Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Isotopic effect in thermodynamic properties of solid neon

Ho Khac Hieu,1, 2, a) Nguyen Phuoc The,1, 2 Huynh Ngoc Toan,1, 2 Dang Thanh Hai,3
and Vu Van Hung4
1)
Institute of Research and Development, Duy Tan University, 03 Quang Trung,
Hai Chau, Da Nang 550000, Viet Nam
2)
Faculty of Environmental and Natural Sciences, Duy Tan University,
03 Quang Trung, Hai Chau, Da Nang 550000, Viet Nam
3)
Vietnam Education Publishing House, 81 Tran Hung Dao, Hoan Kiem, Hanoi,
Viet Nam
4)
VNU University of Education, 144 Xuan Thuy, Cau Giay, Hanoi 123105,
Viet Nam

(Dated: 28 April 2023)

Isotopic effects in mean-square displacement, specific heat at constant pressure and


linear thermal expansion coefficient of solid neon have been considered using the sta-
tistical moment method in statistical mechanics. This approach allows us to study
these effects including the anharmonicity of lattice vibrations. We implement cal-
20 22
culations for Ne and Ne isotopes in temperature range from 0 K to the melting
temperature (24.6 K) using Lennard-Jones potential to describe the interaction be-
tween Ne quantum particles. Our numerical results are compared with those of
path-integral Monte Carlo simulations and previous experimental data showing the
good agreement. This research presents an effective statistical approach to study
isotope effects on thermodynamic quantities of materials.

Keywords: Isotopic effect, Mean-square displacement, Thermal expansion, Specific


heat, Neon

a)
Electronic mail: hieuhk@duytan.edu.vn

1
I. INTRODUCTION

Statistical moment method (SMM) is one of the effective methods in statistical mechan-
ics to study thermodynamic and mechanical properties of materials1–3 . This method offers
several advantages as it considers both the anharmonicity of thermal lattice fluctuations at
high temperatures and the zero-point contribution at low temperatures. It is well-known
that many significant effects including thermal expansion, atomic fluctuation, phonon cou-
plings, pressure dependence of compressibility, and isotopic dependence of structural and
thermodynamic properties,... can be attributed to anharmonicity in solids.

Researchers from various disciplines such as materials science, high-pressure physics, and
geophysics have taken a great interest in rare-gas solids in recent times4–7 . These solids are
stable at high temperature and pressure8 making them popular choices for use as hydrostatic
pressure media in diamond anvil cell (DAC) experiments and as internal pressure standards
in X-ray diffraction (XRD) measurements7 . Moreover, due to their inert nature and simple
cubic structure with a closed-shell electronic configuration similar to that of free atoms, rare-
gas solids become ideal candidates for physical model verification7,9 . At ambient pressure,
neon, the second element in the rare gas group, remains a face-centred cubic (FCC) structure
until its melting point of 24.4 K. Its electronic structure has a relatively wide energy band
gap between the filled 2p-valence states and the 3d -conduction state10 . This characteristics
of neon may causes some interesting physical properties including anharmonic and quantum
behaviors that are distinct from those of the heavier noble gas solids11 . Recently, a qubit
platform has been developed by Zhou et al. based on isolated single electrons which were
trapped on an ultraclean solid neon surface in vacuum12 . Moreover, despite their accuracy
in predicting the physical properties of heavier noble gas solids, many theories may not be
applicable to neon. And this solid serves as a more rigorous assessment of lattice dynamic
properties. Therefore, the investigation of thermodynamic properties, phonon dispersion
and transport properties of this solid is becoming increasingly significant, with the isotopic
composition being one of the factors that affects these properties13–16 .

