This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Paper The following article is Open access

Multi-terminal thermoelectric transport in a magnetic field: bounds on Onsager coefficients and efficiency

and

Published 8 October 2013 © IOP Publishing and Deutsche Physikalische Gesellschaft
, , Focus on Thermoelectric Effects in Nanostructures Citation Kay Brandner and Udo Seifert 2013 New J. Phys. 15 105003 DOI 10.1088/1367-2630/15/10/105003

1367-2630/15/10/105003

Abstract

Thermoelectric transport involving an arbitrary number of terminals is discussed in the presence of a magnetic field breaking time-reversal symmetry within the linear response regime using the Landauer–Büttiker formalism. We derive a universal bound on the Onsager coefficients that depends only on the number of terminals. This bound implies bounds on the efficiency and on efficiency at maximum power for heat engines and refrigerators. For isothermal engines pumping particles and for absorption refrigerators these bounds become independent even of the number of terminals. On a technical level, these results follow from an original algebraic analysis of the asymmetry index of doubly substochastic matrices and their Schur complements.

Export citation and abstract BibTeX RIS

Content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Thermoelectric devices use a coupling between heat and particle currents driven by local gradients in temperature and chemical potential to generate electrical power or for cooling [14]. Since they work without any moving parts, such machines have a lot of advantages compared to their cyclic counterparts, which rely on the periodic compression and expansion of a certain working fluid [5]. However, so far their notoriously modest efficiency prevents a wide-ranging applicability. Although it has been shown that proper energy filtering leads to highly efficient thermoelectric heat engines [6], which, in principle, may even reach Carnot efficiency [7, 8], so far no competitive devices that come even close to this limit are available. Consequently, the challenge of finding better thermoelectric materials has attracted a great amount of scientific interest during recent decades.

Recently, Benenti et al [9] discovered a new option to enhance the performance of thermoelectric engines. Their rather general analysis within the phenomenological framework of linear irreversible thermodynamics reveals that a magnetic field, which breaks time reversal symmetry, could enhance thermoelectric efficiency significantly. In principle, it even seems to be possible to obtain completely reversible transport, i.e. devices that work at Carnot efficiency while delivering finite power output. This spectacular observation prompts the question of whether this option can be realized in specific microscopic models.

An elementary and well established framework for the description of thermoelectric transport on a microscopic level is provided by the scattering approach originally pioneered by Landauer [10]. The basic idea behind this method is to connect two electronic reservoirs (terminals) of different temperature and chemical potential via perfect, infinitely long leads to a central scattering region. By assuming non-interacting electrons, which are transferred coherently between the terminals, it is possible to express the linear transport coefficients in terms of the scattering matrix that describes the motion of a single electron of energy E through the central region. Thus, the macroscopic transport process can be traced back to the microscopic dynamics of the electrons. This formalism can easily be extended to an arbitrary number of terminals [11, 12].

Within a purely coherent two-terminal set-up, current conservation requires a symmetric scattering matrix and hence a symmetric matrix of kinetic coefficients, even in the presence of a magnetic field [13]. Therefore, without inelastic scattering events the broken time-reversal symmetry is not visible on the macroscopic scale. An elegant way to simulate inelastic scattering within an inherently conservative system goes back to Büttiker [14]. He proposed to attach additional, so-called probe terminals to the scattering region, whose temperature and chemical potential are adjusted in such a way that they do not exchange any net quantities with the remaining terminals but only induce phase-breaking.

The arguably most simple case is to include only one probe terminal, which leads to a three-terminal model. Saito et al [15] pointed out that such a minimal set-up is sufficient to obtain a non-symmetric matrix of kinetic coefficients. However, we have shown in a preceding work on the three-terminal system [16] that current conservation puts a much stronger bound on the Onsager coefficients than the bare second law. It turned out that this new bound constrains the maximum efficiency of the model as a thermoelectric heat engine to be significantly smaller than the Carnot value as soon as the Onsager matrix becomes non-symmetric. Moreover, Balachandran et al [17] demonstrated by extensive numerical efforts that our bound is tight.

The strong bounds on Onsager coefficients and efficiency obtained within the three-terminal set-up raise the question of whether they persist if more terminals are included. This problem will be addressed in this paper. We will derive a universal bound on kinetic coefficients that depends only on the number of terminals and gets weaker as this number increases. Only in the limit of infinitely many terminals, this bound approaches the well-known one following from the positivity of entropy production. By specializing these results to thermoelectric transport between two real terminals with the other n − 2 acting as probe terminals, we obtain bounds on the efficiency and the efficiency at maximum power for different variants of thermoelectric devices like heat engines and cooling devices.

Our results follow from analyzing the matrix of kinetic coefficients in the n-terminal set-up and its subsequent specializations to two real and n − 2 probe terminals. On a technical level, we introduce an asymmetry index for a positive semi-definite matrix and compute it for the class of matrices characteristic for the scattering approach. These calculations involve a fair amount of original matrix algebra for doubly substochastic matrices and their Schur complements, which we develop in an extended and self-contained mathematical appendix.

The main part of the paper is organized as follows. In section 2, we introduce the multi-terminal model and recall the expressions for its kinetic coefficients. In section 3, we derive the new bounds on these coefficients. In section 4, we show how these bounds imply bounds on the efficiency and the efficiency at maximum power for heat engines, for refrigerators, for iso-thermal engines and for absorption refrigerators. In contrast to the former two classes, the latter two involve only one type of affinities, namely chemical potential or temperature differences, respectively which implies even stronger bounds. We conclude in section 5.

2. The multi-terminal model

We consider the set-up schematically shown in figure 1. A central scattering region equipped with a constant magnetic field B is connected to n independent electronic reservoirs (terminals) of respective temperature $T_1, \ldots , T_n$ and chemical potential $\mu _1, \ldots \mu _n$ . We assume non interacting electrons, which are transferred coherently between the terminals without any inelastic scattering. In order to describe the resulting transport process within the framework of linear irreversible thermodynamics, we fix the reference temperature T ≡ T1 and chemical potential μ ≡ μ1, and define the affinities

Equation (1)

By Jρα and Jqα we denote the charge and the heat current flowing out of the reservoir α, respectively. Within the linear response regime, which is valid as long as the temperature and chemical potential differences ΔTα and Δμα are small compared to the respective reference values, the currents and affinities are connected via the phenomenological equations [18]

Equation (2)

Here, we introduced the current vector

Equation (3)

with the respective subunits

Equation (4)

Analogously, we divide the matrix of kinetic coefficients

Equation (5)

into the 2 × 2 blocks $\mathbb {L}_{\alpha \beta } \in \mathbb {R}^{2 \times 2}$ ($\alpha ,\beta = 2,\ldots , n$ ), which can be calculated explicitly. By making use of the multi-terminal Landauer formula [11, 12], we get the expression

Equation (6)

where h denotes Planck's constant, e the electronic unit charge,

Equation (7)

the negative derivative of the Fermi function and kB Boltzmann's constant.

