Slip of Alkanes Confined between Surfactant Monolayers Adsorbed on

Mar 14, 2018 - (36) Wang and Zhao(37) extended Eyring's molecular kinetic theory (MKT)(9) to describe the slip behavior. By assigning appropriate valu...
0 downloads 3 Views 2MB Size
Subscriber access provided by Kaohsiung Medical University

Interfaces: Adsorption, Reactions, Films, Forces, Measurement Techniques, Charge Transfer, Electrochemistry, Electrocatalysis, Energy Production and Storage

Slip of Alkanes Confined between Surfactant Monolayers Adsorbed on Solid Surfaces James Patrick Ewen, Sridhar Kumar Kannam, Billy D. Todd, and Daniele Dini Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.8b00189 • Publication Date (Web): 14 Mar 2018 Downloaded from http://pubs.acs.org on March 14, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Slip of Alkanes Confined between Surfactant Monolayers Adsorbed on Solid Surfaces James P. Ewen 1 * +, Sridhar Kumar Kannam 1 2 +, B. D. Todd 2, D. Dini 1 1 – Department of Mechanical Engineering, Imperial College London, London, SW7 2AZ, UK 2 – Department of Mathematics, Swinburne University of Technology, Melbourne, VIC 3122, Australia * – Corresponding author: [email protected] + – These authors contributed equally to this work

Keywords Nonequilibrium molecular dynamics, additives, lubricants, organic friction modifiers, hydrodynamic

1 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 33

Abstract The slip and friction behaviour of n-hexadecane, confined between organic friction modifier (OFM) surfactant films adsorbed on hematite surfaces, have been studied using nonequilibrium molecular dynamics (NEMD) simulations. The influence of OFM type and coverage, as well as the applied shear rate and pressure have been investigated. A measurable slip length is only observed for OFM films with a high surface coverage, which provide smooth interfaces between well-defined OFM and hexadecane layers. Slip commences above a critical shear rate, beyond which the slip length first increases with increasing shear rate and then asymptotes towards a constant value. The maximum slip length increases significantly with increasing pressure. Systems and conditions which show a larger slip length typically give a lower friction coefficient. Generally, the friction coefficient increases linearly with logarithmic shear rate; however, it shows a much stronger shear rate dependency at low pressure than at high pressure. Relating slip and friction, slip only occurs above a critical shear stress, after which the slip length first increases linearly with increasing shear stress and then asymptotes. This behaviour is well described using the extended molecular kinetic theory (MKT) slip model. This study provides a more detailed understanding of the slip of alkanes on OFM monolayers. It also suggests that high coverage OFM films can significantly reduce friction by promoting slip, even when the surfaces are well-separated by a lubricant.

2 ACS Paragon Plus Environment

Page 3 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Introduction The dynamics of flow in nanofluidic devices1 and nanoelectromechanical systems (NEMS)2 can only be accurately described through a detailed understanding of the flow at fluid-solid interfaces.3 In this context, recent experimental and molecular simulation studies have shown that the common assumption in continuum hydrodynamics that fluid adjacent to a sliding surface moves at the same velocity as the surface (no-slip boundary condition) often breaks down at the nanoscale.3 A large proportion of such studies have investigated the slip of alkanes on surfactant monolayers adsorbed on smooth solid surfaces. Many of these studies report a slip length, defined as an extrapolated distance relative to the fluid-solid interface where the tangential velocity component vanishes.3 However, there is significant variation in both the magnitude and the shear rate dependence of the slip lengths from these experiments.3 For example, total internal reflection-fluorescence recovery after photobleaching (TIR-FRAP) experiments suggested a shear rate-independent (measured between 102 - 103 s-1) slip length of approximately 400 nm for n-hexadecane on smooth sapphire surfaces coated with octadecyltrichlorosilane (OTS) self-assembled monolayers (SAMs).4 In surface forces apparatus (SFA) experiments using tertradecane between smooth mica surfaces coated with OTS SAMs, slip was only detected once a critical shear rate was exceeded, above which a linear increase in slip length with log(shear rate) from 0 - 1.4 µm was observed (between 102 - 105 s-1).5 More recently, colloidal probe atomic force microscopy (AFM) has been used to measure the slip length of several n-alkanes (from heptane to hexadecane) on OTS SAM-coated, ultra-smooth silicon wafers.6 They suggested much smaller, shear rate-independent (between 102 - 103 s-1) slip lengths of between 10 - 30 nm, with longer n-alkanes with higher viscosity giving more slip.6 In addition to surfaces coated with surfactant SAMs, slip of alkanes has also been observed on surfactant films formed from solution. For example, TIR-FRAP experiments also showed that the addition of a surfactant (stearic acid) to hexadecane confined between smooth sapphire surfaces led to an increase in slip length from 150 - 350 nm.4 Moreover, adding hexadecylamine to n-alkanes confined between smooth mica surfaces in SFA experiments switched the boundary conditions from no-slip to 3 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 33

partial slip.7 In these experiments, the maximum slip length again increased with increasing n-alkane chain length from octane to tetradecane (1 - 15 nm).7 The large variation between slip length measurements from different experiments on similar systems suggests that different mechanisms may be acting to promote slip.3,6 Experiments which suggested smaller (< 50 nm) slip lengths for alkanes on adsorbed surfactant monolayers6,7 are more consistent with the slip mechanism devised by Tolstoi and reviewed by Blake.8 This mechanism suggests that slip is due to enhanced liquid mobility at the surface and is essentially an extension of Eyring’s activated flow model for bulk liquid viscosity.9 Conversely, the much larger slip lengths measured in some experiments4,5 have subsequently been explained through the formation of gas/vapour films or nanobubbles at the interface,3,6 as initially proposed by de Gennes.10 In addition to experiments, nonequilibrium molecular dynamics (NEMD) simulations have been used extensively to study slip of liquids confined between solid surfaces. For high-slip systems, such as water flowing on graphene surfaces11 or through carbon nanotubes,12 accurate determination of the slip length with NEMD has been shown to be less accurate than equilibrium molecular dynamics (EMD) methods. However, for the partial-slip systems of interest in this study, NEMD simulations have been used extensively to further understand slip phenomena13–16 as well as to estimate slip lengths for a range of confined alkane systems.17–21 Most NEMD studies have investigated slip behaviour in highly confined (< 10 molecular layers) alkane films between atomically-smooth solid surfaces. The slip lengths from these NEMD simulations depend on the system and conditions, but are typically < 10 nm. In these studies, the slip length generally increased with alkane viscosity, slab stiffness, sliding velocity, and pressure.17–21 NEMD simulations of highly confined systems have also shown that the slip length decreases with increasing film thickness20,22 and is drastically reduced when the confining surfaces contain atomic-scale roughness.23,24 SFA and TIR-FRAP experiments have also shown significant reductions in slip length on rougher surfaces.25,26 The slip length of water confined between atomically-smooth surfaces coated by alkylsilane SAMs27,28 and adsorbed

