You are on page 1of 6

pubs.acs.

org/Langmuir Article

Enhanced Evaporation of Ultrathin Water Films on Silicon-


Terminated Si3N4 Nanopore Membranes
Runkeng Liu and Zhenyu Liu*

Cite This: https://doi.org/10.1021/acs.langmuir.1c01212 Read Online

ACCESS Metrics & More Article Recommendations


Downloaded via MASSACHUSETTS INST OF TECHNOLOGY on August 18, 2021 at 12:19:29 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Water evaporation confined in nanoscale is a


ubiquitous phenomenon in nature and has crucial importance in
a broad range of technical applications. With the nonequilibrium
molecular dynamics simulation, we elucidate nanothin water film
evaporation characteristics on a silicon nitride nanopore
membrane considering the effects of pore size and pore chemistry.
Pore chemistry plays the main role in regulating the evaporation
flux. The terminated Si atoms on the pore surface lead to a higher
evaporation intensity than the N ones. We attribute this
enhancement to the transition of the structural properties of
fluid, where liquid molecules are packed loosely and disorderedly
under the inducement of terminated silicon atoms. The findings in
the present work can contribute to the fundamental understanding
of the nanopore-enhanced evaporation process and provide new guidance to the design of advanced nanopore membrane materials.

■ INTRODUCTION
Evaporation on tiny pore structures is the most important
significantly due to the curvature of the liquid/vapor interface.
In addition, Li et al.15 measured the evaporation flux from a
driving force for the liquid transport process in plants. This single Si3N4 nanopore membrane with the diameter ranging
mechanism has been adopted to develop new and critical from 300 to 27 nm. The result showed that the evaporation
techniques, such as membrane-based thermal desalination,1−3 rate increases substantially as the pore size decreases. It can
chip cooling,4−7 solar-driven interfacial evaporation,8,9 and therefore be expected that the evaporation rate would be
nanofluidic generation.10,11 For further improvement of these enhanced as the pore diameter is reduced toward a few
nanopore-evaporation-based devices, it is urgent to reveal the nanometers; however, the existing theoretical models and
enhancement mechanism of nanothin liquid film evaporation experimental approaches are incapable of accurately analyzing
on nanopore structures. the phase transition mechanism under this dimension.
It has been confirmed that ultrathin liquid film evaporation Fortunately, the challenge can be addressed through the
is the most efficient liquid−vapor phase change mode through molecular dynamics (MD) simulation method because it
experimental and theoretical investigations in recent allows us to explore the kinetics and thermodynamics of fluid
years.12−15 Dai et al.16 reported the capillary evaporation and obtain a deeper insight into the physical understanding of
performance on micromembranes, which are made of sintered evaporation behavior from an atomistic perspective. Akkus et
single-layer copper mesh bonded on microchannels. The al.18 conducted the MD simulation to study the evaporation of
microscale pores provide the capillary pressure, and the argon fluid in capillary nanoconduits. Their results showed that
microchannels reduce the flow resistance, which significantly mass transport and net evaporation exist in the adsorbed layer,
enhance the evaporation efficiency. Shi et al.17 created the which was believed to be at the equilibrium of evaporation and
nanoporous synthetic leaves to mimic the natural transpiration condensation in the past. Kieu et al.19 investigated the
process. The result showed that as the water menisci recede evaporation characteristics of water inside the graphene
into the nanopore, they would self-stabilize by locally
concentrating vapor within the dried-up top regions of
nanopores. Lu et al.13,14 explored the interfacial evaporation Received: May 5, 2021
performance with the nanoporous silicon nitride (Si3N4) Revised: August 3, 2021
membrane in which water is transported across pores under
pumping pressure and the gold layer coated at exit is
electrically heated to accelerate the evaporation. The results
suggested that the evaporation process will be facilitated

© XXXX American Chemical Society https://doi.org/10.1021/acs.langmuir.1c01212


A Langmuir XXXX, XXX, XXX−XXX
Langmuir pubs.acs.org/Langmuir Article

Figure 1. (a) Schematic of the computational system consisting of a Si3N4 nanopore, water, and graphene sheet. (b) Three nanopore configurations
considered in this work: Si-terminated, mix-terminated, and N-terminated nanopores. (c) Water density distribution along the z-axis direction on
the top of the nanopore at the beginning of the evaporation process.

