Dataset Viewer
Auto-converted to Parquet
fn
string
text
string
doi
string
title
string
authors
string
__index_level_0__
int64
10.48550_arXiv.0709.0346
###### Abstract We analytically derive the lower bound of the total conformational energy of a protein structure by assuming that the total conformational energy is well approximated by the sum of sequence-dependent pairwise contact energies. The condition for the native structure achieving the lower bound leads to the contact energy matrix that is a scalar multiple of the native contact matrix, i.e., the so-called Go potential. We also derive spectral relations between contact matrix and energy matrix, and approximations related to one-dimensional protein structures. Implications for protein structure prediction are discussed. _Keywords_ protein structure prediction; spectral relations; one-dimensional structures pacs: 87.15.Cc, 87.15.-v, 87.14.Ee ## I Introduction Proteins' biological functions are made possible by their precise three-dimensional (3D) structures, and each 3D structure is determined by its amino acid sequence through the laws of thermodynamics. Therefore, predicting protein structures from their amino acid sequences is important not only for inferring proteins' biological functions, but also for understanding how 3D structures are encoded in such one-dimensional information as amino acid sequence. The problem of protein structure prediction is naturally cast as an optimization problem where a potential function is minimized. Given an appropriate potential function, conformational optimization should yield the native structure as the unique global minimum conformation of the potential function. Thus, the problem has been traditionally divided into two sub-problems: One is to establish an appropriate potential function, and the other is to develop the methods to efficiently search the vast conformational space of a protein. Among various forms of effective energy functions, statistical contact potentials have been widely used. In this Letter, we exclusively treat a class of such contact potentials, neglecting other contributions such as electrostatics and local interactions. Accordingly, a protein conformation is represented as a contact matrix in which the \((i,j)\) element is 1 if the residues \(i\) and \(j\) are in contact in space, otherwise it is 0. Although the contact matrix is a coarse-grained representation of protein conformation, it has been known that the contact matrix contains sufficient information to recover the three-dimensional (native) structure of proteins. It is noted that, for the lattice model of proteins, these representations of protein conformation and energy function are exact. ## II Theory ### Lower bound of contact energy Our fundamental assumption is that the conformational energy of a protein can be somehow expressed in terms of a contact matrix. Now let us assume that the total energy of a protein can be well approximated by the sum of pairwise contact energies between amino acid residues, and that each pairwise contact energy can be decomposed into a sequence-dependent term and a conformation-dependent term. The sequence-dependent term is expressed as a matrix \(\mathcal{E}(S)=(\mathcal{E}_{ij})\) which we call the contact energy matrix, or \(E\)-matrix for short. Each element \(\mathcal{E}_{ij}\) of the \(E\)-matrix represents the energy between the residues \(i\) and \(j\) when they are in contact. This form of the \(E\)-matrix is a very general one: Each element, \(\mathcal{E}_{ij}\), may depend on the entire sequence, \(S\), or it may depend only on the types of the interacting amino acid residues, \(i\) and \(j\), as in the conventional contact potentials. The conformation-dependent term is expressed as another matrix \(\Delta(C)=(\Delta_{ij})\) which we call the contact matrix, or \(C\)-matrix. Each element \(\Delta_{ij}\) of the \(C\)-matrix assumes a value of either 1 or 0, depending on the residues \(i\) and \(j\) are in contact or not, respectively. \[E(C,S) = \frac{1}{2}\sum_{i=1}^{N}\sum_{j=1}^{N}\mathcal{E}_{ij}(S)\Delta _{ij}(C) \tag{1}\] \[= \frac{1}{2}[\mathcal{E}(S),\Delta(C)] \tag{2}\] Based on this assumption, we derive the lower bound for the conformational energy and the conditions for the native structure and \(E\)-matrix toachieve the bound. The Frobenius inner product leads to the matrix \(l_{2}\) norm defined as, for a matrix \(M\), \(\|M\|\equiv[M,M]^{1/2}=(\sum_{i,j}M_{ij}^{2})^{1/2}\). \[\|\Delta(C)\|^{2}=2N_{c}(C) \tag{3}\] As for any inner products, the Frobenius inner product satisfies the Cauchy-Schwarz inequality (\(||[A,B]|\leq\|A\|\|B\|\)) from which we have \[[\mathcal{E},\Delta]\geq-\|\mathcal{E}\|\|\Delta\| \tag{4}\] where the equality holds if and only if \[\mathcal{E}=\varepsilon\Delta \tag{5}\] Although the inequality (Eq. 4) holds for any pair of matrices, we now regard it as the lower bound for conformational energy for a given \(E\)-matrix. For simplicity, we first consider the energy minimization problem for conformations with \(\|\Delta(C)\|\) fixed to the value of the native conformation. It is desirable for the native conformation to satisfy the lower bound and hence its condition Eq.. If the native conformation indeed satisfies the condition Eq., then the elements of the \(E\)-matrix is either \(0\) or \(\varepsilon\) so that only the contacts present in the native conformation are stabilizing. Thus, the native conformation satisfying Eq. is actually a GMEC among any conformations with arbitrary values of \(\|\Delta(C)\|\). An \(E\)-matrix that satisfies Eq. for the native \(C\)-matrix is a kind of the so-called Go potential which has been essential for studying the protein folding problem. At this point, it is still possible that the native structure is not the unique GMEC. For example, if a conformation contains all the native contacts together with some other contacts, this conformation has the same energy as the native conformation. In order for a native conformation to be the unique GMEC, it is required that the total number of contacts of the native conformation is larger than that of any other conformations that contain all the native contacts. From the relation Eq., maximizing the total number of contacts is equivalent to maximizing the norm of the \(C\)-matrix, which in turn implies the minimization of the right-hand side of Eq.. To summarize, for a given \(E\)-matrix, \(\mathcal{E}(S)\), of a protein, its native conformation, \(C_{n}\), achieves the lower bound in Eq. if and only if \(\mathcal{E}(S)=\varepsilon\Delta(C_{n})\) for some \(\varepsilon<0\), and such native structure is the unique GMEC if and only if \(\|\Delta(C_{n})\|\) is the maximum of all possible conformations that contain all the native contacts. Note that the former condition is a relation between \(E\)-matrix and \(C\)-matrix whereas the latter is a condition for a native structure to satisfy. The magnitude of \(\varepsilon\) is not specified here, but it should be determined by other factors such as the folding temperature. It should be noted that a native structure can be the unique GMEC without achieving the lower bound of Eq.. Such a case is made possible either by the limitation of the conformational space imposed by other steric factors such as chain connectivity or excluded volumes, or by inherent inconsistencies of the \(E\)-matrix so that no plausible conformations are allowed to satisfy the lower bounds. ### Spectral relations To examine more closely how the lower bound can be achieved, we next derive a more generous lower bound in a more restricted case. \[\Delta=\sum_{\alpha=1}^{N}\sigma_{\alpha}\mathbf{u}_{\alpha}\mathbf{v}_{\alpha }^{T} \tag{6}\] \(U=(\mathbf{u}_{1},\cdots,\mathbf{u}_{N})\) and \(V=(\mathbf{v}_{1},\cdots,\mathbf{v}_{N})\) are orthogonal matrices. The singular components are sorted in decreasing order of the singular values: \(\sigma_{1}\geq\cdots\geq\sigma_{N}(\geq 0)\). Since \(\Delta\) is real symmetric, the singular values are the absolute values of the eigenvalues of \(\Delta\), and the singular vectors are such that \(\mathbf{u}_{\alpha}=\pm\mathbf{v}_{\alpha}\) where the sign corresponds to that of the respective eigenvalue. \[\mathcal{E}=\sum_{\alpha=1}^{N}\tau_{\alpha}\mathbf{x}_{\alpha}\mathbf{y}_{ \alpha}^{T} \tag{7}\] Since \(\mathcal{E}\) is also real symmetric, the singular components have the same properties as the \(C\)-matrix \(\Delta\). \[(\mathbf{u}_{\alpha}^{T}\mathbf{x}_{\beta})(\mathbf{v}_{\alpha}^{T}\mathbf{y} _{\beta})=-\delta_{\alpha,\beta} \tag{9}\] We now regard this inequality as a lower bound for the conformational energy for a given \(E\)-matrix. For a fixed set of the singular values \(\sigma_{\alpha}\) (\(\alpha=1,\cdots,N\)), if and only if there exists such a conformation that satisfies the condition in Eq., then that conformation is the lowest possible energy conformation. Let \(\lambda_{\alpha}\) and \(\varepsilon_{\alpha}\) (\(\alpha=1,\cdots,N\)) be the eigenvalues of the \(C\)-matrix and \(E\)-matrix, respectively, sorted in the decreasing order of their absolute values. Then \(\sigma_{\alpha}=|\lambda_{\alpha}|\) and \(\tau_{\alpha}=|\varepsilon_{\alpha}|\) for \(\alpha=1,\cdots,N\), and \({\bf u}_{\alpha}\) and \({\bf x}_{\alpha}\) are the eigenvectors of the corresponding matrices. Thus, in terms of eigenvalues and eigenvectors, the lower bound in Eq. is equal to \(\sum_{\alpha}\lambda_{\alpha}\varepsilon_{\alpha}\) with \(\lambda_{\alpha}\varepsilon_{\alpha}\leq 0\) for \(\alpha=1,\cdots,N\). In addition to the condition Eq. for the lower bound of Eq., if \(\Delta\) and \({\cal E}\) are of the same rank, then the numbers of positive, negative, and zero eigenvalues of \(\Delta\) and \(-{\cal E}\) are the same and \({\bf u}_{\alpha}=\pm{\bf x}_{\alpha}\). \[{\cal E}=-S\Delta S^{T}, \tag{10}\] If the conformation that satisfy the condition Eq. is the native structure, the \(E\)-matrix is consistent in the sense that the contributions from all the eigencomponents are stabilizing the native structure (\(\lambda_{\alpha}\varepsilon_{\alpha}\leq 0\)). Since the matrix \(S\) is non-singular, we can "predict" the native structure from the \(E\) matrix as \(\Delta=-S^{-1}{\cal E}S^{-T}\) (if we can construct the appropriate matrix \(S\)). At this point, however, the native structure may not be the GMEC since other conformations with a different set of singular values may have lower energies. In order to compare the energies of conformations with different sets of singular values, we use another inequality: \[-\sum_{\alpha=1}^{N}\sigma_{\alpha}\tau_{\alpha}\geq-\|{\cal E}\|\|\Delta\| \tag{11}\] We note that, in terms of singular values, the matrix norms are expressed as \(\|\Delta\|=(\sum_{\alpha}\sigma_{\alpha}^{2})^{1/2}\) and \(\|{\cal E}\|=(\sum_{\alpha}\tau_{\alpha}^{2})^{1/2}\). Hence, it is clear that the equality in Eq. holds if and only if, in addition to the condition in Eq., there exists a scalar constant \(c\) such that \(\tau_{\alpha}=c\sigma_{\alpha}\) for all \(\alpha=1,\cdots,N\). These conditions are equivalent to Eq.. ### One-dimensional approximations To connect the present results with previous studies, we next introduce two approximations. First, we consider the case where the \(E\)-matrix is well approximated by its principal eigencomponent, that is, \({\cal E}\approx\varepsilon_{1}{\bf x}_{1}{\bf x}_{1}^{T}\). This approximation is motivated by the eigenvalue analysis of the Miyazawa-Jernigan (MJ) contact potential performed by Li et al., and has been employed by others. In this case, the lower bound Eq. is achieved if and only if \({\bf x}_{1}=\pm{\bf u}_{1}\) and \(\varepsilon_{1}\lambda_{1}<0\). This result was previously derived by Cao et al. who subsequently showed that the vector \({\bf x}_{1}\) constructed by using the components of the principal eigenvector of the MJ contact potential is indeed highly correlated with the principal eigenvector of the native contact matrices. Bastolla et al. obtained a similar result, but they also showed that taking the average of such \({\bf x}_{1}\) over evolutionarily related proteins greatly improved the correlation. Since the rank of the contact matrix is in general not 1, Eq. does not hold and the equality in Eq. cannot be satisfied. Consequently, there are attractive interactions between non-native contacts even when \({\bf x}_{1}={\bf u}_{1}\) holds exactly. Nevertheless, Porto et al. have demonstrated that the knowledge of \({\bf u}_{1}\) alone is practically sufficient for reconstructing the native contact matrix of small single-domain proteins. Therefore, construction of effective rank-1 \(E\)-matrices is of great interest. Based on the Porto et al.'s result, it is tempting to postulate that the satisfaction of the lower bound by a rank-1 \(E\)-matrix is sufficient for the native conformation to be the unique GMEC. At present, however, there is no clear connection between the present formulation (energy minimization) and the Porto et al.'s combinatorial algorithm. Another approximation is a kind of mean-field approximations in which the matrix element \({\cal E}_{ij}\) is replaced by its average over column \(\langle{\cal E}_{i\bullet}\rangle\equiv\sum_{j=1}^{N}{\cal E}_{ij}/N\). Let us define \({\bf e}=(\langle{\cal E}_{1\bullet}\rangle,\cdots,\langle{\cal E}_{N\bullet }\rangle)^{T}\) and \({\bf n}=(n_{1},\cdots,n_{N})^{T}\) where \(n_{i}\equiv\sum_{j=1}^{N}\Delta_{ij}\) is the contact number of the \(i\)-th residue. Then, we have the following approximation and the lower bound: \[E(C,S) \approx \frac{1}{2}{\bf e}^{T}{\bf n} \tag{12}\] \[\geq -\frac{1}{2}\|{\bf e}\|\|{\bf n}\| \tag{13}\] This lower bound condition is analogous to Eq., and can be regarded as another kind of the Go potential for one-dimensional protein structure. It has been suggested that contact number vector can significantly constrain the conformational space. Together with other one-dimensional structures, contact number vector is also used for recovering the native structures, and can be accurately predicted. It has been pointed out that the contact number vector is highly correlated with the principal eigenvector of the \(C\)-matrix, which suggests that this mean-field approximation is qualitatively similar to the principal eigenvector approximation introduced above. ## Discussion Using a more restricted, but conventional, form of the \(E\)-matrix where each element \({\cal E}_{ij}\) depends only on the types of \(i\)-th and \(j\)-th residues (e.g., the MJ potential),Vendruscolo et al. have shown that it is impossible for such \(E\)-matrices to stabilize all the native structures in a database. The conventional \(E\)-matrices such as those they studied do not take into account the sequence-dependence beyond a summation of the contributions from residue pairs. In the present study, we assumed a more general form for the \(E\)-matrix, allowing each element \(\mathcal{E}_{ij}\) to depend on the whole amino acid sequence. In practical situations of protein structure prediction, we want to optimize an energy function so that the native conformations of arbitrary proteins achieve the lower bound. Now let us impose this as a requisite for the \(E\)-matrix. Then, there should exist a function, namely \(\mathcal{E}\), that maps each amino acid sequence to the corresponding optimal \(E\)-matrix, that is, the Go potential. Thus, the problem of structure prediction becomes a trivial matter. Currently, most efforts for developing energy functions seem to be focused on accurate estimation of a fixed set of parameters for a given functional form. The present analysis suggests that inferring the function \(\mathcal{E}\) that can generate the Go-like \(E\)-matrices from amino acid sequences is essential if a contact potential is used. The lower bound inequality (Eq. 4) and its condition for the equality (Eq. 5) will serve as the guiding principle for inferring such a function. This approach to structure prediction is apparently similar to machine-learning approaches to contact matrix prediction. Although conventional machine-learning methods are not directly targeted at the optimization of the form of Eq., their prediction accuracy should be indicative of the possibility for identifying the function \(\mathcal{E}\). In the preceding paragraph, we have assumed the existence of the function \(\mathcal{E}\) to construct the optimal contact potential from a given amino acid sequence. What if, however, there is no such function? In fact, the limited success of current contact matrix prediction strongly suggests that this is more likely the case. Such a case implies either that there are proteins for which the lower bound energy cannot be achieved, or that the total energy cannot be sufficiently accurately approximated by Eq.. The former case indicates that some proteins are inherently frustrated, but to a good approximation such proteins should be rather exceptional for natural proteins. The latter case may indicate that multi-body contact interactions and/or other energy components than contact energies are more important. In summary, we have shown that the requirement for the native structure to achieve the lower bound naturally leads to the Go potential and the requirement for such a conformation to be the unique GMEC leads to the native conformation being the most compact one among those containing all the native contacts. These results suggest that protein structure prediction should be possible simply by constructing the optimal energy matrices or that the contact potential alone is not suitable for the problem. Although not yet definitive, the current state of contact prediction as well as recent studies on local interactions suggest that the latter may be the case. Nevertheless, the present results may be useful for evaluating the optimality of potential functions in either case.
