Influence of cholesterol on hydrogen-bond dynamics of water molecules in lipid-bilayer systems at varying temperatures

Kokoro Shikata Division of Chemical Engineering, Department of Materials Engineering Science, Graduate School of Engineering Science, Osaka University, Toyonaka, Osaka 560-8531, Japan    Kento Kasahara kasahara@cheng.es.osaka-u.ac.jp Division of Chemical Engineering, Department of Materials Engineering Science, Graduate School of Engineering Science, Osaka University, Toyonaka, Osaka 560-8531, Japan    Nozomi Morishita Watanabe Division of Chemical Engineering, Department of Materials Engineering Science, Graduate School of Engineering Science, Osaka University, Toyonaka, Osaka 560-8531, Japan    Hiroshi Umakoshi Division of Chemical Engineering, Department of Materials Engineering Science, Graduate School of Engineering Science, Osaka University, Toyonaka, Osaka 560-8531, Japan    Kang Kim kk@cheng.es.osaka-u.ac.jp Division of Chemical Engineering, Department of Materials Engineering Science, Graduate School of Engineering Science, Osaka University, Toyonaka, Osaka 560-8531, Japan    Nobuyuki Matubayasi nobuyuki@cheng.es.osaka-u.ac.jp Division of Chemical Engineering, Department of Materials Engineering Science, Graduate School of Engineering Science, Osaka University, Toyonaka, Osaka 560-8531, Japan
(July 4, 2024)
Abstract

Cholesterol (Chol) plays a crucial role in shaping the intricate physicochemical attributes of biomembranes, exerting considerable influence on water molecules proximal to the membrane interface. In this study, we conducted molecular dynamics simulations on the bilayers of two lipid species, dipalmitoyl phosphatidylcholine (DPPC) and palmitoyl sphingomyelin (PSM); they are distinct with respect to the structures of the hydrogen-bond (H-bond) acceptors. Our investigation focuses on the dynamic properties and H-bonds of water molecules in the lipid-membrane systems, with particular emphasis on the influence of Chol at varying temperatures. Notably, in the gel phase at 303 K, the presence of Chol extends the lifetimes of H-bonds of the oxygen atoms acting as H-bond acceptors within DPPC with water molecules by a factor of 1.5 to 2.5. In the liquid-crystalline phase at 323 K, on the other hand, H-bonding dynamics with lipid membranes remain largely unaffected by Chol. This observed shift in H-bonding states serves as a crucial key to unraveling the subtle control mechanisms governing water dynamics in lipid-membrane systems.

I Introduction

Lipid bilayers, serving as the fundamental architectural frameworks of biological membranes, form stable aggregates through the amphiphilic effect inherent to lipid molecules. The attributes of lipid bilayers vary diversely with the specific type and composition of lipid molecules, giving rise to distinctive structures, such as gel and liquid-crystalline phases. [1] The interactions with water are also a key to self-organization,[2] and water properties are affected by the states of lipid molecules in turn.

Cholesterol (Chol), extensively investigated as a constituent of bio-related membranes, exhibits significant influence on structures of lipid bilayers. [3, 4] In particular, Chol enhances the packing density and rigidity of the lipid, thereby modulating membrane fluidity. Furthermore, water molecules play a crucial role in the structure and function of biological membranes, influencing electrostatic properties, solute exchange, and protein function. [5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16] Thus, it is imperative to elucidate the structure and dynamics of water molecules at the membrane interface, which is expected to differ from those in the bulk owing to the interaction with hydrophilic groups on the lipid head.

Molecular dynamics (MD) simulation is a valuable tool for investigating lipid bilayers, providing molecular-level insights into not only lipid properties but also their interactions with other molecules.  [17, 18, 19, 20, 21, 22, 23, 24, 25, 26] Numerous investigations have also been conducted for water in the interface region, encompassing the distribution of water molecules, reorientation dynamics, mean square displacement, and hydrogen-bond (H-bond) dynamics. [27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40] Specifically, the slowdown of the H-bond dynamics from the bulk to the center of the membrane has been demonstrated. [39, 40] In addition, MD simulations have been used to study the structure and dynamics of lipid bilayers containing phospholipids and Chol. [41, 42, 43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53, 54] These simulations have provided insights into the influence of Chol on a variety of phospholipids differing in the headgroup and tail.

Despite the numerous studies mentioned above, comprehending the water state proximal to lipid membranes, as well as understanding its connection with the membrane state in the presence of Chol, remains a significant problem. Interestingly, experimental observations have indicated that Chol was found to accelerate the water dynamics in the dipalmitoyl phosphatidylcholine (DPPC) membrane interface. [55, 56] Conversely, it has been found that water dynamics decelerate within the interior region of lipid bilayers with increasing Chol concentration. [55]

MD simulations have elucidated that the acceleration of water dynamics at the interface, particularly notable at high Chol concentrations up to 50%, arises from the inhibition of H-bonds between two oxygen atoms of lipid molecules. [49] A more recent MD study conducted a detailed analysis of the H-bond network of water within the DPPC membrane in the presence of Chol. The results unveiled that Chol fosters more bulk-like water at the membrane interface, leading to increased local water density and accelerated water dynamics. [57]

In this study, we conducted MD simulations of two types of lipid bilayers comprising of DPPC and palmitoyl sphingomyelin (PSM) with the presence of Chol. While the Chol concentrations investigated were 0 and 10%, the temperature effect was examined at 303 K and 323 K. The DPPC and PSM membranes are at the gel and liquid-crystalline phases at 303 K and 323 K, respectively. We investigate the microscopic hydration structure and dynamics by considering acceptor sites of lipid molecules, which form H-bonds with the hydrogen atoms of water molecules. Between DPPC and PSM, the choline and phosphate groups are identical, as shown in Fig. 1. There are differences in the degree of carbon chain saturation and the functional group acting as the H-bonding site. Thus, our MD investigations provide insights into H-bonds influenced by Chol, taking into account the molecular structures of the lipids and the environmental effects from the membrane composition and temperature.

II Simulation details

Refer to caption
Figure 1: Structures of the lipid and Chol molecules studied in this paper.
Table 1: Numbers of lipid (DPPC or PSM), Chol, and water molecules in mixture and pure membrane systems.
mixture pure
DPPC / PSM 200 200
Chol 22 -
Water 22000 20000
Refer to caption
Figure 2: Number density distributions of lipid carbon chain (tail), water molecule oxygen (OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT), and Chol along the z𝑧zitalic_z-direction. Solid lines represent pure membrane systems, while dashed lines represent systems containing Chol. (a) and (b) correspond to DPPC, and (c) and (d) to PSM.

The structures of the lipid molecules, DPPC, PSM, and Chol are depicted in Fig. 1. The hydrophilic moiety is common between DPPC and PSM, while they are different for the hydrophobic portion and the distributions of oxygen and nitrogen atoms.

The lipid bilayer system was constructed using CHARMM-GUI, [58, 59, 60, 61, 62] incorporating 200 lipid molecules with 10% of Chol if present, as listed in Table 1. For each lipid and Chol composition, 100 water molecules per molecule of lipid and Chol were added to create the lipid bilayer system. Three different initial configurations were prepared for each composition, employing the CHARMM36 force field for DPPC, PSM, and Chol, [63] and the CHARMM-compatible TIP3P model for water molecules. [64] All the MD simulations were performed using Gromacs 2022.4.  [65]

Refer to caption
Figure 3: (a) Schematic illustration of lipid phosphorus atom position relative to the instantaneous average in the upper and lower leaflet, denoted as zGupper(t)subscriptsuperscript𝑧upper𝐺𝑡z^{\mathrm{upper}}_{G}(t)italic_z start_POSTSUPERSCRIPT roman_upper end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_t ) and zGlower(t)subscriptsuperscript𝑧lower𝐺𝑡z^{\mathrm{lower}}_{G}(t)italic_z start_POSTSUPERSCRIPT roman_lower end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_t ), respectively. zjP(t)zGupper(t)subscriptsuperscript𝑧P𝑗𝑡subscriptsuperscript𝑧upper𝐺𝑡z^{\mathrm{P}}_{j}(t)-z^{\mathrm{upper}}_{G}(t)italic_z start_POSTSUPERSCRIPT roman_P end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) - italic_z start_POSTSUPERSCRIPT roman_upper end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_t ) and zjP(t)zGlower(t)subscriptsuperscript𝑧P𝑗𝑡subscriptsuperscript𝑧lower𝐺𝑡z^{\mathrm{P}}_{j}(t)-z^{\mathrm{lower}}_{G}(t)italic_z start_POSTSUPERSCRIPT roman_P end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) - italic_z start_POSTSUPERSCRIPT roman_lower end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_t ) represent the distances of the P atom of lipid j𝑗jitalic_j, zjP(t)subscriptsuperscript𝑧P𝑗𝑡z^{\mathrm{P}}_{j}(t)italic_z start_POSTSUPERSCRIPT roman_P end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ), from zGupper(t)subscriptsuperscript𝑧upper𝐺𝑡z^{\mathrm{upper}}_{G}(t)italic_z start_POSTSUPERSCRIPT roman_upper end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_t ) and zGlower(t)subscriptsuperscript𝑧lower𝐺𝑡z^{\mathrm{lower}}_{G}(t)italic_z start_POSTSUPERSCRIPT roman_lower end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_t ), respectively, along the z𝑧zitalic_z-direction. (b) and (c) depict the distributions of P(|Δz|)𝑃Δ𝑧P(|\Delta z|)italic_P ( | roman_Δ italic_z | ) for DPPC and PSM, respectively. Snapshots in each panel are taken at 303 K (Left: with Chol, Right: pure).

The equilibration process is described in Table S1 of the supplementary material. In accordance with CHARMM-GUI guidelines, the process involved gradually relaxing restraints imposed on the phosphorus atom and the chiral carbon center of the lipid molecule (Nos. 1-6). The constants of the restraining forces on the z𝑧zitalic_z-coordinate of the phosphorus atom and on dihedral angle concerning the asymmetric center and double bond are denoted as kzsubscript𝑘𝑧k_{z}italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT and kdihsubscript𝑘dihk_{\mathrm{dih}}italic_k start_POSTSUBSCRIPT roman_dih end_POSTSUBSCRIPT, respectively, in Table S1. Subsequently, further equilibration steps were carried out (Nos. 7-11); the computational stability was checked in each of nos. 7-11 and the MD length was gradually increased from no. 7 to 11. Finally, three production runs under NPT𝑁𝑃𝑇NPTitalic_N italic_P italic_T conditions for 10 ns each were performed (No. 12). To examine how the effect of Chol depends on the phase of the lipid membrane (gel vs liquid-crystalline), MD simulations were conducted at 303 K and 323 K for each system. The coordinate system was set so that the z𝑧zitalic_z-axis is normal to the membrane surface, which spans over the x𝑥xitalic_x- and y𝑦yitalic_y-directions.

To confirm the adequacy of the equilibration process, we examined the time evolution of surface area S𝑆Sitalic_S in the x𝑥xitalic_x-y𝑦yitalic_y plane. Figures S1 and S2 of the supplementary material illustrate these results during the 3 μ𝜇\muitalic_μs equilibration at 303 K and 323 K, respectively. While noticeable fluctuations are observed around 1.5 μ𝜇\muitalic_μs in some systems, the area S𝑆Sitalic_S converges to a stable value at approximately 3 μ𝜇\muitalic_μs across all systems. Consequently, equilibration for 3 μ𝜇\muitalic_μs is considered adequate, and a production run was carried out after this equilibration period.

Refer to caption
Figure 4: (a) Schematic illustration of calculations of water molecule distribution ρ(z)𝜌superscript𝑧\rho(z^{\prime})italic_ρ ( italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ). (b) and (c) show the ratios of the water molecule distribution ρ(z)𝜌superscript𝑧\rho(z^{\prime})italic_ρ ( italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ), to the number density of bulk water, ρbulksubscript𝜌bulk\rho_{\mathrm{bulk}}italic_ρ start_POSTSUBSCRIPT roman_bulk end_POSTSUBSCRIPT, for DPPC and PSM, respectively. The values of ρbulksubscript𝜌bulk\rho_{\mathrm{bulk}}italic_ρ start_POSTSUBSCRIPT roman_bulk end_POSTSUBSCRIPT, determined from ρ(z)𝜌𝑧\rho(z)italic_ρ ( italic_z ) as illustrated in Fig. 2, are 33.50 nm-3 and 32.95 nm-3 for DPPC and PSM, respectively. (d) and (e) depict the average numbers of H-bonds using the same zsuperscript𝑧z^{\prime}italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT-axis as (b) and (c), for DPPC and PSM, respectively. The dashed lines at -0.25 nm and 0.8 nm define the boundaries between region 1 and 2 and between region 2 and 3, respectively.

