centered at 0.38·ΘE . A combined system ofDebye and Einstein phonons can be studied provided theEinstein and Debye temperature scales are not so close13.

At low temperatures, namely for ordinary metals T 'ΘD/50 ' liquid 4He temperature, the following relationis valid11:

Cv(T ) = γT +AT 3, (3)

where γ is the Sommerfeld coefficient and A =12π4/5Θ3

D. Here the Debye temperature can be di-rectly obtained via A parameter, whereas the effectivemass m∗ of the charge carriers in a metal can be esti-mated by means of the Sommerfeld coefficient11. Thelatter can be considered the smoking gun, for instance,when discovering materials with heavy-fermion-like be-havior. Moreover, still considering the relevance of high-resolution calorimetry, combining Cp(T ) and the linearthermal expansion coefficient14, αi(T ), and making use ofthe Ehrenfest theorem, see e.g. 11, the uniaxial-pressuredependence of the critical temperature for pressure ap-plied along the i-axis for a second-order phase transitioncan be directly obtained, as follows:(

dTcdPi

)Pi→0

= Vmol · Tc ·∆αi∆C

, (4)

where Vmol is the molar volume, ∆αi and ∆C refer tothe thermal expansion and specific heat (per mol) jumpsat the transition temperature, respectively. The index irefers to the crystallographic direction, along which pres-sure is applied. Rigorously speaking, the Ehrenfest rela-tion is applicable only for mean-field-like phase transi-tion, where both ∆α and ∆C present step-like behavior.Note that the Ehrenfest relation enables us to determinethe pressure dependence of Tc purely via thermodynamicquantities, i.e., dTc/dPi can be estimated without car-rying out any experiment under application of externalpressure.

For a general two-level system, separated by an energygap ∆1, the specific heat is described by the followingexpression:

C =(β∆1)2e−β∆1

(1 + e−β∆1)2, (5)

Energy

E0

E1

E2

Δ1

Δ2

FIG. 2. Schematic representation of the energy levels E0, E1

and E2 considered in the model (Eq. 7). The excitation gaps∆1 and ∆2 are defined as the energy difference relative to theground state energy (E0). See discussion in the main text fordetails.

In such a system, the so-called Schottky anomaly man-ifests itself as a shallow maximum in the specific heatdata as a function of temperature.

After this brief introduction, recalling some funda-mental aspects related to specific heat measurements insolids, well-known from text books, see e.g. 11, below wepresent a multilevel model to describe systems whichpresent multiple peaks in the specific heat data. Interest-ingly enough, such a simple model is capable to describeexperimental electronic specific heat results of correlatedelectrons systems.

II. MATHEMATICAL MODELING

In the frame of the canonical ensemble the sum overstates, namely the partition function (Z) of a system, isgiven by:

Z ≡ Tr(e−βH) =∑n

e−βEn , (6)

where En refers to the nth energy level, i.e. the nth eigen-value of the system’s Hamiltonian H15. As a matter offact, the partition function encodes all information of thephysical system, so that once we know the energy eigen-values of the physical system all thermodynamic observ-ables can be calculated. This in turn is not the case ifone has the system Hamiltonian in hand16, being neces-sary to make its diagonalization to obtain Z and thenthe physical quantities of interest. For the sake of sim-plicity, in order to describe the two-peak like anomaliesobserved experimentally in several materials (see below),

Page 3: Speci c Heat Anomalies in Solids Described by a Multilevel ... · Speci c Heat Anomalies in Solids Described by a Multilevel ... Unesp - Univ Estadual Paulista, Departamento de F

3

FIG. 3. Specific heat (C) as a function of: (a) Temperature (T ) and degeneracy of the energy level 1, labelled g1; (b) T anddegeneracy of the energy level 2, g2; (c) T and energy gap ∆1; (d) T and energy gap ∆2. The projection of each curve shownin the 3D plots is depicted in the various panels. Details are discussed in the main text.

we propose a multilevel model approach. This Ansatzenables us to obtain directly the thermodynamic quanti-ties of interest. In other words, within such an approachwe work in the diagonal representation of the Hamilto-nian H (cf. Eq. 6). We start by considering a three-levelmodel17, with ∆1 = (E1−E0) and ∆2 = (E2−E0), beingE0, E1 and E2 defined as shown in Fig. 2, the partitionsum reads

Z = g0e−βE0 + g1e

−β(E0+∆1) + g2e−β(E0+∆2), (7)

where gi indicates the degeneracy of each energy level(i = 0, 1, 2). The assumption that the energy gaps ∆i,

i.e. energy scales for a generic system, do not depend onthe temperature is quite realistic. In this regard we refer,for instance, to the Schottky model11, where the energygap ∆1 separating the two energy-levels is fixed and thustemperature independent.

