Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Improved Thermoelectric Cooling Based on the Thomson Effect

G. Jeffrey Snyder1 , Eric S. Toberer2 , Raghav Khanna1 , Wolfgang Seifert3


1
Materials Science, California Institute of Technology,
1200 E. California Blvd. Pasadena, CA 91125, USA
2
Department of Physics, Colorado School of Mines, Golden CO 80401, USA and
3
Institute of Physics, University Halle-Wittenberg, D-06099 Halle, Germany
(Dated: June 19, 2012)
Traditional thermoelectric Peltier coolers exhibit a cooling limit which is primarily determined by
the figure of merit, zT. Rather than a fundamental thermodynamic limit, this bound can be traced
to the difficulty of maintaining thermoelectric compatibility. Self-compatibility locally maximizes
the coolers coefficient of performance for a given zT and can be achieved by adjusting the relative
arXiv:1111.5300v3 [cond-mat.mtrl-sci] 16 Jun 2012

ratio of the thermoelectric transport properties that make up zT . In this study, we investigate the
theoretical performance of thermoelectric coolers that maintain self-compatibility across the device.
We find such a device behaves very differently from a Peltier cooler, and term self-compatible
coolers Thomson coolers when the Fourier heat divergence is dominated by the Thomson, as
opposed to the Joule, term. A Thomson cooler requires an exponentially rising Seebeck coefficient
with increasing temperature, while traditional Peltier coolers, such as those used commercially, have
comparatively minimal change in Seebeck coefficient with temperature. When reasonable material
property bounds are placed on the thermoelectric leg, the Thomson cooler is predicted to achieve
approximately twice the maximum temperature drop of a traditional Peltier cooler with equivalent
figure of merit (zT ). We anticipate the development of Thomson coolers will ultimately lead to solid
state cooling to cryogenic temperatures.
PACS numbers: 84.60.Rb, 05.70.Ce, 72.20.Pa, 85.80.Fi

1. INTRODUCTION material in the stage. In practice, the thermal losses


and complications of fabrication limit the performance
Peltier coolers are the most widely used solid state of such devices. The 6-stage cooler of Marlow achieves
cooling devices, enabling a wide range of applications a Tmax of 133 K; this doubling of Tmax compared to
from thermal management of optoelectronics and infra- a single stage cooler is achieved despite using materials
red detector arrays to polymerase chain reaction (PCR) with similar zT [3]. Alternatively, such Tmax with a
instruments. Thermoelectric coolers have been tradition- single-stage CPM cooler would require ZT to be 2.5.
ally understood by means of the Peltier effect, which de- The transport properties across a single thermoelec-
scribes the reversible heat transported by an electric cur- tric leg can be manipulated to improve cooling perfor-
rent. This effect is traditionally understood in terms of mance, although it has been less effective in reducing
absorption or release of heat at the junction of two dis- Tmax than a multi-stage approach. One common strat-
similar materials. The conventional analysis of a Peltier egy is to engineer a change in extrinsic dopant concen-
cooler approximates the material properties as indepen- tration across a thermoelectric element which can sig-
dent of temperature (Constant Property Model (CPM)). nificantly alter , and even . For example, this has
This results in a maximum cooling temperature differ- been demonstrated for thermoelectric generators in n-
ence Tmax for a CPM cooler, which dependent on the type PbTe doped with I [4]. Similar efforts have been
figure of merit ZT of the device [1, 2]. done with cooling materials, as has been reviewed in ref
[5]. The simplest explanation for an improvement is an
ZTc2 in increase in the local zT at some temperatures by spa-
Tmax = (1)
2 tially adjusting the dopant composition within a material
For the best commercial materials this leads to a [6].
Tmax of 65K (single stage) [3], which translates to a Early theoretical work by Sherman et al for TEC found
device ZT at 300K of 0.74. In the CPM the device ZT that different Tmax could be predicted from materials
is equal to the material zT . Material zT depends on the have the same or similar average zT but different temper-
Seebeck coefficient (), temperature (T ), electrical resis- ature dependence of the individual properties , , [7].
2
tivity (), and thermal conductivity (), zT = T . In This demonstrated that optimizing cooler performance is
the CPM, the only way to increase Tmax for a single significantly more complex than simply maximizing zT .
stage is to increase zT , leading to the focus of much ther- More recently, Muller et al. [810] and Bian et al. [11, 12]
moelectric research on improving zT . It is well known used different numerical approaches to predict substan-
that even further cooling to lower temperatures can be tial gains in cooling to Tmax from functionally grading
achieved using multi-stage Peltier coolers [1, 2]. In prin- where an average zT remains constant in an effort to
ciple, each stage can produce additional cooling to lower determine the best approach to functionally grading.
temperatures, regardless of the zT of the thermoelectric Different material classes optimized for different tem-
2

0.1 a) 3
u=s
Th = 300K

reduced COP (r)


