This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Microscopic derivation of transition-state theory for complex quantum systems

Kouichi Hagino Department of Physics, Kyoto University, Kyoto 606-8502, Japan    George F. Bertsch Department of Physics and Institute for Nuclear Theory, Box 351560, University of Washington, Seattle, Washington 98915, USA
Abstract

The decay of quantum complex systems through a potential barrier is often described with transition-state theory, also known as RRKM theory in chemistry. Here we derive the basic formula for transition-state theory based on a generic Hamiltonian as might be constructed in a configuration-interaction basis. Two reservoirs of random Hamiltonians from Gaussian orthogonal ensembles are coupled to intermediate states representing the transition states at a barrier. Under the condition that the decay of the reservoirs to open channels is large, an analytic formula for reaction rates is derived. The transition states act as independent Breit-Wigner resonances which contribute additively to the total transition probability, as is well known for electronic conductance through resonant tunneling states. It is also found that the transition probability is independent of the decay properties of the states in the second reservoir over a wide range of decay widths.

I Introduction

Transition-state theory is ubiquitous in physics and chemistry to calculate reaction and decay rates for many-particle systems in the presence of a barrier hanggi1990 ; tr96 ; weiss2012 ; BW39 ; MR51 . The assumptions in the theory are clear in classical dynamics but less so in the quantum regime. For fermionic systems of equal-mass particles, the Hamiltonian is often formulated in a configuration-interaction (CI) representation. This motivates considering models that exhibit the barrier dynamics in the CI framework to understand conditions to support transition-state approximations.

II Model Hamiltonian

Following previous recent work, we consider here a Hamiltonian composed of three sets of states. The states in the barrier region are represented in the Hamiltonian H2H_{2}. Their precise structure is not specified, but we have in mind a set of configurations with the ground states and their quasiparticle excitations determined by constrained Hartree-Fock or density-functional theory. The other two sets of states contained in Hamiltonians H1H_{1} and H2H_{2} are statistical reservoirs, with their Hamiltonians constructed from the matrices of the Gaussian orthogonal ensemble GOE weidenmuller2009 . For this Hamiltonian, one may think about e.g., a decay from a highly excited configuration, so that the both the pre-saddle and the post-saddle configurations can be treated statistically bertsch-hagino2023 ; polik1990 ; miller1990 ; hernandez1993 . It is important to note that the GOE Ansatz is the only statistical input, and the ensemble is microcanonical rather than canonical. The full Hamiltonian reads

H=(H1V120V12TH2V32T0V32H3),H=\left(\begin{matrix}H_{1}&V_{12}&0\cr V_{12}^{T}&H_{2}&V_{32}^{T}\cr 0&V_{32}&H_{3}\end{matrix}\right), (1)

where V12V_{12} and V23V_{23} are matrices of the coupling interaction between the reservoir states and the states in the bridge Hamiltonian. This model is generalization of the of the model in Ref. weidenmuller2022 , which assumes that H2H_{2} has a single state at the barrier top. See also Ref. weidenmuller2023 for a similar generalization.

To complete a model for reactions, one also needs the coupling matrix elements between HH and the reaction channels. With those ingredients the SS-matrix for transitions from one channel to another can be computed by standard linear algebra manipulations. If one is only interested in reaction probabilities, the linear algebra can be collapsed to a compact formula 111The formula had been used earlier as well meir1992 in the theory of electrical conductivity. mil93 ; da01 ; da95 ; ha08 ; al21 for the transition probability from channel aa to channel bb given by

Tab=|Sab(E)|2=Tr(ΓaG(E)ΓbG(E)).T_{ab}=|S_{ab}(E)|^{2}={\rm Tr}\left(\Gamma_{a}G(E)\Gamma_{b}G^{\dagger}(E)\right). (2)

Here GG is the Green’s function of the Hamiltonian in presence of entrance or decay channels cc at reaction energy EE, 222 It is implicitly assumed in Eq. (3) that the Γc\Gamma_{c} are independent of energy.

G(E)=(HicΓc/2E)1.G(E)=\left(H-i\sum_{c}\Gamma_{c}/2-E\right)^{-1}. (3)

In general the Γc\Gamma_{c} are rank-one matrices (of the same dimension as HH) but for the present model they have only one entry on the diagonal and a block structure given by

Γin=(Γ~in00000000),Γ1=(Γ~100000000),Γ3=(00000000Γ~3)\Gamma_{\rm in}=\left(\begin{matrix}\tilde{\Gamma}_{\rm in}&0&0\cr 0&0&0\cr 0&0&0\end{matrix}\right),\,\Gamma_{1}=\left(\begin{matrix}\tilde{\Gamma}_{1}&0&0\cr 0&0&0\cr 0&0&0\end{matrix}\right),\,\Gamma_{3}=\left(\begin{matrix}0&0&0\cr 0&0&0\cr 0&0&\tilde{\Gamma}_{3}\end{matrix}\right) (4)

depending on which subblock the channel cc connects to. The full Hamiltonian with its coupling to external channels is depicted in Fig. 1.