channel (width in 4−14 nm) with tunable surface wettability


by adjusting the surface charge density. The results suggested
■ COMPUTATIONAL METHODS
In this work, the liquid film evaporation system consists of a graphene
that the evaporation rate shows the nonmonotonic depend- sheet, water, and a Si3N4 nanopore membrane as illustrated in Figure
ence on the hydrophilicity of the channel due to the 1a. The single-layer graphene, completely blocking the penetration of
competition of two different mechanisms: shape of the water molecules, is used as the bottom free boundary. The solid
membrane is modeled using the β-Si3N4 with a hexagon lattice
liquid−gas interface and the attraction between the solid and
structure and has a size of 6.34 × 6.1 × 5.9 nm3. To create the pore on
liquid. Li et al.20 employed the nonequilibrium MD method to the membrane, the Si and N atoms at the center of the membrane
study the self-pumping evaporation process inside the charged were eliminated, and the pore diameter d ranges from 3.4 to 0.5 nm.
carbon nanotube with a diameter of 2.4 nm. These studies The previous studies22−24 have indicated that, due to the hexagonal
focused on examining the capillary evaporation characteristics crystal structure of Si3N4, the possibility to decorate the pore wall
under the confinement condition. However, the work of surface with Si, N, or mixed atoms provides flexibility to design the
nanopore with an objective functionality. As shown in Figure 1b, three
Walther et al.21 showed that there tends to be one liquid film different nanopore configurations will be formed as d varies: Si-
outside the nanopore due to the solid−liquid interaction. terminated, mix-terminated, and N-terminated nanopores. It is of
Instead of inside the nanopore, the evaporation actually occurs interest to reveal how the terminated atoms on the pore surface affect
above the nanopore, but the understanding of the nanothin the thin liquid film evaporation. For clarity, the nanopores are
water film evaporation on nanopore membranes is still limited. classified according to the terminal atom type and the pore diameter
For the nanothin water film, the evaporation rate will be (N-0.53, Mix-0.75, Si-0.89, etc.).
All MD simulations were implemented with an open-source code
significantly influenced by the hydration effect induced by LAMMPS25 in which the calculated interactions include the
membrane surface charge,15 while the existing studies are intramolecular and intermolecular interactions. The intermolecular
incapable of accurately analyzing this kind of evaporation pair interactions consist of the short-range van der Waals term (with a
process. cutoff distance of 10 Å) and the long-range electrostatic one (with a
In this work, by conducting the systematical MD computed reciprocal space distance of 12 Å). Lennard-Jones potential
and Coulombic potential were applied to compute the van der Waals
simulations, the evaporation process of nanothin water film
and electrostatic interactions, respectively. The potential function was

ÅÄÅ ÑÉ
on Si3N4 nanopores with a diameter ranging from 3.4 to 0.5 solved explicitly as shown in eq 1:26

ÅÅij σ yz12 ij σ yz6ÑÑÑ


Åjj ij zz Ñ
nm was examined, for which the terminated atoms on the pore
Å j ij z Ñ
E = 4εijÅÅÅjj zz − jjj zzz ÑÑÑ +
ÅÅÅjk rij z{ j rij z ÑÑ 4πε0rij
surface will vary with pore size. The effects of pore size and qiqj

ÅÇ k { ÑÑÖ
pore chemistry on the evaporation performance were studied
based on the analysis of dynamics and structure properties of (1)
water. The present results identify the key factor behind the where εij and σij are the energy and length parameters, rij is the
superior enhancement of the nanopore-based evaporation distance between particles i and j, qi and qj are the charges on atoms i
process. and j, respectively, and ε0 is the vacuum permittivity. The particle−

B https://doi.org/10.1021/acs.langmuir.1c01212
Langmuir XXXX, XXX, XXX−XXX
Langmuir