Making use of Eq. 7, one can calculate the specific heat,employing the following well-known relations

E = − ∂

∂β(lnZ) , (8)

and

Page 4: Speci c Heat Anomalies in Solids Described by a Multilevel ... · Speci c Heat Anomalies in Solids Described by a Multilevel ... Unesp - Univ Estadual Paulista, Departamento de F

4

C = −β2

(∂E

∂β

), (9)

obtaining

C = β2

∑i

∑j gigj∆i(∆i −∆j)e

−β(∆i+∆j)

[∑i gie

−β∆i ]2 , (10)

where i, j = 0, 1, 2 and ∆0 = 0. Note that considering j= 1 in Eq. 10, the specific heat for the Schottky model(Eq. 5) is nicely restored.

For the sake of completeness, we calculated the sum-mations in Eq. 10 and present below the expression em-ployed in our analysis:

C = β2 g1g0∆21e−β∆1 + g0g2∆2

2e−β∆2

[g0 + g1e−β∆1 + g2e−β∆2 ]2+

+g1g2e

−β(∆1+∆2)[∆1(∆1 −∆2) + ∆2(∆2 −∆1)]

[g0 + g1e−β∆1 + g2e−β∆2 ]2.

(11)

In the present model, the crucial feature refers to therelation between the energy scales of the two excitationgaps. The latter together with the degeneracies gi ofthe energy levels, define whether the system presents aSchottky-like behavior with a single peak/maximum inthe specific heat or if it will exhibit two or even three(in this case a four energy levels is required) peaks ormaxima. More specifically, one can say that in such casestwo and three energy scales govern the Physics of thesystem of interest.

III. RESULTS AND DISCUSSION

In what follows, employing Eq. 11, we discuss the re-quired conditions for the emergence of the two-peakanomalies in the specific heat. In Fig. 3 we present 3-dimensional (3D) plots of the specific heat as a func-tion of temperature and degeneracy of the energy levels1 (Fig. 3-a) and 2 (Fig. 3-b) and, specific heat as a func-tion of temperature and the relative value of the energylevels 1 (Fig. 3-c) and 2 (Fig. 3-d) as well. The optimizedparameters used in such set of fits, appropriate to giverise to the emergence of a double peak in the specificheat are ∆0 = 0, ∆1 = 205 J/mol, ∆2 = 2327 J/mol, g0

= 1, g1 = 2, g2 = 4, i.e. in each 3D plot one of these pa-rameters was varied while the remaining ones were keptconstant. For instance, in Fig. 3-a) T and g1 were usedas the variable parameters. Note that upon increasingthe degeneracy g1 the double peak in the specific heatgradually vanishes. A distinct situation is depicted inFig. 3-b), where the double peak in the specific heat ap-pears as the degeneracy g2 is increased. In Fig. 3-c) andd) we vary the temperature and the energy levels ∆1 and

∆2, respectively. The present findings indicate two dis-tinct ways of catching a double-peak in the specific heat,i.e. either by varying the degeneracy or by increasing theenergy levels ∆1 or ∆2. More specifically, Figs. 3-a) andb) indicate that at least one of the energy levels shouldbe highly degenerated to give way to the double-peak inthe specific heat. Similarly to the Schottky-like anomaly,this behavior suggest that the high degeneracy contributedramatically to the specific heat over a quite restrictedT -window. Such a behavior can be easily understood interms of the system entropy: once the number of acces-sible states is increased and this is the physical situationof a highly degenerated generic system or energy level,the entropy of the system is increased and, as a conse-quence, a double-peak structure shows up in the specificheat, since the entropy of the system is proportional to

