Abstract
Passive transport of molecules through nanopores is characterized by the interaction of molecules with pore internal walls and by a general crowding effect due to the constricted size of the nanopore itself, which limits the presence of molecules in its interior. The molecule–pore interaction is treated within the diffusion approximation by introducing the potential of mean force and the local diffusion coefficient for a correct statistical description. The crowding effect can be handled within the Markov state model approximation. By combining the two methods, one can deal with complex free energy surfaces taking into account crowding effects. We recapitulate the equations bridging the two models to calculate passive currents assuming a limited occupancy of the nanopore in a wide range of molecular concentrations. Several simple models are analyzed to clarify the consequences of the model. Eventually, a biologically relevant case of transport of an antibiotic molecule through a bacterial porin is used to draw conclusions (i) on the effects of crowding on transport of small molecules through biological channels, and (ii) to demonstrate its importance for modelling of cellular transport.
Introduction
Biological nanopores in eukaryotic and prokaryotic cells represent, for those molecules for which cell membranes are impermeable, the gate to diffuse in and out of and between internal organelles [1,2,3]. Specialized protein complexes provide the necessary modulation to the number of transported molecules (the flux/current) and their molecular characteristic (the selectivity). The main objective of a controlled and selective transport is to keep the integrity of the internal environment (membrane potential, concentrations of solutes, pH), key for a correct functioning of cells.
The main general and simple selection rule to achieve controlled transport is based on constriction of the pore size, where an entropic barrier prevents fast entry of molecules. The entropic barrier can be modulated independently by electrostatic and other specific chemical interactions. In biology, a typical example is provided by porins, beta-barrel proteins with an internal water-filled pore. In porins, the diffusive transport of molecules/ions/water is modulated by a constricted shape created by one or more loops folded internally and with different degrees of flexibility [4,5, 6,7]. Investigating the transport of particles under confinement is thus of primary importance to unveil and rationalize the mechanism of biological nanopores functioning [8,9]. At the same time, such biological complexes are a source of inspiration for the development of artificial pores with potential technological applications, such as in molecular sensing [10,11, 12,13].
Recently, the necessity to rationalize and predict the accumulation of molecules in Gram-negative bacteria has emerged as prerequisite to the development of new effective drugs [14,15]. For many small and polar molecules, their accumulation is obtained via a subtle balance between passive diffusion through porins (in and out) and active efflux (out) [16]. Experimentally the accumulation seems not to follow neither simple Fick’s law kinetics for entry nor the Michaelis–Menten kinetics for extrusion [17,18], showing a nonlinear behavior. The Fick’s law was first introduced to describe the free diffusion of noninteracting solute molecules in solvent and then generalized, by introducing the potential of mean force, to the case of membranes with pores, assuming linear concentration dependence. Thus, a proper treatment of saturation due to the interaction of the solute molecules is necessary for the correct use of Fick’s law in a more general and complex model to understand and predict accumulation in finite volumes [15].
The passive transport of molecules through membrane pores is driven by the gradient of the chemical potential and can often be described within the diffusion approximation by using the over-damped Brownian dynamics or the Smoluchowski equation [19]. This model has been widely used for over a century in the form of Nernst–Plank equation for electrodiffusion. It is also behind important classical concepts of membrane physiology such as the membrane potential and current (Goldman–Hodgkin–Katz equation). This is a linear theory that assumes the linear dependence of the molecular flux through the pore on the concentration of the molecules and explains the generalized Fick’s law of the permeability through membranes. The diffusion model also assumes that molecules do not interact with each other inside pores. In other words, it is valid at low concentrations when the probability to have two or more molecules in the pore at the same time is negligible. At high substrate concentrations, especially in the case of favorable pore–substrate interactions, the probability to have two substrate molecules trying to occupy the pore at the same time is not negligible. That leads to the slower growth of the diffusion current vs the substrate concentration, and finally to the saturation, i.e., the independence of the passive current on the concentration.
The saturation can be observed in single-channel electrophysiology (SCE) [20], where the presence of a molecule in the pore is detected via the observations of the corresponding blockages of the ionic current through the pore [21] or via the analysis of the ion current fluctuations at various concentration gradient [20]. The SCE data have been analyzed in terms of the Markov state model, [22], which assumes one or more binding sites for the molecule in the pore and uses the master equations, a discrete Markov process (Markov chain) model, to trace the time course of the occupation probabilities of the binding sites. The simplest two-state Markov model assumes that, at the maximum, one particle at a time can occupy the channel (binding site). It can describe both the linearity with the concentration and the saturated behavior. The parameters of the Markov state model, the transition rates between the states, in principle, can be obtained from the experiment (e.g., electrophysiology). In many cases, it is impossible to calculate directly the transport of molecules through nanopores by using the all-atom molecular dynamics (MD) simulations as the characteristic times (more than milliseconds) are beyond the capabilities of modern high-performance computing (HPC) infrastructures
However, thanks to the use of sophisticated algorithms for accelerating sampling, it is feasible to obtain the free energy surface for the transport of a molecule through a nanopore as well as the diffusion coefficient by using MD simulations and 1–100 microseconds long trajectories [23]. This gives the necessary input data for the diffusion model. In the present work, we discuss in detail how to exploit the free energy surface and the diffusion coefficient of the 1D diffusion model to obtain the rate constants of the corresponding Markov state model. Only by combining the two models, one can obtain the passive current of molecules through pores as function of the gradient concentration (or gradient of chemical potential), which can be accessed experimentally, for example, by the reversal potential method [23], thus also helping the interpretation of experimental results.