III Results and discussion

III.1 Density distributions of lipid, water, and Chol

Initially, we analyzed the structure of the lipid bilayer, as well as the configurations of Chol and water. Figure 2 illustrates the number density distributions ρ(z)𝜌𝑧\rho(z)italic_ρ ( italic_z ) of the lipid carbon chain (tail), water molecular oxygen (OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT), and center of mass of Chol along the z𝑧zitalic_z-direction. Here, z𝑧zitalic_z denotes the distance from the bilayer center, and the center of the bilayer is the center of mass of the lipid molecules. Note that the number density distributions of the upper and lower leaflets were found to be coincident with each other within the margin of errors (data not shown). ρ(z)𝜌𝑧\rho(z)italic_ρ ( italic_z ) in Fig. 2 represents the average of the profiles of both leaflets. As depicted in Fig. 2, Chol is predominantly situated near the carbon chain of the lipid, leading to a broadening of the tail distribution along the z𝑧zitalic_z-direction, particularly evident at 303 K in DPPC. To examine this in detail, the number density distributions of OD7superscriptOD7\mathrm{O}^{\mathrm{D7}}roman_O start_POSTSUPERSCRIPT D7 end_POSTSUPERSCRIPT in DPPC and OS5superscriptOS5\mathrm{O}^{\mathrm{S5}}roman_O start_POSTSUPERSCRIPT S5 end_POSTSUPERSCRIPT in PSM are plotted in Fig. S3 of the supplementary material. These oxygen atoms are part of the hydrophilic functional groups within each lipid molecule, which are located in the innermost part of the lipid membranes (see Fig. 1). Figure S3 of the supplementary material demonstrates that the tail of water density distribution overlaps with those of OD7superscriptOD7\mathrm{O}^{\mathrm{D7}}roman_O start_POSTSUPERSCRIPT D7 end_POSTSUPERSCRIPT in DPPC and OS5superscriptOS5\mathrm{O}^{\mathrm{S5}}roman_O start_POSTSUPERSCRIPT S5 end_POSTSUPERSCRIPT in PSM. In the presence of Chol, particularly for DPPC at 303 K, the distribution of OD7superscriptOD7\mathrm{O}^{\mathrm{D7}}roman_O start_POSTSUPERSCRIPT D7 end_POSTSUPERSCRIPT becomes narrower, hindering the penetration of water molecules into the membrane’s inner regions, as observed in Fig. 2(a). Additionally, Fig. S4 of the supplementary material displays the number density distributions of nitrogen (N+) in the choline group and phosphorus (P) atoms along the z𝑧zitalic_z-direction. The peak intensities of distributions for N+ and P atoms were enhanced, particularly at 303 K in DPPC.

III.2 Fluctuation of the membrane interface

The surfaces of lipid membranes are soft and fluctuate with time. To elucidate the structure of the membrane interface, our focus was directed towards the lipid head, where we examined the distribution of lipid phosphorus atom position relative to the instantaneous interface defined below. The fluctuation of the interface between the membrane and water is seen evidently by employing the instantaneous interface method. [66]

We assign each lipid to either the upper or lower leaflet at each time. Here, Npuppersubscriptsuperscript𝑁upperpN^{\mathrm{upper}}_{\mathrm{p}}italic_N start_POSTSUPERSCRIPT roman_upper end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT denotes the number of lipid molecules in the upper leaflet, zjP(t)superscriptsubscript𝑧𝑗P𝑡z_{j}^{\mathrm{P}}\left(t\right)italic_z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_P end_POSTSUPERSCRIPT ( italic_t ) is the z𝑧zitalic_z-coordinate at time t𝑡titalic_t of the j𝑗jitalic_jth lipid molecule, and zGupper(t)=(1/Npupper)jupperzjP(t)subscriptsuperscript𝑧upper𝐺𝑡1subscriptsuperscript𝑁upperpsubscript𝑗uppersubscriptsuperscript𝑧P𝑗𝑡z^{\mathrm{upper}}_{G}(t)=(1/{N^{\mathrm{upper}}_{\mathrm{p}}})\sum_{j\in{% \mathrm{upper}}}z^{\mathrm{P}}_{j}(t)italic_z start_POSTSUPERSCRIPT roman_upper end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_t ) = ( 1 / italic_N start_POSTSUPERSCRIPT roman_upper end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT ) ∑ start_POSTSUBSCRIPT italic_j ∈ roman_upper end_POSTSUBSCRIPT italic_z start_POSTSUPERSCRIPT roman_P end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) represents the average of the z𝑧zitalic_z-coordinates of phosphorus atoms in the upper leaflet of the lipid bilayer at time t𝑡titalic_t. Similarly, Nplowersubscriptsuperscript𝑁lowerpN^{\mathrm{lower}}_{\mathrm{p}}italic_N start_POSTSUPERSCRIPT roman_lower end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT and zGlower(t)subscriptsuperscript𝑧lower𝐺𝑡z^{\mathrm{lower}}_{G}(t)italic_z start_POSTSUPERSCRIPT roman_lower end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_t ) can be computed for the lower leaflet.

As shown in Fig. 3(a), the deviation of the z𝑧zitalic_z-coordinate of the phosphorus atom of lipid j𝑗jitalic_j from zGupper(t)subscriptsuperscript𝑧upper𝐺𝑡z^{\mathrm{upper}}_{G}(t)italic_z start_POSTSUPERSCRIPT roman_upper end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_t ) is expressed as Δzjupper(t)=zjP(t)zGupper(t)Δsubscriptsuperscript𝑧upper𝑗𝑡subscriptsuperscript𝑧P𝑗𝑡subscriptsuperscript𝑧upper𝐺𝑡\Delta z^{\mathrm{upper}}_{j}(t)=z^{\mathrm{P}}_{j}(t)-z^{\mathrm{upper}}_{G}(t)roman_Δ italic_z start_POSTSUPERSCRIPT roman_upper end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) = italic_z start_POSTSUPERSCRIPT roman_P end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) - italic_z start_POSTSUPERSCRIPT roman_upper end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_G end_POSTSUBSCRIPT ( italic_t ) in the upper leaflet. Similarly, for lipids in the lower leaflet, Δzjlower(t)Δsubscriptsuperscript𝑧lower𝑗𝑡\Delta z^{\mathrm{lower}}_{j}(t)roman_Δ italic_z start_POSTSUPERSCRIPT roman_lower end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) is defined. Then, the time-averaged distribution function of the absolute values |Δz|Δ𝑧\left|\Delta z\right|| roman_Δ italic_z | of Δzjupper(t)Δsubscriptsuperscript𝑧upper𝑗𝑡\Delta z^{\mathrm{upper}}_{j}(t)roman_Δ italic_z start_POSTSUPERSCRIPT roman_upper end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) and Δzjlower(t)Δsubscriptsuperscript𝑧lower𝑗𝑡\Delta z^{\mathrm{lower}}_{j}(t)roman_Δ italic_z start_POSTSUPERSCRIPT roman_lower end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) can be assessed and is denoted as P(|Δz|)𝑃Δ𝑧P(|\Delta z|)italic_P ( | roman_Δ italic_z | ).

Figures 3(b) and 3(c) illustrate the results of P(|Δz|)𝑃Δ𝑧P(|\Delta z|)italic_P ( | roman_Δ italic_z | ) for DPPC and PSM, respectively. In both DPPC and PSM, Chol does not exert discernible effects on P(|Δz|)𝑃Δ𝑧P(|\Delta z|)italic_P ( | roman_Δ italic_z | ) at 323 K. Snapshots captured at 323 K are depicted in Fig. S5 of the supplementary material, revealing a disordered orientation of carbon chains within lipids. This observation signifies a high degree of membrane fluidity, with weak influence of Chol. In contrast, at 303 K, the distribution of P(|Δz|)𝑃Δ𝑧P(|\Delta z|)italic_P ( | roman_Δ italic_z | ) is broader in the pure lipid membrane systems, suggesting that Chol enhances membrane stability and maintains the interface position. The effect of Chol to suppress the interface fluctuations is particularly evident in DPPC, as illustrated in Fig. 3(b) (see also snapshots captured at 303 K in the insets of Figs. 3(b) and 3(c)).

III.3 Classification of water molecules

The analysis of water molecule distribution near the rugged membrane interface was conducted. A precise description of the local water distribution relative to the lipid interface was proposed [22, 23] The location of a water molecule is provided by zsuperscript𝑧z^{\prime}italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, which is defined as the distance from the interface using Voronoi tessellation. Unlike z𝑧zitalic_z, zsuperscript𝑧z^{\prime}italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT takes into account the effects of the fluctuation of the lipid/water interfaces, and the distribution ρ(z)𝜌superscript𝑧\rho(z^{\prime})italic_ρ ( italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) characterizes the layered structures of water molecules.

We propose a simpler method closely resembling Voronoi tessellation. The schematic illustration of the method is described in Fig. 4(a) and the detail is provided as follows: Step 1: Project the oxygen atoms of water molecules and the lipid phosphorus atoms onto the x𝑥xitalic_x-y𝑦yitalic_y plane. Calculate the distance dij=|𝒓iO(t)𝒓jP(t)|subscript𝑑𝑖𝑗subscriptsuperscript𝒓O𝑖𝑡subscriptsuperscript𝒓P𝑗𝑡d_{ij}=|\bm{r}^{\mathrm{O}}_{i}(t)-\bm{r}^{\mathrm{P}}_{j}(t)|italic_d start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = | bold_italic_r start_POSTSUPERSCRIPT roman_O end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) - bold_italic_r start_POSTSUPERSCRIPT roman_P end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) | between the oxygen atom of water molecule i𝑖iitalic_i and the phosphorus atom j𝑗jitalic_j in the x𝑥xitalic_x-y𝑦yitalic_y plane. Identify the lipid phosphorus atom isuperscript𝑖i^{\prime}italic_i start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT that gives the smallest dijsubscript𝑑𝑖𝑗d_{ij}italic_d start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT for water molecule i𝑖iitalic_i. Step 2: Determine ziO(t)subscriptsuperscript𝑧O𝑖𝑡z^{\mathrm{O}}_{i}(t)italic_z start_POSTSUPERSCRIPT roman_O end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) as the z𝑧zitalic_z-coordinate of oxygen atom of water molecule i𝑖iitalic_i and ziP(t)subscriptsuperscript𝑧Psuperscript𝑖𝑡z^{\mathrm{P}}_{i^{\prime}}(t)italic_z start_POSTSUPERSCRIPT roman_P end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_t ) as the z𝑧zitalic_z-coordinate of the phosphorus atom isuperscript𝑖i^{\prime}italic_i start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT nearest to i𝑖iitalic_i in the x𝑥xitalic_x-y𝑦yitalic_y plane. Define the water molecule density of z=ziOziPsuperscript𝑧subscriptsuperscript𝑧O𝑖subscriptsuperscript𝑧Psuperscript𝑖z^{\prime}=z^{\mathrm{O}}_{i}-z^{\mathrm{P}}_{i^{\prime}}italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_z start_POSTSUPERSCRIPT roman_O end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_z start_POSTSUPERSCRIPT roman_P end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT as ρ(z)𝜌superscript𝑧\rho(z^{\prime})italic_ρ ( italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ). Note that the negative zsuperscript𝑧z^{\prime}italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT indicates that the water molecule is located in more inner positions of the membrane than the phosphorus atom.

Refer to caption
Figure 5: (a) Schematic illustration of roosubscript𝑟oor_{\mathrm{oo}}italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT and β𝛽\betaitalic_β for H-bond formed between the oxygen atom (OD5superscriptOD5\mathrm{O}^{\mathrm{D5}}roman_O start_POSTSUPERSCRIPT D5 end_POSTSUPERSCRIPT) of a functional group within DPPC and a water molecule. (b)-(i) 2D PMF, W(roo,β)𝑊subscript𝑟oo𝛽W(r_{\mathrm{oo}},\beta)italic_W ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ), between water oxygen (OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT) and each acceptor oxygen in DPPC with Chol at 303 K. Black points represent saddle points.

Figures 4(b) and 4(c) illustrates the ratio of ρ(z)𝜌superscript𝑧\rho(z^{\prime})italic_ρ ( italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) to the number density of bulk water, ρbulksubscript𝜌bulk\rho_{\mathrm{bulk}}italic_ρ start_POSTSUBSCRIPT roman_bulk end_POSTSUBSCRIPT, for DPPC and PSM, respectively. In contrast to the water molecule density profile ρ(z)𝜌𝑧\rho(z)italic_ρ ( italic_z ) in Fig. 2, which is influenced by the instantaneous fluctuations of lipid membranes, ρ(z)𝜌superscript𝑧\rho(z^{\prime})italic_ρ ( italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) can be a more faithful representation of the water distribution near the rugged surface.