Figure 1.

Figure 1. Sketch of the multi-terminal model for thermoelectric transport.

Standard image High-resolution image

The expression (6) shows that the transport properties of the model are completely determined by the transition probabilities Tαβ(E,B), which obey two important relations. Firstly, current conservation requires the sum rule

Equation (8)

i.e. the transition matrix

Equation (9)

is doubly stochastic for any $E \in \mathbb {R}$ and $\mathbf {B}\in \mathbb {R}^3$ . Secondly, due to time-reversal symmetry, the Tαβ(E,B) have to possess the symmetry

Equation (10)

Notably, for a fixed magnetic field B, the transition matrix $\mathbb {T}(E,\mathbf {B})$ does not necessarily have to be symmetric. This observation will be crucial for the subsequent considerations.

For later purpose, we note that, by combining (5) and (6), $\mathbb {L}(\mathbf {B})$ can be expressed as an integral over tensor products given by

Equation (11)

Here, $\mathbbm {1}$ denotes the identity matrix and $\bar {\mathbb {T}}(E,\mathbf {B})$ arises from $\mathbb {T}(E,\mathbf {B})$ by deleting the first row and column. Consequently, the matrix $\bar {\mathbb {T}}(E,\mathbf {B})$ must be doubly substochastic, which means that all entries of $\bar {\mathbb {T}}(E,\mathbf {B})$ are non-negative and any row and column sums up to a value not greater than 1.

3. Bounds on the kinetic coefficients

3.1. Phenomenological constraints

The phenomenological framework of linear irreversible thermodynamics provides two fundamental constraints on the matrix of kinetic coefficients $\mathbb {L}(\mathbf {B})$ . Firstly, since the entropy production accompanying the transport process described by (2) reads [18]

Equation (12)

the second law requires $\mathbb {L}(\mathbf {B})$ to be positive semi-definite. Secondly, Onsager's reciprocal relations impose the symmetry

Equation (13)

Apart from these constraints, no further general relations restricting the elements of $\mathbb {L}(\mathbf {B})$ at fixed magnetic field B are known. We will now demonstrate that such a lack of constraints leads to profound consequences for the thermodynamical properties of this model. To this end, we split the current vector J into an irreversible and a reversible part given by

Equation (14)

respectively. The reversible part vanishes for B = 0 by virtue of the reciprocal relations (13). However, in situations with B ≠ 0 it can become arbitrarily large without contributing to the entropy production (12). In principle, it would be even possible to have $\dot {S} =0$ and Jrev ≠ 0 simultaneously, i.e. completely reversible transport, suggesting inter alia the opportunity for a thermoelectric heat engine operating at Carnot efficiency with finite power output [9]. This observation raises the question, whether there might be stronger relations between the kinetic coefficients going beyond the well known reciprocal relations (13). In the next section, starting from the microscopic representation (6), we derive bounds on the kinetic coefficients, which prevent this option of Carnot efficiency with finite power.

3.2. Bounds following from current conservation

These bounds can be derived by first quantifying the asymmetry of the Onsager matrix $\mathbb {L}(\mathbf {B})$ . For an arbitrary positive semi-definite matrix $\mathbb {A} \in \mathbb {R}^{m \times m}$ we define an asymmetry index by

Equation (15)

Some of the basic properties of this asymmetry index are outlined in appendix A. We note that a quite similar quantity was introduced by Crouzeix and Gutan [19] in another context.

We will now proceed in two steps. Firstly, we show that the asymmetry index of the matrix of kinetic coefficients $\mathbb {L}(\mathbf {B})$ and all its principal submatrices is bounded from above for any finite number of terminals n. Secondly, we will derive therefrom a set of new bounds on the elements of $\mathbb {L}(\mathbf {B})$ , which go beyond the second law. We note that from now on we notationally suppress the dependence of any quantity on the magnetic field in order to keep the notation slim.

For the first step, we define the quadratic form

Equation (16)

for any $\mathbf {z} \in \mathbb {C}^{2m}$ and any $s\in \mathbb {R}$ . Here, $A\subset \left \{2,\ldots , n\right \}$ denotes a set of m ⩽ n − 1 integers. The matrix $\mathbb {L}_A$ arises from $\mathbb {L}$ by taking all blocks $\mathbb {L}_{\alpha \beta }$ with column and row index in A, i.e. $\mathbb {L}_A$ is a principal submatrix of $\mathbb {L}$ , which preserves the 2 × 2 block structure shown in (5). Comparing (16) with the definition (15) reveals that the minimum s for which Q(z,s) is positive semi-definite equals the asymmetry index of $\mathbb {L}_A$ . Next, by recalling (11) we rewrite the matrix $\mathbb {L}_A$ in the rather compact form

Equation (17)

where $\bar {\mathbb {T}}_A (E) \in \mathbb {R}^{m\times m}$ is obtained from $\bar {\mathbb {T}}(E)$ by taking the rows and columns indexed by the set A. Decomposing the vector z as

Equation (18)

and inserting (17) and (18) into (16) yields

Equation (19)

Here we introduced the vector

Equation (20)

and the Hermitian matrix

Equation (21)

which is positive semi-definite for any

Equation (22)

However, since $\bar {\mathbb {T}}(E)$ is doubly stochastic for any E, the matrix $\bar {\mathbb {T}}_A(E)$ must have the same property and it follows from corollary 2 proven in appendix B:

Equation (23)

Hence, independently of E, $\mathbb {K}_A(E,s)$ is positive semi-definite for any

Equation (24)

Finally, we can infer from (19) that Q(z,s) is positive semi-definite for any s, which obeys (24). Consequently, with (16), we have the desired bound on the asymmetry index of $\mathbb {L}_A$ as

Equation (25)

This bound, which ultimately follows from current conservation, constitutes our first main result.