Numerous papers have extensively explored the influence of isotopic mass difference on the
physical characteristics of solid neon. This mass difference can alter the lattice parameter,
thermal conductivity, vibrational phonon frequencies, energy gap, specific heat, and other
electronic properties. Experimentally, the structural and thermodynamic properties of neon
isotopes have been measured such as specific heats at constant volume and pressure11,17 , lin-

2
ear thermal expansion coefficients (TEC)18 . On the theoretical side, many approaches (such
as self-consistent harmonic approximation of Koehler19,20 , path-integral Monte Carlo simu-
lations (PIMC)21 , and ab initio approach within the Tolpygo model including the electron
shell deformations14 ) have been employed to consider the temperature-dependent thermo-
mechanical properties of neon isotopes.
The aim of this study is to enhance our comprehension of the impact of isotopic effects
on the thermodynamic properties of neon. Using the SMM in statistical mechanics as a
basis1,2 , we investigate the mean-square displacement, linear TEC, and specific heat at
constant pressure of 20 Ne and 22 Ne. The comparison between our obtained results and those
of previous measurements will be made where possible to verify the theory.
This paper is structured as follows. The first section provides an overview of the research
content and current state of research. In Section II, we present a theoretical calculation
considering anharmonic effects to determine the thermodynamic quantities. In Sec. III, we
conduct the numerical calculations to obtain the mean-square displacement, linear TEC and
specific heat at constant pressure as a function of temperature with various isotopic masses.
Additionally, we provide a detailed discussion of the numerical results in this section. Finally,
the conclusions of this research are presented in Sec. IV.

II. THEORETICAL PERSPECTIVE

In the following, we briefly introduce the SMM theory and encapsulate its main formulae
and results. This method has been developed over the years by Tang & Hung and their
colleagues to study effectively the structural, thermodynamic, and mechanical properties of
crystals1,2,22 . Assuming that our system takes the following Hamiltonian as

H b 0 − αVb ,
b =H (1)

where H
b 0 and αVb are, respectively, the harmonic term and anharmonic contribution (caused

by thermal lattice vibrations) of lattice Hamiltonian. Then the Helmholtz free energy of our
system can be obtained by the following general formula1
Z α
ψ = ψ0 − hVb iα dα , hVb iα = −∂ψ/∂α, (2)
0

where ψ0 represents the Helmholtz free energy associated with the Hamiltonian H
b 0 , and

h...iα refers to the expected free energy corresponding to the anharmonic Hamiltonian.

3
As a result of thermal lattice vibrations, an additional force p affects the zeroth central
atom, requiring the total force acting on it to be zero. Therefore, we obtain the following
force balance equation
 2
∂ 3 ϕi0
  
1X ∂ ϕi0 1X
huiα ip + huiα uiχ ip + (3)
2 α,δ ∂uiα ∂uiδ eq 4 α,δ,γ ∂uiα ∂uiδ ∂uiχ eq
∂ 4 ϕi0
 
1 X
+ huiα uiχ uiη ip − p = 0,
12 i,α,χ,η ∂uiα ∂uiδ ∂uiχ ∂uiη eq

where α 6= χ 6= η = x, y or z in the Cartesian coordinate system; huiα ip , huiα uiχ ip and


huiα uiχ uiη ip are, correspondingly, the first, second and third order moments of the atomic
displacement.
Using the recurrence formula in SMM23 , the moments huiα uiχ ip and huiα uiχ uiη ip can be
expressed in terms of huiα ip as

∂ huiα ip ~δαχ
 
~ω θδαχ
huiα uiχ ip = huiα ip huiχ ip + θ + coth − , (4)
∂pχ 2mω 2θ mω 2

∂ huiχ ip
huiα uiχ uiη ip = huiα ip huiχ ip huiη ip + θPαχη huiα ip
∂pη
∂ 2 huiα ip ~ huiη ip δαχ θ huiη ip δαχ
 

+ θ2 + coth − , (5)
∂pχ pη 2mω 2θ mω 2

where Pαχη = 1 only if α = χ = η, otherwise Pαχη = 0, and δαχ is the Kronecker delta
function.
Then we can rewrite the Eq. (1) into the second-order differential equation of the thermal
atomic displacement y ≡ hui ip as2

d2 y dy θ
γθ2 2
+ 3γθy + ky + γ (X − 1) − p = 0, (6)
dp dp k

where
∂ 2 ϕi0
 
1X
k= = mω 2 ; (7)
2 i ∂u2iα eq

γ = 4 (γ1 + γ2 ) , (8)

