-
Notifications
You must be signed in to change notification settings - Fork 0
/
main.tex
1341 lines (1122 loc) · 109 KB
/
main.tex
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
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
%\documentclass[aps,prb,reprint,superscriptaddress, a4paper]{revtex4-1}
%\documentclass[aps,prl,preprint,superscriptaddress]{revtex4-1}
%\documentclass[aps,prl,reprint,groupedaddress]{revtex4-1}
\documentclass[5p]{elsarticle}
% You should use BibTeX and apsrev.bst for references
% Choosing a journal automatically selects the correct APS
% BibTeX style file (bst file), so only uncomment the line
% below if necessary.
%\bibliographystyle{apsrev4-1}
\usepackage{siunitx}
\usepackage[shell]{gnuplottex}
%\usepackage[miktex]{gnuplottex}
\usepackage{float}
\usepackage{hyperref}
\usepackage{pgf}
\usepackage{graphicx}
\usepackage{amsmath}
\usepackage{tikz}
\usepackage{tikz-dimline}
\usepackage{bm}
\usepackage{structuralanalysis}
\usepackage{color}
\usepackage{diagbox}
\usepackage{placeins}
\usepackage{enumitem}
\setlist[description]{leftmargin=1pt,labelindent=1pt,itemsep=-2pt,topsep=1pt}
\renewcommand{\descriptionlabel}[1]{
\hspace\labelsep \upshape #1
}
\definecolor{SERorange}{HTML}{FFA500}
\definecolor{SERred}{HTML}{F00000}
\definecolor{SERgray}{HTML}{909090}
\definecolor{SERwhite}{HTML}{F8F8F8}
\definecolor{SERgreen}{HTML}{34B853}
\definecolor{SERorange2}{HTML}{F64724}
\definecolor{SERyellow}{HTML}{FFCC00}
\definecolor{SERblue}{HTML}{4295F9}
\DeclareSIUnit{\calorie}{cal}
%\usepackage{enumitem}
\begin{document}
% Use the \preprint command to place your local institutional report
% number in the upper righthand corner of the title page in preprint mode.
% Multiple \preprint commands are allowed.
% Use the 'preprintnumbers' class option to override journal defaults
% to display numbers if necessary
%\preprint{}
%Title of paper
\title{Behaviour of $n$-Alkanes Confined between Iron Oxide Surfaces at High Pressure and Shear Rate: A Nonequilibrium Molecular Dynamics Study}
% repeat the \author .. \affiliation etc. as needed
% \email, \thanks, \homepage, \altaffiliation all apply to the current
% author. Explanatory text should go in the []'s, actual e-mail
% address or url should go in the {}'s for \email and \homepage.
% Please use the appropriate macro foreach each type of information
% \affiliation command applies to all authors since the last
% \affiliation command. The \affiliation command should follow the
% other information
% \affiliation can be followed by \email, \homepage, \thanks as well.
\author[SKF,KCL]{Sebasti\'{a}n Echeverri Restrepo}
\ead{sebastianecheverrir@gmail.com}
\author[SKF]{Marcel C. P. van Eijk}
\ead{marcel.van.eijk@skf.com}
\author[IC,CA]{James P. Ewen}
\ead{j.ewen@imperial.ac.uk}
%\author[IC]{Daniele Dini}
\address[SKF]{SKF Research \& Technology Development, Nieuwegein, The Netherlands}
\address[KCL]{Department of Physics, King's College London, Strand, London WC2R 2LS, UK}
\address[IC]{Department of Mechanical Engineering, Imperial College London, London SW7 2AZ, England, UK}
\address[CA]{Corresponding author}
%Collaboration name if desired (requires use of superscriptaddress
%option in \documentclass). \noaffiliation is required (may also be
%used with the \author command).
%\collaboration can be followed by \email, \homepage, \thanks as well.
%\collaboration{}
%\noaffiliation
\begin{abstract}
The behaviour of n-alkanes confined and sheared between iron oxide surfaces has been investigated using nonequilibrium molecular dynamics simulations. The chain extension and orientation, film structure and flow, and friction have been measured for a range of chain lengths under conditions representative of the elastohydrodynamic lubrication regime. At high pressure, the molecules show strong layering and long-range order, suggesting solid-like films. Conversely, high shear rates result in less elongated, layered, and ordered chains; indicating more liquid-like films. Although Couette flow is usually observed for short chains, the flow is often non-linear for long chains. The friction coefficient increases logarithmically with shear rate, but the slope decreases with increasing pressure such that it becomes insensitive to shear rate for long chains.
\end{abstract}
% insert suggested PACS numbers in braces on next line
%\pacs{}
%\maketitle must follow title, authors, abstract, \pacs, and \keywords
\maketitle
% body of paper here - Use proper section commands
% References should be done using the \cite, \ref, and \label commands
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Introduction}
A detailed understanding of the molecular-scale behaviour of fluids confined, pressurised, and sheared between solid surfaces critical in tribology. In particular, for lubricated machine components that are based on elements that roll and slide together (e.g. rolling bearings, gears, constant velocity joints and cam/follower systems), a significant portion of the friction losses originate from the elastohydrodynamic lubrication (EHL) regime. In the EHL regime, thin lubricant films (nm) are subjected to very high pressures (GPa) and shear rates ($> 10^{7}~\SI{}{\per\second}$). These conditions are extremely challenging to probe directly through experiment, and thus the behaviour of lubricants under EHL conditions remains somewhat unclear~\cite{Spikes2014}.
\textcolor{blue}{Alongside experiments, nonequilibrium molecular dynamics (NEMD) simulations can provide unique insights into the behaviour of lubricants under EHL conditions~\cite{Ewen2018}. Since the early 1990s, many studies have investigated the behaviour of lubricants strongly confined ($< 6$ molecular layers) between solid surfaces, where confinement is known to significantly increase the viscosity~\cite{Granick1991}. Early NEMD simulations of thin atomic and molecular fluid films suggested that this viscosity increase was due to vitrification to a glassy state or even crystallization for atomic fluids~\cite{Thompson1992}. Indeed, MD simulations of $n$-dodecane (C$_{12}$) showed a trasition from liquidlike to solidlike behaviour between 6 and 7 molecular layers at ambient temperature and pressure~\cite{Cui2001}. Further MD simulations of C$_{12}$ suggested that the initial increase in viscosity from confinement is due to the formation of crystal bridges between the surfaces, which is followed by a further increase due to tetratic order upon increasing confinement~\cite{Jabbarzadeh2006,Jabbarzadeh2007}.
NEMD simulations of strongly confined systems showed flow behaviour that deviated from the linear (Couette) case, such as slip at the solid walls, as well as slip between molecular layers within the fluid film itself~\cite{Thompson1990}. NEMD comparisons between thin (8 molecular layers) films of different $n$-alkane chain lengths (C$_6$-C$_{80}$) suggested that, at low pressure (\SI{0.1}{\mega\pascal}), interlayer slip within the ordered film occurred and shear stress was sensitive to the chain length~\cite{Koike1998}. Conversely, at high pressure (\SI{100}{\mega\pascal}), boundary slip occurred and the shear stress was insensitive to the chain length~\cite{Koike1998}. For thin $n$-alkane films at high pressure (\SI{500}{\mega\pascal}) between $\alpha$-Fe$_2$O$_3$ surfaces, where no slip occurred, the shear stress was shown to be dependent on chain length (C$_{1}$-C$_{25}$)~\cite{Savio2013a}. The shear stress was also dependant on chain length (C$_{20}$-C$_{1400}$) for thin (7 molecular layers) $n$-alkane films sheared between polymer-like surfaces at low pressure (\SI{10}{\mega\pascal}), where no slip occurred~\cite{Sivebaek2008}. However, for metal-like surfaces, on which the $n$-alkanes showed slip, the shear stress was independent on the chain length under the same conditions~\cite{Sivebaek2008}. For the same systems, the shear stress was also shown to increase monotonically with the sliding velocity for metal-like surfaces but to be independent of sliding velocity on polymer-like surfaces~\cite{Sivebaek2010}. It was also found that shorter (C$_{20}$) chains melted at shear rates, whereas longer chains only showed thermal softening~\cite{Sivebaek2012}. NEMD simulations of very thick ($>$ 400 nm) C$_{6}$ films confined between Fe surfaces under EHL conditions confirmed the applicability of no-slip boundary conditions in this case~\cite{Washizu2010,Washizu2014}.}
Relatively fewer NEMD studies have investigated the behaviour of thicker lubricant films ($\geq 10$ molecular layers) of direct relevance to EHL. Both NEMD simulations~\cite{Thompson1992,Robbins1996} and experiments~\cite{VanAlsten1988} have suggested that there could be similarities between the behaviour of strongly confined films and thicker films subjected to high pressures. Indeed, comprehensive NEMD simulations of relatively thick (35 atomic layers) atomic fluid films showed a range of `nonequilibrium phases' under high shear rate and pressure conditions.~\cite{Heyes2012,Gattinoni2013} This behaviour was reminiscent of previous observations for more strongly confined films at lower pressure~\cite{Thompson1990}. Further NEMD simulations showed that the nonequilibrium phase transitions observed at high pressure can lead to deviations from classical friction laws~\cite{Mackowiak2016}, which has also been observed for thinner $n$-alkane films at lower pressure~\cite{Sivebaek2010}. Recently, unusual friction and nonequilibrium phase behaviour has also been observed in NEMD simulations of relatively thick films of model lubricant and traction fluid molecules under EHL conditions~\cite{Ewen2017a}. To design new lubricant molecules to control EHL friction, a clear understanding of how molecular structure affects friction and nonequilibrium phase behaviour is required.
In the present study, the behaviour of $n$-alkanes confined, pressurised, and sheared between $\text{Fe}_2\text{O}_3$ surfaces (as a model for steel) has been investigated under EHL conditions using NEMD simulations. The effect of $n$-alkane chain length (C$_{16}$-C$_{60}$), applied pressure ($0.5 - 1.5~\SI{}{\giga\pascal}$) and shear rate ($10^{9} - 10^{10} ~\SI{}{\per\second}$) on the chain extension and orientation, film structure and flow, and friction has been studied simultaneously. These results shed further light on the molecular-scale behaviour of model lubricants under EHL conditions.
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Methodology}
\label{method}
Classical MD simulations were performed using the large-scale atomic/molecular massively parallel simulator (LAMMPS) software~\cite{Plimpton1995}. The velocity-Verlet algorithm was used to integrate the equations of motion with a time-step of \SI{1.0}{\femto\second}. All of the initial configurations were generated using the open source LAMMPS\_builder software~\cite{Ewen2017b,Ewen2017}, which utilises Moltemplate~\cite{Jewett2013} and the Atomic Simulation Environment (ASE)~\cite{HjorthLarsen2017}. The systems consist of two parallel atomically-smooth surfaces separated an approximately \SI{8}{\nano\meter} thick $n$-alkane lubricant layer, see Figure \ref{fig:Regions}. The $n$-alkane layer was sufficiently thick ($\geq 16$ molecular layers) such that the lubricant was not strongly layered in the middle of the film and thus any confinement-induced viscosity increase was expected to be negligible~\cite{Gee1990}. Periodic boundary conditions were applied in the \emph{x} and \emph{y} directions.
The lubricant layer consisted of $n$-alkanes of different lengths: $n$-hexadecane (C$_{16}$), $n$-triacontane (C$_{30}$) and $n$-hexacontane (C$_{60}$). For all the cases, a total number of 19200 carbon atoms were inserted into the system; this corresponds to 1200, 640 and 320 molecules for C$_{16}$, C$_{30}$ and C$_{60}$, respectively. In modern engine oils, most lubricant molecules are in the middle of this range, with $n$-alkanes constituting a significant proportion of the various isomers present~\cite{Liang2018}.