2 Bi2Te3

COP ()
CPM
u=s
1
u=s
r < r,max 0
CPM Tmax

0 0 50 100 150
0 100 200 300 T (K)
Relative current density (-u) b) 0.2

reduced COP (r)


Fig. 1: The local reduced coefficient of performance r is
optimized at a specific reduced current density, termed s. If
u=s
u 6= s, the r is less than that predicted by the material zT . 0.1
Here, z = 0.002 K1 , = 200V K 1 .
Bi2Te3

CPM

peratures can also be segmented together to improve per- 0


240 260 280 300
formance of thermoelectric generators but the current Temperature (K)
must also be matched [13]. The analysis of segmenta-
tion strikingly demonstrates that increasing the average
Fig. 2: a) The CPM Peltier cooler and u = s Thomson cooler
zT does not always lead to an increase in overall thermo-
are compared using the same constant z = 0.002 K1 . The
electric efficiency and so an understanding of the thermo- overall device of a CPM cooler crosses zero at a finite tem-
electric compatibility factor is needed to explain device perature, indicating Tmax is reached, while remains pos-
performance [14]. itive for all temperatures for the u = s cooler. b) The local
This paper derives the cooling limit for a single stage, performance of a CPM cooler (r ) is significantly compro-
fully optimized (self-compatible) TEC that functions as mised at both the hot and cold ends. In contrast, r,max is
an infinitely staged cooler. The Fourier heat divergence achieved at all temperatures when u = s. In both panels, the
in such an optimized cooler is found to be dominated performance calculated for an actual Bi2 Te3 Peltier cooler leg
by the Thomson effect rather than the Joule heating is similar to the CPM.
as in traditional Peltier coolers. This new opportunity
presents a new challenge for material optimization based The compatibility approach to optimizing thermoelec-
on compatibility factor rather than only zT . tric cooling arises naturally from an analysis of the ther-
mal and electric transport equations. This method has
been described in detail for thermoelectric generators [15]
2. THEORY and coolers[19] and are reproduced here for TEC. The
method has been experimentally verified [20] and shown
Coolers are characterized by the coefficient of perfor- to reproduce results using a more traditional finite el-
mance ( = Qc /P ), which relates the rate of heat extrac- ement results but with less computational complexity.
tion at the cold end Qc to the power consumption P in This method has been incorporated into several engineer-
the device [15]. For simplicity, but without loss of gener- ing models such as those used by NASA for Radioisotope
ality, a single thermoelectric element can be considered Thermoelectric Generators [21] and Amerigon/BSST for
rather than a complete device. A TEC leg can be treated automotive applications [20, 22]. Consider an infinites-
as an infinite series of infinitesimal coolers, each of which imal section of thermoelectric leg in a temperature gra-
is operating locally with some COP. Scaling this COP to dient and an electric field. The temperature gradient
the local Carnot COP (T /dT ) yields the local reduced will induce a Fourier heat flux (q = T ) across this
coefficient of performance r . [16]. This relationship be- segment. The divergence of this heat (Eq. 3) is equal to
tween local performance across the leg and global COP, the source terms: irreversible Joule heating (j 2 ) and
d
, given in Eq. 2 is derived in the Appendix based on the reversible Thomson heat (T dT jT ), both of which
Ref.[7][17]. While TECs are traditionally analysed using depend on the electric current density (j). From these
a global approach, we have previously shown the utility of two effects, the governing equation for heat flow in vector
a local approach [14, 15, 18]. This local approach leads to notation is
a consideration of material compatibility, as discussed
below.
q = (T ) = j 2 j T (3)
Z Th !
1 1 1 with Joule heat per volume j 2 , Thomson coefficient
= exp dT 1 (2) d
Tc T r (T ) = T dT and Thomson heat per volume j T . The
3

Peltier, Seebeck and Thomson effect are all manifesta- The maximum local r , denoted r,max , occurs when
tions of the same thermoelectric property characterized u = s. The expression for r,max (Eq. 10) is an explicit
d
by . The Thomson coefficient ( = T dT ) describes the function of the material zT and is independent of the
Thomson heat absorbed or released when current flows individual properties , , . This maximum allowable
in the direction of a temperature gradient. local efficiency provides a natural justification for the def-
Restricting the problem to one spatial dimension, Eq. inition of zT as the materials figure of merit.
(3) is typically examined assuming the heat flux and elec-
tric current are parallel [15]. In the typical CPM model
1 + zT 1
used to analyze Peltier coolers, the Thomson effect is zero r,max = (10)
because is constant along the leg ( dTd
= 0). 1 + zT + 1
The exact performance of a thermoelectric leg with One thus wishes to construct devices where, locally, each
(T ), (T ), and (T ) possessing arbitrary temperature segment has u = s and thus r,max is obtained. Glob-
dependence can be straightforwardly computed using the ally, maximum is found when the entire cooler satisfies
reduced variables: relative current density (u) and ther- u = s.
moelectric potential () [18]. The relative current den-
sity u, given in Eq. 4, is primarily determined by the
electrical current density j, which is adjusted to achieve 3. COOLING PERFORMANCE
maximum global COP. The thermoelectric potential is
a state function which simplifies Eq. 2 to Eq. 6 [15]. To compare the cooling performance of traditional
2 Peltier coolers and u = s coolers, we consider coolers
j with equivalent z. Traditional Peltier coolers have typ-
u= (4)
T j ically been analyzed with the constant property model
(CPM), yielding a constant z (where zT is linearly in-
creasing with temperature). We will show that constant
= T + 1/u (5) z, but allowing , , to vary with T , can lead to sub-
stantial improvement in cooling. At the limit of this
variation, we will assume the properties can be varied
to satisfy u = s.
(Tc )
= (6) Performance of a CPM cooler CPM coolers have
(Th ) (Tc ) been extensively studied, typically using a global ap-
proach to the transport behavior. The for a CPM
Changing variables to T via the monotonic function
cooler (operated at optimum j) is given by Eq. 11 [24].
x(T ), Eq. 3 simplifies to the differential equation in u(T ).
Figure 2a shows the of a CPM cooler decreases with
du