Refer to caption
Figure 1: Schematic structure of the model Hamiltonian composed of two reservoirs connected by a bridge Hamiltonian and open to three sets of external channels.

In general, one is interested in the reaction probability Ta3T_{a3} from an entrance channel (a=ina={\rm in}) to all possible decay channels in the second reservoir,

Tin,3=cblock   3|Sin,c|2.T_{{\rm in},3}=\sum_{c\in{\rm block}\,\,\,3}|S_{{\rm in},c}|^{2}. (5)

Due to the block structures of HH and Γc\Gamma_{c} we only need the G13G_{13} block of the Green’s function

G=(G11G12G13G21G22G23G31G32G33)G=\left(\begin{matrix}G_{11}&G_{12}&G_{13}\cr G_{21}&G_{22}&G_{23}\cr G_{31}&G_{32}&G_{33}\end{matrix}\right) (6)

in Eq. (2). As derived in Appendix A, the submatrix G13G_{13} reduces to

G13=G1V12G2V32TG3,G_{13}=G_{1}V_{12}G_{2}V_{32}^{T}G_{3}, (7)

where G1G_{1}, G3G_{3}, and G2G_{2} are given by

G1\displaystyle G_{1} =\displaystyle= (H1iΓ~in/2iΓ~1/2E)1,\displaystyle\left(H_{1}-i\tilde{\Gamma}_{\rm in}/2-i\tilde{\Gamma}_{1}/2-E\right)^{-1}, (8)
G3\displaystyle G_{3} =\displaystyle= (H3iΓ~3/2E)1,\displaystyle\left(H_{3}-i\tilde{\Gamma}_{3}/2-E\right)^{-1}, (9)
G2\displaystyle G_{2} =\displaystyle= (H2V12TG1V12V32TG3V32E)1.\displaystyle\left(H_{2}-V_{12}^{T}G_{1}V_{12}-V_{32}^{T}G_{3}V_{32}-E\right)^{-1}. (10)

Substituting Eq. (7) into Eq. (2), the transmission coefficient is obtained as

Tin,3(E)\displaystyle T_{\rm in,3}(E) =\displaystyle= Tr[Γ~in(G1V12G2V32TG3)Γ~3(G3V32G2V12TG1)],\displaystyle{\rm Tr}[\tilde{\Gamma}_{{\rm in}}(G_{1}V_{12}G_{2}V_{32}^{T}G_{3})\tilde{\Gamma}_{3}(G_{3}^{\dagger}V_{32}G_{2}^{\dagger}V_{12}^{T}G_{1}^{\dagger})],
=\displaystyle= Tr[(V12TG1Γ~inG1V12)G2(V23G3Γ~3G3V32)G2].\displaystyle{\rm Tr}[(V_{12}^{T}G_{1}^{\dagger}\tilde{\Gamma}_{\rm in}G_{1}V_{12})G_{2}(V_{23}G_{3}\tilde{\Gamma}_{3}G_{3}^{\dagger}V_{32})G_{2}^{\dagger}].

We write the elements of the two GOE Hamiltonians as

(Hk)ij=(Hk)ji=vk1+δijrijk,(H_{k})_{ij}=(H_{k})_{ji}=v_{k}\sqrt{1+\delta_{ij}}\,r_{ijk}, (13)

where rijkr_{ijk} is a random number from a Gaussian distribution of unit dispersion, rijk2=1\langle r_{ijk}^{2}\rangle=1. Then the average level density ρk\rho_{k} at E=0E=0 at the centers of the GOE Hamiltonians is given by

ρk=Nk1/2πvk,\rho_{k}=\frac{N_{k}^{1/2}}{\pi v_{k}}, (14)

where NkN_{k} is the dimension of HkH_{k}. We set E=0E=0 for the rest of this paper. Each state in H2H_{2} is assumed to couple to specific states in the GOE reservoirs. We parameterize the couplings as 333We implicitly assume Nk>N2N_{k}>N_{2} for both reservoirs.