pubs.acs.org/Langmuir Article

particle particle−mesh method with a relative accuracy of 10−5 was RESULTS AND DISCUSSION
adopted to calculate the electrostatic interaction when the two
charged atoms are over 12 Å apart. Water molecules were modeled The variations of water and nanopore temperatures during the
using the flexible TIP3P model27 in which the hydrogen−oxygen evaporation process are recorded and shown in Figure 2a. As
bond length is 0.9572 Å with a bond angle of 104.52°. The
CHARMM force field28 was adopted to describe the potential
interactions between the atoms of the nanopore membrane in which
the pairwise interaction is effectively turned off for all pairs of atoms
bonded to each other. In all simulations, the harmonic restraints that
pull each Si/N atom toward its equilibrium position in the
crystallographic membrane were included in the force field, and the
force constant was set to 1.0 kcal/(Å2 mol). The selection of atomic
charges of each silicon/nitrogen atom was based on the previous
reports:29,30 fixed positive charge of 0.7707e was assigned to each Si
atom, while the charge of each N atom was determined by the law of
charge neutrality (ranging from −0.59e to −0.61e). The correspond-
ing potential parameters are listed in Table 1, and the parameters
between two atoms are determined by the Lorentz−Berthelot mixing
rule, i.e., σij = (σi + σj)/2, and εij = (εi + εj)0.5.
Figure 2. (a) Liquid and solid temperatures during the evaporation
process for cases of Si-1.85, Mix-1.78, and N-1.65. (b) Liquid
Table 1. Potential Parameters Used in This Work temperature distribution along the z-axis direction for cases of Si-1.85,
Mix-1.78, and N-1.65.
atom ε (kcal/mol) σ (Å) q (e)
Si 0.31 3.804 0.7707
can be seen, the solid temperature fluctuates around the
N 0.19 3.559 ∼
desired value (400 K) under the action of the thermostat. For
O 0.102 3.188 −0.83
cases with different nanopore configurations, the water
H 0 0 0.415
temperatures all stabilize (with oscillation) at a roughly
C 0.0859 3.3997 0
constant level (373 K) and there exists one temperature
jump between the nanopore membrane and water due to the
Kapitza resistance (27 K). Moreover, as shown in Figure 2b,
The velocity-Verlet integration method was employed with a time there is no obvious difference in water temperature among the
step of 2.0 fs. The periodic boundary conditions were applied to the three nanopore configurations.
lateral direction (the x and y directions) of the system. For each Figure 3a presents the visualized snapshots of water
simulation case, first, the velocities of all water molecules were molecules during the evaporation process. As it can be seen,
generated at the prescribed temperature (300 K) following the the liquid−vapor interface is always above the Si3N4 nanopore
Gaussian distribution, and the arrangement of each atom was adjusted during the whole evaporation process. The evaporation rate
by minimizing the total energy of the system. Then, under the NVT can be measured by counting the number of water molecules
ensemble, all atoms were equilibrated to maintain the system (N) removed from the system. As shown in Figure 3b, it can
temperature at 300 K using the Nose−Hoover thermostat with a
damping coefficient of 0.2 ps−1, and this process lasted for 1 ns.
Finally, the evaporation process was realized by heating the liquid. It
has been demonstrated that the method of exerting the thermostat on
water molecules may not be physically reasonable.31 To solve this
problem, the Nose−Hoover thermostat (target temperature 400 K)
was exerted on the Si and N atoms in the heating region, which
satisfies 2 x 2 + y 2 > 5.4 nm and abs(z) < 1 nm, while the NVE
integration was performed to update the position and velocity for the
other atoms. By thermal transport at the solid−fluid interface, the thin
liquid film is heated and the evaporation process takes place. This step
was carried out for 5 ns, and the trajectories of atoms of the last 4 ns
were collected for the follow-up analysis. To obtain the meaningful
statistics, the properties in following results were calculated by
averaging over five independent runs; for each of them, different
random velocity distributions were initially assigned to water
molecules. During the entire process, a vacuum boundary over the
liquid−vapor interface was defined, where the water molecules will be
removed from the system once arriving at it. This nonequilibrium
condition was used in the previous research studies19,32,33 in which
the vapor molecules run into an infinite vacuum. Figure 1c shows the
water density profile along the z-direction at the beginning of the
evaporation process. The density is calculated by averaging the water
molecules within a 1-Å-thick slab over 0.2 ns. The liquid−vapor
interface (Z = 50 Å ) was defined at the location with a half density of Figure 3. (a) Snapshots of water molecules for case of Si-1.85. (b)
saturated water (0.48 g/cm3). We can observe that the mass density Number of water molecules evaporating from the thin liquid film
of water molecules is very low (<0.0001 g/cm3) as Z > 55 Å , and the during the evaporation process for the case of Si-1.85. (c) Evaporation
distance between the liquid−vapor interface and evaporation region flux (J) of the thin liquid film on nanopores with various pore
(40 Å ) is so far that it does not affect the evaporation performance. diameters (d).

