Anda di halaman 1dari 4

APPLIED PHYSICS LETTERS 91, 122104 共2007兲

Nonlinear Peltier effect in semiconductors


Mona Zebarjadia兲
Department of Electrical Engineering, University of California, Santa Cruz, California 95064, USA
Keivan Esfarjani
Department of Physics, University of California, Santa Cruz, California 95064, USA
Ali Shakouri
Department of Electrical Engineering, University of California, Santa Cruz, California 95064, USA
共Received 12 July 2007; accepted 27 August 2007; published online 18 September 2007兲
Nonlinear Peltier coefficient of a doped InGaAs semiconductor is calculated numerically using the
Monte Carlo technique. The Peltier coefficient is also obtained analytically for single parabolic band
semiconductors assuming a shifted Fermi-Dirac electronic distribution under an applied bias.
Analytical results are in agreement with numerical simulations. Key material parameters affecting
the nonlinear behavior are doping concentration, effective mass, and electron-phonon coupling.
Current density thresholds at which nonlinear behavior is observable are extracted from numerical
data. It is shown that the nonlinear Peltier effect can be used to enhance cooling of thin film
microrefrigerator devices especially at low temperatures. © 2007 American Institute of Physics.
关DOI: 10.1063/1.2785154兴

The Peltier coefficient plays an important role on how bulk semiconductors. Doped semiconductors are the best
good a material is for thermoelectric solid-state refrigeration candidates for thermoelectric applications so it is important
or power generation. In the linear regime, the Peltier coeffi- to understand their behavior at experimentally achievable
cient is independent of the current and it is equal to the high current densities. At scales larger than the electron de
product of the Seebeck coefficient by the absolute tempera- Broglie wavelength, the Boltzmann transport equation 共BTE兲
ture. If we keep increasing the applied fields to high values, is the governing equation. The Monte Carlo 共MC兲 technique
linear relations will no longer be valid. Nonlinear current- is considered as one of the most accurate tools to solve BTE.
voltage characteristics are very common in most active elec- We have developed a Monte Carlo program9 to simulate
tronic devices. On the other hand, nonlinear thermoelectric thermoelectric transport in GaAs family of materials. The
effects have not been investigated in detail. code is three dimensional both in k and r spaces with non-
Kulik1 calculated the electric field dependence of the parabolic multivalley band structure. The scattering mecha-
third-order Peltier coefficient in metals at low temperatures nisms included are ionized and neutral impurities, intravalley
supposing constant inelastic and elastic relaxation times. He polar optical phonons, acoustic phonons, and inter-
showed that this term is proportional to the product of the intravalley nonpolar optical phonons. Pauli exclusion prin-
total relaxation time by the inelastic relaxation time and that ciple is enforced after each scattering process supposing a
shifted Fermi sphere as the local electronic distribution. For
it is inversely proportional to the electron effective mass.
each valley, the electronic temperature is defined locally as
Grigorenko et al.2 calculated the nonlinear Seebeck coeffi-
follows:

再 冉 冎
cient in metals by expanding the distribution function in se-
ries of temperature gradients. They found that higher-order
nonlinear thermoelectric terms are proportional to the square
of the scattering time at the Fermi level. A dimensionless
f v共k, ␮v,Tev兲 = exp
Ev关兩k − kdv共r兲兩兴 − ␮v共r兲
kBTev共r兲
冊 +1
−1
,

parameter ␻ = l0䉮T / T was defined 共l0 is the electron mean


free path and T is the temperature兲 to describe the deviation 2
Tev共r兲 = 兵具Ev关k − kdv共r兲兴典 − 具Ev共r兲典0其 + T. 共1兲
from local equilibrium and the nonlinearity of the system. 3kB
Later they extend their theory to the case of two-dimensional
Here 具Ev共r兲典0 is the local average energy of electrons in equi-
metals.3 Freericks and Zlatic generalized the many-body for-
malism of the Peltier effect to the nonlinear regime.4 Non- librium at zero electric field.kdv共r兲 is the local drift wave
linearity of the thermoelectric effects in lower dimensions, vector, which is the average wave vector of all the particles
such as nanowires5 and point contacts,6 has also been inves- at position r and in valley v, and ␮ is the quasiFermi level.
tigated using the Landauer formalism. Experimentally, non- Details of adding Pauli exclusion principle in highly doped
semiconductors is described in another publication10 where
linearity of the Seebeck coefficient has been observed in a
we showed that using the above definition for electronic tem-
one-dimensional ballistic constriction at low temperatures7
perature results in the correct electronic distribution. The for-
共550 mK兲 and recently in the measurement of the Seebeck
malism works up to high fields, in the regime where nonpa-
coefficient of single molecule junctions.8 rabolic multivalley band structure is valid.
On the theoretical side, there has not been any formalism A uniform lattice temperature is enforced along the
beyond the constant relaxation time approximation to de- sample. The sample is subjected to a voltage difference. The
scribe the nonlinearity of thermoelectric effects in doped resulting potential distribution and current flow are obtained
via the Monte Carlo code coupled with a one-dimensional
a兲
Electronic mail: mona@soe.ucsc.edu Poisson solver. Dirichlet boundary conditions are used for

