A general suite for benchmarking nonadiabatic dynamics

In this tutorial we provide a general suite for benchmarking nonadiabatic dynamics described in ref 1 together with data (this link is the collection) of all figures. Among all the dynamics tested, the robustness of nonadiabatic field (NaF)23 is highlighted.

For a full publication list of the generalized constraint coordinate-momentum phase space (CPS) and NaF please see this page.

For an illustration of the nonadiabatic force on the surface hopping dynamics please see this page (reported in ref2).

1. LVCM (Linear vibronic coupling models)

Figures_and_data

The Hamiltonian of a LVCM (linear vibronic coupling model) reads

(1)H^=H^0+H^C,

where

(2)H^0=k=1Nnucωk2(P¯^k2+R¯^k2),
(3)H^C=n=1F(En+k=1Nnucκk(n)R¯^k)|nn|+nmF(k=1Nnucλk(nm)R¯^k)|nm|.

Note that the relation between the dimensionless weighted normal-mode coordinate/momentum, {R¯k,P¯k}, and the corresponding canonical (mass-weighted) coordinate/momentum, {Rk,Pk}, in the diabatic representation reads

(4)R¯k=ωkRk,P¯k=Pk/ωk

The parameters of LVCMs in the suite are listed as follows, except the parameters of the 7-state 39-mode LVCM of Thymine in ref 4 which are kindly provided by Martha Yaghoubi Jouybari and Fabrizio Santoro.

Table 1. Parameters of 2-state 3-mode LVCM of Pyrazine (unit: eV) 5

ParametersValue
E1,E23.94, 4.84
κ1(1),κ2(1),κ1(2),κ2(2)0.037, -0.105, -0.254, 0.149
λ3(12)0.262
ω1,ω2,ω30.126, 0.074, 0.118

Table 2. Parameters of 24-mode LVCM of Pyrazine (unit: eV)6

ParametersValue
E1,E2-0.4617, 0.4617
λ1(12)0.1825
Modeωκ(1)κ(2)
10.0936  
20.074-0.09640.1194
30.12730.04700.2012
40.15680.15940.0484
50.13470.0308-0.0308
60.34310.0782-0.0782
70.11570.0261-0.0261
80.32420.0717-0.0717
90.36210.0780-0.0780
100.26730.0560-0.0560
110.30520.0625-0.0625
120.09680.0188-0.0188
130.05890.0112-0.0112
140.04000.0069-0.0069
150.17260.0265-0.0265
160.28630.0433-0.0433
170.24840.0361-0.0361
180.15360.0210-0.0210
190.21050.0281-0.0281
200.07780.0102-0.0108
210.22940.0284-0.0284
220.19150.0196-0.0196
230.40000.0306-0.0306
240.38100.0269-0.0269

In our simulation of the pyrazine models, the second diabatic (electronic) state is initially occupied, and the nuclear initial condition is set to be the Wigner distribution of the vibrational ground state,

(5)ρW(R¯,P¯)exp[k=1N(R¯k2+P¯k2)].

The numerically exact results in our data are produced by MCTDH7.

Table 3. Parameters of 3-state 2-mode LVCM of Cr(CO)5 (unit: eV)8

ParametersValue
E1,E2,E30.0424, 0.0424, 0.4344
κ2(1),κ2(2)-0.0328, 0.0328
λ1(12),λ1(23),λ2(13)0.0328, -0.0978, -0.0978
ω1,ω20.0129, 0.0129

In our test of the Cr(CO)5 model, the second diabatic (electronic) state is initially occupied, and the nuclear initial condition is set to be the Wigner distribution of the following Wigner distribution,

(6)ρW(R¯,P¯)exp[k=12((R¯kr)22αk2+2αk2P¯k2)],

where r1=0,r2=14.3514,α1=0.4501, and α2=0.4586. The numerically exact results in our data are produced by MCTDH, taken from ref 8.

In our test of the Thymine model, the ππ2 state is initially occupied with each nuclear mode in its vibrational ground state, i.e., eq (5). The numerically exact results in our data are produced by ML-MCTDH, taken from ref 4.