(V12)ij=v12N11/2δij,(V32)ij=v32N31/2δij.(V_{12})_{ij}=v_{12}N_{1}^{1/2}\delta_{ij},~{}~{}~{}(V_{32})_{ij}=v_{32}N_{3}^{1/2}\delta_{ij}. (15)

This parameterization is not as restrictive as it may seem. Due to the GOE invariance, the couplings can be to arbitrary orthogonal vectors in the GOE spaces. The specific form of the coupling is such that the average matrix element is independent of the dimension NkN_{k}.

The matrices for the decay widths are assumed to be diagonal with elements

(Γ~1)ij=γ1δi,j,(Γ~3)ij=γ3δi,j,(\tilde{\Gamma}_{1})_{ij}=\gamma_{1}\,\delta_{i,j},~{}~{}~{}(\tilde{\Gamma}_{3})_{ij}=\gamma_{3}\,\delta_{i,j}, (16)

except for the entrance channel a=ina={\rm in}, which couples to a single state i=1i=1 in the first reservoir,

(Γ~in)ij=γinδi,1δj,1.(\tilde{\Gamma}_{\rm in})_{ij}=\gamma_{\rm in}\,\delta_{i,1}\delta_{j,1}. (17)

Without loss of generality, the only requirement on Γ~in\tilde{\Gamma}_{\rm in} is that it has rank one within the space of H1H_{1}, as in the “Büttiker probes” of semiconductor transport theorybuttiker1986 ; venugopal2003 ; datta2005 .

III Transition-state theory

We now examine how the average reaction probability depends on the parameters of the model. Since transition-state theory deals with fluxes into or from a statistical reservoir, it is convenient to define a transmission coefficient 𝒯ck\mathcal{T}_{ck} of a channel cc in the second block into the reservoir kk

𝒯ck=2πρkγk\mathcal{T}_{ck}=2\pi\rho_{k}\gamma_{k} (18)

and its sum over channels,

𝒯k=ck𝒯ck.\mathcal{T}_{k}=\sum_{c\in k}\mathcal{T}_{ck}. (19)

For small values of 𝒯ck\mathcal{T}_{ck} it has the physical significance of the transmission probability from the channel cc into the statistical reservoir. As shown in Ref. weidenmuller2022 , it is straightforward to carry out the statistical averaging for Eq. (2) in the limit that 𝒯k/Nk1\mathcal{T}_{k}/N_{k}\gg 1 for both reservoirs. We first examine the Green’s function G2G_{2} and the coupling terms Vk2TGkVk2V_{k2}^{T}G_{k}V_{k2} in it. The averages of Gk(k=1,3)G_{k}~{}(k=1,3) at E=0E=0 444 We note that there are mild restrictions on the ranges of the parameters in Eq. (20). In practice, the widths associated with the individual channels should be large compared to the level spacing in the GOE but small with respect to the boundaries of its eigenspectrum. including the decay-width matrices are given by weidenmuller2022 ; Lo2000 ; Fe2020 ; hagino-bertsch2021

Gkij=iπρkNkδij.\langle G_{k}\rangle_{ij}=i\frac{\pi\rho_{k}}{N_{k}}\delta_{ij}. (20)

The standard deviation of the fluctuations is

SD(Gk)ij=πρkNk(1+i)(2(1+δi,j)Nk𝒯k)1/2.SD(G_{k})_{ij}=\frac{\pi\rho_{k}}{N_{k}}(1+i)\left(\frac{2(1+\delta_{i,j})N_{k}}{\mathcal{T}_{k}}\right)^{1/2}. (21)

The fluctuations go to zero in the limit 𝒯k/Nk1\mathcal{T}_{k}/N_{k}\gg 1 so these terms in G2G_{2} can be replaced by iπvk22ρki\pi v_{k2}^{2}\rho_{k} times the unit matrix. Thus the correlations between G2G_{2} and the other terms in Eq. (LABEL:eq:datta2) vanish, allowing it to be evaluated as

G¯2=(H2V12TG1V12V32TG3V32E)1.\bar{G}_{2}=(H_{2}-V_{12}^{T}\langle G_{1}\rangle V_{12}-V_{32}^{T}\langle G_{3}\rangle V_{32}-E)^{-1}. (22)

The two terms in parentheses in Eq. (LABEL:eq:datta2) are independent of each other so can also be replaced by their ensemble averages.

(V12TG1Γ~inG1V12)ij=γinN1γ12πv122ρ1δij\langle(V_{12}^{T}G_{1}\tilde{\Gamma}_{\rm in}G_{1}^{\dagger}V_{12})_{ij}\rangle=\frac{\gamma_{\rm in}}{N_{1}\gamma_{1}}2\pi v_{12}^{2}\rho_{1}\delta_{ij} (23)
(V32TG3Γ~3G3V32)ij=2πv322ρ3δij.\langle(V_{32}^{T}G_{3}\tilde{\Gamma}_{3}G_{3}^{\dagger}V_{32})_{ij}\rangle=2\pi v_{32}^{2}\rho_{3}\delta_{ij}. (24)

in the limit 𝒯k/Nk1\mathcal{T}_{k}/N_{k}\gg 1 (see Appendix B).