We will now demonstrate that (25) puts indeed strong bounds on the kinetic coefficients. To this end, we extract a 2 × 2 principal submatrix from $\mathbb {L}$ by a two-step procedure, which is schematically summarized in figure 2. In the first step, we consider the 4 × 4 principal submatrix of $\mathbb {L}$ given by

Equation (26)

which arises from $\mathbb {L}$ by taking only the blocks with row and column index equal to α or β. From (25) we immediately get with m = 2

Equation (27)

Next, from (26), we take a 2 × 2 principal submatrix

Equation (28)

where $\left (\mathbb {L}_{\alpha \beta }\right )_{ij}$ with i,j = 1,2 denotes the (i,j)-entry of the block matrix $\mathbb {L}_{\alpha \beta }$ . By virtue of proposition 3 proven in appendix B, the inequality (27) implies

Equation (29)

which is equivalent to requiring the Hermitian matrix

Equation (30)

to be positive semi-definite. Since the diagonal entries of $\tilde {\mathbb {K}}_{\{\alpha ,\beta \}}$ are obviously non-negative, this condition reduces to $ {{\mathrm { Det}}}\, \tilde {\mathbb {K}}_{\{ \alpha ,\beta \} }= K_{11}K_{22} - \left \vert K_{12} \right \vert ^2 \geqslant 0.$ Finally, expressing the Kij again in terms of the Lij yields the new constraint

Equation (31)

This bound that holds for the elements of any 2 × 2 principal submatrix of the full matrix of kinetic coefficients $\mathbb {L}$ , irrespective of the number n of terminals is our second main result. Compared to relation (31), the second law only requires $\tilde {\mathbb {L}}_{\{\alpha ,\beta \}}$ to be positive semi-definite, which is equivalent to L11,L22 ⩾ 0 and the weaker constraint

Equation (32)

Note that the reciprocal relations (13) do not lead to any further relations between the kinetic coefficients contained in $\tilde {\mathbb {L}}_{\{\alpha ,\beta \}}$ for a fixed magnetic field B.

Figure 2.

Figure 2. Schematic illustration of the reduction from $\mathbb{L}$ to $\tilde{\mathbb {L}}_{\alpha \beta }$ . The big square represents $\mathbb {L}$ for the case n = 6, the smaller ones correspond to the 2 × 2 blocks introduced in (5). By taking the bold framed squares, the 4 × 4 matrix $\mathbb {L}_{\{\alpha \beta \}}$ is obtained for the case α = 1 and β = 3. The filled squares represent the elements of the 2 × 2 matrix $\tilde {\mathbb {L}}_{\{\alpha \beta \}}$ introduced in (28) for (i,j) = (1,1) (blue) and (i,j) = (2,1) (green).

Standard image High-resolution image

At this point, we emphasize that the procedure shown here for 2 × 2 principal submatrices of $\mathbb {L}$ could be easily extended to larger principal submatrices. The result would be a whole hierarchy of constraints involving more and more kinetic coefficients. However, (31) is the strongest bound following from (25), which can be expressed in terms of only four of these coefficients.

4. Bounds on efficiencies

In this section, we explore the consequences of the bound (25) on the performance of various thermoelectric devices.

4.1. Heat engine

A thermoelectric heat engine uses heat from a hot reservoir as input and generates power output by driving a particle current against an external field or a gradient of chemical potential [5]. Such an engine can be realized within the multi-terminal model by considering the terminals $3,\ldots , n$ as pure probe terminals, which mimic inelastic scattering events while not contributing to the actual transport process. This constraint reads

Equation (33)

By assuming the matrix

Equation (34)

to be invertible, we can solve the self-consistency relations (33) for $\boldsymbol {\mathcal {F}}_3,\ldots ,\boldsymbol {\mathcal {F}}_n$ obtaining

Equation (35)

After inserting this solution into (2) and identifying the heat current Jq ≡ Jq2 leaving the hot reservoir and the particle current Jρ ≡ Jρ2, we end up with the reduced system

Equation (36)

of phenomenological equations. Here, the effective matrix of kinetic coefficients is given by

Equation (37)

and the affinities $\mathcal {F}_{\rho }\equiv \mathcal {F}_2^{\rho }=\Delta \mu _2/T<0$ and $\mathcal {F}_q\equiv \mathcal {F}_2^q=\Delta T _2/T^2>0$ have to be chosen such that Jρ,Jq ⩾ 0 for the model to work as a proper heat engine. $\mathbb {L}^{{{\mathrm { HE}}}}$ is not a principal submatrix of the full Onsager matrix $\mathbb {L}$ and therefore the bound (25) does not apply directly. However, $\mathbb {L}^{{{\mathrm { HE}}}}$ can be written as the Schur complement $\mathbb {L}/\mathbb {L}_{\{3,\ldots ,n\}}$ (see appendix C for the definition), the asymmetry index of which is dominated by the asymmetry index of $\mathbb {L}$ as proven in proposition 4 of appendix C. Consequently, we have

Equation (38)

or, equivalently,

Equation (39)

This constraint shows that whenever Lρq ≠ Lqρ, the entropy production (12) must be strictly larger than zero, thus ruling out the option of dissipationless transport generated solely by reversible currents for any model with a finite number n of terminals. For any n > 3 this constraint is weaker than (31). The reason is that the Onsager coefficients in (39) are not elements of the full matrix (5) but rather involve the inversion of $\mathbb {L}_{\{3,\ldots ,n\}}$ defined in (34). Still, this constraint is stronger than the bare second law, which requires only

Equation (40)

irrespective of whether or not $\mathbb {L}^{{{\mathrm { HE}}}}$ is symmetric.

The constraint (39) implies a constraint on the efficiency η of such a particle-exchange heat engine [5], which is defined as

Equation (41)

Like for any heat engine, this efficiency is subject to the Carnot-bound ηC ≡ 1 − T/T2, which, in the linear response regime, is given by $\eta _{\mathrm { C}}\approx \Delta T_2/T = T\mathcal {F}_q$ . Following Benenti et al [9], we now introduce the dimensionless parameters

Equation (42)

which allow us to write the maximum efficiency of the engine ηmax (under the condition Jq > 0) in the instructive form [9]

Equation (43)

Restating the new bound (39) in terms of x and y yields

Equation (44)

with

Equation (45)

Consequently, maximizing (43) with respect to y yields the optimal y*(x) = hn(x) and the maximum efficiency

Equation (46)

This bound is plotted in figure 3 as a function of x for an increasing number n of terminals. For n = 3, we recover the result obtained in our preceding work on the three terminal model [16]. In the limit n → , ηmax(x) converges to the bound derived by Benenti et al [9] within a general analysis relying only on the second law. However, for any finite n, ηmax(x) is constrained to be strictly smaller than ηC, as soon as x deviates from 1. Thus, from the perspective of maximum efficiency, breaking the time reversal symmetry is not beneficial.