4 ACS Paragon Plus Environment

Page 5 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

methanol films29,30 has also been measured with NEMD simulations. Several NEMD simulation studies31–33 have probed the structure, flow and friction behaviour of alkanes confined between surfactant monolayers; however, slip lengths in such systems have not yet been successfully quantified. NEMD simulations have also been central in developing more accurate models to describe liquid slip between solid surfaces.34 Thompson and Troian14 developed an equation for slip where, above a critical shear rate, the slip length is a power law function of the shear rate. Conversely, Lichter et al.35 suggested the variable-density Frenkel-Kontorova (vdFK) model, which predicts a plateau of the slip length at high shear rate. This predicts similar behaviour to the model developed by Spikes and Granick to describe data from SFA experiments; in this model the slip length asymptotes at high shear stress rather high than strain rate.36 Wang and Zhao37 extended Eyring’s molecular kinetic theory (MKT) to describe slip behaviour. By assigning appropriate values of the tuneable parameters, the extended MKT model37 is able to describe the important features from the slip models due to Thompson and Troian,14 Lichter et al. (vdFK),35 and Spikes and Granick.36 In addition to accelerating flow in nanofluidic devices and NEMS, it has been proposed that slip could be exploited to reduce friction in macroscopic tribological systems.38 Specifically, slip has been used to rationalise results from macroscopic tribology experiments which showed that the addition of an organic friction modifier (OFM) to n-hexadecane could significantly reduce friction in the hydrodynamic lubrication regime when the surfaces are well-separated.38 OFMs are amphiphilic surfactant molecules that contain nonpolar aliphatic tailgroups attached to polar headgroups. OFMs are based solely on C, H, O, and N atoms, are not environmentally harmful, and do not poison exhaust after-treatment devices. The most widely studied OFM headgroups are carboxylic acids; however, amines, amides and glycerides are more industrially relevant. Commercial OFMs generally contain unbranched aliphatic tailgroups containing 12−20 carbon atoms because of their effective friction reduction, high base oil solubility, and high availability from natural fats and oils.39 OFMs

5 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 33

adsorb to metal, ceramic, or carbon-based surfaces through their polar headgroups and strong, cumulative van der Waals forces between proximal nonpolar tailgroups leads to the formation of incompressible monolayers.39 These monolayers are known to significantly reduce friction and wear in the boundary lubrication regime (where the load is primarily supported by surface asperities) by preventing direct contact between solid surfaces.39 However, it is less clear if OFMs also reduce friction in the hydrodynamic lubrication regime (where the load is supported by the lubricant). When OFM monolayers form on surfaces inside a tribological contact,39 the planes of methyl groups at the end of the vertically adsorbed OFM molecules create non-wetting, oleophobic surfaces40 over which lubricants could slip.38,39 In this study, large-scale NEMD simulations will be used to investigate the friction and slip behaviour of hexadecane, confined between OFM films, under hydrodynamic lubrication conditions. Several different OFM types (stearic acid, stearamide, glycerol mono-stearate) and coverages (1.44 - 4.32 molecules nm-2) will be explored in order to understand their influence on the friction and slip behaviour. The dependence of the slip length and the friction coefficient on the applied shear rate (108 - 1010 s-1) and pressure (0.1 - 1.0 GPa) will also be investigated. This study provides unique insights into the slip of alkanes on surfactant monolayers, as well as the relationship between slip and friction. These results also provide support for previously proposed mechanisms and models for slip in such systems. A deeper understanding of the slip phenomenon expected to be valuable not only for improving molecular design of lubricant additives but also for understanding and controlling flow in nanofluidic devices1 and NEMS.2

6 ACS Paragon Plus Environment

Page 7 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Methodology System Setup A representative example of the systems simulated in this study is shown in Figure 1a. It consists of a lubricant (n-hexadecane) layer confined between two OFM monolayers which are adsorbed on atomically-smooth hematite slabs. The hexadecane layer is sufficiently thick (> 15 molecular diameters) such that no lubricant structuring was evident in the middle of the film32 and thus any confinement-induced viscosity increase was expected to be negligible.41 All systems were constructed using the Materials and Processes Simulations (MAPS) platform from Scienomics SARL. Classical MD simulations were performed using the large-scale atomic/molecular massively parallel simulator (LAMMPS) code.42 Figure 1b shows the three different OFM types considered in these simulations; stearic acid (SA), stearamide (SAm), and glycerol mono-stearate (GMS). SA was selected to allow comparisons to previous experiments and NEMD simulations, whereas SAm and GMS are more commercially relevant.39 The hexadecane lubricant was chosen because of its well-defined properties and common use in both experimental38,43 and modeling33 tribology studies.

7 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 33

Figure 1. Simulation details: setup for compression and sliding simulations (a). Example shown for SA at 4.32 nm-2 after compression, before sliding. Fe atoms are shown in pink, O in red, terminal C in yellow, and the other all C in cyan. Headgroup H are shown in white while H atoms in the alkyl groups are omitted for clarity. Periodic boundary conditions (yellow dotted line) are applied in the x and y directions. Rendered using VMD.44 Chemical structures of OFMs simulated in this study (b): stearic acid (SA), stearic amide (SAm), and glycerol monostearate (GMS).

In all of the MD simulations, (100) slabs of α-iron(III) oxide45 (hematite) with dimensions (xyz) of approximately 55 × 55 × 12 Å were used as the substrates. Periodic boundary conditions were applied in the x and y directions.46 The surface was cleaved such that the Fe/O ratio remained at 2:3 in order to ensure that surfaces with no overall charge were produced.33 Three different OFM surface coverages were considered. The coverage, Γ, is defined as the number of OFM molecules in a given surface area (nm-2). The limiting headgroup area for acids, amides and glycerides (assuming monodentate binding) is around 0.22 nm2, which corresponds to a maximum theoretical coverage of 4.55 nm−2 for the OFMs.47 In these simulations, a high surface coverage (Γ = 4.32 nm−2) is simulated by adsorbing 132 OFM molecules on each ≈ 30 nm2 slab to form a closepacked monolayer. Two other surface coverages are considered: a medium coverage (Γ = 2.88 nm−2), 8 ACS Paragon Plus Environment

Page 9 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

approximately 2/3 of the maximum coverage; and a low coverage (Γ = 1.44 nm−2), around 1/3 of the maximum coverage. In situ-AFM experiments48 and depletion isotherms49 have shown that OFMs with saturated tailgroups (e.g. SA, SAm, and GMS) can form high coverage monolayers on solid surfaces. Multilayer OFM films have also been observed in recent in situ-AFM experiments50 although their role in lubrication remains unclear, and they are not considered in this current study. The OFM molecules were oriented perpendicular to, and initially 3 Å above, the interior surfaces of the two hematite slabs (Figure 1a). This produced OFM films similar to those formed by LangmuirBlodgett experiments.47,51 Between 450 (high coverage) and 650 (low coverage) hexadecane molecules were then randomly distributed in the region between the OFM films. This resulted in a similar total number of atoms (approximately 50,000 including the surface atoms) and film thickness for all of the coverages studied.32 Parameters from the long hydrocarbon-optimized potentials for liquid simulations, all-atom force field (L-OPLS-AA)52 were used to represent the carbon and hydrogen atoms in the alkyl chains (both OFMs and hexadecane) and standard OPLS-AA53,54 parameters were used for the headgroup atoms (see ref.33 for full details). Accurate density and viscosity prediction of bulk hexadecane using L-OPLSAA under ambient and high pressure conditions has been confirmed in a previous publication.55 Lennard-Jones interactions were cut off at 12 Å, and ‘unlike’ interactions were evaluated using the geometric mean mixing rules, as prescribed in the OPLS force field.53 Electrostatic interactions were evaluated using a slab implementation of the particle-particle, particle-mesh (PPPM) algorithm,56 with a relative force accuracy of 10-5. Surface−hexadecane and surface−OFM interacƟons were represented by the Lennard-Jones and Coulomb potentials; the hematite surface parameters selected were developed by Berro et al.57 for alkane adsorption. The hematite slab atoms were ‘frozen’ in the corundum crystal structure45 to facilitate more accurate slip length analysis. Previous NEMD simulations have shown that the use of rigid slabs can lead to unphysical velocity-slip behaviour when they are in direct contact with a non-