From the profile of ρ(z)𝜌superscript𝑧\rho(z^{\prime})italic_ρ ( italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ), the water molecules can be categorized into three regions (regions 1-3), as depicted in Figs. 4(b) and 4(c). Specifically, region 1 represents the region inside the membrane at z<superscript𝑧absentz^{\prime}<italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT < -0.25 nm, region 2 denotes the interface region at -0.25 nm <z<absentsuperscript𝑧absent<z^{\prime}<< italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT < 0.8 nm, and region 3 encompasses the bulk region at z>superscript𝑧absentz^{\prime}>italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT > 0.8 nm. These classifications align with previous studies. [22, 23] Note that the number of DPPC molecules was 128 in Ref. 22, which is slightly smaller than that of our system. Furthermore, similar results of ρ(z)𝜌superscript𝑧\rho(z^{\prime})italic_ρ ( italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) were reported by Elola et al, where 968 DPPC molecules with the united-atom model were simulated. [49]

The water content in region 1 decreases progressively towards the center of the membrane. In the interface region (region 2), a minimum was observed near z=0superscript𝑧0z^{\prime}=0italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 0 nm, with a peak occurring around z=0.4superscript𝑧0.4z^{\prime}=0.4italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = 0.4 nm, for both DPPC and PSM. This peak stems from the tendency for H-bond formation around phosphate groups. A further elucidation on the H-bond rearrangement will be provided in subsequent Sec. III.4. Remarkably, as shown in Fig. 4(b), at 303 K in DPPC, the water content in the interface region (region 2) is larger in the pure lipid membrane system than in the presence of Chol. This observation aligns with the variation in ρ(z)𝜌𝑧\rho(z)italic_ρ ( italic_z ) of water molecules due to Chol, as shown in Fig. 2(a). The pronounced stabilization of the DPPC membrane by Chol at 303 K highlights the significant impact on the hydration structure near the interface. On the contrary, at 323 K in DPPC, but the effect of Chol is less significant. Moreover, for PSM, the impacts of both temperature variations and Chol on ρ(z)𝜌superscript𝑧\rho(z^{\prime})italic_ρ ( italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) are not appreciable, as demonstrated in Fig. 4(c).

Refer to caption
Figure 6: (a) Schematic illustration of roosubscript𝑟oor_{\mathrm{oo}}italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT and β𝛽\betaitalic_β for H-bond formed between the oxygen atom (OS5superscriptOS5\mathrm{O}^{\mathrm{S5}}roman_O start_POSTSUPERSCRIPT S5 end_POSTSUPERSCRIPT) of a functional group within PSM and a water molecule. (b)-(i) 2D PMF, W(roo,β)𝑊subscript𝑟oo𝛽W(r_{\mathrm{oo}},\beta)italic_W ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ), between water oxygen (OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT) and each acceptor oxygen in PSM with Chol at 303 K. Black points represent saddle points.

III.4 H-bond arrangement

The H-bonding states at each of the donor and acceptor sites were analyzed. When investigating H-bond state in MD simulations, a commonly employed approach involves applying a geometric criterion to identify an H-bond between two water molecules. The predominant definition often adopts the distance between oxygen atoms (referred to as roosubscript𝑟oor_{\mathrm{oo}}italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT) and the angle formed by the oxygen atom and the oxygen-hydrogen bond (referred to as β𝛽\betaitalic_β) within a water dimer. [67, 68, 69]

A more comprehensive understanding of the H-bond state can be obtained by analyzing the distribution function of roosubscript𝑟oor_{\mathrm{oo}}italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT and β𝛽\betaitalic_β, denoted as g(roo,β)𝑔subscript𝑟oo𝛽g(r_{\mathrm{oo}},\beta)italic_g ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β )[70, 71, 72, 73] In this context, 2πρroo2sinβg(roo,β)droodβ2𝜋𝜌superscriptsubscript𝑟oo2𝛽𝑔subscript𝑟oo𝛽subscript𝑟oo𝛽2\pi\rho r_{\mathrm{oo}}^{2}\sin\beta g(r_{\mathrm{oo}},\beta)\differential{r_% {\mathrm{oo}}}\differential{\beta}2 italic_π italic_ρ italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_sin italic_β italic_g ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ) roman_d start_ARG italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT end_ARG roman_d start_ARG italic_β end_ARG represents the average number of oxygen atoms acting as H-bond acceptors within the partial spherical shell volume characterized by droodsubscript𝑟oo\mathrm{d}r_{\mathrm{oo}}roman_d italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT and dβd𝛽\mathrm{d}\betaroman_d italic_β at the position (roo,β)subscript𝑟oo𝛽(r_{\mathrm{oo}},\beta)( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ), with the average number density of water molecules, ρ𝜌\rhoitalic_ρ. The logarithm form W(roo,β)=kBTln(g(roo,β))𝑊subscript𝑟oo𝛽subscript𝑘B𝑇𝑔subscript𝑟oo𝛽W(r_{\mathrm{oo}},\beta)=-k_{\mathrm{B}}T\ln{g(r_{\mathrm{oo}},\beta)}italic_W ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ) = - italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T roman_ln ( start_ARG italic_g ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ) end_ARG ) can be interpreted as the two-dimensional potential of mean force (2D PMF). For reference, the 2D PMF W(roo,β)𝑊subscript𝑟oo𝛽W(r_{\mathrm{oo}},\beta)italic_W ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ) of bulk water at 303 K and 323 K with a density of 1 g/cm3 is depicted in Fig. S6 of the supplementary material. The temperature-independent energetically stable state is characterized by roo<0.35subscript𝑟oo0.35r_{\mathrm{oo}}<0.35italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT < 0.35 nm and 0<β<30superscript0𝛽superscript300^{\circ}<\beta<30^{\circ}0 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT < italic_β < 30 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT, which can be considered indicative of H-bond state.

Figures 5(a) and 6(a) provide schematic illustrations of roosubscript𝑟oor_{\mathrm{oo}}italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT and β𝛽\betaitalic_β for H-bond formed between a water molecule and the oxygen atom of a functional group within DPPC and PSM, respectively. Given the presence of H-bonds between water molecules and those with acceptors within lipid molecules, the analysis of 2D PMF was conducted for water molecules and potential acceptors, including the oxygen atoms in DPPC and the oxygen and nitrogen atoms in PSM. Additionally, in the presence of Chol, the analysis was also performed for H-bonds between water molecule and oxygen atom of hydroxy group, denoted as OCholsuperscriptOChol\mathrm{O}^{\mathrm{Chol}}roman_O start_POSTSUPERSCRIPT roman_Chol end_POSTSUPERSCRIPT. A similar 2D PMF analysis was previously done in polymer-water mixtures. [74]

Figure 5 illustrates the 2D PMF, W(roo,β)𝑊subscript𝑟oo𝛽W(r_{\mathrm{oo}},\beta)italic_W ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ), representing the interaction between water molecule as donors and DPPC oxygen atoms as acceptors at 303 K in the presence of Chol. Other results at 303 K in the absence of Chol and at 323 K both in the presence and absence of Chol are displayed in Figs. S7-S9 of the supplementary material. Similarly, in Fig. 6 and Figs. S10-S12 of the supplementary material provide the 2D PMF, W(roo,β)𝑊subscript𝑟oo𝛽W(r_{\mathrm{oo}},\beta)italic_W ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ), for PSM systems. Note that the 2D PMF between water molecules is omitted since the overall profile remains unchanged for both DPPC and PSM, when compared to that of bulk water (see Fig. S6 of the supplementary material). Based on the 2D PMF analysis, we identify potential H-bond acceptors as O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT, O2superscriptO2\mathrm{O}^{\mathrm{2}}roman_O start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, O3superscriptO3\mathrm{O}^{\mathrm{3}}roman_O start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, OD5superscriptOD5\mathrm{O}^{\mathrm{D5}}roman_O start_POSTSUPERSCRIPT D5 end_POSTSUPERSCRIPT, OD7superscriptOD7\mathrm{O}^{\mathrm{D7}}roman_O start_POSTSUPERSCRIPT D7 end_POSTSUPERSCRIPT and OCholsuperscriptOChol\mathrm{O}^{\mathrm{Chol}}roman_O start_POSTSUPERSCRIPT roman_Chol end_POSTSUPERSCRIPT for DPPC, and O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT, O3superscriptO3\mathrm{O}^{\mathrm{3}}roman_O start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, OS4superscriptOS4\mathrm{O}^{\mathrm{S4}}roman_O start_POSTSUPERSCRIPT S4 end_POSTSUPERSCRIPT, OS5superscriptOS5\mathrm{O}^{\mathrm{S5}}roman_O start_POSTSUPERSCRIPT S5 end_POSTSUPERSCRIPT, and OCholsuperscriptOChol\mathrm{O}^{\mathrm{Chol}}roman_O start_POSTSUPERSCRIPT roman_Chol end_POSTSUPERSCRIPT for PSM, respectively. See Fig. 1 for the notations of the acceptor sites in the lipids. As illustrated in Figs. 5 and 6 and Figs. S7-S12 of the supplementary material, the H-bond region is characterized by roo<3.5subscript𝑟oo3.5r_{\mathrm{oo}}<3.5italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT < 3.5 nm and 0<β<30superscript0𝛽superscript300^{\circ}<\beta<30^{\circ}0 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT < italic_β < 30 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT in the 2D PMF, irrespective of the acceptor. Furthermore, the H-bond regions remain unchanged regardless of the presence of Chol or variations in temperature for both DPPC and PSM. Nevertheless, specific oxygen atoms such as O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT, OD5superscriptOD5\mathrm{O}^{\mathrm{D5}}roman_O start_POSTSUPERSCRIPT D5 end_POSTSUPERSCRIPT, and OD7superscriptOD7\mathrm{O}^{\mathrm{D7}}roman_O start_POSTSUPERSCRIPT D7 end_POSTSUPERSCRIPT in DPPC, and O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT and OS5superscriptOS5\mathrm{O}^{\mathrm{S5}}roman_O start_POSTSUPERSCRIPT S5 end_POSTSUPERSCRIPT in PSM, which possess higher negative charges, form more energetically stable states compared to other acceptors. In contrast, oxygen atoms OD4superscriptOD4\mathrm{O}^{\mathrm{D4}}roman_O start_POSTSUPERSCRIPT D4 end_POSTSUPERSCRIPT and OD6superscriptOD6\mathrm{O}^{\mathrm{D6}}roman_O start_POSTSUPERSCRIPT D6 end_POSTSUPERSCRIPT in DPPC, and oxygen atom O2superscriptO2\mathrm{O}^{\mathrm{2}}roman_O start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and nitrogen (N) in PSM, form a second coordination region outside the defined H-bond region. Consequently, these are excluded from further H-bond analysis due to their indeterminate bond characteristics.

Figures 4(d) and 4(e) illustrate the distributions of the average number of H-bonds formed by water molecules at each position, using the zsuperscript𝑧z^{\prime}italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT-axis corresponding to ρ(z)𝜌superscript𝑧\rho(z^{\prime})italic_ρ ( italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ), for DPPC and PSM, respectively. The average number of H-bonds in region 3 converged to 3.33 at 303 K and 3.23 at 323 K, respectively, corresponding to those observed in bulk water at each temperature. In region 2, the average number of H-bonds reaches a maximum value higher than the average observed in the bulk, gradually decreasing towards the interior of the bilayer. This peak position corresponds to that in ρ(z)𝜌superscript𝑧\rho(z^{\prime})italic_ρ ( italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ), as observed in Figs. 4(b) and 4(c). These observations suggest that around phosphate groups, the oxygen atoms within the lipid head group, such as O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT, O2superscriptO2\mathrm{O}^{\mathrm{2}}roman_O start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, and O3superscriptO3\mathrm{O}^{\mathrm{3}}roman_O start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT in DPPC, and O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT and O3superscriptO3\mathrm{O}^{\mathrm{3}}roman_O start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT in PSM, act as acceptors, promoting the formation of H-bonds with water molecules. Remarkably, at 303 K, the average number of H-bonds in region 1 increases with removal of Chol for both DPPC and PSM, with this trend being particularly notable in the DPPC system. However, at 323 K, the average number of H-bonds remains unchanged regardless of the presence or absence of Chol, for both DPPC and PSM. These findings indicate variations in membrane structure induced by Chol impact the propensity for H-bond formation within the membrane.

Refer to caption
Figure 7: Conditional probability C2,j(t)subscript𝐶2𝑗𝑡C_{2,j}(t)italic_C start_POSTSUBSCRIPT 2 , italic_j end_POSTSUBSCRIPT ( italic_t ), representing transition dynamics from region 2222 at the initial time t=0𝑡0t=0italic_t = 0 to either regions 1 or 3 at time t𝑡titalic_t or remaining within the same region 2 in the time interval t𝑡titalic_t [(a) DPPC at 303 K, (b) DPPC at 323 K, (c) PSM at 303 K, and (d) PSM at 323 K].

III.5 Water molecule rearrangement dynamics

We explore the transition dynamics of water molecules among the three regions. We address the dynamics of transition by defining Ci,j(t)subscript𝐶𝑖𝑗𝑡C_{i,j}(t)italic_C start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ( italic_t ) (ij𝑖𝑗i\neq jitalic_i ≠ italic_j) as the conditional probability that when a water molecule is in region i𝑖iitalic_i at time 0, it visits region j𝑗jitalic_j with the number of passing the boundaries of the regions being unity by time t𝑡titalic_t. Further, Ci,i(t)subscript𝐶𝑖𝑖𝑡C_{i,i}(t)italic_C start_POSTSUBSCRIPT italic_i , italic_i end_POSTSUBSCRIPT ( italic_t ) is the probability that the water molecule stays in region i𝑖iitalic_i without visiting the other regions during the time interval between 0 and t𝑡titalic_t. Note that the summation over all possible j𝑗jitalic_j states ensures jCi,j(t)=1subscript𝑗subscript𝐶𝑖𝑗𝑡1\sum_{j}C_{i,j}(t)=1∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ( italic_t ) = 1, conserving the number of water molecules. In practice, the trajectories of water molecules are continuously monitored from t=0𝑡0t=0italic_t = 0, tracking the subsequent transitions from region i𝑖iitalic_i to region j𝑗jitalic_j at each time t𝑡titalic_t.