d 2
 increasing T . With increasing cooling, this decreases
= u2 T + u (7) and reaches zero at Tmax (Eq. 1).
dT dT z
!
1 + zTavg TThc
p
Using this formalism, the reduced coefficient of perfor- 
Tc

CP M
mance (r ) can be simply defined for any point in the = p (11)
T 1 + zTavg + 1
cooler (Eq. 8). Fig. 1 shows this relationship between u
and r . From Eq. 2, it can be shown that is largest To understand what is limiting the CPM cooler at
when r is maximized for every infinitesimal segment Tmax , we derive the local reduced coefficient of per-
along the cooler. Hence, global maximization can be formance CP M
(T ). To obtain CP M
we need u as a
r r
traced back to local optimization [23]. function of T . The solution to differential equation 7 for
CPM is
u z + z1T u + T1 1 1 22
r = = (8) = + (Th T ) (12)
u z (1 u z ) u( u) u(T )2 u2h z
where the value of u at T = Th (uh ) serves as an ini-
The optimum u which maximizes r ( d du = 0) can
r
tial condition. This expression allows u(T ) to be de-
be expressed solely in terms of local material properties
termined for any CPM cooler, regardless of temperature
(Eq. 9). This optimum value of u is defined as the ther-
drop (T Tmax ) and applied current density (j).
moelectric compatibility factor sc for coolers.
The global maximum COP () is obtained when the op-
timum uh from Eq. 13 is employed.
1 + zT 1
sc = (9)
T
1 zTc2 2(Th Tc )
As this paper strictly focuses on coolers, we will refer to = q (13)
uh z T + T z( Th +Tc ) + 1
sc as simply s. h c 2
4

Consideration of Eq. 13 reveals that the maximum Tc 1000


is obtained when 1/uh approaches zero. Figure 3 shows u=s
|u| becoming infinite at Th for the CPM cooler. In this CPM
Bi2Te3

|u| and |s| (V )


-1
limit, Eq. 13 can be simplified to give Eq. 1 with Z = z. u&s
Thus, a local approach to transport yields the classic
CPM limit typically obtained through an evaluation of 100
global transport behavior. s
Combining Eq. 8, 12, and 13 results in r (T ) at Tmax u
for the CPM Peltier cooler (Eq. 14). This expression re-
veals r drops to zero at both ends of the CPM cooler leg,
10
as shown in Figure 2b. This prohibits additional cooling 240 260 280 300
and sets Tmax . Temperature (K)