The mathematical grounds and physical applications of the continuous and discrete Markov processes have been known for decades if not centuries. However, the applications to the passive transport of molecules through nanopores bridging the timescales from the diffusion (ps-ns) to the Markov state (us-ms) are relatively recent, see, e.g., [24,25], and are mainly focused on the theoretical development based on the mean first passage time (MFPT) concept and to the analysis of simplified analytic models. Here, we focus on passive transport through biological nanopores within the bridged multiscale approach – from all-atom MD simulation to the 1D diffusion and two state Markov model (MD-1DD-2MS).
This article is organized as follows. First, we recapitulate the necessary results of the 1D diffusion theory and of the Markov state models. Then, we derive the expression of the kinetic rate constants in terms of the free energy profile and the diffusion coefficient, in original way, without the use of the MFPT concept. We discuss the qualitative properties of the theory on simple analytic models of free energy barriers and binding site and then consider, as a more realistic application, the transport of biologically related molecules through bacterial porins.
Theory
Diffusion model
In many cases, the transport of molecules though nanopores can be adequately described within the diffusion approximation by considering the molecules as over-damped Brownian particles [19,26,27], and assuming the adiabatic separation of the time scales, i.e., the fast degrees of freedom are in the thermodynamic equilibrium, while few slow ones are evolving quasi-stationary according to a stochastic Markov process. Often, it is sufficient to consider only one slow coordinate, e.g., the one characterizing the position of the molecule along the pore axis or the reaction path. The corresponding one-dimensional Fokker–Planck equation (FPE) for the probability density of molecule’s coordinate,
where
where
As far as the PMF,
In the practically important case of the steady-state, the particle distribution is time independent,
where
For small numbers,
For the specific cases, equation (3) has been known and practically used for many decades, e.g., the Fick’s law for the diffusive current through membranes driven by the gradient of the concentration, the Goldman–Hodgkin–Katz equation for ionic current through membrane driven by the gradient of the electric potential and ion concentration difference, etc. The formula for the diffusion current is modified in the case when the radiation boundary conditions are more appropriate [33], or for a single particle in a periodic PMF driven by a constant force [19,34,35].
Equation (3) demonstrates that the PMF enters the current in the exponent, while the diffusion coefficient does only linearly: the dependence of the diffusion current on the PMF is stronger. In many cases, the diffusion coefficient can be effectively taken as a constant,
This latter concept has been successfully applied recently to model the permeation of antibiotics through outer membrane channels of Gram-negative bacteria [7,36].
The central input quantities of the model,
Markov state model
The Markov state kinetic model of the pore–molecule interaction, used, e.g., for the analysis of the channel gating in the electophysiology [47], starts with a set of states defined by different ionic conductance levels of the channel (the “bound states”) due to the presence of substrate molecules. The transitions between these states are assumed to be much faster than the corresponding residence times. The ground state is the open-channel state (optionally, several) without any molecule inside. The transitions between the states are described as a discrete Markov process determined by the corresponding rate constants – the target parameters describing the pore–molecule interaction kinetics [22].
Here, we consider a one-binding-site (two-states) Markov model of the channel, the extension to the many-state model is straightforward.
Let
Additional normalization condition,
Below, we summarize necessary results from the two-state Markov model in the steady-state [22,47,48]. If the particles are added on the both sides at the concentrations, respectively,
The net current of molecules through the pore reads,
With the aforementioned definition, the current is positive for the net transport from cis to trans, and it is negative for the opposite direction.
From the aforementioned results, and from the general requirements of the zero net particle current, in the case of the symmetric bulk solution on the sides of the channel (equal concentrations, no electric field gradient, etc.),
Thus, one finally obtains,
Equation (13) shows that the molecular pore-translocation current obeys Fick’s law for small concentrations only,
In the opposite case of large concentrations, in contrast with the diffusion approximation, the molecular current saturates and reaches its maximum,
Another important difference between the Markov state model and the diffusion model is that the diffusion molecular current is always symmetric with respect to the side of application of the concentration gradient, while the Markov state translocation current is symmetric only at small concentrations. In general, the translocation current depends on the side of application of the concentration gradient,
In the case of asymmetric solvent on cis and trans sides, due, e.g., to the different solvent activity or an applied electric potential for charged molecules, the definition of the in-rates should be modified,
Interestingly, the maximum current at high concentration is given by the same equations (16).
Bridging 1D diffusion and the Markov state models
To connect the kinetic rates of the Markov state model with the PMF and the diffusion coefficient, one can consider the low concentration limit, where the both models are linear with the concentration. Similar arguments have been used by other authors, see, e.g., [24,25]. We also assume that the solute at concentration
The linear-regime particle current through the channel when added to one side (the same for trans and cis) at concentration gradient
Quantities
satisfying
The aforementioned formulas connect the Markov state model with the diffusion approximation.
The translocation current and the channel occupancy at higher concentrations can be calculated by using Markov state model results, equations (10) and (13). In particular, if the molecules are added from one side only, one finds,
It is worth to note again that the two-state Markov state is a time average concept with respect to the continuous diffusion approximation. It assumes that the average time spent by the molecule in each of the states is much larger than the average diffusion time required to cross the barrier between the states. The latter can be estimated as
Only if the aforementioned conditions are satisfied, one can use the obtained kinetic rates (
However, equations (31) for the steady-state tanslocation current are also valid at much weaker conditions. It assumes two main approximations. First, the current, e.g., from cis to trans, is presentable in the form (compared with equation (20)),
where
where
By following the latter approach, one can formally extend the model of at maximum
Application
By representing the PMF as a piece-wise constant function, as shown in Figure 1, one can develop simple analytic models to illustrate (i) how is the molecule–pore interaction, (ii) how the crowding affects the translocation current, and (iii) how the kinetic rates are accessible in the electrophysiology.
