Nanotribology: Simulation Results


Soft and hard lubricants goto top

Constant-force algorithm: hysteresis

const-f.gif

Simulation results for Vss=3, Vsl=1/3, lattice=12×11, Nl=1 to 5:

 

"soft" lubricant: Vll <<Vsl   (Vll = 1/9)

  • two lubricant layers are “glued” to the substrates 

  • sliding occurs at a middle of the lubricant

  • lubricant melts at the onset of sliding

  • stick-slip motion is due to melting/freezing mechanism

"hard" lubricant: Vll >> Vsl   (Vll = 1)

  • lubricant film remains solid at sliding

  • sliding at the lubricant/substrate interface(s)

  • stick-slip motion is due to inertia mechanism

  • crystalline structure → perfect sliding (superlubricity)

hyst-soft.gif

hyst-hard.gif

       Nl=5 (enlarged view)

                           Nl=1 & 5  (enlarged view)

Spring algorithm: stick-slip goto top

draw01.gif

Simulation results for kspring= 3·10-4

 

Soft lubricant: Vll = 1/9
melting/freezing mechanism (vspring= 0.03)
Hard lubricant: Vll = 1 (ideal structure)
inertia mechanism (vspring= 0.01)
spring-soft.gif spring-hard.gif
Nl=3 movie (avi 2 Mb) Nl=3 movie (avi 1.6 Mb)

 

Melting of the confined film goto top

The melting temperature Tm of a thin lubricant film confined between two solid surfaces is higher than the corresponding bulk value, because the confinement decreases the entropy of the film and shifts the bulk melting transition to higher temperatures.

The melting temperature  Tm  as a function of the number of lubricant layers  Nl  for  fload= 0.1.
The filled markers correspond to MD simulation results, while the open markers are the theoretical values.
 

The solid curves describe the fits
Tm= 0.405 + 0.165/Nl  for the hard lubricant and
Tm= 0.045 + 0.350/Nl  for the soft lubricant.

The dependence Tm(Nl)  can be calculated analytically using the empirical Lindemann criterion: the melting starts when the amplitude of mutual displacement of nearest neighboring atoms reaches some threshold value of order 10% of the lattice constant. In the case of a thin film, mutual displacements decrease for the confined film, where the oscillations of the boundary atoms are suppressed. The mutual displacements are maximal at the lubricant/substrate interface for the hard-lubricant case, and at the middle of the soft-lubricant film where the internal interactions are weaker than the interactions with the substrates. Therefore, in the hard-lubricant system the melting starts at the interface, while in the soft-lubricant case the melting starts in the middle of the lubricant.
Thus, a very thin confined film is typically in a solid state. This results in a nonzero static frictional force  fs.

The dynamics of the film is significantly affected by the substrate, both in the solid and in the molten phases. The solid phase, able to sustain shear stress, shows large diffusional motions of the atoms; the molten phase shows a layered structure.
For more details, see O.M. Braun and M. Peyrard, Phys. Rev. E 68 (2003) 011506 "Dynamics and melting of a thin confined film"  (pdf files may be found here).

 


Soft lubricant: melting/freezing mechanism goto top

First, the soft lubricant freezes in a metastable configuration (which is different from stick to stick), where the two utmost lubricant layers are glued to the bottom and top substrates correspondingly and are characterized by a larger atomic concentration (e.g., for <10%) than the middle layers. The structure of the lubricant film has many defects (e.g., see two configurations below). The static threshold  fs  grows with the time of stationary contact (aging of the contact). Because the substrate-lubricant interaction (Vsl=1/3) is much larger than the interaction inside the lubricant (Vll=1/9), the sliding should begin somewhere within the lubricant, by depinning of the defects that stick the substrates.

Configurations of the  Nl=5  lubricant just prior the beginning of sliding

The annealed configuration at  f = 0.087

The stick configuration obtained during stick-slip motion with  vspring= 0.01  at  f = 0.017

Second, there are two different steady-state sliding regimes, the "layer-over-layer sliding" (LoLS) regime, when the lubricant remains in the solid state, and the "liquid-sliding" (LS) regime, when the lubricant is melted due to sliding. These two regimes exist for different intervals of the driving force or velocity.