9 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 33

wetting fluid;18 however, here the flexible OFM films form the interface with the lubricant and are sufficient to prevent such behaviour (see Results and Discussion). The MD equations of motion were integrated using the velocity Verlet algorithm with an integration time step of 1.0 fs. Fast-moving bonds involving hydrogen atoms were constrained with the SHAKE algorithm.58 The Nosé-Hoover thermostat,59,60 with a time relaxation constant of 0.1 ps, was used in order to maintain the target temperature, T = 300 K. The pressure (P = 0.1 - 1.0 GPa) was controlled by applying a constant normal force to the outermost layer of atoms in the upper slab, keeping the z coordinates of the outermost layer of atoms in the lower slab fixed (Figure 1a).34,61

Simulation Procedure First, a density similar to that of liquid hexadecane (0.75 g cm−3) was achieved between the slabs by moving the top slab down at 10 m s−1 and then the system was energy minimized. The system was then pressurized (P = 0.1 - 1.0 GPa), thermostatted in directions perpendicular to the compression (x and y), and allowed to equilibrate at 300 K. Initially, the slab separation varied in a damped harmonic manner, so sliding was not applied until a constant average slab separation was obtained and the average hydrostatic pressure within the hexadecane film was close to its target value (within 1%).32 These compression simulations were generally around 2 ns in duration. After compressive oscillation became negligible, a velocity, vs, was added in the x direction to the outermost layer of atoms in the top slab (Figure 1a), and sliding simulations were conducted for between 2 - 20 ns (depending on vs). Lower sliding velocities required longer simulations for the block-averaged x-velocity profile and friction coefficient to reach a steady state. The values of vs applied were between 1 - 200 m s−1. Accurate determination of the friction coefficient and particularly the slip length becomes challenging at even lower sliding velocities, while shear heating becomes more difficult to control above the selected range.34 A combination of high sliding velocity (shear rate) and pressure are particularly relevant for components which primarily operate in the elastohydrodynamic lubrication (EHL) regime.62 10 ACS Paragon Plus Environment

Page 11 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

A consequence of the rigid slabs is that the thermostat cannot be applied directly to the slab atoms, as is common in confined NEMD simulations.33 Applying the thermostat directly to fluid molecules confined between rigid surfaces has been shown artificially influence their behaviour under sliding conditions,63 and also lead to erroneous slip behaviour.18 Therefore, during the sliding simulations, any heat generated was dissipated using a thermostat acting only on the OFM layers (Figure 1a), applied perpendicular to the sliding direction (y and z). Using this approach, there was a negligible increase in the lubricant film temperature under shear, even at the highest sliding velocity considered. Moreover, the variation in the slip with sliding velocity is consistent with NEMD simulations of alkanes confined between thermostatted, flexible metal slabs (see Results and Discussion).18

11 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 33

Results and Discussion Slip Length Analysis The slip length analysis will be presented first, followed by the shear rate and pressure dependence of slip and friction, and finally the interdependence of slip and friction. Figure 2a shows a representative NEMD system snapshot (SAm, Γ = 4.32 nm-2). Note that, consistent with previous NEMD simulations,33 the tilted OFM monolayers move at the same velocity as the slabs to which they are adsorbed. This is the case all of the simulations and thus the OFM velocity profiles are omitted for clarity. Comparing the red (assumed no-slip) and blue (measured) hexadecane velocity profiles shows that the confined lubricant layer is only partially sheared21 and slip occurs at both of the OFM-hexadecane interfaces, not at the hematite-OFM interfaces (see also Figures 3-6).38,39

Figure 2. (a) Example system snapshot showing slip between OFM (SAm, Γ = 4.32 nm-2) films and confined n-hexadecane layer. Fe atoms are shown in pink, O in red, N in blue, terminal C in yellow, and the other all C in cyan. Headgroup H are shown in white while H atoms in the alkyl groups are omitted for clarity. Rendered using VMD.44 Horizontal red dotted line shows OFM-hexadecane interface. Solid red line shows linear velocity profile when no-slip boundary condition at the OFMhexadecane interface is assumed. Solid blue line represents the measured (slip) velocity profile. (b) Schematic showing the definition of the slip length (purple double headed arrows) and slip velocity (orange double headed arrows) in OFM-lubricated systems.11

12 ACS Paragon Plus Environment

Page 13 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 2b shows the definition of the slip length (purple arrows) from the NEMD simulations. The solid red line in Figure 2b shows the velocity profile that would be obtained within the hexadecane layer assuming no slip. In this case, the OFM layers move at the same velocity as the slabs to which they are adsorbed, the hexadecane layer is completely sheared, and there is a net zero velocity difference between the OFM and hexadecane layers at their interface. The solid blue line in Figure 2b shows a measured velocity profile with a non-zero net velocity difference between the OFM and hexadecane layers, indicating partial slip at the interface. Note that this represents apparent slip rather than true slip since the slip plane is located above an adsorbed layer rather than directly at the solid-liquid interface.3 The slip length (shown in purple) can be calculated by extrapolating the measured (slip) velocity profile to the point at which it intersects the applied slab velocity (in this case 0 and vs) and measuring the distance from the assumed no-slip case (see Figure 2b).4,11 Figures 3-6 show mass density profiles in the z-direction for the hexadecane (orange) and OFM (green) molecules. The measured hexadecane x-velocity profile in the z-direction (blue) and the assumed no-slip velocity profile (red) for a fully sheared hexadecane layer are also shown. Note that these figures are rotated 90⁰ relative to the schematics in Figure 2. The sharp, intense OFM mass density peaks on the far left- and right-hand sides of Figures 3-6 indicate the adsorption of the headgroups on the slabs while the less intense peaks which extend further from the surface are due to the tailgroups.33 Slip occurs when there is a net non-zero velocity difference between the hexadecane layer and the OFM layers at their interface.3 The interface where slip occurs (red vertical dotted lines) can be clearly identified by the intersection of the hexadecane and OFM mass density profiles. The centre of the mass density and velocity profiles have been shifted to zero for clarity. This does not affect the measured slip length since it is an average between the top and bottom slabs.

13 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 33

Figure 3. Effect of OFM surface coverage on the mass density profiles for the OFMs (green) and hexadecane (orange) and the measured velocity profile for hexadecane (blue). Representative example shown for SA; Γ = 1.44 nm-2 (a), 2.88 nm-2 (b), 4.32 nm-2 (c); P = 0.1 GPa; vs = 100 m s-1. Solid red line shows linear velocity profile when no-slip boundary condition at the OFM-hexadecane interface is assumed. Purple double headed arrow between red and blue vertical dotted lines shows the calculated slip length, which is only detectable in (c).