We can cast the formulas in a more transparent notation by defining decay widths of the transition states to the right-hand and left-hand reservoirs as

ΓR=2πv322ρ3\Gamma_{R}=2\pi v_{32}^{2}\rho_{3} (25)

and

ΓL=2πv122ρ1.\Gamma_{L}=2\pi v_{12}^{2}\rho_{1}. (26)

Using Eq. (50) in Appendix B, one obtains

Tin,3\displaystyle\langle T_{\rm in,3}\rangle =\displaystyle= 𝒯in𝒯1ΓLΓRi,j|(G2)ij|2b3(Tbb3Tb),\displaystyle\frac{\mathcal{T}_{\rm in}}{\mathcal{T}_{1}}\,\Gamma_{L}\Gamma_{R}\sum_{i,j}\langle|(G_{2})_{ij}|^{2}\rangle\,\sum_{b\in 3}\left(\frac{T_{b}}{\sum_{b^{\prime}\in 3}T_{b^{\prime}}}\right), (28)
=\displaystyle= 𝒯in𝒯1ΓLΓRi,j|(G2)ij|2\displaystyle\frac{\mathcal{T}_{\rm in}}{\mathcal{T}_{1}}\,\Gamma_{L}\Gamma_{R}\sum_{i,j}\langle|(G_{2})_{ij}|^{2}\rangle

where 𝒯in=2πΓinρ1/N1\mathcal{T}_{\rm in}=2\pi\Gamma_{\rm in}\rho_{1}/N_{1}. It is remarkable that the ensemble average of the transmission coefficient is independent of Γ3\Gamma_{3}, and thus the insensitive property bertsch-hagino2021 ; hagino-bertsch2021 ; bertsch-hagino2023 is realized.

Notice that G2G_{2} in Eq. (10) can be written

G2=(H2i(ΓL/2+ΓR/2)𝟙)1.G_{2}=(H_{2}-i(\Gamma_{L}/2+\Gamma_{R}/2)\mathbb{1})^{-1}. (29)

Then if H2H_{2} is diagonal555In fact, this restriction is not necessary. The matrix H2H_{2} can always be diagonalized by an orthogonal transformation which has no effect on the ensemble average. with matrix elements (H2)ij=Eiδi,j(H_{2})_{ij}=E_{i}\delta_{i,j}, the transmission coefficient becomes

Tin,3=𝒯in𝒯1iΓLΓREi2+(ΓL+ΓR)2/4.\langle T_{\rm in,3}\rangle=\frac{\mathcal{T}_{\rm in}}{\mathcal{T}_{1}}\sum_{i}\frac{\Gamma_{L}\Gamma_{R}}{E_{i}^{2}+(\Gamma_{L}+\Gamma_{R})^{2}/4}. (30)

This is a well-known formula for electron transport through intermediate resonances alhassid2000 ; bertsch2014 . It agrees with an underlying assumption in transition-state models, that the contributions of the individual transition states are additive in the total transmission probability 666This may not be the case for the probability fluxes through the individual transition states..

The formula also shows that the contribution to the transmission coefficient is suppressed when the energy of a bridge state is outside the range of ±(ΓL+ΓR)/2\pm(\Gamma_{L}+\Gamma_{R})/2 around the incident energy. This is marked contrast to models in which the transition state is an internal channel that remains open at all energies above the threshold. To maintain the correspondence to the CI formulation, one would have to include highly excited configurations that carry momentum along a collective coordinate. At some point the model would break down because the coupling matrix element would become small compared to vkv_{k}, the coupling strength of the configurations within the GOE’s.

III.1 Numerical examples

Eq. (21-23,29) are valid in limit 𝒯k/Nk1\mathcal{T}_{k}/N_{k}\gg 1. It is of interest to see how their accuracy is degraded at finite values of these parameters, as well as the sensitivity to other assumptions in the model. Table I shows two aspects of the model Hamiltonian and its reduction to the 𝒯k/Nk1\mathcal{T}_{k}/N_{k}\gg 1 limit. First, one sees that reduction is accurate only to a 20% level despite the seeming large value for 𝒯k/Nk=20\mathcal{T}_{k}/N_{k}=20. The slow convergence can be traced to the r.m.s. fluctuation exhibited in Eq. (21), dying off only as Nk1/2N_{k}^{-1/2}. The table also demonstrates for N2=1N_{2}=1 and 22 that the transmission probability scales quite well with the number of transition states at the same energy, given that their couplings are orthogonal and have the same strength. This is implicit in the reduction to Eq. (30).

