Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Dissipative split-charge formalism: Ohm’s law, Nyquist noise, and non-contact friction

Martin H. Müser Dept. of Materials Science and Engineering, Saarland University, Germany
(October 1, 2024; October 1, 2024)
Abstract

The split-charge equilibration method is extended to describe dissipative charge transfer similarly as the Drude model, whereby the complex-valued frequency-dependent dielectric permitivities or conductivities of dielectrics and metals can be mimicked at non-zero frequencies. To demonstrate its feasibility, a resistor-capacitor circuit is simulated using an all-atom representation for resistor and capacitor. The dynamics reproduce the expected charging process and Nyquist noise, the latter resulting from the thermal voltages acting on individual split charges. The method bears promise to model friction caused by the motion of charged particles past metallic or highly polarizable media.

I Introduction

The electronegativity equalization method [1, 2, 3] and its charge-equilibration (QE) [4] or fluctuating-charge [5] variants are frequently used in molecular simulations to assign partial charges on the fly. It can be seen as a low-key coarse-grained approximation of density-functional theory, where spatial discretization points coincide with atomic positions [1, 2, 3, 6]. In the simplest case, the total energy of a system is a quadratic polynomial in the set of partial charges {Qi}subscript𝑄𝑖\{Q_{i}\}{ italic_Q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } [7, 8], see also Eq. (1). The zero-order coefficient generally depends strongly on the environment and contains, e.g., pair repulsion and many-body cohesion as described in the embedded-atom-method [9]. The first-order term is, depending on background or application, the electronegativity of an isolated atom [1, 2, 3] or the work function of a metal [10], while the second-order term represents the self-interaction of the charge plus Coulomb interactions between atoms. It can be parameterized to reflect either the chemical hardness of an isolated atom [2, 3] or the Thomas-Fermi screening length in a metal [11, 12]. Adding enviroment dependence to the QE parameters [13] promises to increase the transferability, such that properties of both isolated atoms and solids but potentially also any intermediate structure can be faithfully reproduced. However, even if successful, regular QE schemes still face various intrinsic limitations. The two most relevant to this work are that any QE solid has the static dielectric response function of a metal [11]. Moreover, QE is a pure equilibrium approach lacking dissipation caused by time dependence.

The metallicity problem of regular QE approaches could be solved with the split-charge equilibration (SQE) method [14, 11, 15]. It arose as a phenomenological split between the conventional QE and a pure bond-polarizibility model [16]. In brief, SQE allows transfer of fractional charges only through chemical (covalent and metallic) bonds. In non-metallic systems, this transfer is penalized through a bond hardness term in addition to that caused by the chemical or atomic hardness. The latter can also be losely associated with the Anderson U𝑈Uitalic_U parameter describing the strength of the on-site charge interaction in the Hubbard model [17]. The dielectric permittivity of a solid turns out inversely proportional to the bond hardness [11]. In the redox-SQE formalism, in which integer charges can be donated from one atom or ion to another one without bond penalty, (irreversible, contact-induced) electron-transfer dielectrics can be simulated [18]. This feature made it possible to simulate battery discharge and recharge, such that all processes inside the battery were modelled atomistically, including the build-up of voltage between the electrodes [19]. A ’daemon resistor’, which was not modelled explicitly, allowed current to flow externally between the electrodes according to Ohm’s law. In this study, the roles of all-atom and daemon swap: the battery is the daemon, while the electrical circuit component is modeled explicitly.

To achieve a self-containing all-atom representation of resistors and capacitors, we add an inertial and, more importantly, a dissipative (D) term to the SQE formalism. The resulting D-SQE approach, which can be seen as a hybrid of the Drude model, the polarizable-ion model [20, 21], and a related extension of polarizable force fields [22] can mimick frequency-dependent conductivities of metals or dielectric permitivitties of dielectrics over a broad frequency range and not merely their quasi-static limits. Moreover, resistors can be put together using (coarse-grained) atomic building blocks and when subjected to thermal random forces, or rather random voltages, Nyquist noise in capacitors arise naturally.

When broken down to a truly atomic scale, for example, to a hypothetical one-atom-thin wire, the model can certainly not (yet?) reproduce the intricacies of Luttinger liquids [23]. The goal of the current work is to design simple microscopic charge-transfer dynamics that produce the appropriate response functions of bulk systems. Despite its simplicity, the pursued approach can be expected to further improve the reliability of simulations of ions near electrodes [24, 25, 26] or related systems, where charged particles move relative to metallic or highly polarizable surfaces. Thus, this work should be viewed as a feasibility test rather than as an attempt to describe a specific solid. To this end, the model is introduced first. Next, its generic dielectric response functions are derived. Following that, simulations of an RC element driven by either a battery or thermal noise or by a fixed charge moving through the capacitance are presented. Finally, conclusions are drawn.

II Model and theory

II.1 General model

In this study, all grid-points or (pseudo-) atoms participating in charge-transfer are fixed in space. This allows us to ignore the zero-oder term U0({rij})subscript𝑈0subscript𝑟𝑖𝑗U_{0}(\{r_{ij}\})italic_U start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( { italic_r start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT } ), which can be argued to reflect the energy of a system under the constraint that all atoms be neutral, or, rather to have an integer oxidation state. This ’expansion coefficient’ would surely depend on the method used to assign partial charges. To proceed, let us make the unplausible but pragmatic assumption, that we identified one such method approved by readers, referees, and author alike. The relevant, remaining charge-transfer energy then reads

ΔU=iχiQi+12i,j{Jij(rij)QiQj+κijqij2}Δ𝑈subscript𝑖subscript𝜒𝑖subscript𝑄𝑖12subscript𝑖𝑗subscript𝐽𝑖𝑗subscript𝑟𝑖𝑗subscript𝑄𝑖subscript𝑄𝑗subscript𝜅𝑖𝑗superscriptsubscript𝑞𝑖𝑗2\Delta U=-\sum_{i}\chi_{i}Q_{i}+\frac{1}{2}\sum_{i,j}\left\{J_{ij}(r_{ij})Q_{i% }Q_{j}+\kappa_{ij}q_{ij}^{2}\right\}roman_Δ italic_U = - ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_Q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT { italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( italic_r start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ) italic_Q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_Q start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_κ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT } (1)

in the split-charge formalism, where

Qi=nie+jqijsubscript𝑄𝑖subscript𝑛𝑖𝑒subscript𝑗subscript𝑞𝑖𝑗Q_{i}=n_{i}e+\sum_{j}q_{ij}italic_Q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_e + ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT (2)

is the charge of atom i𝑖iitalic_i, while qij=qjisubscript𝑞𝑖𝑗subscript𝑞𝑗𝑖q_{ij}=-q_{ji}italic_q start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = - italic_q start_POSTSUBSCRIPT italic_j italic_i end_POSTSUBSCRIPT is the partial charge or split charge donated from atom j𝑗jitalic_j to atom i𝑖iitalic_i, which is penalized via a bond-hardness term κijsubscript𝜅𝑖𝑗\kappa_{ij}italic_κ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT. For reasons of simplicity, we assume neutral rather than ionic references and thus omit the oxidation state nisubscript𝑛𝑖n_{i}italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT in the following. The χisubscript𝜒𝑖\chi_{i}italic_χ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT denote the electronegativity of atom i𝑖iitalic_i to which a potential bias or the interaction with external charges can be added. The diagonal elements Jiisubscript𝐽𝑖𝑖J_{ii}italic_J start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT represent the self-interaction or (half) the chemical or atomic hardness of atom i𝑖iitalic_i. In the following, mono-atomic model, they will be denoted as κasubscript𝜅a\kappa_{\textrm{a}}italic_κ start_POSTSUBSCRIPT a end_POSTSUBSCRIPT. The off-diagonal elements Ji,jisubscript𝐽𝑖𝑗𝑖J_{i,j\neq i}italic_J start_POSTSUBSCRIPT italic_i , italic_j ≠ italic_i end_POSTSUBSCRIPT represent the Coulomb interactions between atoms. They can be damped at short distance to mirror the delocalization of charge density [4]. Here, they will simply be set to Ji,ji=1/rijsubscript𝐽𝑖𝑗𝑖1subscript𝑟𝑖𝑗J_{i,j\neq i}=1/r_{ij}italic_J start_POSTSUBSCRIPT italic_i , italic_j ≠ italic_i end_POSTSUBSCRIPT = 1 / italic_r start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT, where rijsubscript𝑟𝑖𝑗r_{ij}italic_r start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT is the distance between atoms i𝑖iitalic_i and j𝑗jitalic_j.