10.48550/arXiv.0709.0346
On the optimal contact potential of proteins
Akira R. Kinjo, Sanzo Miyazawa
2,531
10.48550_arXiv.0906.5274
###### Abstract We report a numerical simulation of the rate of crystal nucleation of sodium chloride from its melt at moderate supercooling. In this regime nucleation is too slow to be studied with "brute-force" Molecular Dynamics simulations. The melting temperature of ("Tosi-Fumi") NaCl is \(\sim 1060\)K. We studied crystal nucleation at \(T\)=800K and 825K. We observe that the critical nucleus formed during the nucleation process has the crystal structure of bulk NaCl. Interestingly, the critical nucleus is clearly faceted: the nuclei have a cubical shape. We have computed the crystal-nucleation rate using two completely different approaches, one based on an estimate of the rate of diffusive crossing of the nucleation barrier, the other based on the Forward Flux Sampling and Transition Interface Sampling (FFS-TIS) methods. We find that the two methods yield the same result to within an order of magnitude. However, when we compare the extrapolated simulation data with the only available experimental results for NaCl nucleation, we observe a discrepancy of nearly 5 orders of magnitude. We discuss the possible causes for this discrepancy. ## 1 Introduction Crystallization of salts is a phenomenon of great practical relevance. In fact, it is one of the most important industrial separation processes. But it also plays a crucial role in geological processes that occur on an altogether different time scale. The crystallization process consists of two steps: nucleation and growth. If nucleation is slow compared to the time it takes a crystal to grow to a size comparable to the size of the container, large single crystals will form (an example is rock salt). When nucleation is fast, the resulting solid will form as a fine powder. It is clearly important to be able to predict the rate of nucleation of salts and - at a later stage - to understand the factors that influence nucleation. In the present paper we aim to demonstrate that, with current simulation techniques and currently available force-fields, it is indeed possible to compute the rate of nucleation of a real salt crystal (in the present case NaCl from its melt) This opens the way to "ab-initio" predictions of nucleation rates of many ionic substances. Solutions or melts can often be cooled well below their freezing temperature. The reason is that the formation of small nuclei of the stable crystal phase is an activated process that may be extremely slow. Intuitively, it is easy to understand why crystal nucleation is an activated process, i.e. why there is a free-energy barrier separating the metastable parent phase (the liquid) from the stable crystal phase. The point is that, initially, the formation of small crystalline nuclei costs free energy. But once the crystal nucleus exceeds a critical size, its free energy decreases as it grows. The rate at which crystal nuclei form depends strongly on \(\Delta G_{crit}\), the free-energy required to form a critical nucleus. Classical Nucleation Theory (CNT) is commonly used to estimate the height of the nucleation barrier and to predict the rate of crystal nucleation. According to CNT, the total free energy of a crystallite that forms in a supersaturated solution or melt contains two terms: the first is a "bulk" term that expresses the fact that the solid is more stable than the supersaturated fluid - this term is negative and proportional to the volume of the crystallite. The second is a "surface" term that takes into account the free-energy cost of creating a solid-liquid interface. This term is positive and proportional to the surface area of the crystallite. \[\Delta G=\frac{4\pi}{3}R^{3}\rho_{S}\Delta\mu+4\pi R^{2}\gamma, \tag{1}\] The function \(\Delta G\) goes through a maximum at \(R=2\gamma/(\rho_{S}|\Delta\mu|)\) and the height of the nucleation barrier is \[\Delta G_{crit}=\frac{16\pi}{3}\gamma^{3}/(\rho_{S}|\Delta\mu|)^{2}. \tag{2}\] The crystal-nucleation rate per unit volume, \(I\), depends strongly on \(\Delta G_{crit}\): \[I=\kappa\exp(-\Delta G_{crit}/k_{B}T). \tag{3}\] Here \(\kappa\) is a kinetic prefactor, \(T\) is the absolute temperature and \(k_{B}\) is Boltzmann's constant. \[I=\kappa\,\exp\left[-\frac{16\pi}{3}\gamma^{3}/(\rho_{S}|\Delta\mu|)^{2}\right]. \tag{4}\] Under experimental conditions, nucleation is infrequent on the time scale of typical molecular processes. Yet, when it happens, it proceeds rapidly. This makes it difficult to study the structure and dynamics of crystal nuclei of atoms or small molecules in experiments. In the case of NaCl, the experiments are also complicated by the fact that crystallization occurs at high temperatures. This may explain why there is a scarcity of experimental data on the nucleation of NaCl. To our knowledge, the only data are those of Buckle and Ubbelhode from the 1960's. In these experiments, crystallization in NaCl micro-droplets was observed visually. As the droplet size ((\(\mathcal{O}(3\mu)\)) and time window for the measurement (1-30 seconds - after which the droplets sedimented out of view) - were fixed, the nucleation rate could be determined at one temperature only (905 K for NaCl). At this temperature, the nucleation rate was such that, on average, one nucleus would form during the observation time (\({\cal O}(10s)\)) in a droplet with a volume of order \(10^{-17}m^{3}\). Hence the experimental nucleation rate per unit volume was \(O(10^{16})\)\(m^{-3}s^{-1}\). For experimental nucleation rates of this order of magnitude, brute-force MD simulations are out of the question. The average time it would take for nuclei to form spontaneously in a system consisting of several thousands of particles is of order of \(10^{20}\) seconds. Clearly, this is beyond the scope of MD simulations. The standard solution to circumvent this problem is to perform simulations at much larger undercooling than used in the experiments. Huang et al. performed MD simulations of melting and freezing of a droplet composed of 216 NaCl ions in vacuum: to this end, they performed temperature quenches down to 550K (i.e. approximately half the melting temperature) and found nucleation rates of the order of \(O(10^{36})m^{-3}s^{-1}\), which is 20 orders of magnitude higher than the experimental rate at \(905K\). Another effort to study nucleation at less severe supercooling was made by Koishi et al.. These authors performed an MD simulation of 125000 ions system in vacuum, at temperatures of 740K (i.e. approximately 0.7 the melting temperature \(T_{m}\)) and 640K (i.e. approximately 0.6 the melting temperature \(T_{m}\)). Both free and periodic boundary conditions were used. The estimated nucleation rate at \(740K\) was \(O(10^{35})m^{-3}s^{-1}\), which is virtually the same value as found by Huang et al. at a much larger supercooling. This is surprising because nucleation rates tend to depend very strongly on temperature. This suggests that, at least at the lowest temperatures, the barrier for crystal nucleation is negligible. More in general, crystal nucleation under extreme supercooling need not proceed following the same path as under moderate supercooling. In what follows, we use the technique of refs. based on a combination of umbrella sampling (to determine the barrier height) and a dynamical simulation (to determine the crossing rate). The computing time required for this scheme does not scale exponentially with the nucleation barrier, but it does increase with increasing nucleus size. We also compute the nucleation rate using an algorithm based on the "forward flux sampling" (FFS) and the "Transition Path Samplig" (TIS) techniques and compare it with the one obtained using the method previously mentioned. In the present work we study homogeneous crystal nucleation in the Tosi-Fumi NaCl model for NaCl at two different temperatures, viz. \(T_{1}\)=800K and \(T_{2}\)=825K, corresponding to 25% and 22% undercooling. For this system, we computed the nucleation barrier, examined the structure and shape of the critical nucleus and computed the nucleation rate. ## 2 Methods The Tosi-Fumi rigid-ion interaction potential for NaCl is of the following form, \[U_{ij}(r)=A_{ij}e^{[B(\sigma_{ij}-r)]}-\frac{C_{ij}}{r^{6}}-\frac{D_{ij}}{r^{8}}+ \frac{q_{i}q_{j}}{r}\;, \tag{5}\] This pair potential is written as the sum of a Born-Mayer repulsion, two attractive van der Waals contributions and a Coulomb interaction term. In our simulations, we calculated the Coulomb interactions using the Ewald summations method with a real space cutoff of 10 \(\dot{A}\) and a real space damping parameter of 0.25 \(\dot{A}^{-1}\). We truncated the Van Der Waals part of the potential at 9 \(\dot{A}\), assuming the \(g(r)=1\) beyond this cutoff. The computed number density of ions in the bulk solid at 800K and 825K at \(10^{5}\) Pa was 0.041\(\dot{A}^{-3}\), in agreement with experiment. The density of ions in the supercooled liquid at the same temperature and pressure was 0.034\(\dot{A}^{-3}\). We prepared under cubic boundary conditions a supercooled system of\({}^{3}\) NaCl ion pairs at ambient pressure by cooling it down below the melting temperature. For the present model, Anwar et al. have computed the melting temperature: \(T_{m}\)=(1064 \(\pm\) 14)K, which is very close to the experimental melting temperature (\(T_{m}^{exp}\)=1072K). Using constant-pressure Monte Carlo simulations, we cooled the system down to the temperatures where we studied nucleation: \(T_{1}\)=800K and \(T_{2}\)=825K, corresponding to 25% and 22% undercooling. Note that the experiments on NaCl nucleation where performed at a somewhat higher temperature (16% supercooling). The reason why we could not perform simulations at these higher temperatures is that the critical nucleus would be about twice the size of the nucleus that could be studied without spurious finite-size effects for the system sizes that we employed. At temperatures below 750K, spontaneous nucleation occurred during the simulations. We therefore kept the temperature above this lower limit. Nucleation is an activated process. In steady-state, the nucleation rate per unit volume and time is given by eq. where \(\exp(-\beta\Delta G_{crit})\) is the equilibrium probability per nucleus to find a critical nucleus in the metastable parent phase. \(\kappa\) is a kinetic prefactor. In the case of a diffusive barrier crossing, \(\kappa\) can be expressed as: \[\kappa=\sqrt{\frac{|\Delta\mu|}{6\pi k_{B}Tn_{crit}}}\;\rho_{liq}f^{+}_{n^{crit}} \tag{6}\] \begin{table} \begin{tabular}{||c|c|c|c|c|c||} \hline & \(A_{ij}\) & \(B\) & \(C_{ij}\) & \(D_{ij}\) & \(\sigma_{ij}\) \\ & \([kJ/mol]\) & \([\dot{A}^{-1}]\) & \([\dot{A}^{6}kJ/mol]\) & \([\dot{A}^{8}kJ/mol]\) & \([\dot{A}]\) \\ \hline Na-Na & 25.4435 & 3.1546 & 101.1719 & 48.1771 & 2.340 \\ \hline Na-Cl & 20.3548 & 3.1546 & 674.4793 & 837.0770 & 2.755 \\ \hline Cl-Cl & 15.2661 & 3.1546 & 6985.6786 & 14031.5785 & 3.170 \\ \hline \end{tabular} \end{table} Table 1: Potential parameters for NaCl. It is important to distinguish between \(f^{+}_{ncrit}\), which is the rate at which particles are added to a nucleus with the critical size, and the net flux across the nucleation barrier. In steady state, this net flux is equal the number of nuclei that go from \(ncrit\) to \(ncrit+1\) minus the number that go from \(ncrit+1\) to \(ncrit\). Hence, the actual nucleation rate is a combination of the forward rate \(f^{+}\) and the backward rate \(f^{-}\). However, because of detailed balance, knowledge of \(f^{+}_{ncrit}\) (combined with knowledge of the barrier height and shape) is enough to compute the nucleation rate. Assuming a diffusive attachment or detachment of single particles from the critical nucleus, the forward rate \(f^{+}_{crit}\) at the top of the barrier can be related to the spontaneous fluctuations in the number of particles in a nucleus at the top of the nucleation barrier: \[f^{+}_{crit}=\frac{1}{2}\frac{\langle\Delta n^{2}_{crit}(t)\rangle}{t} \tag{7}\] To estimate \(f^{+}_{crit}\) a series of dynamical trajectories were necessary: after generating a set of uncorrelated configurations at the top of the barrier, we carried out NVT MD simulations using the \(DL\_POLY\) package with a timestep of 0.5 fs. We computed the nucleation rate using equation. Moreover, we computed the nucleation rate per unit time and volume using an algorithm based on path-sampling techniques and compared the results with those obtained using equation. ## 3 Results We computed the free energy barrier for crystal nucleation at \(T_{1}\)=800K and \(T_{2}\)=825K, corresponding to \(\beta\Delta\mu_{1}\)=0.54 and \(\beta\Delta\mu_{2}\)=0.48. The values of \(\beta\Delta\mu\) were estimated numerically by thermodynamic integration from the coexistence temperature and free energies reported by Anwar et al.. Fig.1 shows the computed nucleation barriers as a function of \(n\). As expected, \(\Delta G\) decreases with supersaturation. Around T\(<\)750K the barrier gets sufficiently low that spontaneous nucleation can take place on the time scale of a simulation. The size of the critical nucleus, \(N_{c}\), was estimated according to a fit of the functional form of the CNT, and we found \(N_{c}\approx 120\) ions at \(T_{1}\) and \(N_{c}\approx 150\) ions at \(T_{2}\). Koishi et al. estimated \(N_{c}=120-130\) ions at 640K and 740K, which is surprising in view of the CNT prediction that the size of the critical nucleus scales as \((\gamma/|\Delta\mu|)^{3}\). If we make the usual assumption that \(\gamma\sim\Delta h\) and \(\Delta\mu\approx(\Delta h/T_{m})(T-T_{m})\), where \(\Delta h\) is the entropy of fusion per ion pair, then we would expect that \(N_{c}\sim(1-T/T_{m})^{-3}\) and we would predict that the critical nucleus at 640K should contain a quarter as many particles as those at 800K. Next, we consider the structure and shape of the critical nucleus. Fig.2 shows a snapshot of the critical nucleus at \(T_{2}\)=825K. Note that the crystal presents rudimentary low-index facets. In experiments the existence of such facets was postulated, as they may act as sites for subsequent heterogeneous nucleation. As can be seen from Fig.2, the critical nucleus already shows the charge-ordered rock-salt structure of the bulk phase. Snapshot of the critical nucleus at \(T_{2}\)=825K: the bulk NaCl structure is already evident. The critical crystal nucleus seems to have rudimentary facets, which is in agreement with the interpretation of the experiments in ref.. Free energy barriers \(\Delta G\) as a function of the nucleus size \(n\) for \(T_{1}\)=800K(\(\beta\Delta\mu_{1}=0.54\)) and \(T_{2}\)=825K(\(\beta\Delta\mu_{2}=0.48\)). Error-bars on \(\beta\Delta G\) are of the order of 1 \(k_{B}T\). The dashed curves are fits to the functional form given by CNT. Fig.2 also shows that, in the temperature range that we studied, the critical nucleus is non-spherical. In order to quantify the degree of non-sphericity of the critical nucleus, we expanded its density with respect to the center of mass in rank-four-spherical-harmonics and constructed the quadratic invariant \(S_{4}\): we obtained \(S_{4}(T_{1})\)=0.115 and \(S_{4}(T_{2})\)=0.110. For a simple cube \(S_{4}\)=0.172 and for a sphere \(S_{4}\)=0, therefore the shape of the critical nucleus is closer to a cube than to a sphere. In other words, the critical nucleus already exhibits the morphology of macroscopic NaCl crystals. In order to make sure that there were not finite size effects that resulted in interactions between image cluster, we visually checked that the critical nuclei didn't show a preferred spacial orientation and that the minimum distance between them was bigger than half box. Moreover we calculated the Debye-Huckle screening lenght and found that it was smaller than 1 \(\dot{A}\), showing that the critical nuclei did not even electrostatically interact. We could therefore conclude that there was no induced nucleation due to the interaction between a critical cluster and its own periodic image. Using the computed height of the nucleation barrier and the values of \(\Delta\mu\) as input, we can estimate the surface free-energy density \(\gamma_{ls}\). To this end, we make use of the CNT expression for the barrier height (eqn. 2). However, this expression assumes that the critical nucleus is spherical. It is easy to derive the corresponding expression for a cubical nucleus. The results for both estimates of \(\gamma_{ls}\) are given in Table 2. There exist experimental estimates of \(\gamma_{ls}\) at 905K. These estimates are based on a somewhat questionable CNT expression for the nucleation rate. Moreover, in ref. it is assumed that the critical nucleus is spherical. The experimental estimate of \(\gamma_{ls}\) (\(\gamma_{exp}\)=84.1[\(erg\)\(cm^{-2}\)]) is therefore not based on a direct determination. Nevertheless, in the absence of other experimental data, this is the only number we can compare to. As the table shows, there is a fair agreement between simulation and experiment. The experimental estimate for \(\gamma\) is based on the assumption that the critical nucleus is spherical. If it is cubic, one would obtain the number in the lower right-hand corner. In view of the many uncertainties in the analysis of the experimental data, it is impossible to tell whether the discrepancy between simulation and experiment is significant. \begin{table} \begin{tabular}{||c|c|c||} \hline \hline ## T[K] & \(\gamma_{sphere}\) & \(\gamma_{cube}\) \\ \hline ## 800 & \(98\pm 2\) & \(80\pm 1\) \\ \hline ## 825 & \(99\pm 1\) & \(79\pm 1\) \\ \hline ## 905 & \(84.1^{(e)}\) & \(67.8^{(e)}\) \\ \hline \hline \end{tabular} \end{table} Table 2: Surface free energy density (in \(erg\)\(cm^{-2}\)) assuming spherical and cubical shape for critical nuclei. At 905K we report the experimental value. The entry in the lower right-hand corner is based on the experimental estimate, but assuming a cubical nucleus. Huang et al. estimated the solid-liquid surface free-energy density of NaCl from the nucleation rate at 550K. To achieve this, Huang et al. assumed that the CNT expression for the nucleation rate is valid. Under those assumptions, they obtained: \(\gamma\)=119.6[\(erg~{}cm^{-2}\)] for a spherical nucleus. In the present work, we can compute absolute nucleation rates without making use of CNT. The only assumption we make is that the barrier crossing is diffusive and that the Zeldovitch (pre)factor is well approximated by the form given in eqn.. The Zeldovitch factors were found to be respectively \(Z_{1}\)=0.016 for \(T_{1}\) and \(Z_{2}\)=0.013 for \(T_{2}\). The true Zeldovitch factor may be slightly different, but is in any event expected to be of \({\cal O}(10^{-2})\). From our MD simulations we obtained the following estimates for the forward rates (eq. 7) : \(f_{crit}^{+}\)=0.013 ps\({}^{-1}\) for \(T_{1}\) and \(f_{crit}^{+}\)=0.033 ps\({}^{-1}\) for \(T_{2}\). Combining this information, we can compute the kinetic prefactor of eq. 6: \(\kappa\)(\(T_{1}\))= 6.9\(\times\)10\({}^{36}\)\(m^{-3}s^{-1}\) and \(\kappa\)(\(T_{2}\))= 1.5\(\times\)10\({}^{37}\)\(m^{-3}s^{-1}\). As is to be expected, the kinetic prefactor depends only weakly on temperature. Using eqn. we then calculated the nucleation rate. The results are: \(I(T_{1})\)=3\(\times\)10\({}^{26\pm 1}\)\(m^{-3}s^{-1}\) and \(I(T_{2})\)=4\(\times\)10\({}^{24\pm 1}\)\(m^{-3}s^{-1}\). These nucleation rates are about ten orders of magnitude higher than the estimated experimental rate at 905K (\(O(10^{16})m^{-3}s^{-1}\)). Such a difference is hardly surprising because the nucleation rate is expected to increase rapidly with increasing supercooling. We also computed the nucleation rate using an algorithm based on the path-sampling techniques of refs. (Forward-Flux Sampling (FFS) and Transition-Interface Sampling (TIS)). The value obtained at \(T_{1}=800\)K is \(I_{FFS-TIS}(T_{1})=O(10^{27\pm 2})\)\(m^{-3}s^{-1}\), which agrees surprisingly well with the one obtained using the diffusive barrier-crossing approach. As the path-sampling method does not depend on the choice of the reaction coordinate and does not require prior knowledge of the phase space density, we can conclude that the method based on the free-energy calculation gives us a good estimate for the nucleation rate at this temperature. We have also computed the