Parameters from the long hydrocarbon-optimised potentials for liquid simulations, all-atom force field (L-OPLS-AA)~\cite{Jorgensen1996,Siu2012} were used for the carbon and hydrogen atoms in the alkane chains. This force-field has been shown to correctly describe the viscous behaviour of long alkanes under ambient and high temperature, high pressure conditions~\cite{Ewen2016a}. Moreover, all-atom force-fields have also been shown to be important to accurately model phase transitions in confined alkane systems~\cite{Docherty2010}. Cross-interactions were evaluated using the geometric mean mixing rules~\cite{Jorgensen1996}. All C-H bonds were constrained using the SHAKE algorithm~\cite{Ryckaert1977}. Lennard-Jones interactions were cut off at \SI{12}{\angstrom} and electrostatic interactions were evaluated using a slab implementation of the particle-particle-particle-mesh (PPPM) algorithm~\cite{Yeh1999} with a relative accuracy in the forces of \SI{1e-7}{}.
Hematite $\left(\text{Fe}_2\text{O}_3\right)$ slabs were chosen as a representative model for steel surfaces~\cite{Oh1998}. The hematite slabs~\cite{Maslen1994} were cleaved and oriented such that the $\left[100\right]$ plane is in contact with the lubricant (see Figure \ref{fig:Regions}). The thickness of the surfaces in the \emph{z} direction is \SI{13.7}{\angstrom} and the size of the surface in the \emph{x} and \emph{y} directions are \SI{80.5}{\angstrom} and \SI{78.4}{\angstrom} respectively. The cell dimensions were large enough to prevent the interaction of individual $n$-alkane chains with their own periodic image under shear. The Fe and O surface atoms are restrained in the corundum crystal structure~\cite{Maslen1994} with harmonic bonds between atoms within \SI{3}{\angstrom}. The force constant of these bonds was \SI{130}{\kilo\calorie\per\mol\per\angstrom\squared}, which has been shown previously to be a reasonable compromise between surface rigidity and thermostatting efficiency~\cite{Berro2010}. The surface Fe and O Lennard-Jones and partial charge parameters, which determine the strength of the interactions with the lubricant atoms were taken from ref.~\cite{Savio2012} and ref.~\cite{Berro2010} respectively. These parameters were developed to reproduce experimental desorption energies of $n$-alkanes on hematite surfaces.~\cite{Savio2012}
\subsection{Equilibration}
First, $n$-alkane molecules were inserted between the slabs in an ordered fashion, followed by energy minimisation. Next, a properly equilibrated system was a generated using a simple heat-quench sequence. In this procedure, the initially ordered $n$-alkanes were heated to \SI{2000}{\kelvin} to accelerate diffusion, and left to evolve until equilibrium is reached (\SI{10}{\nano\second}). The system was then quenched back to the target temperature \SI{353}{\kelvin} over a further \SI{10}{\nano\second}. During the heat-quench sequence, the fluid temperature was controlled using a global Langevin thermostat~\cite{Schneider1978}, with a damping constant of \SI{0.1}{\pico\second}, acting in all Cartesian directions.
Three different criteria were considered to ensure that the systems had reached equilibrium. Specifically, verifying that the mean square end-to-end distance, $\left(\left< R^2 \right> \right)$ and the mean square radius of gyration $\left(\left< S^2 \right> \right)$ reached a steady state~\cite{Brown1994}. The final criterion reflects the fact that systems of long chain alkanes which are not properly equilibrated can present deformation on short length scales, and this deformation is only relaxed after the chains have moved a distance equivalent to their own size~\cite{Auhl2003}. In order to verify this, the mean-squared displacement for the C atoms $\left(\text{MSD}_{\text{atom}}\right)$ and for the centre of mass of the alkane chains $\left(\text{MSD}_{\text{CG}}\right)$ were also monitored during equilibration. More details regarding the equilibration procedure can be found in the Appendix. The final equilibrated configurations were used as inputs for the following compression and shear simulations.
\subsection{Compression}
After the systems were properly equilibrated, the pressure was increased by assigning a constant normal (\emph{z}) force to the outer layer of Fe atoms in the top wall while keeping the outer layer of atoms in the bottom wall fixed in the \emph{z} direction (see Figure \ref{fig:Regions}). Three different pressure values were considered; \SI{0.5}{\giga\pascal}, \SI{1.0}{\giga\pascal}, and \SI{1.5}{\giga\pascal}. These pressures are of direct relevance to lubricants in the the EHL regime~\cite{Spikes2014}.
\begin{figure}[htp]
\begin{center}
\begin{tikzpicture}
\node[anchor=south west,inner sep=0] (fig_regions) at (0,0){\includegraphics[width=0.25\textwidth]{DataDump/Images/5_region.png}};
\begin{scope}[x={(fig_regions.south east)},y={(fig_regions.north west)}]
\draw [](0,0.8) node[text width=1cm,align=center]{Top\\Wall} ;
\draw [](0,0.53) node[]{Fluid} ;
\draw [](0,0.27) node[text width=1cm,align=center]{Bottom\\Wall} ;
\draw [decorate,decoration={brace}] (0.17,0.76) -- (0.17,0.86);
\draw [decorate,decoration={brace}] (0.17,0.295) -- (0.17,0.755);
\draw [decorate,decoration={brace}] (0.17,0.19) -- (0.17,0.29);
\node (A) at (0.7,0.9) {} ;
\node (B) at (1.0,1.0) [SERgreen]{Frozen} ;
\draw [SERgreen, <- , line width=0.006*\textwidth] (A) -- (B);
\node (C) at (0.8,0.78) {} ;
\node (D) at (1.2,0.9) [SERorange2] {Thermostat} ;
\draw [SERorange2, <- , line width=0.006*\textwidth] (C) -- (D);
\node (E) at (0.8,0.74) {} ;
\node (F) at (1.2,0.7) [SERyellow]{Free} ;
\draw [SERyellow, <- , line width=0.006*\textwidth] (E) -- (F);
\node (G) at (0.3,0.79) {} ;
\node (H) at (0.6,0.725) []{$v/2$} ;
\draw [ -> , line width=0.004*\textwidth] (G) -- (H);
\node (I) at (0.3,0.23) {} ;
\node (J) at (0.6,0.165) []{$v/2$} ;
\draw [ <- , line width=0.004*\textwidth] (I) -- (J);
\node (K) at (0.62,1.1) {$P$} ;
\node (L) at (0.62,0.85) []{} ;
\draw [ <- , line width=0.008*\textwidth] (L) -- (K);
\end{scope}
\end{tikzpicture}
\caption{Structure and regions of a representative system for these simulations. The system is divided in three regions, namely, the $n$-alkane fluid, and the top and bottom walls. Within each wall, the \textit{frozen} atoms (green) are used to apply the shearing and compressive constraints, and the \textit{thermostat} atoms (orange) to control the temperature. Both the \textit{free} layers (yellow) and the fluid (blue) atoms are unconstrained.}
\label{fig:Regions}
\end{center}
\end{figure}
For the compression and shear stages, the temperature of the system is controlled (T = \SI{353}{\kelvin}) by a Langevin thermostat~\cite{Schneider1978} with a damping coefficient of \SI{0.1}{\pico\second}, acting on the middle layer of atoms in both slabs (\textit{thermostat} in Figure \ref{fig:Regions}). The thermostat was applied only in the direction perpendicular to both the sliding and compression (\emph{y}). This approach has been shown to be more physically accurate compared to application of the thermostat directly to the confined fluid, which can significantly affect friction and flow behaviour, particularly at high shear rates~\cite{Liem1992,Bernardi2010,Yong2013}.
The compression phase is performed in two steps; first, the pressure is linearly incremented over \SI{10}{\nano\second} and then the target pressure is maintained for a further \SI{10}{\nano\second}. The density of the fluid region and the average pressure both reach a steady state well within the simulation time. The average $n$-alkane densities over the final \SI{1}{\nano\second} of the compression phase are presented in Table \ref{tab:rho}. In agreement with experiments data~\cite{Griesbaum2000}, there is an increase in average density with increasing $n$-alkane chain length, as well as with increasing pressure.
\begin{table}
\caption{Effect of confined $n$-alkane chain length and pressure on the average density in the fluid region (\ref{fig:Regions})}.
\centering
\begin{tabular}{c | l l l}
\hline\hline\\ [-2ex]
& \multicolumn{3}{c}{ $\rho \, [\SI{}{\kilogram\per\cubic\meter}]$} \\
\hline\\ [-2ex]
\backslashbox{Chain \\ Length}{P $[\SI{}{\giga\pascal}]$} & 0.5 & 1.0 & 1.5 \\
\hline\\ [-2ex]
C$_{16}$ & 0.87 & 0.93 & 0.98 \\
C$_{30}$ & 0.90 & 0.96 & 1.00 \\
C$_{60}$ & 0.91 & 0.97 & 1.01 \\
\hline\hline \\[-2ex]
\end{tabular}
\label{tab:rho}
\end{table}
\subsection{Shear}
After the $n$-alkanes are equilibrated at the target pressures (\SI{0.5}-\SI{1.5}{\giga\pascal}), the shear response of the lubricant was studied. A shear velocity gradient was imposed on the system by applying an equal and opposite sliding velocity, $v_x = \pm v_s/2$, to the outermost layer of atoms in each slab (see Figure \ref{fig:Regions}) in the \emph{x} direction. The following sliding velocities were considered, $v_s$; \SI{10}{\meter\per\second}, \SI{20}{\meter\per\second}, \SI{50}{\meter\per\second} and \SI{100}{\meter\per\second}. For the simulated film thickness (approx. \SI{8}{\nano\meter}), these correspond to applied shear rates, $\dot{\gamma}$, of approximately 10$^{9}$-10$^{10} \text{s}^{-1}$. While these are above those encountered in real engineering components (10$^{6}$-10$^{8} \text{s}^{-1}$)~\cite{Taylor2017}, lower shear rates do not reach a steady state in the available simulation time for the long chains studied~\cite{Ewen2018}.
To verify that a nonequilibrium steady state was reached, the evolution of $\left< R^2 \right>$, $\left<P_{2}^{xz} \right>$~\cite{Erman1985} and $F_L$ on the surfaces in response to the fluid, were monitored. Lower sliding velocities require a longer simulation time to reach a steady state (up to 200 ns); however, they require a similar sliding distance, in agreement with experimental observations~\cite{Drummond2000}. In general, a steady state is reached in a shorter sliding distance for shorter $n$-alkanes and lower pressures. A representative case is C$_{30}$ at \SI{1.0}{\giga\pascal}, for which the evolution of $\left< R^2 \right>$, $\left<P_{2}^{xz} \right>$, and $F_L$ are presented at four sliding velocities in Figure \ref{fig:SS}.