While split charges are not needed in the QE formalism for the modeling of (singly connected) metals, in which ΔUΔ𝑈\Delta Uroman_Δ italic_U is merely minimized w.r.t. the set of charges {Q}𝑄\{Q\}{ italic_Q }, the locality of charge transfer must be imitated whenever electron dynamics matter. To reflect locality, κijsubscript𝜅𝑖𝑗\kappa_{ij}italic_κ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT is assumed to be zero between nearest neighbors in a metal and approximated as infinitely large otherwhise to make the value and thereby the energy of that split charge vanish. In this zero-order approach, split charges only need to be considered between nearest neighbors.

The driving force or voltage acting on a split charge is Vij=ΔU/qijsubscript𝑉𝑖𝑗Δ𝑈subscript𝑞𝑖𝑗V_{ij}=-\partial\Delta U/\partial q_{ij}italic_V start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = - ∂ roman_Δ italic_U / ∂ italic_q start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT, i.e., when using Qkqij=δikδjksubscript𝑄𝑘subscript𝑞𝑖𝑗subscript𝛿𝑖𝑘𝛿𝑗𝑘\partial Q_{k}\partial q_{ij}=\delta_{ik}-\delta{jk}∂ italic_Q start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ∂ italic_q start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = italic_δ start_POSTSUBSCRIPT italic_i italic_k end_POSTSUBSCRIPT - italic_δ italic_j italic_k,

Vij=Jij(QiQj)κijqij+(χiχj).subscript𝑉𝑖𝑗subscript𝐽𝑖𝑗subscript𝑄𝑖subscript𝑄𝑗subscript𝜅𝑖𝑗subscript𝑞𝑖𝑗subscript𝜒𝑖subscript𝜒𝑗V_{ij}=-J_{ij}(Q_{i}-Q_{j})-\kappa_{ij}q_{ij}+(\chi_{i}-\chi_{j}).italic_V start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = - italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( italic_Q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_Q start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) - italic_κ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT + ( italic_χ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_χ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) . (3)

To model dissipative dynamics locally, rather than globally as, e.g., through Eq. (3) in Ref. [26],

Lijq¨ij+Rijq˙ij+κijqij=Φi+Φj+χiχj+Vth(t)subscript𝐿𝑖𝑗subscript¨𝑞𝑖𝑗subscript𝑅𝑖𝑗subscript˙𝑞𝑖𝑗subscript𝜅𝑖𝑗subscript𝑞𝑖𝑗subscriptΦ𝑖subscriptΦ𝑗subscript𝜒𝑖subscript𝜒𝑗subscript𝑉th𝑡L_{ij}\ddot{q}_{ij}+R_{ij}\dot{q}_{ij}+\kappa_{ij}q_{ij}=-\Phi_{i}+\Phi_{j}+% \chi_{i}-\chi_{j}+V_{\textrm{th}}(t)italic_L start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT over¨ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT + italic_R start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT over˙ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT + italic_κ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = - roman_Φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + roman_Φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_χ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_χ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_V start_POSTSUBSCRIPT th end_POSTSUBSCRIPT ( italic_t ) (4)

is proposed as equation of motion, where Φi=jJijQjsubscriptΦ𝑖subscript𝑗subscript𝐽𝑖𝑗subscript𝑄𝑗\Phi_{i}=\sum_{j}J_{ij}Q_{j}roman_Φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_Q start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the electrostatic potential of atom i𝑖iitalic_i produced by the explicitly treated charges. Lijsubscript𝐿𝑖𝑗L_{ij}italic_L start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT is a pseudo inductance arising due to electronic inertia, while Rijsubscript𝑅𝑖𝑗R_{ij}italic_R start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT absorbs dissipative effects in a phenomenological fashion. This generalization is similar in spirit to the (Lorentz-) Drude model and will thus share many strentghs and limitations. At finite temperature, dissipation entails thermal noise Vth(t)subscript𝑉th𝑡V_{\textrm{th}}(t)italic_V start_POSTSUBSCRIPT th end_POSTSUBSCRIPT ( italic_t ), whose expectation values must satisfy

Vth(t)delimited-⟨⟩subscript𝑉th𝑡\displaystyle\langle V_{\textrm{th}}(t)\rangle⟨ italic_V start_POSTSUBSCRIPT th end_POSTSUBSCRIPT ( italic_t ) ⟩ =\displaystyle== 00\displaystyle 0 (5a)
Vth(t)Vth(t)delimited-⟨⟩subscript𝑉th𝑡subscript𝑉thsuperscript𝑡\displaystyle\langle V_{\textrm{th}}(t)V_{\textrm{th}}(t^{\prime})\rangle⟨ italic_V start_POSTSUBSCRIPT th end_POSTSUBSCRIPT ( italic_t ) italic_V start_POSTSUBSCRIPT th end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ⟩ =\displaystyle== 2kBTRδ(tt)2subscript𝑘𝐵𝑇𝑅𝛿𝑡superscript𝑡\displaystyle 2k_{B}TR\delta(t-t^{\prime})2 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T italic_R italic_δ ( italic_t - italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) (5b)

for an instantaneous Ohmic voltage of Rq˙𝑅˙𝑞R\dot{q}italic_R over˙ start_ARG italic_q end_ARG, according to the fluctuation-dissipation theorem [27]. Here kBTsubscript𝑘𝐵𝑇k_{B}Titalic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T is the thermal energy and δ(tt)𝛿𝑡superscript𝑡\delta(t-t^{\prime})italic_δ ( italic_t - italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) is Dirac’s delta function.