nucleation rate at T=750 K, using the FFS-TIS method. The computed nucleation rate is \(I_{FFS-TIS}(T=740K)\)=\(O(10^{35})\)\(m^{-3}s^{-1}\). This is of the same order of magnitude as the nucleation rate obtained calculated by Koishi et al. using "brute-force" MD. We cannot use the diffusive barrier crossing method at this temperature, as the barrier is too low to avoid spontaneous nucleation during long runs. However, if we assume that the kinetic pre-factor and the surface free-energy density do not vary much with temperature, we can use CNT to extrapolate the nucleation rate from \(800K\) to \(740K\). We find: \(I_{extrap}(T=740K)=O(10^{30})\)\(m^{-3}s^{-1}\) which is considerably lower than the results of the direct calculations. This suggests that an extrapolation procedure based on CNT is not reliable. A summary of our numerical results for the nucleation barriers and rates are given in Table 3. A similar problem occurs if we try to extrapolate our numerical data at \(800K\) and \(825K\) to \(905K\), the temperature of the experiments of ref.. If we can extrapolate our simulation results to the experimental temperature of 905 K (\(\beta\Delta\mu\)=0.3) we obtain an estimated nucleation at 905 K that is \(O(6\times 10^{11})m^{-3}s^{-1}\). This is some nearly five orders of magnitude less than the experimentally observed rate. The discrepancy between simulation and experiment can be due to several reasons. a) There might be an appreciable (but unspecified) error in the experimental estimates (e.g. due to residual heterogeneous nucleation). b) The estimated error in the computed melting temperature of the Tosi-Fumi model is \(\pm 20\)K. Such an uncertainty again easily translates into a variation of the nucleation rate by several orders of magnitude. c) In view of the extreme sensitivity of nucleation rates to the details of the intermolecular potential (see, e.g.), the Tosi-Fumi potential may be inadequate to model nucleation in NaCl, even though it can reproduce the static properties of the solid and liquid NaCl. d) Finally, it it is not quite correct to assume that the kinetic prefactor, the surface free energy and the latent heat of fusion are temperature-independent. We can also compare our calculated nucleation rate at \(800K\) to the rate estimated with CNT. In order to do that, we need to compute the kinetic pre-factor \(\kappa_{CNT}\) that, using the CNT approximations, is: \[\kappa_{CNT}=Z\;\rho_{liq}\;\frac{24D_{S}n_{crit}^{2/3}}{\lambda^{2}}. \tag{8}\] The attachment rate of particles to the critical nucleus (\(f_{n^{crit}}^{+}\)) takes into account the number of available attachment sites on the surface of a spherical nucleus (\(n_{crit}^{2/3}\)) and depends on the jump frequency for bulk diffusion (\(D_{S}/\;\lambda^{2}\)), where \(\lambda\) is the atomic jump distance. Since the functional form of the nucleation barrier can be fitted to the corresponding CNT expression, the computed Zeldovitch factor (Z) coincides with the predicted one. We computed the self-diffusion coefficient with MD simulations using the \(DL\_POLY\) package in the supercooled liquid at \(T_{1}\)=800K. We found \(D_{S}^{Na}=3.4\times 10^{-5}cm^{2}s^{-1}\), in good agreement with an estimate (\(D_{S}^{Na}=2.3\times 10^{-5}cm^{2}s^{-1}\)) based on extrapolation of the available experimental data of Ref. to the temperature \(T_{1}\). Since \(D_{S}^{Na}/D_{S}^{Cl}\approx 1\), we only considered the self-diffusion of the Na\({}^{+}\) ions. We estimated \(\lambda\) as a fitting parameter from the \(f_{n^{crit}}^{+}\) previously calculated; we obtained \(\lambda(T_{1})=10^{2}\dot{A}\). However, considering that the ion size is \(\sigma_{Na}\sim 1.1\;\dot{A}\), this value for the jump distance seems unphysical (\(\lambda\sim 100\sigma\)). Typically, one would expect \(\lambda\) to be of the order of a mean free path. In a molten salt, the mean free path of an ion is certainly less than a particle diameter. This discrepancy also suggests that the CNT picture is inadequate to describe crystal nucleation of NaCl. In summary, we have computed the crystal nucleation rate of sodium chloride from the melt using two independent methods: one based on calculations of the free \begin{table} \begin{tabular}{||c|c|c|c|c|c||} \hline ## T[K] & \(\beta\Delta\mu\) & \(\beta\Delta G_{crit}\) & \(f_{ncrit}^{+}\) & \(I[m^{-3}s^{-1}]\) & \(I_{FFS-TIS}[m^{-3}s^{-1}]\) \\ \hline ## 800 & 0.54 & 24 & 0.013 & 3 \(\times 10^{26\pm 1}\) & \(10^{27\pm 2}\) \\ \hline ## 825 & 0.48 & 29 & 0.033 & 4\(\times 10^{24\pm 1}\) & \(--\) \\ \hline \end{tabular} \end{table} Table 3: Summary of the simulation results for the calculation of the free-energy barrier and the nucleation rate for Tosi-Fumi NaCl. We have found that, to within an order of magnitude, the two approaches yield the same value for the nucleation rate. When we use Classical Nucleation Theory to extrapolate our numerical data to lower temperatures, we observe serious discrepancies with the results of direct calculations. When we use CNT to extrapolate to high temperatures, we find serious discrepancies with the nucleation rates found in experiments. Several factors may contribute to this discrepancy but, at present, it is not yet known which factor is most important. The work of the FOM Institute is part of the research program of FOM and is made possible by financial support from the Netherlands Organization for Scientific Research (NWO). C.V. gratefully acknowledges the financial support provided through the European Community Human Potential Program under contract HPRN-CT-1999-00025, (Nucleus). E.S. gratefully acknowledges the Spanish government for the award of a FPU Ph.D. grant, and the FOM Institute for the hospitality during the period in which this work was carried out. C.V. and E.S. thanks Angelo Cacciuto and Rosalind Allen for valuable discussions and suggestions, and Georgios Boulougouris and Josep Pamies for a critical reading of the manuscript. ## Appendix A Identification of crystalline clusters To distinguish between solid-like and liquid-like particles and identify the particles belonging to a solid cluster, we used the local bond order parameter introduced by Ten Wolde el al.. Although the method we used is the same as the one proposed by Ten Wolde, the definition of a'solid-like' particle is not rigorously the same. The q vector and the thresolds selected were optimazed for the NaCl Tosi-Fumi model. First we computed a normalized complex vector \(q_{4}\) for every particle \(i\). Each component of this vector was given by: \[\vec{q}_{4,m}(i)=\frac{\frac{1}{N_{b}(i)}\sum_{j}^{N_{b}(i)}\Upsilon_{4,m}( \theta_{i,j},\phi_{i,j})}{\vec{q}_{4,m}(i)\cdot\vec{q}_{4,m}^{\,*}(i)},\qquad m =[-4,4] \tag{1}\] Where \(N_{b}(i)\) is the number of neighbours of the particle \(i\) within a cut-off radius of \(4\dot{A}\) (the first minimum in the Na-Cl radial distribution function). Then we computed a scalar product \(q_{4}(i)\cdot q_{4}^{*}(j)\) for every particle \(i\) with each of its neighbours particle \(j\). A particle was considered to be'solid-like' when at least 6 of the scalar products were bigger than 0.35. Finally two'solid-like' particles were considered to be neighbours in the same cluster if they were closer than 3.4 \(\dot{A}\). Ionic fluids are more ordered than Lennard-Jones or Hard Spheres ones, as the radial distribution function shows. However, with the method implemented by TenWolde, we were able to clearly distinguish between solid-like and liquid-like particles in the NaCl Tosi-Fumi model. We enclose a plotA1 that shows the distributions of the number of scalar products bigger than 0.35 at T=800 K for the solid and for the liquid: for values bigger than 6 a particle was considered to be solid-like.
10.48550/arXiv.0906.5274
Rate of Homogeneous Crystal Nucleation in molten NaCl
C. Valeriani, E. Sanz, D. Frenkel
4,593
10.48550_arXiv.1905.06168
## 1 Introduction Synthetic methods allowing one-step C-C bond formation through homogenous-catalyst-mediated transformation of C-H bonds have become increasingly important in both industry and academia. Potential routes for selective C-H bond activation and subsequent C-C bond formation in alkenes and aromatic compounds include hydroarylation by addition of aromatic C-H bonds across an unsaturated C=C bond, or an oxidative coupling that preserves the double bond. The latter reaction, while a highly desirable industrial goal, is challenging from the synthetic point of view. Pioneering examples of Ru-catalyzed coupling of aromatic carbon-hydrogen bonds with olefins were reported two decades ago. Since then (for a review, see Ref.), an increasing number of examples catalyzed by Rh, Ru and Pd have been published, but the mechanistic aspects of the reactions have only been addressed by experimental methods. Motivated by the experimental results of Milstein and coworkers, over a decade ago we started to explore the mechanisms of the concurrent reactions of oxidative coupling and hydroarylation of methyl acrylate (MA) catalyzed by Ru carbonyl complexes (Scheme 1) using hybrid and double hybrid DFT families. These calculations showed that either proton elimination by chloride ion, or hydrogen transfer to coordinated olefin, can serve as the initial step of aromatic C-H bond activation, while oxidative addition mechanism could be excluded. Proton elimination proceeds via transition state **TS1** and initiates oxidative coupling with olefin (**TS2**) according to Scheme 2a. Next, interaction of Ru hydride intermediate **Int3** with MA and HCl regenerates the initial RuCl2 carbonyl (**TS4a**). Alternatively, the catalytic cycle could be closed by interaction of **Int3** with MA and benzene, yielding **Int1 (TS4b)**. Hydrogen transfer to MA (**TS1a** and **TS1b**) causes coexistence of phenyl and one of the two isomeric alkyl ligands in the Ru coordination sphere (**Int4** and **Int5**). This mainly leads to the hydroxylation products via **TS3** and **TS3a** (Scheme 2,b), but coordination of the second olefin molecule followed by oxidative coupling is also possible. ## Scheme 2. Mechanisms of MA interactions with benzene in presence of Ru(II) chloride carbonyl complexes: oxidative coupling (a) and hydroxylation (b). We found that the activation barriers, the relative energies of the key intermediates, and the overall direction of the catalytic reaction strongly depend on the composition of the Ru coordination sphere. Ru complexes that could form in the reaction mixture, and serve as initial species of catalytic cycles,are shown in Scheme 3. In all the complexes, chloride anions are strongly bound to the metal atom; the only exception is RuCl2(CO)4(benzene) (rightmost complex in the 2\({}^{\text{nd}}\) row of Scheme 3) showing CO insertion into one of the Ru-Cl bonds. Moreover, the energetic results, particularly calculated activation barriers, differ between computational approaches and were hard to reconcile with experimental observations. Therefore, for this specific case, we will attempt to obtain rigorous first-principles results by means of coupled cluster theory near the complete basis set limit, and use these results to assess the performance of more affordable computational methods. We believe our results have broader relevance for modeling mechanisms of catalytic reactions mediated by transition metal complexes. The CCSD(T) method5 is considered the "gold standard" of quantum chemistry. However, the computational cost of canonical CCSD(T) calculations scales as O(N7) and becomes prohibitively high for mechanistic studies of practical transition metal catalysis problems. Recently developed domain pair natural orbital methods, such as DLPNO-CCSD(T) of Neese and coworkers6 and PNO-LCCSD(T) of Werner and coworkers,7 scale almost linearly with system size (at least for closed-shell cases) and, in the main group, provide similar accuracy to the corresponding canonical calculation. Recently, benchmark studies of the performance of density functionals for transition metal problems, using DLPNO-CCSD(T) for calibration, have started appearing for reaction energies8 and barrier heights.9 Since the Ru complexes shown in Schemes 2 and 3 are still barely tractable by canonical methods, this enables us to assess the "domain error" for real-size transition metal complexes. In this paper, we assess both DFT and PNO methods against canonical CCSD(T) for the hydroxylation and oxidative coupling of benzene and methyl acrylate (MA) catalyzed by RuCl2-carbonyl complexes, as a representative example for the complex mechanisms of homogeneous catalytic reactions. Having in mind specific computational problems usually addressed in such mechanistic studies, the calculations were divided into four groups: (i) overall reaction energies (Scheme 1) that do not involve transition metals; (ii) relative energies of stable RuCl2 complexes with CO, benzene and MA (Scheme 3) that should reproduce breaking and formation of the metal-ligand bonds and deep alterations of the coordination sphere and electronic structure of the metal atom; (ii) energies of key intermediates along reaction pathways catalyzed by different Ru complexes (Scheme 2) and (iv) barrier heights along these reaction pathways. It would be natural to calculate relative energies of the carbonyl complexes relative to RuCl\({}_{2}\), however, our calculations revealed that it has a triplet ground state, and that the singlet is essentially purely biradical (which is also reflected in the pathological D\({}_{1}\) diagnostic value of 0.435). Discussion of the open-shell calculations are beyond the scope of this preliminary report; therefore, we used RuCl\({}_{2}\)(benzene) as a reference, as RuCl\({}_{2}\) will anyway have no independent existence under the experimental conditions (benzene solvent). At all levels, 1:1 exchange of benzene with CO or MA is energetically unfavorable; all other complexes are exothermic with respect to the reference. The relative energies of key intermediates and transition states along each reaction path were calculated relative to the initial form of the catalyst, (C\({}_{6}\)H\({}_{6}\))(CO)\({}_{n}\)RuCl\({}_{2}\) (n=0-4). ## Computational Methods The Weigend-Ahlrichs basis set family def2-TZVP, def2-TZVP, and def2-QZVPP was used throughout. Reference geometries were optimized at the PBE0-D3BJ/def2-TZVP level using Gaussian 09; identities of transition states were verified by frequency and intrinsic reaction coordinate calculations. At the final geometries, canonical CCSD(T)/def2-TZVPP single-point energy calculations were performed using MOLPRO 2018, both using default frozen cores and including Ru(4s,4p) subvalence orbitals (which _are_ correlated by default in ORCA. We found in this work that the mean absolute effects of Ru(4s,4p) subvalence correlation on carbonyl ligand energies and transition states are both a nontrivial 1.0 kcal/mol.) DLPNO-CCSD(T) and the version with improved iterative triples, DLPNO-CCSD(T1), were calculated with the def-TZVPP basis set using ORCA, likewise DLPNO-CCSD(T)/def2-QZVPP calculations were done for basis set extrapolation using the simple L\({}^{-3}\) formula. TightPNO cutoffs were used to reduce domain discretization error; we found in the present work that DefaultPNO causes errors up to 3.5 kcal/mol in energy differences, and hence do not recommend its use. In addition, single-point DFT calculations with a number of DFT functionals were carried out using ORCA. Aside from PBE0 already mentioned, these include: (a) the Berkeley "combinatorially optimized" B97M-V, oB97X-V and oB97M-V; (b) the M06 family: M06-L, M06 and M06-2X; (c) TPSS and two different hybrids thereof, namely, TPSSh and TPSS0 (10% and 25% HF exchange, respectively); (d) both the original double-hybrid DSD-PBEP86-D3BJ and its reparametrized version revDSD-PBEP86-D4, in the latter, D3BJ also replaced with the very recently published next-generation D4 model. As basis set convergence of double hybrids tends to be dominated by the MP2-like term, we carried out def2-TZVP and def2-QZVPP calculations and applied L\({}^{-3}\) basis set extrapolation, for the remaining DFT functionals we applied def2-TZVPP except for oB97, which were accurate enough that we also tried def2-QZVPP. (Changes are on the order of 1 kcal/mol.) GRID6 was used in all DFT calculations. In all Orca calculations, the RIJCOSX approximation was employed, as well as the RI-MP2 approximation for the double hybrids, in conjunction with the respective appropriate auxiliary basis sets for the def2 family. ## Results and Discussion MAD (mean absolute deviation) and RMSD (root mean square deviation) error statistics with respect to our best CCSD(T) basis set limit estimates are shown in for the four types of energy differences. As expected, the smallest deviations were found for the overall reaction energies, which involve only main group elements. The hybrid functionals of the M06 family (MAD=0.68 and 0.57 kcal/mol for M06 and M06-2X, respectively) and double hybrid revDSD-PBEP86 functionals (MAD=0.64) show the best performance in this group. However, overall for the four criteria, the best results were obtained using oB97M-V and oB97X-V range-separated hybrids as well as by the revDSD-PBEP86 double hybrid, with accuracy between DLPNO-CCSD and DLPNO-CCSD(T). On the 3\({}^{\rm rd}\) rung (meta-GGA) of the Jacob's Ladder, B97M-V performed best for reaction energies and carbonyl complex stabilities; however, M06-L showed similar performance, and TPSS outperformed them for barrier heights (MAD 3.69 vs. 4.54 for M06-L and 5.47 for B97M-V). The revDSD-PBEP86 functional outperforms the original DSD-PBEP86 for all four groups of calculations, whereas DSD-SCAN-D4 improved on DSD-PBEP86-D3BJ only for the carbonyl complexes. The DLNPO-CCSD(T) approach shows very close agreement (MAD=0.35 kcal/mol) with canonical CCSD(T) for the reaction energies, while for other groups there is a bit more daylight between them (MAD=1.04 for carbonyls, 1.56 for intermediates, and 1.58 kcal/mol for TSe). Using the DLNPO-CCSD(T1) approach with improved perturbative triples decreases these statistics to 0.60, 0.80, and 1.02 kcal/mol, respectively. The RMSD/MAD ratios are close to the theoretical value (for a normal distribution) of \(\sqrt[n(\pi)\)\(\approx\)1.2533, for carbonyls and intermediates but much larger for TSes, indicating an outlier (RMSD=0.80, 0.86, 2.05 kcal/mol): we note that TS2-CO2 is essentially biradical (and has D1\(>\)0.4). If we exclude it, MAD and RMSD drop to quite pleasing values of 0.63 and 0.73 kcal/mol, respectively. At all levels except for DLPNO-MP2, the highest RMSD and MAD values were found for the first reaction, i.e. dissociation of benzene and association of CO and MA ligands. Most density functionals tend to overbind the ligands; detailed analysis shows that for one group of functionals (PBE0-D3BJ, SCAN-D3BJ, TPSS, M06L and DSD-PBEP86 family, Fig. 2a) the error becomes greater with increasing number of CO ligands; the other group (M06, M06-2X, TPSS0, TPSSh, and B97, Fig. 2b) exhibits a less pronounced opposite trend, especially in the presence of coordinated benzene. Our findings are consistent (see also our companion paper in the present volume) with the findings of Najibi and Goerigk and ourselves for the very large GMTKN55 main-group benchmark and of Iron and Janes for the MOBH35 transition metal reaction benchmark: Notably, that the range-separated hybrids \(\omega\)B97X-V and \(\omega\)B97M-V acquit themselves particularly well, that revDSD represents an improvement over the original DSD not just for the main group but also transition metals, and that unlike for the main group where empirical double hybrids are clearly superior, they offer no clear advantage over \(\omega\)B97M-V for transition metal reactions. Unlike Iron and Janes, however, who found the new DSD-SCAN double hybrid to be among the best performers for MOBH35, we find it to be inferior to revDSD-PBEP86 and the \(\omega\)B97\(n\)-V family for the present problem. This research was supported by the Israel Science Foundation (grant 1358/15), the Minerva Foundation, and the Helen and Martin Kimmel Center for Molecular Design (Weizmann Institute of Science). RMSD and MAD relative to CCSD(T)/CBS for the four types of energetics considered. Energy deviation relative to CCSD(T)/CBS in different RuCl\({}_{2}\) (CO)\({}_{\text{n}}\) complexes as a function of n.