We will start with a single barrier/binding site inside a pore of length
Analytic single barrier/binding site model
The barrier (binding site) has the width
Here,
The linear diffusion current due to the gradient of concentration
The molecular current is independent of the rectangle position,
Thus, the presence of a barrier for the diffusing molecule decreases the diffusion current, while the binding site increases the diffusion current effectively by reducing the channel length to
The channel occupancy for the cis molecule addition, for the trans addition, and for the symmetric addition reads, respectively:
Here, the distance to the barrier/well from the trans side,
1DD-2MS mapping, large barrier:
Finally, the translocation current through the high barrier reads for the cis and the trans addition, respectively,
At high concentrations,
The saturated translocation current is larger when molecules are added from the side to which the barrier is closer (the barrier at the entry side).
The mapping onto the 2MS model is formal in this case, as molecules stay either on the left side of the barrier or on the right one (two different states). Thus, the obtained transition rate constants do not have real physical meaning. Nevertheless, the steady-state current, equations (46) and (47), can be used for the calculations as far as the assumption of a single (at maximum) molecule in the pore holds.
1DD-2MS mapping, strong binding:
The translocation current through a deep potential well reads, respectively,
At the high concentrations, the currents reach their maximum values,
In contrast with the barrier, the saturated translocation current through the well is larger when the molecules are added from the side opposite to the well (the binding site next to the exit).
In the case of strong binding site nonadjacent to the channel ends, i.e.,
have physical meaning. Indeed, in this case, the binding site is single and the 1DD-2MS bridging conditions, equations (32)–(34), can be satisfied for a range of concentrations from zero till the saturation.
Analytic one-barrier-one-binding site model
Let us suggest that the PMF is a sum of nonoverlapping rectangular barrier of magnitude
The diffusion coefficient,
The channel occupancy for the cis molecule addition, for the trans addition and for the symmetric addition, reads, respectively,
Here, the following “distances” to the barrier/well from the trans and cis sides are introduced for more symmetric formulas,

Schematic diagrams of a pore with (a) a single barrier PMF, (b) a single binding site PMF, (c) a barrier and a binding site PMF, and (d) a two-barriers-one-binding-site PMF. The dotted lines represent the pore length.
Now, we will assume that the barrier and the binding site are strong, i.e.,
That is, in the presence of a strong barrier, the diffusion current is independent on the binding site. The translocation current reads,
At high concentrations, currents reach their maximum values,
As,
Markov chain model
The elementary cases discussed in the two previous subsections allows one to write down the transition rates for a more general situation of a PMF presented as a sequence of binding sites, separated by large rectangular barriers. The molecule moves in the pore by jumping between neighboring binding sites. We number the binding sites from 1 (the first site on the cis site) to
Here,
This mapping of the diffusion onto the Markov chain model assumes that the barriers are large enough for the Markov timescale conditions, like those given by equations (33)–(32), being satisfied.
By using the transition rates obtained by the Markov chain mapping, one can build a kinetic Monte-Carlo scheme [49] to simulate the translocation kinetics. This approach makes it possible to limit the occupancy for each binding site (not the whole pore), thus enhancing the possibility to study molecular crowding in the channel.
Equations (63) and (64) can be interpreted as a diffusion analog of the classical dynamic transition-state theory [26], where
Analytic two-barriers-one-binding site model
A more realistic example is provided by the two-barrier-one-binding-site model, introduced earlier to model the transport of metabolites with binding in specific porins [22,48].
Here, Markov chain mapping formulas, (63)–(66), give (the diffusion coefficient is assumed constant along the pore for simplicity),
Then, one obtains,
The translocation current is given by equations (31). The maximum cis/trans asymmetry,
is determined by the magnitude and the width of the barriers.
This model contains two previous examples as special cases: (i) a symmetric case, where the barriers from the two sides are the same; (ii) a strongly asymmetric case, where the barriers are different.
Thus, in the asymmetric case,
Discussion
Numerical validation
In order to show the potency of the method introduced, we considered a realistic model system, the transport of a beta-lactamase inhibitor (tazobactam) through the main porins expressed in the outer membrane of Escherichia coli, OmpF and OmpC. Molecules as tazobactam are used in combination with beta-lactam antibiotics because they bind strongly to beta-lactamase enzyme inhibiting them from modifying chemically antibiotics and thus revert their antibacterial efficacy [50,51]. Today, there is a wide interest in developing such inhibitors to counteract the spreading of resistant pathogens, and one common problem is to predict their transport through porins. By using multiple walkers, metadynamics simulations we reconstructed the one-dimensional free energy surface for the interaction of tazobactam with the two porins, as done for other inhibitors recently [52]. As we can see in Figure 2, the two FES profiles are a complex combination of local minima and barriers. In particular, we note a main central barrier and some minima on the positive region z (extracellular region), thus reminding the simple model of a barrier and a binding site. By calculating the diffusion current with equation (3) and thus applying equation (31), we obtained the total currents as function of gradient concentration, as reported in Figure 3.

Free energy surface profiles vs tazobactam position along the diffusion axis inside the OmpF and OmpC porins.

Current of tazobactam through the OmpF and OmpC porins versus the concentration of the molecule being added to one side only: Cis to Trans (left panel) and Trans to Cis (right panel) currents.
A comparison of the two currents at constant concentration is not sufficient, as, for example, for the cis to trans current in the left panel of Figure 3. Though the barrier is lower in OmpF, than OmpC, and correspondingly larger the current, in the case of OmpF the saturation occurs before, and at concentrations higher than 10 mM, the current through OmpF becomes lower.