2. Tully models

Figures_and_data

The single avoided crossing (SAC) model9 reads

(7)V11(R)=V22(R)=A[1exp(B|R|)]sgn(R),V12(R)=V21(R)=Cexp(DR2),

where A=0.01,B=1.6,C=0.005,D=1 (in a.u.). The dual avoided crossing (DAC) model9 reads

(8)V11(R)=0,V22(R)=Aexp(BR2)+E0,V12(R)=V21(R)=Cexp(DR2),

with A=0.01,B=1.6,C=0.005,D=1.0 (in a.u.). The extended coupling region (ECR) model 9 reads

(9)V11(R)=V22(R)=E0,V12(R)=V21(R)=C[exp(BR)h(R)+(2exp(BR))h(R)],

where E0=0.0006,B=0.9,C=0.1 (in a.u.), and h(R) denotes the Heaviside step function. In our data, the nuclear DOF is with mass m=2000 (in a.u.). The electronic ground state is selected as the initial state, while the initial condition for the nuclear DOF follows the Wigner distribution

(10)ρW(R,P)exp(α(RR0)2(PP0)2/α)

with width parameter α=1. The coordinate center R0 for SAC, DAC, and ECR models is set to -3.8, -10.0, and -13.0, respectively. The initial momentum P0 is scanned and is used as the x-axis in our figures.

The exact results in our data are provided by DVR10.

3. 3-state photodissociation models

Figures_and_data

The potential matrix elements of anharmonic 3-state photodissociation models11 reads

(11)Vii(R)=Di[1eβi(RRi)]2+Ci,i=1,2,3.Vij(R)=Vji(R)=Aijeαij(RRij)2,i,j=1,2,3; and ij.

The parameters are listed as follows,

Table 4. Parameters of the 3-state photodissociation models (unit: a.u.)11

ParametersModel 1Model 2Model 3
D1,D2,D30.003,0.004,0.0030.020,0.010,0.0030.020,0.020,0.003
β1,β2,β30.65,0.60,0.650.65,0.40,0.650.40,0.65,0.65
R1,R2,R35.0,4.0,6.04.5,4.0,4.44.0,4.5,6.0
C1,C2,C30.00,0.01,0.0060.00,0.01,0.020.02,0.00,0.02
A12,A23,A310.002,0.002,0.00.005,0.0,0.0050.005,0.0,0.005
R12,R23,R313.40,4.80,0.003.66,0.00,3.343.40,0.00,4.97
α12,α23,α3116.0,16.0,0.032.0,0.0,32.032.0,0.0,32.0
Re2.93.32.1

The initial electronic configuration is set to the first diabatic state, and the nuclear DOF is sampled from the Wigner distribution of the ground state,

(12)ρW(R,P)exp(mω(RRe)2P2/mω),

where m=20000 a.u. is the mass of the nuclear DOF, ω=0.005 a.u. is the vibrational frequency of the ground state, and the center Re is listed in Table 4.

The exact results in our data are produced by DVR10.

4. 2-state photodissociation model

Figures_and_data

The potential matrix elements of the anharmonic 2-state photodissociation model111 reads

(13)Vii(R)=Di[1eβi(RRi)]2+Ci,i=1,2.Vij(R)=Vji(R)=Aijeαij(RRij)2,i,j=1,2; and ij.

The parameters are listed as follows,

Table 5. Parameters of the 2-state photodissociation model (unit: a.u.)111

ParametersValues
D1,D20.020,0.003
β1,β20.65,0.65
R1,R24.5,4.4
C1,C20.00,0.02
A120.005
R123.34
α128.0
Re3.3

The initial condition and mass are identical to those of Model 2 of the 3-state photodissociation models.

The exact results in our data are produced by DVR10.

5. Spin-boson models

Figures_and_data

The Hamiltonian of a spin-boson model reads12

(14)H^=H^b+H^s+H^bs=j=1Nb(P^j22+12ωj2R^j2)+[εΔΔε]+[1001]j=1NbcjR^j

We use the discretization scheme for the Ohmic spectral density12 with the Kondo parameter α and the cut-off frequency ωc, which results in the following expressions1314 with Nb discrete modes:

(15)ωj=ωcln(1j1+Nb),cj=ωjαωcNb+1,j=1,,Nb

We employ four specific spin-boson models with ε=Δ=1 at β=5. The parameters15 of the bath DOFs are α{0.1,0.4} and ωc{1,2.5} with Nb=300.