2z(Th T ) 2 ThTT
p
Fig. 3: Both the CPM and Bi2 Te3 coolers have u = s at one
CP M
r,Tmax = p (14) point along the leg. By definition, the u = s is self-compatibile
1 + 2z(Th T )
along the entire leg. The different slope signs of u for CPM
To achieve cryogenic cooling (Tc 0) within the CPM, and u = s reveals that these coolers are fundamentally dis-
zT must approach infinity (Eq. 1). For example, cooling tinct. The curves were generated for an optimized cooler at
Tmax with z = 0.002 K1 , Th = 300 K.
with a single-stage CPM cooler to 10 K would require
zT to be over 1000 if the hot side is 300 K. When is
negative, the net effect of the thermoelectric device is to
supply heat, rather than remove heat, from the cold side. leg. r for this cooler is simply given by Eq. 10. This r
For negative values for the CPM cooler to be obtained is found to be positive for all T , as z is always a posi-
requires certain parts of the cooler to locally possess r < tive real number. Globally, this translates to the analytic
0. Such a result may be surprising at first as this r < 0 maximum for for a cooler where z is defined and lim-
region is made from material possessing positive zT . This ited.
seems particularly odd when compared to the behavior of
staged generators, discussed above. Clearly, single- and To facilitate comparison with CPM, we consider a con-
multi-staged CPM legs exhibit fundamentally different stant z model where the individual properties are ad-
behavior, despite being composed of exactly the same justed to maintain u = s. The constant z approach
material. Such behavior can be rationalized using the yields vanishing zT at low T , consistent with real ma-
thermoelectric compatibility concept. terials. Evaluating (Eq. 2) for a u = s cooler and
Figure 3 shows that the compatibility condition (u = s) the assumption
of constant z, one obtains Eq. 15, where
is maintained at only one point in the CPM cooler. Con- Mi = 1 + zTi with Ti = Th , Tc .
sequently, CPM coolers operate inefficiently (u 6= s) at
both the hot and cold ends. This is demonstrated in Fig-  2  
1 Mh 1 2(Mh Mc )
ure 2b, where r < r,max for all but one point. Once r = exp 1 (15)
goes below zero at low temperature, the thermoelectric u=s Mc 1 (Mh 1)(Mc 1)
device is no longer cooling the cold end and Tmax is
reached (Figure 2a). Inspection of Eq. 15, where Mh > Mc > 1, reveals that
While real coolers do not possess temperature- is always greater than zero for a u = s cooler.
independent properties, the qualitative results for CPM
translate well to traditional Peltier coolers due to their The difference between CPM and u = s coolers can
weak material gradients. Considering a Bi2 Te3 leg with be visualized in Fig. 2a, with the of the Thomson
temperature-dependent properties described in Ref. [25], cooler asymptotically approaching zero with increasing
we find u and s to be quite close to a z-matched CPM T . Figure 2b shows that r for a self-compatible cooler
cooler (Figure 3). Like the CPM cooler, u = s at only one with constant z remains finite and positive throughout
temperature along the leg. This leads to similar r (T ) the device. In contrast, the CPM cooler is operating
for the Bi2 Te3 and CPM coolers, shown in Figure 2b. inefficiently at both the hot and cold ends, limiting its
Within the CPM, large zT results in a high upper temperature range.
limit to r but does not ensure this r,max is achieved. In principle, if u = s can be maintained, the idealized
Generaly, commercial cooling materials such as Bi2 Te3 u = s cooler can achieve an arbitrarily low cold side
and any material that can be described by the CPM temperature as long as the all of the materials have
model will be operating significantly below the r a finite zT . However, the material requirements to
predicted by the zT they possess (Eq. 8, 10). maintain u = s become exceedingly difficult to achieve
as the cooling temperature is reduced and the ultimate
Performance of a u = s cooler We now consider an cooling will be finite, yielding Tc > 0.
idealized cooler which maintains u = s across the entire
5

4. MATERIAL REQUIREMENTS A lower bound to c also arises from the interconnected


nature of the transport properties. We require zT to be
A CPM cooler has a fixed z and performance which finite; thus the electrical conductivity must be large as
is independent of the ratio of individual properties as c tends to zero. In this limit, the electronic component
long as they are constant with respect to temperature of the thermal conductivity (E ) is much larger than the
(Eq. 1). In contrast, a u = s cooler requires dramatic lattice (L ) contribution and E . To satisfy the
2 2
changes in properties with temperature to maintain self- Wiedemann-Franz law (E = LT where L = 3 ke2 is
compatibility. Within the constraint of constant z, con- the Lorenz factor in the free electron limit), c has a
1
sideration of Eq. 9 suggests that the Seebeck coefficient lower bound given by Eq. 20. For example, a z = 300
must be varied across the
p device to maintain u = s. Ad- K1 and Tc = 175 K results in a lower bound to c of
ditionally, as (T ) = z(T )(T ) within a constant z 119 V/K.
model, the product (T )(T ) must also vary across the
device.
The Seebeck coefficient profile (T ) for a u = s cooler 2 kB
2
c2 = LzTc = zTc (20)
with constant z can be solved analytically. Combining 3 e2
Eq. 7 and u = s yields the simple differential equation of The maximum cooling temperature Tc can be solved as
(T ): a function of z, Eg and Th from equations Eq. 17, Eq. 19
  and Eq. 20. For small z the approximate solution
d T d 1+ 1+zT
=T . (16) !
dT 1 + 1 + z T dT z T z 2 Eg2
T Th ln 4 2 2 3
(21)
Solving this equation yields 8 3 kB z Th