Figure 3.

Figure 3. Bounds on the efficiency of the multi-terminal model as a thermoelectric heat engine as functions of the asymmetry parameter x and in units of ηC. The upper panel shows ηmax(x) (see equation (46)), the lower one η*(x) (see equation (49)). In both panels, the blue lines, from bottom to top, belong to models with $n=3,\ldots ,12$ terminals and the solid, black line corresponds to the bound following from the bare second law as obtained by Benenti et al [9]. The dashed line in the lower panel marks the Curzon–Ahlborn limit ηCA = ηC/2.

Standard image High-resolution image

As a second important benchmark for the performance of a heat engine, we consider its efficiency at maximum power η* [2022] obtained by maximizing the power output

Equation (47)

with respect to $\mathcal {F}_{\rho }$ for fixed $\mathcal {F}_q$ . In terms of the dimensionless parameters (42), it reads [9]

Equation (48)

and attains its maximum

Equation (49)

at y*(x) = hn(x). In the lower panel of figure 3, η*(x) is plotted as a function of the asymmetry parameter x. For x = 1, this bound acquires the Curzon–Ahlborn value ηCA ≡ ηC/2. For x ≠ 1, however it can become significantly higher even for a small number n of terminals. Specifically, we observe that η*(x) exceeds ηCA for any n ⩾ 3 in a certain range of x values. For n ⩾ 4, this range includes all x > 1. Furthermore, η*(x) attains its global maximum

Equation (50)

at the finite value $ x = 1/\sin ^2 \left (\frac {\pi }{n}\right )$ . Remarkably, both ηmax(x) and η*(x) approach the same asymptotic value $\eta ^{\infty }\equiv \eta _{\mathrm { C}} \cos ^{2} \left (\frac {\pi }{n}\right )$ for x → ± .

4.2. Refrigerator

In the preceding section, we discussed the performance of the multi-terminal model if it is operated as a heat engine. Quite naturally, we can change the mode of operation of this engine such that it functions as a refrigerator. The resulting device consumes electrical power from which it generates a heat current from the cold to the hot reservoir. Thus, compared to the heat engine, input and output are interchanged and the affinities $\mathcal {F}_{\rho }<0$ and $\mathcal {F}_q>0$ have to be chosen such that both currents Jρ and Jq are negative.

Analogously to the case of the heat engine, we will now show that the bound (39) on the kinetic coefficients constrains the performance of the thermoelectric refrigerator described above. To this end, we will use the coefficient of performance [18]

Equation (51)

as a benchmark parameter. Its upper bound following from the second law is given by $\varepsilon _{\mathrm { C}}\equiv T/\Delta T_2 = 1/(T\mathcal {F}_q)$ , which is the efficiency of the ideal refrigerator. In this sense, εC is the analogue to the Carnot efficiency.

Taking the maximum of ε over $\mathcal {F}_{\rho }$ (under the condition Jρ < 0) while keeping $\mathcal {F}_q$ fixed, yields the maximum coefficient of performance [9]

Equation (52)

Here, we used again the dimensionless parameters defined in (42). Since y is subject to the constraint (44), εmax(x,y) attains its maximum

Equation (53)

with respect to y at y*(x) = hn(x), where hn(x) was introduced in (45). Figure 4 shows ηmax(x) for models with an increasing number of probe terminals n. For any finite n, εC can only be reached for the symmetric value x = 1. The black line follows solely from the second law (40) and would in principle allow to reach εC with finite current for x between −1 and 1. However, like for the heat engine, our analysis reveals that such a high performance refrigerator would need to be equipped with an infinite number of terminals.

Figure 4.

Figure 4. Maximum coefficient of performance εmax(x) (see (53)) of a thermoelectric refrigerator as a function of the asymmetry parameter x. The blue lines from bottom to top represent models with $n=3,\ldots ,12$ terminals. The black curve shows the bound required by the bare second law, which is asymptotically reached in the limit n → .

Standard image High-resolution image

4.3. Isothermal engine

By an isothermal, thermoelectric engine, we understand in this context a device in which one particle current driven by a (negative) gradient in chemical potential drives another one uphill a chemical potential gradient at constant temperature T. In order to implement such a machine within the multi-terminal framework, we put $\mathcal {F}_2^q =\cdots = \mathcal {F}_n^q=0$ . The remaining affinities $\mathcal {F}_2^{\rho },\ldots ,\mathcal {F}_n^{\rho }$ are connected to the particle currents via a reduced set of phenomenological equations given by

Equation (54)

where $(\mathbb {L}_{\alpha \beta })_{11}$ denotes the (11)-entry of the block matrix $\mathbb {L}_{\alpha \beta }$ defined in (6). We note that the heat currents $J^q_2,\ldots , J^q_n$ do not necessarily have to vanish. However, since they do not contribute to the entropy production (12), they are irrelevant in the present analysis. Similar to the treatment of the heat engine, we put Jρ4 = ··· = Jρn = 0, thus considering the terminals $4,\ldots ,n$ as pure probe terminals simulating inelastic scattering events. Consequently, (54) can be reduced further to the generic form

Equation (55)

Here, we have introduced the matrix

Equation (56)

again using the Schur complement defined in appendix C. The affinities $\mathcal {F}^{\rho }_2, \mathcal {F}^{\rho }_3>0$ have to be chosen such that Jρ2 is negative and Jρ3 is positive to ensure that the device pumps particles into the reservoir 2 against the gradient in chemical potential Δμ2.

We will now derive a bound on the elements of $\mathbb {L}^{{{\rm IE}}}$ . By employing expression (11), we can write

Equation (57)

with

Equation (58)

and

Equation (59)

Since $\bar {\mathbb {T}}(E)$ is doubly substochastic for any E, the matrix $\langle \bar {\mathbb {T}}\rangle $ is also doubly substochastic. Therefore, by applying corollary 3 of appendix C, we find

Equation (60)

where $\left (\mathbbm {1}-\langle \bar {\mathbb {T}}\rangle \right )_{\{3,\ldots ,n-1\}}$ denotes the principal submatrix of $\mathbbm {1}-\langle \bar {\mathbb {T}}\rangle $ consisting of all but the first two rows and columns. Expressing (60) in terms of the elements of $\mathbb {L}^{{{\mathrm { IE}}}}$ gives the bound

Equation (61)

We emphasize that, in contrast to the bound (39) we derived for the heat engine, the bound (61) is independent of the number of probe terminals involved in the device.