C https://doi.org/10.1021/acs.langmuir.1c01212
Langmuir XXXX, XXX, XXX−XXX
Langmuir pubs.acs.org/Langmuir Article

be found that each evaporation curve shows a linear evolution


pattern (water evaporates at a constant rate), which suggests
that under the present configuration of the simulated system,
the finite size effect of the periodic computational box can be
negligible34 and the present model is capable of reproducing
the evaporation performance in the correct manner. The slope
of the curve (dN/dτ) was taken to describe the evaporation
flux J, expressed in kilomoles per square meter per seconds
(kmol/m2·s), using the following equation (eq 2):
1 dN
J=
NAA dτ (2)
where NA is the Avogadro constant and A is the liquid/vapor
evaporation interface (cross-section of the thin water film Figure 4. (a) Mean squared displacement (MSD) of interface water
along the z-axis, 6.34 × 6.1 nm2). The calculated evaporation molecules. (b) Diffusion coefficient (D) of interface water molecules
flux was on the order of 6 kmol/m2·s, which is two times as a function of pore diameter (d).
higher than the capillary evaporation inside the graphene
channel (about 2.3 kmol/m2·s, 363 K)19 and the carbon or N-terminated one. The similar variation of the diffusion
nanotube (about 3 kmol/m2·s, 400 K).20 The more intense coefficient and evaporation flux confirms that the elevated
evaporation phenomenon can be attributed to the hydrophilic collision frequency of water molecules near the interface can
nature of the silicon nitride surface. Our previous work has lead to an enhanced evaporation flux.
demonstrated that the boiling heat transfer performance can be To better understand the dynamics properties of water
improved by increasing the hydrophilicity of the solid molecules, we investigated the effect of nanopore configuration
surface.35 Therefore, the hydrophilic nature of the nanopore on the water structure. The density distributions of water on
in the present work enhances the rate of water evaporation. the top of the nanopore are shown in Figure 5. It is observed
The influence of pore size was investigated at a water
temperature of 373 K. Figure 3c shows the evaporation flux of
the thin liquid film on nanopores with various pore diameters.
For the nanopores with the same terminal atoms, the
evaporation rate of water decreases with the decreasing pore
size. Moreover, it can be observed that although the pore size
plays a significant role in the evaporation performance, the
pore chemistry is the main factor that affects the evaporation
characteristics of the thin liquid film. For example, evaporation
flux for the case of Si-0.89 reaches 5.28 kmol/m2·s, while
evaporation flux for that of N-2.13 is only 5.1 kmol/m2·s. This
means that Si-terminated nanopores can enhance the water
evaporation performance compared to the N-terminated
(more than 20%) or mix-terminated ones. This important
finding contradicts with the abovementioned experimental
perception that the evaporation flux increased monotonically
as the pore diameter decreased (pore diameter >27 nm).
As aforementioned, there is no obvious difference in water
temperature among the three nanopore configurations, which
indicates that the heat transfer from the nanopore to water is
not the dominant factor in the evaporation process. In general,
the water molecules at the liquid−vapor interface are the Figure 5. Water density profiles on top of nanopores with equivalent
candidates for the evaporation process. After colliding with pore sizes: (a) Si-1.41, Mix-1.33, N-1.13; (b) Si-1.85, Mix-1.78, N-
1.65; (c) Si-2.33, Mix-2.26, N-2.13; and (d) Si-2.89, Mix-2.78, N-2.63.
neighbor molecules and then obtaining enough kinetic energy,
the water molecule is able to break the intermolecular bond
from the neighbor water molecules and then vaporizes.36 The that the water mass densities on the Si-terminated nanopores
mean squared displacement (MSD), which is the deviation of are always lower than those on the N-terminated and mix-
the position of a molecule with respect to a reference position terminated ones under the approximately equivalent accessible
over time, can be correlated with the mobility of molecules. To pore area, which implies that the water structure of the liquid
further investigate the collision frequency of water molecules, thin film on the Si-terminated nanopore is looser than that on
which dominates the evaporation process, the MSD profile of the N-terminated or mix-terminated ones. On nanopores with
water molecules near the liquid−vapor interface (40 Å < Z < the same terminal atoms, the water density increases with the
60 Å) over 200 ps was computed (Figure 4a) and the self- reduction of pore size. Figure 6 depicts the probability
diffusion coefficient (D) can be calculated from the slope of distribution of water dipole orientation on the top of the
MSD with respect to the time (dMSD/dτ) as depicted in nanopore, P(θ), in which the orientation θ is defined as the
Figure 4b. The result shows that under the conditions of angle between the dipole moment of the water molecule
equivalent pore size, the water molecules on the Si-terminated (defined from the center of two hydrogen atoms to the oxygen
nanopore are more mobile compared with the mix-terminated atom) and the positive z-axis direction as shown in the inset of
D https://doi.org/10.1021/acs.langmuir.1c01212
Langmuir XXXX, XXX, XXX−XXX
Langmuir pubs.acs.org/Langmuir Article