N2N_{2} 𝒯k/Nk1\mathcal{T}_{k}/N_{k}\gg 1 Eq. (2)
1 0.0010 0.00079
2 0.0020 0.00148
Table 1: Comparison of transmission probability calculated by the trace formula Eq. (2) and by the 𝒯k/Nk1\mathcal{T}_{k}/N_{k}\gg 1 reduction, Eq. (28). The parameter values in the reservoir Hamiltonians are (vk,vk2,γk,Nk)=(0.1,0.1,0.1,100)(v_{k},v_{k2},\gamma_{k},N_{k})=(0.1,0.1,0.1,100). The H2H_{2} matrix contains N2N_{2} transition states at energies Ei=0E_{i}=0, and the partial width of the entrance channel is γin=0.01\gamma_{\rm in}=0.01. The ensemble average in Eq. (2) was carried out with 10000 samples; the statistical uncertainties are of the order 1~{}1%.

Next we examine the sensitivity to the decay matrix elements. The dependence on γin\gamma_{\rm in} is trivial as it contributes quadraticly when it is small compared to other widths. The exit decay is independent of γ3\gamma_{3} in the reduced formula. This is tested in Table II, varying γ3\gamma_{3} and keeping the other parameters fixed. One sees that the dependence is quite flat within the boundaries ρ31<γ3<N3v3\rho_{3}^{-1}<\gamma_{3}<\sqrt{N_{3}}v_{3}.

γ3\gamma_{3} Eq. (2)
0.05 0.00074
0.1 0.00079
0.2 0.00083
0.4 0.00085
Table 2: Dependence of transmission probability on the parameter γ3\gamma_{3}. The other parameters are the same as in the caption to Table I for the N2=1N_{2}=1 Hamiltonian.

It is also interesting to see how the formula breaks down when the condition 𝒯k/Nk1\mathcal{T}_{k}/N_{k}\gg 1 is no longer satisfied. When the direct decay of the first reservoir becomes small, more of the probability flux crosses the barrier and Tin,3T_{{\rm in},3} only competes with elastic scattering. Table 3 shows a comparison of the analytic reduction with the full trace evaluation in Eq. (2).

𝒯k/Nk\mathcal{T}_{k}/N_{k} R
1 2.75
2 2.00
6 1.45
10 1.30
20 1.25
Table 3: Dependence of transmission probability on 𝒯1/N1\mathcal{T}_{1}/N_{1} varying the parameter γ1\gamma_{1}. The other parameters are the same as in the caption to Table I for the N2=1N_{2}=1 Hamiltonian. The column RR shows the ratio of the analytic reduction Eq. (28) to value obtained with Eq. (2) taking 10000 samples of the GOE’s. The statistical errors decrease from 2~{}2% for the first entry to 1~{}1% for the last one.

One sees that the analytic reduction becomes quite inaccurate for the smaller values of 𝒯1/N1\mathcal{T}_{1}/N_{1}.

IV summary

While transition-state theory for decay of quantum complex systems is usually derived with a statistical approach, we have successfully derived it starting from a matrix Hamiltonian as is commonly used in configuration-interaction formulations. To this end, we considered two reservoirs described by random matrices. One of the configurations in the first reservoir undergoes transitions to configurations in the second reservoir through bridge configurations between them. A potential barrier may exist for the bridge configurations. This generalizes a model with a single barrier configuration that was discussed by Weidenmüller weidenmuller2022 .

As in Ref. weidenmuller2022 , we have shown that the average transmission coefficient from the entrance configuration to configurations in the second reservoir can be factorized into a product form of the formation and the decay probabilities of transition channels, in the limit of 𝒯3/N31\mathcal{T}_{3}/N_{3}\gg 1. This is also a consequence of the usual starting point of transition state theory, that once the system passes the barrier, it never comes back.

If the condition 𝒯1/N11\mathcal{T}_{1}/N_{1}\gg 1 is also satisfied, the transmission coefficient is further simplified to a product of the population probability of the first reservoir, the transmission coefficient over the barrier, and the decay probability of the configurations in the second reservoir. In that case the transmission coefficient can be expressed in terms of Breit-Wigner resonance decays, as has been long known in nuclear physics and in the field of electron transport.