In the next step we explore the implications of (61) for the performance of the isothermal engine. To this end, we identify the output power of the device as

Equation (62)

and correspondingly the input power as

Equation (63)

Consequently, the efficiency of the isothermal engine reads

Equation (64)

We note that, in the situation considered here, the entropy production (12) reduces to

Equation (65)

and thus the second law $\dot {S}\geqslant 0$ requires η ⩽ 1 for isothermal engines [22].

Optimizing η and Pout (under the condition Jρ3 > 0) with respect to $\mathcal {F}_2^{\rho }$ while keeping $\mathcal {F}_3^{\rho }$ fixed yields the maximum efficiency

Equation (66)

and the efficiency at maximum power

Equation (67)

where we have introduced the dimensionless parameters

Equation (68)

analogous to (42). Using these definitions, the bound (61) translates to

Equation (69)

with

Equation (70)

and ηmax(x,y) as well as η*(x,y) attain their respective maxima with respect to y at y* = h(x). The resulting bounds

Equation (71)

and

Equation (72)

are plotted in figure 5. We observe that the ηmax(x) reaches 1 only for x = 1 and decreases rapidly as the asymmetry parameter x deviates from 1, while η*(x) exceeds the Curzon–Ahlborn value 1/2 for x between 1 and 2 with a global maximum η** = 4/7 at x = 4/3. In contrast to the non-isothermal engines analyzed in the preceding sections, all these bounds do not depend on the number of probe terminals.

Figure 5.

Figure 5. Bounds on benchmark parameters for the performance of the isothermal, thermoelectric engine as functions of the asymmetry parameter x. The right panel shows the maximum efficiency ηmax(x) (see equation (71)), the left one efficiency at maximum power η*(x) (see equation (72)). The black lines follow from the bare second law, the blue lines from the stronger constraint (61). Both, ηmax(x) and η*(x) asymptotically reach the value 1/4. The dashed line in the right plot marks the value 1/2 of η*(x) at the symmetric value x = 1.

Standard image High-resolution image

4.4. Absorption refrigerator

By an absorption refrigerator, one commonly understands a device that generates a heat current cooling a hot reservoir, while itself being supplied by a heat source [23, 24]. The multi-terminal model allows to implement such a device by following a very similar strategy like the one used for the isothermal engine, i.e. we put $\mathcal {F}_2^{\rho }=\cdots =\mathcal {F}_n^{\rho }=0$ and end up with the reduced system of phenomenological equations

Equation (73)

connecting the heat currents with the temperature gradients. Assuming the terminals $4,\ldots ,n$ to be pure probe terminals then leads to

Equation (74)

where $\mathcal {F}_2^q<0$ , $\mathcal {F}_3^q>0$ have to be adjusted such that J2q > 0 and J3q > 0. The matrix $\mathbb {L}^{\mathrm { AR}}$ is given by

Equation (75)

and by following the reasoning of the last section, we can derive the bound

Equation (76)

The efficiency of the absorption refrigerator can be consistently defined as

Equation (77)

Just like for the isothermal engine, after maximizing this efficiency over $\mathcal {F}_2^q$ (under the condition Jq2 > 0), we can derive an upper bound

Equation (78)

from (76). Again, this bound is independent of the number of probe terminals. Figure 6 shows it as a function of the asymmetry parameter x ≡ L23'/L32'.

Figure 6.

Figure 6. Maximum efficiency ηmax(x) (see equation (78)) of the thermoelectric absorption refrigerator as a function of x. The blue line follows by virtue of the constraint (76), the black line by invoking only the second law.

Standard image High-resolution image

For completeness, we emphasize that the efficiency (77) used here differs from the coefficient of performance

Equation (79)

used as a benchmark parameter in [23, 24]. Since ε is unbounded in the linear response regime, maximization with respect to $\mathcal {F}^q_2$ or $\mathcal {F}^q_3$ would be meaningless.

5. Conclusion and outlook

We have studied the influence of broken time-reversal symmetry on thermoelectric transport within the quite general framework of an n-terminal model. Our analytical calculations prove that the asymmetry index of any principal submatrix of the full Onsager matrix defined in (5) is bounded according to (25). This somewhat abstract bound can be translated into the set (31) of new constraints on the kinetic coefficients. Any of these constraints is obviously stronger than the bare second law and cannot be deduced from Onsagers time-reversal argument. Furthermore, we note that it is straightforward to repeat the procedure carried out in section 3.2 for larger principal submatrices, thus obtaining relations analogous to (31), which involve successively higher order products of kinetic coefficients. Investigating this hierarchy of constraints will be left to future work.

After the general analysis of the transport processes in the full multi-terminal set-up, we investigated the consequences of our new bounds on the performance of the model if operated as a thermoelectric heat engine. We found that both the maximum efficiency as well as the efficiency at maximum power are subject to bounds, which strongly depend on the number n of terminals. In the minimal case n = 3, we recover the strong bounds already discussed in [16]. Although our new bounds become successively weaker as n is increased, they prove that reversible transport is impossible in any situation with a finite number of terminals. Only in the limit n →  we are back at the situation discussed by Benenti et al [9], in which the second law effectively is the only constraint. We recall that for n = 3 our bounds can indeed be saturated as Balachandran et al [17] have shown within a specific model. Whether or not it is possible to saturate the bounds for higher n remains open at this stage and constitutes an important question for future investigations.

Like in the case of the heat engine, the bound on the maximum coefficient of performance we derived for the thermoelectric refrigerator becomes weaker as n increases. Interestingly, the situation is quite different for the isothermal engine and the absorption refrigerator considered in sections 4.3 and 4.4. The bounds on the respective benchmark parameters equal those of the three-terminal case irrespective of the actual number of terminals involved. If one assumed that any kind of inelastic scattering could be simulated by a sufficiently large number of probe terminals, one had to conclude that the results shown in figures 5 and 6 were a universal bound on the efficiency of any such device. At least, the results of sections 4.3 and 4.4 suggest a fundamental difference between transport processes under broken time-reversal symmetry that are driven by only one type of affinities, i.e. either chemical potential differences or temperature differences, and those, which are induced by both types of thermodynamic forces.