It is also interesting to calculate the rate of the cis → trans and trans → cis currents, as provided in Figure 4. As we illustrated with in the aforementioned simple examples, the two current are symmetric at low concentration, while at high concentration, there is an asymmetry modulated by the position of minima and barrier, see equations (57), (58), and (31):
The current trans → cis is always superior to the cis →trans, because the highest barrier is on the trans side.
The presence of minima is responsible for the saturation of the current at high concentration, and the change in current is more drastic on the cis → trans current because the minima are on the cis side.

Ratio between the Cis to Trans and Trans to Cis current of tazobactam (being added to one side, Cis or Trans, respectively) through the OmpF and OmpC porins.
Experimental validation
Our model can be employed to guide and plan experiments till the interpretation of results. A quick look at recent published data shows that kanamycin, a positively charged molecule, is transported through both OmpF/OmpC [53]. At high negative external potential applied, the association rate in OmpF saturates, contrarily to OmpC. Because kanamycin is a molecule charged positive, a high negative potential applied means to force a large association rate toward the pore, corresponding to the case described here as an increase in the gradient concentration. Interestingly, the obtained results are described by the top panel of Figure 3. Increasing more and more the external potential we would see higher association rate through OmpC than OmpF.
Further, the behavior described in Figure 4, an asymmetric transport upon asymmetric addition of molecules in solution, was measured for maltoheaxose through the maltoporin trimeric channel [21]. This sugar molecules has higher association rate when added on the trans side, exactly as for tazobactam through OmpF/OmpC. In that case, the asymmetry is visible already at 0.01 mM concentration, because of the high specificity of the maltoporin channel for sugar molecules.
Conclusion
We combined the diffusion equation with the Markov state model to obtain expression for the passive current at any concentrations, thus taking into account the effect of crowding (saturation) on transport. It is demonstrated that, in a biologically relevant case of transport of antibiotics through bacterial porins, the transition between Fick’s-low regime (when the influx current is linear with the antibiotic concentration gradient) to the saturated influx current (independent on the concentration) can happen at the 0.1–1 mM concentrations. The stronger is the interaction between the molecule and the pore, the lower is the transitional concentration between the two transport regimes. This is particularly important to take into account when a model of whole-cell accumulation of molecules is designed and experimental results have to be interpreted [54].
Further, some biological conclusion can be drawn by looking at the interaction of particles inside the pores, that reflects the structural features of nanopores. For example, from the free energy profiles presented for the real case, we can conclude that by expressing these channels bacteria maximize the outward current. This means that in case of antibiotics, they can penetrate with a given inward current when present at high concentration on the cis side. Because the periplasmic space has a small volume, a low amount of molecules accumulated with the inward current are sufficient to saturate the interior [15]. However, when the external pressure of concentration disappears on the cis side, the outward current empties the periplasmic space easily. Thus, it is important to keep an almost constant external concentration for antibiotics to be effective on the small periplasmic volume.
Acknowledgements
MC thanks partial support from the Translocation consortium (http://www.translocation.com) through the IMI Joint Undertaking under grant agreement no. 115525, resources which are composed of financial contribution from the European Union’s seventh framework programme (FP7/2007-2013) and EFPIA companies’ kind contribution. IVB acknowledges the financial support by DAAD (Deutscher Akademischer Austausch Dienst) during his research stay period at Jacobs University Bremen. SM is funded through the bilateral Russian (RFBR)-Italian(CNR) research project No. 20-58-7802(RFBR) and B55F21000620005 (CNR). The authors gratefully acknowledge the computing time granted through the ISCRA-B project “PREDICT” on the supercomputer Marconi100 at CINECA, IT.
-
Conflict of interest: The authors state no conflict of interest.
-
Data availability statement: The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.
References
[1] Vergalli J, Bodrenko IV, Masi M, Moynié L, Acosta-Gutierrez S, Naismith JH, et al. Porins and small-molecule translocation across the outer membrane of Gram-negative bacteria. Nature Rev Microbiol. 2019 Dec;33(3):1831–913. 10.1038/s41579-019-0294-2Search in Google Scholar PubMed