∂ 4 ϕi0 ∂ 4 ϕi0
   
1 X 1X
γ1 = , γ2 = , (9)
48 i ∂u4iα eq 8 i ∂u2iα ∂u2iδ eq

4

where X = (~ω/2θ) coth ~ω/2θ , θ = kB T (~ and kB denote, respectively, the reduced
Planck constant and the Boltzmann constant), the notation eq represents the thermal aver-
age computed over the equilibrium ensemble2 , and m is the atomic mass.
p
Here it is significant to mention that ω = k/m is the frequency of atomic vibration
which should be treated as the Einstein frequency in the Einstein model. Therefore, the
Debye frequency ωD and Debye temperature θD can be obtained by using the following
approximations √ r
√ √ 2~ k
ωD ≈ 2ω, θD ≈ 2θE = , (10)
kB m
r
~ω ~ k
where θE = = is the Einstein temperature of the crystal.
kB kB m
Because the supplemental force p is small and arbitrary, the solution of the differential
Eq. (6) can be found approximately as1,2

y (T ) ≈ y0 (T ) + A1 p + A2 p2 , (11)

where y0 is the atomic displacement in the case of no supplemental force p. It has the value
as1 r
2γθ2
y0 (T ) ≈ A, (12)
3k 3
with
γ 2 θ2 γ 3 θ3 γ 4 θ4 γ 5 θ5 γ 6 θ6
A = a1 + a 2 + a 3 + a 4 + a 5 + a6 (13)
k4 k6 k8 k 10 k 12
where the detailed expressions of ai (i = 1, 6) can be found in Refs. 1
Furthermore, applying Eq. (2) for a crystal with FCC structure having N atoms, the
Helmholtz free energy of the system has been derived as1,2
3N θ2
  
2 2γ1 X
ψ = U0 + ψ0 + γ2 X − 1+ +
k2 3 2
  
2θ 4 2 2
 X
+ 2 γ X − 2 γ1 + 2γ1 γ2 (1 + X) 1 + , (14)
k 3 2 2
where U0 and ψ0 = 3N θ [x + ln (1 − e−2x )] are, respectively, the total interaction potential
energy and the harmonic free energy of the system.
Moreover, using Eq. (4) we obtain the temperature-dependent analytical expression of
atomic MSD function as the following2
θ
u2iα (T ) = huiα i2 (T ) + θA1 + (X − 1) (15)
k
θ
≈ y02 (T ) + θA1 + (X − 1) ,
k

5
where
2γ 2 θ2
   
1 X
A1 = 1+ 1+ (X + 1) . (16)
k k4 2
For isotropic crystal (such as FCC structure), the three-dimensional mean-square atomic
deviation (represented here as hu2 i) has a simple relation with the one-dimensional version
hu2iα i as

u2 (T ) = 3 u2iα (T ) (17)

= 3y02 (T ) + 3θA1 + (X − 1) .
k

The limitations of the atomic MSD at low and high temperatures are
*) At low temperature limit θ → 0:

3~ω
u2 T →0 K
= . (18)
2k

*) At high temperature limit X = x coth x → 1, and therefore

2 3θ 18γ 2 θ3 2γθ2
u T →∞
= + + 3 A. (19)
k k5 k

And from the Helmholtz free energy ψ of the system, we can derive the expressions of
TEC as √
2kB χT 1 ∂ 2 ψ
 
kB χT ∂P
αT = =− , (20)
3 ∂θ V 3a2 3N ∂θ∂a
where P = − (∂ψ/∂V )T , and χT = −1/V0 (∂V /∂P )T is the isothermal compressibility.
Furthermore, we can also derive the specific heats at constant volume CV and pressure
CP for FCC system as24
 
2θ h γ1  γ1 2 2
i
CV = 3N kB Y + 2 2γ2 + XY + (1 + Y ) − γ2 Y + 2X Y , (21)
k 3 3

9T V α2
CP = CV + , (22)
χT
where Y = x2 / sinh2 x.

III. RESULTS AND DISCUSSION

In this section, from the derived analytical expressions, we implement numerical calcu-
20 22
lations for the two isotopes Ne and Ne to investigate the isotope effects on the atomic
MSD function, linear TEC and specific heat at constant pressure in the temperature range

6
TABLE I. Debye frequency and Debye temperature of 20 Ne and 22 Ne isotopes.

Isotopes ωD (×1012 Hz)a θD (K)a θD (K)b u2 0


(×10−2 Å2 )a u2 0
(×10−2 Å2 )c
20 Ne 9.0388 69.0391 70.0 7.46
22 Ne 8.6181 65.8262 66.7 7.12
Ne 8.9984 68.7310 – 7.43 7.05–8.58

a This work
b Debye model13
c First-principles calculations14 .