We emphasize that technically all our results ultimately rely on the sum rules (8) for the elements of the transmission matrix. These constraints reflect the fundamental law of current conservation, which should be seen as the basic physical principle behind our bounds. Therefore the validity of these bounds is not limited to the quantum realm. It rather extends to any model, quantum or classical, for which the kinetic coefficients can be expressed in the generic form (6). Some specific examples for quantum mechanical models which fulfill this requirement are discussed in [17, 25]. A classical model belonging to this class was recently introduced by Horvat et al [26].

In summary, we have achieved a fairly complete picture of thermoelectric transport under broken time reversal symmetry in systems with non-interacting particles for which the Onsager coefficients can be expressed in the Landauer–Büttiker form (6). However, fully interacting systems, which require to go beyond the single particle picture, are not covered by our analysis yet. Exploring these systems remains one of the major challenges for future research.

Acknowledgments

We gratefully acknowledge stimulating discussions with K Saito and support of the ESF through the EPSD network.

Appendix A.: Quantifying the asymmetry of positive semi-definite matrices

We first recall the definition (15)

Equation (A.1)

of the asymmetry index of an arbitrary positive semi-definite matrix $\mathbb {A}\in \mathbb {R}^{m\times m}$ . Below, we list some of the basic properties of this quantity, which can be inferred directly from its definition.

(Basic properties of the asymmetry index).

Proposition 1 For any positive semi-definite $\mathbb {A}\in \mathbb {R}^{m\times m }$ and λ > 0, we have

Equation (A.2)

and

Equation (A.3)

with equality if and only if $\mathbb {A}$ is symmetric. If $\mathbb {A}$ is invertible, it holds additionally

Equation (A.4)

Furthermore, we can easily prove the following two propositions, which are crucial for the derivation of our main results.

(Convexity of the asymmetry index).

Proposition 2 Let $\mathbb {A},\mathbb {B} \in \mathbb {R}^{m \times m}$ be positive semi-definite, then

Equation (A.5)

Proof. By definition A.1 the matrices

Equation (A.6)

with $s \equiv \max \left \{ \mathcal {S}\left ( \mathbb {A} \right ),\mathcal {S}\left ( \mathbb {B} \right )\right \}$ both are positive semi-definite. It follows that

Equation (A.7)

is also positive semi-definite and hence $\mathcal {S}\left ( \mathbb {A}+\mathbb {B} \right ) \leqslant s$ .   □

(Dominance of principal submatrices).

Proposition 3 Let $\mathbb {A}\in \mathbb {R}^{m \times m}$ be positive semi-definite and $\bar {\mathbb {A}}\in \mathbb {R}^{p \times p}$ (p < m) a principal submatrix of $\mathbb {A}$ , then

Equation (A.8)

Proof. By definition A.1

Equation (A.9)

is positive semi-definite. Consequently the matrix

Equation (A.10)

which constitutes a principal submatrix of $\mathbb {K}$ , is also positive semi-definite and therefore $\mathcal {S}( \bar {\mathbb {A}})\leqslant \mathcal {S}\left ( \mathbb {A} \right )$ .   □

Appendix B.: Bound on the asymmetry index for special classes of matrices

Theorem 1. Let $\mathbb {P} \in \left \{0,1 \right \}^{m \times m}$ be a permutation matrix and $\mathbbm {1}$ the identity matrix, then the matrix $\mathbbm {1}- \mathbb {P}$ is positive semi-definite on $\mathbb {R}^m$ and its asymmetry index fulfills

Equation (B.1)

Proof. We first show that $\mathbbm {1}-\mathbb {P}$ is positive semi-definite. To this end, we note that the matrix elements of $\mathbb {P}$ are given by $\left ( \mathbb {P} \right )_{ij} =\delta _{i\pi (j)}$ , where πSm is the unique permutation associated with $\mathbb {P}$ and Sm the symmetric group on the set $\left \{1,\ldots ,m\right \}$ . Now, with $\mathbf {x} \equiv \left (x_1, \ldots ,x_m \right )^{\mathrm { t}} \in \mathbb {R}^m$ we have

Equation (B.2)

Equation (B.3)

We now turn to the second part of theorem 1. For any $\mathbf {z}\equiv (z_1,\ldots ,z_m)\in \mathbb {C}^m$ and s ⩾ 0, we define the quadratic form

Equation (B.4)

Equation (B.5)

By definition A.1 the minimum s, for which Q(z,s) is positive semi-definite, equals the asymmetry index of $\mathbbm {1}-\mathbb {P}$ . This observation enables us to derive an upper bound for $\mathcal {S}\left ( \mathbbm {1}-\mathbb {P} \right )$ . To this end, we make use of the cycle decomposition

Equation (B.6)

of π, where $i_1, \ldots ,i_k \in \left \{1, \ldots , m \right \}$ , πl(i) is defined recursively by

Equation (B.7)

k denotes the number of independent cycles of and nr the length the rth cycle. By virtue of this decomposition, (B.4) can be rewritten as

Equation (B.8)

Equation (B.9)

Equation (B.10)

where, for convenience, we introduced the notation z[x] ≡ zx. Next, we define the vectors $\tilde {\mathbf {z}}_r\in \mathbb {C}^{n_r}$ with elements $\left (\tilde {\mathbf {z}}_r\right )_j\equiv z\left [\pi ^{j-1}(i_r)\right ]$ and the Hermitian matrices $\mathbb {H}_{n_r}(s)\in \mathbb {C}^{n_r \times n_r}$ with matrix elements

Equation (B.11)

where periodic boundary conditions nr + 1 = 1 for the indices $i,j=1,\ldots ,n_r$ are understood. These definitions allow us to cast (B.10) in the rather compact form

Equation (B.12)

Obviously, any value of s for which all the $\mathbb {H}_{n_r}(s)$ are positive semi-definite serves as a lower bound for $\mathcal {S}\left (\mathbbm {1}-\mathbb {P} \right )$ . Moreover, we can calculate the eigenvalues of $\mathbb {H}_{n_r}(s)$ explicitly. Inserting the ansatz $\mathbf {v}\equiv (v_1,\ldots , v_{n_r})^{\mathrm { t}}\in \mathbb {C}^{n_r}$ into the eigenvalue equation

Equation (B.13)

yields

Equation (B.14)

where again periodic boundary conditions vnr + 1 = v1 are understood. This recurrence equation can be solved by standard techniques. We put $v_j \equiv \exp \left (2\pi \mathrm {i}\kappa j / n_r\right )$ with $(\kappa = 1, \ldots , n_r)$ and obtain the eigenvalues