Third, the behavior of a very thin lubricant film with two or less atomic layers, differs qualitatively from the behavior of thicker films.

 

Two sliding regimes (LoLS and LS) for  Nl>2  goto top

Figure soft5:

(a) the kinetic frictional force,

(b) the lubricant effective temperature, and

(c) the thickness of the lubricant film

as functions of the driving velocity for the  Nl=5 system.

 

Stars show the data obtained with the spring algorithm, while circles and dash curves show the results obtained with the constant-force algorithm (open circles are for the layer-over-layer sliding regime, and solid circles are for the liquid-sliding regime).

Layer-over-layer sliding. In the LoLS state the lubricant takes the well-defined layered structure; the utmost lubricant layers are glued to the corresponding substrates and move together with them, while the other (middle) layers slide (creep) one over another.

Figure soft7: Distribution of the x components of atomic velocities across the film in the LoLS regime at  f = 0.008;  see also the corresponding  movie  (animated gif, 3 Mb).

Bottom panel shows the distribution of atomic concentration across the lubricant.

 

The lubricant effective temperature grows with the velocity but remains lower than the melting one (Tmelt 0.11  for the five-layer soft flat system). Therefore, the lubricant is in a "solid" state; however, there is exchange of atoms between different lubricant layers in this state. Due to atomic exchange, the 2D atomic concentrations in the layers may vary during sliding and, when two adjacent layers occur to have close concentrations, they become commensurate, lock together and move with the same velocity, while the main sliding occurs at the most "incommensurate" interface (e.g., see Fig.soft7).

The layer-over-layer sliding is stable for the velocities  0.05 < vtop< 0.3  and forces  0.003 < f < 0.011  (for the five-layer film at  fload= 0.1; these intervals of velocities/forces shift to higher values as the load increases or Nl decreases).  The system locks at lower forces/velocities, while at larger ones the lubricant melts and the system goes to the liquid-sliding regime.

 

Liquid sliding. The LS regime appears with the increase of the driving velocity  (vspring> 0.3)  or force  (f  > 0.01  for  Nl=5).  At lower values of the forces/velocities, e.g., for  f = 0.011  and  vtop< 1  or  ~1,  the system exhibits smooth sliding with two utmost layers approximately ordered and glued to the corresponding substrates (the "glued" lubricant layer may exhibit a slow creep motion relative the substrate), while the middle layers are melted. When the force/velocity increases to  f = 0.012  or  vtop>> 0.3,  the glued layers melt as well. In the LS regime the x-component of the atomic velocities exhibits a linear gradient across the melted film. The sliding heats the lubricant to a temperature larger than the melting one, so that the driving itself maintains the lubricant in the molten state.

 

The transition from stick-slip to smooth sliding with the increase of driving velocity:

Figure soft26: Dynamics of the  Nl=3  flat system with the attached spring (kspring= 3×104) for three driving velocities:  v=0.03 (left column),  v=0.1 (middle column), and v=0.3 (right column).

 

The top row shows the spring force,

the second row shows the velocity of the top substrate,

the third row shows the lubricant thickness, and

the bottom row shows the effective lubricant temperature (note that  Tmelt 0.15  for the three-layer soft flat system).

 

The corresponding configurations are shown below

(see also the dependences for the five-layer film)

stick configuration for stick-slip with  vspring= 0.03

LoLS, smooth sliding at  vspring= 0.3

LS, smooth sliding at  vspring= 1

 

Onset of sliding goto top

The threshold force  fs  as well as the system dynamics at the onset of sliding first of all depend on the starting configuration. Namely, if the starting configuration contains a small concentration of defects (dislocations) that pin the substrates so that  fs  only slightly exceeds the kinetic frictional force, then the interstitial defective atoms are pulled back into the lubricant layers, the lubricant film is ordered, and the sliding begins with the layer-over-layer mechanism. On the other hand, when the starting configuration is far from the ordered configuration (e.g., as the annealed configuration shown above), then at the onset of sliding the lubricant has to be reorganized by plastic deformation, its effective temperature grows above the melting one, and the lubricant film melts almost immediately with the beginning of sliding. See also an example (*.mov 12 Mb, needs QuickTime installed).