Figure 7 presents the results of C2,j(t)subscript𝐶2𝑗𝑡C_{2,j}(t)italic_C start_POSTSUBSCRIPT 2 , italic_j end_POSTSUBSCRIPT ( italic_t ), illustrating the transition dynamics of water molecules originating from region 2 at t=0𝑡0t=0italic_t = 0. Note that the sum C2,1(t)+C2,2(t)+C2,3(t)=1subscript𝐶21𝑡subscript𝐶22𝑡subscript𝐶23𝑡1C_{2,1}(t)+C_{2,2}(t)+C_{2,3}(t)=1italic_C start_POSTSUBSCRIPT 2 , 1 end_POSTSUBSCRIPT ( italic_t ) + italic_C start_POSTSUBSCRIPT 2 , 2 end_POSTSUBSCRIPT ( italic_t ) + italic_C start_POSTSUBSCRIPT 2 , 3 end_POSTSUBSCRIPT ( italic_t ) = 1 holds at all times t𝑡titalic_t, as explained in the definition of Ci,j(t)subscript𝐶𝑖𝑗𝑡C_{i,j}(t)italic_C start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ( italic_t ). Except for DPPC at 303 K, C2,j(t)subscript𝐶2𝑗𝑡C_{2,j}(t)italic_C start_POSTSUBSCRIPT 2 , italic_j end_POSTSUBSCRIPT ( italic_t ) is not affected by the presence or absence of Chol and the transition rates from region 2 to regions 1 and 3 are common between the systems with and without Chol. In contrast, for DPPC at 303 K, the decay of C2,2(t)subscript𝐶22𝑡C_{2,2}(t)italic_C start_POSTSUBSCRIPT 2 , 2 end_POSTSUBSCRIPT ( italic_t ) exhibits a slower rate in the presence of Chol. Furthermore, Chol alters the fraction of water molecules transitioning to their respective destination. Specifically, the population of water molecules transitioning from region 2 to region 1 decreases by approximately 5% in the presence of Chol, while the transition to region 3 increases by a similar proportion. At 323 K, the saturated values of C2,1(t)subscript𝐶21𝑡C_{2,1}(t)italic_C start_POSTSUBSCRIPT 2 , 1 end_POSTSUBSCRIPT ( italic_t ) and C2,3(t)subscript𝐶23𝑡C_{2,3}(t)italic_C start_POSTSUBSCRIPT 2 , 3 end_POSTSUBSCRIPT ( italic_t ) resemble those of DPPC without Chol at 303 K. These observations indicate the significant impact of membrane structure variations on water molecule dynamics. For DPPC at 303 K, in particular, Chol enhance the tendency of keeping water molecules in the interface region (region 2) and relocating them towards the bulk region (region 3). This suggests the reduction of water molecule exchanges between the interface region (region 2) and the inner side of the membrane (region 1).

Figure S13 of the supplementary material illustrates the results of C1,j(t)subscript𝐶1𝑗𝑡C_{1,j}(t)italic_C start_POSTSUBSCRIPT 1 , italic_j end_POSTSUBSCRIPT ( italic_t ), which represent the transition dynamics from region 1111 at the initial time t=0𝑡0t=0italic_t = 0 to regions 2 or remaining within the same region 1111 at subsequent time t𝑡titalic_t. Here, C1,3(t)subscript𝐶13𝑡C_{1,3}(t)italic_C start_POSTSUBSCRIPT 1 , 3 end_POSTSUBSCRIPT ( italic_t ) is excluded given that the transition from region 1 to region 3 inevitably passes through region 2, ensuring the relationship, C1,1(t)+C1,2(t)=1subscript𝐶11𝑡subscript𝐶12𝑡1C_{1,1}(t)+C_{1,2}(t)=1italic_C start_POSTSUBSCRIPT 1 , 1 end_POSTSUBSCRIPT ( italic_t ) + italic_C start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT ( italic_t ) = 1. Interestingly, the transition from region 1 to region 2 exhibits slower dynamics in PSM compared to in DPPC at 303 K and 323 K. This observation may be linked to the shorter H-bond lifetime τHBsubscript𝜏HB\tau_{\mathrm{HB}}italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT of water molecules in the PSM system than with DPPC, as elucidated in the subsequent Sec. III.6. In addition, Chol further retards these dynamics, particularly evident at 303 K for both DPPC and PSM. This observation suggests that Chol, situated within the membrane interior, exerts a notable influence on the dynamics of water molecules within region 1.

Refer to caption
Figure 8: Dependency of the ratio of H-bond lifetimes, τHBsubscript𝜏HB\tau_{\mathrm{HB}}italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT, with and without Chol, denoted as τHB(Chol)/τHB(pure)subscript𝜏HBCholsubscript𝜏HBpure\tau_{\mathrm{HB(Chol)}}/\tau_{\mathrm{HB(pure)}}italic_τ start_POSTSUBSCRIPT roman_HB ( roman_Chol ) end_POSTSUBSCRIPT / italic_τ start_POSTSUBSCRIPT roman_HB ( roman_pure ) end_POSTSUBSCRIPT for (a) DPPC and (b) PSM.

III.6 Chol influence on H-bond lifetime

Finally, we conducted an analysis of the H-bonding dynamics involving lipid molecules, Chol, and water molecules to elucidate the timescale of H-bond lifetime, by focusing on the acceptor oxygen atoms, such as O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT, O2superscriptO2\mathrm{O}^{\mathrm{2}}roman_O start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, O3superscriptO3\mathrm{O}^{\mathrm{3}}roman_O start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, OD5superscriptOD5\mathrm{O}^{\mathrm{D5}}roman_O start_POSTSUPERSCRIPT D5 end_POSTSUPERSCRIPT, and OD7superscriptOD7\mathrm{O}^{\mathrm{D7}}roman_O start_POSTSUPERSCRIPT D7 end_POSTSUPERSCRIPT in DPPC, and O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT, O3superscriptO3\mathrm{O}^{\mathrm{3}}roman_O start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, OS4superscriptOS4\mathrm{O}^{\mathrm{S4}}roman_O start_POSTSUPERSCRIPT S4 end_POSTSUPERSCRIPT, and OS5superscriptOS5\mathrm{O}^{\mathrm{S5}}roman_O start_POSTSUPERSCRIPT S5 end_POSTSUPERSCRIPT in PSM, and OCholsuperscriptOChol\mathrm{O}^{\mathrm{Chol}}roman_O start_POSTSUPERSCRIPT roman_Chol end_POSTSUPERSCRIPT within Chol. The H-bond time correlation function PHB(t)subscript𝑃HB𝑡P_{\mathrm{HB}}(t)italic_P start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT ( italic_t ) is defined as

PHB(t)=hi,j(t)hi,j(0)hi,j(0),subscript𝑃HB𝑡delimited-⟨⟩subscript𝑖𝑗𝑡subscript𝑖𝑗0delimited-⟨⟩subscript𝑖𝑗0\displaystyle P_{\mathrm{HB}}(t)=\dfrac{\left\langle h_{i,j}(t)h_{i,j}(0)% \right\rangle}{\left\langle h_{i,j}(0)\right\rangle},italic_P start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT ( italic_t ) = divide start_ARG ⟨ italic_h start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ( italic_t ) italic_h start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ( 0 ) ⟩ end_ARG start_ARG ⟨ italic_h start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ( 0 ) ⟩ end_ARG , (1)

where hi,j(t)subscript𝑖𝑗𝑡h_{i,j}(t)italic_h start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT ( italic_t ) equals 1 if water molecule i𝑖iitalic_i is H-bonded with acceptor oxygen j𝑗jitalic_j at time t𝑡titalic_t, otherwise 0. [75, 68, 67] We computed PHB(t)subscript𝑃HB𝑡P_{\mathrm{HB}}(t)italic_P start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT ( italic_t ) using the Monte-Carlo bootstrap method, which employs a non-parametric approach to statistical inference. [76]

Figures S14 and S15 of the supplementary material show the PHB(t)subscript𝑃HB𝑡P_{\mathrm{HB}}(t)italic_P start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT ( italic_t ) results for acceptor oxygen atoms in the DPPC and PSM systems, respectively. Notably, as the temperature decreases, the decay of PHB(t)subscript𝑃HB𝑡P_{\mathrm{HB}}(t)italic_P start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT ( italic_t ) slows down for each acceptor oxygen atom, with a significant effect observed for OCholsuperscriptOChol\mathrm{O}^{\mathrm{Chol}}roman_O start_POSTSUPERSCRIPT roman_Chol end_POSTSUPERSCRIPT in Chol. However, the impact of temperature variation on H-bond breakages between water molecules is negligible for both DPPC and PSM, owing to the abundance of H-bonding partners of OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT in the bulk. Furthermore, the influence of Chol is more pronounced in DPPC compared to PSM, particularly at 303 K. As illustrated in Figs. S15 and S16, the H-bond correlation between OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT and OCholsuperscriptOChol\mathrm{O}^{\mathrm{Chol}}roman_O start_POSTSUPERSCRIPT roman_Chol end_POSTSUPERSCRIPT exhibits a slower dynamics in PSM compared to DPPC.

The H-bond time correlation function, PHB(t)subscript𝑃HB𝑡P_{\mathrm{HB}}(t)italic_P start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT ( italic_t ), is approximated using the Kohlrausch–Williams–Watts (KWW) function, PHB(t)exp[(t/τKWW)βKWW]similar-to-or-equalssubscript𝑃HB𝑡superscript𝑡subscript𝜏KWWsubscript𝛽KWWP_{\mathrm{HB}}(t)\simeq\exp[-(t/\tau_{\mathrm{KWW}})^{\beta_{\mathrm{KWW}}}]italic_P start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT ( italic_t ) ≃ roman_exp [ - ( italic_t / italic_τ start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_β start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ]. The fitting results are depicted in Figs. S14 and S15 of the supplementary material for DPPC and PSM, respectively. The H-bond lifetime τHBsubscript𝜏HB\tau_{\mathrm{HB}}italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT is evaluated by integrating PHB(t)subscript𝑃HB𝑡P_{\mathrm{HB}}(t)italic_P start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT ( italic_t ), yielding

τHB=0PHB(t)dt=τKWWβKWWΓ(1βKWW),subscript𝜏HBsuperscriptsubscript0subscript𝑃HB𝑡𝑡subscript𝜏KWWsubscript𝛽KWWΓ1subscript𝛽KWW\displaystyle\tau_{\mathrm{HB}}=\int_{0}^{\infty}P_{\mathrm{HB}}(t)% \differential{t}=\frac{\tau_{\mathrm{KWW}}}{\beta_{\mathrm{KWW}}}\Gamma\left(% \frac{1}{\beta_{\mathrm{KWW}}}\right),italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT ( italic_t ) roman_d start_ARG italic_t end_ARG = divide start_ARG italic_τ start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT end_ARG start_ARG italic_β start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT end_ARG roman_Γ ( divide start_ARG 1 end_ARG start_ARG italic_β start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT end_ARG ) , (2)

where Γ()Γ\Gamma(\cdots)roman_Γ ( ⋯ ) denotes the Gamma function. The raw data of τKWWsubscript𝜏KWW\tau_{\mathrm{KWW}}italic_τ start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT, βKWWsubscript𝛽KWW\beta_{\mathrm{KWW}}italic_β start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT, and τHBsubscript𝜏HB\tau_{\mathrm{HB}}italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT are summarized in Tables S2-S7 of the supplementary material for both DPPC and PSM. To highlight the influence of Chol on H-bond lifetime τHBsubscript𝜏HB\tau_{\mathrm{HB}}italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT, the dependency of the ratio between τHBsubscript𝜏HB\tau_{\mathrm{HB}}italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT with and without Chol, denoted as τHB(Chol)/τHB(pure)subscript𝜏HBCholsubscript𝜏HBpure\tau_{\mathrm{HB(Chol)}}/\tau_{\mathrm{HB(pure)}}italic_τ start_POSTSUBSCRIPT roman_HB ( roman_Chol ) end_POSTSUBSCRIPT / italic_τ start_POSTSUBSCRIPT roman_HB ( roman_pure ) end_POSTSUBSCRIPT, on acceptor oxygen at 303 K and 323 K is shown in Fig. 8.

