The Journal of Chemical Physics

The Journal of Chemical Physics

Published by AIP Publishing

Online ISSN: 1089-7690

·

Print ISSN: 0021-9606

Disciplines: Chemistry; Chemistry, Physical and theoretical; Chimie; Chimie physique et théorique; Chimie physique et théorique; Fysische chemie; Physics; Physique

Journal websiteAuthor guidelines

Top-read articles

130 reads in the past 30 days

Representative XPS spectrum for the Ni 2p XPS of Ni(OH)2; the dotted curve represents the measured XPS, while the solid curve is a theoretical modeling of the XPS; the main 2p3/2 and 2p1/2 peaks are labeled.
Schematic models for (a) the Ni(OH)6 cluster model of the Ni(OH)2 crystal shown in panel (b); see the text for details.
Comparison of the theoretical prediction of the Ni(2p) XPS for Ni(OH)2 modeled with the Ni(OH)6 cluster at the OSA level of theory represented by the solid curve, with the measurement represented by the dotted curve with contributions from individual theoretical ionic multiplets in color. The main 2p3/2 and 2p1/2 peaks are labeled. See the text for details.
Comparison of the theoretical prediction of the Ni(2p) XPS for Ni(OH)2 at the OCSA level of theory with measurement; see the caption of Fig. 3 and the text for details.
Comparison of the theoretical prediction of the Ni(2p) XPS for Ni(OH)2 at the OCSA shake optimized level of theory with measurement; see the caption of Fig. 3 and the text for details.

+3

Chemical information from XPS: Theory and experiment for Ni(OH)2

October 2024

·

204 Reads

·

·

Sebastian T. Mergelsberg

·

[...]

·

Download

Aims and scope


The Journal of Chemical Physics publishes cutting edge research in all areas of modern physical chemistry and chemical physics. The Journal also publishes brief communications of significant new findings, perspectives on the latest advances in the field, and Special Topics.

Recent articles


Schematic of different approaches to rare event sampling in molecules. The objective is to calculate the kinetics of going from state A to state B. In enhanced sampling approaches such as OPES flooding, external bias (solid green shade) is deposited in the initial state basin. In path sampling methods like weighted ensemble, multiple trajectory segments (blue arrows) are generated through a resampling procedure. In the integrated sampling method, both are performed simultaneously.
In weighted ensemble, many trajectory segments sample the conformational space. In integrated sampling, the kinetics is only computed by tracing a successful transition back to its origin in the initial state. One of the two such traces is depicted in a gray dashed line. The unbiased transition time is then obtained by reweighting the timescale of the trace based on the external bias deposited along its path.
Convergence of the mean first passage time of the C7eq to C7ax transition in alanine dipeptide. The uncertainty (light blue shade) is computed as the 95% confidence interval from ten independent sets of integrated sampling (IS) simulations. The reference MFPT and its uncertainty are shown in black dashed and dotted lines, respectively.
(a) Bias deposition scheme along the HLDA CV for sampling the unfolding transitions in chignolin mini-protein. (b) Representative unfolding trajectories, sampled using OPES-flooding and integrated sampling, projected along the HLDA CV. Representative trajectories were chosen as the ones with rescaled transition time closest to the estimated MFPT from all trajectories. In the case of integrated sampling, the simulation time axis is equivalent to the molecular time in the weighted ensemble. (c) Convergence of the unfolding time obtained from integrated sampling. The uncertainty (light blue shade) is computed as the 95% confidence interval from three independent simulations. The unfolding time from the unbiased reference simulation is 2.2 ± 0.4 μs. The reference MFPT and its uncertainty are shown in black dashed and dotted lines, respectively.
(a) Bias deposition scheme along the z-axis CV for sampling the dissociation of the OAMe-G4 host guest complex. The biasing scheme is designed blindly by only depositing bias in the bound state minimum and not the intermediate. (b) Convergence of the ligand residence time of G4 obtained from integrated sampling. The uncertainty (light blue shade) is computed as the 95% confidence interval from the three independent simulations. The 95% confidence interval of the 30 OPESf simulations (∼500 ns) is 0.88–1.81 μs (estimated using the method proposed by Kaminsky⁵⁴). The reference MFPT from OPES-flooding and its uncertainty are shown in black dashed and dotted lines, respectively. (c) Projection of the trajectory with the highest weight on the 2D free energy surface along the biasing CV (z) and the Deep-LDA CV trained to model the water hydration behavior.⁵² Although we did not bias the water, the most probable trace of the integrated sampling algorithm sampled the minimum free energy pathway in the 2D space. Similar plots for the 10 most probable trajectory traces are provided in the supplementary material.
Integrating path sampling with enhanced sampling for rare-event kinetics
  • Article
  • Publisher preview available

December 2024

·

13 Reads

Studying the kinetics of long-timescale rare events is a fundamental challenge in molecular simulation. To address this problem, we propose an integration of two different rare-event sampling philosophies: biased enhanced sampling and unbiased path sampling. Enhanced sampling methods, e.g., metadynamics, can facilitate the crossing of free energy barriers by applying an external bias potential. On the contrary, path sampling methods like weighted ensemble do not apply any biasing force but accelerate the exploration of rugged free energy surfaces through trajectory resampling. We show that a judicious combination of the weighted ensemble with a metadynamics-like algorithm can synergize the strengths and mitigate the deficiencies of path sampling and enhanced sampling approaches. The resulting integrated sampling (IS) algorithm improves the computational efficiency of calculating the kinetics of peptide conformational transitions, protein unfolding, and the dissociation of a ligand–receptor complex. Furthermore, the IS approach can direct sampling along the minimum free energy pathway even when the collective variable used for biasing is suboptimal. These advantages make the IS algorithm suitable for studying the kinetics of complex molecular systems of biological and pharmaceutical relevance.


(a) Exact survival function for a hyperexponential distribution [Eq. (6)]. The dashed gray lines show slopes of −k1 and −k2 for comparison. The dotted black line highlights a specific T* = 0.115. (b) Speedup (green) and estimated MFPT (blue) as a function of the timer using Eqs. (3) and (4) and the analytical survival function. The dotted black line highlights a timer of T* = 0.115.
Estimated MFPT (top panel) and speedup (bottom panel) as a function of T* for the hyperexponential distribution. The circles, horizontal lines, boxes, and whiskers in the top panel show the mean, median, IQR, and extreme values within 1.5 IQR of the first and third quartiles, respectively. The shaded region shows the IQR of 1000 batches of 100 unbiased simulations each. The speedup was calculated for a single processor (green squares) and 100 processors (orange triangles).
The Pareto distribution: Estimated MFPT (top) and estimated α (bottom) as a function of the number of first-passage samples. The dashed black lines give the analytic MFPT and α in the corresponding panels. The circles, horizontal lines, boxes, and whiskers show the mean, median, IQR, and extreme values within 1.5 IQR of the first and third quartiles, respectively. The dotted lines and the shaded regions show the mean and IQR, respectively, for 1000 sets of 2000 unbiased measurements each.
(a) The three-state system. The white dashed line marks the first-passage criterion. (b) The survival function at times <10ns. The dashed line marks the time t=1k1+k−1 and the dotted line shows the decay at rate κ=k−1k1+k−1k2. (c) Estimated MFPT (top), timescale t′ (center), and speedup (bottom) as a function of the timer. Speedup was calculated for simulations on a single processor (green squares) or on 100 parallel processors (orange triangles). The circles, horizontal lines, boxes, and whiskers show the mean, median, IQR, and extreme values within 1.5 IQR of the first and third quartiles, respectively. The dotted lines and shaded areas show the mean values and IQR for 100 bootstrapping sets of 100 unbiased simulations each, respectively (1000 independent unbiased trajectories were collected in total). The dashed line in the middle panel shows t=1k1+k−1.
(a) Ball-and-stick and cartoon representations of four stable configurations of a nine-residue alanine peptide. The white, gray, blue, and red spheres represent the hydrogen, carbon, nitrogen, and oxygen atoms, respectively. (b) Free energy along an end-to-end-based CV and an H-bond-based CV. The white dotted lines define the four metastable states. (c) Estimated MFPT (top) for different timers. The MFPT of 1000 unbiased trajectories is indicated with a black dotted line in the top panel, with the gray shading showing ±1/1000 the standard deviation—the estimated error. The circles, horizontal lines, boxes, and whiskers show the mean, median, IQR, and extreme values within 1.5 IQR of the first and third quartiles, respectively. The speedup using a single or 100 processors is plotted in the bottom panel with green squares and orange triangles, respectively.
Inference of non-exponential kinetics through stochastic resetting