 

 

Very thin lubricant films (Nl=2 or 1) goto top

The system behavior changes qualitatively for very thin lubricant films,  Nl ≤ 2, when all  lubricant atoms interact directly with the substrates. The lubricant films with two or less atomic layers are characterized by much larger values of the static threshold  fs  (which now does not change with time, i.e., there are no aging of the contact) and the critical velocity vc, than those for more thick films described above.

Two-layer lubricant film. Here, the two lubricant layers are glued each to the corresponding substrate, and the lubricant takes the two-layer crystalline structure  at stick configurations (e.g., see the figure below). Because both layers are made of the same material, they are almost commensurate, thus the sliding should start from the breaking of bonds between these layers. For the two-layer soft flat system, the lubricant melts at  Tmelt 0.23 and freezes (when the temperature decreases) at  Tfreeze 0.19. Because the static threshold force is large for the two-layer system, at the onset of sliding the effective lubricant temperature increases well above Tmelt. Therefore, the sliding always corresponds to the LS regime, and the hysteresis, as well as stick-slip, occurs due to the melting-freezing mechanism. The film melts at slips, although it remains well layered in three layers (see typical configurations below).

stick configuration for stick-slip with  vspring= 0.1

LS, smooth sliding at  vspring= 1

 

In the LS state, the bottom lubricant layer is almost glued to the immobile bottom substrate,  moving with a very small average velocity corresponded to creep motion with respect the bottom substrate, the top lubricant layer in the same manner is almost glued to the top substrate, while the middle layer moves with the vtop/2 velocity (in average) as demonstrated in Fig.soft32 below.

Figure soft32: The x-component of atomic velocities  vx  versus  z  for all lubricant and substrate atoms in the smooth sliding regime for  Nl=2  and  vspring= 1.

One-layer lubricant film. On the other hand, the  Nl=1  system behaves essentially similar to the hard-lubricant system described below. Although the lubricant is heated due to sliding up to temperatures Tlubr~ 0.1 to 0.4, this temperature is still well below the melting temperature Tmelt 0.57 of the one-layer film. Therefore, the hysteresis, as well as the stick-slip, exist due to inertia mechanism. Analyzing MD trajectories, at the onset of sliding we first observe the motion of two domain walls, which soon transform into a channel of moving lubricant atoms, and then to the "running" state of all lubricant atoms.

Figure soft37:

(a) the velocity of the top substrate,

(b) the effective lubricant temperature, and

(c) its thickness

as functions of the applied force for the one-layer lubricant film.

 

Solid symbols are for flat surfaces, and

open symbols, for the curved geometry.

 

Indicate the huge hysteresis: when the driving grows, the sliding starts at  f > 0.1,  but the backward transition to the locked state, when the force decreases, occurs at  f ~ 0.02

The same remains true for the case of the curved top substrate. In this case we observed at the onset of sliding a solitonic mechanism, which operates in the narrow region due to a bump ahead of it, so that the concentration in the narrow region increases and becomes incommensurate with the substrates.

 

stick configuration of the Nl=1 lubricant with the curved geometry just prior the beginning of sliding at   f = 0.102 < fs 0.105

LoLS, smooth sliding at  f = 0.06 when vtop= 1.99

 


Hard lubricant: Perfect sliding goto top

Flat surfaces and solid lubricant with an ideal crystalline structure: the universal dependence.

In the solid-sliding regime, when the top rigid substrate with one attached s-layer moves as a whole with the average velocity vtop, the bottom rigid substrate with one attached s-layer does not move at all, and the lubricant film moves as a whole with the velocity vlub= (1/2)vtop,  the washboard frequency is equal to  ωwash= 2πvlub/a = πvtop/a,  where a=as=3. The balance of forces for the top substrate + one attached s-layer is  FNS f = Nal ml η*(vtop) vlub,  where we have introduced the total viscous damping coefficient  η*(vtop)  for an atom in the utmost lubricant layer.  In the "perfect-sliding" approximation the atoms in the utmost lubricant layers feel only the external damping   ηext(vtop) ≈ η1(zlubr) [ηph(ωwash) + ηeh]  due to energy exchange with the rigid substrates. Assuming that all the damping within the lubricant is due to the external one,  η*(vtop) = ηext(vtop), we obtain the "universal" ("perfect-sliding") dependence
vtop(uni)(f)  =  2NS