Figure 3 shows the effect of OFM surface coverage on the OFM and hexadecane mass density profiles as well as the hexadecane velocity profile. Comparing Figures 3a-c, the OFM films become substantially thicker and more strongly layered with increasing coverage. Both sum frequency

14 ACS Paragon Plus Environment

Page 15 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

generation (SFG) spectroscopy64 and radial distribution functions (RDFs) from NEMD simulations33 have also shown that the OFMs form solid-like films at high coverage. There is much less overlap of the OFM and hexadecane mass density profiles at high coverage (Figure 3c) than at medium (Figure 3b) and low coverage (Figure 3a), indicating reduced interdigitation.33 Consequentially, under all of the conditions studied, the velocity profiles only show a measurable slip length at high coverage (4.32 nm-2). In Figures 3a and 3b, the inferred no-slip velocity profile (red) overlaps with the measured velocity profile (blue). Conversely, in Figure 3c, there is clear separation between the measured profile and the no-slip profile, allowing the slip length (0.9 nm) to be calculated. This observation agrees with experimental results which have shown much larger slip lengths for high coverage OFM films compared to low coverage films,25 as well as those which have shown a transition from no-slip to partial slip when high coverage films were allowed to form on the surface.4,7

15 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 33

Figure 4. Effect of sliding velocity on the mass density profiles for the OFMs (green) and hexadecane (orange) and the measured velocity profile for hexadecane (blue). Representative example shown for SA; vs = 10 m s-1 (a), 20 m s-1 (b), 50 m s-1 (c); Γ = 4.32 nm-2; P = 0.1 GPa. Solid red line shows linear velocity profile when no-slip boundary condition at the OFM-hexadecane interface is assumed. Purple double headed arrow between red and blue vertical dotted lines shows the calculated slip length.

Figure 4 shows the effect of sliding velocity (or shear rate) on the mass density and velocity profiles. As with previous NEMD simulations,32,33 the mass density profiles are insensitive to sliding velocity in the range tested. The velocity profiles show an increase in slip length with increasing sliding velocity;

16 ACS Paragon Plus Environment

Page 17 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

from 0.3 nm at 10 m s-1 (Figure 4a), to 0.6 nm at 20 m s-1 (Figure 4b), to 0.8 nm at 50 m s-1 (Figure 4c), and 0.9 nm at 100 m s-1 (Figure 3c). This observation is consistent with several previous experiments5,7 and NEMD simulations which also showed a general increase in slip length with increasing sliding velocity.17–21

Figure 5. Effect of OFM type on the mass density profiles for the OFMs (green) and hexadecane (orange) and the measured velocity profile for hexadecane (blue) at low pressure (P = 0.1 GPa). Representative example shown for; SA (a) SAm (b), GMS (c); Γ = 4.32 nm-2; vs = 100 m s-1. Solid red line shows linear velocity profile when no-slip boundary condition at the OFM-hexadecane interface is assumed. Purple double headed arrow between red and blue vertical dotted lines shows the calculated slip length.

17 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 33

Figure 5 shows the effect of OFM type on the mass density and velocity profiles at low pressure (0.1 GPa). The OFM mass density profiles for SA (Figure 5a) and SAm (Figure 5b) are very similar but the GMS profile (Figure 5c) shows that its films are slightly thicker owing to the larger headgroup size. There is much less overlap of the hexadecane and OFM mass density profiles for SAm (Figure 5b) and GMS (Figure 5c) compared to SA (Figure 5a), indicating reduced interdigitation. At 0.1 GPa and 100 m s-1, this leads to larger slip lengths for high coverage SAm (1.1 nm) and GMS (1.3 nm) films compared to SA (0.9 nm).

18 ACS Paragon Plus Environment

Page 19 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

Figure 6. Effect of OFM type on the mass density profiles for the OFMs (green) and hexadecane (orange) and the measured velocity profile for hexadecane (blue) at high pressure (P = 1.0 GPa). Representative example shown for; SA (a), SAm (b), GMS (c); Γ = 4.32 nm-2; vs = 100 m s-1. Solid red line shows linear velocity profile when no-slip boundary condition at the OFM-hexadecane interface is assumed. Purple double headed arrow between red and blue vertical dotted lines shows the calculated slip length.

Figure 6 shows the effect of OFM type on the mass density and velocity profiles at high pressure (1.0 GPa). Similar to at low pressure, SAm (Figure 6b) and GMS (Figure 6c) films are less interdigitated than SA (Figure 6a) films leading to larger slip lengths. Comparing Figures 5 and 6 shows the effect of pressure on the structure and flow behaviour. For all of the OFMs, the mass density profiles show that there is a reduction in film thickness ≈ 10 % moving from low pressure (Figure 5) to high 19 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 33

pressure (Figure 6). The mass density profiles also indicate stronger layering both in the OFM and particularly the hexadecane films at high pressure relative to low pressure. The increased ‘first layer density’ observed here for the hexadecane layer at higher pressure has been correlated with larger slip lengths in previous NEMD simulations.19 Comparing Figures 5a and 6a suggests that, for SA, interdigitation between the hexadecane and OFM layers is reduced at higher pressure, owing to the denser OFM films. For all of the OFMs at 100 m s-1, the slip length increases significantly moving from low to high pressure; 0.9 nm (Figure 5a) to 2.6 nm (Figure 6a) for SA, 1.1 nm (Figure 5b) to 4.3 nm (Figure 6b) for SAm, and 1.3 nm (Figure 5c) to 4.8 nm (Figure 6c) for GMS. The increase in slip length with increasing pressure is consistent with previous NEMD simulations20 and tribology experiments which showed the same trend.65

Shear Rate and Pressure Dependence of Slip and Friction Figure 7a shows the change in slip length with log(shear rate) for high coverage (Γ = 4.32 nm-2) SA, SAm and GMS films at low (0.1 GPa) and high (1.0 GPa) pressure. Note that here the applied shear rate is calculated using the no-slip velocity profile and the hexadecane layer thickness rather than the overall film thickness (see Figure 2b). No slip occurs below a critical shear rate, above which the slip length increases linearly with log(shear rate) and then asymptotes towards a constant value. The critical shear rate decreases from ≈ 109 s-1 at low pressure, to ≈ 108 s-1 at high pressure. The critical shear rates from these simulations are relatively high with respect to those observed in previous experiments for similar systems3 as well as the shear rates experienced in real components.66 The slip length asymptotes to approximately 1 nm at low pressure and between 2-5 nm at high pressure, depending on the OFM type. SAm and GMS give consistently larger slip lengths than SA, particularly at high pressure, due to reduced OFM-hexadecane interdigitation (see Figure 5). The magnitudes of these slip lengths are somewhat lower than those measured for alkanes on adsorbed surfactant monolayers in AFM and SFA experiments (10 - 30 nm), performed under milder pressure and shear rate conditions.6,7 NEMD simulations of water slip on alkyl monolayers also underpredicted the slip length by a similar degree compared to experiment.27,28 The exact reason for this discrepancy 20 ACS Paragon Plus Environment