Connecting the resulting dissipative SQE model and the Drude model is most easily done by discretizing a metal into cubic elements, as this avoids the need for book keeping of Cartesian indices. Alternatively, let us consider polononium, which is the only element condensing into a simple cubic (sc) lattice. An instantaneous current of an elementary charge e𝑒eitalic_e per time τ𝜏\tauitalic_τ between each two nearest neighbors along one principle direction implies a current density of j=eρve(1/a03)(a0/τ)𝑗𝑒𝜌𝑣𝑒1superscriptsubscript𝑎03subscript𝑎0𝜏j=e\rho v\to e(1/a_{0}^{3})(a_{0}/\tau)italic_j = italic_e italic_ρ italic_v → italic_e ( 1 / italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) ( italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_τ ) in the Drude model, where a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the bond length. Replacing e/τ𝑒𝜏e/\tauitalic_e / italic_τ with q˙˙𝑞\dot{q}over˙ start_ARG italic_q end_ARG means that the SQE current density is q˙/a02˙𝑞superscriptsubscript𝑎02\dot{q}/a_{0}^{2}over˙ start_ARG italic_q end_ARG / italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Ohm’s law is thus satisfied correctly if the split-charge resistance is parameterized as R=ρa0𝑅𝜌subscript𝑎0R=\rho a_{0}italic_R = italic_ρ italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, where ρ𝜌\rhoitalic_ρ is the resistivity. Similarly, the kinetic energy mv2/2𝑚superscript𝑣22mv^{2}/2italic_m italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2, m𝑚mitalic_m being the (effective) electron mass turns into Lq˙2/2𝐿superscript˙𝑞22L\dot{q}^{2}/2italic_L over˙ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 with L=ma02/e2𝐿𝑚superscriptsubscript𝑎02superscript𝑒2L=ma_{0}^{2}/e^{2}italic_L = italic_m italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The numerical value of L𝐿Litalic_L can be assigned to target the plasma frequency. Before addressing this and other dielectric properties in more detail, the default unit system and useful dimensionless constants are introduced.

The units for mass [m]delimited-[]𝑚[m][ italic_m ], charge [Q]delimited-[]𝑄[Q][ italic_Q ], and length [l]delimited-[]𝑙[l][ italic_l ] are the electron mass mesubscript𝑚em_{\textrm{e}}italic_m start_POSTSUBSCRIPT e end_POSTSUBSCRIPT, e𝑒eitalic_e, and a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, respectively. The temperature T𝑇Titalic_T is expressed as thermal energy, while [E]=e2/(4πεa0)delimited-[]𝐸superscript𝑒24𝜋𝜀subscript𝑎0[E]=e^{2}/(4\pi\varepsilon a_{0})[ italic_E ] = italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 4 italic_π italic_ε italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) is the unit of energy. Occasionally, we will revert to extended SI units. This will be marked by an upper index SI on the r.h.s. of pertinent equations. Otherwhise, we use reduced units. For sc α𝛼\alphaitalic_α-Po, non-standard base units are a0=3.345subscript𝑎03.345a_{0}=3.345italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 3.345 Å, [E]=4.305delimited-[]𝐸4.305[E]=4.305[ italic_E ] = 4.305 eV, and [T]=49,960delimited-[]𝑇49960[T]=49,960[ italic_T ] = 49 , 960 K, while selected derived units are [t]=0.3842delimited-[]𝑡0.3842[t]=0.3842[ italic_t ] = 0.3842 fs, resistivity [ρ]=3.455μΩ[\rho]=3.455~{}\mu\Omega\cdot[ italic_ρ ] = 3.455 italic_μ roman_Ω ⋅m, and L=3.971𝐿3.971L=3.971italic_L = 3.971 pF.

A central number of a mono-atomic SQE model is the dimensionless chemical hardness Jiisubscript𝐽𝑖𝑖J_{ii}italic_J start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT. Its value for polonium can be estimated from its electron affinity A8.42eV𝐴8.42eVA\approx 8.42~{}\text{eV}italic_A ≈ 8.42 eV and its ionization energy I1.9eV𝐼1.9eVI\approx 1.9~{}\text{eV}italic_I ≈ 1.9 eV to be Jii=A+I10.3subscript𝐽𝑖𝑖𝐴𝐼10.3J_{ii}=A+I\approx 10.3italic_J start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT = italic_A + italic_I ≈ 10.3 eV/e2superscript𝑒2e^{2}italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, i.e., 2.4absent2.4\approx 2.4≈ 2.4 in reduced units. This exceeds the Madelung constant of the rocksalt structure (rs), αM=1.7476subscript𝛼M1.7476\alpha_{\text{M}}=1.7476italic_α start_POSTSUBSCRIPT M end_POSTSUBSCRIPT = 1.7476. Thus, even when using undamped Coulomb interactions, a spontaneous symmetry breaking from sc to rs is energetically unfavorable. Since rs is the softest charge-transfer direction for the sc lattice [10], the pursued simple parameterization makes the energy of a finite system be positive definite in the split charges.

II.2 Frequency-dependent dielectric constant

In former work [11], it was shown that the wave-number-dependent, static dielectric permittivity of the SQE model on the simple-cubic lattice is given by

ε~rSI(𝐤)=1+1ε0a0(κb+κak2a02)superscriptsubscript~𝜀rSI𝐤11subscript𝜀0subscript𝑎0subscript𝜅bsubscript𝜅asuperscript𝑘2superscriptsubscript𝑎02\tilde{\varepsilon}_{\textrm{r}}^{\text{SI}}(\mathbf{k})=1+\frac{1}{% \varepsilon_{0}a_{0}(\kappa_{\textrm{b}}+\kappa_{\textrm{a}}k^{2}a_{0}^{2})}over~ start_ARG italic_ε end_ARG start_POSTSUBSCRIPT r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT SI end_POSTSUPERSCRIPT ( bold_k ) = 1 + divide start_ARG 1 end_ARG start_ARG italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_κ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT + italic_κ start_POSTSUBSCRIPT a end_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG (6)

in the SI unit system. Here, κbsubscript𝜅b\kappa_{\textrm{b}}italic_κ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT is the split-charge or bond (b) hardness, which replaces the nearest-neighbor κijsubscript𝜅𝑖𝑗\kappa_{ij}italic_κ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT. Since the diverging next-nearest κijsubscript𝜅𝑖𝑗\kappa_{ij}italic_κ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT lead to (infinitesimally) small split-charges and bond-charge energies, their effect can be ignored. The validity of Eq. (6) hinges on various approximations, such as undamped Coulomb interactions and environment-independent κasubscript𝜅a\kappa_{\text{a}}italic_κ start_POSTSUBSCRIPT a end_POSTSUBSCRIPT, which, however can be accounted for in principle. In fact, simple-cubic-lattice specific discretization corrections to the continuum Coulomb (C) interaction, J~C(𝐤)1/k2proportional-tosubscript~𝐽C𝐤1superscript𝑘2\tilde{J}_{\textrm{C}}(\mathbf{k})\propto 1/k^{2}over~ start_ARG italic_J end_ARG start_POSTSUBSCRIPT C end_POSTSUBSCRIPT ( bold_k ) ∝ 1 / italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, have proved important to reproduce the Thomas-Fermi screening length ζ𝜁\zetaitalic_ζ [11]. It is also noted that κbsubscript𝜅b\kappa_{\textrm{b}}italic_κ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT is supposedly the most difficult but also the most important coefficient to be made environment dependent, as it will depend not only on the distance but also on local coordination numbers or bond order.