Transition-state theory is a landmark framework for decays of quantum complex systems, but conditions for transition-state theory to work have not yet been well clarified. The microscopic derivation based on the random matrix approach shown in this paper provides a necessary condition for transition-state theory to work. Such consideration would be important in the decay of complex systems at energies close to barrier tops.

Acknowledgements.
We thank Hans Weidenmüller for useful discussions. This work was supported in part by JSPS KAKENHI Grant Numbers JP19K03861 and JP23K03414.

Appendix A the Green’s function for a block-tridiagonal Hamiltonian

We invert the matrix

HiΓ~in/2iΓ~1/2iΓ~3/2E=(H~1V120V12TH~2V230V23TH~3),H-i\tilde{\Gamma}_{\rm in}/2-i\tilde{\Gamma}_{1}/2-i\tilde{\Gamma}_{3}/2-E=\left(\begin{matrix}\tilde{H}_{1}&V_{12}&0\cr V_{12}^{T}&\tilde{H}_{2}&V_{23}\cr 0&V_{23}^{T}&\tilde{H}_{3}\end{matrix}\right), (31)

where H~i\tilde{H}_{i} are defined as

H~1\displaystyle\tilde{H}_{1} \displaystyle\equiv H1iΓin/2iΓ1/2E,\displaystyle H_{1}-i\Gamma_{\rm in}/2-i\Gamma_{1}/2-E, (32)
H~2\displaystyle\tilde{H}_{2} \displaystyle\equiv H2E,\displaystyle H_{2}-E, (33)
H~3\displaystyle\tilde{H}_{3} \displaystyle\equiv H1iΓ3/2E.\displaystyle H_{1}-i\Gamma_{3}/2-E. (34)

The Green’s function (6) satisfies the relation,

(H~1V120V12TH~2V32T0V32H~3)(G11G12G13G21G22G23G31G32G33)=(100010001),\left(\begin{matrix}\tilde{H}_{1}&V_{12}&0\cr V_{12}^{T}&\tilde{H}_{2}&V_{32}^{T}\cr 0&V_{32}&\tilde{H}_{3}\end{matrix}\right)\left(\begin{matrix}G_{11}&G_{12}&G_{13}\cr G_{21}&G_{22}&G_{23}\cr G_{31}&G_{32}&G_{33}\end{matrix}\right)=\left(\begin{matrix}1&0&0\cr 0&1&0\cr 0&0&1\end{matrix}\right), (35)

from which one finds

H~1G13+V12G32T=0,\displaystyle\tilde{H}_{1}G_{13}+V_{12}G_{32}^{T}=0, (36)
V12TG13+H~2G23+V23G33=0,\displaystyle V_{12}^{T}G_{13}+\tilde{H}_{2}G_{23}+V_{23}G_{33}=0, (37)
V32G23+H~3G33=1.\displaystyle V_{32}G_{23}+\tilde{H}_{3}G_{33}=1. (38)

From Eqs. (36) and (38), G13G_{13} and G33G_{33} read

G13=H~11V12G23,G_{13}=-\tilde{H}_{1}^{-1}V_{12}G_{23}, (39)

and

G33=H~31H~31V32G23,G_{33}=\tilde{H}_{3}^{-1}-\tilde{H}_{3}^{-1}V_{32}G_{23}, (40)

respectively. Substituting these into Eq. (37), one obtains

G23=(H~2V12TH~11V12V32TH~31V32)1V32TH~31.G_{23}=-(\tilde{H}_{2}-V_{12}^{T}\tilde{H}_{1}^{-1}V_{12}-V_{32}^{T}\tilde{H}_{3}^{-1}V_{32})^{-1}V_{32}^{T}\tilde{H}_{3}^{-1}. (41)

Combining Eqs. (39) and (41), one finally obtains Eq. (7).

Following a similar procedure, one can also derive

G11=G1+G1V12G2V12TG1.G_{11}=G_{1}+G_{1}V_{12}G_{2}V_{12}^{T}G_{1}. (42)

Appendix B Ensemble average of VGΓ~GVTVG\tilde{\Gamma}G^{\dagger}V^{T}

In this Appendix, we evaluate the ensemble average of a matrix VGΓ~GVTVG\tilde{\Gamma}G^{\dagger}V^{T}, where the elements of VV are Gaussian-distributed random numbers with vij2=v2\langle v_{ij}^{2}\rangle=v^{2}, Γ~\tilde{\Gamma} is a constant times the unit matrix Γ~ij=γδi,j\tilde{\Gamma}_{ij}=\gamma\delta_{i,j}, and GG is the Green’s function G=(HiΓ~/2)1G=(H-i\tilde{\Gamma}/2)^{-1}. Here HH is a sample of the N×NN\times N GOE with a level density ρ0\rho_{0} in the center of its spectrum. We follow Refs. bertsch-hagino2021 ; hagino-bertsch2021 to carry out the ensemble averaging. We first express the elements of the Green’s function as

Gij=λϕiλϕjλEλiγ/2,G_{ij}=\sum_{\lambda}\frac{\phi^{\lambda}_{i}\phi^{\lambda}_{j}}{E_{\lambda}-i\gamma/2}, (43)

where EλE_{\lambda} are the eigenvalues of the Hamiltonian HH and ϕλ\phi^{\lambda} are the corresponding eigenfunctions.