the area and can be estimated via ∆S =∫ Tb

TaC/TdT ,

where Ta and Tb indicate two distinct temperatures inthe range of interest. In other words, the system entropyis increased at the expense of the satellite peak in the spe-cific heat. Nevertheless, as discussed above, the specificheat is a bulk property. Thus in order to map the energylevels of a generic system and its degeneracy, microscopicexperimental techniques like electron spin and magneticresonance18, for instance, are required. We stress thatthe two-maxima observed in the specific heat, obtainedfrom the above-discussed Ansatz, is a universal signa-ture of any system with two distinct characteristic energyscales that differ from each other at least by roughly oneorder of magnitude. In fact, such features in the spe-cific heat are related to various ongoing topics of interestin condensed matter physics. In particular, 4f -electronsbased magnetic systems19 show spectacular properties inthis regard.

Hereafter we focus on examples of various materialsclasses, in which a two-peak like anomaly in the spe-cific heat versus temperature have been observed exper-imentally. We start with the heavy-fermion compoundCe3Pd20Si6, a magnetic field-induced quantum criticalsystem21. For this material, J. Custers et al. observedthe presence of two distinct peaks centered at TQ '0.5 K and TN ' 0.25 K under an external magnetic fieldof 0.5 T in the electronic specific heat. These authorsattributed the features at TQ and TN to a antiferro-quadrupolar magnetic order of the Ce 4f orbitals locatedon the 8c site and to an antiferromagnetic order, respec-tively. Yet, the authors point out that this scenario iscompatible with the Γ8 quartet and Γ7 doublet groundstates due to the difference of the crystal-field splittingof the Ce atoms on the 4a and 8c lattice sites22. Thesmall energy difference of only ∆T ' 0.25 K ' 20µeVbetween the transition temperatures TQ and TN sug-gests that for Ce3Pd20Si6 a subtle change in the sys-tem total energy is enough to change the character ofthe Ce 4f orbitals. Interestingly enough, the suppressionof TQ gives rise to a magnetic field induced quantumphase transition. Also for CeAuGe, CeCuGe, CeCuSi20,TbPO4

23, A3Cu3(PO4)4 (A = Ca, Sr) and Cu(3–

Page 5: Speci c Heat Anomalies in Solids Described by a Multilevel ... · Speci c Heat Anomalies in Solids Described by a Multilevel ... Unesp - Univ Estadual Paulista, Departamento de F

5

0 5 0 1 0 0 1 5 0 2 0 0 2 5 0 3 0 00

2

4

6

E x p e r i m e n t a l d a t a f r o m R e f . [ 1 9 ] F i t s u s i n g E q . ( 1 1 )

C 4f (J

/mol

K)

T ( K )

FIG. 4. Electronic specific heat experimental data C4f (cir-cles) and theoretical fits (solid lines) for CeAuGe20 usingEq. 11 with kB = 8.314 J/(mol.K), g0 = 1, g1 = 2, g2 = 4,∆1 = 205 J/mol (282 J/mol) for the solid pink line (solid blueline) and ∆2 = 2327 J/mol.

Chloropyridine)2(N3)224, Cu(en)2Ni(CN)4

25, URu2Si226

where a magnetic transition followed by the appear-ance of superconductivity is observed, (Ce1−xLax)3Al27,PrOs4Sb12

28 and UPt329 with two superconducting tran-

sitions, double peaks in the specific have been reported.We now present the results of fits that can be comparedto experimental data set for the specific heat of a partic-ular heavy fermion compound. In Fig.4 we show exper-imental literature results of the electronic specific heatfor CeAuGe20 together with theoretical fits employingEq. 11. It is worth mentioning that the ordinary phononiccontribution to the specific heat were subtracted by theauthors of Ref. 20, being thus shown in Fig.4 the bare elec-tronic specific heat contribution originated from the 4f -electrons (C4f ). Interestingly enough, despite the non-interacting gas picture we have adopted in this work,the present approach is capable to describe an elec-tronic system, in which correlation effects are present asdiscussed above. In order to determine the nature ofthe electronic/magnetic “excitations” responsible for theemergence of a double-peak anomaly in the specific heatdata of CeAuGe, as well for the various materials above-mentioned, microscopic/spectroscopic data are requiredand constitute a topic out of the scope of the presentwork. It is worth mentioning that inelastic neutron scat-tering investigations20, carried out on CeAuGe at 15 K,revealed the presence of pronounced crystal field excita-tions at 24.3 meV (' 282 K), which roughly correspondsto one of the energy scales, namely ∆1 = 205 J/mol,employed in one of the fits (solid pink line) depictedin Fig. 4. Note that a fit using ∆1 = 282 J/mol (solidblue line in Fig. 4) does not deviate so much from theexperimental data. Hence, we figure out that the en-