Page 21 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

remains unclear, but could be because the current simulations are representative of the slip mechanism proposed by Tolstoi and reviewed by Blake.8 Alternative slip mechanisms, such as the formation of multilayer OFM films50 or nanobubbles,10 could be the route of the larger slip lengths and lower critical shear rates observed in some SFA and TIR-FRAP experiments of similar systems.4,5 The asymptotic behaviour of the slip length at high shear rates is consistent with results from previous NEMD simulations17,18 and AFM experiments6 which have studied alkane slip. Such behaviour can be rationalised through a transition from ‘defect slip’ to ‘global slip’17 and captured using both the vdFK35 and extended MKT37 slip models. An increase in slip length with increasing pressure has also both been noted in previous NEMD simulations20 and tribology experiments.65 This observation suggests that, as the pressure is increased, the viscous friction between individual hexadecane molecules increases more than friction at the OFM-hexadecane interface.3 Both AFM6 and SFA7 experiments have shown that longer chain alkanes with a larger viscosity give larger slip lengths on surfactant monolayers. It would of interest in future studies to use NEMD to investigate the slip of larger alkanes with viscosities closer to real lubricants on surfactant monolayers.

Figure 7. Variation in the slip length (a) and friction coefficient (b) with log10(shear rate) for; SA (blue), SAm (red), GMS (green); 0.1 GPa (filled symbols) and 1.0 GPa (open symbols); Γ = 4.32 nm-2. Error bars in (b) show the standard deviation between block-averaged friction coefficient values.

21 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 33

The friction behaviour of the OFM films and its relation to slip was also investigated. The kinetic friction coefficient, μ, was obtained using the extended Amontons−Coulomb law under the high load approximation: FL/FN = F0/FN + μ ≃ μ. Here, FL and FN are respectively defined as the block-averaged lateral force (shear stress) and normal force acting on each hematite slab in response to the fluid during sliding, and F0 is the load-independent Derjaguin offset, representing adhesive surface forces. Previous NEMD simulations of the friction between OFM films separated by a lubricant layer have confirmed the validity of this approximation.32,33 In agreement with experimental results,4 the friction coefficient is greater at low (1.44 nm-2) and medium (2.88 nm-2) coverage, where no slip was observed (Figure 3a-b), compared at high coverage (4.32 nm-2), where a slip length could be measured (Figure 3c). The differences in friction between the high coverage case and the low and medium coverage cases are broadly similar to those observed in previous NEMD simulations of OFM films.33 Figure 7b shows the change in the friction coefficient with log(shear rate) for high coverage (Γ = 4.32 nm-2) SA, SAm, and GMS films at low (0.1 GPa) and high (1.0 GPa) pressure. The friction coefficient generally increases linearly with log(shear rate), in agreement with stress-augmented thermal activation theory67 and tribology experiments.43,51 The friction coefficient shows a much stronger shear rate dependence at low pressure than at high pressure. At low shear rates (< 109 s-1), the friction coefficient is larger at high pressure than at low pressure, whereas at higher shear rates, the reverse is true. Since the slip length is always greater at high pressure, this suggests that slip may reduce friction more effectively in the global slip regime.17,35 A decrease in both the magnitude of the friction coefficient and the slope of its increase with sliding velocity at higher pressure has also been observed in tribology experiments of SA dissolved in hexadecane under hydrodynamic lubrication conditions.38 NEMD simulations of pure lubricant molecules under elastohydrodynamic lubrication (EHL) conditions, where slip occurs within the film itself rather than at the solid-fluid interface, also showed similar friction behaviour.62

22 ACS Paragon Plus Environment

Page 23 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

The slip length (Figure 7a) and the friction coefficient (Figure 7b) both generally increase with increasing shear rate; however, at a given shear rate, systems with a larger slip length generally give a lower friction coefficient. This supports the postulate that OFM films can promote slip38 and thus reduce friction in the hydrodynamic regime.38,65 At high pressure, the friction coefficient drops slightly upon the transition from defect slip to global slip.17,35 These results also suggest that friction reduction by OFMs in the hydrodynamic regime will be greatest when the slip length is maximised through a combination of high pressure and high shear rate, as are typical in the EHL regime.62 For all of the conditions studied, SA gives a significantly higher friction coefficient than SAm and GMS. This observation is in agreement with previous NEMD simulations33 and friction experiments under boundary lubrication conditions.43 This behaviour can be most clearly explained by comparing the mass density profile in Figure 5a with those in Figures 5b and 5c; this shows that, compared to SA films, SAm and GMS films are more solid-like and less interdigitated.33 The reduced interdigitation leads to significantly larger slip lengths for SAm and GMS compared to SA (Figure 7a) and ultimately lower friction (Figure 7b). This finding is expected to be useful in designing new OFMs to control friction in a range of applications.

Interdependence of Slip and Friction It is difficult to establish quantitative relationships between slip and friction from these NEMD simulations because shear thinning of the hexadecane film will also reduce friction particularly under the high pressures and shear rates studied.68 However, it was possible to study the change in slip length and slip velocity with shear stress and compare the trends with previously proposed models for slip. Figure 8a shows the relationship between the shear stress and the slip length for high coverage (4.32 nm-2) OFM films at low pressure (0.1 GPa). Slip only occurs above a critical shear stress, after which the slip length increases linearly with shear stress, and then asymptotes, in agreement with the slip model proposed by Spikes and Granick.36 The critical shear increases with

23 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 33

increasing hexadecane-OFM interdigitation from GMS, to SAm, to SA (Figure 5) and is < 0.01 GPa for all of the OFMs at low pressure.

Figure 8. The effect of shear stress on the slip length (a) and slip velocity (b) for high coverage (Γ = 4.32 nm-2) SA (blue), SAm (red), and GMS (green) films at low pressure (0.1 GPa). Note the logarithmic x-axis in (b). Dotted lines in (a) are guides for the eye, dotted lines in (b) use the extended molecular kinetic theory (MKT) slip model developed by Wang and Zhao37 (Equation 1). Fitting parameters given in Table 1. Arrows on the x-axis show estimates of the critical shear stress.

The slip length values gathered from Figures 3-6 can also be used to calculate slip velocities (orange arrows in Figure 2).11 Figure 8b shows the effect of shear stress on the slip velocity for high coverage (4.32 nm-2) OFM films at low pressure (0.1 GPa). The slip velocity-shear stress behaviour is well described using Eyring’s molecular kinetic theory (MKT), extended to include the critical shear stress and energy dissipation at the interface by Wang and Zhao.37 In this model (Equation 1), the slip velocity is given by, 

      

(Equation 1)



where  is the shear stress,  is a characteristic shear stress,  is a dissipation factor, and  is a characteristic velocity.37 In this study, the product of  and  was used in the fitting for simplicity. In order to fit the dotted lines in Figure 8, both  and   increase with increasing interdigitation, from GMS, to SAm, to SA (Table 1). This suggests that the potential energy barriers for slip to occur