December 2024

·

1 Read

We present an inference scheme of long timescale, non-exponential kinetics from molecular dynamics simulations accelerated by stochastic resetting. Standard simulations provide valuable insight into chemical processes but are limited to timescales shorter than ∼1μs. Slower processes require the use of enhanced sampling methods to expedite them and inference schemes to obtain the unbiased kinetics. However, most kinetics inference schemes assume an underlying exponential first-passage time distribution and are inappropriate for other distributions, e.g., with a power-law decay. We propose an inference scheme that is designed for such cases, based on simulations enhanced by stochastic resetting. We show that resetting promotes enhanced sampling of the first-passage time distribution at short timescales but often also provides sufficient information to estimate the long-time asymptotics, which allows the kinetics inference. We apply our method to a model system and a peptide in an explicit solvent, successfully estimating the unbiased mean first-passage time while accelerating the sampling by more than an order of magnitude.


Excited states from GW/BSE and Hartree–Fock theory: Effects of polarizability and transition type on accuracy of excited state energies

December 2024

·

6 Reads

GW and Bethe–Salpeter equation (BSE) methods are used to calculate energies of excited states of organic molecules in the Quest-3 database [Loos et al., J. Chem. Theory Comput. 16, 1711 (2020)]. The self-energy in the GW approximation is conventionally calculated using the RPA polarizability. Inclusion of a screened electron–hole interaction in the polarizability was recently shown to improve predictions of experimental ionization energies in organic molecules [C. H. Patterson, J. Chem. Theory Comput. 20, 7479 (2024)]. Self-energies from RPA or screened time-dependent Hartree–Fock (TDHF) polarizabilities in the GW/BSE method are used to calculate 141 singlet excited states in Quest-3. Theoretical best estimate excited state energies from the CC3 coupled cluster method and aug-cc-pVTZ basis sets are used to benchmark GW/BSE and CIS calculations using the same molecular geometries and basis sets. Differences between GW/BSE or CIS excited state energies and best estimate values show that there are systematic variations in the accuracies of excited state energies classified as ππ*, nπ*, πR (Rydberg), or nR character. The origin of these variations is the accuracy of self-energies of states of nonbonding vs π bonding character. In particular, N or O lone pair states require large self-energy corrections owing to strong orbital relaxation in the localized hole state, while π states have smaller corrections. Self-energies from a screened TDHF vs RPA polarizability are typically over(under)estimated for nonbonding states, leading to under(over)estimation of energies of excited states of nπ* or nR character.


Definition of population-based nonlinear signals and quantities in nested and cogwheel phase cycling. (a) Schemes of a three-pulse sequence with time delays τ and t (left) and a four-pulse sequence with delays τ, T, and t (right). Nonlinear signals are described by the absolute coherence orders X, Y, and Z (where “Q” denotes “quantum”) and a set of phase coefficients α, β, γ, and δ of the individual pulse phases φ1, φ2, φ3, and φ4, respectively. In nested (NES) phase cycling, pulse phases are incremented sequentially according to an A × B × C × D scheme, whereas in cogwheel (COG) phase cycling, all pulse phases are varied simultaneously but with pulse-specific increments. (b) Visualization of 27-fold (1 × 3 × 3 × 3) nested phase cycling and (c) 20-fold cogwheel phase-cycling “COG20(0, 1, 6, 10).”
Cogwheel phase cycling vs nested phase cycling in the simulation of an exemplary three-pulse experiment. (a) Pulse sequence (black) and signal phase coefficients (blue) of the rephasing 2Q–1Q signal with probed coherences depicted in gray. (b) Energy level scheme of the model system with ground (0), singly excited (1 and 2), and doubly excited (3) states that are connected by transition dipole moments μfi, where f denotes the final and i the initial state, and which all have the same magnitude (red arrows). The laser spectrum (red Gaussian) covers the transition energies of the 0–1 and the 0–2 transitions equally. (c) Real part of the rephasing 2Q–1Q 2D spectrum in case of 21-fold nested phase cycling and (d) 18-fold cogwheel phase cycling. (e) Difference spectrum, obtained from subtracting the spectrum of (d) from that of (c). All 2D spectra are drawn with nine linearly spaced contour lines.
The laser spectra for the NES27(1, 3, 3, 3) measurement (shaded orange area) as well as the COG20(0, 1, 6, 10) and COG20(0, 1, 10, 6) measurements (shaded purple area) are shown with the linear absorption spectrum of rhodamine 700 (green).
Real parts of the rephasing 1Q–1Q 2D spectra of rhodamine 700 obtained for different T delays. The upper row shows 2D spectra acquired by NES27(1, 3, 3, 3) at (a) T = 0 fs, (b) T = 80 fs, and (c) T = 280 fs. The lower row displays 2D spectra acquired by COG20(0, 1, 6, 10) at (d) T = 0 fs, (e) T = 80 fs, and (f) T = 280 fs. Integrating the signal amplitude within a region of interest (a square between 1.864 and 1.920 eV, not shown) for each T yields oscillations with a period of 140 fs in both (g) NES27(1, 3, 3, 3) and (h) COG20(0, 1, 6, 10). All 2D spectra are drawn with six linearly spaced contour lines.
Real parts of the nonrephasing 1Q–1Q 2D spectra of rhodamine 700 obtained for different T delays. The upper row shows 2D spectra acquired by NES27(1, 3, 3, 3) at (a) T = 0 fs, (b) T = 80 fs, and (c) T = 280 fs. The lower row shows 2D spectra acquired by COG20(0, 1, 10, 6) at (d) T = 0 fs, (e) T = 80 fs, and (f) T = 280 fs. All 2D spectra are drawn with six linearly spaced contour lines.
Cogwheel phase cycling in population-detected optical coherent multidimensional spectroscopy

December 2024

·

11 Reads

An integral procedure in every coherent multidimensional spectroscopy experiment is to suppress undesired background signals. For that purpose, one can employ a particular phase-matching geometry or phase cycling, a procedure that was adapted from nuclear magnetic resonance (NMR) spectroscopy. In optical multidimensional spectroscopy, phase cycling has been usually carried out in a “nested” fashion, where pulse phases are incremented sequentially with linearly spaced increments. Another phase-cycling approach that was developed for NMR spectroscopy is “cogwheel phase cycling,” where all pulse phases are varied simultaneously in increments defined by so-called “winding numbers.” Here we explore the concept of cogwheel phase cycling in the context of population-based coherent multidimensional spectroscopy. We derive selection rules for resolving and extracting fourth-order and higher-order nonlinear signals by cogwheel phase cycling and describe how to perform a numerical search for the winding numbers for various population-detected 2D spectroscopy experiments. We also provide an expression for a numerical search for nested phase-cycling schemes and predict the most economical schemes of both approaches for a wide range of nonlinear signals. The signal selectivity of the technique is demonstrated experimentally by acquiring rephasing and nonrephasing fourth-order signals of a laser dye by both phase-cycling approaches. We find that individual nonlinear signal contributions are, in most cases, captured with fewer steps by cogwheel phase cycling compared to nested phase cycling.