We revert to the unit system of this study by replacing ε0subscript𝜀0\varepsilon_{0}italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT with 1/4π14𝜋1/4\pi1 / 4 italic_π. Since all equations are linear, the static bond hardness κb(ω=0)subscript𝜅b𝜔0\kappa_{\textrm{b}}(\omega=0)italic_κ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT ( italic_ω = 0 ) only needs to be replaced with κb(ω)=ω2LiωR+κb(0)subscript𝜅b𝜔superscript𝜔2𝐿i𝜔𝑅subscript𝜅b0\kappa_{\textrm{b}}(\omega)=-\omega^{2}L-\text{i}\omega R+\kappa_{\textrm{b}}(0)italic_κ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT ( italic_ω ) = - italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_L - i italic_ω italic_R + italic_κ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT ( 0 ) to yield

ε~r(𝐤,ω)=1+4π/a0Lω2iRω+κb+κak2a2.subscript~𝜀r𝐤𝜔14𝜋subscript𝑎0𝐿superscript𝜔2i𝑅𝜔subscript𝜅bsubscript𝜅asuperscript𝑘2superscript𝑎2\tilde{\varepsilon}_{\textrm{r}}(\mathbf{k},\omega)=1+\frac{4\pi/a_{0}}{-L% \omega^{2}-\text{i}R\omega+\kappa_{\text{b}}+\kappa_{\text{a}}k^{2}a^{2}}.over~ start_ARG italic_ε end_ARG start_POSTSUBSCRIPT r end_POSTSUBSCRIPT ( bold_k , italic_ω ) = 1 + divide start_ARG 4 italic_π / italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG - italic_L italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - i italic_R italic_ω + italic_κ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT + italic_κ start_POSTSUBSCRIPT a end_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (7)

For a loss-free metal, R=κb=0𝑅subscript𝜅b0R=\kappa_{\text{b}}=0italic_R = italic_κ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT = 0, it follows that the dielectric permittivity obeys ϵr=1ωp2/ω2subscriptitalic-ϵr1subscriptsuperscript𝜔2psuperscript𝜔2\epsilon_{\textrm{r}}=1-\omega^{2}_{\textrm{p}}/\omega^{2}italic_ϵ start_POSTSUBSCRIPT r end_POSTSUBSCRIPT = 1 - italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT p end_POSTSUBSCRIPT / italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT at scales large compared to atomic spacings (k0𝑘0k\to 0italic_k → 0), where ωp=4π/(a0L)subscript𝜔p4𝜋subscript𝑎0𝐿\omega_{\textrm{p}}=4\pi/(a_{0}L)italic_ω start_POSTSUBSCRIPT p end_POSTSUBSCRIPT = 4 italic_π / ( italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_L ) is the plasma frequency. Thus, when used to model the dielectric properties of a metal (κb=0subscript𝜅b0\kappa_{\text{b}}=0italic_κ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT = 0) using a simple-cubic discretization, L𝐿Litalic_L can be parameterized to reproduce the plasma frequency.

A second resonance frequency could be realized, for example, by introducing next-nearest split charges. Another generalization would be to account for atomic polarizability. In other words, the D-SQE formalism allows for a quite flexible parameterization of ε0~(𝐤,ω)~subscript𝜀0𝐤𝜔\tilde{\varepsilon_{0}}(\mathbf{k},\omega)over~ start_ARG italic_ε start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ( bold_k , italic_ω ). However, the flexibility already contained in Eq. (7) should suffice for most practical applications.

To overcome the limitation to sc structures, the derivation of Eq. (6) must be repeated for other (Bravais) lattices. This leads to a rescaling of the parameters L𝐿Litalic_L, R𝑅Ritalic_R, κssubscript𝜅s\kappa_{\textrm{s}}italic_κ start_POSTSUBSCRIPT s end_POSTSUBSCRIPT, and κasubscript𝜅a\kappa_{\text{a}}italic_κ start_POSTSUBSCRIPT a end_POSTSUBSCRIPT in Eq. (7) with 2D/Z2𝐷𝑍2D/Z2 italic_D / italic_Z and/or with a0D/Vecsuperscriptsubscript𝑎0𝐷subscript𝑉eca_{0}^{D}/V_{\text{ec}}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_D end_POSTSUPERSCRIPT / italic_V start_POSTSUBSCRIPT ec end_POSTSUBSCRIPT, where Z𝑍Zitalic_Z is the coordination number, D𝐷Ditalic_D the spatial dimension, and Vecsubscript𝑉ecV_{\textrm{ec}}italic_V start_POSTSUBSCRIPT ec end_POSTSUBSCRIPT the volume of the elementary cell. However, for Eq. (7) to be accurate damping and continuum corrections may have to be applied. Exploring this in more detail is beyond the scope of the present feasibility test with its focuses on a generic demonstrator, which is discussed next.

II.3 Demonstrator model

The D-SQE model’s ability to capture charge-transfer dynamics is examined using a generic, all-atom resistor-capacitor (RC) setup. The RC element is coupled to an external voltage, thermal fluctuations causing Nyquist noise, or a fixed charge passing through the capacitor, as shown in Fig. 1.

Refer to caption
Refer to caption
Refer to caption
Figure 1: Set-up of the model in three modes of operation. Spheres represent atoms, with colors indicating charges. Split charges connect adjacent atoms and wire ends.

All atoms and bonds of the RC element are treated equally. Parameters are given without units; however, their values are inspired by those describing polonium. This choice is not significant, as the unit system could be based on mesoscopic rather than atomistic discretization elements. Specifically, the nearest-neighbor spacing is a0=1subscript𝑎01a_{0}=1italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1, while the atomic hardnesses is set to κa=2.4subscript𝜅a2.4\kappa_{\textrm{a}}=2.4italic_κ start_POSTSUBSCRIPT a end_POSTSUBSCRIPT = 2.4. Atoms have a zero electronegativity, except for the two terminal atoms, for which we set χ=±V/2𝜒plus-or-minus𝑉2\chi=\pm V/2italic_χ = ± italic_V / 2 when applying a voltage V𝑉Vitalic_V. All split-charges are given a unity pseudo inductance, for the author’s inability to find a literature value for polonium’s plasma frequency. Any split-charge resistance is set to Rsq=0.1245subscript𝑅sq0.1245R_{\text{sq}}=0.1245italic_R start_POSTSUBSCRIPT sq end_POSTSUBSCRIPT = 0.1245.

A capacitor plate has a target radius of r15a0𝑟15subscript𝑎0r\approx 15a_{0}italic_r ≈ 15 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT resulting in Np=717subscript𝑁p717N_{\text{p}}=717italic_N start_POSTSUBSCRIPT p end_POSTSUBSCRIPT = 717 atoms per plate. The two plates are separated by a distance of d=3a0𝑑3subscript𝑎0d=3a_{0}italic_d = 3 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, yielding a continuum-theory capacitance of C=Npa02/(4πd)19𝐶subscript𝑁psuperscriptsubscript𝑎024𝜋𝑑19C=N_{\text{p}}a_{0}^{2}/(4\pi d)\to 19italic_C = italic_N start_POSTSUBSCRIPT p end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 4 italic_π italic_d ) → 19. A better value for C𝐶Citalic_C is obtained by equally distributing a total unit charge of opposite sign on the two plates and computing the energy, which gives C=26.4𝐶26.4C=26.4italic_C = 26.4. Further corrections have opposite signs. If factors related to atomic hardness were included, the estimated capacitances would decrease, while accounting for the flexibility of some (excess) charge densities on the wires and the plates’ rim would have the opposite effect.

Each wire strand consists of 39 atoms leading to a total external resistance of R=(2×39+1)Rsq𝑅2391subscript𝑅sqR=(2\times 39+1)R_{\text{sq}}italic_R = ( 2 × 39 + 1 ) italic_R start_POSTSUBSCRIPT sq end_POSTSUBSCRIPT given the extra split charge between wire terminals. However, Ohmic dissipation also occurs in the capacitance, even if current flows in a plane rather than through a wire. In order to reduce the relative importance of that contribution to the overall resistance, split charges were also placed between next-nearest neighbors. Note that this eliminates (a little more than) one split-charge resistance at each of the four wire corners as well as that of the split-charge connecting the wire ends to the plates. An improved guess for the total resistance thus is R=73Rsq9.1𝑅73subscript𝑅sq9.1R=73~{}R_{\textrm{sq}}\to 9.1italic_R = 73 italic_R start_POSTSUBSCRIPT sq end_POSTSUBSCRIPT → 9.1.