Nal
  f

ml ηext(vtop)
 
which depends neither on the number of lubricant layers nor on the substrate masses ms and mS , because it corresponds to the steady state (while a delay of response of  vtop  when  f  varies non-adiabatically, has to depend on the masses).
  • üat small forces the effective friction is very  small, ηeff ~ v4
  • üplateau at vtop~ 1, when ωwash is within the phonon zone of the substrates
  • üfmax 7  and  vmax 10.8 (these values depend on ωm) further increase of  f  leads to unstable motion, because the energy injected into the system, cannot be dissipated
mp2fig09.gif

The lubricant is heated due to driving.

The distributions of velocities in all cases can be well approximated by Gaussian curves if we use different "temperatures" for the lubricant and the s-atomic substrate layers as well as for different degrees of freedom.
Definition: Ta = m( vαvα)2, where  ...  is for the averaging over time and over all atoms in a given layer, etc.

Simulation results (see figure to the right):

mp2fig06.gif

Hard lubricant: Energy losses  goto top

When one part of a system moves with respect to another part with a relative velocity v, then the rate at which work is done is equal to ε = vf, where f  is the total force acting on the former. This may be used to define the energy losses of a given atom as

εi = – 1

2
 
i 
(vi vi) fii
 
where the sum is over all atoms of the system including the s-atoms of the rigid substrates, and  fii  is the force acting on the i-th atoms from the  i′-th one.
Illustration: let us consider an ideal case, when the bottom substrate is immobile, the top substrate moves with a velocity  v  in the  x  direction, the lubricant film moves as a whole with the velocity  v/2, and the force  F  is applied to the top substrate. In this case we have
Flub/top= –Ftop/lub= Fbot/lub= –Flub/bot= F, and we obtain for the losses
εtop = –(1/2)(–F)·( v (1/2)v ) = (1/4) Fv,
εlub = –(1/2)[ F·( (1/2)v v ) F·(1/2)v ] = (1/2) Fv,
εbot = –(1/2)F·( 0 – (1/2)v ) = (1/4) Fv,
thus the total losses ε = Fv = the work done by the external force as expected.
Simulation results (see figure to the right): the anharmonicity is large, and main energy losses are at the interfaces.

 

Hard lubricant: melting of the confined film goto top

If  T >Tm, then the lubricant is liquid, the static friction is zero,  fs= 0,  there are no hysteresis, and the kinetic friction is much higher than for the perfect-sliding case.
a61.gif
 

Figure: distribution of  x-velocity across the lubricant for Nl= 5  (f = 0.01)  in

  • the solid state (T = 0.4, perfect sliding), see movie (avi 0.46 Mb)
  • the molten state (T = 0.5, liquid sliding), see movie (avi 0.43 Mb)

Hard lubricant: Role of temperature goto top

a46.gif

a47.gif

      enlarged view

       enlarged view

 

The solid lubricant with the ideal structure provides the smallest frictional force (superlubricity). If the lubricant film is in a liquid state, the frictional force is much larger. When the lubricant is frozen again, it takes an amorphous structure because, due to the contact with the substrates, the energy is removed from the lubricant very fast, and the lubricant film has no time to order. The frictional force in the case of an amorphous lubricant is typically larger than for the molten one.

Ideal crystalline configuration
(provides the perfect sliding)
Configuration after melting/freezing
(“amorphous” lubricant)
tc-5i01.gif as5_ini.gif

 

Hard lubricant: Transition from stick-slip to smooth sliding goto top

 

Perfect sliding: System dynamics for the  Nl=3 hard system (kspring= 3·104)  for three values of the driving velocity:

v=0.01 (stick-slip)  → 

v=0.03 (irregular/chaotic)  → 

v=0.1 (smooth perfect sliding)

 

"Amorphous" lubricant: the frictional force vs time for three values of the spring velocity for  Nl=2  and  Nl=5  with flat geometry

 

Hard lubricant: Phenomenological theory of kinetic friction goto top

The results obtained with MD simulations, however, still remain on an empirical level. Being quite time consuming, MD provides the value of friction only for a given set of model parameters and hardly may explain general trends and laws, e.g., the dependence of friction on the temperature, sliding velocity, shape of lubricant molecules, parameters of the interaction between the lubricant and substrates, etc. The kinetic friction appears due to energy flow from the sliding interface into the bulk of substrates, which finally is to be converted to heat. The energy losses emerge mainly because of creation of phonons; the rate of this process depends on many factors, such as the density of phonon states, positions and velocities of the lubricant atoms relative the surfaces, which in turn depend on the sliding velocity, the heating of the interface due to sliding, etc. Below we outline a phenomenological approach which allows us to explain the simulation results presented above, as well as to predict the friction properties of similar systems without (or before) MD simulations.