0–24.6 K. For quantum crystals, the interaction between two neighbor atoms can be well
described by simple Lennard-Jones potential function. For Ne crystal, its atoms are treated
as quantum particles interacting through the (6-12) Mie-Lennard-Jones potential ϕ (r) =
4 (σ/r)12 − (σ/r)6 , where the equilibrium value σ = 2.7012 Å, and the dispersion energy
 

 = 72.09 × 10−16 ergs20 . With the aim of implementing numerical calculations, we assume
the interaction potential which does not depend on the isotopic composition. It means that
20 22
the above potential can be applied for both isotopes Ne and Ne of neon.

17

20
16 Ne (PIMC)
22
Ne (PIMC)
15 20
Ne
22
14 Ne
MSD <u 2 > (Å 2 )

13

12

11

10

7
0 5 10 15 20 25
Temperature (K)

FIG. 1. Temperature-dependent MSDs of 20 Ne and 22 Ne isotopes (solid and dashed lines, respec-
tively). Open square and circle symbols are the PIMC results13 .

We will first examine the impact of the isotopic mass on the Debye frequency and Debye
temperature of solid neon. In Table I, we report our calculations and collected data from

7
publications of Debye frequency and temperature of 20 Ne and 22 Ne isotopes and natural neon
(by considering the average mass of natural neon 20.18 amu). The Debye temperatures of
20 22
Ne and Ne isotopes in our SMM approach are, respectively, 69.0391 K and 65.8262 K.
These values are great consistent with those derived from Debye model13 . The differences
20 22
of Debye temperatures obtained in our calculations and Debye model for Ne and Ne are
1.4% and 1.3%, respectively.
20 22
In Fig. 1, we show the temperature-dependent MSDs of the two isotopes Ne and Ne
obtained by SMM approach. The PIMC results (open squares and circles)13 are also shown
for comparison. The good agreement between our calculations and PIMC simulations is
found. As it can be observed from this figure, the zero-point vibration contribution is
included in the MSD function at low temperature. Particularly, the zero-point vibration
20 22
contributions to MSDs of Ne, Ne isotopes and natural neon are hu2 i0 (20 Ne) = 7.46 ×
10−2 Å2 , hu2 i0 (22 Ne) = 7.12 × 10−2 Å2 , and hu2 i0 (Ne) = 7.43 × 10−2 Å2 , respectively. As
the temperature increases, the MSD changes linearly with respect to temperature. The
MSDs of isotopes exhibit discrepancies due to the difference in their masses, with isotopes
of greater mass having less spatial delocalization. With lighter mass, the MSD curve of 20 Ne
22 20
is higher than one of Ne. As expected, the difference between MSDs of two isotopes Ne
and 22 Ne can be obviously observed at low temperature. Furthermore, Fig. 1 shows that the
disparity between MSDs diminishes gradually at high temperatures (above 17 K) and the
noticeable difference can be observed at below 12 K. This is caused by the disparity in the
quantum zero-point vibration contributions of the two isotopes where the atomic vibrational
amplitude of the lighter isotope is greater than that of the heavier one. With the increase
of temperature, the predominance of anharmonicity leads to the increasing of oscillation
amplitude, causing the reduction of the isotopic effect on MSD.
In order to evaluate the isotopic impact on the divergence of MSD function, we compute
20 22
the temperature-dependent ratio between the MSDs of Ne and Ne isotopes. The results
of delocalization ratio δ = MSD(22 Ne)/MSD(20 Ne) of our SMM approach and PIMC simula-
tions (open circles)13 are presented in Fig. 2. We can see in Fig. 2 that the SMM calculations
of delocalization ratio δ very close to those of PIMC simulations for temperatures exceeding
8 K. In the temperature range 10–24.6 K, the ratio increases drastically toward the value
of 1. At low temperature region (below 8 K), there is a clear difference between SMM and
PIMC approach. The error of the δ values derived from the SMM approach grows rapidly as
temperature approaches zero. In our work, this ratio does almost not change and converges