A few words on the split charge between the wire ends may be appropriate. As long as no voltage is applied between these two ’atoms,’ their bond is simply longer but has the same resistance as any other bond. Turning the wire ends into (daemon) battery terminals requires adding the voltage V𝑉Vitalic_V to the battery split charge, which, however, now represents ionic rather than electronic charge donation. (This added voltage is the critical factor rather than assigning electronegativities of ±V/2plus-or-minus𝑉2\pm V/2± italic_V / 2 to the two terminal atoms.) While it would have been physically meaningful to assign a different, potentially much larger value to the battery’s internal resistance than to regular bonds, the original value was kept primarily so that the circuit’s resistance did not depend on the specific mode of operation. Similar comments applies to the hardness of the terminal atoms.

Before starting the charging simulation, the system is equilibrated with an open switch, preventing (ionic) charge transfer between the battery terminals at ’negative times.’ This results in a marginal charge transfer between each electrode and the half RC element it is connected to.

The Python code developped to simulate the demonstrator model is avaiblable under https://github.com/mueser/D-SQE. The charging simulations of 200 000 time steps took about 92 s on a two-year old MacBook Pro when atoms with a nearest-neighbor distance were connected through a total of Nsq=2 825subscript𝑁sq2825N_{\text{sq}}=2\,825italic_N start_POSTSUBSCRIPT sq end_POSTSUBSCRIPT = 2 825 split charges. These numbers changed to 108 s and Nsq=5,533subscript𝑁sq5533N_{\text{sq}}=5{,}533italic_N start_POSTSUBSCRIPT sq end_POSTSUBSCRIPT = 5 , 533 when also next-nearest neighbors were connected. The 15% increase of total computing time, after Nsqsubscript𝑁sqN_{\text{sq}}italic_N start_POSTSUBSCRIPT sq end_POSTSUBSCRIPT was essentially doubled, implies that the excess computing time related to split charges was merely about 30%. Thus, the advantage of avoiding explicit split charges while reproducing proper dissociation limits with a charge transfer model (QTM), such as the one proposed by Gergs [28], is primarily related to memory, not computation speed. Despite its elegance, it remains unclear how to incorporate a targeted dielectric response function into QTM.

III Results

Before formally discussing the results in this section, we encourage the reader to view the movies M1–M5 uploaded to the electronic supplement (and for the time before publication being available at https://github.com/mueser/D-SQE), as this may facilitate the understanding of the text. M1 illustrates the initial stages of the charging process of the RC element. M2 extends the observation over a longer time span but with lower temporal resolution, while M3 captures the oscillatory charging dynamics when the split-charge resistivity is set to zero. M4 highlights the dynamics under thermal noise. Finally, M 5 shows the effects in the presence of a fixed charge passing through the capacitor.

The charging of the capacitor at zero temperature is addressed first. As a reminder, the ordinary differential equation describing the charging procees of a macroscopic RC element coupled to a direct current battery is

RQ˙+1CQ=VΘ(tt0).𝑅˙𝑄1𝐶𝑄𝑉Θ𝑡subscript𝑡0R\dot{Q}+\frac{1}{C}Q=V\Theta(t-t_{0}).italic_R over˙ start_ARG italic_Q end_ARG + divide start_ARG 1 end_ARG start_ARG italic_C end_ARG italic_Q = italic_V roman_Θ ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) . (8)

Its solution for the initial condition Q(t=t0)=0𝑄𝑡subscript𝑡00Q(t=t_{0})=0italic_Q ( italic_t = italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) = 0 is

Q(t)=CV{1exp((tt0)/τ)}Θ(tt0),𝑄𝑡𝐶𝑉1𝑡subscript𝑡0𝜏Θ𝑡subscript𝑡0Q(t)=CV\{1-\exp(-(t-t_{0})/\tau)\}\Theta(t-t_{0}),italic_Q ( italic_t ) = italic_C italic_V { 1 - roman_exp ( - ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / italic_τ ) } roman_Θ ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) , (9)

with a relaxation time of τ=RC𝜏𝑅𝐶\tau=RCitalic_τ = italic_R italic_C. Fig. 2 shows that this equation fits the data quite well. Moreover, the asymptotic charge of Q=27.36𝑄27.36Q=27.36italic_Q = 27.36 is surprisingly close to the value of Q=26.4𝑄26.4Q=26.4italic_Q = 26.4, which was ’predicted’ in Sect. II.3. Since the voltage is unity, capacitance and equilibrium charge assume identical numerical values. From the relaxation time τ=248.7𝜏248.7\tau=248.7italic_τ = 248.7, an effective resistance of R=τ/C9.09𝑅𝜏𝐶9.09R=\tau/C\to 9.09italic_R = italic_τ / italic_C → 9.09 can be deduced. This corresponds almost exactly to the simple estimate of R=9.1𝑅9.1R=9.1italic_R = 9.1, which was also made in Sect. II.3. Such a close agreement must benefit to some degree from fortuitous error cancelation: The resistance of each corner was slightly over- and that of the capacitor underesimtated, each time by an amount, which certainly exceeds the 0.1% deviation between the simulation result and the back-of-the-enveloppe estimate. Here, and in the following, we would deem deviations between simulation results and expectations as acceptable, whose order does not exceed a0/r0.067subscript𝑎0𝑟0.067a_{0}/r\to 0.067italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_r → 0.067, since this ratio gives the order of magnitude of either the number of wire or rim atoms relative to the number of atoms per plate.

Refer to caption
Figure 2: Time dependence of the charge Q𝑄Qitalic_Q as obtained in the D-SQE model (symbols) and according to the solution of Q(t)𝑄𝑡Q(t)italic_Q ( italic_t ) (lines), see Eq. (9). The inset highlights the early-time behavior.

Two further observations are worth discussing. First, the electrical circuit is completed at time t=0𝑡0t=0italic_t = 0. However, the charge on the capacitor’s plates starts increasing noticeably only at t011.0subscript𝑡011.0t_{0}\approx 11.0italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 11.0. This delay results from the time it takes for the information to travel from the ’switch,’ in this case the battery, to the front atom of the capacitance. Since the calculations for the dielectric permittivity does not transfer directly to the one-dimensional wire geometry, we do not know the precise wave speed c𝑐citalic_c with which the split-charge travels through the wire. However, it is plausible that it is of the order of κaa/L1.55subscript𝜅a𝑎𝐿1.55\sqrt{\kappa_{\textrm{a}}a/L}\to 1.55square-root start_ARG italic_κ start_POSTSUBSCRIPT a end_POSTSUBSCRIPT italic_a / italic_L end_ARG → 1.55.

A second interesting observation is that the plate connected to the cathode is marginally positive before the circuit is completed. This occurs because the cathode has a higher electronegativity, attracting ’electrons’ to it even before charge-neutralizing ’ions’ can flow between the two terminal atoms. Once the switch is closed, charge compensation occurs. Current flows until the chemical potential of the atomic charges on the two opposing plates differs by the voltage applied to the wire/battery terminals.