A discretized representation for Monte Carlo simulation of deformed semiflexible chains

In this study, we present a novel orientation discretization approach based on the rhombic triacontahedron for Monte Carlo simulations of semiflexible polymer chains, aiming at enhancing structural analysis through rheo-small-angle scattering (rheo-SAS). Our approach provides a more accurate representation of the geometric features of semiflexible chains under deformation, surpassing the capabilities of traditional lattice structures. Validation against the Kratky–Porod chain system demonstrated superior consistency, underscoring its potential to significantly improve the precision of uncovering geometric details from rheo-SAS data. This approach opens new avenues for investigating the conformations of semiflexible polymers and mechanically induced phase transitions in more complex polymer structures, offering deeper insights into their behavior under various conditions.


Dielectric properties of nanoconfined water

The dielectric function of a dipolar liquid exhibits a strong wavenumber dependence in the bulk homogeneous state. Such a behavior seems to suggest the possibility of a strong system size dependence of the dielectric constant (DC) of a nanoconfined liquid, although details have been revealed only recently. The dielectric properties of nanoconfined water, indeed, show a marked sensitivity not only to the size and shape (dielectric boundaries) of confinement but also to the nature of surface–water interactions. For geometries widely studied, namely, water confined in a narrow slit, nanocylinder, and nanospherical cavity, the asymptotic approach to the bulk value of the DC with the increase in confinement size is found to be surprisingly slow. This seems to imply the appearance of a dipolar cross correlation length, much larger than the molecular length-scale of water. In narrow slits and narrow cylinders, the dielectric function becomes both inhomogeneous and anisotropic, and the longitudinal and transverse components display markedly different system size dependencies. This sensitivity can be traced back to the dependence of the DC on the ratio of the mean square dipole moment fluctuation to the volume of the system. The observed sensitivity of collective dipole moment fluctuations to the length scale of confinement points to the possibility of using DC to estimate the orientational correlation length scale, which has been an elusive quantity. Furthermore, the determination of volume also requires special consideration when the system size is in nanoscale. We discuss these and several other interesting issues along with several applications that have emerged in recent years.


Density as a function of molality at T = 298.15 K and 1 bar for NaCl aqueous solutions using the Madrid-2019(TIP4P/2005) (blue circles) and the Madrid-2019(TIP4P/Ice) (red squares) force fields. The solid black line is a fit of the experimental data taken from Ref. 146.
Difference between Δρ [see Eq. (4)] of the Madrid-2019(TIP4P/2005) model and Δρ of the Madrid-2019(TIP4P/Ice) model divided by the studied molality of each salt. When deviations are smaller than ±2.5, we employ circles. The triangles are for deviations between ±2.5 and ±5, and the squares are for deviations larger than ±5.
Difference between the viscosity at each concentration and the viscosity of pure water for each model as a function of concentration for aqueous NaCl solutions at 298.15 K and 1 bar. Blue circles: Madrid-2019(TIP4P/2005) force field (η0 = 0.85 mPa s). Red squares: Madrid-2019(TIP4P/Ice) force field (η0 = 1.63 mPa s). The solid line is a fit of experimental data taken from Refs. 155 and 156.
Molality of the aqueous solution phase as a function of the simulation time for different salts evaluated in this work using the Madrid-2019(TIP4P/Ice) model at 1 bar and 256 K for NaCl and KCl and 1 bar and 253 K for MgCl2. Black circles: results for NaCl. Red squares: results for KCl. Blue diamonds: results for MgCl2.
Freezing point depression (at 1 bar) for the different salt aqueous solutions evaluated in this work. (a) ΔT [i.e., Tf(m) − Tf(m = 0) with Tf(m = 0) being the freezing temperature of ice Ih, 273.15 K from experiments, 270 K for the TIP4P/Ice model, and 250 K for the TIP4P/2005] as a function of concentration of an NaCl solution. (b)–(d) Freezing temperature Tf(m) as a function of concentration for (b) NaCl, (c) KCl, and (d) MgCl2 aqueous solutions. The results using the Madrid-2019(TIP4P/2005) model are represented by the blue circles and are taken from Ref. 120 for NaCl and from Ref. 124 for KCl and MgCl2. The results of this work using the Madrid-2019(TIP4P/Ice) model are represented by the red squares. The continuous black lines are the fit of the experimental data taken from Refs. 157 and 158 for NaCl and from Ref. 159 for KCl and MgCl2. The dashed lines are guide to the eye lines for the simulation results.
On the compatibility of the Madrid-2019 force field for electrolytes with the TIP4P/Ice water model

December 2024

·

26 Reads

The Madrid-2019 force field was recently developed to perform simulations of electrolytes in water. The model was specifically parameterized for TIP4P/2005 water and uses scaled charges for the ions. In this work, we test the compatibility of the Madrid-2019 force field with another water model: TIP4P/Ice. We shall denote this combination as Madrid-2019(TIP4P/Ice) force field. The key idea of this combination is to keep the ion–ion (Madrid-2019) and water–water (TIP4P/Ice) interactions unaltered with respect to the original models and taking the Lennard-Jones parameters for the ion–water interactions from the Madrid-2019 force field. By implementing this approach, we have maintained a reasonably good performance of the model regarding the densities and structural features of aqueous solutions, albeit yielding a moderately higher viscosity than the original model. However, the standout achievement of this new combination lies in its effective reproduction of the absolute values of the freezing temperatures of a number of ionic aqueous solutions, which could also be useful when studying hydrate formation from a two-phase system containing an aqueous solution in contact with a gas.


Relevant potential energy curves adopted from Refs. 28–30. Two different pathways leading to the Ar2(1,2) channel are indicated by blue and red arrows, respectively.
(a) Scheme of the experimental apparatus. DWP: dual-wavelength plate; QWP: quarter-wave plate. (b) Schematic representation of angular streaking for electrons in elliptically polarized two-color laser fields with ϕL = 0 and π, respectively. (c) Relative phase-dependent yield of the Ar2(1,2) channel as a function of the sum-momentum of two ionic fragments in the z-axis.
(a) Momentum distributions of fragments Ar²⁺ from the Ar2(1,2) channel, induced by the elliptical two-color laser pulses with ϕL = 0. (b) Measured KER spectrum of the Ar2(1,2) channel, fitted with the sum of two Gaussian functions. (c) KER dependent momentum angular distributions of fragments Ar²⁺. ϕ is defined as in (a). (d) Momentum angular distributions of ionic fragments Ar²⁺ integrated over different KER ranges. Panels (e) and (f) correspond to panels (c) and (d), but for ϕL = π.
Schematic of the two pathways involving different intermediates that lead to the Ar2(1,2) channel, induced by elliptical two-color laser pulses with ϕL = 0. Steps involving enhanced ionization are marked by green arrows.
Relative phase-dependent asymmetry parameters of the ejected fragments Ar²⁺ as a function of the KER. The top panel shows the asymmetry parameters for the high- and low-KER regions, fitted with sinusoidal curves.
Dissociative triple ionization of argon dimers in elliptically polarized two-color laser fields

December 2024

·

15 Reads

Using ion–ion coincidence measurements, we experimentally investigate the dissociative triple ionization of argon dimers in relative phase controlled elliptically polarized two-color femtosecond laser fields. By examining the kinetic energy release-dependent momentum angular distribution of the ejected ionic fragments, two distinct pathways, each associated with different intermediates, are identified. Control over the emission directions of the ionic fragments is achieved by varying the relative phase of the elliptical two-color laser fields. Notably, the two pathways exhibit nearly opposite asymmetries in their dependence on the relative phase. Our findings demonstrate the critical role of intermediates in the dissociative triple ionization of dimers and provide a method to distinguish the intermediates involved in this process.