8
0.985

0.98 Present theory


PIMC
0.975

=MSD( 22 Ne)/MSD( 20 Ne)


0.97

0.965

0.96

0.955

0.95

0.945

0.94
0 5 10 15 20 25
Temperature (K)

FIG. 2. Temperature dependent-ratio between the MSDs of 22 Ne and 20 Ne.

p p p
to the square-root of the inverse mass ratio m (20 Ne)/ m (22 Ne) = 20/22 ≈ 0.9535 of
these two isotopes, like the harmonic expectation at T = 0 K. The low-temperature limit
for the delocalization ratio in PIMC simulations is lower than the expected value of SMM
approach, about 0.946.

10-3
1.2

Exp
20
1 Ne
22
Ne

0.8
(K -1 )

0.6
T

0.4

0.2

0
0 5 10 15 20 25
Temperature (K)

FIG. 3. Temperature dependent linear TEC of 22 Ne and 20 Ne. Experimental thermal expansions
for free-standing samples of solid neon from 2 to 14 K are also reported for comparison18 .

To get a further picture of the isotopic effects on thermodynamic properties of solid


neon, we determine the linear TECs and specific heats at constant pressure of two 20 Ne and
22
Ne isotopes as functions of temperature. In Fig. 3&4, we show these two thermodynamic

9
20 22
quantities of Ne and Ne isotopes along with the experimental measurements. It can
be observed from Fig. 3&4, our theoretical predictions derived from SMM approach follow
very well the experimental data11,18 . The calculated data points obtained for 22 Ne are softly
20
higher than those corresponding to Ne. The difference between TECs (and specific heats
20 22
CP ) of Ne and Ne is remarkable at temperature beyond 8 K. The disparity disappears
progressively at low temperatures. This phenomenon can be explained by the raising of
anharmonicity caused by thermal lattice vibrations at high temperature.

30

20
Ne (Exp)
22
25 Ne (Exp)
20
Ne
22
Ne
20
Cp

15

10

0
0 5 10 15 20 25
Temperature (K)

FIG. 4. Temperature dependent specific heats CP of 22 Ne and 20 Ne. The measurements of the two
isotopes in a calorimeter employing a mechanical heat switch are also displayed in the temperature
range 2.5–23.5 K11 .

IV. CONCLUSIONS

In present work, we have employed the SMM approach to investigate the influence of
isotopic mass difference on the mean-square displacements, linear TECs, and specific heats
at constant pressure of solid neon isotopes. Numerical computations of these thermodynamic
20 22
properties were conducted for both Ne and Ne isotopes from zero Kelvin to its melting
temperature. The discrepancy in MSDs, linear TECs, and specific heats at constant pressure
20 22
between Ne and Ne isotopes have been attributed to the isotopic effects taking into
account the anharmonicity of thermal lattice vibrations. The favorable consistency between
our findings and the available experimental data attests to the potential of SMM as a novel

10
approach for determining the thermodynamic quantities of materials.

CONFLICT OF INTEREST

We have no conflicts of interest to disclose.

DATA AVAILABILITY

The data that support the findings of this study are available within the article.