10.48550/arXiv.1905.06168
Coupled Cluster Benchmark of New Density Functionals and Domain Pair Natural Orbital Methods: Mechanisms of Hydroarylation and Oxidative Coupling Catalyzed by Ru(II) Chloride Carbonyls
Irena Efremenko, Jan M. L. Martin
4,128
10.48550_arXiv.2205.03691
###### Abstract The time-dependent exchange-correlation potential has the unusual task of directing fictitious non-interacting electrons to move with exactly the same probability density as true interacting electrons. This has intriguing implications for its structure, especially in the non-perturbative regime, leading to step and peak features that cannot be captured by bootstrapping any ground-state functional approximation. We review what has been learned about these features in the exact exchange-correlation potential of time-dependent density functional theory in the past decade or so, and implications for the performance of simulations when electrons are driven far from any ground-state. ## I Introduction Time-resolved dynamics of electrons in molecules and solids have become increasingly relevant over the past decades. A description beyond equilibrium electronic structure and excitation spectra is necessary in many applications of fundamental and technological importance: photovoltaic processes, photocatalysis, radiation damage in biomolecules, nanoscale conductance devices, time-resolved pump-probe spectroscopies, and strong light-matter coupling for quantum-based technologies. Scalable theoretical methods to computationally model coupled electron, ion, and photon dynamics help in interpreting and predicting experiments, and to suggest new systems with improved functionalities. Arguably, the electron-component of the problem is the most challenging: there are many of them, they interact with each other as well as with the nuclei and light fields, and they require a quantum mechanical description, unlike nuclei and photons for which, for different reasons, a classical description may suffice. To this end, time-dependent density functional theory (TDDFT) has emerged as a method of choice. TDDFT is inherently a "real-time" method in that dynamics is woven into the formulation from the very beginning with the foundational theorem developed for general time-dependent evolution of arbitrary initial states. While perturbations around the ground-state yield a formalism for excited state energies, their couplings, and response, it is particularly the generality and practical efficiency of the real-time formulation that has led to the possibility of applications on large systems which could not be done otherwise, e.g.. In a sense, TDDFT is an extension of its ground-state counterpart (DFT). The central object is the one-body density, \(n(\mathbf{r},t)\), which is the probability density of finding any one electron at point \(\mathbf{r}\) in space at time \(t\): \(n(\mathbf{r},t)=N\sum_{\sigma_{1}\ldots\sigma_{N}}\int d^{3}r_{2}...d^{3}r_{N} |\Psi(\pi\sigma_{1},\mathbf{r}_{2}\sigma_{2}...r_{N}\sigma_{N})|^{2}\) where \(\sigma_{i}\) represents the spin coordinate, and \(N\) is the number of electrons in the system. Just like DFT, TDDFT allows one to bypass having to solve the analytically and computationally complicated many-body time-dependent Schrodinger equation (TDSE) and instead only requires one to solve a set of time-dependent single-particle equations involving a modified one-body potential whose solutions yield the same time-dependent one-body density as that of the true system. This drastically reduces the amount of computational effort in obtaining quantum many-body dynamics. This is an audacious concept: the idea that a set of non-interacting electrons reproduces the density of interacting electrons seems fantastical, especially in the time-dependent case, where the motion of the electrons driven by some external perturbation or by nuclear motion is affected by their mutual repulsion in an intricate dance. The one-body potential somehow directs the non-interacting electrons to evolve with the same one-body density as the interacting electrons. Recent work has shown that this choreography results in dynamical step and peak structures that have a non-local density-dependence in time and in space, and are completely missed by the commonly used "adiabatic" approximations. Such approximations utilize ground-state exchange-correlation (xc) functionals, and although ground-state potentials also may feature steps and peaks, they appear only in particular situations such as static correlation or fractionally charged systems, while the TDDFT steps appear quite generically. In this review, we revisit what is known about non-adiabatic features of the exact xc potential in TDDFT, focussing on the dynamical steps and peaks. We begin with a brief reminder of the fundamental theory of TDDFT in Sec. II, including a discussion of memory-dependence. Section III demonstrates the dynamical peaks and steps that arise from memory-dependence, by numerical examples as well as analysis of the equations, identifying a "kinetic component" in the xc potential as responsible for these features, and discussing the role of the local accelerations in the system. In Section IV we make a case study on the helium atom, showing that these features, which have largely been discussed in one-dimensional (1D) systems, persist just as strongly in three-dimensions, and verifying the relevance of the analyses previously made. We summarize in Section V. ## II TddFT formalism At the heart of TDDFT lies the Runge-Gross theorem which, for time-dependent problems, plays an analogous role to the Hohenberg-Kohn theorem for ground-state problems. Namely, for a fixed particle-particle interaction and statistics, it establishes a one-to-one mapping between the possibly time-dependent external potential acting on the electrons and the time-dependent density for a given initial state: \[\Psi:n\leftrightarrow v_{\rm ext} \tag{1}\] As a consequence, all physical observables can be expressed as functionals of the density and initial state: for a given \(\Psi\), \(n\) points to a unique \(v_{\rm ext}\) which points to a unique unique Hamiltonian, which in theory points to a unique time-dependent wavefunction, from which any observable can be extracted in principle. But this can be of little more than theoretical interest unless we know how to obtain the density of the interacting system and the observables of interest. As in ground-state DFT, one maps the interacting system to a fictitious non-interacting system, the Kohn-Sham (KS) system, whose orbitals are required to reproduce the density at any instant. The KS orbitals evolve in the one-body KS potential, \(v_{\rm S}({\bf r},t)\), according to a single-particle TDSE: \[\left(-\frac{\nabla^{2}}{2}+v_{\rm S}({\bf r},t)\right)\phi_{i}({\bf r},t)=i \partial_{t}\phi_{i}({\bf r},t)\,. \tag{2}\] (Atomic units are used throughout). The KS potential is written as the sum of three terms in a similar way to DFT: \[v_{\rm S}({\bf r},t)=v_{\rm ext}({\bf r},t)+v_{\rm H}[n]({\bf r},t)+v_{\rm XC}[n;\Psi_{0},\Phi_{0}]({\bf r},t) \tag{3}\] We note that there is much freedom in selecting the initial KS state \(\Phi_{0}\): it can be any wavefunction that reproduces the density of the true interacting initial state and its first time-derivative. The vast majority of applications of TDDFT are for linear response where a perturbative limit of Eqs.- yields a Dyson-like equation, or matrix equations, in the frequency domain, whose solution gives excitation energies and oscillator strengths. The success of TDDFT in this regime, with the available approximate xc functionals, is incontrovertible, e.g. Refs, however not without plemishes for certain classes of excitations, e.g. double-excitations and charge-transfer excitations, which can create havoc for the black-box use of the method in coupled electron-ion dynamics. But the fully non-perturbative regime is of particular interest in TDDFT, due to the much harsher computational scaling of alternative methods. Armed with a good approximation to the xc potential, solving Eq. with Eq. yields a good approximation to the density of the physical interacting system, obtained from non-interacting KS electrons, through \(n({\bf r},t)=\sum_{i=1}^{N}|\phi_{i}({\bf r},t)|^{2}\). Observables that are directly related to the density, such as the dipole moment, can be directly extracted, while further approximations would be needed to extract other observables, such as momentum distributions, double-ionization cross-sections, and current-densities, where the rotational part differs in general from that of the KS system. More often than not, these observables are approximated by those of the KS system; an open question is how errors from the xc functional itself compare with the (usually unacknowledged) errors from the observable evaluation. An exact expression for the xc potential can be derived by equating the second time-derivative of the density, \(\tilde{n}({\bf r},t)\), for the KS system with that of the interacting system. While the Heisenberg equation of motion for the density gives the continuity equation \(\dot{n}=-\nabla\cdot{\bf j}\), \(\tilde{n}({\bf r},t)\) can then be obtained from the equation of motion for the current-density \({\bf j}({\bf r},t)\). Subtracting the equation obtained for the interacting system from that for the KS system results in the decomposition of the xc potential, \(v_{\rm XC}({\bf r},t)\) into kinetic (T) and interaction (W) terms \(v_{\rm XC}({\bf r},t)=v_{\rm C}^{\rm T}({\bf r},t)+v_{\rm XC}^{\rm W}({\bf r},t)\), that have the following structure : \[\nabla\cdot\left(n\nabla v_{\rm XC}^{\rm W}\right)=\nabla\cdot\left(n({\bf r },t)\int n_{\rm XC}({\bf r},{\bf r}^{\prime},t)\nabla w(|{\bf r}^{\prime}-{ \bf r}|)d^{3}{\bf r}^{\prime}\right) \tag{4}\] \[\nabla\cdot\left(n\nabla v_{c}^{\rm T}\right)=\nabla\cdot\left({\cal D}_{{\bf r }^{\prime},{\bf r}}\Delta\rho_{1}({\bf r}^{\prime},{\bf r},t)|_{{\bf r}^{ \prime}={\bf r}}\right)\,, \tag{5}\] (with the \(({\bf r},t)\)-dependences on the left-hand-side understood) where \(n_{\rm XC}({\bf r},{\bf r}^{\prime},t)\) is the xc hole, defined through the diagonal two-body density matrix \(\rho_{2}({\bf r},{\bf r}^{\prime};{\bf r},{\bf r}^{\prime})=N(N-1)\int dr_{3}...dr _{N}|\Psi({\bf r},{\bf r}^{\prime},{\bf r}_{3}..r_{N})|^{2}=n({\bf r}^{\prime}, t)\left(n({\bf r},t)+n_{\rm XC}({\bf r},{\bf r}^{\prime},t)\right)\) and \(\Delta\rho_{1}({\bf r}^{\prime},{\bf r},t)=\rho_{1}({\bf r}^{\prime},{\bf r},t )-\rho_{1,S}({\bf r}^{\prime},{\bf r},t)\) is the difference between the spin-summed one-body density matrix of the true interacting system \(\rho_{1}({\bf r}^{\prime},{\bf r},t)\) and that of the Kohn-Sham system, \(\rho_{1,S}({\bf r}^{\prime},{\bf r},t)\). The differential operator \({\cal D}_{{\bf r}^{\prime},{\bf r}}=\frac{1}{4}(\nabla^{\prime}-\nabla)(\nabla^ {2}-\nabla^{\prime 2})\). Eqs.- give an exact expression for the xc potential in terms of the exact xc hole, and the exact and KS one-body reduced matrices. They are useful for analysis of TDDFT and understanding errors in approximations. Further, they offer a starting point for approximations: although \(\rho_{1,S}({\bf r}^{\prime},{\bf r},t)\) is accessible in a KS evolution, \(\rho_{1}\) and \(n_{\rm XC}\) need to be approximated in terms of KS quantities. We will also return to this in Sec. III. ### Memory: History and Initial-State Dependence In traditional wavefunction-based quantum mechanics, knowing the wavefunction \(\Psi(\mathbf{r}_{1}...\mathbf{r}_{N},t)\) at any time \(t\) is enough to know all properties of the system at that time; its value at earlier times is not required. The same cannot be said about the one-body density. While the TDDFT reformulation of the many-body problem in terms of its one-body time-dependent density \(n(\mathbf{r},t)\) drastically simplifies the problem from the computational viewpoint, it inevitably introduces some complications. In particular, the xc potential is memory-dependent in that the functional \(v_{\mathrm{xc}}[n,\Psi_{0},\Phi_{0}](\mathbf{r},t)\) depends not only on the instantaneous density \(n(\mathbf{r},t)\) but also on the history of the density \(n(\mathbf{r},t^{\prime}<t)\) and the initial states \(\Psi_{0},\Phi_{0}\). This follows directly from the Runge-Gross theorem Eq.: the mapping is between the density- and potential- functions over space and time, and is one-to-one for a given initial state. This would mean \(v_{\mathrm{ext}}(\mathbf{r},t)\) functionally depends on the density over all times but due to causality it depends only the history of the density, not its future. The initial-state dependence means that the same time-dependent density can be obtained by propagating in two different potentials if the systems begin in different initial states. Applying Eq. to the KS system yields that \(v_{\mathrm{s}}(\mathbf{r},t)\) functionally depends on the history of the density and the KS initial state, and thus \(v_{\mathrm{xc}}(\mathbf{r},t)=v_{\mathrm{s}}(\mathbf{r},t)-v_{\mathrm{ext}}( \mathbf{r},t)-v_{\mathrm{ir}}(\mathbf{r},t)\) functionally depends on both the true and KS states: \(v_{\mathrm{xc}}[n;\Psi,\Phi](\mathbf{r},t)\). As mentioned earlier, one can begin in any initial KS state that reproduces the density of the initial interacting state and its first time-derivative; the structure of the exact xc potential has a strong dependence on this choice. There is an intimate connection between history-dependence and initial-state-dependence that can be useful. What we take as the "initial" time can be re-set, and if we know the wavefunctions at the reset time, then we can evaluate the xc potential on the domain of those wavefunctions and a truncated history, i.e. \[v_{\mathrm{xc}}[n;\Psi_{0},\Phi_{0}](\mathbf{r},t)=v_{\mathrm{xc}}[n_{t^{\prime }};\Psi_{t^{\prime}},\Phi_{t^{\prime}}](\mathbf{r},t),\ \ \ t^{\prime}\leq t \tag{6}\] A useful consequence is that if the xc potential is found at some time \(t\) for a particular dynamics of a system, then in _any_ dynamics where the exact same interacting and KS states happen to be reached at some time, then the xc potential at that time will be the same. We will exploit this in Section IV to emphasize the generality of the features of the xc potential found there. Almost all calculations today however completely neglect memory. They use an adiabatic approximation, in which the instantaneous density is input into a ground-state approximation: \(v_{\mathrm{xc}}^{A}[n;\Psi_{0},\Phi_{0}](\mathbf{r},t)=v_{\mathrm{xc}}^{ \mathrm{g.s.}}\left[n(t)\right](\mathbf{r})\). Such an approximation has two sources of error: one arising from the approximation made for the ground-state functional, and the other from the adiabatic approximation itself. To isolate the error from the adiabatic approximation itself, the adiabatically-exact approximation is defined as \(v_{\mathrm{xc}}^{\mathrm{adia-xc}}[n;\Psi_{0},\Phi_{0}](\mathbf{r},t)=v_{ \mathrm{xc}}^{\mathrm{exact-g.s.}}[n(t)](\mathbf{r})\) which can be a useful analysis tool in cases where the exact ground-state xc potential can be computed (usually model 1D systems). Due to the lack of memory, the adiabatic approximation leads to large errors in some applications, sometimes failing completely, but in other cases it has been found to yield good predictions, even when the system is far from a ground-state. It is not completely understood why: possible reasons include, that the adiabatic approximation satisfies a number of exact conditions that are important in the time-dependent case, that in some applications a strong external field dominates over xc effects in driving the dynamics and that partial compensation of self-interaction in the Hartree potential, even at the ground-state level, is enough, especially when the observables involve averaging over the details of the density distribution. Recently it was further argued that an indicator of the expected success or failure of the adiabatic approximation lies in the natural orbital occupation numbers: if the initial KS state has a configuration close to that of the true initial state, and the natural orbital occupation numbers of the true system do not evolve significantly in time, the adiabatic approximation may make good predictions even for strongly non-perturbative dynamics. ### Examples showing the relevance of memory Model systems have been crucial in understanding cases where the adiabatic approximation fails, because numerically exact solutions are available and because the exact xc potential can be extracted to compare with approximations. shows the errors that adiabatic approximations make in a variety of studies. The top panel shows the electronic dipole for field-free evolution of a state that is a 50:50 superposition of the ground and first-excited singlet state of a 1D soft-Coulomb-interacting He atom; such a state may be reached from a ground state driven by a field that is then turned off, for example. In that scenario, the natural choice for the KS initial state is a Slater determinant, which for our two-electron system, is a doubly-occupied spatial orbital. As seen in the figure, the adiabatic exact-exchange (AEXX) and local density approximations (ALDA) do not get the period of oscillations correct, and display additional beat frequencies. In Sec. III, we will see that the exact xc potential has non-adiabatic step and peak features that even the adiabatically-exact approximation lacks. Such features have an even more notable effect in electron-scattering, illustrated in panel b in This shows the reflection probability of an electron initially in a gaussian wavepacket moving towards a target 1D H atom. The approximations severely underestimate the reflection of the wavepacket (\(N_{R}\) is the numberof electrons on the right of the atom). The dashed and solid lines represent different choices of initial KS state, both of which have the same density as the true state: In one, a Slater determinant is chosen for example to simulate the return of an ionized electron to its parent, while in the other, the initial KS state is a two-orbital state which has the same configuration as the interacting state with one electron in the bound target ion, and the other in a gaussian wavepacket. Although there are significant differences in the details of the two time-dependent densities ensuing from these states when propagated under ALDA or AEXX, with the Slater determinant demonstrating spurious oscillations and reproducing the exact dynamics less accurately initially, neither of them capture the eventual reflection even qualitatively. The exact xc potential shows a step and valley structure that is essential for this effect, and is missing in the adiabatic approximations (see also Sec. III). This may explain the underestimated predictions of scattering probabilities and energy transfer that has been observed in real systems. Panel c of gives an example of a resonantly-driven charge-transfer out of the ground-state. A molecule is modeled via an asymmetric double-well, in which the ground-state has two electrons in the left-well. Applying a weak field that is resonant with an excitation to a charge-transfer state triggers a Rabi oscillation, as evident by the large change in the dipole seen as the molecule reaches the charge-transfer state. Although the adiabatic TDDFT approximations shown yield excellent values for the charge-transfer