\pgfmathsetmacro{\SERFigwidth}{.035\linewidth}
\pgfmathsetmacro{\SERFigheight}{.036\linewidth}
\begin{figure}[htp]
\begin{center}
\begin{gnuplot}[terminal=epslatex, terminaloptions={size \SERFigwidth cm, \SERFigheight cm color solid}]
set multiplot layout 3,1 rowsfirst
set colors podo
set format x ' '
unset xlabel
set key top right
set key samplen 0.0
set ylabel "$\\left< R^2\\right> \\,[\\SI{}{\\square\\angstrom}]$"
set label at graph 0.03,0.89 left "a)"
set lmargin at screen 0.25; set rmargin at screen 0.9
set tmargin at screen 1.00; set bmargin at screen 0.7
set ytics 280,40,400
plot [:700][260:430] 'DataDump/Shear/C30/1.0GPa/10m_s/e2e2.plot_s' u (($1/1e6)*10):($2) w l lw 3 lt 3 notitle ,\
'DataDump/Shear/C30/1.0GPa/10m_s/e2e2.plot_s1' u (($1/1e6)*10):($2) w l lw 3 lt 3 notitle ,\
'DataDump/Shear/C30/1.0GPa/20m_s/e2e2.plot_s' u (($1/1e6)*20):($2) w l lw 3 lt 4 notitle ,\
'DataDump/Shear/C30/1.0GPa/20m_s/e2e2.plot_s1' u (($1/1e6)*20):($2) w l lw 3 lt 4 notitle ,\
'DataDump/Shear/C30/1.0GPa/50m_s/e2e2.plot_s' u (($1/1e6)*50):($2) w l lw 3 lt 5 notitle ,\
'DataDump/Shear/C30/1.0GPa/100m_s/e2e2.plot_s' u (($1/1e6)*100):($2) w l lw 3 lt 10 notitle
unset label
set tmargin at screen 0.7; set bmargin at screen 0.4
set ytics 0.3, 0.1, 0.6
set ylabel "$\\left<P_{2}^{xz}\\right>$"
set label at graph 0.03,0.89 left "b)"
plot [:700][0.28:0.63] 'DataDump/Shear/C30/1.0GPa/10m_s/P2XZ_s.plot' u (($1/1e6)*10):($2) w l lw 3 lt 3 notitle ,\
'DataDump/Shear/C30/1.0GPa/10m_s/P2XZ_s.plot1' u (($1/1e6)*10):($2) w l lw 3 lt 3 notitle ,\
'DataDump/Shear/C30/1.0GPa/20m_s/P2XZ_s.plot' u (($1/1e6)*20):($2) w l lw 3 lt 4 notitle ,\
'DataDump/Shear/C30/1.0GPa/20m_s/P2XZ_s.plot1' u (($1/1e6)*20):($2) w l lw 3 lt 4 notitle ,\
'DataDump/Shear/C30/1.0GPa/50m_s/P2XZ_s.plot' u (($1/1e6)*50):($2) w l lw 3 lt 5 notitle , \
'DataDump/Shear/C30/1.0GPa/100m_s/P2XZ_s.plot' u (($1/1e6)*100):($2) w l lw 3 lt 10 notitle
unset label
set tmargin at screen 0.4; set bmargin at screen 0.1
set ytics 30, 30, 150
set ylabel "$F_L \\, [\\SI{}{\\mega\\pascal}]$"
set format x '%g'
set xlabel "Sliding distance [\\SI{}{\\nano\\meter}]"
set label at graph 0.03,0.89 left "c)"
plot [:700][60:160] 'DataDump/Shear/C30/1.0GPa/10m_s/fc_ave.dump' u (($1/1e6)*10):($2/10) w l lw 3 lt 3 title "\\SI{10}{\\meter\\per\\second}" ,\
'DataDump/Shear/C30/1.0GPa/10m_s/fc_ave.dump1' u (($1/1e6)*10):($2/10) w l lw 3 lt 3 notitle ,\
'DataDump/Shear/C30/1.0GPa/20m_s/fc_ave.dump' u (($1/1e6)*20):($2/10) w l lw 3 lt 4 title "\\SI{20}{\\meter\\per\\second}" ,\
'DataDump/Shear/C30/1.0GPa/20m_s/fc_ave.dump1' u (($1/1e6)*20):($2/10) w l lw 3 lt 4 notitle ,\
'DataDump/Shear/C30/1.0GPa/50m_s/fc_ave.dump' u (($1/1e6)*50):($2/10) w l lw 3 lt 5 title "\\SI{50}{\\meter\\per\\second}" ,\
'DataDump/Shear/C30/1.0GPa/100m_s/fc_ave.dump' u (($1/1e6)*100):($2/10) w l lw 3 lt 10 title "\\SI{100}{\\meter\\per\\second}"
unset multiplot
\end{gnuplot}
\caption{Evolution of the a) mean square end-to-end distance, $\left< R^2 \right>$, b) average segmental orientation, $\left<P_{2}^{xz}\right>$, and c) average lateral (friction) force $F_L$ for C$_{30}$ at a pressure of \SI{1.0}{\giga\pascal}.}
\label{fig:SS}
\end{center}
\end{figure}
%\FloatBarrier
Since the initial configuration is the same for all of the sliding velocities, they all have an initial value of $\left< R^2 \right> = \SI{283}{\angstrom\squared}$. After this point, $\left< R^2 \right> $ increases and then asymptotes towards a steady state value. This suggests that the chains generally unfold when shear is applied. The steady state $\left< R^2 \right>$ generally decreases with increasing sliding velocity. This is consistent with previous confined NEMD simulations of long $n$-alkanes at high shear rates~\cite{Cho2017}.
For ideal chains in the bulk, the average segmental orientation, $\langle P_{2}^{xz}\rangle=0.25$. This suggests a uniform distribution of the orientation of the chains, which corresponds to an average orientation angle of \SI{45}{\degree}. When under confinement, $n$-alkane molecules close to the surfaces tend to align parallel to them~\cite{Cho2017}, meaning that the angle between the chain and the surface is close to zero, and hence $\langle P_{2}^{xz}\rangle$ increases. In the case of C$_{30}$ at \SI{1.0}{\giga\pascal}, $\langle P_{2}^{xz}\rangle=0.32$ before shear is applied (see Figure \ref{fig:SS}b), indicating that there is already a preferential orientation due to the presence of the surfaces. When sliding is initiated, the chains begin to align with the flow direction, further increasing $\langle P_{2}^{xz}\rangle$. Generally, $\langle P_{2}^{xz}\rangle$ decreases with increasing sliding velocity, this is discussed in the next section.
The shear stress was monitored through the average lateral (friction) force $F_L$ (or shear stress) acting on the outermost layer of atoms (frozen atoms in Figure \ref{fig:Regions}) in the top and bottom slabs (divided by the surface area) in response to the fluid. At the onset of sliding, $F_L$ increases rapidly to reach a maximum value and then decreases before reaching a steady state, indicating `stress overshoot' behaviour~\cite{Jeong2017}. The sliding distance needed for $F_L$ to reach a steady state is consistent with that needed for $\left< R^2 \right> $ and $\left<P_{2}^{xz} \right> $ (Figure \ref{fig:SS}). Thus $F_L$ reaches a minimum only when the $n$-alkane chains extend and align with the flow. Generally, $F_L$ increases with increasing sliding velocity, this is discussed in the next section.
Once the simulations reach a steady state (Figure \ref{fig:SS}), they are run for a further \SI{5}{\nano\second} to \SI{20}{\nano\second} while maintaining a constant sliding velocity and applied pressure. During the final \SI{2}{\nano\second} of the simulation, $\left< R^2 \right>$, $\left<P_{2}^{xz} \right> $, and $F_L$ are measured and averaged. These results are presented and discussed in the next section.
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Results and Discussion}
First, the effect of chain length, pressure, and shear rate on the chain extension, $\left< R^2 \right>$ and orientation, $\left<P_{2}^{xz} \right> $ are studied in Section \ref{ext}. The change in film structure and flow are investigated through mass density and velocity profiles as well as radial distribution functions in Section \ref{str}. Finally, the dependency of the friction on the chain length, pressure, and shear rate are presented in Section \ref{fri}.
\subsection{Chain extension and orientation}
\label{ext}
Longer chains will inherently have higher absolute values of $\left< R^2 \right> $, so to compare the variation of $\left< R^2 \right> $ between the different chain lengths, the mean square end-to-end distance was normalised \emph{per C-C bond}; $\left< R^2 \right>/\left(N_\text{bonds}\right)^2$. The steady state $\left< R^2 \right> $ per C-C bond is presented as a function of shear rate for the different chain lengths and pressures studied in Figure \ref{fig:e2e2_v}.
\pgfmathsetmacro{\SERFigwidth}{.035\linewidth}
\pgfmathsetmacro{\SERFigheight}{.026\linewidth}
\begin{figure}[htp]
\begin{center}
\begin{gnuplot}[terminal=epslatex, terminaloptions={size \SERFigwidth cm, \SERFigheight cm color solid}]
set colors podo
set key above
set xlabel "$\\dot{\\gamma} \\, \\left[ \\SI{}{\\per\\second} \\right]$"
set ylabel "$\\left< R^2 \\right>/\\left(N_\\text{bonds}\\right)^2$ [\\SI{}{\\square\\angstrom}]"
set ytics 0.1,0.1,0.7
set key noautotitle
set format x '$10^{%L}$'
set logscale x
p [:2e10][0.25:0.65] 'DataDump/Shear/Compiled.plot8' i 0 u ($3/($12*1e-10)):($6/($1-1)**2):($7/($1-1)**2) title "$\\SI{0.5}{\\giga\\pascal}$" lt 1 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 1 u ($3/($12*1e-10)):($6/($1-1)**2):($7/($1-1)**2) title "$\\SI{1.0}{\\giga\\pascal}$" lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 1 u ($3/($12*1e-10)):($6/($1-1)**2) w l notitle lt 2 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 2 u ($3/($12*1e-10)):($6/($1-1)**2):($7/($1-1)**2) title "$\\SI{1.5}{\\giga\\pascal}$" lt 3 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 2 u ($3/($12*1e-10)):($6/($1-1)**2) w l notitle lt 3 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 3 u ($3/($12*1e-10)):($6/($1-1)**2):($7/($1-1)**2) notitle lt 1 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 0 u ($3/($12*1e-10)):($6/($1-1)**2) w l title "C$_{16}$" lt 1 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 3 u ($3/($12*1e-10)):($6/($1-1)**2) w l title "C$_{30}$" lt 1 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 4 u ($3/($12*1e-10)):($6/($1-1)**2) w l notitle lt 2 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 4 u ($3/($12*1e-10)):($6/($1-1)**2):($7/($1-1)**2) notitle lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 5 u ($3/($12*1e-10)):($6/($1-1)**2) w l notitle lt 3 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 5 u ($3/($12*1e-10)):($6/($1-1)**2):($7/($1-1)**2) notitle lt 3 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 6 u ($3/($12*1e-10)):($6/($1-1)**2):($7/($1-1)**2) notitle lt 1 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 6 u ($3/($12*1e-10)):($6/($1-1)**2) w l title 'C$_{60}$' lt 1 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 7 u ($3/($12*1e-10)):($6/($1-1)**2) w l notitle lt 2 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 7 u ($3/($12*1e-10)):($6/($1-1)**2):($7/($1-1)**2) notitle lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 8 u ($3/($12*1e-10)):($6/($1-1)**2) w l notitle lt 3 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 8 u ($3/($12*1e-10)):($6/($1-1)**2):($7/($1-1)**2) notitle lt 3 lc 0 ps 2
\end{gnuplot}
\caption{Mean square end-to-end distance $\left(\left< R^2 \right> \right)$ per C-C bond as a function of the shear rate, $\dot{\gamma}$. Error bars, calculated from the standard deviation between block-averages, are omitted for clarity, but are of a similar size to the symbols.}
\label{fig:e2e2_v}
\end{center}
\end{figure}
\pgfmathsetmacro{\SERFigwidth}{.035\linewidth}
\pgfmathsetmacro{\SERFigheight}{.026\linewidth}
\begin{figure}[htp]
\begin{center}
\begin{gnuplot}[terminal=epslatex, terminaloptions={size \SERFigwidth cm, \SERFigheight cm color solid}]
set colors podo
set key above
set xlabel "$\\dot{\\gamma} \\, \\left[ \\SI{}{\\per\\second} \\right]$"
set ylabel "$\\left<P_{2}^{xz}\\right>$"
set key noautotitle
set format x '$10^{%L}$'
set logscale x
p [:2e10][] 'DataDump/Shear/Compiled.plot8' i 0 u ($3/($12*1e-10)):($8) title "$\\SI{0.5}{\\giga\\pascal}$" lt 1 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 1 u ($3/($12*1e-10)):($8) title "$\\SI{1.0}{\\giga\\pascal}$" lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 1 u ($3/($12*1e-10)):($8) w l notitle lt 2 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 2 u ($3/($12*1e-10)):($8) title "$\\SI{1.5}{\\giga\\pascal}$" lt 3 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 2 u ($3/($12*1e-10)):($8) w l notitle lt 3 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 3 u ($3/($12*1e-10)):($8) notitle lt 1 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 0 u ($3/($12*1e-10)):($8) w l title "C$_{16}$" lt 1 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 3 u ($3/($12*1e-10)):($8) w l title "C$_{30}$" lt 1 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 4 u ($3/($12*1e-10)):($8) w l notitle lt 2 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 4 u ($3/($12*1e-10)):($8) notitle lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 5 u ($3/($12*1e-10)):($8) w l notitle lt 3 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 5 u ($3/($12*1e-10)):($8) notitle lt 3 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 6 u ($3/($12*1e-10)):($8) notitle lt 1 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 6 u ($3/($12*1e-10)):($8) w l title 'C$_{60}$' lt 1 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 7 u ($3/($12*1e-10)):($8) w l notitle lt 2 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 7 u ($3/($12*1e-10)):($8) notitle lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 8 u ($3/($12*1e-10)):($8) w l notitle lt 3 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 8 u ($3/($12*1e-10)):($8) notitle lt 3 lc 0 ps 2
\end{gnuplot}
\caption{Average segmental orientation, $\left<P_{2}^{xz}\right>$, as a function of the shear rate, $\dot{\gamma}$. Error bars, calculated from the standard deviation between block-averages, are omitted for clarity, but are of a similar size to the symbols.}
\label{fig:P2_v}
\end{center}
\end{figure}
Figure \ref{fig:e2e2_v} shows that $\left< R^2 \right>/\left(N_\text{bonds}\right)^2$ generally decreases with increasing shear rate, suggesting less extended chain conformations. This has also been observed in previous bulk~\cite{Cui1996} and confined~\cite{Sivebaek2008,Cho2017} NEMD simulations of similar $n$-alkanes at high shear rates. This decrease has been attributed to strong intermolecular collisions which lead to intense chain rotation and tumbling dynamics at higher shear rates~\cite{Cho2017}. Longer chains are relatively less extended than shorter chains at all pressures and shear rates studied, as expected for $n$-alkanes in the absence of confinement and shear. Pressure has a negligible effect on $\left< R^2 \right>/\left(N_\text{bonds}\right)^2$ for the systems and condiions studied here.