gives an indication of the important parameters but
 
1 + zT 1 2
(T ) = 0 exp . (17) quickly becomes inaccurate for zT above 0.1.
1 + zT 1 + zT 1
Material limits to performance With these bounds on
With this expression for (T ), it is possible to eval- material properties, we consider the Tmax of a u = s
uate s(T ) with Eq. 9. Figure 3 shows the variation in s cooler. Figure 2 suggests that the of a u = s cooler
required for a u = s cooler with constant z. The self- remains positive for all temperature. However, obtaining
compatible cooler modeled in Figure 3 has z = 0.002; materials with the required properties limits Tmax to a
60 K of cooling results in a change in s of one order of finite value. Fig. 4 compares the Tmax solution for u =
magnitude. s and CPM coolers with the same z. Here, the maximum
The approximation for small zT yield a simple expres- Seebeck coefficient is set by the band gap (Eg = 13 eV ),
sion for (T ), given by Eq. 18. per Eq. 19. The u = s cooler provides significantly higher
  Tmax than the CPM cooler with the same zT , nearly
d 4 4 twice the Tmax for Eg = 3 eV .
(ln (T )) = (T ) exp (18) Spatial dependence of material properties These an-
dT zT 2 zT
alytic results are possible because the compatibility ap-
This reveals that should be very large at the hot end proach does not require an exact knowledge of the spatial
and must decrease to a low value at the cold end. This profile for the material properties. Nevertheless, it is pos-
exponentially varying (T ) required to maintain u = s sible solve for the spatial dependence of the u = s cooler,
for constant z is anticipated to be the limiting factor in given some material constraints. To determine x(T ), we
real coolers and place bounds on the maximum cooling integrate Eq. 4, recalling we have assumed constant cross-
obtainable. We consider the realistic range of below. sectional area (j(x) = const.), obtaining Eq. 22.
Large values of are found in lightly doped semicon-
Th
1
Z
ductors and insulators with large band gaps (Eg ) that
x(T ) = udT (22)
effectively have only one carrier type, thereby preventing j T
compensated thermopower from two oppositely charged
conducting species. Using the relationship between peak Thus, the natural approach to cooler design within the
and Eg of Goldsmid (Eq. 19) allows an estimate for the u = s approach is to determine the temperature depen-
highest (Th ) we might expect at the hot end, h [26]. dence of the material properties, and then determine the
Good thermoelectric materials with band gap of 1 eV are required spatial dependence from the resulting u(T ) and
common while 3 eV should be feasible. For a cooler with (T ).
an ambient hot side temperature, this would suggest h Figure 5a shows an example of the Seebeck distribution
should be 1-5 mV/K. Maintaining zT at such large (x) along the leg that will provide the necessary (T ),
will require materials with both extremely high electronic where a constant L = 0.5 W/mK is assumed. The of
mobility and low lattice thermal conductivity. Figure 5a spans the range permitted by Eq. 19 and 20.
In a real device the spatial profile of thermoelectric
h = Eg /(2eTh ) (19) properties will need to be carefully engineered. If this
6

rapidly changing (x) is achieved by segmenting differ- 5. COOLER PHASE SPACE


ent materials, low electrical contact resistance is required
between the interfaces. We anticipate such control of The improved performance of a u = s cooler is not
semiconductor materials may require thin film methods simply an incremental improvement, but rather we find
on active bulk thermoelectric substrates. CPM and u = s coolers operate in fundamentally differ-
ent phase-spaces. Here, by phase space we refer to the
class of solutions defined by the sign of the Fourier heat
200 divergence ( q in Equation 3). The Fourier heat di-
Eg vergence in a cooler contains both the Joule (j 2 ) and
150 3 eV Thomson ( j T ) terms.
2 eV
We begin by considering the Fourier heat divergence in
Tmax (K)

1 eV u=s
CPM and Bi2 Te3 coolers and then compare this behavior
100 to u = s coolers. In the typical CPM model to analyze
CPM Peltier coolers, = 0 as there is no variation in . In a
50 CPM cooler, q is thus greater than zero. This can
Bi2Te3
be seen by the downward concavity of the temperature
0 distribution in Figure 5b. In a typical Peltier cooler (e.g.
0 0.2 0.4 0.6 0.8 1 1.2 1.4 Bi2 Te3 ), the concavity is the same as the CPM cooler and
zTh
thus the divergence is likewise positive. This is because
the Thomson term is always less than the Joule term in
Fig. 4: The maximum temperature drop Tmax of a u = s a conventional thermoelectric cooler.
Thomson cooler exceeds that of a Peltier cooler with the same
z. Large band gap, Eg , thermoelectric materials are necessary
In contrast, a u = s cooler changes the sign of the
at the hot junction improves the performance (Th = 300K). Fourier heat divergence such that q is less than zero.
This can be readily visualized in Figure 5b, where the
concavity of the u = s cooler is opposite the CPM and
Bi2 Te3 coolers. This difference in concavity must come
from the Thomson term being positive and greater than
the Joule heating term. The large magnitude of Thom-
son term is understandable with the exponentially rising
a) Seebeck coefficient seen in Figure 5a. The reversibility of
the Thomson effect requires that for q to be less than
Seebeck coef. (mV/K)

1
u=s zero, the hot end must have a high || relative to the cold
end, and not vice versa. This translates to a requirement
Bi2Te3
for such that j T > j 2 .
We can also express the Fourier heat divergence in
CPM
0.1 terms of reduced variables.

0 0.2 0.4 0.6 0.8 1 1 1 du


Distance along leg q = j = 2 j T (23)
u u dT
b)
Manipulation with Eq. 4 produces a form where the
Temperature (K)

300 sign of u and directions of j and T are irrelevant.