ergy scales ∆i, employed in our model, give us indirecthints of the physical phenomena associated with the spe-cific heat maxima. In general terms, the description ofthe nature of a phase transition, which usually manifestsitself as sharp maximum in the specific heat data (seeFig. 4) in the low-temperature window, as well as thephysical mechanism responsible for the broad maximumin the high-temperature range require the combination ofcomplementary experimental techniques. Yet, it is worthmentioning that on the lower temperature flank, namelyaround the phase transition critical temperature Tc, thediscrepancy between fits and the experimental data canbe attributed to the absence of critical fluctuations, seee.g. Refs. 30–32 and references cited therein, not negligibleupon approaching Tc, in our Ansatz. In the next sectionwe discuss a degenerate Schottky model.

IV. DEGENERATE SCHOTTKY MODEL

Following the proposal of modelling and understandingof the double peak anomalies in specific heat data, in thisSection we focus on the description of a degenerate Schot-tky model. To this end, we consider a two level system(ε1 = − 1

2∆, ε2 = 12∆, with respective degeneracies Ω1,2)

plus a spin 12 Zeeman splitting contribution ±µBH, cf.

scheme shown in the inset of Fig. 5. The partition func-tion and equation for the energy (obtained using Eq. 8)per particle is, disregarding an additive constant, givenrespectively by33:

Z = Ω1e−βε1(e+βm0H + e−βm0H) +

Ω2e−βε2(e+βm0H + e−βm0H), (12)

E = −mH − ∆e−β∆

(1 + Ωe−β∆), (13)

where Ω = Ω1/Ω2, H is an external magnetic field andβ refers to the inverse temperature. The magnetizationm and the specific heat per particle are in turn givenrespectively by:

m = µBtanh(βµBH), (14)

CV =∂E

∂T= − ∂

∂T(mH) +

Ω(β∆)2e−β∆

(1 + Ωe−β∆)2. (15)

For an ordinary paramagnet, cf. discussion in Section I(see Eq. 5), the magnetic contribution to the specific heatis Schottky type, namely it is described by the secondterm of Eq. 15 with Ω = 1 and an energy gap ∆ = 2µBH,so that we have two decoupled Schottky like systems.We assume now that a ferromagnetic phase transition,

Page 6: Speci c Heat Anomalies in Solids Described by a Multilevel ... · Speci c Heat Anomalies in Solids Described by a Multilevel ... Unesp - Univ Estadual Paulista, Departamento de F

6

0 4 8 1 2 1 6 2 00 . 00 . 51 . 01 . 52 . 02 . 53 . 0

2 m ( T )

2 m ( T )∆

ε1 , Ω 1

∆ = 10, Ω = 5 ∆ = 20, Ω = 5 ∆ = 10, Ω = 10

Sp

ecific

heat

(a.u.)

T ( K )

ε2 , Ω 2

FIG. 5. Main panel: specific heat as a function of temper-ature for various energy gaps ∆ and degeneracy ratio Ω =Ω1/Ω2, cf. indicated in the label. Inset: energy scheme of thedegenerate Schottky model. See details in the main text.

which is a typical many body effect, takes place at acertain critical temperature Tc within the Curie-Weissmolecular mean-field model33. We introduce then themacroscopic molecular field phenomenological parameterλ and an effective magnetic field (Heff ), as follows:

Heff = Hext + λm. (16)

At zero external magnetic field m is calculated in aself-consistent fashion via the following equation:

m = µB tanh(βµBλm), (17)

being thus now the magnetic gap ∆ = 2µBH = 2µBλma function of the order parameter m. Scaling the mag-netization to µB , the energies (and temperatures) to Tc,with Tc ≡ λµB , we have:

CV = − ∂

∂T

(m2)