Figure 8(a) illustrates the ratio τHB(Chol)/τHB(pure)subscript𝜏HBCholsubscript𝜏HBpure\tau_{\mathrm{HB(Chol)}}/\tau_{\mathrm{HB(pure)}}italic_τ start_POSTSUBSCRIPT roman_HB ( roman_Chol ) end_POSTSUBSCRIPT / italic_τ start_POSTSUBSCRIPT roman_HB ( roman_pure ) end_POSTSUBSCRIPT, ranging from 1.5 to 2.5, excluding OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT, at 303 K for DPPC. Moreover, τHBsubscript𝜏HB\tau_{\mathrm{HB}}italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT becomes large with the internal oxygen atoms within the membrane, such as OD5superscriptOD5\mathrm{O}^{\mathrm{D5}}roman_O start_POSTSUPERSCRIPT D5 end_POSTSUPERSCRIPT and OD7superscriptOD7\mathrm{O}^{\mathrm{D7}}roman_O start_POSTSUPERSCRIPT D7 end_POSTSUPERSCRIPT. In contrast, O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT and O3superscriptO3\mathrm{O}^{\mathrm{3}}roman_O start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT exhibits a relatively faster H-bond lifetime, showing that the H-bond dynamics is less susceptible to the presence or absence of Chol near the aqueous region. However, in the case of PSM, the influence of cholesterol on H-bond lifetime is limited, except for OS5superscriptOS5\mathrm{O}^{\mathrm{S5}}roman_O start_POSTSUPERSCRIPT S5 end_POSTSUPERSCRIPT, at both 303 K and 323 K, as observed in Fig. 8(b). The slower dynamics observed for OS5superscriptOS5\mathrm{O}^{\mathrm{S5}}roman_O start_POSTSUPERSCRIPT S5 end_POSTSUPERSCRIPT in PSM can be attributed to its position as the innermost oxygen atom within the membrane, rendering it more susceptible to Chol than other oxygen atoms.

Finally, we examine the energetic aspect of the H-bond lifetime τHBsubscript𝜏HB\tau_{\mathrm{HB}}italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT. Specifically, the activation energy ΔGΔ𝐺\Delta Groman_Δ italic_G of the H-bond is estimated by the free energy difference between the most stable and saddle points on the 2D PMF, W(rOO,β)𝑊subscript𝑟OO𝛽W(r_{\mathrm{OO}},\beta)italic_W ( italic_r start_POSTSUBSCRIPT roman_OO end_POSTSUBSCRIPT , italic_β ) (see Figs. 5 and  6, and Figs. S7-S12 of the supplementary material). We plotted the relationship between τHBsubscript𝜏HB\tau_{\mathrm{HB}}italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT and ΔG/kBTΔ𝐺subscript𝑘B𝑇\Delta G/k_{\mathrm{B}}Troman_Δ italic_G / italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T in Fig. S16 of the supplementary material. Assuming the Arrhenius equation, τHB=Aexp(ΔG/kBT)subscript𝜏HB𝐴Δ𝐺subscript𝑘B𝑇\tau_{\text{HB}}=A\exp({\Delta G}/{k_{\text{B}}T})italic_τ start_POSTSUBSCRIPT HB end_POSTSUBSCRIPT = italic_A roman_exp ( start_ARG roman_Δ italic_G / italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T end_ARG ), we determined the constant A𝐴Aitalic_A by fitting. The straight lines with A=0.183𝐴0.183A=0.183italic_A = 0.183 ps and A=0.385𝐴0.385A=0.385italic_A = 0.385 ps are shown for DPPC and PSM systems, respectively, in Figs. S16(a) and S16(b) of the supplementary material. The positive correlation between τHBsubscript𝜏HB\tau_{\text{HB}}italic_τ start_POSTSUBSCRIPT HB end_POSTSUBSCRIPT and ΔG/kBTΔ𝐺subscript𝑘B𝑇\Delta G/k_{\mathrm{B}}Troman_Δ italic_G / italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T is demonstrated for both DPPC and PSM systems, although the deviation from the Arrhenius equation is noticeable. This result suggests that a more accurate energetic description of H-bond breakage involving a large number of molecules necessitates additional detailed variables beyond the two variables of distance rOOsubscript𝑟OOr_{\mathrm{OO}}italic_r start_POSTSUBSCRIPT roman_OO end_POSTSUBSCRIPT and angle β𝛽\betaitalic_β.

IV Conclusions

In this study, we employed MD simulations to investigate the influence of Chol on water molecule behavior within lipid membranes, with a specific focus on systems comprising DPPC and PSM. While lipid membrane structures are not susceptible to the presence of Chol at 323 K, Chol at 303 K serves to stabilize carbon chains, thereby reducing structural fluctuations at the membrane interface, particularly for DPPC.

The spatial distribution of water molecules surrounding the membrane can be classified into three distinct regions: the membrane interior, the interface, and the bulk. The transition of water from the interface to the bulk is facilitated by Chol for the DPPC system at 303 K. In contrast, the presence of Chol induces entrapment of water molecules within the membrane, leading to reduced rates of transition to the interface region from the interior region.

Our exploration into the dynamic attributes of water molecules in lipid-membrane systems, considering the influence of Chol and temperature variations has yielded insights into the intricate interplay at the membrane interface. DPPC is more susceptible to modifications at 303 K, significantly influencing H-bond dynamics within the membrane. Specifically, at 303 K in DPPC, Chol was found to markedly increase the H-bonding lifetime, particularly impacting internal oxygen atoms.

It is important to note the discrepancy in Chol density between our study, which utilized up to 10%, and 50% employed by Elola et al[49] This emphasizes the need to further investigate H-bonding dynamics in lipid-membrane systems under conditions compatible to real Chol contents in future research. While Elola et al. reported accelerated water dynamics near the interface region, our findings have not directly corroborated these observations. However, we observed a notable migration tendency of water molecules from region 2 (interface region) to region 3 (bulk region) in DPPC in the presence of Chol, as illustrated in Fig. 7(a). Therefore, future studies should focus on systematically varying both temperature and Chol content to provide a comprehensive understanding of H-bonding dynamics.

Supplementary material

The supplementary material include equilibration scheme of MD simulations (Table S1), time evolution of surface area S𝑆Sitalic_S in the x𝑥xitalic_x-y𝑦yitalic_y plane during the equilibration (Fig. S1 and S2), number density distribution of OD7superscriptOD7\mathrm{O}^{\mathrm{D7}}roman_O start_POSTSUPERSCRIPT D7 end_POSTSUPERSCRIPT in DPPC and OS5superscriptOS5\mathrm{O}^{\mathrm{S5}}roman_O start_POSTSUPERSCRIPT S5 end_POSTSUPERSCRIPT in PSM along the z𝑧zitalic_z-direction (Fig. S3), number density distribution of nitrogen and phosphorus atoms along the z𝑧zitalic_z-direction (Fig. S4), MD snapshots taken at 323 K (Fig. S5), 2D PMF between water molecules in bulk water (Fig. S6), 2D PMF between water oxygen and acceptor oxygen atoms in lipid molecules (Fig. S7-S12), conditional probability C1,j(t)subscript𝐶1𝑗𝑡C_{1,j}(t)italic_C start_POSTSUBSCRIPT 1 , italic_j end_POSTSUBSCRIPT ( italic_t ) (Fig. S13), H-bond time correlation function PHB(t)subscript𝑃HB𝑡P_{\mathrm{HB}}(t)italic_P start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT ( italic_t ) (Figs. S14 and S15), raw data of τKWWsubscript𝜏KWW\tau_{\mathrm{KWW}}italic_τ start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT, βKWWsubscript𝛽KWW\beta_{\mathrm{KWW}}italic_β start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT, and τHBsubscript𝜏HB\tau_{\mathrm{HB}}italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT (Table S2-S7), and the relationship between the H-bond lifetime τHBsubscript𝜏HB\tau_{\mathrm{HB}}italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT and activation energy normalized by thermal energy, ΔG/kBTΔ𝐺subscript𝑘B𝑇\Delta G/k_{\mathrm{B}}Troman_Δ italic_G / italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T (Fig. S16).

Acknowledgements.
This work was supported by JSPS KAKENHI Grant-in-Aid Grant Nos. JP21H04628, JP21H05249, JP22H04542, JP22K03550, JP23H01924, JP23K26617, JP23H02622, JP23K27313, and JP24H01719. We are grateful to the Fugaku Supercomputing Project (Nos. JPMXP1020230325 and JPMXP1020230327) and the Data-Driven Material Research Project (No. JPMXP1122714694) from the Ministry of Education, Culture, Sports, Science, and Technology and to Maruho Collaborative Project for Theoretical Pharmaceutics. The numerical calculations were performed at Research Center for Computational Science, Okazaki Research Facilities, National Institutes of Natural Sciences (Project: 24-IMS-C051) and at the Cybermedia Center, Osaka University.

AUTHOR DECLARATIONS

Conflict of Interest

The authors have no conflicts to disclose.

Data availability statement

The program codes and data that support the findings of this study are openly available in Zenodo at https://doi.org/10.5281/zenodo.11273701. Further data are available from the corresponding author upon reasonable request.