CPM
Bi2Te3 j2 d 2
u=s q = u (24)
2u4 dT
200

Thus the sign of q is determined by the sign of d|u|


dT ,
0 0.2 0.4 0.6 0.8 1 which is valid for both p and n-type elements regardless
Tc Distance along leg Th
of the sign of u.
The Fourier heat divergence criterion is a convenient
Fig. 5: a) The Seebeck coefficient of the u = s Thomson definition to distinguish these two regions of thermoelec-
cooler varies exponentially, while it is by definition constant tric cooling in experimental data. The Peltier cooling
for the CPM cooler. As a degenerately doped semiconductor, region, defined by q > 0, is found in the phase space
the Seebeck coefficient of commercial Bi2 Te3 increases gradu-
ally with temperature. b) The curvature of T (x) for the CPM where d|u|
dT > 0. Likewise, the Thomson cooling region
Peltier cooler temperature profile is opposite that of the u = s defined by q < 0 is the phase space where d|u|dT < 0.
cooler because of the different sign of the Fourier heat diver- The constant relative current u(T ) = const. separates
gence. Again, similar behavior is found between the CPM the Thomson-type and from the Peltier-type solutions to
and commercial Bi2 Te3 cooler. the differential equation. In Figure 3, the CPM and u = s
7

cooler have opposite slopes, indicating these coolers ex- cooling than increasing zT .
ist in separate regions of the cooling phase space. This Minor improvements in thermoelectric cooling beyond
result is consistent with our discussion above concerning increasing average zT by increasing the Thomson effect
the concavity of T (x) in Figure 5. in a functionally graded material were predicted as early
For clarity, we suggest coolers which are predominately as 1960 [7]. Similarly Muller et al. describe modest gains
in the Thomson phase-space, q < 0 but may not have in cooling from functionally grading [810] where mate-
u = s be referred to as Thomson coolers. Similarly, rial properties are allowed to vary in a constrained way
Peltier coolers should refer to coolers operating in the such that the average zT remains constant. Such addi-
usual q > 0 Fourier heat divergence phase-space tional constraints can keep the analysis within the Peltier
where Joule heating dominates. region, preventing a full optimization to a u = s solution.
This understanding of phase space for u = s and CPM Bian et al, [11, 12] propose a thermoelectric cooler with
coolers enables us to hypothesize that the performance significantly enhanced Tmax using a rapidly changing
advantages of u = s coolers extends to imperfect Thom- Seebeck coefficient in at least one region. The method
son coolers. We expect such coolers possess two primary of Bian et al focuses on the redistribution of the Joule
advantages over traditional Peltier coolers. First, for a heat rather than a consideration of the Thomson heat
given material zT , performance (Tmax and ) of the or the effect of compatibility. Nevertheless, the region of
Thomson cooler is greater (Figure 2). The Tmax solu- rapidly changing Seebeck coefficient would also create a
tion for the u = s cooler is compared to a Peltier cooler significant Thomson effect and likely place that segment
with the same material assumption for z in Fig. 4. Here, of the cooler in the Thomson cooler phase space while
the maximum Seebeck coefficient is set by the band gap other segments would function like a CPM Peltier cooler.
(Eg = 1 3 eV ), per Eq. 19. The Thomson cooler pro- In a traditional single-stage (or segmented) thermo-
vides significantly higher Tmax than the Peltier cooler electric device, the current flow and the heat flow are
with the same zT , nearly twice the Tmax for Eg = 3 eV . collinear and flow through the same length and cross-
Second, in a Thomson cooler, the temperature minimum sectional area of thermoelement. This leads to the com-
is not limited by zT explicitly like it is in a traditional patibility requirement between the optimal current den-
Peltier coolers. sity and optimal heat flux to achieve optimal efficiency.
In a multi-stage (cascaded) device the thermal and elec-
trical circuits become independent and so the compati-
6. DISCUSSION bility requirement is avoided between stages.
A transverse Peltier cooler also decouples the current
Efficiency improvements from staging and maintaining and heat flow by having them transport in perpendicular
u = s also exists for thermoelectric generators, but the directions. Gudkin showed theoretically that the trans-
improvement is small (< 10% compared to CPM). This verse thermoelectric cooler, could function as an infinite
is because the u does not typically vary by more than a cascade by the appropriate geometrical shaping of the
factor of two across the device. However, in a TEC the thermoelement [27]. The different directions of heat and
compatibility requirement is much more critical. When current flow enable, in principle, an adjustment of geom-
operating a TEC to maximum temperature difference, etry to keep both heat and electric current flow indepen-
the temperature gradient varies from zero to very high dently optimized. Cooling of 23 K using a rectangular
values, which means u will have a much broader range block was increased to 35 K using a trapezoidal cross-
(Figure 3) in a TEC than in a generator. Thus, unless section [28].
compatibility is specifically considered, the poor compat-
ibility will greatly reduce the performance of the thermo-
electric cooler, and this results in the Tmax limit well 7. CONCLUSION
known for Peltier coolers.
In real materials, changing material composition also Here, we compare self-compatible coolers with CPM
changes zT so the effect of maximizing average zT is and commercial Bi2 Te3 thermoelectric coolers. Signif-
difficult to decouple from the effect of compatibility. As icant improvements in cooling efficiency and maximum
such, efforts which are focused on maximizing zT will cooling are achieved for equivalent z when the cooler is
generally fail to create a material with u = s and may self-compatible. Such improvement is most pronounced
only marginally increase Tmax . Conversely, focusing on when the goal is to achieve maximum temperature dif-
u = s without consideration of zT could rapidly lead to ference, rather than high coefficient of performance at
unrealistic materials requirements. small temperature difference. Optimum material profiles
In this new analysis we have focused on the compat- are derived for self-compatible Thomson coolers and re-
ibility criterion, u = s, with constant z (as opposed to alistic material constraints are used to bound the per-
zT [16]) to demonstrate the differences between a Thom- formance. Self-compatible coolers are found to operate
son and a Peltier cooler typically analyzed with the CPM in a fundamentally distinct phase space from traditional
model. Generally, achieving u = s in a material with fi- Peltier coolers. The Fourier heat divergence of Thomson
nite zT , is more important to achieve low temperature coolers is dominated by the Thomson effect, while this
8