excitation energies, they completely fail to capture the charge-transfer dynamics. Again, dynamical step features develop as the electron transfers (see Sec. III). But adiabatic TDDFT generally fails at resonant driving, even to local excitations, because it violates the fundamental condition that resonant excitation frequencies of a system should not shift with the instantaneous state; with adiabatic approximations, the density-dependence of the KS potential leads to the response of a non-equilibrium state having spuriously-shifted poles, while the exact generalized xc kernel requires a frequency-dependence to correct this spurious shift. Although it has not yet been explicitly shown, it is likely this frequency-dependence is related to dynamical steps and peaks in the time-domain. This could lead to some unreliability in TDDFT simulations of pump-probe spectroscopy; absorption peak-shifts have been observed in a number of molecules. ## III Dynamical steps and peaks One common thread across all the examples in Sec. II.1, is the presence of prominent step and peak features in the exact xc potential, which are significant on the scale of the total KS potential. Three examples illustrating substantial errors in adiabatic TDDFT propagation. a) Dipole moment in field-free propagation of a 50:50 superposition of the ground and first excited state in a 1D He atom, showing the exact, against AEXX and ALDA. The initial KS orbital is chosen to be a Slater determinant. b) Scattering of an electron off a 1D H-atom; the top panel shows the density upon approach, while the lower panel shows the integral of the electron density to the right of the atom (\(x\geq-5\)); dashed curves for the Slater determinant initial state, and solid for a two-orbital state. c) Resonantly-driven charge-transfer out of the ground-state of the double-well shown. Neither self-interaction-corrected LDA (SIC-LDA) nor AEXX capture the dynamics, despite yielding good approximations for their excitations. The exact xc potential was computed by numerically inverting the propagation operator viewed as a function of the potential to target the exact density. In the case of a Slater determinant formed with a doubly-occupied orbital, the KS orbital is directly related to the exact density, and then the xc potential can alternatively be computed using analytical formulas (see also Sec. IV). A salient feature common to all the exact xc potentials over the range of different dynamics are step- and peak-like structure which are absent in the adiabatic approximations, including the adiabatically-exact (although partially reproduced in case c)). These features are therefore non-adiabatic and their absence in adiabatic approximations yield inaccurate density-dynamics. Given that in this case the KS state is a doubly-occupied orbital, \(v_{\rm x}=-v_{\rm n}/2\) has a simple well structure that smoothly cradles the density (the exact \(n({\bf r},t)\) is shown as blue dotted), the step and peak are features of the correlation potential which often dominates the KS potential. In the case shown, the KS state has a fundamentally different structure to the true state; instead, if the KS state is chosen with the same configuration of the true state, the step features are smaller but still appear, although their impact on the ensuing dynamics is less. In we display the xc potential and density for the scattering example of Fig 1b, for the case when \(\Phi=\Psi\). The middle panel shows again a prominent peak and step feature just to the right of \(x=-10\)a.u. at the shoulder of the electron density (top panel), which persists over time and contains the effective correlation needed for reflecting part of the density of the two KS electrons. Adiabatic approximations miss this structure and fail to reflect: the top right and middle panels show the density and xc potential for the ALDA in red. The feature appears in the \(v_{\rm C}^{\rm T}\) component of the exact potential, and the lower panels show that while the adiabatically-exact approximation accurately captures \(v_{\rm XC}^{\rm W}\), it hardly resembles \(v_{\rm C}^{\rm T}\) at all. The figure also plots a non-adiabatic approximation, \(v_{\rm XC}^{\rm S}\), which replaces the exact xc hole and one-body density-matrix in Eqs.- with their KS counterparts; this yields a reasonable approximation to \(v_{\rm XC}^{\rm W}\) but gives zero for \(v_{\rm C}^{\rm T}\). Finally, turning to the example of resonantly-driven charge-transfer of Fig. 1c, we plot in the exact correlation potential at the time the charge-transfer state is reached. A step appears in the adiabatically-exact potential, however it has the wrong height. Ref. shows that the heights of the two steps in the limit of large separation between the donor (D) and acceptor (A) can be given by the differences in ionization potential and electron affinity indicated in the figure; the adiabatically-exact step has a size equal to the \((N_{D}-1)\)-electron derivative discontinuity, and is smaller than that of the exact. Non-adiabatic step and peak features in \(v_{\rm XC}\) for the examples of (a) Snapshots of \(v_{\rm XC}(x,t)\) and \(n(x,t)\) for the field-free dynamics of the superposition state of the 1D He atom, over a half-Rabi cycle, whose dipole appears in The adiabatically-exact and ALDA potentials lack the prominent step and peak structures of the exact. (b) Snapshots of the density and potentials for the e-H scattering problem of The top panel shows the density at a time when the electron begins to reflect from the atom and partially transmit to the left, with the grey being the initial density for reference; the exact (black, left panel), ALDA (red, right panel), \(v_{\rm XC}^{\rm S}\) (blue, right panel). The middle panels show the exact \(v_{\rm XC}(x,t)\) at that time (black), ALDA (red), \(v_{\rm XC}^{\rm S}\) (blue), and grey shows the exact \(v_{\rm XC}(x,0)\). The lower left shows the adiabatically-exact approximation (orange) for the exact \(v_{\rm XC}^{\rm W}\) component (black), and lower right \(v_{\rm C}^{\rm T}\) component. Reproduced from Ref. with permission from the Royal Society of Chemistry.) The exact and adiabatically-exact correlation potentials at the final time when the charge-transfer state is reached, for the resonantly-driven charge-transfer dynamics of transfer, there is also an oscillatory dynamical step similar in nature to that in part (a). Thus these non-adiabatic step and peak features have been demonstrated on a wide range of dynamics, and we now briefly mention some beyond those above. Ref. found them in the exact KS potential for a model describing the propagation of a single electron through an infinite semiconductor wire, using a quasi-particle wavepacket of nonzero crystal momentum added to the ground-state of a semiconductor. Refs. studied field-induced tunneling in a system of two or three spinless electrons in 1D. Ref. showed barrier structures that were essential for autoionization processes in a 1D He atom. Refs. considered Hubbard models, while Ref. demonstrated their importance for single-electron transport through a quantum dot using an Anderson model. Although computationally more convenient to demonstrate on 1D model systems, in Section IV we use the example of the real (three-dimensional) Helium atom to illustrate how the dynamical non-adiabatic features in the exact \(v_{\rm XC}\) shown in the 1D helium atom persist just as strongly in the three-dimensional case. Step and peak features are no strangers to DFT. In the ground-state, they appear in situations associated with "fractional charge" regions, for example when a molecule is near a metal surface, related to the derivative-discontinuity. They appear in interatomic regions of dissociating molecules where they are signatures of static correlation. Without the interatomic step, which is equal to the ionization potential difference between the two fragments, the KS system dissociates to unphysical fractional charges. Associated with the onset of the interatomic step is a peak, and recently a secondary peak was found to appear in the very low density region far to the side of the molecule that heralds the descent of the step back to zero asymptotically. These step and peak features appear in the correlation potential and can be analyzed in terms of changes in the conditional amplitudes of the interacting and KS systems, defined through a factorization, \(\Psi({\bf r}_{1},{\bf r}_{2}...{\bf r}_{N})=\sqrt{\frac{n({\bf r}_{1})}{N}} \Psi_{\rm cond}({\bf r}_{2}...{\bf r}_{N})\). Although the exact exchange potential displays step features when the orbital dominating the density switches to one with a different asymptotic decay, it does not capture the interatomic step in dissociating diatomic molecules which is a strong correlation effect. In the absence of nodal planes of the highest occupied molecular orbital (HOMO), the steps go back down to zero asymptotically, which is a striking difference with the non-adiabatic steps in the time-dependent case. Even when there are nodal planes, steps, peaks, and diverging behavior appear as one traverses across the nodal plane of the HOMO in cases where the density has isotropic decay away from the system. We stress here that the exact ground-state xc potential captures these features but does not capture the non-adiabatic steps in non-equilibrium TDDFT. In the linear response regime, steps have appeared that counter the electric field across long-range molecules, and in the xc kernel near charge-transfer excitations associated with a derivative-discontinuity. Steps were found to be essential in Coulomb blockade phenomena, again related to fractional charges and the derivative-discontinuity. These cases involve perturbations around the ground-state, and the steps are associated with fractional charge effects. Beyond the response regime, dynamical steps were found in ionization processes, where an adiabatic approximation that depended on the fraction of charge remaining locally near the parent atom was able to approximate them. The non-adiabatic steps we are discussing here are a distinct phenomenon to all these cases: they are not related to fractional charges, ionization, nor is an external field necessary, and they are missing from any adiabatic approximation, that is, they cannot be captured by using any ground-state approximation. They appear in the kinetic component of the xc potential, Eq., and are instead related to the local KS velocities and accelerations in the system, as we will discuss next. ### Relationship to local velocity and acceleration Earlier analysis on the two-electron systems have shown that the dynamical steps tend to be associated with the spatial integral of the local KS acceleration in the system, while the peak structures are associated with a maximum in the local KS velocity, or at maximum curvature in the density in regions of low density. To see this, consider first the two-electron spin-singlet where the KS initial state is chosen to be a Slater determinant. \[\nabla\cdot{\bf j}=\nabla\cdot(n({\bf r},t)\nabla\alpha({\bf r},t))=-\frac{ \partial}{\partial t}n({\bf r},t)\,. \tag{8}\] We note that the KS current-density \({\bf j}_{\rm s}({\bf r},t)=n\nabla\alpha({\bf r},t)\) may differ from the true current-density \({\bf j}({\bf r},t)\) by a rotational component (although in 1D they are identical, \({\bf j}_{\rm s}({\bf r},t)={\bf j}({\bf r},t)\)). Inverting Eq. yields the exact KS potential: \[v_{\rm s}({\bf r},t) = \frac{\nabla^{2}\sqrt{n({\bf r},t)}}{2\sqrt{n({\bf r},t)}}-\frac{| \nabla\alpha({\bf r},t)|^{2}}{2}-\frac{\partial\alpha({\bf r},t)}{\partial t}\] \[= \frac{\nabla^{2}\sqrt{n({\bf r},t)}}{2\sqrt{n({\bf r},t)}}-\frac{1 }{2}u^{2}({\bf r},t)-\int_{{\bf r}_{0}}^{{\bf r}}\frac{\partial{\bf u}({\bf s} ^{\prime},t)}{\partial t}\cdot\hat{\bf s}^{\prime}ds^{\prime}\]where the second line is written in terms of local KS velocities and accelerations and is defined up to a global spatial constant. The second term in Eq. is directly related to the local KS velocity, \(\mathbf{u}(\mathbf{r},t)=\frac{\mathbf{\dot{s}}_{\mathbf{\mathbf{\mathbf{\mathbf{ \mathbf{\mathbf{\mathbf{\mathbf{\mathbf{\mathbf{\mathbf{\mathbf{\mathbf{\mathbf{\mathbf{ \mathbf{\mathbf{\mathbf{\mathbf{\mathbf{ \mathbf{ }}}}}}}}}}}}}}}{\mathbf{n}( \mathbf{r},t)}\), and this term tends to produce peak structures, especially near density minima. The third term is related to a spatial-integral of the local KS acceleration, and relative to some reference position \(\mathbf{r}_{0}\), it is given by \(\dot{\alpha}(\mathbf{r},t)-\dot{\alpha}(\mathbf{r}_{0},t)=\int_{\mathbf{r}_{0 }}^{\mathbf{r}}\partial_{t}\mathbf{u}(\mathbf{s}^{\prime},t)\cdot\dot{\mathbf{ s}}^{\prime}ds^{\prime}\) where \(\mathbf{s}^{\prime}\) represents a path from \(\mathbf{r}_{0}\) to \(\mathbf{r}\). This shows that when there is a localized peak in the KS acceleration this term develops a step-like feature there. It should however be noted that the first term of Eq. may also have structures that one can associate with local peaks and local steps: these can partially cancel or enhance the velocity and acceleration terms. For example, in the trivial one-electron case one may get peaks in the local velocity and acceleration but \(v_{\textsc{xc}}\) is zero, since the local curvature of the density can also display local steps and peaks especially near density-minima. But steps that straddle across one side of a localized density to the other, leading to different asymptotic values of the potential in different directions, have a height given by the spatial integral of the localized KS acceleration across the system. We will show an example of this Sec. IV for the (three-dimensional) He atom. Eq. holds for the special case of two-electron singlet systems where the KS initial state is chosen to be a Slater determinant. For a general choice and for \(N\)-electrons, the same expression holds but where the density, velocity, and acceleration are instead those of one of the occupied orbitals. Different orbitals yield different terms individually but the sum of the terms must be the same for each orbital; an example can be found in Ref.. ### Approximations for the dynamical steps and peaks: Importance of \(v_{\textsc{xc}}^{\mathrm{T}}\) Although it is tempting to build an approximation directly from Eq., it is not clear how to, because hiding in the expression is the external potential, whose functional dependence would somehow need to be subtracted out. In practise, the external potential is input as the potential function (_not_ density functional) that is physically applied to the system, and only the xc and Hartree potentials are considered as functionals of the density and initial states. In fact this is essential, since without explicit input of the external potential one could not propagate with Eq.: it defines the physics of the problem being solved, and if instead it was represented by a density functional, causality and predictivity issues arise. We want an approximation for only the xc potential, not for the entire KS potential. Still, it is interesting to note that an approximation for the non-adiabatic part of the full KS potential itself has been explored, by transforming to the local instantaneous rest frame of the density, taking the adiabatic approximation to be exact in such a frame, and transforming back. This results in an approximation that adds precisely the last term of Eq. to an adiabatic approximation. \[-\frac{\partial}{\partial t}\int^{x}\frac{j(x^{\prime},t)}{n(x^{\prime},t)}dx ^{\prime}=v_{\mathrm{ext}}(x,t)-\int^{x}\frac{\frac{\partial j^{2}(x^{\prime},t)}{\partial x^{\prime}}}{2n^{2}(x^{\prime},t)}+\frac{\mathcal{Q}(x^{\prime},t)}{n(x^{\prime},t)}dx^{\prime} \tag{10}\] Instead, Eqs.- can be used to develop approximations. Moreover, analyses based on this decomposition have revealed that the non-adiabatic dynamical features figure more prominently in the kinetic term as compared to the interaction term. As demonstrated in one of the examples in Fig. 2, the step structures are a feature of the kinetic component \(v_{\textsc{c}}^{\mathrm{T}}\) and are absent in adiabatic approximations to this term. While adiabatic approximations may tend to approximate \(v_{\textsc{xc}}^{\mathrm{W}}\) well, they do a poor job of capturing the structure in \(v_{\textsc{c}}^{\mathrm{T}}\). This is not surprising given the form of the two terms: While the interaction term benefits from smoothing through the integral, the appearance of multiple gradients in the kinetic component can create large step and peak features that evolve in time, especially in regions where the density has local minima. One such class of approximations is the density-matrix coupled approximation of Refs.. Here, the correlated one-body density matrix needed in \(v_{\textsc{c}}^{\mathrm{T}}\) is replaced by an approximated one computed from the first equation of the BBGKY hierarchy, and propagated alongside the KS calculation. In the BBGKY equation the two-body density matrix is approximated by that of the KS system. This ensures that the diagonal part of this density matrix stays equal to the density of the KS system at all times, while the off-diagonal terms can differ from the KS one-body density matrix (as would be the case in an exact calculation). Even though this approach satisfies most exact conditions known in TDDFT, and is able to capture the elusive dynamical steps and peaks, it remains numerically too unstable to be of practical use. An improved approach within this class of approximations would include an approximation for two-body density matrix as a time-dependent functional of either or both the KS and correlated one-body density matrices. Eq. suggests that the steps are a probe of the difference between the local character of the true and KS density-matrices near the diagonal (Eq. 5). Choosing initial KS states to minimize this difference may improve the performance of the adiabatic approximations at short times. The idea that the performance of an adiabatic approximation is connected with how small the variation of natural orbital occupation numbers is was investigated in Ref.. Ref. found that the largest step structures appeared at local minima of the largest occupation numbers. However it should also be noted that the impact of a large localized dynamical step or peak on the dynamics may be relatively small if the structure oscillates rapidly in time or is very localized in a region of small density. ## IV Case study: dynamics in the helium atom The examples illustrated in the previous section involved 1D systems, and a question then arises: are the non-adiabatic steps and peaks equally prominent in real three-dimensional systems? It is often said that correlation effects are enhanced in reduced dimensionality. A sensible comparison would be to compare the steps and peaks in a three-dimensional system undergoing analogous dynamics to one of the 1D examples. We find the exact interacting state at time \(t\) in terms of the eigenstates of the He atom: \(1^{1}S_{0}\), denoted \(\Psi_{0}\), and singlet first excited state \(2^{1}P_{1}\) that has angular quantum numbers \(L=1\) and \(M=0\), denoted here \(\Psi_{1}\). \[|\Psi(t)\rangle=\frac{1}{\sqrt{1+|a|^{2}}}\left(|\Psi_{0}\rangle+ae^{-i\omega t }|\Psi_{1}\rangle\right) \tag{11}\] The parameter \(a\) allows us to tune the proportion of the excited state \(\Psi_{1}\) that goes into the superposition. For the analog to the 1D He example in Figs. 1a and 2a, \(a=1\), the 50:50 superposition. \[n({\bf r},t)=\frac{1}{1+|a|^{2}}\left(n_{0}({\bf r})+|a|^{2}n_{1}({\bf r})+2an _{01}({\bf r})\cos(\omega t)\right) \tag{12}\] We note that although this may seem to be a very special case of dynamics, the results for the xc potential at any given time \(t\) in fact apply to a far wider class of dynamics: due to Eq. the xc potentials we will find apply in any situation where the instantaneous true state is given by Eq. 11 and an initial Slater determinant is chosen for the KS state. This means that independently of how the system reached instantaneous state at some time \(t\) the xc potential at that time is the same. To find the exact xc potential, we will need to invert the time-dependent KS equation for the three-dimensional case, but first we need to find the exact densities of the interacting eigenstates and the transition density in order to construct the target density Eq.. We use the results of Refs. to find the exact density. There, the system is solved using the close-coupling method, in which the angular part of the exact wavefunction is written in terms of coupled spherical harmonics while the radial part is obtained by the finite element discrete variable representation method. ### Extracting the exact xc potential Here we choose the initial KS state as a Slater determinant: this is the natural choice if the state Eq. is reached from applying an external field to a ground-state and then turning the field off. One would use ground-state DFT to find the initial KS orbitals, and by the ground-state theorems, this is a Slater determinant. Since the KS evolution involves a one-body Hamiltonian, the state remains a single Slater determinant. This choice of initial KS state is the most relevant choice for simulations of general dynamics of the atom when beginning in the ground state. This is often the case for simulating experiments involving laser-driven dynamics, for example, or in photovoltaics where one models the initial photo-excitation process itself; and, as mentioned above, the xc potential we find at time \(t\) is the same for any situation in which the true state Eq. 11 is reached and a KS Slater determinant is chosen. For the case of the He atom, the KS state is then a doubly-occupied orbital. The exact xc potential is then obtained from subtracting \(v_{\rm ext}=-2/|{\bf r}-{\bf r}^{\prime}|\) and \(v_{\rm n}({\bf r},t)\) from Eq. 3. Further, one can isolate the correlation component by noting that for this KS state, \(v_{\rm x}({\bf r},t)=-v_{\rm n}({\bf r},t)/2\). Thus, finding the exact xc potential reduces to solving Eq. for \(\alpha({\bf r},t)\) and inserting it into Eq.. We note that for a different choice of initial KS state, e.g. using a two-configuration state that is more similar to that of the actual interacting state, the inversion to find \(v_{\rm xc}\) involves an iterative numerical procedure; some examples for the 1D analog of the He dynamics can be found in Refs. (see also discussion in Sec. III). This could be a more natural state to begin the KS calculation in some situations, e.g. if the state was prepared in such a superposition at the initial time. The importance of judiciously choosing the KS initial state when using an adiabatic approximation has been realized and exploited in strong-field charge-migration simulations. ### Numerical Details We observe that the second equality in Eq. has the form of a Sturm-Liouville type equation, which yields a unique solution for \(\alpha({\bf r},t)\) for a given boundary condition. Furthermore, thanks to the azimuthal symmetry of our density (\(M=0\) at all times), we need solve this in effectively two dimensions, \((r,\theta)\). \[\alpha({\bf r}\rightarrow\infty,t)=0\quad\mbox{and}\quad\frac{\partial}{\partial \theta}\alpha({\bf r},t)|_{\theta=\pi,0}=0\,. \tag{13}\] We use a rectangular computational domain for the grid (\(r,\theta\)), extending from \(0\to R=30\) a.u. in \(r\) (the density is negligible this far from the nucleus) and \(0\rightarrow\pi\) in \(\theta\), and the boundary conditions, Eq. \[\alpha(r=R,\theta,\varphi,t) = 0\] \[\frac{\partial}{\partial\theta}\alpha(r,\theta,\varphi,t)|_{\theta =\pi} = 0\quad\quad\frac{\partial}{\partial\theta}\alpha(r,\theta,\varphi,t)|_{ \theta=0}=0 \tag{14}\] With the finite-difference scheme the matrix representation of the derivative operator is highly sparse, and consequently the semi-local operator \(\nabla\cdot(n({\bf r},t)\nabla)\) is also, coupling only a few nearby grid points. The dimensions of the matrix are \((N_{\theta}N_{r})\times(N_{\theta}N_{r})\) where \(N_{\theta}=50\) and \(N_{r}=301\) are the number of grid points in the (\(\theta,r\)) computational domain. The high sparsity also allows for efficient matrix inversion. Despite the computational efficiency, caution is required to avoid numerical inaccuracies especially where the density becomes small. To ensure that our conclusions are robust, we restrict our analysis to regions where the inversion is fairly accurate. This is ensured by checking that the action of the matrix representing \(\nabla\cdot n({\bf r},t)\nabla\) on the solution vector \(\alpha({\bf r},t)\) agrees with the right-hand-side of Eq.. In addition to the uniqueness entering the solution of Eq. as stated before, choosing the initial condition as \(\alpha({\bf r},0)=0\) fixes our initial state as \(\phi({\bf r},0)=\sqrt{n({\bf r},0)/2}\). The Runge-Gross theorem then ensures that there is a unique \(v_{\rm xC}({\bf r},t)\) that reproduces the exact \(n({\bf r},t)\) and yields a unique \(\alpha({\bf r},t)\) at later times. It is not _a priori_ obvious that the unique solution to Eq. with the boundary-condition Eq. applied with time \(t\) as a parameter is compatible with the TDKS evolution, but the results do evolve smoothly in time. Moreover, the rapid decay of the density at large \(r\) in the left-hand side of Eq. implies that this leads to the matrix elements corresponding to these points being killed off which renders the boundary conditions at \(r=R\) superfluous. The numerical inversion of the matrix operator \(\nabla\cdot n({\bf r},t)\nabla\) subject to the boundary conditions Eq. 13 described above, produces the solution of Eq. for \(\alpha({\bf r},t)\). When used in Eq. this yields the KS potential, \(v_{\rm s}({\bf r},t)\). \[\nabla^{2}v_{\rm H}({\bf r},t)=-4\pi n({\bf r},t)\,, \tag{15}\] ### Results The problem possesses several symmetry features which we leverage to simplify the analysis. The azimuthal symmetry mentioned earlier and the fact that the states that enter the superposition are those for which \(L=0\) and \(L=1\), imply that the density, current, and potentials in the lower half-plane (\(\pi/2<\theta<\pi\)) evolve in exactly the same way as those in the upper half-plane (\(0<\theta<\pi/2\)) but a half-cycle out of phase, i.e. if \(O(r,\theta,t)\) represents the above quantities, then \(O(r,\pi-\theta,t)=O(r,\theta,t+T/2)\). Another consequence of the simple form of the superposition is the fact \(O({\bf r},T-t)=O({\bf r},t)\). It is therefore sufficient to look at time-snapshots only over a half cycle in one of the octants; the discussion will carry over to the other octants in accordance with the above conditions. In we plot the snapshots of the xc potential, \(v_{\rm xC}({\bf r},t)\) at different fractions of the period of oscillation, \(T=2\pi/\omega=8.057\) a.u.in the octant spanned by \([r=]\times[\theta=(\pi/2,\pi)]\). We observe the xc-potential displays a prominent peak and a step across \(r\) at \(t=0\) in the region swept by \(\pi/2<\theta<\pi\) that decrease in magnitude over the first half-period time until vanishing from the octant and appearing on the other side of \(\theta=\pi/2\). These features have been shown to be completely missing in adiabatic approximations for 1D cases (e.g. Fig. 1) yet they often dominate the KS potential. Here we find these non-adiabatic features persist just as prominently, in analogous dynamics of the real three-dimensional He atom. This justifies that conclusions drawn from previous studies involving 1D systems do apply to real systems as well and that these strong correlation effects do not arise from dimensional reduction. We expect that the lack of the step and peak features in adiabatic approximations lead to less structured density profiles, as was seen in 1D cases. It is important to mention here that at any instant of Exact xc potential, \(v_{\rm xC}({\bf r},t)\) for field-free evolution of the 50:50 superposition state (\(a=1\)) of the He atom in the range \(\pi/2<\theta<\pi\) at times \(t=0,T/6,T/4\), and \(T/2\) The flatness of the KS potential asymptotically in each octant is consistent with a zero current-density there. Our inversion to find the potential very close to \(\theta=\pi/2\) is not reliable near the density's local minimum along this line: Looking across the \(\theta-\)direction, we observe a sharp step and peak in \(v_{c}\) at \(\theta=\pi/2\) at large \(r\), which yields a force that prevents the KS current vectors from crossing the \(xy\)-plane, in a manner consistent with the behavior of the true current vectors. These features make a numerical inversion near \(\theta=\pi/2\) difficult. At \(\theta=\pi/2\), the density of the \(P\)-state \(\Psi_{1}\) vanishes, and the large and sharp change in the potential reminds us of the divergent behavior along the highest-occupied molecular orbital nodal plane found in the asymptotic region of the ground-state potential mentioned in Sec. III for cases where the density has the same decay in all directions. But in our case the density does decay differently along this plane than in other directions, and it cannot be captured by any adiabatic approximation. As discussed in Sec. III, we find that a significant contribution to the peak is related to the local KS velocity, while the step when a cut is taken across a fixed \(\theta\) is related to the radial integral of the local acceleration, \(\dot{\alpha}(r,t)=\int^{r}\partial_{t}\mathbf{u}(\mathbf{r}^{\prime},t).\hat{ \mathbf{r}}^{\prime}dr^{\prime}\), and likewise for the step at \(\pi/2\) in the \(\theta\)-direction. We will explicitly demonstrate this in Sec. IV.4. Taking different superpositions of the ground and excited states is further evidence that the step and peak features are universally present in real three-dimensional systems. The first and third columns of shows the KS and correlation potentials at the initial time, when \(a\) in Eq. 11 is changed through \(0,1,2,\infty\), going from a pure ground-state, through to the purely excited state; these two limits are of course time-independent. The correlation potential in the purely excited case (top panel, third row) has a barrier with a structure that is not dissimilar to the 1D case in magnitude and in shape (Figs. 3 and 4 in Ref.); the KS potential must be such to maintain the constant excited \({}^{1}P\) density at all times with a non-interacting doubly-occupied orbital. (For all cases, \(v_{\mathrm{s}}\) goes asymptotically like a constant \(-1/r\); the different scales makes this behavior less apparent in some cases.) ### "Non-Interacting Helium" As observed in Sec. III, it is the kinetic component \(v_{\mathrm{C}}^{\mathrm{T}}\) that is largely responsible for the non-adiabatic step and peak features, and that even the best adiabatic approximation to this term, the adiabatically-exact, does not capture them. This component of the xc potential depends on spatial derivatives of the difference between the true and KS one-body density matrix (Eq.), and so is significantly affected by the choice of initial KS state, as discussed earlier. In fact, even considering the (unphysical) limit of zero electron-interaction in the He atom, \(v_{\mathrm{xc}}=v_{\mathrm{C}}^{\mathrm{T}}\) is non-zero when the initial KS state is chosen differently to that of this non-interacting atom. This raises the theoretical question of how do the steps and peaks compare in such a system to the physical situation, that is, how are these affected by electron-interaction itself, as opposed to configurational effects. To this end, we consider here a "noninteracting He" atom, and prepare our system in a superposition of the ground state and an excited state such that the configuration is the same as that of the actual interacting He atom considered in the previous subsection, except for the fact that the wavefunctions that go into the superposition are products of the ordinary hydrogenic wavefunctions. That is, in Eq., instead of the interacting ground state, \(|\Psi_{0}\rangle=|1^{1}S_{0}\rangle\) and the first excited state Comparison between interacting and non-interacting helium atom. Columns 1 & 3 correspond to the interacting case while columns 2 & 4 correspond to the non-interacting case. Top panel shows \(n(\mathbf{r},t)\) (dashed red), \(v_{\mathrm{s}}(\mathbf{r},t)\) (dashed blue) and correlation \(v_{\mathrm{C}}(\mathbf{r},t)\) (solid blue) for purely ground state (\(a=0\), columns 1 & 2) and purely excited state (\(a=\infty\), columns 3 & 4), at a cross section taken at \(\theta=0.75\pi\)\({}^{\mathrm{Noted}}\). Middle panel: \(n(\mathbf{r},t)\), \(v_{\mathrm{S}}(\mathbf{r},t)\) and \(v_{\mathrm{C}}(\mathbf{r},t)\) plotted for superposition states 50:50 (\(a=1\), columns 1 & 2) and 20:80 (\(a=2\), columns 3 & 4) at time \(t=0\). Lower panel: Same quantities as middle panel for \(t=T/8\). \(|\Psi_{1}\rangle=|2^{1}P_{1}\rangle\), we prepare our system in the following superposition \[|\Psi^{}(t)\rangle=\frac{1}{\sqrt{1+|a|^{2}}}\left(|\Psi^{}_{0}\rangle+ae^{ -i\omega t}|\Psi^{}_{1}\rangle\right) \tag{16}\] where \(\Psi^{}_{0}(r_{1},r_{2})=\phi_{1}(r_{1})\phi_{1}(r_{2})\) and \(\Psi^{}_{1}=\frac{1}{\sqrt{2}}[\phi_{1}(r_{1})\phi_{2}(r_{2})+\phi_{1}(r_{ 2})\phi_{2}(r_{1})]\) are composed in terms of the products of hydrogenic wavefunctions \(\phi_{1}\) and \(\phi_{2}\) \[\phi_{1}(r)=\sqrt{\frac{8}{\pi}}e^{-2r},\;\;\;\phi_{2}(r)=\frac{\cos\theta}{ \sqrt{\pi}}re^{-r} \tag{17}\] Following the same steps as involved in the solution of Eq. we analyse the exact KS potential \(v_{\rm s}\) along with the exact correlation potential \(v_{\rm c}({\bf r},t)\) for this non-interacting version of the Helium atom. The second and fourth columns of Figure present these potentials for \(a=0,1,2\) and \(\infty\). Comparing with the first and third columns we see that they display structures that resemble the ones found in the interacting case. We find that the peaks and the steps follow a similar trend to that displayed by the interacting system when the proportions of ground and excited states are changed. The appearance of such dominating steps in the correlation potential here is fundamentally linked to the difference in configurations of the interacting superposition state and KS Slater determinant rather than being a direct consequence of electron interaction. Tuning down the electron-interaction dampens the peak but the step remains. Both these features have a role in somehow nudging non-interacting electrons to be in separate lobes of the density, with the peak containing dynamical Coulomb interaction effects further enhancing the separation. An accurate approximation for the kinetic component \(v_{\rm c}^{\rm T}\) is required to capture these features. Finally, we demonstrate with this non-interacting system, the decomposition of the exact KS potential in the second line of Eq.. In particular, the second term depends on the local velocity, and yields a peak structure, while the third term dependent on a spatial integral of the local acceleration is primarily responsible for the step. Having analytic expressions for the "true" wavefunction gives analytic expressions for \(n({\bf r},t)\) and \({\bf j}({\bf r},t)\), and so allow us calculate the local acceleration and velocity in a straightforward way. verifies the assertions above. We note that the first term in Eq., that depends on the curvature of the instantaneous density, also can yield peaks near the minimum of the density; in fact at the initial time, where the current-density is zero, the peak structure arises predominantly from the first term. The first term can also yield localized step-like structures, but it is only the third term that can yield a step structure that asymptotes to different constants away from a localized density. Interestingly, in this case, \({\bf j}_{\rm s}({\bf r},t)={\bf j}({\bf r},t)\), in contrast to the physical interacting He atom where snapshots of their difference can be found in Ref.. ## V Conclusions and outlook TDDFT tasks the potential driving the non-interacting electronic system with reproducing the exact time-dependent density of interacting electrons at all times. As a result, the exact time dependent xc potential must choreograph a rather unusual dance through a landscape of dynamical steps and peaks that nudge non-interacting electrons to evolve with the same density as Coulomb-interacting ones. In this work, we have reviewed aspects of what is known about these steps and demonstrated them on a real three-dimensional system, the He atom. They are distinct from step and peak features that arise in the ground-state case which tend to be associated with fractional charge and static correlation issues. The dynamical steps and peaks instead appear generically in dynamics far from the ground-state, associated with a kinetic component to the xc potential that is sensitive to the difference of the exact and KS one-body density matrices near the diagonal. They have a non-local dependence on the density in time, and are completely absent in any adiabatic approximation. They also have a non-local dependence on the density in space, as can be seen e.g. from the different asymptotes of the potential. The entanglement of spatial and time non-locality in TDDFT has been pointed out in extended systems in linear response, where approximations that build in frequency-dependence while remaining local in space violate the zero-force theorem, and here we see the time- and space- entanglement is a generic feature in non-perturbative dynamics of finite systems as well. The lack of these structures in approximations in use today lead to errors, including spurious peak-shifting that can muddle interpretation of pump-probe spectra. (a) Decomposition of \(v_{\rm s}({\bf r},t)\): Terms 1-3 are the three terms in Eq. plotted as a function of \(\theta\) at r=2.0 a.u. (b) Same three terms now plotted across as a function of \(r\) at \(\theta=0.75\pi\). In both (a) and (b), the density, \(n(r,t)\) (dotted magenta) as a function of \(\theta\) is plotted on a different scale at the same r. Going beyond the adiabatic approximations, functionals that have explicit dependence on the instantaneous orbitals incorporate memory (e.g. time-dependent exact exchange but the orbital-dependence of these commonly only involves exchange which is inadequate to capture these dynamical steps, as they appear in \(v_{c}^{\mathrm{T}}\). Building practical non-adiabatic approximations for these features have so far proven elusive, although the recent density-matrix coupled approximations show it may be possible. The hope is that such approximations would increase the reliability of real-time TDDFT for dynamics of electrons in non-perturbative fields. The challenge is certainly worth it.