https://pubs.acs.org/10.1021/acs.langmuir.1c01212

Notes
The authors declare no competing financial interest.

■ ACKNOWLEDGMENTS
This work was supported by the National Natural Science
Foundation of China through grant no. 51676124 and
Shanghai International Science and Technology Cooperation
Project through grant no. 18160743900.

■ REFERENCES
(1) Surwade, S. P.; Smirnov, S. N.; Vlassiouk, I. V.; Unocic, R. R.;
Veith, G. M.; Dai, S.; Mahurin, S. M. Water Desalination Using
Nanoporous Single-Layer Graphene. Nat. Nanotechnol. 2015, 10,
459−464.
(2) He, S. M.; Chen, C. J.; Kuang, Y. D.; Mi, R. Y.; Liu, Y.; Pei, Y.;
Figure 6. Probability distribution of water dipole orientation on top Kong, W. Q.; Gan, W. T.; Xie, H.; Hitz, E.; Jia, C.; Chen, X.; Gong,
of nanopores with equivalent pore sizes: (a) Si-1.41, Mix-1.33, N- A.; Liao, J. M.; Li, J.; Ren, Z. J.; Yang, B.; Das, S.; Hu, L. B. Nature-
1.13; (b) Si-1.85, Mix-1.78, N-1.65; (c) Si-2.33, Mix-2.26, N-2.13; Inspired Salt Resistant Bimodal Porous Solar Evaporator for Efficient
and (d) Si-2.89, Mix-2.78, N-2.63. and Stable Water Desalination. Energy Environ. Sci. 2019, 12, 1558−
1567.
(3) Hu, H. W.; Li, Q.; Liu, S.; Fang, T. Y. Molecular Dynamics
Figure 6. On the Si-terminated nanopore, the P(θ) has the Study on Water Vapor Condensation and Infiltration Characteristics
maximum value when θ is close to 90°, which indicates that in Nanopores with Tunable Wettability. Appl. Surf. Sci. 2019, 494,
water molecules tend to be parallel to the nanopore surface 249−258.
and the water structure is disordered. However, on the N- (4) Hanks, D. F.; Lu, Z. M.; Sircar, J.; Kinefuchi, I.; Bagnall, K. R.;
terminated and mix-terminated nanopores, there is a maximum Salamon, T. R.; Antao, D. S.; Barabadi, B.; Wang, E. N. High Heat
of P(θ) as θ < 90°, which means that water molecules prefer to Flux Evaporation of Low Surface Tension Liquids from Nanoporous
Membranes. ACS Appl. Mater. Interfaces 2020, 12, 7232−7238.
point toward the nanopore and the water structure is ordered.
(5) Wang, Q. Y.; Chen, R. K. Widely Tunable Thin Film Boiling
These results suggest that the terminated Si atoms on the pore Heat Transfer through Nanoporous Membranes. Nano Energy 2018,
surface plays a primary role in the water structure. For the Si- 54, 297−303.
terminated nanopore, the structure of water is looser and more (6) Wang, Q. Y.; Chen, R. K. Ultrahigh Flux Thin Film Boiling Heat
disordered compared to the other nanopores. It is conducive to Transfer through Nanoporous Membranes. Nano Lett. 2018, 18,
the thermal motion of water molecules and then leads to an 3096−3103.
enhanced evaporation performance. (7) Poudel, S.; Zou, A.; Maroo, S. C. Evaporation Dynamics in