divergence in Peltier coolers is of the opposite sign, in- By conservation of energy (or recursive relation Eq 25),
dicating the Joule heating is the dominant effect. This the heat exiting the n pumps, Qn is the heat pumped by
analysis opens a new strategy for solid state cooling and the first pump plus the total power, Eq 29.
creates new challenges for material optimization.
Qn = Q0 + P (31)
Acknowledgments

Combining equations 30, 31, and 28, it is straightfor-


We thank AFOSR MURI FA9550-10-1-0533 for sup-
ward to show
port. E. S. T. acknowledges support from the U.S. Na-
tional Science Foundation MRSEC program - REMR-
n
SEC Center, Grant No. DMR 0820518. 1 Y 1
(1 + )= (1 + ) (32)
i=1
i

8. APPENDIX
This can be transformed into a summation by use of a
The metric for summing the efficiency and coefficient natural logarithm
of performance of these thermodynamic processes is not
a simple summation because energy is continuously be- n
ing supplied or removed so that neither the heat nor en- 1 X 1
ln(1 + )= ln(1 + ) (33)
ergy flow is a constant. The derivation, attributed to i=1
i
Zener[17], of the coefficient of the performance Eq. 2 and
efficiency summation metric for a continuous system in
one-dimension given here is based on Zener and similar While 1/i should become very small with small Ti =
derivations given in [7][29][30][19]. Ti Ti1 the reduced coefficient of performance r,i
Consider n heat pumps (or heat engines) connected in should remain finite
series such that the heat entering the ith pump, Qi , is the
same as the heat exiting the i 1 pump, namely Qi1 . 1 1 Ti
Then by conservation of energy = (34)
i r,i Ti
Qi = Qi1 + Pi (25)
Assuming a monotonic temperature distribution (for a
where Pi is the power entering the pump i. The coeffi- simple TEC Ti > Ti1 for all i) the sum in Eq. 33 can be
cient of performance n of the pump n is defined by converted to an integral in the limit that n , where
1 Pi Ti = (Th Tc )/n and ln(1 + x) x for small x
= (26)
i Qi1
Z Th
1 1
This gives the recursive relation ln(1 + )= dT (35)
Tc T r (T )
1
Qi = Qi1 (1 + ) (27)
i If the temperature distribution is not monotonic but can
which can be solved for Qn as be divided up into monotonic segments, each of these
monotonic segments can be individually transformed into
n
Y 1 integrals.
Qn = Q0 (1 + ) (28)
i For a generator (as opposed to a cooler or heat pump)
i=1
power is extracted rather than added at each segment
The total power added to the system P from the n in the series. Then the sign of the power in equation
pumps is 25 is negative for a generator with the efficiency given
by = P/Q0 = 1/ and reduced efficiency r =
n
X 1/r . Thus the above method can be used to derive
P = Pi (29) the analogous equation for generator efficiency:
i=1

The coefficient of performance for the n pumps, , with Z Th


r
heat Qc = Q0 entering from the cold side is defined by ln(1 ) = dT (36)
Tc T
1 P
= (30)
Q0
9

[1] Goldsmid, H. J. Introduction to Thermoelectricity and technology. In Conference on Thermoelectricity., 3