[2] Ujwal R, Cascio D, Colletier JP, Faham S, Zhang J, Toro L, et al. The crystal structure of mouse VDAC1 at 2.3 Åresolution reveals mechanistic insights into metabolite gating. Proc National Acad Sci. 2008;105(46):17742–7. 10.1073/pnas.0809634105Search in Google Scholar PubMed PubMed Central
[3] van den Berg B, Prathyusha Bhamidimarri S, Dahyabhai Prajapati J, Kleinekathöfer U, Winterhalter M. Outer-membrane translocation of bulky small molecules by passive diffusion. Proc National Acad Sci. 2015 Jun;112(23):E2991–9. Available from: http://www.pnas.org/lookup/doi/10.1073/pnas.1424835112. 10.1073/pnas.1424835112Search in Google Scholar PubMed PubMed Central
[4] Biswas S, Mohammad MM, Patel DR, Movileanu L, van den Berg B. Structural insight into OprD substrate specificity. Nature Struct & Mol Biol. 2007 Nov;14(11):1108–9. 10.1038/nsmb1304Search in Google Scholar PubMed
[5] Ferrara LGM, Wallat GD, Moynié L, Dhanasekar NN, Aliouane S, Acosta-Gutierrez S, et al. MOMP from campylobacter jejuni is a trimer of 18-stranded β-barrel monomers with a Ca(2+) ion bound at the constriction zone. J Mol Biol. 2016 Nov;428(22):4528–43. 10.1016/j.jmb.2016.09.021Search in Google Scholar PubMed PubMed Central
[6] Pathania M, Acosta-Gutierrez S, Bhamidimarri SP, Baslé A, Winterhalter M, Ceccarelli M, et al. Unusual constriction zones in the major Porins OmpU and OmpT from Vibrio cholerae. Structure. 2018;26(5):708–21. 10.1016/j.str.2018.03.010Search in Google Scholar PubMed
[7] Acosta-Gutierrez S, Ferrara L, Pathania M, Masi M, Wang J, Bodrenko I, et al. Getting drugs into gram-negative bacteria: Rational rules for permeation through general porins. ACS Infect Diseases. 2018 Aug;4(10):1487–98. 10.1021/acsinfecdis.8b00108Search in Google Scholar PubMed
[8] Marbach S, Dean DS, Bocquet Lxr. Transport and dispersion across wiggling nanopores. Nature Phys. 2018 Jul;14(11):1–6. 10.1038/s41567-018-0239-0Search in Google Scholar
[9] Gravelle S, Joly L, Ybert C, Bocquet L. Large permeabilities of hourglass nanopores: from hydrodynamics to single file transport. J Chem Phys. 2014 Nov;141(18):18C526. 10.1063/1.4897253Search in Google Scholar PubMed
[10] Willems K, Van Meervelt V, Wloka C, Maglia G. Single-molecule nanopore enzymology. Philosoph Trans R Soc London B Biol Sci. 2017 Aug;372(1726):20160230. 10.1098/rstb.2016.0230Search in Google Scholar PubMed PubMed Central
[11] Chowdhury R, Ren T, Shankla M, Decker K, Grisewood M, Prabhakar J, et al. PoreDesigner for tuning solute selectivity in a robust and highly permeable outer membrane pore. Nature Commun. 2018 Sep;9(1):3661–10. 10.1038/s41467-018-06097-1Search in Google Scholar PubMed PubMed Central
[12] Qing Y, Ionescu SA, Pulcu GS, Bayley H. Directional control of a processive molecular hopper. Science. 2018 Aug;361(6405):908–12. 10.1126/science.aat3872Search in Google Scholar PubMed
[13] Ouldali H, Sarthak K, Ensslen T, Piguet F, Manivet P, Pelta J, et al. Electrical recognition of the twenty proteinogenic amino acids using an aerolysin nanopore. Nature Biotechnol. 2020 Jan;38:1–10. 10.1038/s41587-019-0345-2Search in Google Scholar PubMed PubMed Central
[14] Richter MF, Drown BS, Riley AP, Garcia A, Shirai T, Svec RL, et al. Predictive compound accumulation rules yield a broad-spectrum antibiotic. Nature Publishing Group. 2017 May;545(7654):299–304. 10.1038/nature22308Search in Google Scholar PubMed PubMed Central
[15] Brönstrup HPVFSKHMAGRAHM, Fetz V, Hotop Sk, García-Rivera MA, Heumann A, Brönstrup M. Quantification of uptake in gram-negative bacteria. Anal Chem. 2018 Nov;91(3):1863–72. 10.1021/acs.analchem.8b03586Search in Google Scholar PubMed
[16] Masi M, Réfrégiers M, Pos KM, Pagès JM. Mechanisms of envelope permeability and antibiotic influx and efflux in Gram-negative bacteria. Nature Microbiol. 2017 Feb;2(3):17001–7. 10.1038/nmicrobiol.2017.1Search in Google Scholar PubMed
[17] Westfall DA, Krishnamoorthy G, Wolloscheck D, Sarkar R, Zgurskaya HI, Rybenkov VV. Bifurcation kinetics of drug uptake by Gram-negative bacteria. PLoS One. 2017 Sep;12(9):e0184671. 10.1371/journal.pone.0184671Search in Google Scholar PubMed PubMed Central
[18] Saha P, Sikdar S, Krishnamoorthy G, Zgurskaya HI, Rybenkov VV. Drug Permeation against efflux by two transporters. ACS Infect Diseases. 2020 Feb;6:747–58. 10.1021/acsinfecdis.9b00510Search in Google Scholar PubMed PubMed Central
[19] Hänggi P, Talkner P, Borkovec M. Reaction-rate theory: fifty years after Kramers. Rev Mod Phys. 1990 Apr;62(2):251–341. Available from: https://link.aps.org/doi/10.1103/RevModPhys.62.251. 10.1103/RevModPhys.62.251Search in Google Scholar
[20] Nestorovich EM, Danelon C, Winterhalter M, Bezrukov SM. Designed to penetrate: time-resolved interaction of single antibiotic molecules with bacterial pores. Proc Nat Acad Sci US Am. 2002 Jul;99(15):9789–94. 10.1073/pnas.152206799Search in Google Scholar PubMed PubMed Central