REFERENCES

1
N. Tang and V. V. Hung, Phys. Stat. Sol. (b) 149, 511 (1988), ISSN 1521-3951, URL
http://dx.doi.org/10.1002/pssb.2221490212.
2
K. Masuda-Jindo, V. V. Hung, and P. D. Tam, Phys. Rev. B 67, 094301 (2003), URL
http://link.aps.org/doi/10.1103/PhysRevB.67.094301.
3
K. Masuda-Jindo, S. R. Nishitani, and V. Van Hung, Phys. Rev. B 70, 184122 (2004),
URL http://link.aps.org/doi/10.1103/PhysRevB.70.184122.
4
W. A. Caldwell, J. H. Nguyen, B. G. Pfrommer, F. Mauri, S. G. Louie, and R. Jean-
loz, Science 277, 930 (1997), ISSN 0036-8075, URL https://science.sciencemag.org/
content/277/5328/930.
5
D. Errandonea, B. Schwager, R. Boehler, and M. Ross, Phys. Rev. B 65, 214110 (2002),
URL https://link.aps.org/doi/10.1103/PhysRevB.65.214110.
6
W. Wei, X. Li, N. Sun, S. N. Tkachev, and Z. Mao, American Mineralogist 104, 1650
(2019), ISSN 0003-004X, URL https://doi.org/10.2138/am-2019-7033.
7
A. Dewaele, A. D. Rosa, N. Guignot, D. Andrault, J. E. F. S. Rodrigues, and G. Garbarino,
Scientific Reports 11, 15192 (2021), ISSN 2045-2322, URL https://doi.org/10.1038/
s41598-021-93995-y.
8
S. Ono, Scientific Reports 10, 1393 (2020), ISSN 2045-2322, URL https://doi.org/10.
1038/s41598-020-58252-8.
9
A. N. Singh, J. C. Dyre, and U. R. Pedersen, The Journal of Chemical Physics 154, 134501
(2021), URL https://doi.org/10.1063/5.0045398.

11
10
Y. guang He, X. zhang Tang, and Y. kang Pu, Physica B: Condensed Matter 405, 4335
(2010), ISSN 0921-4526, URL https://www.sciencedirect.com/science/article/
pii/S0921452610007362.
11
E. Somoza and H. Fenichel, Phys. Rev. B 3, 3434 (1971), URL https://link.aps.org/
doi/10.1103/PhysRevB.3.3434.
12
X. Zhou, G. Koolstra, X. Zhang, G. Yang, X. Han, B. Dizdar, X. Li, R. Divan, W. Guo,
K. W. Murch, et al., Nature 605, 46 (2022), ISSN 1476-4687, URL https://doi.org/
10.1038/s41586-022-04539-x.
13
C. P. Herrero, Phys. Rev. B 65, 014112 (2001), URL https://link.aps.org/doi/10.
1103/PhysRevB.65.014112.
14
E. P. Troitskaya, V. V. Chabanenko, and E. E. Horbenko, Physics of the Solid State 51,
2121 (2009), ISSN 1090-6460, URL https://doi.org/10.1134/S1063783409100229.
15
J. Eckert, W. B. Daniels, and J. D. Axe, Phys. Rev. B 14, 3649 (1976), URL https:
//link.aps.org/doi/10.1103/PhysRevB.14.3649.
16
Y. J. Gu, W. L. Quan, G. Yang, M. J. Tan, L. Liu, and Q. F. Chen, Phys. Rev. E 102,
043214 (2020), URL https://link.aps.org/doi/10.1103/PhysRevE.102.043214.
17
H. Fenichel and B. Serin, Phys. Rev. 142, 490 (1966), URL https://link.aps.org/doi/
10.1103/PhysRev.142.490.
18
J. C. Holste and C. A. Swenson, Journal of Low Temperature Physics 18, 477 (1975),
ISSN 1573-7357, URL https://doi.org/10.1007/BF00116138.
19
H. D. Jones, Phys. Rev. Lett. 23, 766 (1969), URL https://link.aps.org/doi/10.
1103/PhysRevLett.23.766.
20
D. Acocella, G. K. Horton, and E. R. Cowley, Phys. Rev. B 61, 8753 (2000), URL https:
//link.aps.org/doi/10.1103/PhysRevB.61.8753.
21
C. P. Herrero, Journal of Physics: Condensed Matter 15, 475 (2003), URL https://dx.
doi.org/10.1088/0953-8984/15/3/312.
22
H. K. Hieu and V. V. Hung, Modern Physics Letters B 25, 1041 (2011), URL http:
//www.worldscientific.com/doi/abs/10.1142/S0217984911026760.
23
V. V. Hung, D. D. Phuong, N. T. Hoa, and H. K. Hieu, Thin Solid Films 583, 7
(2015), ISSN 0040-6090, URL http://www.sciencedirect.com/science/article/pii/
S0040609015002552.
24
N. Tang and V. V. Hung, physica status solidi (b) 161, 165 (1990),
https://onlinelibrary.wiley.com/doi/pdf/10.1002/pssb.2221610115, URL https:

12
//onlinelibrary.wiley.com/doi/abs/10.1002/pssb.2221610115.

13

You might also like