■ CONCLUSIONS
In this work, we explored the ultrathin water film evaporation
Buried Nanochannels with Micropores. Langmuir 2020, 36, 7801−
7807.
(8) Tao, P.; Ni, G.; Song, C. Y.; Shang, W.; Wu, J. B.; Zhu, J.; Chen,
G.; Deng, T. Solar-Driven Interfacial Evaporation. Nat. Energy 2018,
on the silicon nitride nanopores. Through MD simulations, we 3, 1031−1041.
confirm that the pore chemistry plays an important role in (9) Ito, Y.; Tanabe, Y.; Han, J. H.; Fujita, T.; Tanigaki, K.; Chen, M.
evaporation performance and that Si-terminated nanopores W. Multifunctional Porous Graphene for High-Efficiency Steam
show the best among all possible nanopore configurations. The Generation by Heat Localization. Adv. Mater. 2015, 27, 4302−4307.
terminated Si atoms on the nanopore wall surface can realize a (10) Xue, G. B.; Xu, Y.; Ding, T. P.; Li, J.; Yin, J.; Fei, W. W.; Cao, Y.
disordered and loose water structure and induce a more active Z.; Yu, J.; Yuan, L. Y.; Gong, L.; Chen, J.; Deng, S. Z.; Zhou, J.; Guo,
thermal motion of water molecules, which gives rise to a higher W. L. Water-Evaporation-Induced Electricity with Nanostructured
evaporation rate compared to N-terminated and mix- Carbon Materials. Nat. Nanotechnol. 2017, 12, 317−321.
terminated ones. The findings in the present work reveal the (11) Lao, J. C.; Wu, S.; Gao, J.; Dong, A. P.; Li, G. J.; Luo, J. Y.
Electricity Generation Based on a Photothermally Driven Ti3C2Tx
mechanism behind the enhancement of nanopore-based MXene Nanofluidic Water Pump. Nano Energy 2020, 70, No. 104481.
evaporation and provide new guidance for the design of (12) Nazari, M.; Masoudi, A.; Jafari, P.; Irajizad, P.; Kashyap, V.;
innovative nanopore membrane materials.


Ghasemi, H. Ultrahigh Evaporative Heat Fluxes in Nanoconfined
Geometries. Langmuir 2019, 35, 78−85.
AUTHOR INFORMATION (13) Lu, Z. M.; Wilke, K. L.; Preston, D. J.; Kinefuchi, I.; Chang-
Corresponding Author Davidson, E.; Wang, E. N. An Ultrathin Nanoporous Membrane
Zhenyu Liu − School of Mechanical Engineering, Shanghai Evaporator. Nano Lett. 2017, 17, 6217−6220.
Jiao Tong University, Shanghai 200240, China; (14) Lu, Z. M.; Kinefuchi, I.; Wilke, K. L.; Vaartstra, G.; Wang, E. N.
orcid.org/0000-0002-9214-9798; Email: zhenyu.liu@ A Unified Relationship for Evaporation Kinetics at Low Mach
Numbers. Nat. Commun. 2019, 10, 2368.
sjtu.edu.cn
(15) Li, Y. X.; Chen, H. W.; Xiao, S. Y.; Alibakhshi, M. A.; Lo, C. W.;
Author Lu, M. C.; Duan, C. H. Ultrafast Diameter-Dependent Water
Runkeng Liu − School of Mechanical Engineering, Shanghai Evaporation from Nanopores. ACS Nano 2019, 13, 3363−3372.
(16) Dai, X. M.; Yang, F. H.; Yang, R. G.; Lee, Y. C.; Li, C.
Jiao Tong University, Shanghai 200240, China Micromembrane-Enhanced Capillary Evaporation. Int. J. Heat Mass
Complete contact information is available at: Transfer 2013, 64, 1101−1108.

E https://doi.org/10.1021/acs.langmuir.1c01212
Langmuir XXXX, XXX, XXX−XXX
Langmuir pubs.acs.org/Langmuir Article

(17) Shi, W.; Vieitez, J. R.; Berrier, A. S.; Roseveare, M. W.;