0003-6951/2007/91共12兲/122104/3/$23.00 91, 122104-1 © 2007 American Institute of Physics


122104-2 Zebarjadi, Esfarjani, and Shakouri Appl. Phys. Lett. 91, 122104 共2007兲

冕 T
Te
cvdT = ␶E␴F2 ⬇
3n
2
kB共Te − T兲. 共5兲

Here cv is the heat capacity per unit volume, Te is the elec-


tronic temperature, ␶E is the energy relaxation time 关Eq. 共5兲
can be taken as the definition of ␶E兴, ␴ is the electrical con-
ductivity, and F is the electric field. After substituting Eqs.
共4兲 and 共5兲 into Eq. 共3兲, we find that the Peltier coefficient is
␧d + 5/2kBTe − ␮
⌸= , 共6兲
e

FIG. 1. 共Color online兲 Comparison of Peltier 共dots with errorbar兲 and See-
beck 共multiply by the absolute temperature, squares兲 coefficients obtained
from the MC simulation and the analytical results. Figure confirms that the
⌸=−
␮ 5kBT
e
+
2e
m
+ 3 2 1+
2e n
10␶E 2
3␶av
J ,冉 冊 共7兲

Onsager relation is satisfied. Results are reported for In0.53Ga0.47As at room


where ␶av is defined as ␶av = 具E␶共E兲典 / 具E典; ␶共E兲 is the charac-
temperature.
teristic time which describes how the distribution function
relaxes.11
Poisson solver and periodic boundary conditions are sup- In degenerate limit, we have
posed in all directions for MC simulation. The Peltier coef-
ficient is defined as JQ = 兩⌸Je兩ⵜT=0. In the linear transport re- ប2q2e 3PL共5/2,− e␤e␮兲 3 3␲2 共kBTe兲2
= ␤e␮ ⬇ ␮ + . 共8兲
gime, the Peltier coefficient can be calculated analytically 2m ␤ePL共3/2,− e 兲 5 10 ␮
共see for example, Ref. 9兲. A simple test of the program is to
check the agreement between MC data and analytical results. Again Te can be related to ␶E by
This is confirmed in Fig. 1. The same band structure and
relaxation times are used in both cases. In another MC pro-
gram, we enforce a linear temperature drop along the same
冕 T
Te
cvdT = ␶E␴F2 ⬇
␲2 2
k g共␮兲共T2e − T2兲,
6 B
共9兲

bulk sample and we calculate the electrochemical difference where g共␮兲 is the density of states per unit volume at the
of the hot and cold side under open voltage conditions. The Fermi level. Finally for degenerate case the Peltier coeffi-
Seebeck coefficient is defined as ⵜ␮ ¯ = 兩eS ⵜ T兩J=0, where ␮¯ cient is
= ␮ + eV. In Fig. 1 we have also reported the result of the
Seebeck coefficient obtained from the later program. This
confirms the satisfaction of the Onsager relation and there- ⌸⬃
␲2 共kBT兲2
2 ␮e
m
+ 3 2 1+4
2e n

␶E 2
␶av
J . 冊 共10兲
fore the consistency of simulations.
Supposing a shifted Fermi-Dirac distribution for elec- Decreasing total scattering rates result in stronger non-
trons, the Peltier coefficient is obtainable analytically. After a linear transport, by which we mean the current is not linearly
second order Taylor expansion of the distribution function proportional to the electric field 关Eq. 共11兲 below兴. However,
about kd 共drift wavevector兲, one finds: it does not affect the nonlinearity of the Peltier coefficient as
much, since the Peltier coefficient is the ratio of two nonlin-
ear currents and the effect of increasing scattering rates can-
5 ប2q2e ⳵2vd nq2 cels.
n v d␧ d + n v d + ␧d Tr 2 e