Comparison of various DFT functionals for H atom of fractional occupancies with a minimum basis.
Components of the DFT energy for He atom with fractional occupancy(n) from 1 to 2.
Density functional theory for fractional charge: Locality, size consistency, and exchange-correlation

Jing Kong

We show that the exact universal density functional of integer electronic charge leads to an extension to fractional charge in an asymptotic sense when it is applied to a system made of asymptotically separated densities. The extended functional is asymptotically local and is said to be i-local. The concept of i-locality is also applicable to nuclear external potentials, and a natural association exists between the localities of a density and a set of nuclei. Applying the functional to a system with nuclei distributed in two asymptotically separated locales requires an explicit search of the electronic charge at each locale with the constraint of the global charge. The determined number of electrons at each locale can be fractional. The molecular size consistency principle is realized as the result of the search. It is physically sensible to extend the molecule concept to include a fractional number of electrons (called fractional molecule henceforth) as a localized observable. The physical validity of fractional molecules is equivalent to the asymptotic separability of molecules, a basic assumption in molecular research. A one-to-one mapping between a fractional molecule’s density and external potential is shown to exist with a nondegenerate condition. The global one-to-one mapping required by the Hohenberg–Kohn first theorem is realized through the aforementioned global search for molecular charges. Furthermore, the well-known piecewise linearity of the universal functional with respect to the number of electrons is necessary for an approximate i-local universal functional to be broadly accurate for any integer number of electrons. The Kohn–Sham (KS) noninteracting kinetic energy functional for a fractional molecule is well-defined and has the same form as that for a system of an integer number of electrons. It is shown to be i-local. A nondegenerate, noninteracting ensemble v-representable fractional density is simultaneously noninteracting wavefunction representable. A constrained search over those representing wavefunctions yields the definition of an exchange–correlation functional pertaining to fractional occupancies of KS orbitals. The functional is shown to be an upper bound to the formal KS exchange–correlation energy of a fractional molecule and includes a strong correlation. It yields the correct result for a well-designed example of effective fractional occupancies in the literature.


Structure of the four clusters studied in this work: cluster 1: [CeO7Ti12]38+; cluster 2: [Ce4O22Ti21]56+; cluster 3: [Ce7O31Ti24]62+; and cluster 4: [Ce10O40Ti33]92+. In all cases, the CO molecule is positioned perpendicular to the surface and aligns with the central Ce. (Cerium: green; titanium: blue; oxygen: red.)
fc-MP2/def2-TZVPP CO harmonic vibrational frequency. Cluster size and shape dependence analysis considering only the relaxation of CO (a) and that of CO–Ce (b). Results for cluster 1: [CO@CeO7Ti12]38+; cluster 2: [CO@Ce4O22Ti21]56+; cluster 3: [CO@Ce7O31Ti24]62+; cluster 4: [CO@Ce10O40Ti33]92+ are displayed for surfaces obtained using the BEEF-vdW, PBE-U, and HSE06 functionals.
fc-MP2 CO harmonic vibrational frequency. Cluster-size vs basis-set size analysis considering only the relaxation of the CO structure (a) and that of the CO–Ce trinomial (b). Results for cluster 1: [CO@CeO7Ti12]38+; cluster 2: [CO@Ce4O22Ti21]56+; cluster 3: [CO@Ce7O31Ti24]62+; cluster 4: [CO@Ce10O40Ti33]92+ obtained employing the HSE06(2 × 2) surface.
CO adsorption on CeO2(111): A CCSD(T) benchmark study using an embedded-cluster model

December 2024

·

4 Reads

Juana Vázquez Quesada

·

Sarah Bernart

·

Felix Studt

·

[...]

·

Karin Fink

A benchmark model that combines an embedded-cluster approach for ionic surfaces with wavefunction-based methods to predict the vibrational frequencies of molecules adsorbed on surfaces is presented. As a representative case, the adsorption of CO on the lowest index non-polar and most stable facet of CeO2, that is, (111) was studied. The CO harmonic vibrational frequencies were not scaled semiempirically but explicitly corrected for anharmonic effects, which amount to about 25 cm−1 with all tested methods. The second-order Møller–Plesset perturbation method (MP2) tends to underestimate the CO harmonic frequency by about 40–45 cm−1 in comparison with the results obtained with the coupled-cluster singles and doubles with perturbational treatment of triple excitation method [CCSD(T)] and independently from the basis set used. The best estimate for the CO vibrational frequency (low-coverage case) differs by 12 cm−1 with the experimental value obtained by infrared reflexion absorption spectroscopy of 1 monolayer CO adsorbed on the oxidized CeO2(111) surface. In addition, a conservative estimate of the adsorption energy of about −0.22 ± −0.07 eV obtained at the CCSD(T) level confirms the physisorption character of the adsorption of CO on the CeO2(111) surface.


Time-resolved Coulomb explosion imaging of vibrational wave packets in alkali dimers on helium nanodroplets

Vibrational wave packets are created in the lowest triplet state 13Σu+ of K2 and Rb2 residing on the surface of helium nanodroplets, through non-resonant stimulated impulsive Raman scattering induced by a moderately intense near-infrared laser pulse. A delayed, intense 50-fs laser pulse doubly ionizes the alkali dimers via multiphoton absorption and thereby causes them to Coulomb explode into a pair of alkali ions Ak⁺. From the kinetic energy distribution P(Ekin) of the Ak⁺ fragment ions, measured at a large number of delays, we determine the time-dependent internuclear distribution P(R, t), which represents the modulus square of the wave packet within the accuracy of the experiment. For both K2 and Rb2, P(R, t) exhibits a periodic oscillatory structure throughout the respective 300 and 100 ps observation times. The oscillatory structure is reflected in the time-dependent mean value of R, ⟨R⟩(t). The Fourier transformation of ⟨R⟩(t) shows that the wave packets are composed mainly of the vibrational ground state and the first excited vibrational state, in agreement with numerical simulations. In the case of K2, the oscillations are observed for 300 ps, corresponding to more than 180 vibrational periods with an amplitude that decreases gradually from 0.035 to 0.020 Å. Using time-resolved spectral analysis, we find that the decay time of the amplitude is ∼260 ps. The decrease is ascribed to the weak coupling between the vibrating dimers and the droplet.