+Ω(β∆)2e−β∆

(1 + Ωe−β∆)2 , (18)

where m is the solution of m = tanh(βm). It is wellknown that m ≡ 0 for T > 1 and m(T ) 6= 0 otherwise.Also for small m (for T near and less than 1) we havem2 ≈ 3(1 − T ) so CV exhibits a jump δCV = 3 atT = 1. Fig. 5 (main panel) depicts the specific heat as afunction of temperature for various energy gaps ∆ (in Tc

units) and degeneracy ratio Ω, namely ∆ = 10 and Ω =5 (black solid line); ∆ = 20 and Ω = 5 (blue solid line);∆ = 10 and Ω = 10 (red solid line) employing Eq. 17.The low-T peak, with a mean-field like shape, remainsunaffected upon tuning ∆(>> 1) and Ω. The Schottkymaximum, however, becomes broader and shifts to highertemperatures as ∆ is increased, being the peak inten-sity also increased as Ω increases cf. expected for a two-level model, a feature that can be understood in termsof entropy arguments as discussed in a previous section.It is worth mentioning that between the ferromagnetictransition and Schottky maximum is exponentially small,i.e. the ferromagnetic phase transition produces a jumpin the specific heat data on top of a non-magnetic contri-bution. For ∆ values close to unity (not shown in Fig. 5)the high-T maximum tail evolves into the low-T , beingthe sharp mean-field like jump ∆C = 3 at T = 1 pre-served. Thus, our analysis show that a simple Schottkymodel coupled to a ferromagnetic phase transition is alsoamenable to fit emerging double peak/maxima featuresin measured specific heat data.

V. CONCLUSIONS

To summarize, we have employed a multilevel model todescribe specific heat data as a function of temperaturewith two maxima. Our results suggest that: i) a ratioof ∆2/∆1 ≈ 10 between the energy levels and ii) the en-ergy levels degeneracy, govern the two-peaks regime inthe specific heat. Apart from the nature of the entitiesresponsible for the emergence of the maxima in the spe-cific heat data, this simple model describes nicely recentliterature results for the CeAuGe compound. Further-more, the specific heat of a degenerate Schottky modelwas calculated. Our analysis demonstrate that doublepeak/maxima features in specific heat data can also stemfrom a two-level model coupled to a ferromagnetic phasetransition. We trust that our findings pave the way to-wards the description of experimental specific heat datafor other systems in both soft and condensed matterPhysics.

ACKNOWLEDGEMENTS

MdS and RP acknowledge financial support fromthe Sao Paulo Research Foundation – Fapesp (GrantsNo. 2011/22050-4 and 2011/17253-3, respectively) andNational Council of Technological and Scientific De-velopment – CNPq (Grants No. 308977/2011-4 and305472/2014-3).

[email protected]; Present Address: Institute of Semi-conductor and Solid State Physics, Johannes Kepler Uni-

versity Linz, Austria.

Page 7: Speci c Heat Anomalies in Solids Described by a Multilevel ... · Speci c Heat Anomalies in Solids Described by a Multilevel ... Unesp - Univ Estadual Paulista, Departamento de F

7

1 M. de Souza, A. Bruehl, J. Mueller, P. Foury-Leylekian,A. Moradpour, J.-P. Pouget, and M. Lang, Physica B:Condensed Matter 404, 494 (2009).

2 R. S. Manna, M. de Souza, A. Bruehl, J. A. Schlueter, andM. Lang, Phys. Rev. Lett. 104, 016403 (2010).

3 M. Pregelj, A. Zorko, O. Zaharko, Z. Kutnjak, M. Jagodic,Z. Jaglicic, H. Berger, M. de Souza, C. Balz, M. Lang, andD. Arcon, Phys. Rev. B 82, 144438 (2010).

4 A. Jesche, C. Krellner, M. de Souza, M. Lang, andC. Geibel, Phys. Rev. B 81, 134525 (2010).

5 A. Jesche, T. Forster, J. Spehling, M. Nicklas, M. de Souza,R. Gumeniuk, H. Luetkens, T. Goltz, C. Krellner, M. Lang,J. Sichelschmidt, H.-H. Klauss, and C. Geibel, Phys. Rev.B 86, 020501 (2012).