 

The energy balance approach. The most natural way of  fk  calculation is through the energy balance arguments: the energy  dEin/dt = Fvtop = Nsubfvtop pumped into the system per time unit due to external driving, must be equal to the energy  dEdiss/dt  dissipated in the substrates. The only way of energy dissipation in the model is through the viscous damping term  mlη(z,v)v  in Langevin equations of motion. When the lubricant has an effective temperature Tlub due to heating by the substrate as well as by the external driving, then its atoms move with a thermal velocity  vth=(kBTlub/ml)1/2 ~ 0.3−- 0.7  at temperatures  Tlub~ 0.1 − 0.5. When a lubricant atom is near a substrate at a distance  zl  from the nearest surface and moves with an average velocity  vl  with respect to it, then it losses per time unit the energy
ε(vl ; Tlub)  =  mlη1(zl)
+∞

-∞ 
dv η2(v) v2 P(vvl; Tlub) ,
(ph3)
where  P(v;T) = (ml /2πkBT)1/2exp(−mlv2/2kBT)  is the Maxwell distribution. In the thermodynamic equilibrium state, vl= 0  and  Tlub= Tsub,  this loss must be compensated by "energy gain" due to the stochastic force in Langevin equation, which acts on the mobile atoms from the substrate thermostat. Let  N'at  be the number of atoms in the lubricant layer just adjoined to the substrate surface, and vlx the average x-velocity of atoms in this layer relative the substrate, while vly= vlz= 0  for the motion along y and z. The total energy loss due to friction can be estimated as
  dEdiss

dt
 ≈  sN'atE1(vlx) ,   where
(ph4)
E1(vlx)  =  ε(vlx; Tlub) + 2ε(0; Tlub) 3ε(0; Tsub) .
(ph5)

The first contribution in E1 comes from the x-motion, and the second one, from the motion along y and z. The last term in Eq.(ph5) describes the energy gain coming from the thermostat to mobile atoms due to action of the stochastic force; this contribution has to be subtracted from the frictional losses. In the thermodynamic equilibrium state the energy gain must be equal to the energy loss due to the thermostat, ε(0;Tsub). We assume that this contribution (from the stochastic force emerged due to the thermostat) remains the same in the nonequilibrium driven state. The factor s=2  in front of the r.h.s. of Eq.(ph4) appears in the case of symmetric sliding, where there are two sliding lubricant/substrate interfaces. In the asymmetric case, when there is only one sliding interface, one has to put  s=1  and

  dEdiss

dt
   N'at [ E1(vlx) + E1(0) ] ,
(ph6)
where the first contribution comes from the sliding interface and the second one, from the stick lubricant/substrate interface.
From the equality  dEin/dt = dEdiss/dt  we finally obtain for the kinetic friction force
fkml G η1(zlF(vtop) ,
(ph7)
where  G = sN'at/Nsub  and  η1(zl)  are "geometrical" factors which only indirectly depend on the velocity and temperature through a change of the lubricant structure due to sliding (typically  G < 1.2  and  η1~ 0.1  in our model). The last factor in Eq.(ph7) is the main one that determines the dependence of the kinetic friction on the driving velocity and temperature, and is defined (for the symmetric sliding,  s=2) as
F(vtop)  =  vtop-1
+∞