10.48550/arXiv.2205.03691
The Exact Exchange-Correlation Potential in Time-Dependent Density Functional Theory: Choreographing Electrons with Steps and Peaks
Davood Dar, Lionel Lacombe, Neepa T. Maitra
3,624
10.48550_arXiv.1110.5423
###### Abstract We consider the flow-driven translocation of single polymer chains through nanochannels. Using analytical calculations based on the de Gennes blob model and mesoscopic numerical simulations, we estimate the threshold flux for the translocation of chains of different number of monomers. The translocation of the chains is controlled by the competition between entropic and hydrodynamic effects, which set a critical penetration length for the chain before it can translocate through the channel. We demonstrate that the polymers show two different translocation regimes depending on how their length under confinement compares to the critical penetration length. For polymer chains longer than the threshold, the translocation process is insensitive to the number of monomers in the chain as predicted in Sakaeue _et al._, _Euro. Phys. Lett._, **72** 83. However, for chains shorter than the critical length we show that the translocation process is strongly dependent on the length of the chain. We discuss the possible relevance of our results to biological transport. ## I Introduction The passage of polymer chains through nanochannels is an ubiquitous process in nature. Biopolymers, such as DNA and RNA, have to cross a multitude of barriers to perform different biological functions, for example, in translocating through cellular membrane pores or when ejecting from viral capsids. Considerable interest has arisen in the details of the translocation process due to a vast array of practical applications that include the potential sequencing of DNA chains, the sorting of biopolymers using smart entropic traps or in sieving processes in the pharmaceutical and food industries. The translocation of a polymer chain through a nanochannel can be described as a three-stage process: the chain must first find the pore, then enter it and finally move through it. It is the second stage, of a polymer entering a long nanochannel, that we concentrate on here. Recent theoretical progress has shown that the entry of a polymer chain into a narrow channel driven by a fluid flow can be regarded as a tunnelling phenomenon, in which the entropic cost of squeezing the chain into the pore is opposed by the energy gain provided by the driving hydrodynamic force. The outcome of such competition is a free energy barrier, which the polymer has to surpass in order that translocation takes place. According to the de Gennes blob model for confined chains, the barrier is overcome once the chain has been pushed a distance \(y^{*}\) into the channel, at which point, the hydrodynamic force wins over the entropic pressure. Therefore, the strength and position of the the free energy barrier can be controlled by varying the driving volumetric flux, \(J\). Increasing \(J\) has the effect of shifting the position of the barrier closer to the channel entrance, up to a critical flux, \(J_{c}\), at which the barrier height becomes comparable to the thermal energy and translocation takes place. \[J_{c}\sim\frac{k_{B}T}{\eta}, \tag{1}\] Remarkably, the threshold flux given by eq. is independent of the degree of polymerisation of the chain, \(N\). Physically, this occurs because the free energy barrier is overcome when a major portion of the chain is still outside the pore. This behaviour thus belongs to a long-chain regime, where the polymer always reaches the position of the barrier before being completely brought into the channel. Conversely, there is a short-chain regime, in which the whole chain is pushed inside the pore without reaching \(y^{*}\). In this case, the barrier still arises from the competition of the entropic pressure and the hydrodynamic force as before, but with the crucial difference that its magnitude and position are determined by the length of the confined chain. The practical consequence is that the translocation process becomes \(N\)-dependent, a feature of potential interest, for example, in sorting chains according to their number of monomers. In this paper we shall focus on the existence of two regimes of translocation between long and short chains, and will examine the effect of the size of the chain on the threshold flux in each regime. Applying the energy barrier approach of Ref. we obtain a different scaling relation for the threshold translocation flux for long and short chains and test our predictions against numerical simulations. The numerical modelling of polymer-solvent dynamics has made considerable progress during the last few years. In particular, coarse-grained models for the fluid that offer a physical coupling with polymer chains have been developed, leading to a reliable representation of the dynamics of the system whilst retaining an efficient numerical performance. Given that the large-\(N\) limit relevant to the blob model is difficult to access directly, we shall adopt a coarse-grained representation introduced by Dunweg _et al._ and described by Usta _et al._, which is based on a lattice-Boltzmann model for the fluid coupled to a bead-spring representation of the polymer chain through an effective Stokes drag. This approach has been used to study the lateral migration of chains in channel flows under external forces and pressure gradients. More recently, Markesteijn _et al._ used the same model to study the forced translocation of long chains into narrow pores in the limit where the bead-spring model is expected to give the blob-model behaviour. As expected, they confirmed the scaling of eq. showing that the coarse-grained model is indeed able to capture the main mechanisms at play in the translocation process. The rest of this paper is organised as follows: in Section II we present scaling arguments for the position and height of the energy barrier, and for the threshold translocation flux in the long-chain and short-chain regimes. In Section III we describe the lattice-Boltzmann algorithm and the bead-spring model for the polymer chain, and list the set of parameters used to carry out the numerical simulations. Our numerical results are presented in Section IV. After describing the simulation setup in Section IV.1, in Section IV.2 we study the translocation process for long chains, demonstrating the presence of the energy barrier. We then focus on the cross-over to the short-chain regime in Section IV.3, and on the short-chain translocation process which gives rise to a length-dependent threshold flux, in Section IV.4. Finally, in Section V we present the discussion and conclusions of this work. ## II Entry of a polymer chain into a nanopore: blob model In this section we present scaling arguments to predict the threshold flux allowing the translocation of linear polymer chains through a narrow channel. As we have anticipated above, we will examine two regimes for the translocation corresponding to long and short chains. ### Confined chains in equilibrium We start by reviewing the scaling argument of Sakaue _et al._ for long, blob-like linear chains. Consider a linear polymer chain of ideal radius \(R_{0}\simeq aN^{1/2}\), composed of \(N\) monomers of size \(a\). In equilibrium, the Flory radius of the free chain is \(R\simeq aN^{3/5}\). The size of the chain changes when it is confined in a channel of width \(D<R\). As depicted in Fig. 1, the chain stretches due to the effect of confinement. \[\frac{F}{k_{B}T}\simeq\frac{L^{2}}{R_{0}^{2}}+\frac{N^{2}a^{3}}{LD^{2}}, \tag{2}\] and reads, \[L\simeq D\left(\frac{a}{D}\right)^{5/3}N\simeq D\left(\frac{R}{D}\right)^{5/3}. \tag{3}\] In the de Gennes' blob picture, the confined chain accommodates itself into \(M\) blobs of uniform size \(\xi\). The number of blobs then obeys the relation \(M=LD^{2}/\xi^{3}\). Within each blob the effect of confinement is unimportant. Hence, the size of the blob scales as \(\xi\simeq aP^{3/5}\), where \(P\) is the number of monomers in each blob. \[\phi=\frac{Na^{3}}{LD^{2}}=\frac{Pa^{3}}{\xi^{3}},\] from which it follows that for a linear chain, the size of the blob is comparable to the pore size, _i.e.,_ \[\xi\simeq D. \tag{4}\] ### Long-chain translocation We now consider a partly confined chain in the presence of a driving flow. For a weak driving flux, the conformation of the chain outside the channel is close to equilibrium. Hence, the translocation process is controlled by the forces acting on the confined part of the chain. This corresponds to the inside approach in the terminology of Sakaue _et al._ and is valid as long as the length of the pore, \(L_{p}\), is larger than its thickness, \(D\). Schematic representation of a long blob-like polymer chain under confinement. The chain is confined up to a distance \(y\), adopting a configuration of \(M\) stacked blobs of size \(\xi\). For a linear chain, the size of a blob scales as \(\xi\sim D\). When subject to a driving flux, \(J\), the competition between the hydrodynamic drag and the entropic pressure sets a barrier to translocation, located at \(y^{*}\). Once it has reached the position of the barrier the polymer is able to translocate. The confinement of the chain has an entropic penalty of the order of \(k_{B}T\) per blob. Thus, for a chain composed of \(M\) blobs which penetrates a distance \(y\) into the channel, as shown in Fig. 1, there is an energy cost \[F_{S}\simeq k_{B}T\frac{yD^{2}}{\xi^{3}}=Ak_{B}T\left(\frac{y}{D}\right), \tag{5}\] As expected, the entropic cost increases with \(y\), given that a larger number of monomers are pushed into the channel. Countering the entropic cost is the fluid flow, which tends to drag the chain further into the channel. The hydrodynamic drag per blob scales as \(\eta u\xi\), where \(u=J/D^{2}\) is the typical velocity in the translocation direction inside the channel. For \(M\) blobs, the change in free energy corresponds to the work done by the fluid to displace the chain up to a distance equal to \(y\), \[F_{H}\simeq-\eta u\int_{0}^{y}M(y^{\prime})\xi(y^{\prime})\mathrm{d}y^{\prime} =-\frac{B\eta J}{2}\left(\frac{y}{D}\right)^{2}, \tag{6}\] As before, we introduce a proportionality constant, \(B\). Adding the contributions given by eqs. \[\frac{\Delta F}{k_{B}T}(y)=F_{S}+F_{H}=Ak_{B}T\left(\frac{y}{D}\right)-\frac{B \eta J}{2}\left(\frac{y}{D}\right)^{2}. \tag{7}\] The competition between the entropic and hydrodynamic terms gives an energy barrier, \(\Delta F^{*}\), located at \(y=y^{*}\). Differentiation of eq. \[y^{*}=\left(\frac{A}{B}\right)\frac{k_{B}T}{\eta J}D, \tag{8}\] to obtain its magnitude, \[\frac{\Delta F^{*}}{k_{B}T}=\frac{A}{2}\left(\frac{y^{*}}{D}\right)=\frac{C}{2 }\frac{k_{B}T}{\eta J}, \tag{9}\] In order that translocation proceeds, the chain must overcome this barrier. \[P=\kappa_{D}\tau_{m}\exp{\left(-\frac{\Delta F^{*}}{k_{B}T}\right)}, \tag{10}\] We define the threshold flux by \[P(J_{c})=P_{c}, \tag{11}\] Inverting eq. \[J_{c}\sim\frac{k_{B}T}{\eta}.\] As we had anticipated, \(J_{c}\) does not depend on \(N\). This is because the polymer reaches the position of the barrier, \(y^{*}\), before it is completely confined in the channel, _i.e._, the translocation is triggered when \(N^{*}<N\) monomers have been pushed into the channel. To obtain the critical number of monomers, \(N^{*}\), it suffices to replace \(L\) by \(y^{*}\) in eq., from which \[N^{*}\simeq\frac{y^{*}}{D}\left(\frac{D}{a}\right)^{5/3}\simeq\left(\frac{A}{B }\right)\left(\frac{D}{a}\right)^{5/3}\frac{k_{B}T}{\eta J}, \tag{12}\] We note that in Ref. the scaling for the threshold flux was obtained by setting \(\Delta F^{*}\simeq k_{B}T\) in eq.. This corresponds to a free energy barrier whose height is overcome by pushing only one blob into the pore, with a corresponding number of monomers \(N_{D}\equiv N^{*}(y^{*}=D)\simeq(D/a)^{5/3}\). ### Short-chain translocation The short-chain regime corresponds to the limit \(L<y^{*}\), or equivalently, to \(N<N^{*}\), where the chain is completely confined _before_ overcoming the energy barrier given by eq.. From eqs. and, for a given chain length, this corresponds to weak fluxes and/or wide pores. Instead of being located at \(y^{*}\), the energy barrier corresponds to a penetration \(y=L\), given that there is no extra cost for pushing the chain into the channel any further than its own length. In terms of the number of beads, the free energy barrier obeys \(\Delta F^{*}=\Delta F(L(N))\), and follows from eq., \[\frac{\Delta F^{*}}{k_{B}T}=A\left(\frac{a}{D}\right)^{5/3}\left(1-\frac{1}{2 }\frac{N}{N^{*}}\right)N, \tag{13}\] and to write the dependence on \(N\) and \(N^{*}\) explicitly. According to this result, for a given constant value of the flux, the barrier can be reached more easily by decreasing the number of beads. Conversely, increasing \(N\) has the effect of increasing the strength of the barrier, up to \(N=N^{*}\), where one crosses over to the long-chain regime, and eq. reduces to eq. as expected. The threshold flux can be calculated from the probability of translocation, which follows after combining eqs. and. Using the criterion given by eq., we obtain \[\frac{\eta J_{c}}{k_{B}T} =2\left(\frac{A}{B}\right)\left(\frac{D}{a}\right)^{10/3} \tag{14}\] \[\times\left(\frac{\left(\frac{a}{D}\right)^{5/3}N+\frac{\log(P_{c })}{A}-\frac{\log(\kappa_{D}\tau_{m})}{A}}{N^{2}}\right).\]The scaling should be valid in the range \(N_{D}<N<N^{*}\), which corresponds to \(D<R\) and \(L<y^{*}\). In this range, the threshold flux increases monotonically with the number of monomers. This behaviour can be traced back to the \(N\)-dependent terms in eq., where the linear term, corresponding to the entropic cost, grows faster than the hydrodynamic gain, provided that \(N/N^{*}<1\). ## III Numerical method To further study the cross-over from long- to short-chain translocation discussed in Section II, we use the numerical method introduced in Ref.. This is a hybrid scheme that couples a lattice-Boltzmann fluid to a bead-spring model for the polymer chain. Our work is based on the original implementation of the code, _susp3d_, which was kindly provided by the developers. Here we present the main features of the numerical method. For a more detailed description, the reader is referred to the original papers. ### Lattice Boltzmann Method In the lattice-Boltzmann algorithm the fluid dynamics follows from the evolution of the particle velocity distribution function \(f_{i}\), which is defined for a discretised velocity set \(\{\vec{c}_{i}\}\). For a given velocity vector, \(f_{i}\) is proportional to the average number of particles moving in the direction of \(\vec{c}_{i}\). The dynamics of \(f_{i}\) is given by the lattice-Boltzmann equation, \[f_{i}(\vec{r}+\vec{c}_{i}\Delta t,t+\Delta t)-f_{i}(\vec{r},t)=\Delta_{i}(\vec {r},t)+F_{i}^{\rm ext}, \tag{15}\] Here we use the D3Q19 model, which consists of a cubic lattice with a set of nineteen velocity vectors in three dimensions. The lattice spacing, \(\Delta x\), is uniform along the lattice axes. The model has three possible magnitudes of the velocity vectors, \(\{|\vec{c}_{i}|\}=(0,1,\sqrt{2})\Delta x/\Delta t\). Accordingly, the \(f_{i}\) have a corresponding weight \(a_{c_{i}}\) that satisfies the condition \(\sum_{i}a_{c_{i}}=1\). A suitable choice of the weights for the lattice model used here is \(a_{0}=1/3\), \(a_{1}=1/18\) and \(a_{\sqrt{2}}=1/36\). The dynamics expressed by eq. is composed of two steps. First, the distribution function undergoes a collision step, where fluid particles exchange momentum according to the collision operator, \(\Delta_{i}\), and are driven by the term \(F_{i}^{\rm ext}\), which plays the role of a body force. Following this collision stage, the \(f_{i}\) are propagated to neighbouring sites in a streaming step, corresponding to the left-hand side of the equation. The mapping between the lattice-Boltzmann scheme and the hydrodynamic equations follows from the definition of the hydrodynamic variables as moments of the \(f_{i}\). \[\sum_{i}f_{i}=\rho,\quad\sum_{i}f_{i}\vec{c}_{i}=\rho\vec{v},\quad\mbox{and} \quad\sum_{i}f_{i}\vec{c}_{i}\vec{c}_{i}=\Pi, \tag{16}\] In the absence of external forces, the system relaxes towards equilibrium through the collision stage in eq.. For the lattice-Boltzmann model used in this paper, this is done by defining the post-collision distribution function, \(f_{i}^{*}=f_{i}+\Delta_{i}\), which can be expressed as an expansion in the fluid velocity: \[f_{i}^{*}=a_{c_{i}}\left(\rho+\frac{\rho\vec{v}\cdot\vec{c}_{i}}{c_{s}^{2}}+ \frac{(\rho\vec{v}\vec{v}+\Pi^{\rm neq,*}):(\vec{c}_{i}\vec{c}_{i}-c_{s}^{2} \mathbf{1})}{2c_{s}^{4}}\right), \tag{17}\] Similarly, the forcing term, \(F_{i}^{\rm ext}\), can be expanded in powers of the fluid velocity (cf. Ref.). The post-collisional momentum flux tensor, \(\Pi^{\rm neq,*}\), describes the relaxation towards the equilibrium momentum flux tensor, \(\Pi^{\rm eq}\), according to \[\Pi^{\rm neq,*}=(1+\lambda)\bar{\Pi}^{\rm neq}+\frac{1}{3}(1+\lambda_{\nu})( \Pi^{\rm neq}:\mathbf{1})\mathbf{1}, \tag{18}\] The parameters \(\lambda\) and \(\lambda_{\nu}\) characterise the relaxation timescales of the lattice-Boltzmann fluid. By performing a Chapman-Enskog expansion of eq., corresponding to the limit of long length and timescales compared to the lattice spacing and the relaxation timescale of the fluid, the lattice-Boltzmann scheme leads to the continuity equation for the fluid density and the Navier-Stokes equations with second order corrections in the velocity for the fluid momentum. \[\eta=-\rho c_{s}^{2}\Delta t\left(\frac{1}{\lambda}+\frac{1}{2}\right), \tag{19}\] and \[\eta_{\nu}=-\frac{2\rho c_{s}^{2}}{3}\Delta t\left(\frac{1}{\lambda}+\frac{1} {2}\right), \tag{20}\] Solid boundaries in the simulation box are implemented using the well-known bounce-back rules. These correspond to a reflection of any distribution function propagating to a solid node back to the fluid node it came from at the streaming stage in eq.. As a consequence, a stick condition for the velocity is recovered approximately halfway between the fluid node and the solid node. ### Polymer chain The linear polymer is modelled by a chain composed of \(N\) beads joined by freely rotating bonds. \[U_{\rm el}(r)=k(r-a)^{2}, \tag{21}\] The short-range excluded-volume interactions are modelled by a truncated DLVO potential, \[U_{\rm DLVO}=U_{0}\frac{\exp(-\kappa_{\rm DH}r)}{r}, \tag{22}\] The position vector of the \(i\)-th polymer bead, \(\vec{x}_{i}\), evolves in time according to \[m\ddot{\vec{x}}_{i}=-\sum_{j\neq i}\vec{\nabla}_{ij}U+\vec{F}_{i}, \tag{23}\] The term \(\vec{F}_{i}\) contains the viscous hydrodynamic force that couples the bead to the lattice-Boltzmann fluid, \[\vec{F}_{i}=-\xi_{0}(\vec{x}_{i}-\vec{v}(\vec{x}_{i}))+\vec{F}_{i}^{\rm r}. \tag{24}\] This expression includes the Stokes drag, \(\xi_{0}=6\pi\eta r_{\rm H}\), with \(r_{H}\) being the hydrodynamic radius of the beads, and a random force \(F_{i}^{\rm r}\) that satisfies the fluctuation-dissipation relation, \[\langle\vec{F}_{i}^{\rm r}(t)\vec{F}_{i}^{\rm r}(t^{\prime})\rangle=2k_{B}T\xi _{0}\delta(t-t^{\prime})\mathbf{1}. \tag{25}\] While the polymer beads move in continuous space, the lattice-Boltzmann fluid is only defined at the lattice nodes. In order to calculate the coupling force, \(\vec{F}_{i}\), the model uses a linear interpolation scheme to estimate the fluid velocity, \(\vec{v}\), at the position of the bead, \(\vec{x}_{i}\). The discretisation of the lattice gives an _effective_ hydrodynamic radius, \(r_{\rm H}^{\rm eff}\), which differs from the input hydrodynamic