References

  • Nagle and Tristram-Nagle [2000] J. F. Nagle and S. Tristram-Nagle, “Structure of lipid bilayers,” Biochim. Biophys. Acta Biomembr. 1469, 159–195 (2000).
  • Israelachvili [2011] J. N. Israelachvili, Intermolecular and Surface Forces, 3rd ed. (Academic Press, Burlington, MA, 2011).
  • Mouritsen and Zuckermann [2004] O. G. Mouritsen and M. J. Zuckermann, “What’s so special about cholesterol?” Lipids 39, 1101–1113 (2004).
  • de Meyer and Smit [2009] F. de Meyer and B. Smit, “Effect of cholesterol on the structure of a phospholipid bilayer,” Proc. Natl. Acad. Sci. U.S.A. 106, 3654–3658 (2009).
  • Marrink et al. [1996] S.-J. Marrink, D. P. Tieleman, A. R. van Buuren, and H. J. C. Berendsen, “Membranes and water: An interesting relationship,” Faraday Discuss. 103, 191–201 (1996).
  • Pratt and Pohorille [2002] L. R. Pratt and A. Pohorille, “Hydrophobic Effects and Modeling of Biophysical Aqueous Solution Interfaces,” Chem. Rev. 102, 2671–2692 (2002).
  • Higgins et al. [2006] M. J. Higgins, M. Polcik, T. Fukuma, J. E. Sader, Y. Nakayama, and S. P. Jarvis, “Structured Water Layers Adjacent to Biological Membranes,” Biophys. J. 91, 2532–2542 (2006).
  • Raschke [2006] T. M. Raschke, “Water structure and interactions with protein surfaces,” Curr. Opin. Struct. Biol. Theory and Simulation/Macromolecular Assemblages, 16, 152–159 (2006).
  • Ziegler and Vernier [2008] M. J. Ziegler and P. T. Vernier, “Interface Water Dynamics and Porating Electric Fields for Phospholipid Bilayers,” J. Phys. Chem. B 112, 13588–13596 (2008).
  • Disalvo et al. [2008] E. A. Disalvo, F. Lairion, F. Martini, E. Tymczyszyn, M. Frías, H. Almaleck, and G. J. Gordillo, “Structural and functional properties of hydration and confined water in membrane interfaces,” Biochim. Biophys. Acta Biomembr. 1778, 2655–2670 (2008).
  • Cheng et al. [2013] C.-Y. Cheng, J. Varkey, M. R. Ambroso, R. Langen, and S. Han, “Hydration dynamics as an intrinsic ruler for refining protein structure at lipid membrane interfaces,” Proc. Natl. Acad. Sci. U.S.A. 110, 16838–16843 (2013).
  • Zhou and Cross [2013] H.-X. Zhou and T. A. Cross, “Influences of Membrane Mimetic Environments on Membrane Protein Structures,” Annu. Rev. Biophys. 42, 361–392 (2013).
  • Disalvo [2015] E. A. Disalvo, ed., Membrane Hydration: The Role of Water in the Structure and Function of Biological Membranes, Subcellular Biochemistry, Vol. 71 (Springer, Cham, Cham, 2015).
  • Jungwirth [2015] P. Jungwirth, “Biological Water or Rather Water in Biology?” J. Phys. Chem. Lett. 6, 2449–2451 (2015).
  • Laage, Elsaesser, and Hynes [2017] D. Laage, T. Elsaesser, and J. T. Hynes, “Water Dynamics in the Hydration Shells of Biomolecules,” Chem. Rev. 117, 10694–10725 (2017).
  • Chattopadhyay et al. [2021] M. Chattopadhyay, E. Krok, H. Orlikowska, P. Schwille, H. G. Franquelim, and L. Piatkowski, “Hydration Layer of Only a Few Molecules Controls Lipid Mobility in Biomimetic Membranes,” J. Am. Chem. Soc. 143, 14551–14562 (2021).
  • Berkowitz and Raghavan [1991] M. L. Berkowitz and K. Raghavan, “Computer simulation of a water/membrane interface,” Langmuir 7, 1042–1044 (1991).
  • Pastor [1994] R. W. Pastor, “Molecular dynamics and Monte Carlo simulations of lipid bilayers,” Curr. Opin. Struct. Biol. 4, 486–492 (1994).
  • Marrink and Berendsen [1994] S.-J. Marrink and H. J. C. Berendsen, “Simulation of water transport through a lipid membrane,” J. Phys. Chem. 98, 4155–4168 (1994).
  • Zhou and Schulten [1995] F. Zhou and K. Schulten, “Molecular Dynamics Study of a Membrane-Water Interface,” J. Phys. Chem. 99, 2194–2207 (1995).
  • Jakobsson [1997] E. Jakobsson, “Computer simulation studies of biological membranes: Progress, promise and pitfalls,” Trends Biochem. Sci. 22, 339–344 (1997).
  • Pandit, Bostick, and Berkowitz [2003] S. A. Pandit, D. Bostick, and M. L. Berkowitz, “An algorithm to describe molecular scale rugged surfaces and its application to the study of a water/lipid bilayer interface,” J. Chem. Phys. 119, 2199–2205 (2003).
  • Berkowitz, Bostick, and Pandit [2006] M. L. Berkowitz, D. L. Bostick, and S. Pandit, “Aqueous Solutions next to Phospholipid Membrane Surfaces: Insights from Simulations,” Chem. Rev. 106, 1527–1539 (2006).
  • Matubayasi, Shinoda, and Nakahara [2008] N. Matubayasi, W. Shinoda, and M. Nakahara, “Free-energy analysis of the molecular binding into lipid membrane with the method of energy representation,” J. Chem. Phys. 128, 195107 (2008).
  • Marrink et al. [2019] S. J. Marrink, V. Corradi, P. C. Souza, H. I. Ingólfsson, D. P. Tieleman, and M. S. Sansom, “Computational Modeling of Realistic Cell Membranes,” Chem. Rev. 119, 6184–6226 (2019).
  • Karathanou and Bondar [2022] K. Karathanou and A.-N. Bondar, “Algorithm to catalogue topologies of dynamic lipid hydrogen-bond networks,” Biochim. Biophys. Acta 1864, 183859 (2022).
  • Alper, Bassolino-Klimas, and Stouch [1993] H. E. Alper, D. Bassolino-Klimas, and T. R. Stouch, “The limiting behavior of water hydrating a phospholipid monolayer: A computer simulation study,” J. Chem. Phys. 99, 5547–5559 (1993).
  • Pasenkiewicz-Gierula et al. [1997] M. Pasenkiewicz-Gierula, Y. Takaoka, H. Miyagawa, K. Kitamura, and A. Kusumi, “Hydrogen Bonding of Water to Phosphatidylcholine in the Membrane As Studied by a Molecular Dynamics Simulation:  Location, Geometry, and Lipid-Lipid Bridging via Hydrogen-Bonded Water,” J. Phys. Chem. A 101, 3677–3691 (1997).
  • Feller [2000] S. E. Feller, “Molecular dynamics simulations of lipid bilayers,” Curr. Opin. Colloid Interface Sci. 5, 217–223 (2000).
  • Lopez et al. [2004] C. F. Lopez, S. O. Nielsen, M. L. Klein, and P. B. Moore, “Hydrogen Bonding Structure and Dynamics of Water at the Dimyristoylphosphatidylcholine Lipid Bilayer Surface from a Molecular Dynamics Simulation,” J. Phys. Chem. B 108, 6603–6610 (2004).
  • Bhide and Berkowitz [2005] S. Y. Bhide and M. L. Berkowitz, “Structure and dynamics of water at the interface with phospholipid bilayers,” J. Chem. Phys. 123, 224702 (2005).
  • Volkov et al. [2006] V. V. Volkov, F. Nuti, Y. Takaoka, R. Chelli, A. M. Papini, and R. Righini, “Hydration and Hydrogen Bonding of Carbonyls in Dimyristoyl-Phosphatidylcholine Bilayer,” J. Am. Chem. Soc. 128, 9466–9471 (2006).
  • von Hansen, Gekle, and Netz [2013] Y. von Hansen, S. Gekle, and R. R. Netz, “Anomalous Anisotropic Diffusion Dynamics of Hydration Water at Lipid Membranes,” Phys. Rev. Lett. 111, 118103 (2013).
  • Srivastava and Debnath [2018] A. Srivastava and A. Debnath, “Hydration dynamics of a lipid membrane: Hydrogen bond networks and lipid-lipid associations,” J. Chem. Phys. 148, 094901 (2018).
  • Calero and Franzese [2019] C. Calero and G. Franzese, “Membranes with different hydration levels: The interface between bound and unbound hydration water,” J. Mol. Liq. 273, 488–496 (2019).
  • Lee et al. [2019] E. Lee, A. Kundu, J. Jeon, and M. Cho, “Water hydrogen-bonding structure and dynamics near lipid multibilayer surface: Molecular dynamics simulation study with direct experimental comparison,” J. Chem. Phys. 151, 114705 (2019).
  • An et al. [2021] X. An, A. Majumder, J. McNeely, J. Yang, T. Puri, Z. He, T. Liang, J. K. Snyder, J. E. Straub, and B. M. Reinhard, “Interfacial hydration determines orientational and functional dimorphism of sterol-derived Raman tags in lipid-coated nanoparticles,” Proc. Natl. Acad. Sci. U.S.A. 118, e2105913118 (2021).
  • Higuchi et al. [2021] Y. Higuchi, Y. Asano, T. Kuwahara, and M. Hishida, “Rotational Dynamics of Water at the Phospholipid Bilayer Depending on the Head Groups Studied by Molecular Dynamics Simulations,” Langmuir 37, 5329–5338 (2021).
  • Malik and Debnath [2021] S. Malik and A. Debnath, “Dehydration induced dynamical heterogeneity and ordering mechanism of lipid bilayers,” J. Chem. Phys. 154, 174904 (2021).
  • Malik, Karmakar, and Debnath [2023] S. Malik, S. Karmakar, and A. Debnath, “Relaxation time scales of interfacial water upon fluid to ripple to gel phase transitions of bilayers,” J. Chem. Phys. 158, 114503 (2023).
  • Tu, Klein, and Tobias [1998] K. Tu, M. L. Klein, and D. J. Tobias, “Constant-Pressure Molecular Dynamics Investigation of Cholesterol Effects in a Dipalmitoylphosphatidylcholine Bilayer,” Biophys. J. 75, 2147–2156 (1998).
  • Chiu et al. [2002] S. W. Chiu, E. Jakobsson, R. J. Mashl, and H. L. Scott, “Cholesterol-Induced Modifications in Lipid Bilayers: A Simulation Study,” Biophys. J. 83, 1842–1853 (2002).
  • Hofsäß, Lindahl, and Edholm [2003] C. Hofsäß, E. Lindahl, and O. Edholm, “Molecular Dynamics Simulations of Phospholipid Bilayers with Cholesterol,” Biophys. J. 84, 2192–2206 (2003).
  • Pandit, Bostick, and Berkowitz [2004] S. A. Pandit, D. Bostick, and M. L. Berkowitz, “Complexation of Phosphatidylcholine Lipids with Cholesterol,” Biophys. J. 86, 1345–1356 (2004).
  • Alwarawrah, Dai, and Huang [2010] M. Alwarawrah, J. Dai, and J. Huang, “A Molecular View of the Cholesterol Condensing Effect in DOPC Lipid Bilayers,” J. Phys. Chem. B 114, 7516–7523 (2010).
  • Saito and Shinoda [2011] H. Saito and W. Shinoda, “Cholesterol Effect on Water Permeability through DPPC and PSM Lipid Bilayers: A Molecular Dynamics Study,” J. Phys. Chem. B 115, 15241–15250 (2011).
  • Sodt, Pastor, and Lyman [2015] A. J. Sodt, R. W. Pastor, and E. Lyman, “Hexagonal Substructure and Hydrogen Bonding in Liquid-Ordered Phases Containing Palmitoyl Sphingomyelin,” Biophys. J. 109, 948–955 (2015).
  • Boughter et al. [2016] C. T. Boughter, V. Monje-Galvan, W. Im, and J. B. Klauda, “Influence of Cholesterol on Phospholipid Bilayer Structure and Dynamics,” J. Phys. Chem. B 120, 11761–11772 (2016).
  • Elola and Rodriguez [2018] M. D. Elola and J. Rodriguez, “Influence of Cholesterol on the Dynamics of Hydration in Phospholipid Bilayers,” J. Phys. Chem. B 122, 5897–5907 (2018).
  • Pantelopulos and Straub [2018] G. A. Pantelopulos and J. E. Straub, “Regimes of Complex Lipid Bilayer Phases Induced by Cholesterol Concentration in MD Simulation,” Biophys. J. 115, 2167–2178 (2018).
  • Päslack et al. [2019] C. Päslack, J. C. Smith, M. Heyden, and L. V. Schäfer, “Hydration-mediated stiffening of collective membrane dynamics by cholesterol,” Phys. Chem. Chem. Phys. 21, 10370–10376 (2019).
  • Kumari, Kumari, and Kashyap [2019] P. Kumari, M. Kumari, and H. K. Kashyap, “Counter-effects of Ethanol and Cholesterol on the Heterogeneous PSM–POPC Lipid Membrane: A Molecular Dynamics Simulation Study,” J. Phys. Chem. B 123, 9616–9628 (2019).
  • Elkins et al. [2021] M. R. Elkins, A. Bandara, G. A. Pantelopulos, J. E. Straub, and M. Hong, “Direct Observation of Cholesterol Dimers and Tetramers in Lipid Bilayers,” J. Phys. Chem. B 125, 1825–1837 (2021).
  • Antila et al. [2022] H. S. Antila, A. Wurl, O. S. Ollila, M. S. Miettinen, and T. M. Ferreira, “Rotational decoupling between the hydrophilic and hydrophobic regions in lipid membranes,” Biophys. J. 121, 68–78 (2022).
  • Cheng et al. [2014] C.-Y. Cheng, L. L. C. Olijve, R. Kausik, and S. Han, “Cholesterol enhances surface water diffusion of phospholipid bilayers,” J. Chem. Phys. 141, 22D513 (2014).
  • Pyne, Pyne, and Mitra [2022] S. Pyne, P. Pyne, and R. K. Mitra, “Addition of cholesterol alters the hydration at the surface of model lipids: A spectroscopic investigation,” Phys. Chem. Chem. Phys. 24, 20381–20389 (2022).
  • Oh, Oh, and Weaver [2020] M. I. Oh, C. I. Oh, and D. F. Weaver, “Effect of Cholesterol on the Structure of Networked Water at the Surface of a Model Lipid Membrane,” J. Phys. Chem. B 124, 3686–3694 (2020).
  • Jo et al. [2008] S. Jo, T. Kim, V. G. Iyer, and W. Im, “CHARMM-GUI: A web-based graphical user interface for CHARMM,” J. Comput. Chem. 29, 1859–1865 (2008).
  • Jo et al. [2009] S. Jo, J. B. Lim, J. B. Klauda, and W. Im, “CHARMM-GUI Membrane Builder for Mixed Bilayers and Its Application to Yeast Membranes,” Biophys. J. 97, 50–58 (2009).
  • Brooks et al. [2009] B. R. Brooks, C. L. Brooks, A. D. Mackerell, L. Nilsson, R. J. Petrella, B. Roux, Y. Won, G. Archontis, C. Bartels, S. Boresch, A. Caflisch, L. Caves, Q. Cui, A. R. Dinner, M. Feig, S. Fischer, J. Gao, M. Hodoscek, W. Im, K. Kuczera, T. Lazaridis, J. Ma, V. Ovchinnikov, E. Paci, R. W. Pastor, C. B. Post, J. Z. Pu, M. Schaefer, B. Tidor, R. M. Venable, H. L. Woodcock, X. Wu, W. Yang, D. M. York, and M. Karplus, “CHARMM: The biomolecular simulation program,” J. Comput. Chem. 30, 1545–1614 (2009).
  • Wu et al. [2014] E. L. Wu, X. Cheng, S. Jo, H. Rui, K. C. Song, E. M. Dávila-Contreras, Y. Qi, J. Lee, V. Monje-Galvan, R. M. Venable, J. B. Klauda, and W. Im, “CHARMM-GUI Membrane Builder toward realistic biological membrane simulations,” J. Comput. Chem. 35, 1997–2004 (2014).
  • Lee et al. [2016] J. Lee, X. Cheng, J. M. Swails, M. S. Yeom, P. K. Eastman, J. A. Lemkul, S. Wei, J. Buckner, J. C. Jeong, Y. Qi, S. Jo, V. S. Pande, D. A. Case, C. L. Brooks, A. D. MacKerell, J. B. Klauda, and W. Im, “CHARMM-GUI Input Generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM Simulations Using the CHARMM36 Additive Force Field,” J. Chem. Theory Comput. 12, 405–413 (2016).
  • Huang and MacKerell Jr [2013] J. Huang and A. D. MacKerell Jr, “CHARMM36 all-atom additive protein force field: Validation based on comparison to NMR data,” J. Comput. Chem. 34, 2135–2145 (2013).
  • Jorgensen et al. [1983] W. L. Jorgensen, J. Chandrasekhar, J. D. Madura, R. W. Impey, and M. L. Klein, “Comparison of simple potential functions for simulating liquid water,” J. Chem. Phys. 79, 926–935 (1983).
  • Abraham et al. [2015] M. J. Abraham, T. Murtola, R. Schulz, S. Páll, J. C. Smith, B. Hess, and E. Lindahl, “GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers,” SoftwareX 1–2, 19–25 (2015).
  • Willard and Chandler [2010] A. P. Willard and D. Chandler, “Instantaneous Liquid Interfaces,” J. Phys. Chem. B 114, 1954–1958 (2010).
  • Luzar and Chandler [1996a] A. Luzar and D. Chandler, “Effect of Environment on Hydrogen Bond Dynamics in Liquid Water,” Phys. Rev. Lett. 76, 928–931 (1996a).
  • Luzar and Chandler [1996b] A. Luzar and D. Chandler, “Hydrogen-bond kinetics in liquid water,” Nature 379, 55–57 (1996b).
  • Laage and Hynes [2006] D. Laage and J. T. Hynes, “A Molecular Jump Mechanism of Water Reorientation,” Science 311, 832–835 (2006).
  • Kumar, Schmidt, and Skinner [2007] R. Kumar, J. R. Schmidt, and J. L. Skinner, “Hydrogen bonding definitions and dynamics in liquid water,” J. Chem. Phys. 126, 204107 (2007).
  • Kikutsuji, Kim, and Matubayasi [2018] T. Kikutsuji, K. Kim, and N. Matubayasi, “How do hydrogen bonds break in supercooled water?: Detecting pathways not going through saddle point of two-dimensional potential of mean force,” J. Chem. Phys. 148, 244501 (2018).
  • Kikutsuji, Kim, and Matubayasi [2019] T. Kikutsuji, K. Kim, and N. Matubayasi, “Consistency of geometrical definitions of hydrogen bonds based on the two-dimensional potential of mean force with respect to the time correlation in liquid water over a wide range of temperatures,” J. Mol. Liq. 294, 111603 (2019).
  • Kikutsuji, Kim, and Matubayasi [2021] T. Kikutsuji, K. Kim, and N. Matubayasi, “Transition pathway of hydrogen bond switching in supercooled water analyzed by the Markov state model,” J. Chem. Phys. 154, 234501 (2021).
  • Shikata et al. [2023] K. Shikata, T. Kikutsuji, N. Yasoshima, K. Kim, and N. Matubayasi, “Revealing the hidden dynamics of confined water in acrylate polymers: Insights from hydrogen-bond lifetime analysis,” J. Chem. Phys. 158, 174901 (2023).
  • Rapaport [1983] D. Rapaport, “Hydrogen bonds in water: Network organization and lifetimes,” Mol. Phys. 50, 1151–1162 (1983).
  • Efron [1992] B. Efron, “Bootstrap Methods: Another Look at the Jackknife,” in Breakthroughs in Statistics, edited by S. Kotz and N. L. Johnson (Springer, New York, NY, 1992) pp. 569–593.