6 R. Lagos, A. Stein-Barana, and G. Cabrera, Physica C:Superconductivity 309, 170 (1998).

7 L. A. Macedo and R. E. Lagos, Journal of Magnetism andMagnetic Materials 226, 105 (2001).

8 A. Einstein, Annalen der Physik 22, 180 (1907).9 P. Foury-Leylekian, S. Petit, I. Mirebeaub, G. Andre,

M. de Souza, M. Lang, E. Ressouchee, A. Moradpour, andJ.-P. Pouget, Physical Review B 88, 024105 (2013).

10 P. Debye, Ann. Phys. 39, 789 (1912).11 E. S. R. Gopal, Specific Heats at Low Temperatures (Hey-

wood, New York, 1966).12 R. Pathria, Statistical Mechanics, International series of

monographs in natural philosophy (Pergamon Press, 1977).13 R. P. Hermann, F. Grandjean, and G. J. Long, American

Journal of Physics 73, 110 (2005).14 M. de Souza and J.-P. Pouget, Journal of Physics: Con-

densed Matter 25, 343201 (2013).15 L. Landau and E. Lifshitz, Statistical Physics, Bd. 5 (El-

sevier Science, 2013).16 R. B. Laughlin and D. Pines, Proceedings of the National

Academy of Sciences 97, 28 (2000).17 P. H. Meijer, J. H. Colwell, and B. Shah, American Journal

of Physics 41, 332 (1973).18 A. Abragam and B. Bleaney, Electron Paramagnetic Res-

onance of Transition Ions, International series of mono-graphs on physics (Clarendon P., 1970).

19 C. Pfleiderer, Rev. Mod. Phys. 81, 1551 (2009).20 B. Sondezi-Mhlungu, D. Adroja, A. Strydom, S. Paschen,

and E. Goremychkin, Physica B: Condensed Matter 404,3032 (2009).

21 J. Custers, K.-A. Lorenzer, M. Mueller, A. Prokofiev,A. Sidorenko, H. Winkler, A. Strydom, Y. Shimura,T. Sakakibara, R. Yu, Q. Si, and S. Paschen, Nature Ma-terials 11, 189 (2012).

22 T. Goto, T. Watanabe, S. Tsuduku, H. Kobayashi,Y. Nemoto, T. Yanagisawa, M. Akatsu, G. Ano, O. Suzuki,N. Takeda, A. Dnni, and H. Kitazawa, Journal of thePhysical Society of Japan 78, 024716 (2009).

23 R. Hill, J. Cosier, and S. Smith, Solid State Communica-tions 26, 17 (1978).

24 T. Nakanishi and S. Yamamoto, Phys. Rev. B 65, 214418(2002).

25 J. Cernak, M. Orendac, I. Potocnak, J. Chomic,A. Orendacova, J. Skorsepa, and A. Feher, CoordinationChemistry Reviews 224, 51 (2002).

26 T. Palstra, A. Menovsky, J. v. d. Berg, A. Dirkmaat,P. Kes, G. Nieuwenhuys, and J. Mydosh, Phys. Rev. Lett.55, 2727 (1985).

27 Y. Y. Chen, Y. D. Yao, B. C. Hu, C. H. Jang, J. M.Lawrence, H. Huang, and W. H. Li, Phys. Rev. B 55,5937 (1997).

28 R. Vollmer, A. Faißt, C. Pfleiderer, H. v. Lohneysen, E. D.Bauer, P.-C. Ho, V. Zapf, and M. B. Maple, Phys. Rev.Lett. 90, 057001 (2003).

29 R. A. Fisher, S. Kim, B. F. Woodfield, N. E. Phillips,L. Taillefer, K. Hasselbach, J. Flouquet, A. L. Giorgi, andJ. L. Smith, Phys. Rev. Lett. 62, 1411 (1989).

30 L. Bartosch, M. de Souza, and M. Lang, Phys. Rev. Lett.104, 245701 (2010).

31 M. de Souza, P. Foury-Leylekian, A. Moradpour, J.-P.Pouget, and M. Lang, Phys. Rev. Lett. 101, 216403(2008).

32 M. de Souza and L. Bartosch, Journal of Physics: Con-densed Matter 27, 053203 (2015).

33 F. Reif, Fundamentals of Statistical and Thermal Physics:(Waveland Press, 2009).