More surprisingly, Figure \ref{fig:P2_v} shows that there is also a general decrease of $\left<P_{2}^{xz} \right> $ with increasing shear rate. This indicates an increase of the average angle of segments relative to the shearing direction at higher shear rate (see Appendix), meaning that the chains are less aligned with flow direction. Although such behaviour has also been observed in bulk NEMD simulations of $n$-alkanes at high shear rates~\cite{Padilla1992}, the opposite trend is more commonly reported~\cite{Cho2017,Cui1996}. This discrepancy can be rationalised through the high pressures applied in these simulations which lead to the fluids becoming solid-like, which is discussed further in the following section. Previous experiments~\cite{Drummond2002} and NEMD simulations~\cite{Cho2017} have shown that $n$-alkanes close to the confining walls align with the shear direction. Under the high pressures applied in these NEMD simulations, the ordered region extends further towards the centre of the film (Figure \ref{fig:VelProf_MDP2}), leading to a high value of $\left<P_{2}^{xz} \right> $. As the shear rate is increased, the film becomes more liquid-like (Figure \ref{fig:RDF}), and the thickness of the ordered region decreases (Figure \ref{fig:VelProf_MDP2}), and thus $\left<P_{2}^{xz} \right> $ also decreases. Consistent with previous bulk NEMD simulations~\cite{Cui1996}, longer chains tend be more aligned with the shearing direction, while the chains are slightly less aligned at higher pressure.
% Maybe have this section first
\subsection{Film structure and flow}
\label{str}
The radial distribution function (RDF), $g(r)$ for the C atoms was calculated using a cut-off distance of \SI{20}{\angstrom} (see Figure \ref{fig:RDF}). Note that C atoms separated from each other by one or two C-C bonds were excluded from the calculations.
Peak splitting in the RDFs is observed for all of the systems and conditions studied, which shows different ranges of order and is indicative of glassy behaviour~\cite{Bulacu2007,Valencia-Jaime2019}.
%
Three intramolecular RDF peaks are observed at C-C separations of: $d_1 \approx \SI{3.2}{\angstrom}$, $d_2 \approx \SI{3.9}{\angstrom}$ and $d_3 \approx \SI{5.1}{\angstrom}$ (see \ref{fig:RDF_Peaks}). The $d_1$ and $d_2$ peaks can be mostly attributed to C atoms separated by three bonds in $gauche$ and $trans$ conformations respectively~\cite{Bulacu2007}. The latter peak is much more intense, indicating that most of the chains are in extended conformations. The $d_3$ peak is mostly due to C atoms separated by four bonds, and its high intensity further suggests that the chains are in mostly extended conformations. These observations are consistent with the trends shown previously for $\left< R^2 \right>/\left(N_\text{bonds}\right)^2$ (Figure \ref{fig:e2e2_v}).
There are additional peaks at larger distances in Figure \ref{fig:RDF}, which are indicative of long-range order and thus solid-like films~\cite{Tsuchiya2001,Kalyanasundaram2009}. These peaks are more intense at higher pressure and lower sliding velocity; suggesting more solid-like films under these conditions. Moreover, comparing the size of the $d_1$ and $d_2$ peaks in Figure \ref{fig:RDF}, it is clear that there is a smaller trans/gauche ratio at high pressure and lower speed, a further indication of more solid-like films under these conditions~\cite{Kavitha2007}. The relative peak intensities also suggest that longer chains are more solid-like compared to shorter chains, which is expected from their higher melting points (C$_{16}$ = 291 K, C$_{30}$ = 339 K, C$_{60}$ = 373 K).
\pgfmathsetmacro{\SERFigwidth}{.035\linewidth}
\pgfmathsetmacro{\SERFigheight}{.04\linewidth}
\begin{figure}[htp]
\begin{center}
\begin{gnuplot}[terminal=epslatex, terminaloptions={size \SERFigwidth cm, \SERFigheight cm color solid}]
set colors podo
set multiplot layout 4,1 rowsfirst
set format x ' '
unset xlabel
set key samplen 0.0
set label at graph 0.03,0.89 left "a) 0.5 GPa, \\SI{10}{\\meter\\per\\second}"
# set lmargin at screen 0.25; set rmargin at screen 0.9
set tmargin at screen 1.00; set bmargin at screen 0.775
set ylabel "g(r)"
set key noautotitle
set xrange [0:20]
set yrange [0:2.5]
set ytics 0,1,4
plot 'DataDump/Shear/C16/0.5GPa/10m_s_RDF/all_c_rdf_s.dump' u ($2):($3/1.8) notitle w l lw 2 dt 1,\
'DataDump/Shear/C30/0.5GPa/10m_s_RDF/all_c_rdf_s.dump' u ($2):($3/1.8) notitle w l lw 2 dt 2,\
'DataDump/Shear/C60/0.5GPa/10m_s_RDF/all_c_rdf_s.dump' u ($2):($3/1.8) notitle w l lw 2 dt 3
unset label
set tmargin at screen 0.775; set bmargin at screen 0.550
set label at graph 0.03,0.89 left "b) 0.5 GPa, \\SI{100}{\\meter\\per\\second}"
set ylabel "g(r)"
set ytics 0,1,4
set key noautotitle
plot 'DataDump/Shear/C16/0.5GPa/100m_s_RDF/all_c_rdf_s.dump' u ($2):($3/1.8) notitle w l lw 2 dt 1,\
'DataDump/Shear/C30/0.5GPa/100m_s_RDF/all_c_rdf_s.dump' u ($2):($3/1.8) notitle w l lw 2 dt 2,\
'DataDump/Shear/C60/0.5GPa/100m_s_RDF/all_c_rdf_s.dump' u ($2):($3/1.8) notitle w l lw 2 dt 3
unset label
set tmargin at screen 0.550; set bmargin at screen 0.325
set label at graph 0.03,0.89 left "c) 1.5 GPa, 0.5 GPa, \\SI{10}{\\meter\\per\\second}"
set ylabel "g(r)"
set ytics 0,1,4
set key noautotitle
plot 'DataDump/Shear/C16/1.5GPa/10m_s_RDF/all_c_rdf_s.dump' u ($2):($3/1.8) notitle w l lw 2 dt 1,\
'DataDump/Shear/C30/1.5GPa/10m_s_RDF/all_c_rdf_s.dump' u ($2):($3/1.8) notitle w l lw 2 dt 2,\
'DataDump/Shear/C60/1.5GPa/10m_s_RDF/all_c_rdf_s.dump' u ($2):($3/1.8) notitle w l lw 2 dt 3
unset label
set format x '%g'
set xlabel "r [\\SI{}{\\angstrom}]"
set tmargin at screen 0.325; set bmargin at screen 0.10
set label at graph 0.03,0.89 left "d) 1.5 GPa, \\SI{100}{\\meter\\per\\second}"
set ylabel "g(r)"
set ytics 0,1,4
set key noautotitle
plot 'DataDump/Shear/C16/1.5GPa/100m_s_RDF/all_c_rdf_s.dump' u ($2):($3/1.8) title 'C$_{16}$' w l lw 2 dt 1,\
'DataDump/Shear/C30/1.5GPa/100m_s_RDF/all_c_rdf_s.dump' u ($2):($3/1.8) title 'C$_{30}$' w l lw 2 dt 2,\
'DataDump/Shear/C60/1.5GPa/100m_s_RDF/all_c_rdf_s.dump' u ($2):($3/1.8) title 'C$_{60}$' w l lw 2 dt 3
unset label
unset multiplot
\end{gnuplot}
\caption{Radial distribution function (RDF), $g(r)$, measured during steady state sliding for the three chain lengths studied: C$_{16}$, C$_{30}$ and C$_{60}$. a) \SI{0.5}{\giga\pascal}, \SI{10}{\meter\per\second}, b) \SI{0.5}{\giga\pascal}, \SI{100}{\meter\per\second}, c) \SI{1.5}{\giga\pascal}, \SI{10}{\meter\per\second} and d) \SI{1.5}{\giga\pascal}, \SI{100}{\meter\per\second}.