The ijij element of VGΓ~GVTVG\tilde{\Gamma}G^{\dagger}V^{T} then reads,

(VGΓ~GVT)ij=ijmmλλViiϕiλϕmλΓ~mmϕmλϕjλVjj(Eλiγ/2)(Eλ+iγ/2).(VG\tilde{\Gamma}G^{\dagger}V^{T})_{ij}=\sum_{i^{\prime}j^{\prime}mm^{\prime}\lambda\lambda^{\prime}}\frac{V_{ii^{\prime}}\phi^{\lambda}_{i^{\prime}}\phi^{\lambda}_{m}\tilde{\Gamma}_{mm^{\prime}}\phi^{\lambda^{\prime}}_{m^{\prime}}\phi^{\lambda^{\prime}}_{j^{\prime}}V_{j^{\prime}j}}{(E_{\lambda}-i\gamma/2)(E_{\lambda^{\prime}}+i\gamma/2)}. (44)

The sum over mm^{\prime} is evaluated as

mΓ~mmϕmλ=γϕmλ,\sum_{m^{\prime}}\tilde{\Gamma}_{mm^{\prime}}\phi^{\lambda^{\prime}}_{m^{\prime}}=\gamma\phi^{\lambda^{\prime}}_{m}, (45)

since Γ~\tilde{\Gamma} is assumed to be proportional to the unit matrix. Next the orthogonality of the eigenvectors λ\lambda and λ\lambda^{\prime} permits the sum over λ\lambda^{\prime} to be dropped with replacement λ\lambda^{\prime} by λ\lambda. Then Eq. (44) reduces to

(VGΓ~GVT)ij=γijλ(ViiVjjϕiλϕjλEλ2+γ2/4).(VG\tilde{\Gamma}G^{\dagger}V^{T})_{ij}=\gamma\sum_{i^{\prime}j^{\prime}}\sum_{\lambda}\left(\frac{V_{ii^{\prime}}V_{jj^{\prime}}\phi^{\lambda}_{i^{\prime}}\phi^{\lambda}_{j^{\prime}}}{E_{\lambda}^{2}+\gamma^{2}/4}\right). (46)

Next we take the ensemble average of the factor in parentheses. One of the properties of the GOE is that fluctuations about the average level densities are small, so the ensemble averages of the numerator and denominator are uncorrelated. The denominator average is bertsch-hagino2021 ; hagino-bertsch2021

λ1Eλ2+γ2/4=2πρ0γ.\left\langle\sum_{\lambda}\frac{1}{E_{\lambda}^{2}+\gamma^{2}/4}\right\rangle=2\pi\frac{\rho_{0}}{\gamma}. (47)

For the numerator, we first notice that the dot products 𝑽iϕλiViiϕiλ\mbox{\boldmath$V$}_{i}\cdot\mbox{\boldmath$\phi$}_{\lambda}\equiv\sum_{i^{\prime}}V_{ii^{\prime}}\phi_{i^{\prime}}^{\lambda} are Gaussian distributed with 𝑽iϕλ=vriλ\mbox{\boldmath$V$}_{i}\cdot\mbox{\boldmath$\phi$}_{\lambda}=v\,r_{i\lambda}, where riλr_{i\lambda} is a random number satisfying riλ=0\langle r_{i\lambda}\rangle=0 and riλriλ=δi,iδλλ\langle r_{i\lambda}r_{i^{\prime}\lambda^{\prime}}\rangle=\delta_{i,i^{\prime}}\delta_{\lambda\lambda^{\prime}}. One thus obtains

(𝑽iϕλ)(𝑽jϕλ)=v2δi,j,\langle(\mbox{\boldmath$V$}_{i}\cdot\mbox{\boldmath$\phi$}_{\lambda})(\mbox{\boldmath$V$}_{j}\cdot\mbox{\boldmath$\phi$}_{\lambda})\rangle=v^{2}\delta_{i,j}, (48)

which leads to

(VGΓ~GVT)ij=2πv2ρ0δi,j.\left\langle(VG\tilde{\Gamma}G^{\dagger}V^{T})_{ij}\right\rangle=2\pi v^{2}\rho_{0}\,\delta_{i,j}. (49)