The system occupies the first diabatic electronic state, and the nuclear DOFs are sampled from their equilibrium Wigner density. The numerically exact results in our data15 are calculated by eHEOM16.

6. Site-exciton models

Figures_and_data

The Hamiltonian of site-exciton models reads17

(16)H^=H^s+H^b+H^sb,

where H^b=n,j12(P^nj2+ωnj2R^nj2) and H^sb=n,jcnjR^nj|nn|. We employ the Debye spectral density

(17)J(ω)=2λωcωωc2+ω2

for each state, where λ and ωc denote the bath reorganization energy and the characteristic frequency, respectively. The corresponding discretization scheme181920 is

(18)ωnj=ωctan[π2(1jNb+1)]cnj=2λNb+1ωnj,n=1,,F;j=1,,Nb

For the 7-state FMO model21, the system Hamiltonian reads

(19)H^s=(1241087.75.55.96.713.79.987.71253030.88.20.711.84.35.530.81221053.52.29.66.05.98.253.51232070.717.063.36.70.72.270.71248081.11.313.711.89.617.081.11263039.79.94.36.063.31.339.712440)cm1,

The bath reorganization energy is λ=35 cm1, the characteristic frequency is ωc=106.14 cm1, and Nb=50 for each state.

For 3-state singlet-fission model of pentacene2223, the system Hamiltonian reads

(20)H^s=[20050050300500500] meV

The bath reorganization energy is λ=0.1 eV, the characteristic frequency is ωc=0.18 eV, and Nb=200 for each state.

In these two site-exciton models, the bath DOFs are sampled from their equilibrium Wigner distribution, and the first diabatic electronic state is initially occupied.

The numerically exact results in our data1 are produced by TD-DMRG24 in our FMO model, and by HEOM25 in our SF model.

7. Atom-in-cavity model

Figures_and_data

The total Hamiltonian for the atom-in-cavity models2627 can be decomposed into three parts

(21)H^=H^a+H^p+H^c

where

(22)H^a=n=1Fεn|nn|
(23)H^p=j=1N12(P^j2+ωj2R^j2)
(24)H^c=nmF(j=1Nωjλj(r0)R^j)μnm|nm|

The quantity μnm=μmn denotes the transition dipole moment between the n-th and m-th atomic energy levels. The coupling strength between the atom and the j-th optical field mode is given by

(25)λj(r0)=2/ε0Lsin(jπr0/L),

where L=236200 a.u. represents the cavity length, r0=L/2 corresponds to the cavity center, and ε0 is the vacuum permittivity. The mode frequency is determined by ωj=jπc/L, with c=137.036 a.u. to be the light speed in vacuum. We employ 400 standing wave modes and model the hydrogen atom system as a three-level system. The energy levels and transition dipole moments (in atomic units) are: ε1=0.6738, ε2=0.2798, ε3=0.1547, μ12=1.034, μ23=2.536, and μ13=0. A reduced two-level system is also considered, where only the lowest two energy levels of the hydrogen atom is retained.

Initially the highest atomic state is occupied, with each optical field mode in their optical vacuum state with Wigner distribution reading

(26)ρW(Rj,Pj)exp(Pj2/ωjωjRj2)

The exact results are produced by truncated configuration interaction and are taken from refs2627.

8. Holstein model

Figures_and_data

The Hamiltonian of the 1-dim Holstein model2829 is given by

(27)H^=H^s+H^p+H^c

where the electronic system is described by a nearest-neighbor tight-binding model:

(28)H^s=nEna^na^n+V(a^n+1a^n+a^na^n+1)

Here, a^n and a^n denote the fermionic creation and annihilation operators at the n-th lattice site, while En and V represent the on-site energy and the coupling between neighboring sites, respectively. For a periodic chain with a finite number of lattice sites L, one typically sets En=0 and assumes a^n+L=a^n.