- 
dv η2(vv2 [ P(vvlx; Tlub) + 2P(v; Tlub) − 3P(v; Tsub) ].
(ph8)
If we take into account only the minimal contribution  ηmin  in Eq.(ph2), then the factor  F  becomes equal to
Fmin(vtop)  =  ηmin é
ë
vlx2 + 3kB

ml
(TlubTsub) ù
û
  / vtop,
(ph9)
which grows linearly with the velocity (at  Tlub= Tsub) as well as with the lubricant temperature (at fixed  vtop  and  Tsub).
In a general case, the factor  F  grows with the driving velocity and temperature too. Its value is determined by the velocity  vlx  and the lubricant temperature Tlub. It is easy to find vlx in the case of solid lubricant:  vlx= vtop/2  for symmetric sliding and  vlx= vtop  for asymmetric sliding. In the case of a liquid lubricant (the LS regime), one has  0 < vlx< vtop/2. As for the lubricant temperature, its value depends on the sliding velocity due to sliding-induced heating.

Thus, we have to take into account two issues: first, the lubricant is always heated due to driving, and second, the lubricant structure depends on the sliding steady state, so that the kinetic friction force is determined by a rather delicate balance of these factors.

 

The geometrical factor. While the factor  F  in Eq.(ph7) increases with the driving velocity and temperature, the geometrical factors  G  and  η1  decrease thus compensating the growth. First, the lubricant thickness d  increases with the lubricant temperature due to thermal expansion. As a result, the distance of lubricant atoms from the nearest surface will increase with temperature too, e.g., as  zl zl0+ βzlTlub. The increase of  zl  will result in the decrease of the coefficient  η1. Second, the geometrical factor  G  is directly proportional to  N'at,  i.e., to the number of atoms in the lubricant layer just adjoined to the substrate surface. If  N'l  is the effective number of layers in a given configuration, then  N'at can be found from the conservation of the total number of atoms, N'atN'l= NatNl. For the perfect crystalline structure of the lubricant we have N'at= Nat. In the case of hard lubricant with "amorphous" structure Nl < N'l Nl+1 so that N'at< Nat,  or  N'at= αNNat,  where αN< 1. The case of liquid lubricant, when the lubricant structure is changed with the driving velocity and temperature, is more involved: the number of atoms that interact with the substrate, N'at, should decrease when the film thickness grows. For  Tlub> Tmelt  we found the dependence

N'at = αN Nat / [1+ βN (TlubTmelt)]2/3.
(ph11)
The geometrical factor G, which is directly proportional to  N'at,  may strongly decrease with temperature and driving velocity, especially in the LS regime.

 

The lubricant temperature. The very important issue is that the lubricant temperature increases with the driving velocity. It can be found analytically with the help of energy balance arguments similarly to the approach used above: an energy R+ pumped into the lubricant (per atom and per time unit), must be equal to the dissipated energy R. The energy flux from the lubricant to the substrates, which emerges due to the difference of their temperatures, may be defined as

R  =  6ml


+∞

- 

dv η(z,v) v2 [ P(v,Tlub) − P(v,Tsub) ].

(ph15)

The pumped energy R+ emerges due to the "shaking" of the lubricant during sliding. This rate can be estimated with the help of linear response theory: Let the lubricant be perturbed by an oscillating force of an amplitude  f0  and a frequency  ω0. When one lubricant layer slides over another one with a relative velocity  v0,  it is disturbed with a "lubricant washboard frequency"  ω0=2πv0/r,  where  r  is an average interatomic distance in the lubricant. Then, the rate R+  can be found as

R+  =  πs f02

4ml
 
+∞

- 
dv ρlub(2πv/r) [ P(vv0,Tlub) − P(v,Tlub) ],
(ph16)
where  ρlub(ω)  is the (local) density of phonon states in the lubricant. The value  v0  in Eq.(ph16) is directly proportional to the driving velocity  vtop,  although it has to be defined in different ways for different types of sliding. In the limit  vtop 0,  Eq.(ph16) gives  R+ v02,  which leads to the correct low-velocity relation  ΔT vtop2.  At higher velocities, the value of R+ depends on the lubricant phonon spectrum. Extracting all necessary parameters from the simulation data, we obtained the following dependences:

 

The perfect sliding.

Figure ph12: Phenomenological dependences for the perfect sliding of the Nl=5 solid lubricant for f0= 0.2. The symbols correspond to simulation data.

(a) Dependence of the kinetic friction force on the sliding velocity at Tsub= 0.1 (open diamonds) and Tsub= 0.4 (red solid circles). 