We also need the ensemble average of VGΓ~inGVTVG\tilde{\Gamma}_{\rm in}G^{\dagger}V^{T} with G=(HiΓ~/2iΓ~in/2)1G=(H-i\tilde{\Gamma}/2-i\tilde{\Gamma}_{\rm in}/2)^{-1}, where Γ~in\tilde{\Gamma}_{\rm in} is given by Eq. (17). In this case, the sum over mm is restricted to a single state in Eq. (44). Due to the invariance of the averages under unitary transformations, the sum over m,mm,m^{\prime} becomes ϕλ,inϕλ,in=δλ,λ/N\langle\phi_{\lambda,{\rm in}}\phi_{\lambda^{\prime},\rm{in}}\rangle=\delta_{\lambda,\lambda^{\prime}}/N. The final result is

(VGΓ~inGVT)ij=ΓinNγ2πv2ρ0δi,j.\left\langle(VG\tilde{\Gamma}_{\rm in}G^{\dagger}V^{T})_{ij}\right\rangle=\frac{\Gamma_{\rm in}}{N\gamma}2\pi v^{2}\rho_{0}\,\delta_{i,j}. (50)

.

References

  • (1) P. Hänggi, P. Talkner, and M. Borkovec, Rev. Mod. Phys. 62, 251 (1990).
  • (2) D.G. Truhlar, B.C. Garrett, and S.J. Klippenstein, J. Phys. Chem. 100 12771 (1996).
  • (3) U. Weiss, Quantum Dissipative Systems 4th edition, (World Scientific, Singapore, 2012).
  • (4) N. Bohr and J.A. Wheeler, Phys. Rev. 56, 426 (1939).
  • (5) R.A. Marcus and O.K. Rice, J. Phys. and Colloid Chem. 55, 894 (1951).
  • (6) H.A. Weidenmüller, Rev. Mod. Phys. 81, 539 (2009).
  • (7) G.F. Bertsch and K. Hagino, Phys. Rev. C107, 044615 (2023).
  • (8) W.F. Polik, D.R. Guyer, W.H. Miller, and C.B. Moore, J. Chem. Phys. 92, 3471 (1990).
  • (9) W.H. Miller, R. Hernandez, C.B. Moore, and W.F. Polik, J. Chem. Phys. 93, 5657 (1990).
  • (10) R. Hernandez, W.H. Miller, C.B. Moore, and W.F. Polik, J. Chem. Phys. 99, 950 (1993).
  • (11) H.A. Weidenmüller, Phys. Rev. E105, 044143 (2022).
  • (12) H.A. Weidenmüller, arXiv:2311.08030.
  • (13) Y. Alhassid, Rev. Mod. Phys. 72, 895 (2000).
  • (14) P.S. Damle, A.W. Ghosh, and S. Datta, Phys. Rev. B 64, 201403 (2001).
  • (15) S. Datta, Electronic Transport in Mesoscopic Systems, (Cambridge University Press, Cambridge, 1995), Eq. (3.5.20).
  • (16) W.H. Miller, Acc. Chem. Res. 26, 174 (1993).
  • (17) H. Haug and A-P. Jauho, Quantum kinetics in transport and optics of semiconductors, (Spring, Berlin, 2008), Eq. (12.37).
  • (18) Y. Alhassid, G.F. Bertsch, and P. Fanto, Ann. Phys. 424 168381 (2021).
  • (19) Y. Meir and N.S. Wingreen, Phys. Rev. Lett. 68, 2512 (1992), Eq. (7).
  • (20) M. Büttiker, Phys. Rev. Lett. 57, 1761 (1986).
  • (21) R. Venugopal, M. Paulsson, S. Goasguen, S. Datta, and M.S. Lundstrom, J. of App. Phys. 93, 5613 (2003).
  • (22) S. Datta, Quantum Transport: Atom to Transistor (Cambridge University Press, Cambridge, 2005).
  • (23) K. Hagino and G.F. Bertsch, Phys. Rev. E104, L052104 (2021).
  • (24) O.I. Lobkis, R.L. Weaver, and I. Rozkhov, J. Sound Vib. 237, 281, (2000).
  • (25) S.B. Fedeli and Y.V. Fyodorov, J. Phys. A: Math. Theor. 53 165701 (2020).
  • (26) G.F. Bertsch and K. Hagino, J. Phys. Soc. Jpn. 90, 114005 (2021).
  • (27) G.F. Bertsch, arXiv: 1407.1899 [nucl-th].