[21] Kullman L, Winterhalter M, Bezrukov SM. Transport of maltodextrins through maltoporin: a single-channel study. Biophys J. 2002 Feb;82(2):803–12. Available from: http://eutils.ncbi.nlm.nih.gov/entrez/eutils/elink.fcgi?dbfrom=pubmed&id=11806922&retmode=ref&cmd=prlinks. 10.1016/S0006-3495(02)75442-8Search in Google Scholar
[22] Schwarz G, Danelon C, Winterhalter M. On translocation through a membrane channel via an internal binding site: kinetics and voltage dependence. Biophys J. 2003 May;84(5):2990–8. 10.1016/S0006-3495(03)70025-3Search in Google Scholar
[23] Ghai I, Pira A, Scorciapino MA, Bodrenko I, Benier L, Ceccarelli M, et al. General method to determine the flux of charged molecules through nanopores applied to beta-lactamase inhibitors and OmpF. J Phys Chem Lett. 2017 mar;8(6):1295–301. Available from: http://pubs.acs.org/doi/10.1021/acs.jpclett.7b00062. 10.1021/acs.jpclett.7b00062Search in Google Scholar
[24] Bauer WR, Nadler W. Molecular transport through channels and pores: effects of in-channel interactions and blocking. Proc National Acad Sci. 2006;103(31):11446–51. Available from: https://www.pnas.org/content/103/31/11446. 10.1073/pnas.0601769103Search in Google Scholar
[25] Bezrukov SM, Berezhkovskii AM, Szabo A. Diffusion model of solute dynamics in a membrane channel: mapping onto the two-site model and optimizing the flux. J Chem Phys. 2007;127(11):115101. Available from: https://doi.org/10.1063/1.2766720. 10.1063/1.2766720Search in Google Scholar
[26] Kramers HA. Brownian motion in a field of force and the diffusion model of chemical reactions. Physica. 1940 Apr;7(4):284–304. Available from: http://linkinghub.elsevier.com/retrieve/pii/S0031891440900982. 10.1016/S0031-8914(40)90098-2Search in Google Scholar
[27] Gardiner CW. Handbook of Stochastic Methods for Physics, Chemistry, and the Natural Sciences. Berlin: Springer; 2004. 10.1007/978-3-662-05389-8Search in Google Scholar
[28] Landau LD, Lifshitz EM, Pitaevskiĭ LP, Sykes JB, Kearsley M, Statistical Physics. Part 1 3rd ed. Amsterdam; London: Elsevier Butterworth Heinemann; 2008. Published in 2 parts. Search in Google Scholar
[29] Zwanzig R. Diffusion past an entropy barrier. J Phys Chem. 1992 May;96(10):3926–30. Available from: http://pubs.acs.org/doi/abs/10.1021/j100189a004. 10.1021/j100189a004Search in Google Scholar
[30] Kalinay P, Percus JK. Corrections to the Fick-Jacobs equation. Phys Rev E. 2006 Oct;74(4):041203. Available from: https://link.aps.org/doi/10.1103/PhysRevE.74.041203. 10.1103/PhysRevE.74.041203Search in Google Scholar PubMed
[31] Bradley RM. Diffusion in a two-dimensional channel with curved midline and varying width: Reduction to an effective one-dimensional description. Phys Rev E. 2009 Dec;80(6):061142. Available from: https://link.aps.org/doi/10.1103/PhysRevE.80.061142. 10.1103/PhysRevE.80.061142Search in Google Scholar PubMed
[32] Berezhkovskii A, Szabo A. Time scale separation leads to position-dependent diffusion along a slow coordinate. J Chem Phys. 2011 Aug;135(7):074108. Available from: http://aip.scitation.org/doi/10.1063/1.3626215. 10.1063/1.3626215Search in Google Scholar
[33] Berezhkovskii AM, Pustovoit MA, Bezrukov SM. Channel-facilitated membrane transport: Transit probability and interaction with the channel. J Chem Phys. 2002 June;116(22):9952–6. Available from: http://aip.scitation.org/doi/10.1063/1.1475758. 10.1063/1.1475758Search in Google Scholar
[34] Stratonovich RL. Radiotekh Elektron (Moscow). 1958;3:497. Search in Google Scholar
[35] Reguera D, Luque A, Burada PS, Schmid G, Rubí JM, Hänggi P. Entropic splitter for particle separation. Phys Rev Lett. 2012 Jan;108(2):020604. Available from: https://link.aps.org/doi/10.1103/PhysRevLett.108.020604. 10.1103/PhysRevLett.108.020604Search in Google Scholar
[36] Acosta-Gutierrez S, Bodrenko IV, Ceccarelli M. The influence of permeability through bacterial porins in whole-cell compound accumulation. Antibiotics (Basel, Switzerland) 2021 May;10(6):635. Available from: https://www.mdpi.com/2079-6382/10/6/635/htm. 10.3390/antibiotics10060635Search in Google Scholar
[37] Im W, Roux B. Ion permeation and selectivity of OmpF Porin: A theoretical study based on molecular dynamics, brownian dynamics, and continuum electrodiffusion theory. J Mol Biol. 2002 Sep;322(4):851–69. Available from: http://www.sciencedirect.com/science/article/pii/S0022283602007787, http://linkinghub.elsevier.com/retrieve/pii/S0022283602007787. 10.1016/S0022-2836(02)00778-7Search in Google Scholar
[38] Hummer G. Position-dependent diffusion coefficients and free energies from Bayesian analysis of equilibrium and replica molecular dynamics simulations. New J Phys. 2005 Feb;7:34. Available from: http://stacks.iop.org/1367-2630/7/i=1/a=034?key=crossref.e1c728f07773a70730fc6b9bedef5646. 10.1088/1367-2630/7/1/034Search in Google Scholar