The D-SQE model no longer dissipates energy when the split-charge resistance Rsqsubscript𝑅sqR_{\text{sq}}italic_R start_POSTSUBSCRIPT sq end_POSTSUBSCRIPT is set to zero. In this case, the model could be called the SQE-ω𝜔\omegaitalic_ω method, in analogy to the so-called ACKS2ω𝜔\omegaitalic_ω model [22], which motivates an on-site frequency dependent SQE model from the bottom up rather than in a top-down fashion. (To complete the isomorphism to the ACKS2 description, point dipoles and potentially higher-order multipoles would have to be added to the SQE description, while the resistance would have to be set to zero.) Omitting the resistance changes the nature of the circuit from an RC element to an LC element, where L stands for inductor. In the given case, its inductance L𝐿Litalic_L would be a collective pseudo inductance, which has a similar effect on the dynamics as a magentic inductance. Its value would be close to the sum of the inertia of individual split charges. The corresponding oscillations of the LC element are depicted in Fig. 3 for the same initial condition as for the RC circuit. Small deviations from a single-sinusoidal oscillations are noticeable. They are due to a coupling of the capacitor’s (symmetrized) charge to eigenmodes other than the slowest one, even if the slowest eigenmode clearly dominates the signal for the given initial boundary condition.

Refer to caption
Figure 3: Time dependence of the charge Q𝑄Qitalic_Q as obtained in the SQE-ω𝜔\omegaitalic_ω model (symbols) and according to the formal solution Q(t)=Q0{1cos(ωt)}𝑄𝑡subscript𝑄01𝜔𝑡Q(t)=Q_{0}\{1-\cos(\omega t)\}italic_Q ( italic_t ) = italic_Q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT { 1 - roman_cos ( italic_ω italic_t ) }.

The Nyquist noise was also investigated. To this end, the same setup was used as before, except that both wire end atoms were assigned the same chemical potential as the other atoms, and no voltage was applied. The measured observable is the time auto-correlation function (ACF) of the capacitor’s charge Q𝑄Qitalic_Q, from which the noise spectrum can be deduced via a Fourier transform:

CQQ(Δt)=Q(t)Q(t+Δt).subscript𝐶𝑄𝑄Δ𝑡delimited-⟨⟩𝑄𝑡𝑄𝑡Δ𝑡C_{QQ}(\Delta t)=\langle Q(t)Q(t+\Delta t)\rangle.italic_C start_POSTSUBSCRIPT italic_Q italic_Q end_POSTSUBSCRIPT ( roman_Δ italic_t ) = ⟨ italic_Q ( italic_t ) italic_Q ( italic_t + roman_Δ italic_t ) ⟩ . (10)

This ACF is not yet fully defined, because the choice of charge is not unique. It can be that of the capacitor plates connected to the anode Qasubscript𝑄aQ_{\textrm{a}}italic_Q start_POSTSUBSCRIPT a end_POSTSUBSCRIPT or to the cathode Qcsubscript𝑄cQ_{\textrm{c}}italic_Q start_POSTSUBSCRIPT c end_POSTSUBSCRIPT or the symmetrized charge Qs=(QaQc)/2subscript𝑄ssubscript𝑄asubscript𝑄c2Q_{\text{s}}=(Q_{\text{a}}-Q_{\text{c}})/2italic_Q start_POSTSUBSCRIPT s end_POSTSUBSCRIPT = ( italic_Q start_POSTSUBSCRIPT a end_POSTSUBSCRIPT - italic_Q start_POSTSUBSCRIPT c end_POSTSUBSCRIPT ) / 2. In a macroscopic capacitor, all three correlation functions would yield indistinguishable results, while for a finite system the one based on Qssubscript𝑄sQ_{\textrm{s}}italic_Q start_POSTSUBSCRIPT s end_POSTSUBSCRIPT yields a different result than the other two. Those of the individual plates are indentical in our system, because the entire set-up is ideally symmetric. Of course, the ACF of the symmetrized charge must be used to deduce the (quasi-) static capacitance of our circuit via the equipartition theorem

C=CQQ(Δt=0)kBT.𝐶subscript𝐶𝑄𝑄Δ𝑡0subscript𝑘𝐵𝑇C=\frac{C_{QQ}(\Delta t=0)}{k_{B}T}.italic_C = divide start_ARG italic_C start_POSTSUBSCRIPT italic_Q italic_Q end_POSTSUBSCRIPT ( roman_Δ italic_t = 0 ) end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG . (11)

Fig. 4 shows results deduced from a run with 100,000 time steps for equilibration and 64 million time steps, Δt=0.1Δ𝑡0.1\Delta t=0.1roman_Δ italic_t = 0.1, for observation using the D-SQE method. The fit to the ACF of the symmetrized charge on 30<t<50030𝑡50030<t<50030 < italic_t < 500 yields a value of C(0)=0.1702𝐶00.1702C(0)=0.1702italic_C ( 0 ) = 0.1702 and a relaxation time of τ=242𝜏242\tau=242italic_τ = 242. CQQ(0)subscript𝐶𝑄𝑄0C_{QQ}(0)italic_C start_POSTSUBSCRIPT italic_Q italic_Q end_POSTSUBSCRIPT ( 0 ) can be converted in a capacitance of C=CQQ(0)/kBT28.37𝐶subscript𝐶𝑄𝑄0subscript𝑘𝐵𝑇28.37C=C_{QQ}(0)/k_{B}T\to 28.37italic_C = italic_C start_POSTSUBSCRIPT italic_Q italic_Q end_POSTSUBSCRIPT ( 0 ) / italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T → 28.37. This is within 4% of the directly measured value. The correlation time of τ=242𝜏242\tau=242italic_τ = 242 is even within 2.5% of the relaxation time deduced earlier. Thus, the thermal voltage of the 5,533 simulated split charges adds up to that of an individual large resistor. Given the small deviation between ’continuum theory’ and simulations, one could conclude that the continuum limit of text-book RC circuits would be an excellent description of a 1,554-atom containing circuit, if electrons only behaved like a Fermi liquid in atom-thin wires.

Refer to caption
Figure 4: Charge auto-correlation function (ACF), see Eq. (10), for the charge on one plate (anode), Qasubscript𝑄aQ_{\text{a}}italic_Q start_POSTSUBSCRIPT a end_POSTSUBSCRIPT, and the symmetrized (symm.) charge (QaQc)/2subscript𝑄asubscript𝑄c2(Q_{\text{a}}-Q_{\text{c}})/2( italic_Q start_POSTSUBSCRIPT a end_POSTSUBSCRIPT - italic_Q start_POSTSUBSCRIPT c end_POSTSUBSCRIPT ) / 2, at a temperature of T=0.006𝑇0.006T=0.006italic_T = 0.006. The dashed line is a fit to the data at 20<t<50020𝑡50020<t<50020 < italic_t < 500 to the ACF of the symmetrized charge with an exponential function.

While the D-SQE model could be used to model response functions of regular dielectrics and metals, it is primarily meant as a tool to properly capture the dynamics of ions moving past polarizable but dissipative media. To show the feasibility of this application, the force acting on a unit charge moving at a constant velocity of v=0.01𝑣0.01v=0.01italic_v = 0.01 is computed and shown in Fig. 5. The motion of the unit charge is confined to the radial direction of the capacitor as shown in the bottom panel of Fig. 1, specifically midway between the two planes and parallel to the arrow. The point charge experiences a large attractive force shortly before entering the capacitor, a rather small force while being between the plates and again an attractive force shortly after exiting. In the capacitor, atomic discreteness of the two (commensurate) plates causes oscillations of the force with the period of the lattice. These oscillations are rather minor in particular in light of the distance between the charge and either plate being merely 3/2 times the lattice constant.