24 ACS Paragon Plus Environment

Page 25 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

are larger when there is more interdigitation, similar to the effect of stronger solid-fluid interactions seen in previous NEMD studies.37 Intuitively, the critical shear stress values obtained for these partial-slip systems (3.0-7.2 MPa) are much larger than observed in NEMD simulations of high-slip (water inside carbon nanotubes) systems (0.003-0.2 MPa).37 Table 1. Extended MKT37 (Equation 1) fitting parameters used in Figure 8b and asymptotic values of the slip length and estimate of the critical shear stress from Figure 8a for all OFMs (P = 0.1 GPa)

 (GPa)    (nm)  (GPa)

SA 0.0065 0.33 0.8 0.0072

SAm 0.0050 0.18 1.1 0.0051

GMS 0.0045 0.17 1.3 0.0030

An important remaining question following these NEMD simulations is the resilience of slip with respect to surface roughness. The presence of nanoscale roughness has been shown to significantly reduce the slip length of alkanes confined by solid surfaces in NEMD simualtions23,24 and experiments.25,26 However, recent NEMD simulations showed that high coverage OFM films can give smooth, low friction interfaces even on surfaces with realistic nanoscale roughness.69 Thus, it is expected that, unlike for NEMD simulations of alkanes confined between bare surfaces with nanoscale roughness, such surfaces coated with high coverage OFM films should still show slip. Although this is beyond the scope of this current study, it is certainly an interesting area to explore and it will be pursued in future NEMD investigations.

25 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 33

Summary and Conclusions In this study, the slip and friction behaviour of n-hexadecane, confined between organic friction modifier (OFM) surfactant films adsorbed on hematite surfaces, have been studied using nonequilibrium molecular dynamics (NEMD) simulations. The influence of the OFM type (stearic acid, stearamide, glycerol mono-stearate) and coverage (1.44 - 4.32 nm-2), as well as the applied shear rate (108 - 1010 s-1) and pressure (0.1 - 1.0 GPa) have been investigated. The slip length is found to be highly sensitive to the OFM type and coverage as well as the applied shear rate and pressure. A measurable slip length is only observed for OFM films with a high surface coverage, which give a smooth interface between well-defined OFM and hexadecane layers. At low and medium coverage, the hexadecane and OFM layers are significantly interdigitated, which prevents slip, resulting in higher friction. Slip only occurs above a critical shear rate, which depends on the applied pressure as well as the OFM type. Above the critical shear rate, the slip length increases with increasing shear rate and subsequently asymptotes towards a constant value. The maximum slip length increases significantly with increasing pressure from ≈ 1 nm at 0.1 GPa to 2 - 5 nm at 1.0 GPa. As has been noted for other systems, there seems to be a lower propensity for slip in these NEMD simulations, which show relatively high critical shear rates and low slip lengths, compared to previous experiments. This suggests that different slip mechanisms could be acting in these experiments. For a given nonequilibrium state point, systems which show a larger slip length typically give a lower friction coefficient. Generally, the friction coefficient increases linearly with logarithmic shear rate, in accordance with stress-augmented thermal activation theory. However, the friction coefficient shows a much stronger shear rate dependence at low pressure (0.1 GPa), where only modest slip lengths are measured, than at high pressure (1.0 GPa), where the slip length is much larger. At low shear rates, the friction coefficient is higher at high pressure than at low pressure, whilst at high shear rates, the reverse is true. Glycerol mono-stearate and stearamide films are less interdigitated

26 ACS Paragon Plus Environment

Page 27 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

than stearic acid films, leading to significantly larger slip lengths and lower friction coefficients. This finding is expected to be useful in designing new surfactants to control hydrodynamic friction and flow in a range of applications. The friction and slip behaviour are interrelated and, in agreement with the slip model due to Spikes and Granick, slip only occurs above a critical shear stress, after which the slip length increases linearly with shear stress and then asymptotes towards a constant value. At low pressure, the slip velocity-shear stress behaviour is well described using the sinh relationship in Eyring’s molecular kinetic theory (MKT) model for slip, extended to include a critical shear stress and energy dissipation at the interface by Wang and Zhao.37 This study has provided a more detailed understanding of how alkane slip occurs on surfactant monolayers adsorbed on solid surfaces. Indeed, to our knowledge, it is the first to successfully measure the slip length in such systems using NEMD. These simulations have also provided compelling evidence that OFMs can significantly reduce friction even when the surfaces are wellseparated, as in the hydrodynamic lubrication regime. The simulations also suggest that friction reduction by OFMs will be greatest when the slip length is maximised through a combination of high pressure and high shear rate, as typical in the elastohydrodynamic lubrication regime. Future NEMD studies will probe the sensitivity of hydrodynamic friction and slip to lubricant viscosity and surface roughness.

27 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 33

Acknowledgements J. P. E. acknowledges the financial support of the Engineering and Physical Sciences Research Council (EPSRC) via a Doctoral Prize Fellowship and a CASE Studentship co-funded by the Shell University Technology Centre (UTC) for Fuels and Lubricants. D. D. also thanks the EPSRC for an Established Career Fellowship EP/N025954/1 and Grant EP/P030211/1. S. K. K. was funded by a strategic grant from the office of the Deputy Vice Chancellor Research and Development at Swinburne University of Technology. The authors thank Hugh A. Spikes for useful discussions. We acknowledge the use of Imperial College London Research Computing Service (RCS) and Swinburne University of Technology HPC service. All data reported in the manuscript can be obtained by emailing the corresponding author or [email protected].

28 ACS Paragon Plus Environment

Page 29 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

References (1)

Sparreboom, W.; Van Den Berg, A.; Eijkel, J. C. T. Principles and Applications of Nanofluidic Transport. Nat. Nanotechnol. 2009, 4 (11), 713–720.

(2)

Cao, B. Y.; Sun, J.; Chen, M.; Guo, Z. Y. Molecular Momentum Transport at Fluid-Solid Interfaces in MEMS/NEMS: A Review. Int. J. Mol. Sci. 2009, 10 (11), 4638–4706.

(3)

Neto, C.; Evans, D. R.; Bonaccurso, E.; Butt, H. J.; Craig, V. S. J. Boundary Slip in Newtonian Liquids: A Review of Experimental Studies. Rep. Prog. Phys. 2005, 68 (12), 2859–2897.

(4)

Pit, R.; Hervet, H.; Leger, L. Direct Experimental Evidence of Slip in Hexadecane: Solid Interfaces. Phys. Rev. Lett. 2000, 85 (5), 980–983.

(5)

Zhu, Y.; Granick, S. Rate-Dependent Slip of Newtonian Liquid at Smooth Surfaces. Phys. Rev. Lett. 2001, 87 (9), 96105.

(6)

McBride, S. P.; Law, B. M. Viscosity-Dependent Liquid Slip at Molecularly Smooth Hydrophobic Surfaces. Phys. Rev. E 2009, 80 (6), 60601.

(7)

Zhu, Y.; Granick, S. No-Slip Boundary Condition Switches to Partial Slip When Fluid Contains Surfactant. Langmuir 2002, 18 (26), 10058–10063.

(8)

Blake, T. D. Slip between a Liquid and a Solid: D.M. Tolstoi’s (1952) Theory Reconsidered. Colloids Surf. 1990, 47, 135–145.

(9)

Eyring, H. Viscosity, Plasticity, and Diffusion as Examples of Absolute Reaction Rates. J. Chem. Phys. 1936, 4, 283–291.