Supplementary Material

Influence of cholesterol on hydrogen-bond dynamics of water molecules in lipid-bilayer systems at varying temperatures

Kokoro Shikata, Kento Kasahara, Nozomi Morishita Watanabe, Hiroshi Umakoshi, Kang Kim, and Nobuyuki Matubayasi

Division of Chemical Engineering, Department of Materials Engineering Science, Graduate School of Engineering Science, Osaka University, Toyonaka, Osaka 560-8531, Japan

Table S1: Equilibration scheme.
No. process dt𝑑𝑡dtitalic_d italic_t [fs] time [ns] integrater kz[kJ/molnm2]subscript𝑘𝑧delimited-[]kJmolsuperscriptnm2k_{z}~{}[\mathrm{kJ/mol\,nm^{2}}]italic_k start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT [ roman_kJ / roman_mol roman_nm start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] kdih[kJ/molrad2]subscript𝑘dihdelimited-[]kJmolsuperscriptrad2k_{\mathrm{dih}}~{}\mathrm{[kJ/mol\,rad^{2}]}italic_k start_POSTSUBSCRIPT roman_dih end_POSTSUBSCRIPT [ roman_kJ / roman_mol roman_rad start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ]
1 Energy Minimization - 5000 (steps) steepest descent 1000 1000
2 NVT𝑁𝑉𝑇NVTitalic_N italic_V italic_T 1 0.125 Leap-Flog 1000 1000
3 NVT𝑁𝑉𝑇NVTitalic_N italic_V italic_T 1 0.125 Leap-Flog 400 400
4 NPT𝑁𝑃𝑇NPTitalic_N italic_P italic_T 1 0.125 Leap-Flog 400 200
5 NPT𝑁𝑃𝑇NPTitalic_N italic_P italic_T 2 0.5 Leap-Flog 200 200
6 NPT𝑁𝑃𝑇NPTitalic_N italic_P italic_T 2 0.5 Leap-Flog 40 100
7 NPT𝑁𝑃𝑇NPTitalic_N italic_P italic_T 2 0.5 Leap-Flog 0 0
8 NPT𝑁𝑃𝑇NPTitalic_N italic_P italic_T 2 10 Leap-Flog 0 0
9 NPT𝑁𝑃𝑇NPTitalic_N italic_P italic_T 2 100 Leap-Flog 0 0
10 NPT𝑁𝑃𝑇NPTitalic_N italic_P italic_T 2 500 Leap-Flog 0 0
11 NPT𝑁𝑃𝑇NPTitalic_N italic_P italic_T 2 3000 Leap-Flog 0 0
12 Production (NPT𝑁𝑃𝑇NPTitalic_N italic_P italic_T) 2 10 Leap-Flog 0 0
Refer to caption
Figure S1: Time evolution of surface area S𝑆Sitalic_S in the x𝑥xitalic_x-y𝑦yitalic_y plane during the equilibration at 303K. (a) and (b): DPPC; (c) and (d): PSM. (a) and (b) refer to the systems with Chol, while (c) and (d) are for those without Chol. Systems 01, 02, and 03 correspond to three distinct runs with different initial configurations.
Refer to caption
Figure S2: Time evolution of surface area S𝑆Sitalic_S in the x𝑥xitalic_x-y𝑦yitalic_y plane during the equilibration at 323K. (a) and (b): DPPC; (c) and (d): PSM. (a) and (b) refer to the systems with Chol, while (c) and (d) are for those without Chol. Systems 01, 02, and 03 correspond to three distinct runs with different initial configurations.
Refer to caption
Figure S3: Number density distributions of OD7D7{}^{\text{D}7}start_FLOATSUPERSCRIPT D 7 end_FLOATSUPERSCRIPT in DPPC and OS5S5{}^{\text{S}5}start_FLOATSUPERSCRIPT S 5 end_FLOATSUPERSCRIPT in PSM along the z𝑧zitalic_z-direction. For comparison, number density distributions of the water oxygen atoms (Oww{}^{\text{w}}start_FLOATSUPERSCRIPT w end_FLOATSUPERSCRIPT) are plotted, which are same as those in Fig. 2. Solid lines represent pure membrane systems, while dashed lines represent systems containing Chol. (a) and (b) refer to DPPC, and (c) and (d) refer to PSM.
Refer to caption
Figure S4: Number density distributions of nitrogen (N+), and phosphorus (P) atoms along the z𝑧zitalic_z-direction. Solid lines represent pure membrane systems, while dashed lines represent systems containing Chol. (a) and (b) refer to DPPC, and (c) and (d) refer to PSM.
Refer to caption
Figure S5: Snapshots at 323 K for (a) DPPC with Chol, (b) DPPC without Chol, (c) PSM with Chol, and (d) PSM without Chol.
Refer to caption
Figure S6: 2D PMF, W(roo,β)𝑊subscript𝑟oo𝛽W(r_{\mathrm{oo}},\beta)italic_W ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ), between water molecules in bulk water at 1 g/cm3 at (a) 303 K and (b) 323 K.
Refer to caption
Figure S7: 2D PMF, W(roo,β)𝑊subscript𝑟oo𝛽W(r_{\mathrm{oo}},\beta)italic_W ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ), between water oxygen (OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT) and each acceptor oxygen in DPPC with Chol at 323 K. Black points represent saddle points.
Refer to caption
Figure S8: 2D PMF, W(roo,β)𝑊subscript𝑟oo𝛽W(r_{\mathrm{oo}},\beta)italic_W ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ), between water oxygen (OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT) and each acceptor oxygen in DPPC without Chol at 303 K. Black points represent saddle points.
Refer to caption
Figure S9: 2D PMF, W(roo,β)𝑊subscript𝑟oo𝛽W(r_{\mathrm{oo}},\beta)italic_W ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ), between water oxygen (OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT) and each acceptor oxygen in DPPC without Chol at 323 K. Black points represent saddle points.
Refer to caption
Figure S10: 2D PMF, W(roo,β)𝑊subscript𝑟oo𝛽W(r_{\mathrm{oo}},\beta)italic_W ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ), between water oxygen (OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT) and each acceptor oxygen in PSM with Chol at 323 K. Black points represent saddle points.
Refer to caption
Figure S11: 2D PMF, W(roo,β)𝑊subscript𝑟oo𝛽W(r_{\mathrm{oo}},\beta)italic_W ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ), between water oxygen (OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT) and each acceptor oxygen in PSM without Chol at 303 K.
Refer to caption
Figure S12: 2D PMF, W(roo,β)𝑊subscript𝑟oo𝛽W(r_{\mathrm{oo}},\beta)italic_W ( italic_r start_POSTSUBSCRIPT roman_oo end_POSTSUBSCRIPT , italic_β ), between water oxygen (OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT) and each acceptor oxygen in PSM without Chol at 323 K. Black points represent saddle points.
Refer to caption
Figure S13: Conditional probability C1,j(t)subscript𝐶1𝑗𝑡C_{1,j}(t)italic_C start_POSTSUBSCRIPT 1 , italic_j end_POSTSUBSCRIPT ( italic_t ), representing transition dynamics from region 1111 at the initial time t=0𝑡0t=0italic_t = 0 to region 2 or remaining within the same region 1 during the time interval t𝑡titalic_t [(a) DPPC at 303 K, (b) DPPC at 323 K, (c) PSM at 303 K, and (d) PSM at 323 K].
Refer to caption
Figure S14: H-bond time correlation function PHB(t)subscript𝑃HB𝑡P_{\mathrm{HB}}(t)italic_P start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT ( italic_t ) with respect to acceptor oxygens in the DPPC systems. The solid line represents the result of fitting with the KWW function, PHB(t)exp[(t/τKWW)βKWW]similar-to-or-equalssubscript𝑃HB𝑡superscript𝑡subscript𝜏KWWsubscript𝛽KWWP_{\mathrm{HB}}(t)\simeq\exp[-(t/\tau_{\mathrm{KWW}})^{\beta_{\mathrm{KWW}}}]italic_P start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT ( italic_t ) ≃ roman_exp [ - ( italic_t / italic_τ start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_β start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ].
Refer to caption
Figure S15: H-bond time correlation function PHB(t)subscript𝑃HB𝑡P_{\mathrm{HB}}(t)italic_P start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT ( italic_t ) with respect to acceptor oxygens in the PSM systems. The solid line represents the result of fitting with the KWW function, PHB(t)exp[(t/τKWW)βKWW]similar-to-or-equalssubscript𝑃HB𝑡superscript𝑡subscript𝜏KWWsubscript𝛽KWWP_{\mathrm{HB}}(t)\simeq\exp[-(t/\tau_{\mathrm{KWW}})^{\beta_{\mathrm{KWW}}}]italic_P start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT ( italic_t ) ≃ roman_exp [ - ( italic_t / italic_τ start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_β start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ].
Table S2: τKWWsubscript𝜏KWW\tau_{\mathrm{KWW}}italic_τ start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT of acceptor oxygen atoms in the DPPC systems with and without Chol. The error is provided at the standard error and is not shown when it is smaller than 0.010.010.010.01 ps.
303 K [ps] 323 K [ps]
pure Chol 10 % pure Chol 10 %
OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT 1.211.211.211.21 1.181.181.181.18 0.940.940.940.94 0.930.930.930.93
O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT 29.28±0.02plus-or-minus29.280.0229.28\pm 0.0229.28 ± 0.02 41.7±0.2plus-or-minus41.70.241.7\pm 0.241.7 ± 0.2 14.2714.27{14.27}14.27 15.07±0.01plus-or-minus15.070.01{15.07\pm 0.01}15.07 ± 0.01
O2superscriptO2\mathrm{O}^{\mathrm{2}}roman_O start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT 24.97±0.04plus-or-minus24.970.0424.97\pm 0.0424.97 ± 0.04 40.0±0.3plus-or-minus40.00.340.0\pm 0.340.0 ± 0.3 8.818.818.818.81 9.359.359.359.35
O3superscriptO3\mathrm{O}^{\mathrm{3}}roman_O start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 22.26±0.02plus-or-minus22.260.0222.26\pm 0.0222.26 ± 0.02 29.6±0.1plus-or-minus29.60.129.6\pm 0.129.6 ± 0.1 10.3210.32{10.32}10.32 10.7010.70{10.70}10.70
OD5superscriptOD5\mathrm{O}^{\mathrm{D5}}roman_O start_POSTSUPERSCRIPT D5 end_POSTSUPERSCRIPT 51.1±0.1plus-or-minus51.10.151.1\pm 0.151.1 ± 0.1 88.7±0.5plus-or-minus88.70.588.7\pm 0.588.7 ± 0.5 18.20±0.02plus-or-minus18.200.02{18.20\pm 0.02}18.20 ± 0.02 19.74±0.02plus-or-minus19.740.02{19.74\pm 0.02}19.74 ± 0.02
OD7superscriptOD7\mathrm{O}^{\mathrm{D7}}roman_O start_POSTSUPERSCRIPT D7 end_POSTSUPERSCRIPT 40.2±0.1plus-or-minus40.20.140.2\pm 0.140.2 ± 0.1 81.1±0.4plus-or-minus81.10.481.1\pm 0.481.1 ± 0.4 15.01±0.01plus-or-minus15.010.01{15.01\pm 0.01}15.01 ± 0.01 16.74±0.02plus-or-minus16.740.02{16.74\pm 0.02}16.74 ± 0.02