[39] Wilson MA, Nguyen TH, Pohorille A. Combining molecular dynamics and an electrodiffusion model to calculate ion channel conductance. J Chem Phys. 2014 Dec;141(22):22D519. Available from: http://aip.scitation.org/doi/10.1063/1.4900879. 10.1063/1.4900879Search in Google Scholar
[40] Berezhkovskii AM, Makarov DE. Communication: coordinate-dependent diffusivity from single molecule trajectories. J Chem Phys. 2017 Nov;147(20):201102. Available from: http://aip.scitation.org/doi/10.1063/1.5006456. 10.1063/1.5006456Search in Google Scholar
[41] Torrie GM, Valleau JP. Monte Carlo free energy estimates using nonBoltzmann sampling: Application to the sub-critical Lennard-Jones fluid. Chem Phys Lett. 1974 Oct;28(4):578–81. Available from: http://linkinghub.elsevier.com/retrieve/pii/0009261474801090. 10.1016/0009-2614(74)80109-0Search in Google Scholar
[42] Laio A, Parrinello M. Escaping free-energy minima. Proc Natl Acad Sci. 2002 Oct;99(20):12562–6. Available from: http://www.jstor.org/stable/3073262, http://www.pnas.org/cgi/doi/10.1073/pnas.202427399. 10.1073/pnas.202427399Search in Google Scholar PubMed PubMed Central
[43] Tiwary P, Parrinello M. From metadynamics to dynamics. Phys Rev Lett. 2013 Dec;111(23):230602. Available from: https://link.aps.org/doi/10.1103/PhysRevLett.111.230602. 10.1103/PhysRevLett.111.230602Search in Google Scholar PubMed
[44] Acosta-Gutierrez S, Scorciapino MA, Bodrenko IV, Ceccarelli M. Filtering with electric field: the case of E. coli porins. J Phys Chem Lett. 2015 May;6(10):1807–12. Available from: http://pubs.acs.org/doi/10.1021/acs.jpclett.5b00612. 10.1021/acs.jpclett.5b00612Search in Google Scholar
[45] D’Agostino T, Salis S, Ceccarelli M. A kinetic model for molecular diffusion through pores. Biochim Biophys Acta - Biomembr. 2016 Jul;1858(7):1772–7. Available from: http://dx.doi.org/10.1016/j.bbamem.2016.01.004, http://linkinghub.elsevier.com/retrieve/pii/S0005273616000055. Search in Google Scholar
[46] Bajaj H, Acosta Gutierrez S, Bodrenko I, Malloci G, Scorciapino MA, Winterhalter M, et al. Bacterial outer membrane porins as electrostatic nanosieves: exploring transport rules of small polar molecules. ACS Nano. 2017 Jun;11(6):5465–73. Available from: http://pubs.acs.org/doi/10.1021/acsnano.6b08613. 10.1021/acsnano.6b08613Search in Google Scholar
[47] Colquhoun D, Hawkes AG. Relaxation and fluctuations of membrane currents that flow through drug-operated channels. Proc R Soc B Biol Sci. 1977 Nov;199(1135):231–62. 10.1098/rspb.1977.0137Search in Google Scholar
[48] Nekolla S, Andersen C, Benz R. Noise analysis of ion current through the open and the sugar-induced closed state of the LamB channel of Escherichia coli outer membrane: evaluation of the sugar binding kinetics to the channel interior, Biophys J. 1994 May;66(5):1388–97. 10.1016/S0006-3495(94)80929-4Search in Google Scholar
[49] Ceccarelli M, Vargiu AV, Ruggerone P. A kinetic Monte Carlo approach to investigate antibiotic translocation through bacterial porins. J Phys Condensed Matter Inst Phys J. 2012 Mar;24(10):104012. 10.1088/0953-8984/24/10/104012Search in Google Scholar PubMed
[50] Papp-Wallace KM. The latest advances in beta-lactam/beta-lactamase inhibitor combinations for the treatment of Gram-negative bacterial infections. Expert Opinion Pharmacotherapy. 2019 Nov;20(17):2169–84. Available from: https://doi.org/10.1080/14656566.2019.1660772. 10.1080/14656566.2019.1660772Search in Google Scholar PubMed PubMed Central
[51] Davies DT, Leiris S, Zalacain M, Sprynski N, Castandet J, Bousquet J, et al. Discovery of ANT3310, a novel broad-spectrum serine beta-Lactamase inhibitor of the Diazabicyclooctane class, which strongly potentiates meropenem activity against Carbapenem-resistant enterobacterales and Acinetobacter baumannii. J Med Chem. 2020 Dec;63(24):15802–20. Available from: https://pubs.acs.org/doi/10.1021/acs.jmedchem.0c01535. 10.1021/acs.jmedchem.0c01535Search in Google Scholar PubMed
[52] Pira A, Scorciapino MA, Bodrenko IV, Bosin A, Acosta-Gutierrez S, Ceccarelli M. Permeation of beta-Lactamase inhibitors through the general porins of gram-negative bacteria. Molecules 2020 Dec;25(23):5747. 10.3390/molecules25235747Search in Google Scholar PubMed PubMed Central
[53] Bafna JA, Sans-Serramitjana E, Acosta-Gutierrez S, Bodrenko IV, Hörömpöli D, Berscheid A. Kanamycin uptake into Escherichia coli is facilitated by OmpF and OmpC Porin channels located in the outer membrane. ACS Infect Diseases. 2020 Jul;6(7):1855–65. 10.1021/acsinfecdis.0c00102Search in Google Scholar PubMed
[54] Zgurskaya HI, Rybenkov VV. Permeability barriers of Gram-negative pathogens. Ann NY Acad Sci. 2020 Jan;1459(1):5–18. 10.1111/nyas.14134Search in Google Scholar PubMed PubMed Central
© 2022 Igor V. Bodrenko et al., published by De Gruyter
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Research Articles
- Diagnostic accuracy of genetic markers for identification of the Lr46/Yr29 “slow rusting” locus in wheat (Triticum aestivum L.)