(10)

De Gennes, P. G. On Fluid/wall Slippage. Langmuir 2002, 18 (9), 3413–3414.

(11)

Kumar Kannam, S.; Todd, B. D.; Hansen, J. S.; Daivis, P. J. Slip Length of Water on Graphene: Limitations of Non-Equilibrium Molecular Dynamics Simulations. J. Chem. Phys. 2012, 136 (2), 24705.

(12)

Kumar Kannam, S.; Todd, B. D.; Hansen, J. S.; Daivis, P. J. How Fast Does Water Flow in Carbon Nanotubes? J. Chem. Phys. 2013, 138 (9), 94701.

(13)

Thompson, P. A.; Robbins, M. O. Shear Flow near Solids: Epitaxial Order and Flow Boundary Conditions. Phys. Rev. A 1990, 41 (12), 6830–6837.

(14)

Thompson, P. A.; Troian, S. M. A General Boundary Condition for Liquid Flow at Solid Surfaces. Nature 1997, 389 (6649), 360–362.

(15)

Barrat, J. L.; Bocquet, L. Large Slip Effect at a Nonwetting Fluid-Solid Interface. Phys. Rev. Lett. 1999, 82 (23), 4671–4674.

(16)

Jabbarzadeh, A.; Atkinson, J. D.; Tanner, R. I. Wall Slip in the Molecular Dynamics Simulation of Thin Films of Hexadecane. J. Chem. Phys. 1999, 110 (5), 2612–2620.

(17)

Martini, A.; Roxin, A.; Snurr, R. Q.; Wang, Q.; Lichter, S. Molecular Mechanisms of Liquid Slip. J. Fluid Mech. 2008, 600, 257–269.

(18)

Martini, A.; Hsu, H. Y.; Patankar, N. A.; Lichter, S. Slip at High Shear Rates. Phys. Rev. Lett. 2008, 100 (20), 206001.

(19)

Vadakkepatt, A.; Dong, Y. L.; Lichter, S.; Martini, A. Effect of Molecular Structure on Liquid Slip. Phys. Rev. E 2011, 84 (6), 66311. 29 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 33

(20)

Fillot, N.; Berro, H.; Vergne, P. From Continuous to Molecular Scale in Modelling Elastohydrodynamic Lubrication: Nanoscale Surface Slip Effects on Film Thickness and Friction. Tribol. Lett. 2011, 43 (3), 257–266.

(21)

Savio, D.; Fillot, N.; Vergne, P.; Zaccheddu, M. A Model for Wall Slip Prediction of Confined NAlkanes: Effect of Wall-Fluid Interaction Versus Fluid Resistance. Tribol. Lett. 2012, 46 (1), 11– 22.

(22)

Washizu, H.; Hyodo, S.; Ohmori, T.; Nishino, N.; Suzuki, A. Macroscopic No-Slip Boundary Condition Confirmed in Full Atomistic Simulation of Oil Film. Tribol. Online 2014, 9 (2), 45–50.

(23)

Gao, J. P.; Luedtke, W. D.; Landman, U. Structures, Solvation Forces and Shear of Molecular Films in a Rough Nano-Confinement. Tribol. Lett. 2000, 9 (1–2), 3–13.

(24)

Jabbarzadeh, A.; Atkinson, J. D.; Tanner, R. I. Effect of the Wall Roughness on Slip and Rheological Properties of Hexadecane in Molecular Dynamics Simulation of Couette Shear Flow between Two Sinusoidal Walls. Phys. Rev. E 2000, 61 (1), 690–699.

(25)

Zhu, Y.; Granick, S. Limits of the Hydrodynamic No-Slip Boundary Condition. Phys. Rev. Lett. 2002, 88 (10), 106102.

(26)

Schmatko, T.; Hervet, H.; Leger, L. Effect of Nanometric-Scale Roughness on Slip at the Wall of Simple Fluids. Langmuir 2006, 22 (16), 6843–6850.

(27)

Huang, D. M.; Sendner, C.; Horinek, D.; Netz, R. R.; Bocquet, L. Water Slippage versus Contact Angle: A Quasiuniversal Relationship. Phys. Rev. Lett. 2008, 101 (22), 226101.

(28)

Chinappi, M.; Casciola, C. Intrinsic Slip on Hydrophobic Self-Assembled Monolayer Coatings. Phys. Fluids 2010, 22 (4), 42003.

(29)

Nakaoka, S.; Yamaguchi, Y.; Omori, T.; Kagawa, M.; Nakajima, T.; Fujimura, H. Molecular Dynamics Analysis of the Velocity Slip of a Water and Methanol Liquid Mixture. Phys. Rev. E 2015, 92 (2), 22402.

(30)

Nakaoka, S.; Yamaguchi, Y.; Omori, T.; Joly, L. Molecular Dynamics Analysis of the Friction between a Water-Methanol Liquid Mixture and a Non-Polar Solid Crystal Surface. J. Chem. Phys. 2017, 146 (17), 174702.

(31)

Gupta, S. A.; Cochran, H. D.; Cummings, P. T. Shear Behavior of Squalane and Tetracosane under Extreme Confinement .1. Model, Simulation Method, and Interfacial Slip. J. Chem. Phys. 1997, 107 (23), 10316–10326.

(32)

Doig, M.; Warrens, C. P.; Camp, P. J. Structure and Friction of Stearic Acid and Oleic Acid Films Adsorbed on Iron Oxide Surfaces in Squalane. Langmuir 2014, 30 (1), 186–195.

(33)

Ewen, J. P.; Gattinoni, C.; Morgan, N.; Spikes, H.; Dini, D. Nonequilibrium Molecular Dynamics Simulations of Organic Friction Modifiers Adsorbed on Iron Oxide Surfaces. Langmuir 2016, 32 (18), 4450–4463.

(34)

Ewen, J. P.; Heyes, D. M.; Dini, D. Advances in Nonequilibrium Molecular Dynamics Simulations of Lubricants and Additives. Friction 2018, https://doi.org/10.1007/s40544-0180207-9.

(35)

Lichter, S.; Roxin, A.; Mandre, S. Mechanisms for Liquid Slip at Solid Surfaces. Phys. Rev. Lett. 2004, 93 (8), 86001.

(36)

Spikes, H.; Granick, S. Equation for Slip of Simple Liquids at Smooth Solid Surfaces. Langmuir 2003, 19 (12), 5065–5071. 30 ACS Paragon Plus Environment

Page 31 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

(37)

Wang, F.-C.; Zhao, Y.-P. Slip Boundary Conditions Based on Molecular Kinetic Theory: The Critical Shear Stress and the Energy Dissipation at the Liquid–solid Interface. Soft Matter 2011, 7 (18), 8628.

(38)

Choo, J. H.; Forrest, A. K.; Spikes, H. A. Influence of Organic Friction Modifier on Liquid Slip: A New Mechanism of Organic Friction Modifier Action. Tribol. Lett. 2007, 27 (2), 239–244.

(39)

Spikes, H. Friction Modifier Additives. Tribol. Lett. 2015, 60 (1), 5.

(40)

Hare, E. F.; Zisman, W. A. Autophobic Liquids and the Properties of Their Adsorbed Films. J. Phys. Chem. 1955, 59 (4), 335–340.