}
\label{fig:RDF}
\end{center}
\end{figure}
%\FloatBarrier
\begin{figure}[htp]
\begin{center}
\begin{tikzpicture}
\node[anchor=south west,inner sep=0] (fig_axes) at (0,0){\includegraphics[width=0.8\linewidth]{DataDump/Images/RDF_Peaks.png}};
%%%%
\filldraw[fill=SERgray, draw=black] (6,2) circle (0.385*3/4) node {C};
\filldraw[fill=SERwhite, draw=black] (5,2) circle (0.185) node {H};
\draw node[] (d1_A) at (3.6,-0.5) {};
\draw node[] (d1_B) at (4.92,-0.5) {};
\draw node[] (d2_A) at (1.52,-0.5) {};
\draw node[] (d3_A) at (1.4,1.99) {};
\draw node[] (d3_B) at (2.89,1.59) {};
\draw node[minimum width=10,minimum height=20,fill=white,rotate=53] (Cut1) at (0.069,0.85) {};
\draw node[minimum width=10,minimum height=20,fill=white,rotate=53] (Cut2) at (6.985,0.45) {};
\draw node[] (Cut1) at (0.0,0.5) {...};
\draw node[] (Cut2) at (7.05,0.5) {...};
\dimline[line style = {line width=0.7},extension start length=-0.35,extension end length=-0.35] {(d1_A)}{(d1_B)}{$d_1$};
\dimline[line style = {line width=0.7},extension start length=-0.24,extension end length=0.0] {(d2_A)}{(d1_A)}{$d_3$};
\dimline[line style = {line width=0.7},extension start length=0.54,extension end length=0.54] {(d3_A)}{(d3_B)}{$d_2$};
\end{tikzpicture}
\caption{Distances $d_1$, $d_2$ and $d_3$ between pairs of carbon atoms corresponding to the first three peaks of the radial distribution functions for all the systems. }
\label{fig:RDF_Peaks}
\end{center}
\end{figure}
Figure \ref{fig:VelProf_MDP2} shows the atomic mass density profile in the \emph{z}-dimension and \emph{x}-velocity profile in the \emph{z}-dimension profiles for the shortest and longest chains (C$_{16}$-C$_{60}$) and lowest and highest pressures (0.5-\SI{1.5}{\giga\pascal}) and sliding velocities (10-\SI{100}{\meter\per\second}) studied.
The mass density profiles in Figure \ref{fig:VelProf_MDP2} show the through-thickness layering of the fluids, which is strongest close to the walls. Generally, longer chains (C$_{60}$ in blue) show stronger layering compared to shorter chains (C$_{16}$ in purple), suggesting they form more solid-like films~\cite{Ewen2017a}. Comparing Figure \ref{fig:VelProf_MDP2}a with \ref{fig:VelProf_MDP2}b and Figure \ref{fig:VelProf_MDP2}c with \ref{fig:VelProf_MDP2}d shows that the films become less layered (particularly in the centre of the film), and thus more liquid-like at higher sliding velocity. Comparing Figure \ref{fig:VelProf_MDP2}a with \ref{fig:VelProf_MDP2}c and Figure \ref{fig:VelProf_MDP2}b with \ref{fig:VelProf_MDP2}d shows that the films become more layered and thus more solid-like at higher pressure.
At the onset of shear, the velocity profiles show deviations from planar Couette flow for several of the systems and conditions studied; however, in most cases, a Couette profile develops once a nonequilibrium steady state is reached (see example in Appendix). Figure \ref{fig:VelProf_MDP2} shows steady state velocity profiles for the lowest and highest chain length, pressure, and sliding velocity investigated. At low sliding velocity and low pressure (Figure \ref{fig:VelProf_MDP2}a), the flow for C$_{16}$ (purple) remains Couette-like, whereas C$_{60}$ (blue) shows some central localisation (CL). CL describes the outer solid-like regions of the fluid moving at the same velocity as the neighbouring walls, with only the centre of the film being sheared~\cite{Heyes2012}. The transition from Couette flow to CL has been observed for viscous polymers using in-contact photoluminescence experiments under EHL conditions~\cite{Galmiche2016}. It has also been observed in previous NEMD simulations of atomic fluids~\cite{Heyes2012,Gattinoni2013,Mackowiak2016} as well as alkanes and traction fluids~\cite{Ewen2017a} under EHL conditions. At high sliding velocity and low pressure (Figure \ref{fig:VelProf_MDP2}b), Couette flow is maintained for C$_{16}$ and while C$_{60}$ shows a linear velocity profile, there is a small amount of boundary slip at the fluid-wall interface (slip length $\leq$ 1 nm). Boundary slip has also been observed in previous NEMD simulations of thin fluid films~\cite{Jabbarzadeh1999,Sivebaek2010,Ta2017,Porras-Vazquez2018}. However, NEMD simulations of thicker films more commonly show no-slip boundary conditions~\cite{Washizu2010,Washizu2014}. At low sliding velocity and high pressure, (Figure \ref{fig:VelProf_MDP2}c), C$_{16}$ shows Couette flow while C$_{60}$ shows CL. Figure \ref{fig:VelProf_MDP2}d shows that C$_{16}$ also slips at high pressure ($\approx$ 10 nm). Moreover, the slip length increases ($\approx$ 40 nm) with increasing pressure for C$_{60}$, which is consistent with previous NEMD simulations~\cite{Ta2017} and in-contact photoluminescence experiments~\cite{Ponjavic2014} of different fluids.
%-1.39716 , 81.0918
%-1.39716 , 81.0918
%-1.51672, 75.8545
%-1.51672, 75.8545
%-1.34156, 81.2584
%-1.34156, 81.2584
%-1.47486, 72.8981
%-1.47486, 72.8981
%-1.5611, 70.1875
%-1.5611, 70.1875
%-1.55172, 74.2276
%-1.55172, 74.2276
%-1.5611, 70.1875
%-1.5611, 70.1875
\pgfmathsetmacro{\SERFigwidth}{.035\linewidth}
\pgfmathsetmacro{\SERFigheight}{.02\linewidth}
\begin{figure}[htp]
\begin{center}
\begin{gnuplot}[terminal=epslatex, terminaloptions={size \SERFigwidth cm, \SERFigheight cm color solid}]
set colors podo
binwidth=2
binwidth2=4
bin(x,width)=width*floor(x/width)
set format y2 '%g'
set label at graph 0.03,0.89 left "a)"
set ylabel "x Velocity $\\left[ \\SI{}{\\meter\\per\\second} \\right]$"
set y2label "AMD $\\left[ \\SI{}{\\gram\\per\\cubic\\centi\\meter} \\right]$ "
set key noautotitle
set xrange [-50:50]
set y2range [0:3]
set ytics -5,2.5,5
set ytics nomirror
set y2tics 0,1,4
set xlabel "z coordinate [\\SI{}{\\angstrom}]"
plot [][-6:6] 'DataDump/Shear/C16/0.5GPa/10m_s/fluid_mdp_z_s.dump1' u ($2-39.84732):($4) w l lt 1 axes x1y2,\
'DataDump/Shear/C16/0.5GPa/10m_s/vel_prof_x_s.dump1' u (bin($2-39.84732,binwidth2)):($4*100) smooth unique lt 1 dashtype 3 lw 5,\
'DataDump/Shear/C60/0.5GPa/10m_s/fluid_mdp_z_s.dump1' u ($2-37.16889):($4) w l lt 3 axes x1y2,\
'DataDump/Shear/C60/0.5GPa/10m_s/vel_prof_x_s.dump1' u (bin($2-37.16889,binwidth)):($4*100) smooth unique lt 3 dashtype 3 lw 5
\end{gnuplot}
\begin{gnuplot}[terminal=epslatex, terminaloptions={size \SERFigwidth cm, \SERFigheight cm color solid}]
set colors podo
binwidth=2
bin(x,width)=width*floor(x/width)
set format y2 '%g'
set label at graph 0.03,0.89 left "b)"
set ylabel "x Velocity $\\left[ \\SI{}{\\meter\\per\\second} \\right]$"
set y2label "AMD $\\left[ \\SI{}{\\gram\\per\\cubic\\centi\\meter} \\right]$ "
set key noautotitle
set xrange [-50:50]
set y2range [0:3]
set ytics -50,25,50
set ytics nomirror
set y2tics 0,1,4
set xlabel "z coordinate [\\SI{}{\\angstrom}]"
plot [][-60:60] 'DataDump/Shear/C16/0.5GPa/100m_s/fluid_mdp_z_s.dump' u ($2-41):($4) w l lt 1 axes x1y2,\
'DataDump/Shear/C60/0.5GPa/100m_s/vel_prof_x_s.dump' u (bin($2-41,binwidth)):($4*100) smooth unique lt 1 dashtype 3 lw 5,\
'DataDump/Shear/C60/0.5GPa/100m_s/fluid_mdp_z_s.dump' u ($2-39.11):($4) w l lt 3 axes x1y2,\
'DataDump/Shear/C16/0.5GPa/100m_s/vel_prof_x_s.dump' u (bin($2-39.11,binwidth)):($4*100) smooth unique lt 3 dashtype 3 lw 5
\end{gnuplot}
\begin{gnuplot}[terminal=epslatex, terminaloptions={size \SERFigwidth cm, \SERFigheight cm color solid}]
set colors podo
binwidth=4
binwidth2=3
bin(x,width)=width*floor(x/width)
set format y2 '%g'
set label at graph 0.03,0.89 left "c)"
set ylabel "x Velocity $\\left[ \\SI{}{\\meter\\per\\second} \\right]$"
set y2label "AMD $\\left[ \\SI{}{\\gram\\per\\cubic\\centi\\meter} \\right]$ "
set key noautotitle
set xrange [-50:50]
set y2range [0:3]
set ytics -5,2.5,5
set ytics nomirror
set y2tics 0,1,4
set xlabel "z coordinate [\\SI{}{\\angstrom}]"
plot [][-6:6] 'DataDump/Shear/C16/1.5GPa/10m_s/fluid_mdp_z_s.dump1' u ($2-35.7116):($4) w l lt 1 axes x1y2,\
'DataDump/Shear/C16/1.5GPa/10m_s/vel_prof_x_s.dump1' u (bin($2-35.7116,binwidth2)):($4*100) smooth unique lt 1 dashtype 3 lw 5,\
'DataDump/Shear/C60/1.5GPa/10m_s/fluid_mdp_z_s.dump1' u ($2-34.3132):($4) w l lt 3 axes x1y2,\
'vel_prof_x_s.dump3' u (bin($2-34.3132,binwidth)):($4*100) smooth unique lt 3 dashtype 3 lw 5
\end{gnuplot}
\begin{gnuplot}[terminal=epslatex, terminaloptions={size \SERFigwidth cm, \SERFigheight cm color solid}]
set key top center
set colors podo
binwidth=2
bin(x,width)=width*floor(x/width)
set format y2 '%g'
set label at graph 0.03,0.89 left "d)"
set ylabel "x Velocity $\\left[ \\SI{}{\\meter\\per\\second} \\right]$"
set y2label "AMD $\\left[ \\SI{}{\\gram\\per\\cubic\\centi\\meter} \\right]$ "
set key noautotitle
set xrange [-50:50]
set y2range [0:3]
set ytics -50,25,50
set ytics nomirror
set y2tics 0,1,4
set xlabel "z coordinate [\\SI{}{\\angstrom}]"
plot [][-60:60] 'DataDump/Shear/C16/1.5GPa/100m_s/fluid_mdp_z_s.dump' u ($2-36.3379):($4) w l lt 1 axes x1y2 title "C$_{16}$",\
'DataDump/Shear/C16/1.5GPa/100m_s/vel_prof_x_s.dump' u (bin($2-36.3379,binwidth)):($4*100) smooth unique lt 1 dashtype 3 lw 5,\
'DataDump/Shear/C60/1.5GPa/100m_s/fluid_mdp_z_s.dump1' u ($2-34.8132):($4) w l lt 3 axes x1y2 title "C$_{60}$",\
'DataDump/Shear/C60/1.5GPa/100m_s/vel_prof_x_s.dump1' u (bin($2-34.8132,binwidth)):($4*100) smooth unique lt 3 dashtype 3 lw 5
\end{gnuplot}
\caption{Velocity (solid lines) and atomic mass density (dotted lines) during steady state sliding for two chain lenghts: C$_{16}$ and C$_{60}$. a) \SI{0.5}{\giga\pascal}, \SI{10}{\meter\per\second}, b) \SI{0.5}{\giga\pascal}, \SI{100}{\meter\per\second}, c) \SI{1.5}{\giga\pascal}, \SI{10}{\meter\per\second} and d) \SI{1.5}{\giga\pascal}, \SI{100}{\meter\per\second}}
\label{fig:VelProf_MDP2}
\end{center}
\end{figure}
\subsection{Friction}
\label{fri}
\textcolor{blue}{Figure \ref{fig:FL_FN} shows that shear stress increases with increasing pressure. This is in agreement with the Amontons-Coulomb friction law under the high load approximation; $F_L/F_N = F_0/F_N + \mu$, where $F_L$ and $F_N$ are respectively the mean lateral force and normal force acting on the outer layer of atoms in each slab in response to the fluid, $\mu$ is the friction coefficient, and $F_0$ is the load-independent Derjaguin offset, which represents adhesive surface forces~\cite{Eder2013,Ewen2017}. The $\mu$ and $F_0$ were respectively calculated from the gradient and intercept of the linear fits in Figure \ref{fig:FL_FN} for the different systems and conditions studied. Friction anisotropy has been observed in previous experiments~\cite{Kristiansen2012} and NEMD simulations~\cite{Jabbarzadeh2016} of strongly-confined $n$-alkanes, but was not observed for the relatively thick films studied here.}
\pgfmathsetmacro{\SERFigwidth}{.035\linewidth}
\pgfmathsetmacro{\SERFigheight}{.026\linewidth}
\begin{figure}[htp]
\begin{center}
\begin{gnuplot}[terminal=epslatex, terminaloptions={size \SERFigwidth cm, \SERFigheight cm color solid}]