First-principles study of electron dynamics of MoS2 under femtosecond laser irradiation from deep ultraviolet to near-infrared wavelengths

Huimin Qi

·

Jinshi Wang

·

Zongwei Xu

·

Fengzhou Fang

Time-dependent density functional theory was employed to investigate the electron dynamics of MoS2 following femtosecond pulse irradiation. The study concerned the effects of laser wavelength, intensities, and polarization and elucidated the ionization mechanisms across the intensity range of 10¹⁰–10¹⁴ W/cm². As laser intensity increases, MoS2 irradiated with an infrared (IR) laser (800 nm) deviates from single-photon absorption at lower intensities compared to that subjected to an ultraviolet (UV) laser (266 nm), and nonlinear effects in the current arise at lower intensities for the 800 nm laser. At a wavelength of 266 nm, MoS2 irradiated with an a-axis polarized laser deposited more energy and generated more electron–hole pairs compared to c-axis polarization. Rate equations were used to estimate the total number of excited electrons in MoS2 and the corresponding plasma frequency. Simulation results indicate that the damage threshold of the UV laser is higher than that of the IR laser, which contradicts the experimental results. This outcome suggests that the mechanism of material damage induced by the UV femtosecond laser near the damage threshold is independent of optical breakdown. The findings of this research are significant for enhancing the performance of MoS2-based photodetectors and optimizing their stability and reliability in high-power, short-wavelength laser applications.


Nonlinear interferometer schematic. A–F refer to the path lengths. IR-BS is an infrared beam splitter, vis-BS is the visible beam splitter, BC is the beam combiner, R is the reference, and S is the sample. ω1, ω2, ωSF is the frequency of the three beams involved: the visible (ω1), the infrared (ω2), and the sum frequency (ωSF), respectively. White light is a broad spectrum white-light source. I and II refer to the two output beams. The red colored cone symbolizes the ultrafast pulse, red for the infrared and blue for the sum frequency.
Normalized, centered interference intensity for GaAs vs Z-cut quartz is independent of the wavelength. Constancy indicates a successful, coordinated IR-BS, BS, and BC shift. The constant value varies with BS shift ΔV from setup; four representative plots are shown. ΔV = 0 is the instrument phase position. (Kelly green) ΔV = 5.0 V, (orange) ΔV = 80 V, (dark green) ΔV = 105 V, and (blue) ΔV = 60 V.
Relative sample-reference phase varies with BS displacement from calibration. The linear interferometer is locked on at −90° for best stability. Data are indicated with the filled circles; the fit is with a solid line. The reference phase [Eq. (10)] is calculated from the intercept.
Measuring complex SFG: Characterizing a phase reference

Ziqing Xiong

·

Rebecca G. Lynch

·

Emma F. Gubbins

·

Mary Jane Shultz

Reactions and interactions at interfaces play pivotal roles in processes ranging from atmospheric aerosols influencing climate to battery electrodes determining charge–discharge rates to defects in catalysts controlling the fate of reactants to the outcome of biological processes at membrane interfaces. Tools to probe these surfaces at the atomic-molecular level are thus critical. Chief among non-invasive probes is the vibrational spectroscopy sum frequency generation (SFG). The complex signal amplitude generated by SFG requires techniques to interfere the unknown amplitude with a well-characterized one. An interferometric method is described to characterize the signal from any nonresonant reference material. The technique is demonstrated by measuring the phase of polycrystalline GaAs, chosen due to the strong signal and insensitivity to surface contamination. With a 515 nm visible field, the phase of GaAs is 54.5° ± 0.5°. The capability of choosing a reference based solely on its signal intensity enables probing a wide range of interfaces.


A nonequilibrium kinetic model of high-resolution vibrational energy transfer in RDX from selective IR excitation

December 2024

·

11 Reads

In this paper, we develop a model based on a second quantization—with anharmonic phonon scattering and the phonon Boltzmann transport equation—to study precise high-resolution nonequilibrium vibrational energy transfer (VET) under selective phonon excitation in cyclotrimethylene trinitramine. We simulate mid-infrared pump–probe spectroscopy and observe a prompt appearance (<1 ps) of broad-spectrum intensity, which agrees well with experimental data in the literature. The selective excitation of phonons at different frequencies reveals distinct VET pathways and the kinetic evolution of mode occupations as the system reaches a new equilibrium temperature. Three types of transition mechanisms are found to play outsized roles in terms of the amount of energy transferred and the transfer rate: (1) vibrational modes close to the excited frequencies typically respond faster and reach higher temperatures regardless of the excitation frequency; (2) the overtone pathway connecting the modes near 550 and 1150 cm⁻¹ is an important bridge between far- and mid-IR; and (3) fast aggregation of energy at 2800 cm⁻¹ mediates transfer to/from high frequencies through a second overtone pathway involving modes near 1400 cm⁻¹. In addition, by monitoring the temperature of the N–N/N–O stretching modes, strong coupling between those modes and the C–H stretching modes is found. The coupling likely draws the vibrational modes close to both the proton transfer transition state for HONO elimination and the N–N stretching for bond cleavage. The high-resolution understanding of the nonequilibrium kinetics of phonons provides important insight into the energy transfer and initiation mechanisms of molecular solids due to external stimuli.