OCholsuperscriptOChol\mathrm{O}^{\mathrm{Chol}}roman_O start_POSTSUPERSCRIPT roman_Chol end_POSTSUPERSCRIPT - 110±1plus-or-minus1101110\pm 1110 ± 1 - 18.3±0.1plus-or-minus18.30.118.3\pm 0.118.3 ± 0.1
Table S3: βKWWsubscript𝛽KWW\beta_{\mathrm{KWW}}italic_β start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT of acceptor oxygen atoms in the DPPC systems with and without Chol. The error is not shown since it is smaller than 0.010.010.010.01.
303 K 323 K
pure Chol 10 % pure Chol 10 %
OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT 0.600.600.600.60 0.610.610.610.61 0.600.600.600.60 0.600.600.600.60
O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT 0.530.530.530.53 0.510.510.510.51 0.560.560.560.56 0.550.550.550.55
O2superscriptO2\mathrm{O}^{\mathrm{2}}roman_O start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT 0.340.340.340.34 0.320.320.320.32 0.370.370.370.37 0.370.370.370.37
O3superscriptO3\mathrm{O}^{\mathrm{3}}roman_O start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 0.510.510.510.51 0.480.480.480.48 0.520.520.520.52 0.520.520.520.52
OD5superscriptOD5\mathrm{O}^{\mathrm{D5}}roman_O start_POSTSUPERSCRIPT D5 end_POSTSUPERSCRIPT 0.420.420.420.42 0.390.390.390.39 0.450.450.450.45 0.440.440.440.44
OD7superscriptOD7\mathrm{O}^{\mathrm{D7}}roman_O start_POSTSUPERSCRIPT D7 end_POSTSUPERSCRIPT 0.410.410.410.41 0.380.380.380.38 0.440.440.440.44 0.440.440.440.44
OCholsuperscriptOChol\mathrm{O}^{\mathrm{Chol}}roman_O start_POSTSUPERSCRIPT roman_Chol end_POSTSUPERSCRIPT - 0.370.370.370.37 - 0.420.420.420.42
Table S4: τHBsubscript𝜏HB\tau_{\mathrm{HB}}italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT of acceptor oxygen atoms in the DPPC systems with and without Chol. The error is provided at the standard error and is not shown when it is smaller than 0.010.010.010.01 ps.
303 K [ps] 323 K [ps]
pure Chol 10 % pure Chol 10 %
OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT 1.811.811.811.81 1.761.761.761.76 1.421.421.421.42 1.401.401.401.40
O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT 52.5±0.5plus-or-minus52.50.552.5\pm 0.552.5 ± 0.5 81±4plus-or-minus81481\pm 481 ± 4 23.8±0.2plus-or-minus23.80.223.8\pm 0.223.8 ± 0.2 25.4±0.3plus-or-minus25.40.325.4\pm 0.325.4 ± 0.3
O2superscriptO2\mathrm{O}^{\mathrm{2}}roman_O start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT 135±3plus-or-minus1353135\pm 3135 ± 3 280±30plus-or-minus28030280\pm 30280 ± 30 37.3±0.5plus-or-minus37.30.537.3\pm 0.537.3 ± 0.5 40.9±0.5plus-or-minus40.90.540.9\pm 0.540.9 ± 0.5
O3superscriptO3\mathrm{O}^{\mathrm{3}}roman_O start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 43.0±0.7plus-or-minus43.00.743.0\pm 0.743.0 ± 0.7 64±3plus-or-minus64364\pm 364 ± 3 19.2±0.2plus-or-minus19.20.219.2\pm 0.219.2 ± 0.2 20.1±0.2plus-or-minus20.10.220.1\pm 0.220.1 ± 0.2
OD5superscriptOD5\mathrm{O}^{\mathrm{D5}}roman_O start_POSTSUPERSCRIPT D5 end_POSTSUPERSCRIPT 151±3plus-or-minus1513151\pm 3151 ± 3 330±30plus-or-minus33030330\pm 30330 ± 30 46.1±0.5plus-or-minus46.10.546.1\pm 0.546.1 ± 0.5 51.4±0.7plus-or-minus51.40.751.4\pm 0.751.4 ± 0.7
OD7superscriptOD7\mathrm{O}^{\mathrm{D7}}roman_O start_POSTSUPERSCRIPT D7 end_POSTSUPERSCRIPT 125±3plus-or-minus1253125\pm 3125 ± 3 310±20plus-or-minus31020310\pm 20310 ± 20 38.4±0.4plus-or-minus38.40.438.4\pm 0.438.4 ± 0.4 44.4±0.5plus-or-minus44.40.544.4\pm 0.544.4 ± 0.5
OCholsuperscriptOChol\mathrm{O}^{\mathrm{Chol}}roman_O start_POSTSUPERSCRIPT roman_Chol end_POSTSUPERSCRIPT - 450±60plus-or-minus45060450\pm 60450 ± 60 - 53±3plus-or-minus53353\pm 353 ± 3
Table S5: τKWWsubscript𝜏KWW\tau_{\mathrm{KWW}}italic_τ start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT of acceptor oxygen atoms in the PSM systems with and without Chol. The error is provided at the standard error and is not shown when it is smaller than 0.010.010.010.01 ps.
303 K [ps] 323 K [ps]
pure Chol 10 % pure Chol 10 %
OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT 1.191.191.191.19 1.181.181.181.18 0.920.920.920.92 0.910.910.910.91
O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT 53.2±0.1plus-or-minus53.20.153.2\pm 0.153.2 ± 0.1 65.8±0.2plus-or-minus65.80.265.8\pm 0.265.8 ± 0.2 27.1±0.1plus-or-minus27.10.127.1\pm 0.127.1 ± 0.1 28.41±0.02plus-or-minus28.410.02{28.41\pm 0.02}28.41 ± 0.02
O3superscriptO3\mathrm{O}^{\mathrm{3}}roman_O start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 39.9±0.1plus-or-minus39.90.139.9\pm 0.139.9 ± 0.1 45.3±0.1plus-or-minus45.30.145.3\pm 0.145.3 ± 0.1 17.68±0.03plus-or-minus17.680.03{17.68\pm 0.03}17.68 ± 0.03 19.14±0.03plus-or-minus19.140.03{19.14\pm 0.03}19.14 ± 0.03
OS4superscriptOS4\mathrm{O}^{\mathrm{S4}}roman_O start_POSTSUPERSCRIPT S4 end_POSTSUPERSCRIPT 55.1±0.3plus-or-minus55.10.355.1\pm 0.355.1 ± 0.3 62.9±0.4plus-or-minus62.90.462.9\pm 0.462.9 ± 0.4 20.3±0.1plus-or-minus20.30.120.3\pm 0.120.3 ± 0.1 21.08±0.03plus-or-minus21.080.03{21.08\pm 0.03}21.08 ± 0.03
OS5superscriptOS5\mathrm{O}^{\mathrm{S5}}roman_O start_POSTSUPERSCRIPT S5 end_POSTSUPERSCRIPT 236±1plus-or-minus2361236\pm 1236 ± 1 400±2plus-or-minus4002400\pm 2400 ± 2 88.5±0.4plus-or-minus88.50.488.5\pm 0.488.5 ± 0.4 109.4±0.4plus-or-minus109.40.4{109.4\pm 0.4}109.4 ± 0.4
OCholsuperscriptOChol\mathrm{O}^{\mathrm{Chol}}roman_O start_POSTSUPERSCRIPT roman_Chol end_POSTSUPERSCRIPT - 670±20plus-or-minus67020670\pm 20670 ± 20 - 63.9±0.7plus-or-minus63.90.763.9\pm 0.763.9 ± 0.7
Table S6: βKWWsubscript𝛽KWW\beta_{\mathrm{KWW}}italic_β start_POSTSUBSCRIPT roman_KWW end_POSTSUBSCRIPT of acceptor oxygen atoms in the PSM systems with and without Chol. The error is not shown since it is smaller than 0.010.010.010.01.
303 K 323 K
pure Chol 10 % pure Chol 10 %
OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT 0.610.610.610.61 0.610.610.610.61 0.600.600.600.60 0.600.600.600.60
O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT 0.480.480.480.48 0.460.460.460.46 0.500.500.500.50 0.500.500.500.50
O3superscriptO3\mathrm{O}^{\mathrm{3}}roman_O start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 0.440.440.440.44 0.420.420.420.42 0.460.460.460.46 0.450.450.450.45
OS4superscriptOS4\mathrm{O}^{\mathrm{S4}}roman_O start_POSTSUPERSCRIPT S4 end_POSTSUPERSCRIPT 0.340.340.340.34 0.330.330.330.33 0.360.360.360.36 0.360.360.360.36
OS5superscriptOS5\mathrm{O}^{\mathrm{S5}}roman_O start_POSTSUPERSCRIPT S5 end_POSTSUPERSCRIPT 0.390.390.390.39 0.370.370.370.37 0.410.410.410.41 0.410.410.410.41
OCholsuperscriptOChol\mathrm{O}^{\mathrm{Chol}}roman_O start_POSTSUPERSCRIPT roman_Chol end_POSTSUPERSCRIPT - 0.310.310.310.31 - 0.360.360.360.36
Table S7: τHBsubscript𝜏HB\tau_{\mathrm{HB}}italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT of acceptor oxygen atoms in the PSM systems with and without Chol. The error is provided at the standard error and is not shown when it is smaller than 0.010.010.010.01 ps.
303 K [ps] 323 K [ps]
pure Chol 10 % pure Chol 10 %
OwsuperscriptOw\mathrm{O}^{\mathrm{w}}roman_O start_POSTSUPERSCRIPT roman_w end_POSTSUPERSCRIPT 1.771.771.771.77 1.741.741.741.74 1.381.381.381.38 1.361.361.361.36
O1superscriptO1\mathrm{O}^{\mathrm{1}}roman_O start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT 113±4plus-or-minus1134113\pm 4113 ± 4 153±7plus-or-minus1537153\pm 7153 ± 7 54±2plus-or-minus54254\pm 254 ± 2 57.3±0.7plus-or-minus57.30.757.3\pm 0.757.3 ± 0.7
O3superscriptO3\mathrm{O}^{\mathrm{3}}roman_O start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 104±2plus-or-minus1042104\pm 2104 ± 2 129±6plus-or-minus1296129\pm 6129 ± 6 43±1plus-or-minus43143\pm 143 ± 1 47±1plus-or-minus47147\pm 147 ± 1
OS4superscriptOS4\mathrm{O}^{\mathrm{S4}}roman_O start_POSTSUPERSCRIPT S4 end_POSTSUPERSCRIPT 320±30plus-or-minus32030320\pm 30320 ± 30 390±30plus-or-minus39030390\pm 30390 ± 30 95±6plus-or-minus95695\pm 695 ± 6 98±2plus-or-minus98298\pm 298 ± 2
OS5superscriptOS5\mathrm{O}^{\mathrm{S5}}roman_O start_POSTSUPERSCRIPT S5 end_POSTSUPERSCRIPT 860±80plus-or-minus86080860\pm 80860 ± 80 1600±100plus-or-minus16001001600\pm 1001600 ± 100 280±20plus-or-minus28020280\pm 20280 ± 20 350±20plus-or-minus35020350\pm 20350 ± 20
OCholsuperscriptOChol\mathrm{O}^{\mathrm{Chol}}roman_O start_POSTSUPERSCRIPT roman_Chol end_POSTSUPERSCRIPT - 6000±1000plus-or-minus600010006000\pm 10006000 ± 1000 - 290±30plus-or-minus29030290\pm 30290 ± 30
Refer to caption
Figure S16: τHBsubscript𝜏HB\tau_{\mathrm{HB}}italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT vs. ΔG/kBTΔ𝐺subscript𝑘B𝑇\Delta G/k_{\mathrm{B}}Troman_Δ italic_G / italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T for DPPC (a) and PSM (b) systems. The activation energy ΔGΔ𝐺\Delta Groman_Δ italic_G is determined by the free energy difference between the most stable and saddle points on the 2D PMF, W(rOO,β)𝑊subscript𝑟OO𝛽W(r_{\mathrm{OO}},\beta)italic_W ( italic_r start_POSTSUBSCRIPT roman_OO end_POSTSUBSCRIPT , italic_β ) (see Figs. 5 and 6, and Figs. S7-S12). The straight line represents the Arrhenius equation, τHB=Aexp(ΔG/kBT)subscript𝜏HB𝐴Δ𝐺subscript𝑘B𝑇\tau_{\mathrm{HB}}=A\exp(\Delta G/k_{\mathrm{B}}T)italic_τ start_POSTSUBSCRIPT roman_HB end_POSTSUBSCRIPT = italic_A roman_exp ( start_ARG roman_Δ italic_G / italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T end_ARG ), where A𝐴Aitalic_A is determined by fitting. The values of A𝐴Aitalic_A are 0.183 ps and 0.385 ps for DPPC and PSM systems, respectively.