Equation (B.15)

For any fixed s ⩾ 0, the function

Equation (B.16)

is non-negative for x∈[x*,2π] and strictly negative for x∈(0,x*) with

Equation (B.17)

Therefore, all the eigenvalues λκ of $\mathbb {H}_{n_r}(s)$ are non-negative, if and only if

Equation (B.18)

Solving (B.18) for s gives the equivalent condition

Equation (B.19)

Since nr ⩽ m and therefore 2π/nr ⩾ 2π/m, we can conclude that any of the $\mathbb {H}_{n_r}(s)$ is positive semi-definite for any

Equation (B.20)

thus establishing the desired result (B.1).   □

Corollary 1. Let $\mathbb {T} \in \mathbb {R}^{m \times m}$ be doubly stochastic, then the matrix $\mathbbm {1}- \mathbb {T}$ is positive semi-definite and its asymmetry index fulfills

Equation (B.21)

Proof. The Birkhoff-theorem (see [27, p 549]) states that for any doubly stochastic matrix $\mathbb {T}\in \mathbb {R}^{m\times m}$ there is a finite number of permutation matrices $\mathbb {P}_1, \ldots \mathbb {P}_N \in \{0,1 \}^{m \times m}$ and positive scalars $\lambda _1, \ldots , \lambda _N \in \mathbb {R}$ such that

Equation (B.22)

Hence, we have

Equation (B.23)

and consequently $\mathbbm {1} - \mathbb {T}$ must be positive semi-definite by virtue of theorem 1. Furthermore, using proposition 2 and again theorem 1 gives the bound (B.21).   □

Theorem 2. Let $\bar {\mathbb {P}}\in \left \{0,1\right \}^{m\times m}$ be a partial permutation matrix, i.e. any row and column of $\bar {\mathbb {P}}$ contains at most one non-zero entry and all of these non-zero entries are 1. Then, the matrix $\mathbbm {1} - \bar {\mathbb {P}}$ is positive semi-definite and its asymmetry index fulfills

Equation (B.24)

Proof. Let q be the number of non-vanishing entries of $\bar {\mathbb {P }}$ . If q = 0, $\bar {\mathbb {P}}$ equals the zero matrix and there is nothing to prove. If q = m, $\bar {\mathbb {P}}$ itself must be a permutation matrix and lemma 1 provides that $\mathbbm {1}-\bar {\mathbb {P}}$ is positive semi-definite as well as the bound

Equation (B.25)

which is even stronger than (B.24). If 0 < q < m, there are two index sets $A\subset \{1,\ldots , m\}$ and $B\subset \{1,\ldots , m\}$ of equal cardinality m − q, such that the rows of $\bar {\mathbb {P}}$ indexed by A and the columns of $\bar {\mathbb {P}}$ indexed by B contain only zero entries. Clearly, in this case, $\bar {\mathbb {P}}$ is not a permutation matrix. Nevertheless, we can define a bijective map

Equation (B.26)

in such a way that $\bar {\mathbb {P}}$ can be regarded as a representation of $\bar { \pi }$ . To this end, we denote by $\{\mathbf {e}_1,\ldots , \mathbf {e}_{m}\}$ the canonical basis of $\mathbb {R}^{m}$ and define $\bar {\pi }: \; i \mapsto \bar {\pi }(i)$ such that

Equation (B.27)

This definition naturally leads to the cycle decomposition

Equation (B.28)

Here, we introduced two types of cycles. The ones in round brackets, which we will term complete, are just ordinary permutation cycles, which close by virtue of the condition $\bar {\pi }^{n_r} (i_r) =i_r$ and therefore must be contained completely in the set

Equation (B.29)

The cycles in rectangular brackets, which will be termed incomplete, do not close, but begin with a certain $j_{\bar {r}}$ taken from the set

Equation (B.30)

and terminate after $\bar {n}_{\bar {r}} -1$ iterations with $\bar {\pi }^{\bar {n}_{\bar {r}}-1}(j_{\bar {r}})$ , which is contained in

Equation (B.31)

Figure B.1 shows a schematic visualization of the two different types of cycles. We note that, since the map $\bar {\pi }$ , is bijective the cycle decomposition (B.28) is unique up to the choice of the ir and any element of

Equation (B.32)

shows up exactly once.

For the next step, we introduce the vectors

Equation (B.33)

as well as the bordered matrix

Equation (B.34)

Obviously, all rows and columns of $\mathbb {B}$ sum up to 1 and all off-diagonal entries are non-negative. Hence, with $B_{ij}\equiv \left (\mathbb {B}\right )_{ij}$ , we have for any $\mathbf {x}\in \mathbb {R}^m$

Equation (B.35)

i.e. the matrix $\mathbbm {1}-\mathbb {B}$ is positive semi-definite. Since $\mathbbm {1}-\bar {\mathbb {P}}$ is a principal submatrix of $\mathbbm {1}-\mathbb {B}$ , (B.35) implies in particular that $\mathbbm {1}-\bar {\mathbb {P}}$ is positive semi-definite, thus establishing the first part of lemma 2.

We will now prove the bound (B.24) on the asymmetry index of $\mathbbm {1}-\bar {\mathbb {P}}$ . To this end, for any $\mathbf {z}\in \mathbb {C}^{m+1}$ we associate the matrix $\mathbb {B}$ with the quadratic form

Equation (B.36)

Equation (B.37)

and notice that the minimum s for which $\bar {Q}(\mathbf {z},s)$ is positive semi-definite equals the asymmetry index of $\mathbbm {1}-\mathbb {B}$ . Furthermore, since $\mathbbm {1}-\bar {\mathbb {P}}$ is a principal submatrix of $\mathbbm {1} -\mathbb {B}$ , proposition 3 implies that this particular value of s is also an upper bound on the asymmetry index of $\mathbbm {1}-\bar {\mathbb {P}}$ . Now, by inserting the decomposition

Equation (B.38)

into (B.37) while keeping in mind the definition (B.27), we obtain

Equation (B.39)

By realizing

Equation (B.40)

and making use of the cycle decomposition (B.28), we can rewrite (B.39) as

Equation (B.41)

thus explicitly separating contributions from complete and incomplete cycles. Finally, since we have

Equation (B.42)

Equation (B.43)

by employing the definitions

Equation (B.44)

Equation (B.45)

(B.41) can be written as

Equation (B.46)

where the matrices $\mathbb {H}_{n}(s)$ are defined in (B.11). Since we have already shown for the proof of lemma 1 that $\mathbb {H}_n(s)$ is positive semi-definite for any $s \geqslant \cot \left ( \pi / n \right )$ , we immediately infer from (B.46) that $\bar {Q}(\mathbf {z},s)$ is positive semi-definite for any

Equation (B.47)

Since $\max \{n_r,n_{\bar {r}}+1\}\leqslant m+1$ , we finally end up with

Equation (B.48)

   □

Figure B.1.