The phonon Hamiltonian describes a bosonic bath environment:

(29)H^p=jnωj(b^jnb^jn+1/2)

The term H^c encompasses all electron-phonon coupling effects:

(30)H^c=n,jcj(1)(b^jn+b^jn)a^na^n

Considering the low carrier concentration regime, where the system contains only a single electron. In this case, the fermionic operators act within the single-particle Fock space, and the system Hamiltonian becomes isomorphic to a multistate Hamiltonian30. Consequently, the fermionic operators can be expressed in terms of state representations:

(31)a^n=|01,,1n,,0L01,,0n,,0L||n0~|,a^n=|01,,0n,,0L01,,1n,,0L||0~n|

or equivalently,

(32)a^na^m=|nm|

We set V=0.083 eV,d=R=7.2  and L=213132, where the unmentioned parameters are used in the calculation of correlation function. Other parameters are listed as follows.

Table 6. Parameters of the Holstein model in refs3132

Modeωj (meV)gj=cj(1)/ωj
110.00.96
227.00.38
378.00.25
4124.00.20
5149.00.15
6167.00.31
7169.00.13
8190.00.20
9198.00.31

We guide readers to Section S5-S6 of Supporting Information of ref1 for details of initial condition and correlation function. The exact results in our data are taken from ref31, produced by TD-DMRG.


9. References

 