Materials characterization of the ITO/ZnO/ITO/Si memristor device. (a) X-ray diffraction (XRD) pattern of the sputtered ITO/ZnO/ITO structure on a Si substrate, confirming the polycrystalline structure of the ZnO functional layer. (b) X-ray photoelectron spectroscopy (XPS) survey spectra providing an overview of the elemental composition of the ITO/ZnO/ITO layers. (c) XPS depth profile analysis showing the atomic percentages of oxygen (O), zinc (Zn), indium (In), and tin (Sn) at different etching times, revealing the distribution of these elements throughout the layers. Deconvolution of the (d) In 3d, (e) Zn 2p. (f) Sn 3d, and (g) O 1s XPS spectra across a range of etching times from 0 to 160 s, with data taken every 20 s, illustrating the different chemical compositions and states of In, Zn, Sn, and oxygen within the ITO/ZnO/ITO layers.
Electrical characteristics of the ITO/ZnO/ITO/Si memristor: (a) Typical I–V sweeps showing the forming and reset processes at ∼4.0 and −0.9 V, respectively. (b) Box plot depicting the forming and reset voltages measured from 16 memristor devices. (c) Reproducible bipolar switching behaviors observed over 100 repeated cycles under positive 2.0 V and negative −2 V voltage sweeps. (d) Endurance characteristics of the memristor devices. (e) Histograms of the set and reset voltage distributions fitted with Gaussian curves. (f) Box plot showing the resistance distribution of the HRS and LRS measured from six randomly selected memristor devices. (g) Typical threshold switching behavior operated at an ICC of 500 μA over 100 repeated cycles. (h) Box plot of the LRS and HRS measured from six memristor devices for endurance properties. (i) Histogram of the LRS and HRS resistance distributions fitted with Gaussian curves. (j) Analog I–V sweeps obtained by varying the reset-stop voltage from −0.7 to −1.7 V in −0.05 V steps under ICC of 500 μA. (k) Resistance as a function of the reset-stop voltage. (l) Analog I–V sweeps obtained by varying the reset-stop voltage from −1.0 to −2.0 V in −0.02 V steps under ICC of 800 μA. (m) Resistance as a function of reset-stop voltage. (n) Analog I–V curves with incremental steps of ICC from 100 μA to 1 mA. (o) Resistance levels as a function of ICC for different steps.
Analog synaptic characteristics of ITO/ZnO/ITO/Si memristor. (a) Conductance states observed over 50 potentiation/depression (P/D) cycles with set voltages between 1.4 and 1.7 V, reset voltages ranging from −1.45 to −1.85 V, a pulse width of 20 μs, an interval of 10 μs, and 30 repeated cycles per test. (b) Gain in excitatory post-synaptic current (EPSC) as a function of set (1.4–1.7 V) and reset (−1.45 V to −1.85 V) pulse amplitudes. (c) Variation in conductance states with pulse widths ranging from 10 to 50 μs, using a set amplitude of 1.8 V, a reset amplitude of 1.85 V, and 30 repeated cycles per test. (d) EPSC gain compared to set and reset pulse widths (10–50 μs). (e) PPF index in relation to pulse intervals ranging from 10 to 1000 μs, using constant pulse amplitudes of 1.8 V and a width of 20 μs. (f) Spike-amplitude-dependent plasticity (SADP) plotted against ten sequential pulse stimuli, each with a constant width of 20 μs and an interval of 10 μs. (g) Average change in current (ΔI) following consecutive pulses (ΔI = I2 − I1) and after the tenth pulse (ΔI = I10 − I1) for various pulse amplitudes (from 1.8 to 1.92 V), demonstrating PPF and post-tetanic potentiation (PTP). (h) SRDP plotted against ten sequential pulse stimuli, each with a constant width of 20 μs and an amplitude of 1.92 V. (i) SRDP as a function of change in current (ΔI), showing PPF and PTP behaviors. (j) EPSC current response measured at frequencies ranging from 0.1 to 0.5 MHz. (k) SNDP under different pulse amplitudes (1.74–1.8 V), with a constant width of 20 μs and an interval of 10 μs. (l) SNDP index vs the number of spikes for various pulse amplitudes (1.74–1.8 V). (m) SNDP at different pulse widths. (n) SNDP index vs the number of spikes at different pulse widths (10–30 μs).
Implementation of an RC system using STM characteristics of the ITO/ZnO/ITO/Si memristor device. (a) Schematic of the psychological multistore memory model, illustrating the transition from STM to LTM. (b) Transition from STM to LTM by modulating pulse amplitudes in the range of 2.5–3.0 V. (c) Transition from STM to LTM by varying pulse widths from 200 to 800 μs. (d) Schematic of the RC structure, which includes three layers: input, reservoir, and output, corresponding to the input [x(t)], reservoir state [u(t)], and output [y(t)], respectively. The input weights remain fixed while the weights connecting the reservoir and output layers undergo training. (e) Implementation of number recognition through the RC system by applying electrical pulses to five different devices. Input pulse sequences were fed into the memristors, and the corresponding reservoir states (xi) were recorded. (f)–(i) Current responses to individual pulses from five pulse sequences used for digit recognition of 5 × 4-pixel numbers: digits “0,” “1,” “5,” and “9.”
Handwritten digit recognition using an ITO/ZnO/ITO/Si memristor device-based RC system. (a) Input image of the handwritten digit “8.” (b) Implementation diagram of the RC layer. (c) Schematic of the CNN used for MNIST handwritten digit classification, including convolutional and pooling layers with multiple output channels; the number of outputs is indicated by the first number under each layer label, and the last three layers are fully connected. (d) Current response for the 16 states of the 4-bit memristor system, demonstrating the RC implementation. (e) Simulated pattern recognition accuracy for the 28 × 28 MNIST dataset, achieving up to 97.82% after training with 16 conductance states using sequential training with CNN; ten cycles were repeated per epoch to ensure reliability and reproducibility. (f) Confusion matrix showing predicted vs actual labels from the memristor-based 16-conductance-state 4-bit system during the testing phase of the trained pattern recognition system.
Reservoir computing and advanced synaptic plasticity of sputter-deposited ZnO memristors with controllable threshold and nonvolatile switching behavior

December 2024

·

14 Reads

This study presents an ITO/ZnO/ITO/Si memristor fabricated via reactive sputtering for use in advanced analog synaptic plasticity and reservoir computing (RC) systems. The proposed device exhibited stable threshold and nonvolatile switching characteristics by effectively controlling the current compliance (ICC) limit. Multilevel data storage was achieved through controlled multistate switching via reset-stop voltage and ICC. X-ray diffraction analysis confirmed the formation of a polycrystalline ZnO film with a 12:8 oxygen-to-argon ratio, which facilitated the generation of oxygen-vacancy conductive filaments. The memristor effectively replicated key synaptic characteristics such as long-term potentiation, long-term depression, spike-amplitude/width-dependent plasticity, spike-rate-dependent plasticity, and the transition from short-term to long-term memory. The RC system processed binary 4-bit codes and recognized different digits, achieving 98.84% accuracy in handwritten digit recognition using a convolutional neural network simulation, highlighting its potential for efficient image processing applications.


A schematic representation of the factor (−1)ⁿ arising from converting an nth order commutator to its Poisson bracket. See the Appendix for details.
On the relation between the velocity- and position-Verlet integrators

The difference and similarity between the velocity- and position-Verlet integrators are discussed from the viewpoint of their Hamiltonian representations for both linear and nonlinear systems. For a harmonic oscillator, the exact Hamiltonians reveal that positional trajectories generated by the two integrators follow an identical second-order differential equation and thus can be matched by adjusting initial conditions. In contrast, the series expansion of the Hamiltonians for the nonlinear discrete dynamics clearly indicates that the two integrators differ fundamentally. These analytical results are confirmed by simple numerical simulations of harmonic and anharmonic oscillators.


Deviations from ideality in solutions of dicarboxylic acid salts modeled within a BiMSA theory for flexible chains

The binding mean spherical approximation theory is used to describe the thermodynamic properties of dicarboxylic acid salts by adding a chain term in the free energy. The dianions in these solutions are modeled as flexible charged chains composed of two, three, or four spheres. Five aqueous solutions of such salts are studied in different concentration ranges: dipotassium oxalate, disodium malonate, disodium succinate, potassium tartrate, and sodium tartrate. A description of the experimental deviations from ideality (osmotic and activity coefficients) for these salts is obtained. The model is compared with a previous one that does not include a chain contribution. It is found that the model with a chain contribution provides a more physically sound framework.


Mixing small proteins with lipids and cholesterol

December 2024

·

15 Reads

Many ternary mixtures composed of saturated and unsaturated lipids with cholesterol (Chol) exhibit a region of coexistence between liquid-disordered (Ld) and liquid-ordered (Lo) domains, bearing some similarities to lipid rafts in biological membranes. However, biological rafts also contain many proteins that interact with the lipids and modify the distribution of lipids. Here, we extend a previously published lattice model of ternary DPPC/DOPC/Chol mixtures by introducing a small amount of small proteins (peptides). We use Monte Carlo simulations to explore the mixing phase behavior of the components as a function of the interaction parameter representing the affinity between the proteins and the saturated DPPC chains and for different mixture compositions. At moderate fractions of DPPC, the system is in a two-phase Ld + Lo coexistence, and the proteins exhibit a simple partition behavior between the phases that depends on the protein–lipid affinity parameter. At low DPPC compositions, the mixture is in Ld phase with local nanoscopic ordered domains. The addition of proteins with sufficiently strong attraction to the saturated lipids can induce the separation of a distinct Lo large domain with tightly packed gel-like clusters of proteins and saturated lipids. Consistent with the theory of phase transitions, we observe that the domain sizes grow when the mixture composition is in the vicinity of the critical point. Our simulations show that the addition of a small amount of proteins to such mixtures can cause their size to grow even further and lead to the formation of metastable dynamic Lo domains with sizes comparable to biological rafts.