Figure B.1. Schematic illustration of the cycle decomposition (B.28). The green circle represents the set $\{1,\ldots ,m\}\setminus B$ , the blue one the set $\{1,\ldots ,m\}\setminus A$ . The black dots symbolize the elements of the respective sets and the arrows show the action of the map $\bar {\pi }$ . While the dashed arrows form a complete cycle, the solid ones combine to an incomplete cycle.

Standard image High-resolution image

Corollary 2. Let $\bar {\mathbb {T}}\in \mathbb {R}^{m\times m}$ be doubly substochastic, then the matrix $\mathbbm {1} - \bar {\mathbb {T}}$ is positive semi-definite and its asymmetry index fulfills

Equation (B.49)

Proof. It can be shown that any doubly substochastic matrix is the convex combination of a finite number of partial permutation matrices $\bar {\mathbb {P}}_k$ (see [28, p 165]), i.e. we have

Equation (B.50)

with

Equation (B.51)

Consequently, it follows

Equation (B.52)

Using the same argument with lemma 2 instead of lemma 1 in the proof of corollary 1 completes the proof of corollary 2.   □

Appendix C.: Bound on the asymmetry index of the Schur complements

For $\mathbb {A}\in \mathbb {C}^{m\times m}$ partitioned as

Equation (C.1)

with non-singular $\mathbb {A}_{22}\in \mathbb {R}^{p\times p}$ , the Schur complement of $\mathbb {A}_{22}$ in $\mathbb {A}$ is defined by (see [29, p 18])

Equation (C.2)

Regarding the asymmetry index, we have the following proposition.

(Dominance of the Schur complement).

Proposition 4 Let $\mathbb {A}\in \mathbb {R}^{m\times m}$ be a positive semi-definite matrix partitioned as in (C.1), where $\mathbb {A}_{22}\in \mathbb {R}^{p\times p}$ is non-singular, then the matrix $\mathbb {A}/\mathbb {A}_{22}$ is positive semi-definite and its asymmetry index fulfills

Equation (C.3)

Proof. By assumption and by definition (A.1), we have for any $\mathbf {z}\in \mathbb {C}^m$

Equation (C.4)

Putting

Equation (C.5)

yields

Equation (C.6)

and

Equation (C.7)

   □

For the special class of matrices considered in corollary 2, the assertion of proposition 4 can be even strengthened. Before being able to state this stronger result, we need to prove the following lemma.

Lemma 1. Let $\bar {\mathbb {T}}\in \mathbb {R}^{m\times m}$ be a doubly substochastic matrix and $\mathbb {S}\equiv \mathbbm {1}-\bar {\mathbb {T}}$ be partitioned as

Equation (C.8)

where $\mathbb {S}_{22}\in \mathbb {R}^{p\times p}$ is non-singular, then there is a doubly substochastic matrix $\bar {\mathbb {T}}_{m-p}\in \mathbb {R}^{(m-p)\times (m-p)}$ , such that

Equation (C.9)

Proof. We start with the case p = 1. Let $\bar {T}_{ij}$ be the matrix elements of $\bar {\mathbb {T}}$ , then the matrix elements of $\mathbb {S}/\mathbb {S}_{22}$ are given by

Equation (C.10)

with $k,l=1,\ldots ,m-1$ . Obviously, we have

Equation (C.11)

Furthermore, since by assumption

Equation (C.12)

it follows

Equation (C.13)

Analogously, we find

Equation (C.14)

Next, we investigate the sign pattern of the $\left (\mathbb {S}/\mathbb {S}_{22}\right )_{kl}$ . Firstly, for k ≠ l, we have

Equation (C.15)

Secondly, we rewrite the $\left (\mathbb {S}/\mathbb {S}_{22}\right )_{kk}$ as

Equation (C.16)

The numerator appearing on the right-hand side can be written as

Equation (C.17)

which is a principal minor of $\mathbbm {1}-\bar {\mathbb {T}}$ . Since, by corollary 2, $\mathbbm {1}-\bar {\mathbb {T}}$ is positive semi-definite, we end up with

Equation (C.18)

From the sum rules (C.11), (C.13) and (C.14) and the constraints (C.15) and (C.18), we deduce that $\mathbbm {1} -\mathbb {S}/\mathbb {S}_{22}$ is doubly substochastic and thus we have proven lemma 1 for p = 1. We now continue by induction. To this end, we assume that lemma 1 is true for p = q. For p = q + 1 the matrix $\mathbb {S}_{22}\in \mathbb {R}^{(q+1)\times (q+1)}$ can be partitioned as

Equation (C.19)

with $\mathbb {W}_{22}\in \mathbb {R}^{q\times q}$ , $W_{11}\in \mathbb {R}$ and accordingly $\mathbf {W}_{12}, \mathbf {W}_{21}\in \mathbb {R}^q$ . The Crabtree–Haynsworth quotient formula (see [29, p 25]), allows us to rewrite $\mathbb {S}/\mathbb {S}_{22}$ as

Equation (C.20)

A direct calculation shows that $\mathbb {S}_{22}/\mathbb {W}_{22}\in \mathbb {R}$ is the lower right diagonal entry of $\mathbb {S}/\mathbb {W}_{22}$ (see [29, p 25] for details). Furthermore, by the induction hypothesis, there is a doubly substochastic matrix $\bar {\mathbb {T}}_{m-q}\in \mathbb {R}^{(m-q)\times (m-q)}$ , such that

Equation (C.21)

Thus, (C.20) reduces to the case p = 1, for which we have already proven lemma 1.   □

From lemma 1 and corollary 2, we immediately deduce

Corollary 3. Let $\bar {\mathbb {T}}$ , $\mathbb {S}$ and $\mathbb {S}_{22}$ be as in lemma 1, then

Equation (C.22)
Please wait… references are loading.
10.1088/1367-2630/15/10/105003