- NADPH-derived ROS generation drives fibrosis and endothelial-to-mesenchymal transition in systemic sclerosis: Potential cross talk with circulating miRNAs
- Effect of omega-3 fatty acids on the telomere length: A mini meta-analysis of clinical trials
- Analysis of differentially expressed genes and signaling pathways involved in atherosclerosis and chronic obstructive pulmonary disease
- The epigenetic dimension of protein structure
- Gamma-induced mutants of Bacillus and Streptomyces display enhanced antagonistic activities and suppression of the root rot and wilt diseases in pulses
- Corticosterone potentiates ochratoxin A-induced microglial activation
- Supercomplex supercomplexes: Raison d’etre and functional significance of supramolecular organization in oxidative phosphorylation
- Insights into functional connectivity in mammalian signal transduction pathways by pairwise comparison of protein interaction partners of critical signaling hubs
- The effects of supplementation of Nannochloropsis oculata microalgae on biochemical, inflammatory and antioxidant responses in diabetic rats
- Molecular epidemiology of human papillomavirus in pregnant women in Burkina Faso
- Review Articles
- Interaction of cervical microbiome with epigenome of epithelial cells: Significance of inflammation to primary healthcare
- Seaweeds’ pigments and phenolic compounds with antimicrobial potential
- The capture of host cell’s resources: The role of heat shock proteins and polyamines in SARS-COV-2 (COVID-19) pathway to viral infection
- Erratum
- Erratum to “Plant growth-promoting properties of bacterial endophytes isolated from roots of Thymus vulgaris L. and investigate their role as biofertilizers to enhance the essential oil contents”
- Special Issue on XXV Congress of the Italian Society for Pure and Applied Biophysics
- Low-temperature librations and dynamical transition in proteins at differing hydration levels
- The phosphoinositide PI(3,5)P2 inhibits the activity of plant NHX proton/potassium antiporters: Advantages of a novel electrophysiological approach
- Targeted photoimmunotherapy for cancer
- Calorimetry of extracellular vesicles fusion to single phospholipid membrane
- Calcium signaling in prostate cancer cells of increasing malignancy
- Oxygen diffusion pathways in mutated forms of a LOV photoreceptor from Methylobacterium radiotolerans: A molecular dynamics study
- A photosensitizing fusion protein with targeting capabilities
- Ion channels and neuronal excitability in polyglutamine neurodegenerative diseases
- Is styrene competitive for dopamine receptor binding?
- Diffusion of molecules through nanopores under confinement: Time-scale bridging and crowding effects via Markov state model
- Quantitative active super-resolution thermal imaging: The melanoma case study
- Innovative light sources for phototherapy
- Electrophysiological study of the effects of side products of RuBi-GABA uncaging on GABAA receptors in cerebellar granule cells
- Subcellular elements responsive to the biomechanical activity of triple-negative breast cancer-derived small extracellular vesicles
Articles in the same Issue
- Research Articles
- Diagnostic accuracy of genetic markers for identification of the Lr46/Yr29 “slow rusting” locus in wheat (Triticum aestivum L.)
- NADPH-derived ROS generation drives fibrosis and endothelial-to-mesenchymal transition in systemic sclerosis: Potential cross talk with circulating miRNAs
- Effect of omega-3 fatty acids on the telomere length: A mini meta-analysis of clinical trials
- Analysis of differentially expressed genes and signaling pathways involved in atherosclerosis and chronic obstructive pulmonary disease
- The epigenetic dimension of protein structure
- Gamma-induced mutants of Bacillus and Streptomyces display enhanced antagonistic activities and suppression of the root rot and wilt diseases in pulses
- Corticosterone potentiates ochratoxin A-induced microglial activation
- Supercomplex supercomplexes: Raison d’etre and functional significance of supramolecular organization in oxidative phosphorylation
- Insights into functional connectivity in mammalian signal transduction pathways by pairwise comparison of protein interaction partners of critical signaling hubs
- The effects of supplementation of Nannochloropsis oculata microalgae on biochemical, inflammatory and antioxidant responses in diabetic rats
- Molecular epidemiology of human papillomavirus in pregnant women in Burkina Faso
- Review Articles
- Interaction of cervical microbiome with epigenome of epithelial cells: Significance of inflammation to primary healthcare
- Seaweeds’ pigments and phenolic compounds with antimicrobial potential
- The capture of host cell’s resources: The role of heat shock proteins and polyamines in SARS-COV-2 (COVID-19) pathway to viral infection
- Erratum
- Erratum to “Plant growth-promoting properties of bacterial endophytes isolated from roots of Thymus vulgaris L. and investigate their role as biofertilizers to enhance the essential oil contents”
- Special Issue on XXV Congress of the Italian Society for Pure and Applied Biophysics
- Low-temperature librations and dynamical transition in proteins at differing hydration levels
- The phosphoinositide PI(3,5)P2 inhibits the activity of plant NHX proton/potassium antiporters: Advantages of a novel electrophysiological approach
- Targeted photoimmunotherapy for cancer
- Calorimetry of extracellular vesicles fusion to single phospholipid membrane
- Calcium signaling in prostate cancer cells of increasing malignancy
- Oxygen diffusion pathways in mutated forms of a LOV photoreceptor from Methylobacterium radiotolerans: A molecular dynamics study
- A photosensitizing fusion protein with targeting capabilities
- Ion channels and neuronal excitability in polyglutamine neurodegenerative diseases
- Is styrene competitive for dopamine receptor binding?
- Diffusion of molecules through nanopores under confinement: Time-scale bridging and crowding effects via Markov state model
- Quantitative active super-resolution thermal imaging: The melanoma case study
- Innovative light sources for phototherapy
- Electrophysiological study of the effects of side products of RuBi-GABA uncaging on GABAA receptors in cerebellar granule cells
- Subcellular elements responsive to the biomechanical activity of triple-negative breast cancer-derived small extracellular vesicles