set colors podo
set key above
set xlabel "$F_N \\, [\\SI{}{\\mega\\pascal}]$"
set ylabel "$F_L \\, [\\SI{}{\\mega\\pascal}]$"
set ytics -20,20,120
set key noautotitle
p [0:1600][-25:] 'DataDump/Shear/Compiled_p.plot8' i 0 u ($2*1000):($4/10) w p notitle lt 1 lc 1 ps 2,\
'DataDump/Shear/Compiled_p.plot8' i 0 u (NaN):(NaN) w p title "$\\SI{10}{\\meter\\per\\second}$" lt 1 lc 0 ps 2,\
'DataDump/Shear/Compiled_p.plot8' i 1 u ($2*1000):($4/10) w p notitle lt 2 lc 1 ps 2,\
'DataDump/Shear/Compiled_p.plot8' i 1 u (NaN):(NaN) w p title "$\\SI{20}{\\meter\\per\\second}$" lt 2 lc 0 ps 2,\
0.0825796*x+(-10.8789) w l notitle lt 1 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled_p.plot8' i 2 u ($2*1000):($4/10) w p notitle lt 3 lc 1 ps 2,\
'DataDump/Shear/Compiled_p.plot8' i 2 u (NaN):(NaN) w p title "$\\SI{50}{\\meter\\per\\second}$" lt 3 lc 0 ps 2,\
0.0813269*x+(1.62006) w l notitle lt 1 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled_p.plot8' i 3 u ($2*1000):($4/10) w p notitle lt 4 lc 1 ps 2,\
'DataDump/Shear/Compiled_p.plot8' i 3 u (NaN):(NaN) w p title "$\\SI{100}{\\meter\\per\\second}$" lt 4 lc 0 ps 2,\
0.0793452*x+(10.9365) w l notitle lt 1 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled_p.plot8' i 4 u ($2*1000):($4/10) w p notitle lt 1 lc 2 ps 2,\
0.0831989*x+(-22.0636) w l title "C$_{16}$" lt 1 lc 1 lw 2 dt 1,\
0.0827557*x+(-7.60443) w l title "C$_{30}$" lt 1 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled_p.plot8' i 5 u ($2*1000):($4/10) w p notitle lt 2 lc 2 ps 2,\
0.0839157*x+(-2.96795) w l notitle lt 1 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled_p.plot8' i 6 u ($2*1000):($4/10) w p notitle lt 3 lc 2 ps 2,\
0.0813649*x+(9.62516) w l notitle lt 1 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled_p.plot8' i 7 u ($2*1000):($4/10) w p notitle lt 4 lc 2 ps 2,\
0.0795552*x+(16.7165) w l notitle lt 1 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled_p.plot8' i 8 u ($2*1000):($4/10) w p notitle lt 1 lc 3 ps 2,\
0.0867396*x+(-4.45137) w l title "C$_{60}$" lt 1 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled_p.plot8' i 9 u ($2*1000):($4/10) w p notitle lt 2 lc 3 ps 2,\
0.0823248*x+(4.65988) w l notitle lt 1 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled_p.plot8' i 10 u ($2*1000):($4/10) w p notitle lt 3 lc 3 ps 2,\
0.0765719*x+(18.4519) w l notitle lt 1 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled_p.plot8' i 11 u ($2*1000):($4/10) w p notitle lt 4 lc 3 ps 2,\
0.0781266*x+(21.6555) w l notitle lt 1 lc 3 lw 2 dt 3
\end{gnuplot}
\caption{Lateral (friction) force, $F_L$, as a function of the applied pressure, $F_N$, on the outer layer of atoms in the top and bottom slabs during steady state sliding.}
\label{fig:FL_FN}
\end{center}
\end{figure}
%\FloatBarrier
\pgfmathsetmacro{\SERFigwidth}{.035\linewidth}
\pgfmathsetmacro{\SERFigheight}{.026\linewidth}
\begin{figure}[htp]
\begin{center}
\begin{gnuplot}[terminal=epslatex, terminaloptions={size \SERFigwidth cm, \SERFigheight cm color solid}]
set key above
set colors podo
set ylabel "$F_L \\, [\\SI{}{\\mega\\pascal}]$"
set ytics 0,20,150
set key noautotitle
set xlabel "$\\dot{\\gamma} \\, \\left[ \\SI{}{\\per\\second} \\right]$"
set format x '$10^{%L}$'
set logscale x
p [:2e10][0:150] 'DataDump/Shear/Compiled.plot8' i 0 u ($3/($12*1e-10)):($4/10):($5/10) w p title "$\\SI{0.5}{\\giga\\pascal}$" lt 1 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 1 u ($3/($12*1e-10)):($4/10):($5/10) w p title "$\\SI{1.0}{\\giga\\pascal}$" lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 1 u ($3/($12*1e-10)):($4/10) w l notitle lt 2 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 2 u ($3/($12*1e-10)):($4/10):($5/10) w p title "$\\SI{1.5}{\\giga\\pascal}$" lt 3 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 2 u ($3/($12*1e-10)):($4/10) w l notitle lt 3 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 3 u ($3/($12*1e-10)):($4/10):($5/10) w p notitle lt 1 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 0 u ($3/($12*1e-10)):($4/10) w l title "C$_{16}$" lt 1 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 3 u ($3/($12*1e-10)):($4/10) w l title "C$_{30}$" lt 1 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 4 u ($3/($12*1e-10)):($4/10) w l notitle lt 2 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 4 u ($3/($12*1e-10)):($4/10):($5/10) w p notitle lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 5 u ($3/($12*1e-10)):($4/10) w l notitle lt 3 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 5 u ($3/($12*1e-10)):($4/10):($5/10) w p notitle lt 3 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 6 u ($3/($12*1e-10)):($4/10):($5/10) w p notitle lt 1 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 6 u ($3/($12*1e-10)):($4/10) w l title 'C$_{60}$' lt 1 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 7 u ($3/($12*1e-10)):($4/10) w l notitle lt 2 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 7 u ($3/($12*1e-10)):($4/10):($5/10) w p notitle lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 8 u ($3/($12*1e-10)):($4/10) w l notitle lt 3 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 8 u ($3/($12*1e-10)):($4/10):($5/10) w p notitle lt 3 lc 0 ps 2
\end{gnuplot}
\caption{Shear stress, $F_L$, versus logarithmic shear rate, $\dot{\gamma}$, during steady state sliding. Error bars, calculated from the standard deviation between the trajectory time-averages, are omitted for clarity, but are of a similar size to the symbols.}
\label{fig:FL_v1a}
\end{center}
\end{figure}
\pgfmathsetmacro{\SERFigwidth}{.035\linewidth}
\pgfmathsetmacro{\SERFigheight}{.026\linewidth}
\begin{figure}[htp]
\begin{center}
\begin{gnuplot}[terminal=epslatex, terminaloptions={size \SERFigwidth cm, \SERFigheight cm color solid}]
set key above
set colors podo
set ylabel "$\\mu$"
set key noautotitle
set ytics 0.04,0.01,0.2
set xlabel "$\\dot{\\gamma} \\, \\left[ \\SI{}{\\per\\second} \\right]$"
set format x '$10^{%L}$'
set logscale x
p [:2e10][0.04:0.13] 'DataDump/Shear/Compiled.plot8' i 0 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w p title "$\\SI{0.5}{\\giga\\pascal}$" lt 1 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 1 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w p title "$\\SI{1.0}{\\giga\\pascal}$" lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 1 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w l notitle lt 2 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 2 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w p title "$\\SI{1.5}{\\giga\\pascal}$" lt 3 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 2 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w l notitle lt 3 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 3 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w p notitle lt 1 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 0 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w l title "C$_{16}$" lt 1 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 3 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w l title "C$_{30}$" lt 1 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 4 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w l notitle lt 2 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 4 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w p notitle lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 5 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w l notitle lt 3 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 5 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w p notitle lt 3 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 6 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w p notitle lt 1 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 6 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w l title 'C$_{60}$' lt 1 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 7 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w l notitle lt 2 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 7 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w p notitle lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 8 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w l notitle lt 3 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 8 u ($3/($12*1e-10)):(($4/10)/($2*1000)) w p notitle lt 3 lc 0 ps 2
\end{gnuplot}
\caption{Friction coefficient $\mu$ as a function of logarithmic shear rate, $\dot{\gamma}$, during steady state sliding.}
\label{fig:FL_va}
\end{center}
\end{figure}
\pgfmathsetmacro{\SERFigwidth}{.035\linewidth}
\pgfmathsetmacro{\SERFigheight}{.026\linewidth}
\begin{figure}[htp]
\begin{center}
\begin{gnuplot}[terminal=epslatex, terminaloptions={size \SERFigwidth cm, \SERFigheight cm color solid}]
set key above
set colors podo
set ylabel "$F_0 \\, [\\SI{}{\\mega\\pascal}]$"
set key noautotitle
set xlabel "$\\dot{\\gamma} \\, \\left[ \\SI{}{\\per\\second} \\right]$"
set format x '$10^{%L}$'
set logscale x
p [:2e10][] 'DataDump/Shear/Compiled.plot8' i 0 u ($3/($12*1e-10)):(($14)) w p notitle lt 1 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 0 u ($3/($12*1e-10)):(($14)) w l title "C$_{16}$" lt 1 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 3 u ($3/($12*1e-10)):(($14)) w p notitle lt 1 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 3 u ($3/($12*1e-10)):(($14)) w l title "C$_{30}$" lt 1 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 6 u ($3/($12*1e-10)):(($14)) w p notitle lt 1 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 6 u ($3/($12*1e-10)):(($14)) w l title 'C$_{60}$' lt 1 lc 3 lw 2 dt 3
\end{gnuplot}
\caption{Derjaguin offset $F_0$ as a function of logarithmic shear rate, $\dot{\gamma}$, during steady state sliding.}
\label{fig:Derj}
\end{center}
\end{figure}
Figure \ref{fig:FL_va} shows the change in the shear stress, $F_L$, with logarithmic sliding velocity and Figure \ref{fig:FL_va} shows the same data as a friction coefficient, $\mu$. In general, longer chains give higher $F_L$ than shorter chains in \ref{fig:FL_va}, as expected due to their higher bulk viscosity~\cite{Zhang2017}. This trend has also been observed in previous NEMD simulations of thinner films at lower pressure~\cite{Koike1998,Sivebaek2008,Sivebaek2010,Savio2012}. The difference in $F_L$ between chain lengths decreases with increasing shear rate, such that they are almost identical at the highest shear rate studied ($\approx \SI{e10}{\per\second}$). It is important to note that tribological experiments~\cite{Ewen2017a,Zhang2017} and engineering components~\cite{Taylor2017} operate at lower shear rates than those accessible through NEMD simulations~\cite{Ewen2018}. Thus a larger increase in $F_L$ with increasing chain length can be expected under these lower shear rates than was observed here.