Refer to caption
Figure 5: Main graph: Force on a unit charge moving at velocity v=0.01𝑣0.01v=0.01italic_v = 0.01 along the symmetry axis of a plane-parallel capacitor with a radius of 15a015subscript𝑎015~{}a_{0}15 italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Inset: Damping force (scaled by a factor of 1,000) obtained by symmetrizing the force Fd=(F(x)F(x))/2subscript𝐹d𝐹𝑥𝐹𝑥2F_{\text{d}}=-(F(x)-F(-x))/2italic_F start_POSTSUBSCRIPT d end_POSTSUBSCRIPT = - ( italic_F ( italic_x ) - italic_F ( - italic_x ) ) / 2.

Zooming into the main panel of Fig. 5 near x=0𝑥0x=0italic_x = 0 reveals a small deviation from a (pseudo-) steady state. This is mainly due to the interaction of the moving ion with the two front atoms of the wire being connected to capacitors. The effect is clearly visible in movie M3, when the front atom of the upper wire briefly becomes blue when the patch indicating induced charge density is centered near the wire.

Any conservative force between charge and capacitor must obey F(x)=F(x)𝐹𝑥𝐹𝑥F(x)=-F(-x)italic_F ( italic_x ) = - italic_F ( - italic_x ) due to the symmetry of the set-up. However, the motion of the charge causes a dynamical response in the capacitor plates, leading to dissipation – to some minor degree even without explicit Ohmic damping. In the split-charge model, this would be because the split-charges in the wire-capacitor system will be in motion after a charge has traveled through them, in close analogy to the damping caused by atom/solid-surface scattering [29]. The kinetic and potential energy contained in this motion will have to be provided by the (force acting on the) moving charge.

The (frictional) work done on the system when moving between two symmetry-related points is x0x0dxF(x)superscriptsubscriptsubscript𝑥0subscript𝑥0d𝑥𝐹𝑥\int_{-x_{0}}^{x_{0}}\!\text{d}xF(x)∫ start_POSTSUBSCRIPT - italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT d italic_x italic_F ( italic_x ) so that the symmetrized force Fd(x)=(F(x)+F(x))/2subscript𝐹d𝑥𝐹𝑥𝐹𝑥2F_{\textrm{d}}(x)=-(F(x)+F(-x))/2italic_F start_POSTSUBSCRIPT d end_POSTSUBSCRIPT ( italic_x ) = - ( italic_F ( italic_x ) + italic_F ( - italic_x ) ) / 2 can be interpreted as an effective, instantaneous damping force as long as the moving charge’s velocity is constrained to be constant. It is shown in the inset of Fig. 5. Two contributions can be separated: One that is due to the charging / decharging of the capacitor upon approach and retraction of the charge and another one due to the dragging of mirror charges through the capacitance.

Although investigating the dissipation caused by the moving charge in detail would certainly be an interesting exercise, which might be done in some loose analogy to the modeling of viscous contact hysteresis, we restrict ourselves to a minimum analysis here with a single reference velocity. Reducing the velocity by a factor of two makes the damping Fd(x)subscript𝐹d𝑥F_{\textrm{d}}(x)italic_F start_POSTSUBSCRIPT d end_POSTSUBSCRIPT ( italic_x ) be almost exactly half of that of the reference calculation. If the resistivity is halfed, only the damping for |x|rless-than-or-similar-to𝑥𝑟|x|\lesssim r| italic_x | ≲ italic_r is halfed, while the hysteretic loss related to approach and retraction is reduced less. Thus, motion within the capacitor is impeded by Ohmic resistance, while the approach-retraction hysteresis has additional contributions.

Given the results on the Stokes friction, the damping force satisfies Fdmγvsubscript𝐹d𝑚𝛾𝑣F_{\text{d}}\approx m\gamma vitalic_F start_POSTSUBSCRIPT d end_POSTSUBSCRIPT ≈ italic_m italic_γ italic_v with mγ0.024.741017𝑚𝛾0.024.74superscript1017m\gamma\approx 0.02\to 4.74\cdot 10^{-17}italic_m italic_γ ≈ 0.02 → 4.74 ⋅ 10 start_POSTSUPERSCRIPT - 17 end_POSTSUPERSCRIPT N/s for the used parameterization and set-up. This would give a damping coefficient of γ4.11𝛾4.11\gamma\approx 4.11italic_γ ≈ 4.11 GHz in case of a lithium ion and thus a slip time of 0.243 ns. For comparison, the electronic damping coefficient of nitrogen sliding past a lead surface is merely 0.05 GHz [30].

IV Conclusions

In this work, dissipative dyanmics were added to the split-charge model in a phenomenological fashion. It allows the modeling of targeted dielectric properties in the framework of charge-equilibration processes by introducing Ohmic damping to the transfer of charge through a chemical bond. However, the approach is not restricted to the atomic scale. It can be defined in a way such that any linear-response, dielectric permittivity of a metal or a dielectric medium can be reproduced at coarse scales. The philosophy of the approach is similar to the Drude model. However, the D-SQE model reflects the charging of heterogeneous or contacting solids in a natural way, while the Drude model would have to be generalized to reflect local work functions. Once the bond hardness is finite, the model is close in spirit to the Lorentz model. However, the D-SQE model is meant to describe charge-transfer polarizability while the Lorentz model mimicks on-site polarizability. Of course, inducible point dipoles can be added to the model, as mentioned above. To alleviate computational burden, it may be advisable to use large pseudo-inductances, as this allows large time steps to be used. This would impair the accuracy of modeling conductivity near optical frequencies, while the Ohmic response, which is typically the most relevant for molecular dynamics applications, would remain unaffected.

One shortcoming of the current formulation is that the Ohmic resistance is added in an ad-hoc fashion. In reality, Ohmic resistance has several origins, one of which is the scattering of phonons and electrons [31]. Such scattering can also occur naturally in the pursued approach, e.g., when the pseudo-inductance is a time dependent quantity due to fluctuating bond lengths. Temporo-spatial variations of the pseudo-inductance densities would most naturally lead to the scattering of split-charge waves through ths system. As long as ions and split charges are treated classicaly and the temperature is below the Debye temperature, that scattering should be small compared to the real one. However, it might be possible to emulate the phonon-electron scattering after quantizating the relevant degrees of freedom, for example, through Feynman’s path-centroid density [32]. Such an approch has successfully reproduced damped dynamics of a commensurate, quantum mechanical Frenkel-Kontorova model, including the gap closure at sufficiently small mass [33].