Surinach, D. A.; Srijanto, B. R.; Collier, C. P.; Boreyko, J. B. Self-
Stabilizing Transpiration in Synthetic Leaves. ACS Appl. Mater.
Interfaces 2019, 11, 13768−13776.
(18) Akkus, Y.; Koklu, A.; Beskok, A. Atomic Scale Interfacial
Transport at an Extended Evaporating Meniscus. Langmuir 2019, 35,
4491−4497.
(19) Kieu, H. T.; Liu, B.; Zhang, H.; Zhou, K.; Law, A. W. K.
Molecular Dynamics Study of Water Evaporation Enhancement
through a Capillary Graphene Bilayer with Tunable Hydrophilicity.
Appl. Surf. Sci. 2018, 452, 372−380.
(20) Li, J.; Gao, S.; Long, R.; Liu, W.; Liu, Z. C. Self-Pumped
Evaporation for Ultra-Fast Water Desalination and Power Generation.
Nano Energy 2019, 65, No. 104059.
(21) Walther, J. H.; Ritos, K.; Cruz-Chu, E. R.; Megaridis, C. M.;
Koumoutsakos, P. Barriers to Superfast Water Transport in Carbon
Nanotube Membranes. Nano Lett. 2013, 13, 1910−1914.
(22) Heiranian, M.; Farimani, A. B.; Aluru, N. R. Water Desalination
with a Single-Layer MoS2 Nanopore. Nat. Commun. 2015, 6, 8616.
(23) Shen, J. W.; Li, J. C.; Liu, F.; Zhang, L.; Liang, L. J.; Wang, H.
B.; Wu, J. Y. A Molecular Dynamics Study on Water Desalination
Using Single-Layer MoSe2 Nanopore. J. Membr. Sci. 2020, 595,
No. 117611.
(24) Ang, E. Y. M.; Ng, T. Y.; Yeo, J.; Lin, R. M.; Liu, Z. S.;
Geethalakshmi, K. R. Investigations on Different Two-Dimensional
Materials as Slit Membranes for Enhanced Desalination. J. Membr. Sci.
2020, 598, No. 117653.
(25) Plimpton, S. Fast Parallel Algorithms for Short-Range
Molecular Dynamics. J. Comput. Phys. 1995, 117, 1−19.
(26) Azamat, J.; Khataee, A. Removal of nitrate ion from water using
boron nitride nanotubes: Insights from molecular dynamics
simulations. Comput. Theor. Chem. 2016, 1098, 56−62.
(27) Price, D. J.; Brooks, C. L., III A Modified TIP3P Water
Potential for Simulation with Ewald Summation. J. Chem. Phys. 2004,
121, 10096−10103.
(28) Comer, J.; Dimitrov, V.; Zhao, Q.; Timp, G.; Aksimentiev, A.
Microscopic Mechanics of Hairpin DNA Translocation through
Synthetic Nanopores. Biophys. J. 2009, 96, 593−608.
(29) Ho, C.; Qiao, R.; Heng, J. B.; Chatterjee, A.; Timp, R. J.; Aluru,
N. R.; Timp, G. Electrolytic Transport through a Synthetic
Nanometer-Diameter Pore. Proc. Natl. Acad. Sci. U. S. A. 2005, 102,
10445−10450.
(30) Hilder, T. A.; Gordon, D.; Chung, S. H. Salt Rejection and
Water Transport through Boron Nitride Nanotubes. Small 2009, 5,
2183−2190.
(31) Zhao, K. W.; Wu, H. Y. Fast Water Thermo-Pumping Flow
across Nanotube Membranes for Desalination. Nano Lett. 2015, 15,
3664−3668.
(32) Srinivasan, S.; Das, S. R.; Balasubramanian, G. Transient
Evaporation of Water Thin Film over Nanostructured Graphene.
Appl. Surf. Sci. 2019, 495, No. 143545.
(33) Zhang, H.; Liu, B.; Kieu, H. T.; Wu, M. S.; Zhou, K.; Law, A.
W. K. Coarse-Grained Molecular Dynamics Study of Membrane
Distillation through Meso-Size Graphene Channels. J. Membr. Sci.
2018, 558, 34−44.
(34) Cohen-Tanugi, D.; Grossman, J. C. Water Desalination across
Nanoporous Graphene. Nano Lett. 2012, 12, 3602−3608.
(35) Liu, R. K.; Liu, Z. Y. Study of Boiling Heat Transfer on
Concave Hemispherical Nanostructure Surface with MD Simulation.
Int. J. Heat Mass Transfer 2019, 143, No. 118534.
(36) Nagata, Y.; Usui, K.; Bonn, M. Molecular Mechanism of Water
Evaporation. Phys. Rev. Lett. 2015, 115, No. 236102.

F https://doi.org/10.1021/acs.langmuir.1c01212
Langmuir XXXX, XXX, XXX−XXX

You might also like