Generally, $F_L$ increases linearly with logarithmic shear rate in Figure \ref{fig:FL_va}, as is commonly observed for lubricants in macroscopic EHL friction experiments~\cite{Spikes2014,Ewen2017a}. Such behaviour has also been reported in previous NEMD simulations of thin (6 molecular layers) C$_{20}$ films between polymer-like surfaces at low pressure (\SI{10}{\mega\pascal})~\cite{Sivebaek2010}. In Figure \ref{fig:FL_va}, the slope of the increase in $F_L$ with shear rate generally decreases with increasing pressure. This has also been observed in NEMD simulations of linear~\cite{Ta2017} and branched~\cite{Ewen2017a} alkanes under similar conditions.
In Figure \ref{fig:Derj}, $F_0$ increases with linearly with logarithmic shear rate for all of the $n$-alkanes studied. ~\cite{Sivebaek2016}
Non-zero values for $F_0$ occur when the lubricant is less ordered.~\cite{Eder2013}
Surface effects~\cite{Ta2015a,Ta2015b}.
$n$-hexadecane~\cite{Jabbarzadeh1998} and ~\cite{Tseng2009}.
In Figure \ref{fig:FL_v1a}, $\mu$ increases with linearly with logarithmic shear rate at low pressure (0.5 GPa) for all of the $n$-alkanes studied. At higher pressure (1.0-1.5 GPa), $\mu$ still increases with shear rate for C$_{16}$, but with a shallower slope. As a result, at high shear rates, $\mu$ is higher at low pressure than at high pressure. For the longer alkanes (C$_{30}$-C$_{60}$) at high pressure, $\mu$ becomes insensitive to shear rate, in line with Coulomb's law for dry friction. The transition to this behaviour, above the limiting shear stress (LSS), has been observed for several model lubricants in macroscopic friction experiments~\cite{Martinie2016a}. It has also been observed for bulky traction fluid molecules in NEMD simulations~\cite{Ewen2017a,Porras-Vazquez2018}. Similar behaviour has also been reported in previous NEMD simulations of thin C$_{60}$ films between polymer-like surfaces at low pressure (\SI{10}{\mega\pascal}) which showed boundary slip~\cite{Sivebaek2010}.
LSS behaviour has been attributed to a range of effects, including the glass transition~\cite{Porras-Vazquez2018} and the consequent change in flow behaviour.~\cite{Ewen2017a} In this study, non-linear flow profiles occur at the onset of shear for several of the systems and conditions studied; however, a Couette profile usually develops once a nonequilibrium steady state is reached (see Appendix). However, non-linear flow persists into the steady state for some of the systems and conditions studied (Figure \ref{fig:VelProf_MDP2}). For example, boundary slip occurs at the highest shear rates studied (Figure \ref{fig:VelProf_MDP2}) for C$_{60}$, which has been shown to reduce $\mu$~\cite{Sivebaek2010,Savio2012}. Moreover, the longer $n$-alkanes (C$_{30}$-C$_{60}$) show CL at high pressure, which has also been linked to the LSS~\cite{Ewen2017a}. However, LSS behaviour is also observed in these simulations for systems and conditions in which the flow remains Couette-like.
EHL conditions can lead to large temperature rises in the fluid film as well as at the sliding surfaces~\cite{Spikes2014}. Figure \ref{fig:T_v} shows the change in average film temperature (details in Appendix) with logarithmic shear rate. At the lowest shear rate studied, the temperature in the fluid region is the same as the thermostatted surfaces (\SI{353}{\kelvin}). As predicted by the Archard equation for EHL temperature rise~\cite{Archard1959}, there is a linear increase in temperature with increasing sliding velocity (inset in Figure~\ref{fig:T_v}). The temperature rise is insensitive to chain length and pressure in the ranges studied. Under similar conditions, $\Delta$T is significantly larger than in previous NEMD simulations of thinner $n$-alkane films~\cite{Berro2011,Ta2017}. Moreover, the `critical shear rate' at which a detectable $\Delta$T occurs is much lower than in these simulations. These observations are expected since thicker films mean that phonons generated in the centre of the film during shearing have further to travel before they can be dissipated through the thermostatted surfaces. Indeed, the Archard equation predicts a linear increase in temperature rise with film thickness~\cite{Archard1959}. The $\Delta$T values in Figure~\ref{fig:T_v} are smaller than in recent NEMD simulations of thicker films of branched and cyclic alkanes which give higher $\tau$~\cite{Ewen2019}.
The increase of the temperature at high shear rates will decrease the fluid viscosity which will in turn reduce friction. However, the rather modest $\Delta$T values in Figure \ref{fig:T_v} are insufficient to explain the drastic changes in friction behaviour shown in Figure \ref{fig:FL_v1a}~\cite{Berro2011,Ewen2019}. However, the solid-like films at high pressure become more liquid-like at high shear rates. The reduction in $\left<P_{2}^{xz} \right> $ and $\left< R^2 \right>$ with increasing shear rate (Figures \ref{fig:e2e2_v} and \ref{fig:P2_v}) indicate that the chains are less elongated and less orientated with the flow direction, which is indicative of more liquid-like behaviour. Moreover, the radial distribution functions (RDFs) shown in Figure \ref{fig:RDF}) indicate that there is less long range order at higher shear rates, further suggesting more liquid-like films. Thus, in addition to changes in flow behaviour, shear heating and the consequent transition from solid-like to liquid-like films with increasing shear rate could also contribute to the LSS behaviour observed in Figure \ref{fig:FL_v1a}.
\pgfmathsetmacro{\SERFigwidth}{.035\linewidth}
\pgfmathsetmacro{\SERFigheight}{.03\linewidth}
\begin{figure}[htp]
\begin{center}
\begin{gnuplot}[terminal=epslatex, terminaloptions={size \SERFigwidth cm, \SERFigheight cm color solid}]
set multiplot
set colors podo
set key above
set xlabel "$\\dot{\\gamma} \\, \\left[ \\SI{}{\\per\\second} \\right]$"
set ylabel "$\\Delta T \\, [\\SI{}{\\kelvin}]$"
set format x '$10^{%L}$'
set key noautotitle
set log x
set yrange [-5:70]
set ytics 00,20,100
p [:2e10][] 'DataDump/Shear/Compiled.plot8' i 0 u ($3/($12*1e-10)):($10-336.663383258371) w p title "$\\SI{0.5}{\\giga\\pascal}$" lt 1 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 1 u ($3/($12*1e-10)):($10-328.661940829585) w p title "$\\SI{1.0}{\\giga\\pascal}$" lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 1 u ($3/($12*1e-10)):($10-328.661940829585) w l notitle lt 2 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 2 u ($3/($12*1e-10)):($10-330.68385912044) w p title "$\\SI{1.5}{\\giga\\pascal}$" lt 3 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 2 u ($3/($12*1e-10)):($10-330.68385912044) w l notitle lt 3 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 3 u ($3/($12*1e-10)):($10-317.877393053473) w p notitle lt 1 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 0 u ($3/($12*1e-10)):($10-336.663383258371) w l title "C$_{16}$" lt 1 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 3 u ($3/($12*1e-10)):($10-317.877393053473) w l title "C$_{30}$" lt 1 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 4 u ($3/($12*1e-10)):($10-324.873226886557) w l notitle lt 2 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 4 u ($3/($12*1e-10)):($10-324.873226886557) w p notitle lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 5 u ($3/($12*1e-10)):($10-327.756576261869) w l notitle lt 3 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 5 u ($3/($12*1e-10)):($10-327.756576261869) w p notitle lt 3 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 6 u ($3/($12*1e-10)):($10-320.835055572214) w p notitle lt 1 lc 0 ps 2 ,\
'DataDump/Shear/Compiled.plot8' i 6 u ($3/($12*1e-10)):($10-320.835055572214) w l title 'C$_{60}$' lt 1 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 7 u ($3/($12*1e-10)):($10-323.253763818091) w l notitle lt 2 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 7 u ($3/($12*1e-10)):($10-323.253763818091) w p notitle lt 2 lc 0 ps 2,\
'DataDump/Shear/Compiled.plot8' i 8 u ($3/($12*1e-10)):($10-326.935222638681) w l notitle lt 3 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 8 u ($3/($12*1e-10)):($10-326.935222638681) w p notitle lt 3 lc 0 ps 2,\
set size 0.45,0.4
set origin 0.15,0.4
set format x
set lmargin 5
set xlabel "x Velocity $\\left[ \\SI{}{\\meter\\per\\second} \\right]$"
set ylabel
unset logscale
set ytics 0,30,100
set xtics 0,25,100
p [0:110][] 'DataDump/Shear/Compiled.plot8' i 0 u ($3):($10-336.663383258371) w p notitle lt 1 lc 0 ps 1,\
'DataDump/Shear/Compiled.plot8' i 1 u ($3):($10-328.661940829585) w p notitle lt 2 lc 0 ps 1,\
'DataDump/Shear/Compiled.plot8' i 1 u ($3):($10-328.661940829585) w l notitle lt 2 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 2 u ($3):($10-330.68385912044) w p notitle lt 3 lc 0 ps 1,\
'DataDump/Shear/Compiled.plot8' i 2 u ($3):($10-330.68385912044) w l notitle lt 3 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 3 u ($3):($10-317.877393053473) w p notitle lt 1 lc 0 ps 1 ,\
'DataDump/Shear/Compiled.plot8' i 0 u ($3):($10-336.663383258371) w l notitle lt 1 lc 1 lw 2 dt 1,\
'DataDump/Shear/Compiled.plot8' i 3 u ($3):($10-317.877393053473) w l notitle lt 1 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 4 u ($3):($10-324.873226886557) w l notitle lt 2 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 4 u ($3):($10-324.873226886557) w p notitle lt 2 lc 0 ps 1,\
'DataDump/Shear/Compiled.plot8' i 5 u ($3):($10-327.756576261869) w l notitle lt 3 lc 2 lw 2 dt 2,\
'DataDump/Shear/Compiled.plot8' i 5 u ($3):($10-327.756576261869) w p notitle lt 3 lc 0 ps 1 ,\
'DataDump/Shear/Compiled.plot8' i 6 u ($3):($10-320.835055572214) w p notitle lt 1 lc 0 ps 1 ,\
'DataDump/Shear/Compiled.plot8' i 6 u ($3):($10-320.835055572214) w l notitle lt 1 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 7 u ($3):($10-323.253763818091) w l notitle lt 2 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 7 u ($3):($10-323.253763818091) w p notitle lt 2 lc 0 ps 1,\
'DataDump/Shear/Compiled.plot8' i 8 u ($3):($10-326.935222638681) w l notitle lt 3 lc 3 lw 2 dt 3,\
'DataDump/Shear/Compiled.plot8' i 8 u ($3):($10-326.935222638681) w p notitle lt 3 lc 0 ps 1
unset multiplot
\end{gnuplot}
\caption{Average temperature rise of the fluid region, $\Delta$T, as a function of logarithmic shear rate, $\dot{\gamma}$. Inset shows the same data versus linear sliding velocity, v$_s$. Error bars, calculated from the standard deviation between the trajectory time-averages, are omitted for clarity, but are of a similar size to the symbols.}
\label{fig:T_v}
\end{center}
\end{figure}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Conclusions}
\label{sec:Conc}
In this study, the behaviour of $n$-alkane films confined ($\geq 16$ molecular layers) and sheared between atomically-smooth hematite slabs has been investigated under EHL conditions using NEMD simulations. The effect of the alkane chain length (C$_{16}$-C$_{60}$), pressure ($0.5$-$1.5~\SI{}{\giga\pascal}$), and shear rate ($10^{9}$-$10^{10}~\SI{}{\per\second}$) on the chain extension and orientation, film structure and flow, shear heating, and friction have been studied. An accurate all-atom force-field was used to ensure that the viscosity was close to the experimental value.