(b) Dependence of the kinetic friction on the substrate temperature for three values of the sliding velocity vtop= 0.1 (black), 0.3 (red), and 1 (blue).

Here  vlx= vtop/2,  s=2,  G=1.212,  and from the simulation data we extracted other parameters:  βz 3.6,  zl0 5.21,  βzl 0.3,  and  zs 2.12.

Note that if we ignore all temperature dependences, we obtain in the  Tsub= 0  case the "universal" dependence described above.

 

The lubricant with the "amorphous" structure.

Figure ph13: Phenomenological dependences of the kinetic friction force on the sliding velocity at  Tsub= 0  (blue curve and open diamonds, the symbols correspond to simulation data) and  Tsub= 0.3  (red curve and solid circles) for the "amorphous" lubricant.

Here  αN= Nl /(Nl+1) 0.833 (simulation suggests  αN 0.81), and from the simulation data we extracted other parameters:  βz 2.33,  zl0 5.22,  βzl 0.333,  and  zs 2.112. Note, however, that we had to take different values for the fitting parameter  f0,  namely,  f0 = 1.4 for  Tsub= 0  but  f0= 1.8  for  Tsub= 0.3.

The sliding becomes asymmetric at low driving velocities  vtop< 0.3;  in this case we put  s=1  and  vlx= vtop. Unfortunately, we were unable to obtain a good quantitative agreement with simulation using the procedure described above. Instead, a good agreement was achieved with an (artificial?) dependence  Tlub 2vtop3.

The T=0 movies can be found here (animated gif, 8.6 Mb).

 

The LS (liquid sliding) regime. When the confined film is molten, its thickness changes with the temperature and driving velocity, and the  factor strongly decreases with temperature. To find  N'at, we calculated the distribution of atomic concentration around the first lubricant layer  P1(z)  (see Fig.ph10) and then integrated it,  N'at=C dz P1(z);  the dependence  N'at(Tlub)  can be fitted with a good accuracy by Eq.(ph11). Then, we have to know the average velocity  vlx  of the first lubricant layer with respect to the substrate. Using the distribution of x-velocities across the lubricant obtained in simulation, we schematized it as shown in Fig.ph16. The dependence  vx(z)  is linear within the lubricant, but undergoes jumps at the substrate/lubricant interfaces. Using this picture, we assumed that  vlx  can be described by the expression

vlx = αvtop (zlzs)/
~
d
,
(ph20)
which includes a phenomenological parameter  α.  The values of  α  extracted from the simulation data, may be fitted by the dependence  α = 2.53 − 0.71Tlub.

Figure ph10: Distribution of atomic concentration across the lubricant in the liquid-sliding regime (schematically).

Figure ph16: Schematic presentation of distribution of the x-velocity of lubricant atoms across the lubricant in the LS regime.

 

In the result, we obtained the following dependences for the LS regime:

Figure ph18: Dependence of the kinetic friction force

(a) on the driving velocity at a fixed substrate temperature,  T = 0.3 (black), 0.4 (blue), 0.5 (red), and 0.6 (magenta), and

(b) on the substrate temperature at a fixed driving velocity  vdrive= 0.75 (red) and 0.25 (black) for the LS regime.

Symbols correspond to simulation data, while solid curves describe the phenomenological dependences.

Here  βz 36.4,  zs 2.09,  βN 2.84,  Tmelt= 0.44,  αN = 0.65,  zl0= 5.41,  βzl 0.313,  and  f0 4.3.

 

Discussion and conclusion. Thus, the simulation results can be satisfactorily explained with the help of the phenomenological approach. Almost all parameters that we extracted from simulation data, may in principle be calculated from first principles or estimated at least, except the two fitting parameters  f0  and α.  The first parameter, the coefficient  f0  in Eq.(ph16), describes the amplitude of the oscillating ("shaking") force responsible for the sliding-induced heating of the lubricant. Note that the value of  f0  correlates with the structure of the lubricant: the smallest value (f0= 0.2) was found for the perfect sliding regime, while much larger values  (f0= 1.4 to 1.8), for the "amorphous" lubricant structure. The largest value (f0= 4.3) was obtained for the liquid lubricant. We could suggest that the value of  f0  is proportional to "defectivity" of the lubricant film structure, although we have no clear understanding of this problem, and it requires further investigation. Second, the friction force in the LS regime strongly depends on the velocity of the first lubricant layer relatively to the substrate. According to Eq.(ph20),  vlx  is determined by the parameter α,  i.e., by the gradient of x-velocity at the substrate/lubricant interface. This question is closely related to the "slip" or "no-slip" behavior of a liquid flow near a solid surface. The latter is characterized by the so-called slip length defined as

Ls = vlx

/

 

dvx


dz

 

|
|


z=zl

.