1 B. Wu, B. Li, X. He, X. Cheng, J. Ren and J. Liu, "Nonadiabatic field: A conceptually novel approach for nonadiabatic quantum molecular dynamics ", J. Chem. Theory Comput. 21, 3775-3813 (2025).
2 B. Wu, X. He and J. Liu, "Nonadiabatic Field on Quantum Phase Space: A Century after Ehrenfest", J. Phys. Chem. Lett., 15(2), 644-658 (2024).
3 X. He, X. Cheng, B. Wu and J. Liu, "Nonadiabatic Field with Triangle Window Functions on Quantum Phase Space", J. Phys. Chem. Lett., 15(20), 5452-5466 (2024).
4 J. A. Green, M. Y. Jouybari, D. Aranda, R. Improta, and F. Santoro, “Nonadiabatic absorption spectra and ultrafast dynamics of DNA and RNA photoexcited nucleobases”, Molecules 26, 1743 (2021).
5 R. Schneider and W. Domcke, “S1-S2 conical intersection and ultrafast S2→S1 internal-conversion in pyrazine”, Chem. Phys. Lett. 150, 235–242 (1988).
6 S. Krempl, M. Winterstetter, H. Plöhn, and W. Domcke, “Path-integral treatment of multi-mode vibronic coupling”, J. Chem. Phys. 100, 926–937 (1994).
7 G. A. Worth, M. H. Beck, A. Jackle and H.-D. Meyer, The MCTDH Package, Version 8.2, (2000). H.-D. Meyer Version 8.3 (2002), Version 8.4 (2007). O. Vendrell and H.-D. Meyer Version 8.5 (2013). Version 8.5 contains the ML-MCTDH algorithm. See http://mctdh.uni-hd.de. (accessed on November 1st, 2023) Used version: 8.5.14.
8 G. A. Worth, G. Welch, and M. J. Paterson, “Wavepacket dynamics study of Cr(CO)5 after formation by photodissociation: Relaxation through an (E ⊕ A) ⊗ e Jahn–Teller conical intersection”, Mol. Phys. 104, 1095–1105 (2006).
9 J. C. Tully, “Molecular-dynamics with electronic-transitions”, J. Chem. Phys. 93, 1061–1071 (1990).
10 D. T. Colbert and W. H. Miller, "A novel discrete variable representation for quantum mechanical reactive scattering via the S-matrix Kohn method", J. Chem. Phys. 96, 1982-1991 (1992).
11 E. A. Coronado, J. Xing, and W. H. Miller, “Ultrafast non-adiabatic dynamics of systems with multiple surface crossings: a test of the Meyer-Miller Hamiltonian with semiclassical initial value representation methods”, Chem. Phys. Lett. 349, 521–529 (2001).
12 A. J. Leggett, S. Chakravarty, A. T. Dorsey, M. P. A. Fisher, A. Garg, and W. Zwerger, “Dynamics of the dissipative two-state system”, Rev. Mod. Phys. 59, 1–85 (1987).
13 N. Makri, “The linear response approximation and its lowest order corrections: An influence functional approach”, J. Phys. Chem. B 103, 2823–2829 (1999).
14 P. L. Walters, T. C. Allen, and N. Makri, “Direct determination of discrete harmonic bath parameters from molecular dynamics simulations”, J. Comput. Chem. 38, 110–115 (2017).
15 X. He, Z. Gong, B. Wu and J. Liu, "Negative zero-point-energy parameter in the Meyer–Miller mapping model for nonadiabatic dynamics", J. Phys. Chem. Lett. 12, 2496–2501 (2021).
16 Z. Tang, X. Ouyang, Z. Gong, H. Wang and J. Wu, "Extended hierarchy equation of motion for the spin-boson model", J. Chem. Phys. 143, 224112 (2015).
17 M. N. Yang and G. R. Fleming, "Influence of phonons on exciton transfer dynamics: Comparison of the Redfield, Forster and modified Redfield equations", Chem. Phys. 275, 355-372 (2002).
18 H. Wang, X. Song, D. Chandler, and W. H. Miller, “Semiclassical study of electronically nonadiabatic dynamics in the condensed-phase: spin-boson problem with Debye spectral density”, J. Chem. Phys. 110, 4828–4840 (1999).
19 M. Thoss, H. Wang, and W. H. Miller, “Self-consistent hybrid approach for complex systems: application to the spin-boson model with Debye spectral density”, J. Chem. Phys. 115, 2991–3005 (2001).
20 I. R. Craig, M. Thoss, and H. Wang, “Proton transfer reactions in model condensed phase environments: accurate quantum dynamics using the multilayer multiconfiguration time-dependent Hartree approach”, J. Chem. Phys. 127, 144503 (2007).
21 A. Ishizaki and G. R. Fleming, "Theoretical examination of quantum coherence in a photosynthetic system at physiological temperature", Proc. Natl. Acad. Sci. U. S. A. 106, 17255-17260 (2009).
22 W.-L. Chan, T. C. Berkelbach, M. R. Provorse, N. R. Monahan, J. R. Tritsch, M. S. Hybertsen, D. R. Reichman, J. Gao and X.-Y. Zhu, "The quantum coherent mechanicsm for singlet fission: Experiment and theory", Acc. Chem. Res. 46, 1321-1329 (2013).
23 G. H. Tao, "Electronically nonadiabatic dynamics in singlet fission: A quasi-classical trajectory simulation", J. Phys. Chem. C 118, 17299-17305 (2014).
24 J. J. Ren, W. T. Li, T. Jiang, Y. H. Wang and Z. G. Shuai, "Time-dependent density matrix renormalization group method for quantum dynamics in complex systems", Wiley Interdiscip. Rev. Comput. Mol. Sci. 12, e1614 (2022).
25 Y. Tanimura and R. Kubo, "Time evolution of a quantum system in contact with a nearly Gaussian-Markoffian noise bath", J. Phys. Soc. Jpn. 58, 101-114 (1989).
26 N. M. Hoffmann, C. Schäfer, A. Rubio, A. Kelly, H. Appel, "Capturing vacuum fluctuations and photon correlations in cavity quantum electrodynamics with multitrajectory Ehrenfest dynamics.", Phys. Rev. A 99, 063819 (2019).
27 N. M. Hoffmann, C. Schäfer,N. Säkkinen, A. Rubio, H. Appel, A. Kelly, "Benchmarking semiclassical and perturbative methods for real-time simulations of cavity-bound emission and interference", J. Chem. Phys. 151, 244113 (2019).
28 T. Holstein, “Studies of polaron motion: Part II. The “small” polaron”, Ann. Phys. 8, 343–389 (1959).
29 K. Hannewald and P. Bobbert, “Ab initio theory of charge-carrier conduction in ultrapure organic crystals”, Appl. Phys. Lett. 85, 1535–1537 (2004).
30 J. Liu, “Isomorphism between the multi-state Hamiltonian and the second-quantized many-electron Hamiltonian with only 1-electron interactions”, J. Chem. Phys. 146, 024110 (2017).
31 W. Li, J. Ren, and Z. Shuai, “Finite-temperature TD-DMRG for the carrier mobility of organic semiconductors”, J. Phys. Chem. Lett. 11, 4930–4936 (2020).
32 W. Li, J. Ren, and Z. Shuai, “A general charge transport picture for organic semiconductors with nonlocal electron-phonon couplings”, Nat. Commun. 12, 4260 (2021).