冉 冊
−␮ 3 2md ⳵k 6

冉 冊
⌸= + , 共2兲 ␲2 ne␶共kBT兲2 ne3␶av
3
␶E 3
e 1 ⳵ vd
2
JQ ⬃ F+ 1+4 F . 共11兲
en vd + q2e Tr 2 2 ␮m 2m 2
␶av
6 ⳵k
In both degenerate and nondegenerate limits, nonlinear
where q = k − kd, q2e = 兺qq2 f q / 兺q f q, vd = 兩共1 / ប兲共⳵␧ / ⳵k兲兩k=kd, and Peltier is proportional to the effective mass and it is inversely
1 / md = 兩共1 / ប2兲共⳵2␧ / ⳵k2兲兩k=kd. The Taylor expansion becomes proportional to the square of the carrier concentration. We
exact for the quadratic dispersion and the Peltier coefficient numerically checked the validity of these proportionalities
simplifies to for the intermediate doping concentrations and we found that
these relations are valid even in the intermediate regime.

⌸=
1
e

− ␮ + ␧d +
5 ប2q2e
3 2m
. 冊 共3兲
Figure 2 shows the results obtained from the Monte
Carlo simulation for a parabolic band structure. In the non-
degenerate limit, the curves are compared with analytical
In nondegenerate limit, the third term in the Peltier co- expression 关Eq. 共6兲兴. ␧d, Te, and ␮ were extracted from the
efficient becomes MC data. One might argue that the agreement we obtained in
this figure is due to the assumption of a shifted Fermi-Dirac
distribution in both cases. To show that this is a correct hy-
ប2q2e 3PL共5/2,− e␤e␮兲 3 pothesis, results obtained using the standard method of en-
= ␤e␮ ⬇ k BT e , 共4兲
2m ␤ePL共3/2,− e 兲 2 forcing Pauli exclusion principle without any assumption on
the electronic distribution 关known as Lugli-Ferry method12
where PL is the polylog function defined as PL共n , z兲 LF兴 are also plotted. The agreement between the LF method
⬁ k n
= 兺k=1 z / k . To relate the electronic temperature to the relax- and the other data suggests that the distribution function is a
ation time, we use the energy conservation, shifted Fermi-Dirac. In Fig. 2 we have also reported the re-
122104-3 Zebarjadi, Esfarjani, and Shakouri Appl. Phys. Lett. 91, 122104 共2007兲