(Springer-Verlag, 2010). 22 (Wiley, New York, 1960).
[2] Heikes, R. R. & Ure, R. W. Thermoelectricity: Science [18] Snyder, G. J. & Ursell, T. S. Thermoelectric efficiency
and Engineering. (Interscience, New York, 1961). and compatibility. Phys. Rev. Lett. 91, 148301 (2003).
[3] Marlow Industries, Inc. Six state - thermoelectric mod- [19] Seifert, W., Muller, E. & Walczak, S. Generalized ana-
ules. http://www.marlow.com (2011). lytic description of onedimensional non-homogeneous TE
[4] Gelbstein, Y., Dashevsky, Z. & Dariel, M. High perfor- cooler and generator elements based on the compatibility
mance n-type PbTe-based materials for thermoelectric approach. In Rogl, P. (ed.) 25th International Conference
applications. Physica B: Condensed Matter 363, 196 on Thermoelectrics, 714 719 (IEEE, Picataway, 2006).
205 (2005). [20] Crane, D. T. & Bell, L. E. Design to maximize perfor-
[5] Kuznetsov, V. Functionally Graded Materials for Ther- mance of a thermoelectric power generator with a dy-
moelectric Applications (In: CRC Handbook of Thermo- namic thermal power source. Journal of Energy Re-
electrics, CRC Press, 2006). sources Technology 131, 012401 (2009). URL http:
[6] Mahan, G. D. Inhomogeneous thermoelectrics. Journal //link.aip.org/link/?JRG/131/012401/1.
of Applied Physics 70, 4551 (1991). [21] Wood, E. G. et al. Simulating the Gradually Deteriorat-
[7] Sherman, B., Heikes, R. & R.W. Ure, J. Calculation of ing Performance of an RTG. NASA Tech Briefs 45252,
efficiency of thermoelectric devices. J. Appl. Phys. 31, 1 5152 (2008).
16 (1960). [22] Crane, D. T. An introduction to system-level, steady-
[8] Muller, E., Drasar, C., Schilz, J. & Kaysser, W. A. Func- state and transient modeling and optimization of high-
tionally graded materials for sensor and energy applica- power-density thermoelectric generator devices made of
tions. Materials Science and Engineering A 362, 17 39 segmented thermoelectric elements. J. Elect. Mater. 40,
(2003). 561569 (2011).
[9] Muller, E., Walczak, S. & Seifert, W. Optimization [23] Seifert, W., Zabrocki, K., Snyder, G. & Muller, E. The
strategies for segmented Peltier coolers. phys. stat. sol. compatibility approach in the classical theory of thermo-
(a) 203, 2128 2141 (2006). electricity seen from the perspective of variational calcu-
[10] Muller, E., Karpinski, G., Wu, L. M., Walczak, S. & lus. phys. stat. sol. (a) 207, 760765 (2010).
Seifert, W. Separated effect of 1D thermoelectric mate- [24] Heikes, R. R. & Roland W. Ure, J. Thermoelectricity:
rial gradients. In Rogl, P. (ed.) 25th Int. Conf. Thermo- Science and Engineering (Interscience Publishers, Inc.,
electrics, 204 209 (IEEE, Picataway, 2006). New York, 1961).
[11] Bian, Z. & Shakouri, A. Beating the maximum cool- [25] Seifert, W., Ueltzen, M. & Muller, E. One-dimensional
ing limit with graded thermoelectric materials. Applied modelling of thermoelectric cooling. Phys. Stat. Sol. A
Physics Letters 89, 212101 (2006). 194, 277290 (2002).
[12] Bian, Z., Wang, H., Zhou, Q. & Shakouri, A. Maximum [26] Goldsmid, H. J. & Sharp, J. W. Estimation of the ther-
cooling temperature and uniform efficiency criterion for mal band gap of a semiconductor from seebeck measure-
inhomogeneous thermoelectric materials. Physical Re- ments. J. Electron. Mater. 28, 869 (1999).
view B 75, 245208 (2007). [27] Gudkin, T. S., Iordanishviliu, E. K. & Fiskind, E. E. Sov.
[13] Moizhes, B., Shishkin, Y. P., Petrov, A. & Kolomoet, Phys. Semicond. 11, 1048 (1977).
L. A. Sov. Phys. Tech. Phys. 7, 336 (1962). [28] Gudkin, T. S., Iordanishviliu, E. K. & Fiskind, E. E.
[14] Snyder, G. J. Application of the compatibility factor to With reference to theory of anisotropic thermoelectric
the design of segmented and cascaded thermoelectric gen- cooler. Sov. Phys. Tech, Phys. Lett. 4, 844 (1978).
erators. Applied Physics Letters 84, 24362438 (2004). [29] Freedman, S. I. Thermoelectric power generation. In
[15] Snyder, G. J. Thermoelectric Power Generation: Effi- Sutton, G. W. (ed.) Direct Energy Conversion, vol. 3
ciency and Compatibility (In: CRC Handbook of Ther- of Inter-University Electronic Series, chap. 3, 105180
moelectrics, CRC Press, 2006). (McGraw-Hill Book Company, 1966).
[16] Seifert, W. et al. Maximum performance in self- [30] Harman, T. C. & Honig, J. M. Thermoelectric and ther-
compatible thermoelectric elements. J. Mater. Res. 26, momagnetic effects and applications (McGraw-Hill Book
19331939 (2011). Company, New York, 1967).
[17] Zener, C. The impact of thermoelectricity upon science

You might also like