radius, \(r_{\rm H}\). Here we follow the same procedure as in Ref., and choose \(r_{\rm H}\) in order to obtain the desired value of \(r_{\rm H}^{\rm eff}\). Once the hydrodynamic force has been exerted on the monomer, momentum conservation is enforced by exerting a force of equal magnitude back onto the fluid. ### Parameter Values Our objetive is to carry out numerical simulations of polymer chains in the limit where the bead-spring model gives the blob-model behaviour presented in Section II. Such regime has been validated when the monomer size matches the blob size, \(a\simeq\xi\), for linear chains, where \(\xi\simeq D\). Therefore, the bead-spring model should give the blob-limit behaviour for \(a\simeq D\). On the other hand, the hydrodynamic radius of the beads, \(r_{\rm H}^{\rm eff}\), is limited to small values, corresponding to the point-particle coupling between the chain and the lattice-Boltzmann fluid. With these conditions in mind, we fix the bead diameter to \(2r_{\rm H}^{\rm eff}=\Delta x/2\), while \(a=\Delta x\) and \(D=2\Delta x\). The remaining model parameters are chosen as follows: we work at a fixed temperature, \(k_{B}T=0.1\). To prevent chain crossings the spring constant is taken as \(k=300k_{B}T/\Delta x^{2}\). Parameter values for the excluded-volume potential are fixed to \(U_{0}=k_{B}T\Delta x\) and \(\kappa_{\rm DH}=80/\Delta x\), which ensure that inter-monomer repulsions are larger than \(k_{B}T\) at distances comparable to \(r_{\rm H}^{\rm eff}\). To match the hydrodynamic diameter to the effective bead diameter, we fix \(r_{\rm H}=0.32\Delta x\) following the calibration procedure presented in Ref.. In order to resolve the hydrodynamics correctly, one needs to ensure that the distribution function relaxes on a faster timescale than the diffusive timescale of the polymer chain. This condition can be satisfied by setting the parameters in the collision operator in eq. to \(\lambda=-1\) and \(\lambda_{\nu}=-1\). We set the fluid density to \(\rho=36\), and the timestep and lattice spacing in the lattice-Boltzmann fluid to \(\Delta t=1\) and \(\Delta x=1\). Using these values, the dynamic viscosity in simulation units is \(\eta=6\). ## IV Numerical results ### System Geometry and Initial Conditions The geometry of the system is depicted in We consider two identical rectangular ducts of dimensions \(L_{x}=108\), \(L_{y}=14\) and \(L_{z}=10\), separated by a square pore of width \(D=2\) and length \(L_{p}=24\). Solid walls are imposed in the \(y\) and \(z\) directions, while periodic boundary conditions are enforced in the \(x\) direction. For small Reynolds numbers, one expects that the volumetric flux, \(J\), caused by applying a uniform body force Schematic representation of the simulation box. to the fluid, \(f\), obeys Darcy's Law: \[J=\frac{kS}{\eta}f, \tag{26}\] In order to verify the validity of eq., we measure the local flux for different applied forcings. The results, illustrated in Fig. 3, show a linear dependence of \(J\) on \(f\) as expected. A linear fit to the data gives \(kS\simeq 238\). For a given forcing, we find that the flux does not vary significantly whether it is measured inside or outside the pore thus confirming that, for the range of forcings considered here, the lattice-Boltzmann fluid behaves as an incompressible liquid. Given that we are interested in the scaling of the critical flux, we use values of the forcing that give a ratio between hydrodynamic and thermal effects in the range \(0<\eta J/k_{B}T<1.5\). We initially tether the polymer chain at a position \(y=y_{0}\) inside the channel as depicted in the inset of Fig. 4, and let the system equilibrate for 5\(\times 10^{4}\) time steps. The equilibration period allows the relaxation of the chain. Evidence for this is presented in the same figure, which shows the number of beads inside the pore, \(N_{y_{0}}\), as a function of \(y_{0}\) for \(N=128\). For a confined chain in equilibrium, we have, from eq., \(N_{y_{0}}\simeq(y_{0}/D)(D/a)^{5/3}\). Our results show that the number of beads increases linearly with the tethering position, as expected. ### Energy Barrier In order to confirm that the polymer translocation is controlled by surpassing an energy barrier, we first perform simulations of long polymer chains subject to a fixed driving flux while varying the initial tethering position, \(y_{0}\). We consider different values of \(y_{0}\) in the range \(0<y_{0}<L_{p}\), and set the number of monomers to \(N=128\), thereby ensuring that the major part of the chain lies outside the pore. Simulations are carried out by letting the chain equilibrate as before. Once the chain has equilibrated, it is released from its tethering point and the body force is applied uniformly to the fluid. The chain then is either carried down the pore by the underlying fluid flow and eventually translocates to the opposite duct, or is ejected from the pore back into the original chamber. Simulations are run for \(5\times 10^{5}\) timesteps, which is a sufficiently long timescale to identify successful or failed translocation events. We observe a clear transition from non-translocating (\(P=0\)) to translocating (\(P=1\)) chains as \(y_{0}\) is increased. This confirms the presence of an energy barrier, which is progressively approached as the chain is pushed further into the channel. The probability curves are shifted to the left as one increases the imposed flux, \(J\). This indicates a shift of the position of the barrier, \(y^{*}\), closer to the pore entrance caused by a higher driving hydrodynamic force. As a criterion we define the position of the barrier as \(P(y^{*})\approx 1\), and plot the measured values of \(y^{*}\) as a function of \(k_{B}T/\eta J\) in the inset of As expected from eq., we observe a linear growth. A fit of the data shown in the figure to the function \(y^{*}(x)=(A/B)Dx\) gives an estimate of the numerical prefactor in eq., \(A/B\simeq 4.4\). The imposed flux sets the position and height of the energy barrier. Given that we fix the chain to a prescribed position inside the channel, the height of the barrier is reduced by an amount determined by \(y_{0}\). Volumetric flux, \(J\), as a function of the body force, showing that Darcy’s law holds. Number of beads in the confined part of the chain as a function of the initial tethering position. Inset: schematic representation of the tethered chain. \(y-y_{0}\) in eq. and performing the integration in eq. from \(y^{\prime}=y_{0}\) to \(y^{\prime}=y\). \[\frac{\Delta F^{*}}{k_{B}T}=\frac{A}{2}\left(\frac{y^{*}}{D}\right)\left(1+g \left(\frac{y_{0}}{y^{*}}\right)\right). \tag{27}\] we can calculate the probability of translocation. This predicts \(\log P\sim\Delta F^{*}/k_{B}T\) as observed in As expected, all data points collapse onto the same curve, which is linear in the free energy. A fit to the data gives an estimate for the numerical prefactor in eq., \(A\simeq 1.8\), which together with the measured value of \(A/B\), gives the amplitude of the hydrodynamic contribution to the free energy in eq., \(B\simeq 0.41\). The inset in shows how the theoretical prediction accurately captures the main features of the translocation process. The probability of translocation increases to unity as \(y_{0}/y^{*}\to 1\). The spread of the curves is dictated by the \(\sim k_{B}T/\eta J\) prefactor in eq., increasing the likelihood of translocation at a given \(y_{0}/y^{*}\) for larger fluxes, or lower temperatures. ### Cross-over from long to short chains So far we have discussed the translocation of long chains, which can reach the position of the barrier while a significant amount of monomers remain outside the pore. The threshold flux is thus controlled only by the number of monomers that it takes to reach the barrier, \(N^{*}\), and not by the total number of monomers in the chain, \(N\). This picture breaks down for shorter chains, where our theory predicts that the scaling of the critical flux becomes dependent on \(N\). In order to explore this effect, we have carried out simulations at a fixed initial tethering position, \(y_{0}\), and driving flux, while decreasing the total number of beads in the chain, \(N\). We fix the initial tethering position to \(y_{0}=L_{p}/2=12\), for which the number of beads inside the pore after equilibration is \(N_{y_{0}}\simeq 18\). We carry out the simulations in the range \(16<N<64\), where we expect to observe the cross-over. The probability always increases with the applied flux as expected. Logarithm of the translocation probability as a function of the free energy barrier magnitude. Symbols correspond to numerical simulations. The solid line is a linear fit to the data. Inset: translocation probability as a function of \(y_{0}/y^{*}\). Numerical results (symbols) show a good agreement with the theoretical prediction (solid lines). Probability of translocation as a function of the number of beads in the chain at different values of the imposed flux. Simulation results correspond to symbols, while the theoretical model is indicated by the solid lines where we set \(N_{y_{0}}=20\) and \(A=1.8\). Probability of translocation as a function of \(y_{0}\). Inset: linear growth of the position of the energy barrier with \(k_{B}T/\eta J\). For smaller \(N\) the probability of translocation increases strongly with decreasing \(N\) until it saturates at \(N\simeq 20\) to \(P\to 1\), where all polymer chains translocate even under the weakest flow. This behaviour is in agreement with our theoretical model. For large \(N\), the critical number of beads to overcome the barrier follows from eq., and obeys \(N^{*}\sim(D/a)^{5/3}(k_{B}T/\eta J).\) Therefore, the cross-over to the short-chain regime (\(N<N^{*}\)) is observed at larger chain lengths as \(J\) is decreased. From eq. we also have \(N^{*}\sim y^{*}\), from which it is possible to estimate the critical number of beads corresponding to each driving flux by interpolating the data shown in Once each value of \(N^{*}\) is known, we can determine the energy barrier as a function of \(N\) from eq.. In order to include the effect of \(y_{0}\), we follow the same procedure as that used to obtain eq.. \[\frac{\Delta F^{*}}{k_{B}T}=A\left(\frac{a}{D}\right)^{5/3}\left[N\left(1- \frac{1}{2}\frac{N}{N^{*}}\right)+\frac{1}{2}N^{*}g\left(\frac{N_{y_{0}}}{N^ {*}}\right)\right]. \tag{28}\] Using this expression for the free energy barrier an estimate for the probability of translocation follows from eq.. The observed saturation length \(N\simeq 20\) is indeed in good agreement with our estimate \(N_{y_{0}}\simeq 18\), at which the free energy barrier given by eq. vanishes. As shown in Fig. 7, we obtain a good agreement between the theory and simulations. ### Critical flux for small chains We now turn our attention to the dependence of the critical flux, \(J_{c}\), on the number of monomers in the short-chain regime. To find the threshold flux, we carry out simulations of the translocation process at a fixed number of beads while increasing the flux. We fix \(y_{0}=12\) as before, and consider the range \(0<\eta J/k_{B}T<1\), for which the probability curves cross over from complete rejection of the chains to complete translocation. For \(N>32\) the \(N\)-independence of the long-chain regime is recovered, as curves fall on top of each other. Our results in the long-chain regime are consistent with recent experimental studies of polymer translocation in nanochannels carried out by Beguin _et al._. In their experiments they measure the rejection coefficient of chain translocation, which is related to the probability of translocation by \({\cal R}\sim 1-P\). Their experiments give the same smooth transition from rejection to translocation of the chains in the range \(0<\eta J/k_{B}T<1\), very close to our simulation results. For \(N<32\) the probability curves shift upwards and become increasingly plateau-like as the number of monomers is decreased, indicating the cross-over to the short-chain regime. In order to quantify the cross-over from the short- to the long-chain regime, we use the criterion given by eq. to interpolate \(J_{c}\) from each of the curves shown in The threshold flux is shown in as a function of \(N\) for six different values of the threshold probability. As expected, the overall behaviour does not depend on the particular choice of \(P_{c}\). The threshold flux saturates for large \(N\), corresponding to the long-chain regime, and shows a marked decrease as \(N\to N_{y_{0}}\), where the chain becomes completely confined inside the channel. Using eqs., and, we can estimate the threshold flux as a function of the number of monomers in the chain, \[\frac{\eta J_{c}}{k_{B}T} \simeq 2\frac{A}{B}\left(\frac{D}{a}\right)^{10/3}\] \[\times\left(\frac{\left(\frac{a}{D}\right)^{5/3}(N-N_{y_{0}})+ \frac{\log(P_{c})}{A}-\frac{\log(\kappa_{D}T_{m})}{A}}{N^{2}-N_{y_{0}}^{2}} \right).\] Both prefactors, \(A\) and \(B\), depend on the intrinsic properties of the chain and not on the particular translocation regime. We therefore set them to the values that we have measured previously. Remarkably, we find a good quantitative agreement by using no new fit parameters in our theoretical prediction. ## V Summary and discussion The results presented in this paper show that the translocation of polymer chains through narrow pores exhibits two different regimes depending on the length of the chain. As previously proposed in Ref., our numerical results show that the translocation process is controlled by overcoming a free energy barrier characterised by a critical penetration of the polymer inside the pore, \(y^{*}\). For long chains, this length scale always outruns the polymer length under confinement, \(L\). Probability of translocation as a function of the imposed flux for different number of monomers in the chain. Conversely, for small chains, where the confined length of the polymer is smaller than \(y^{*}\), we have shown that the translocation process is controlled by \(L\), and the process becomes \(N\)-dependent. A qualitatively similar conclusion should apply to the passage of the chains through two dimensional narrow slits. For even smaller chains, where \(L\) is comparable to the size of the pore, \(D\), and for very small fluxes (\(\eta J/k_{B}T\simeq 10^{-1}\)), we have observed a reduction of the probability of translocation of the chains. This is caused by the effect of diffusion of the chain inside the pore, which can favour the ejection of small chains at very low forcings. However, this regime falls out of the free energy barrier picture presented in this paper, and we therefore leave it for future exploration. An application of the dependence of the translocation probability of short chains on \(N\) is the potential sorting of the chains according to their number of monomers. In the short-chain regime the polymer must be completely pushed in before translocating through the pore. Given that only a small number of blobs can be pushed into the pore by fluctuations, one expects that only very short chains translocate by virtue of thermal effects in this regime. For longer chains, but still in the short-chain regime, the translocation must be assisted, for example, by a molecular motor that is able to push the chain deeper into the pore. In an aqueous solution at room temperature, the critical flux in the long chain regime follows the scaling of eq. and is of order \(J_{c}\sim 10^{-18}\) m\({}^{3}\) s\({}^{-1}\). For a micrometre sized channel the corresponding velocity, \(v_{c}\simeq J_{c}/D^{2}\), is of order \(10^{-6}\) m s\({}^{-1}\). Such a small value suggests that chains can easily translocate through microfluidic chambers at normal operation velocities, which are typically of centimetres to metres per second. However, recent experimental measurements of cytoplasmic streaming in micrometre cell channels show that the streaming velocities are in the range of 10 \(\mu\)m s\({}^{-1}\). For weaker fluxes than \(J_{c}\), one enters the short-chain translocation regime. Therefore, it is feasible that the translocation regime proposed in this paper might be at play in biological systems as a regulator of selective translocation. Here we have considered polymers in a good solvent using the DLVO potential to implement short-ranged excluded-volume interactions with the wall. Understanding the details of electrostatic effects, with regards both to the interaction with the wall and the interactions between segments, and how they are affected by the salt concentration, is an important issue for future work, relevant, in particular, to the dynamics of DNA in practical applications. Finally, we comment on the feasibility of an experimental confirmation of our prediction. Beguin _et at._ have recently performed an experimental study of the flow-driven translocation of hydrosoluble polymers into nanopores. In their experiments, they consider chains whose radius of gyration is \(R\approx 84\) nm, and which are forced across pores \(23-35\) nm in radius and 6.5 \(\mu\)m in length. Their results for the rejection coefficient, \({\cal R}=1-\exp{(\Delta F^{*}/k_{B}T)}\), show a good agreement with the long-chain limit of the de Gennes model, where they use a similar expression to eq.. According to our prediction, the short chain regime would be observable for chains whose length under confinement is smaller than the position of the barrier. This traduces into a cross-over radius of gyration \(R^{*}\simeq D(k_{B}T/\eta J)^{3/5}\). Taking \(D=70\) nm and \(\eta J/k_{B}T\approx 0.8\), which correspond to experimental conditions reported in ref., we estimate \(R^{*}\approx 80\) nm, which is close to the radius of gyration of the polymers used in the experiments. This supports the feasibility of future experimental work to verify the theoretical predictions presented in this paper.
10.48550/arXiv.1110.5423
Length-dependent translocation of polymers through nanochannels
Rodrigo Ledesma-Aguilar, Takahiro Sakaue, Julia M. Yeomans
4,204
10.48550_arXiv.2305.08835
"###### Abstract\n\nWe describe a multiple electronic state adaptation of the mapping approach to su(...TRUNCATED)
10.48550/arXiv.2305.08835
A multi-state mapping approach to surface hopping
Johan E. Runeson, David E. Manolopoulos
1,918
10.48550_arXiv.1208.0397
"###### Abstract\n\nWe report a study of the structure of droplets of colloidal gels containing diss(...TRUNCATED)
10.48550/arXiv.1208.0397
Aggregation Patterns of Salt Crystalizing in Drying Colloidal Solvents
Moutushi Dutta Choudhury, Sayanee Jana, Sruti Dutta, Sujata Tarafdar
5,250
10.48550_arXiv.1209.0195
"###### Abstract\n\nWe calculate accurate eigenvalues and eigenfunctions of the Schrodinger equation(...TRUNCATED)
10.48550/arXiv.1209.0195
Bound states for the quantum dipole moment in two dimensions
Paolo Amore, Francisco M Fernández
3,558
10.48550_arXiv.2411.00990
"### Active-space\n\nModeling electronic structure problems with quantum computing is a complex task(...TRUNCATED)
10.48550/arXiv.2411.00990
"Exploiting the Variational Quantum Eigensolver for Determining Ground State Energy of Protocatechui(...TRUNCATED)
"Gleydson Fernandes de Jesus, Erico Souza Teixeira, Lucas Queiroz Galvão, Maria Heloísa Fraga da S(...TRUNCATED)
4,743
10.48550_arXiv.2407.18842
"## Appendix C RT-NEO-TDDFT Ehrenfest Dynamics with Traveling Proton Basis\n\n#### c.3.1 Traveling P(...TRUNCATED)
10.48550/arXiv.2407.18842
"Lagrangian Formulation of Nuclear-Electronic Orbital Ehrenfest Dynamics with Real-time TDDFT for Ex(...TRUNCATED)
Jianhang Xu, Ruiyi Zhou, Tao E. Li, Sharon Hammes-Schiffer, Yosuke Kanai
6,619
End of preview. Expand in Data Studio
README.md exists but content is empty.
Downloads last month
17