FIG. 2. 共Color online兲 Q refers to quadratic band dispersion and Full refers FIG. 3. 共Color online兲 Linear and nonlinear theory predictions of the cool-
to multivalley nonparabolic band structure. Results obtained from the Monte ing efficiency of InGaAs at T = 300 K and T = 77 K. For each temperature
Carlo simulation by dividing the thermal current to the electrical current the results are reported for the corresponding optimum dopings of the linear
共circles兲 are shown in comparison with the analytical expression 关Eq. 共6兲兴 transport theory, which are 1018 and 5 ⫻ 1015 cm−3 for T = 300 K and T
using the Fermi level and electronic temperature obtained from the simula- = 77 K, respectively.
tion 共triangular兲. Results obtained using LF method are also shown to con-
firm the validity of the approach 共squares兲. Solid lines are obtained from a
more realistic band structure 共nonparabolic兲. These are plotted for four dif- ing at T = 300 K is 1018 cm−3 and at T = 77 K is 5
ferent carrier concentrations. The above data are reported for n-type ⫻ 1015 cm−3. Linear and nonlinear cooling curves 关Eq. 共12兲兴
In0.53Ga0.47As at room temperature. At high carrier concentrations Peltier are plotted in Fig. 3 for these optimum doping concentra-
coefficient tends to be linear.
tions. According to the figure, cooling efficiency is enhanced
by 20% at room temperature and by 700% at T = 77 K.
sults from a more realistic band structure 关multivalley non- In summary, nonlinear Peltier coefficient is calculated
parabolic uses E共1 + ␣E兲 = ប2k2 / 2m with ␣⌫ = 1.307, ␣L analytically and numerically. Results show that nonlinearity
= 0.691, and ␣x = 0.202 eV−1兴. occurs when electronic temperature starts to exceed the lat-
When electronic temperature is higher than the lattice tice temperature. Electronic heating is stronger when the
temperature, nonlinear behavior is observable. Nonlinearity electron heat capacity is low 共that is the case for low doping
is stronger for lower carrier concentrations. The reason is concentrations兲 and when e-ph coupling is weak. Nonlinear
that at high concentrations, where the system is almost de- Peltier coefficient is independent of the ambient temperature
generate, the electron heat capacity is large and therefore and it is proportional to the electronic mass and inversely
much larger fields are required to heat up electrons. The e-ph proportional to the square of carrier concentration. The cur-
coupling is another factor that determines the nonlinearity of rent threshold at which the Peltier coefficient becomes non-
the system. In materials with large e-ph coupling, electrons linear depends on the carrier concentration. For InGaAs non-
tend to thermalize faster with the lattice, therefore no heating linearity starts at 104 A / cm2 for n = 1016 cm−3 and it
takes place and transport stays linear. Figure 2 shows that for increases to 105 A / cm2 for n = 1017 cm−3. These currents are
low carrier concentrations, nonlinear Peltier is relevant at achievable experimentally in thin film devices. The nonlinear
currents on the order of 105 A cm−2 which is achievable in Peltier effect can improve the cooling performance of thin
thin film thermoelectric elements. film InGaAs microrefrigerators by 700% at 77 K.
At low temperatures the linear part of the Peltier coeffi-
cient decreases significantly. However, MC simulations show This work was supported by ONR MURI Thermionic
that the nonlinear part of the Peltier coefficient does not Energy Conversion Center.
change as much. Therefore the nonlinear contribution be- 1
comes important in analyzing the efficiency of cryogenic O. Kulik, J. Phys.: Condens. Matter 6, 9737 共1994兲.
2
A. N. Grigorenko, P. I. Nikitin, D. A. Jelski, and T. F. George, J. Appl.
solid state coolers and it can enhance their performance. The Phys. 69, 3375 共1991兲.
temperature difference created along a bulk sample due to an 3
A. N. Grigorenko, P. I. Nikitin, D. A. Jelski, and T. F. George, Phys. Rev.
applied current can be obtained by B 42, 7405 共1990兲.

冉 冊
4
J. K. Freericks and V. Zlatic, Condens. Matter Phys. 9, 603 共2006兲.
d 1 5
E. N. Bogachek, A. G. Scherbakov, and U. Landman, Phys. Rev. B 60,
⌬T = − R1AJ2 + ⌸1J 共linear兲, 11678 共1999兲.
k 2 6
M. A. Çipiloğlu, S. Turgut, and M. Tomak, Phys. Status Solidi B 241,

⌬T =
d
k
冋1
− 共R1 + R3J2兲AJ2 + 共⌸1 + ⌸3J2兲J
2
册 共nonlinear兲,
7

8
2575 共2004兲.
A. S. Dzurak, C. G. Smith, L. Martin-Moreno, D. A. Ritchie, G. A. C.
Jones, and D. G. Hasku, J. Phys.: Condens. Matter 5, 8055 共1993兲.
P. Reddy, S. Y. Jang, R. A. Segalman, and A. Majumdar, Science 315,
共12兲 1568 共2007兲. In the supplementary material, the measured thermovoltage
as a function of temperature gradient is clearly nonlinear when a tempera-
where A is the area, k is the thermal conductivity, R is the ture difference of 20– 30 ° C is applied across single long-chain molecules.
resistance, ⌸1 is the linear Peltier, and ⌸3 is the third order 9
M. Zebarjadi, A. Shakouri, and K. Esfarjani, Phys. Rev. B 74, 195331
Peltier coefficient 共the total Peltier coefficient is ⌸ = ⌸1 10
共2006兲.
+ ⌸3J2兲. Figure 3 shows the effect of including nonlinear M. Zebarjadi, C. Bulutay, K. Esfarjani, and A. Shakouri, Appl. Phys. Lett.
90, 092111 共2007兲.
contribution of the Peltier coefficient in the calculation of the 11
M. Lundstrom, Fundamentals of Carrier Transport, 2nd ed. 共Cambridge
cooling curve at room temperature and a low temperature of University Press, Cambridge, UK, 2000兲, Chap. 3, p. 132.
12
77 K. According to the linear transport theory optimum dop- P. Lugli and D. K. Ferry, IEEE Trans. Electron Devices 32, 2431 共1985兲.

Anda mungkin juga menyukai