(41)

Gee, M. L.; McGuiggan, P. M.; Israelachvili, J. N.; Homola, A. M. Liquid to Solidlike Transitions of Molecularly Thin Films under Shear. J. Chem. Phys. 1990, 93 (3), 1895–1906.

(42)

Plimpton, S. Fast Parallel Algorithms for Short-Range Molecular-Dynamics. J. Comput. Phys. 1995, 117 (1), 1–19.

(43)

Campen, S.; Green, J.; Lamb, G.; Atkinson, D.; Spikes, H. On the Increase in Boundary Friction with Sliding Speed. Tribol. Lett. 2012, 48 (2), 237–248.

(44)

Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual Molecular Dynamics. J. Mol. Graph. Model. 1996, 14 (1), 33–38.

(45)

Maslen, E. N.; Streltsov, V. A.; Streltsova, N. R.; Ishizawa, N. Synchrotron X-Ray Study of the Electron-Density in Alpha-Fe2O3. Acta Crystallogr. B 1994, 50, 435–441.

(46)

Ashurst, W. T.; Hoover, W. G. Dense-Fluid Shear Viscosity via Nonequilibrium Molecular Dynamics. Phys. Rev. A 1975, 11 (2), 658–678.

(47)

Ramachandran, S.; Tsai, B. L.; Blanco, M.; Chen, H.; Tang, Y. C.; Goddard, W. A. SelfAssembled Monolayer Mechanism for Corrosion Inhibition of Iron by Imidazolines. Langmuir 1996, 12 (26), 6419–6428.

(48)

Campen, S.; Green, J. H.; Lamb, G. D.; Spikes, H. A. In Situ Study of Model Organic Friction Modifiers Using Liquid Cell AFM; Saturated and Mono-Unsaturated Carboxylic Acids. Tribol. Lett. 2015, 57 (2), 18.

(49)

Wood, M. H.; Casford, M. T.; Steitz, R.; Zarbakhsh, A.; Welbourn, R. J. L.; Clarke, S. M. Comparative Adsorption of Saturated and Unsaturated Fatty Acids at the Iron Oxide/Oil Interface. Langmuir 2016, 32 (2), 534–540.

(50)

Hirayama, T.; Kawamura, R.; Fujino, K.; Matsuoka, T.; Komiya, H.; Onishi, H. Cross-Sectional Imaging of Boundary Lubrication Layer Formed by Fatty Acid by Means of FrequencyModulation Atomic Force Microscopy. Langmuir 2017, 33 (40), 10492−10500.

(51)

Briscoe, B. J.; Evans, D. C. B. The Shear Properties of Langmuir-Blodgett Layers. Proc. R. Soc. A 1982, 380 (1779), 389.

(52)

Siu, S. W. I.; Pluhackova, K.; Bockmann, R. A. Optimization of the OPLS-AA Force Field for Long Hydrocarbons. J. Chem. Theory Comput. 2012, 8 (4), 1459–1470.

(53)

Jorgensen, W. L.; Maxwell, D. S.; Tirado-Rives, J. Development and Testing of the OPLS AllAtom Force Field on Conformational Energetics and Properties of Organic Liquids. J. Am. Chem. Soc. 1996, 118 (45), 11225–11236.

(54)

Price, M. L. P.; Ostrovsky, D.; Jorgensen, W. L. Gas-Phase and Liquid-State Properties of Esters, Nitriles, and Nitro Compounds with the OPLS-AA Force Field. J. Comput. Chem. 2001, 31 ACS Paragon Plus Environment

Langmuir 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 33

22 (13), 1340–1352. (55)

Ewen, J. P.; Gattinoni, C.; Thakkar, F. M.; Morgan, N.; Spikes, H.; Dini, D. A Comparison of Classical Force-Fields for Molecular Dynamics Simulations of Lubricants. Materials. 2016, 9 (8), 651.

(56)

Yeh, I. C.; Berkowitz, M. L. Ewald Summation for Systems with Slab Geometry. J. Chem. Phys. 1999, 111 (7), 3155–3162.

(57)

Berro, H.; Fillot, N.; Vergne, P. Molecular Dynamics Simulation of Surface Energy and ZDDP Effects on Friction in Nano-Scale Lubricated Contacts. Tribol. Int. 2010, 43 (10), 1811–1822.

(58)

Ryckaert, J. P.; Ciccotti, G.; Berendsen, H. J. C. Numerical-Integration of Cartesian Equations of Motion of a System with Constraints: Molecular-Dynamics of N-Alkanes. J. Comput. Phys. 1977, 23 (3), 327–341.

(59)

Nosé, S. A Molecular-Dynamics Method for Simulations in the Canonical Ensemble. Mol. Phys. 1984, 52 (2), 255–268.

(60)

Hoover, W. G. Canonical Dynamics: Equilibrium Phase-Space Distributions. Phys. Rev. A 1985, 31 (3), 1695–1697.

(61)

Gattinoni, C.; Mackowiak, S.; Heyes, D. M.; Branka, A. C.; Dini, D. Boundary-Controlled Barostats for Slab Geometries in Molecular Dynamics Simulations. Phys. Rev. E 2014, 90 (4), 43302.

(62)

Ewen, J.; Gattinoni, C.; Zhang, J.; Heyes, D. M.; Spikes, H. A.; Dini, D. On the Effect of Confined Fluid Molecular Structure on Nonequilibrium Phase Behaviour and Friction. Phys. Chem. Chem. Phys. 2017, 19 (27), 17883–17894.

(63)

Bernardi, S.; Todd, B. D.; Searles, D. J. Thermostating Highly Confined Fluids. J. Chem. Phys. 2010, 132 (24), 244706.

(64)

Watanabe, S.; Nakano, M.; Miyake, K.; Sasaki, S. Analysis of the Interfacial Molecular Behavior of a Lubrication Film of N-Dodecane Containing Stearic Acid under Lubricating Conditions by Sum Frequency Generation Spectroscopy. Langmuir 2016, 32 (51), 13649–13656.

(65)

Ponjavic, A.; Wong, J. S. S. The Effect of Boundary Slip on Elastohydrodynamic Lubrication. RSC Adv. 2014, 4 (40), 20821–20829.

(66)

Taylor, R.; de Kraker, B. Shear Rates in Engines and Implications for Lubricant Design. Proc. Inst. Mech. Eng. Part J. 2017, 231 (9), 1106–1116.

(67)

Spikes, H. A. Stress-Augmented Thermal Activation: Tribology Feels the Force. Friction 2018, 6 (1), 1–31.

(68)

Tseng, H. C.; Wu, J. S.; Chang, R. Y. Shear Thinning and Shear Dilatancy of Liquid NHexadecane via Equilibrium and Nonequilibrium Molecular Dynamics Simulations: Temperature, Pressure, and Density Effects. J. Chem. Phys. 2008, 129, 14502.

(69)

Ewen, J. P.; Echeverri Restrepo, S.; Morgan, N.; Dini, D. Nonequilibrium Molecular Dynamics Simulations of Stearic Acid Adsorbed on Iron Surfaces with Nanoscale Roughness. Tribol. Int. 2017, 107, 264–273.

32 ACS Paragon Plus Environment

Page 33 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Langmuir

TOC Graphic

33 ACS Paragon Plus Environment