Starting from ordered chains, a heat-quench cycle was used to equilibrate the systems by monitoring the end-to-end distance and radius of gyration. It was also ensured that the chains moved at least their own size by monitoring the mean square displacement. When shear was applied, the end-to-end distance, segmental orientation, and friction were monitored to determine when a nonequilibrium steady state had been attained. The simulation results were in agreement with previous experiments which suggested that the evolution to steady state sliding in $n$-alkane films is governed by the sliding distance rather than the sliding time.
At higher pressure, the films show more layering and ordering in the mass density profiles and RDF respectively, suggesting that they become more solid-like. Conversely, at higher shear rates, the chains become less elongated, aligned, layered, and ordered; indicating the films become more liquid-like as the fluid film temperature rise increases. Longer chains are generally more solid-like, in accordance with their higher melting point.
For all of the systems and conditions studied, the shear stress increases linearly with increasing pressure. Longer chains, with higher viscosity give higher friction, particularly at low shear rates where the films remain solid-like. The shear stress is more sensitive to pressure than to chain length in the ranges studied. For short chains, the flow remains mostly Couette-like, but some boundary slip occurs at the highest pressures and shear rates studied. Conversely, longer chains consistently show non-linear flow (both boundary slip and CL) even under comparatively mild conditions.
At low pressure, the friction coefficient increases linearly with logarithmic shear rate, as is commonly observed for lubricants in experiments and simulations. For short chains, the slope of this increase becomes shallower as the pressure increases. For longer chains, the friction coefficient becomes insensitive to shear rate at high pressure, suggesting that the LSS was reached. A detailed study of the fluid properties reveals that a combination of non-linear flow and shear-induced melting are responsible for this behaviour for the systems and conditions studied here.
The results of this study suggest that $n$-alkanes can show deviations from the commonly assumed flow and friction behaviour at high pressure and shear rate. These deviations are more significant for longer $n$-alkanes, which have higher viscosity and melting points.
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section*{Acknowledgments}
S.E.R. acknowledges the support of the European Commission through the Marie Curie Industry-Academia Partnerships and Pathways (IAPP) iBETTER Project: \url{http://cordis.europa.eu/project/rcn/109976_en.html}. J.P.E acknowledges the Engineering and Physical Sciences Research Council (EPSRC) for funding via a Doctoral Prize Fellowship and grant EP/P030211/1. The figures showing atomic configurations where generated using the OVITO software~\cite{Stukowski2010b}. Some of the simulations were facilitated through the Imperial College London Research Computing Service (RCS).
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section*{Bibliography}
\bibliography{MyCollection}
%\bibliographystyle{abbrv}
%\bibliographystyle{unsrt}
\bibliographystyle{ieeetr}
% \begin{figure*}
% \begin{center}
% \begin{tikzpicture}
% \node[anchor=south west,inner sep=0] (fig_All_M_Profiles) at (0,0){\includegraphics[width=\textwidth]{Data/Images/All_M_Profiles.png}};
% \begin{scope}[x={(fig_All_M_Profiles.south east)},y={(fig_All_M_Profiles.north west)}]
% \end{scope}
% \end{tikzpicture}
% \caption{}
% \label{fig:All_M_Profiles}
% \end{center}
% \end{figure*}
% \begin{figure*}
% \begin{center}
% \begin{tikzpicture}
% \node[anchor=south west,inner sep=0] (fig_All_RPF_Profiles) at (0,0){\includegraphics[width=\textwidth]{Data/Images/All_RPF_Profiles.png}};
% \begin{scope}[x={(fig_All_RPF_Profiles.south east)},y={(fig_All_RPF_Profiles.north west)}]
% \end{scope}
% \end{tikzpicture}
% \caption{}
% \label{fig:All_RPF_Profiles}
% \end{center}
% \end{figure*}
% \begin{figure*}
% \begin{center}
% \begin{tikzpicture}
% \node[anchor=south west,inner sep=0] (fig_All_T_Profiles) at (0,0){\includegraphics[width=\textwidth]{Data/Images/All_T_Profiles.png}};
% \begin{scope}[x={(fig_All_T_Profiles.south east)},y={(fig_All_T_Profiles.north west)}]
% \end{scope}
% \end{tikzpicture}
% \caption{}
% \label{fig:All_T_Profiles}
% \end{center}
% \end{figure*}
% \begin{figure*}
% \begin{center}
% \begin{tikzpicture}
% \node[anchor=south west,inner sep=0] (fig_All_V_Profiles) at (0,0){\includegraphics[width=\textwidth]{Data/Images/All_V_Profiles.png}};
% \begin{scope}[x={(fig_All_V_Profiles.south east)},y={(fig_All_V_Profiles.north west)}]
% \end{scope}
% \end{tikzpicture}
% \caption{}
% \label{fig:All_V_Profiles}
% \end{center}
% \end{figure*}
\appendix
\section{Definitions}
\subsection{Segmental Orientation}
In MD simulations using atomically-detailed force-fields, the segmental orientation of $n$-alkane chains can be monitored through the orientation of the chemical bonds. The segmental orientation \cite{Erman1985} of the fluid in terms of the second Legendre polynomial $\left<P_2\right>$ is defined as:
\begin{equation}\label{eq:P_2}
\left\langle P_2 \right\rangle =\frac{3\left\langle \cos ^2 \alpha \right\rangle - 1}{2},
\end{equation}
\noindent where $\alpha$ is the angle between a chain segment and a reference axis. In practice, $\left\langle \cos ^2 \alpha \right\rangle$ is calculated in the following way:
\begin{equation}
\left\langle \cos ^2 \alpha \right\rangle = \frac{1}{N_\text{bonds}} \sum_{b=1}^{N_\text{bonds}} \left(\frac{\textbf{r}^b_{ij}}{r^b_{ij}} \cdot \mathbf{\hat \imath} \right)^2
\end{equation}
in which $N_\text{bonds}$ is the total number of bonds, $\textbf{r}^b_{ij}$ is the difference between the positions of the two beads that form the bond $b$, and $\mathbf{\hat \imath}$ is the direction of the main axis.
In this study, the changes of the orientation in the shearing direction are monitored, through the segmental orientation with respect to $x$ projected in the $xz$ plane: $\langle P_{2}^{xy}\rangle$. Note that, by definition, if all the chains are parallel to the shearing direction $\langle P_{2}^{xz}\rangle=1 $; if all the chains are perpendicular to the surfaces $\langle P_{2}^{xz} \rangle=-0.5$, and, interestingly, if all the chains are oriented at \SI{45}{\degree} or if they are randomly oriented $\langle P_{2}^{xz}\rangle=0.25 $
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\subsection{End-to-end distance}
The mean square end-to-end distance of a single chain is defined as the square of the magnitude of the vector that points from one end of a chain to the other end. For a system containing several chains, the mean square end-to-end distance, $\left< R^2 \right>$, is given by the following equation~\cite{Brown1994}:
\begin{equation}
\left< R^2 \right> = \left<\left( \mathbf{r}_1 - \mathbf{r}_N \right)^2\right>
\label{eq:e2e2}
\end{equation}
\noindent where, for each chain, $N$ is the number of atoms and $ \mathbf{r}_i$ is the position vector of atom $i$.
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\subsection{Radius of gyration}
The mean square radius of gyration $\left(\left< S^2 \right>\right)$ is the average squared distance of each atom in the $n$-alkane chain from its centre of mass, as defined in the following equation\cite{Brown1994}:
\begin{equation}
\left< S^2 \right> = \frac{\left< \sum_{i=1}^{N} \left ( \lVert \mathbf{r}_i - \mathbf{r}_{\text{com}} \rVert \right)^2 \right>}{N}
\end{equation}
\noindent where, for each chain, $N$ is the number of atoms, $ \mathbf{r}_i$ is the position vector of atom $i$ and $\mathbf{r}_{\text{com}}$ is the centre of mass.
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\subsection{Mean squared displacement}
The mean squared displacement (MSD) of an atom is a measure of the deviation between its current and its initial position. The MSD is usually averaged over the number of atoms $\left(N\right)$ and its slope when plotted over time can be used to quantify the diffusion coefficient~\cite{Auhl2003}. The MSD is defined as:
\begin{equation}
\text{MSD} = \frac{1}{N}\sum_{i=1}^{N} \left( \mathbf{r}_i\left(t\right)-\mathbf{r}_i\left(0\right)\right)^2
\end{equation}
\noindent where $\mathbf{r}_i\left(t\right)$ is a vector containing the coordinates of atom $i$ at time $t$.
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\subsection{Average film temperature}
As common in NEMD simulations, the temperature of the lubricant is calculated via the kinetic energy of the atoms but excluding the velocity component in the direction of sliding (\emph{x}) to avoid contributions that are not related to the thermal vibration of the atoms~\cite{Gattinoni2013}. The measurements are block-averaged every \SI{100}{\femto\second}, the $\Delta$T values shown in Figure \ref{fig:T_v} are the time-averaged values from the final \SI{1}{\nano\second} of the sliding phase.
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\section{Equilibration}
Figure \ref{fig:Steps} shows the stages involved in the simulation procedure. Starting from an artificially ordered chains, the a heat-quench cycle is used to equilibrate the system, followed by compression and shear (see \ref{method}). Further details regarding the equilibration are shown below.
The evolution of $\left< R^2 \right>$ and $\left< S^2 \right>$ for C$_{16}$, C$_{30}$ and C$_{60}$ at \SI{2000}{\kelvin} are presented on the main part of Figures \ref{fig:e2e2} and \ref{fig:rg2}. For the case of C$_{60}$, which has lower diffusivity, these two quantities reach a steady state after approximately \SI{2e5}{} steps (\SI{0.2}{\nano\second}). The insets of the figures show the evolution of the same quantities after reducing the temperature to \SI{353}{\kelvin}; after \SI{1e7}{} steps (\SI{10}{\nano\second}), both $\left< R^2 \right>$ and $\left< S^2 \right>$ have reached their equilibrium values for the three chain lengths considered.
\begin{figure*}[htp]
\begin{center}