Interfacial vs confinement effects in the anisotropic frequency-dependent dielectric, THz and IR response of nanoconfined water

December 2024

·

14 Reads

We investigate the anisotropic frequency-dependent dielectric, THz and IR response of liquid water confined between two planar graphene sheets with force-field- and density-functional-theory-based molecular dynamics simulations. Using spatially resolved anisotropic spectra, we demonstrate the critical role of the volume over which the spectral response is integrated when reporting spatially averaged electric susceptibilities. To analyze the spectra, we introduce a unique decomposition into bulk, interfacial, and confinement contributions, which reveals that confinement effects on the spectra occur only for systems with graphene separation below 1.4 nm, for all frequencies. Based on this decomposition, we discuss the molecular origin of the main absorption features of nanoconfined water from the GHz to the IR regime. We show that, at low frequencies, the 15 GHz Debye peak of interfacial water is redshifted due to a slowdown of collective water reorientations. At high frequencies, the OH stretch at 100 THz blue shifts and a signature of free OH groups emerges, while the HOH bend mode at 50 THz is redshifted. Strikingly, in nanoconfinement, the 20 THz libration band shifts to below 15 THz and broadens drastically, spanning two orders of magnitude in frequency. These results are rationalized by the collective water motion and the structure of the hydrogen-bond network at the water–graphene interface and in two-dimensional water layers, which reveals the intricate behavior of nanoconfined water and its spectral properties.


Results of PIMC simulations and extrapolated values for the density [(a) and (c)] and speed of sound [(b) and (d)] at the state point (80 K, 90 MPa). (a) and (b) Extrapolation of results for systems of 108 particles to the quantum-mechanical limit as a function of P⁻² for the PIMC levels 2B and 2B+3B. For the speed of sound, results for the analytical virial estimator, finite-difference virial estimator, and primitive estimator are given. (c) and (d) Extrapolation of the results in the quantum-mechanical limit and results of semiclassical simulations using the Feynman–Hibbs correction as a function of the inverse particle number to the thermodynamic limit. Solid lines and dashed lines represent linear fits to the data.
Comparison of estimated uncertainties in all ten investigated properties obtained with the primitive and virial estimators at the two state points (80 K, 15 MPa) and (80 K, 90 MPa) for model 2B+3B.
Results for the density at the state point (223.15 K, 38.28 MPa). (a) Extrapolation of results for the density from PIMC simulations in the quantum-mechanical limit and of results of semiclassical simulations using the Feynman–Hibbs correction as a function of the inverse particle number for models 2B and 2B+3B to the thermodynamic limit. (b) Relative deviations of the results for the density in the thermodynamic limit, experimental data of McLinden and Lösch-Will, and VEoS-n from the EoS as a function of pressure.
Relative deviations of PIMC results at the level 2B+3B-Vir(FD), results of semiclassical MC simulations using the Feynman–Hibbs correction with model 2B+FH+3B, experimental data from the literature, and the VEoS-n for the density (a) and speed of sound (b) in the thermodynamic limit from the EoS as a function of pressure at the isotherm 80 K.
Relative deviations of our results for ten different thermodynamic properties [(a)–(j)] at levels 2B-Cl, 2B+3B-Cl, 2B-PI, 2B+3B-PI, and 2B+FH+3B at the isotherm 80 K from the EoS as a function of pressure.
Calculation of thermodynamic properties of helium using path integral Monte Carlo simulations in the NpT ensemble and ab initio potentials

December 2024

·

11 Reads

We apply the methodology of Lustig, with which rigorous expressions for all thermodynamic properties can be derived in any statistical ensemble, to derive expressions for the calculation of thermodynamic properties in the path integral formulation of the quantum-mechanical isobaric–isothermal (NpT) ensemble. With the derived expressions, thermodynamic properties such as the density, speed of sound, or Joule–Thomson coefficient can be calculated in path integral Monte Carlo simulations, fully incorporating quantum effects without uncontrolled approximations within the well-known isomorphism between the quantum-mechanical partition function and a classical system of ring polymers. The derived expressions are verified by simulations of supercritical helium above the vapor–liquid critical point at selected state points using recent highly accurate ab initio potentials for pairwise and nonadditive three-body interactions. We observe excellent agreement of our results with the most accurate experimental data for the density and speed of sound and a reference virial equation of state for helium in the region where the virial equation of state is converged. Moreover, our results agree closer with the experimental data and virial equation of state than the results of semiclassical simulations using the Feynman–Hibbs correction for quantum effects, which demonstrates the necessity to fully include quantum effects by path integral simulations. Our results also show that nonadditive three-body interactions must be accounted for when accurately predicting thermodynamic properties of helium by solely theoretical means.


Time evolution of the trefoil knot of size Kn(t) (in red) in a chain with N = 1024 segments at shear rate γ̇=0.005 (Wi ≈1071) in Couette flow. Thus, (a) Kn(55τMD) = 186, (b) Kn(176τMD) = 58, (c) Kn(200τMD) = 12, and (d) Kn(775τMD) = 7. The knotted section (as determined by our automated procedure) is depicted in red. All plots are prepared with VMD.⁵⁷
(a) Average knot size Kn decreasing with time t in a chain with N = 512 segments for several shear rates, or Weissenberg numbers Wi. (b) The same as in (a) for N = 1024. (c) The onset of knot tightening at critical shear rate Wicr ≈ 7.97 for chain length N = 512. (d) The largest eigenvalue of the mean-squared gyration radius ⟨Rg2(t)⟩ for a chain with N = 1024 and several Weissenberg numbers. The inset shows the variation in the eigenvalues of the tensor ⟨Rg2⟩: λ1(t) > λ2(t) > λ3(t), indicating the transformation of the coil into a stretched string with time for Wi ≈ 1071.
Tumbling effects for chains of size N = 512. (a) Trefoil length distributions; only configurations after t = 1000τMD are analyzed to exclude the dynamic tightening phase. Tight knots are strongly favored at high Weissenberg numbers due to stretching Couette flow, but the distribution shows a long tail indicating that knots occasionally increase in size. (b) Kymographs of the knot start and end positions compared to the relative x-position of the knot center xrel along the chain. Three individual simulation runs where tumbling occurred are shown: Knot sizes fluctuate and grow whenever the knot center reaches the most extreme x-positions (xrel ≈ 100% or xrel ≈ 0%). (c) Snapshots of the simulation run shown in (b) for Wi = 478.38, with 3× zooms onto the knotted regions. (i) The chain is completely stretched and the knot is in a non-diffusive tightened state. (ii) and (iii) The chain end with larger x-coordinate folds as tumbling begins, leaving the knot at xrel ≈ 100%. At this point, knot sizes start to fluctuate. (iv) The tumble is almost complete as the knot reaches xrel ≈ 0%.
Structural and topological analysis of a representative simulation of a polymer globule with N = 512 beads in Couette flow at γ̇=0.75. (a) Conformation of an initially globular polymer and a shear-induced pearl necklace configurations oriented along the Y-axis perpendicular to flow (corresponding to the contact map (b-iv). The continuous coloring scheme from red to white to blue indicates the monomer index. (b) Contact maps displaying monomers that come within 5σ of other monomers, averaged over 5% of the simulation run starting at (i) t = 0, (ii) t = 900τMD, (iii) t = 1800τMD, and (iv) t = 2850τMD. (c) Variation of the eigenvalues λ1, λ2, λ3 of the tensor Rg2(t) with time indicating ongoing unfolding and refolding of the globule along the Y (black), and X-axes (red). (d) Kymograph indicating the start and end of the knotted section in the course of the simulation run.
Effect of simple shear on knotted polymer coils and globules

We explore the effect of Couette flow on knotted linear polymer chains with extensive molecular dynamics simulations. Hydrodynamic interactions are accounted for using multi-particle collision dynamics. The polymer chain, originally containing a simple trefoil knot at rest, is described by a coarse-grained bead-spring model in a coil or globular state. We demonstrate that under shear existing loosely localized knots in polymer coils typically tighten to several segments beyond a certain shear rate threshold. At large shear rates, the polymer undergoes a tumbling-like motion during which knot sizes can fluctuate. In contrast, sheared knotted globules unwind into a convoluted pearl-necklace structure of sub-globules that folds back onto itself and in which knot types change over time.