Even without quantization, challenges remain for the D-SQE approach when used in regular potential or force-field based simulations, i.e., when the bonds related to a split charge have varying lengths and can break or form. Once a bond becomes long, κbsubscript𝜅b\kappa_{\textrm{b}}italic_κ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT should increase and the pseudo-inductance (roughly linearly proportional) with it. Such a scaling is not only physically meaningful, because electronic eigenfrequencies do not diverge when chemical bonds break, but it is also numerically desirable, as it does not require the time steps to be made extremely small. Nonetheless, the equation of motion would be a modified Langevin equation with an extra damping term proportional to m˙q˙˙𝑚˙𝑞\dot{m}\dot{q}over˙ start_ARG italic_m end_ARG over˙ start_ARG italic_q end_ARG. This might be a minor task compared to making force fields ’learn’ realistic values for pseudo inductances and bond polarizabilites from calculations as those presented by Cheng and Verstraelen [22]

References

  • Nalewajski [1984] R. F. Nalewajski, Electrostatic effects in interactions between hard (soft) acids and bases, J. Am. Chem. Soc. 106, 944–945 (1984).
  • Mortier et al. [1985] W. J. Mortier, K. Van Genechten, and J. Gasteiger, Electronegativity equalization: application and parametrization, J. Am. Chem. Soc. 107, 829 (1985).
  • Mortier et al. [1986] W. J. Mortier, S. K. Ghosh, and S. Shankar, Electronegativity-equalization method for the calculation of atomic charges in molecules, J. Am. Chem. Soc. 108, 4315–4320 (1986).
  • Rappe and Goddard [1991] A. K. Rappe and W. A. Goddard, Charge equilibration for molecular dynamics simulations, J. Phys. Chem. 95, 3358–3363 (1991).
  • Rick et al. [1994] S. W. Rick, S. J. Stuart, and B. J. Berne, Dynamical fluctuating charge force fields: Application to liquid water, J. Chem. Phys. 101, 6141–6156 (1994).
  • Verstraelen et al. [2013] T. Verstraelen, P. W. Ayers, V. Van Speybroeck, and M. Waroquier, Acks2: Atom-condensed Kohn-Sham DFT approximated to second order, J. Chem. Phys. 13810.1063/1.4791569 (2013).
  • Müser et al. [2022] M. H. Müser, S. V. Sukhomlinov, and L. Pastewka, Interatomic potentials: achievements and challenges, Adv. Phys.: X 810.1080/23746149.2022.2093129 (2022).
  • Jensen [2023] F. Jensen, Unifying charge-flow polarization models, J. Comput. Chem. 19, 4047–4073 (2023).
  • Streitz and Mintmire [1994] F. H. Streitz and J. W. Mintmire, Electrostatic potentials for metal-oxide surfaces and interfaces, Phys. Rev. B 50, 11996–12003 (1994).
  • Müser [2012] M. H. Müser, The chemical hardness of molecules and the band gap of solids within charge equilibration formalisms: Toward force field-based simulations of redox reactions, EPJ B 8510.1140/epjb/e2012-21081-8 (2012).
  • Nistor and Müser [2009] R. A. Nistor and M. H. Müser, Dielectric properties of solids in the regular and split-charge equilibration formalisms, Phys. Rev. B 7910.1103/physrevb.79.104303 (2009).
  • Scalfi et al. [2020] L. Scalfi, T. Dufils, K. G. Reeves, B. Rotenberg, and M. Salanne, A semiclassical Thomas–Fermi model to tune the metallicity of electrodes in molecular simulations, J. Chem. Phys. 15310.1063/5.0028232 (2020).
  • Bhattarai et al. [2019] H. Bhattarai, K. E. Newman, and J. D. Gezelter, Polarizable potentials for metals: The density readjusting embedded atom method (DR-EAM), Phys. Rev. B 9910.1103/physrevb.99.094106 (2019).
  • Nistor et al. [2006] R. A. Nistor, J. G. Polihronov, M. H. Müser, and N. J. Mosey, A generalization of the charge equilibration method for nonmetallic materials, J. Chem. Phys. 12510.1063/1.2346671 (2006).
  • Verstraelen et al. [2009] T. Verstraelen, V. V. Speybroeck, and M. Waroquier, The electronegativity equalization method and the split charge equilibration applied to organic systems: Parametrization, validation, and comparison, J. Chem. Phys. 131, 044127 (2009).
  • Chelli et al. [1999] R. Chelli, P. Procacci, R. Righini, and S. Califano, Electrical response in chemical potential equalization schemes, J. Chem. Phys. 111, 8569 (1999).
  • Hubbard [1963] J. Hubbard, Proc. R. Soc. Lond. A Math. Phys. Sci. 276, 238–257 (1963).
  • Dapp and Müser [2013] W. B. Dapp and M. H. Müser, Towards time-dependent, non-equilibrium charge-transfer force fields: Contact electrification and history-dependent dissociation limits, EPJ B 8610.1140/epjb/e2013-40047-x (2013).
  • Dapp and Müser [2013] W. B. Dapp and M. H. Müser, Redox reactions with empirical potentials: Atomistic battery discharge simulations, J. Chem. Phys. 13910.1063/1.4817772 (2013).
  • Lewis and Catlow [1985] G. V. Lewis and C. R. A. Catlow, Potential models for ionic oxides, J. Phys. C: Solid State Phys. 18, 1149–1161 (1985).
  • Cipcigan et al. [2019] F. Cipcigan, J. Crain, V. Sokhan, and G. Martyna, Electronic coarse graining: Predictive atomistic modeling of condensed matter, Rev. Mod. Phys. 91, 025003 (2019).
  • Cheng and Verstraelen [2022] Y. Cheng and T. Verstraelen, A new framework for frequency-dependent polarizable force fields, J. Chem. Phys. 15710.1063/5.0115151 (2022).
  • Haldane [1981] F. D. M. Haldane, “Luttinger liquid theory” of one-dimensional quantum fluids. i. properties of the Luttinger model and their extension to the general 1d interacting spinless fermi gas, J. Phys. C: Solid State Phys. 14, 2585–2609 (1981).
  • Ahrens-Iwers et al. [2022] L. J. V. Ahrens-Iwers, M. Janssen, S. R. Tee, and R. H. Meißner, Electrode: An electrochemistry package for atomistic simulations, J. Chem. Phys. 15710.1063/5.0099239 (2022).
  • Goloviznina et al. [2024] K. Goloviznina, J. Fleischhaker, T. Binninger, B. Rotenberg, H. Ers, V. Ivanistsev, R. Meissner, A. Serva, and M. Salanne, Accounting for the quantum capacitance of graphite in constant potential molecular dynamics simulations, Adv. Mater. 10.1002/adma.202405230 (2024).
  • Holoviak et al. [2024] S. Holoviak, I. Dabo, and S. Sinnott, Simulation of electrochemical oxidation in aqueous environments under applied voltage using classical molecular dynamics, J. Phys. Chem. A 128, 2236–2244 (2024).
  • Kubo [1966] R. Kubo, The fluctuation-dissipation theorem, Rep. Prog. Phys. 29, 255–284 (1966).
  • Gergs et al. [2021] T. Gergs, F. Schmidt, T. Mussenbrock, and J. Trieschmann, Generalized method for charge-transfer equilibration in reactive molecular dynamics, J. Comput. Chem. 17, 6691–6704 (2021).
  • Adelman and Doll [1976] S. A. Adelman and J. D. Doll, Generalized Langevin equation approach for atom/solid-surface scattering: General formulation for classical scattering off harmonic solids, J. Chem. Phys. 64, 2375 (1976).
  • Dayo et al. [1998] A. Dayo, W. Alnasrallah, and J. Krim, Superconductivity-dependent sliding friction, Phys. Rev. Lett. 80, 1690–1693 (1998).
  • Ashcroft and Mermin [1976] N. W. Ashcroft and N. D. Mermin, Solid State Physics (Holt, Rinehart and Winston, New York, 1976).
  • Cao and Voth [1994] J. Cao and G. A. Voth, The formulation of quantum statistical mechanics based on the feynman path centroid density. ii. dynamical properties, J. Chem. Phys. 100, 5106–5117 (1994).
  • Krajewski and Müser [2004] F. R. Krajewski and M. H. Müser, Quantum creep and quantum-creep transitions in 1d sine-Gordon chains, Phys. Rev. Lett. 9210.1103/physrevlett.92.030601 (2004).