(a) Setup of the equilibrium simulation system with a liquid–liquid interface controlled by a miscibility parameter η. A pressure control piston (PC) and a semipermeable phantom wall (PW) were located on each end of the system in the x-direction (left and right). (b) Side snapshots of the equilibrium systems for the free-energy calculation by the thermodynamic integration using the (i) extended dry surface (DS) and (ii) extended PW methods. For the extended DS method (i), the miscibility parameter η was varied while keeping the PWs at x1pw and −x1pw far from the liquid; thus, the PWs as well as the PCs are shown only on the top panel of (i). For the PW method (ii), the PW positions xpw and −xpw were varied while keeping η unchanged.
(a) Distributions of the diagonal stress components τyy (black) and τxx (green) calculated by the VA and the densities ρα (blue) and ρβ (red) of α and β components for the systems at η = 0.01, 0.4, 0.8, and 1. Enlarged snapshots of the systems around the interface are also shown. (b) Interfacial tension γαβ calculated from the stress distribution by Eq. (10) as the mechanical route. The value of 2γLV obtained in an independent system is also displayed.
Comparison between the relative interfacial tension −(γαβ − 2γLV) obtained with the mechanical route and work of isolation Wiso for various η. Error bars for WisoDS are not shown for better visualization here (see Fig. 5). Side snapshots in Fig. 1(b) with the corresponding (η, xpw) value are appended for some systems.
(a) Force on the right phantom wall (PW) per unit area upon the calculation of the work of isolation WisoPW(η) by the extended-PW method in the completely miscible case (η = 1). The inset corresponds to Fig. 1(a-ii). (b) Schematic of the force balance among the force on the left and right PWs, the pressure of the two single component liquids, and that of the mixed liquid in the center between the two PWs.
Mechanical and thermodynamic routes to the liquid–liquid interfacial tension and mixing free energy by molecular dynamics

December 2024

·

12 Reads

Rei Ogawa

·

Hiroki Kusudo

·

Takeshi Omori

·

[...]

·

In this study, we carried out equilibrium molecular dynamics (EMD) simulations of the liquid–liquid (LL) interface between two different Lennard-Jones components with varying miscibility, where we examined the relation between the interfacial tension and the free energy to completely isolate the two liquids using both a mechanical and thermodynamic approach. Using the mechanical approach, we obtained a stress distribution around a quasi-one-dimensional EMD system with a flat LL interface. From the stress distribution, we calculated the LL interfacial tension based on Bakker’s equation, which uses the stress anisotropy around the interface, and measured how it varied with miscibility. The second approach uses thermodynamic integration by enforcing quasi-static isolation of the two liquids to calculate the free energy. This uses the same EMD systems as the mechanical approach, with both extended dry-surface and phantom-wall (PW) schemes applied. When the two components were immiscible, the mechanical interfacial tension and isolation free energy were in good agreement. When the components were miscible, the values were significantly different. From the result of PW for the case of completely mixed liquids, the difference was attributed to the additional free energy required to separate the binary mixture into single components against the osmotic pressure prior to the complete detachment of the two components. This provides a new route to obtain the free energy of mixing.


Consistent theoretical description of nuclear spin long-lived states decay under conditions of reversible ligand–protein binding

Determining the stability constant of the complex formed by an organic ligand with a protein is the first stage in the screening of new drugs. Nuclear spin long-lived states, in particular the singlet state, can be used to study the reversible binding of ligands to proteins. In a complex with a protein, the spins of the ligand interact with the spins of the protein, the system of protein and ligand nuclei can relax by a dipole–dipole mechanism, and the lifetime of the singlet state is strongly reduced. In this theoretical study, a system of encounter theory equations with the condition of fast relaxation in free protein was solved to determine the lifetime of the LLS in the presence of protein. It was shown that in the limit of fast chemical exchange, the relaxation of the LLS of the ligand nuclei due to dipole interaction with the protein nuclei is reduced to relaxation by the mechanism of dipole interaction with one proton of the protein, which is located at some effective distance from the ligand nuclei. Numerical calculations were made to test the applicability of the approximations used to process the experimental lifetime dependencies on the ligand concentration and external field, and it was shown that these approximations coincide with the limit of fast exchange in strong and weak magnetic fields, but not in the medium field. An analytical expression for the lifetime of the singlet state of ligand nuclei in an arbitrary magnetic field in the absence of protein was obtained.


Variance-based wave function optimization within the unrestricted doubly occupied configuration interaction framework: A half-projection treatment

Javier Garcia

·

Diego R. Alcoba

·

Alicia Torre

·

[...]

·

Gustavo E. Massaccesi

The energy-variance-based optimization procedures have proven to be useful tools to describe N-electron spectra. However, the resulting wave functions usually present spin-contaminant contributions. The goal of this work is to reduce the spin contamination of the results arising from the unrestricted doubly occupied configuration interaction method in its energy variance minimization version [Alcoba et al., J. Chem. Phys. 160, 164107 (2024)]. We propose to incorporate the half-projection technique, which allows removing the spin components with even or odd spin quantum number of an approximate N-electron wave function, into the framework of the unrestricted doubly occupied configuration interaction treatment. This implementation can be carried out following several possible ways, whose results are analyzed in detail, in order to show the behavior of each procedure. Numerical determinations performed on selected strongly correlated N-electron systems, in ground and excited states, allow us to assess the most suitable procedure.


Uncovering the lung cancer mechanisms through the chromosome structural ensemble characteristics and nucleation seeds

Wen-Ting Chu

·

Jin Wang

Lung cancer is one of the most common cancers in humans. However, there is still a need to understand the underlying mechanisms of a normal cell developing into a cancer cell. Here, we develop the chromosome dynamic structural model and quantify the important characteristics of the chromosome structural ensemble of the normal lung cell and the lung cancer A549 cell. Our results demonstrate the essential relationship among the chromosome ensemble, the epigenetic marks, and the gene expressions, which suggests the linkage between chromosome structure and function. The analysis reveals that the lung cancer cell may have a higher level of relative ensemble fluctuation (micro CFI) and a higher degree of phase separation between the two compartments than the normal lung cell. In addition, the significant conformational “switching off” events (from compartment A to B) are more than the significant conformational “switching on” events during the lung cancerization. We identify “nucleation seeds” or hot spots in chromosomes, which initiate the transitions and determine the mechanisms. The hot spots and interaction network results reveal that the lung cancerization process (from normal lung to A549) and the reversion process have different mechanisms. These investigations have revealed the cell fate determination mechanism of the lung cancer process, which will be helpful for the further prevention and control of cancers.


Journal metrics


3.1 (2023)

Journal Impact Factor™


7.4 (2023)

CiteScore™


24 days

Submission to first decision


0.9 (2023)

Immediacy Index


0.07284 (2023)

Eigenfactor®


0.908 (2023)

Article Influence Score


USD 2,750

Article processing charge

Editors