In turn, the slip length is determined by the liquid-surface interaction, i.e., either the surfaces are wetting or non-wetting by the lubricant (the latter situation corresponds to the hard-lubricant system described here).

 

Although the phenomenological theory presented above still uses some parameters, it allows us to explain the simulation results. In particular, it naturally explains the increase of kinetic friction with the driving velocity. Then, it explains the temperature dependence of the friction, which emerges due to interplay of two factors  F  and  1:  the first factor increases with temperature, while the geometrical one,  1,  decreases with  T. Thus, the phenomenological theory allows us to predict, at least qualitatively, the behavior of other tribological systems in a general framework.

 

For more details of the phenomenological approach, see O.M. Braun, Phys. Scr. 78 (2008) 015802 "Phenomenological theory of kinetic friction for the solid lubricant film(pdf files may be found here).

 


Role of substrate geometry goto top

In the results presented above, the top and bottom substrates were identical. Now let us consider the tribosystem, where the substrates are not identical. In this case, the sliding becomes asymmetric, and essential changes of sliding mechanisms occur in the case of soft lubricants.

 

Curved surface goto top

Note: magnitude of frictional forces, as well as the critical velocity  vc,  for the hard lubricant are in about ten times smaller, than for the soft lubricant.

                 low driving velocity:

                 stick-slip motion

high driving velocity:

smooth sliding

Nl=3    Vsl=1/3
fig02.gif

 

soft lubricant:

Vll << Vsl

 

 

hard lubricant:

Vll >> Vsl

 

                           configurations / movies (mpeg)        

soft lubricant (Vll=0.2): melting/freezing

stick-slip motion (v=0.3)  movie (6.6 Mb)

soft lubricant (Vll=0.2)

smooth sliding (v=1)  movie (3.4 Mb)

 

hard lubricant (Vll=1): inertia mechanism

stick-slip motion (v=0.03)  movie (3.4 Mb)

hard lubricant (Vll=1):

smooth sliding (v=0.1)  movie (4.5 Mb)

 

Stepped surface goto top

 

nsteps=3,  smooth LoLS,  vspring=0.1

nsteps=3,  smooth LS,  vspring=1

 

Rotated surface goto top

Figure rot14: Friction force as a function of the misfit angle φ  (i.e., when the substrates are rotated one with respect to another on the angle φ).

 

Diamonds and solid curve describe the static friction,

dotted curves show the kinetic friction force at different driving velocities (from v = 0.1 to 3) as indicated in legend.

 

Star symbols correspond to the regime of "solid sliding".

 

Panel (a) is for T = 0, and (b) is for the "room" temperature T = 0.025.

As expected, in the LS regime (e.g., for v = 1 and 3 in Fig.rot14) the friction is almost independent of φ. However, the static friction, as well as the kinetic friction in the LoLS regime, may change on more than one order in magnitude, when the misfit angle varies. The highest friction is for φ=0. For all other angles studied, the sliding always corresponds to the LoLS regime at low driving velocities. Contrary to the φ=0 case, now the sliding is typically asymmetric, the sliding takes place at a single interface only, between the middle and one of the attached lubricant layers, i.e., the middle layers move either with the top or bottom substrate. The middle layers may remain ordered during sliding or, for other values of φ, they may be 2D melted (in the former case, the friction is lower).

 

For few special misfit angles, namely, for  φ = 15.4, 27.2 and 42.8 indicated by stars in Fig.rot14, we observed the "superlubricity" characterized by a very low friction. In these cases, the lubricant film is solid and ordered during sliding, and moves together with one of the substrates. This is the "solid sliding" (SS) regime. However, the lubricant is not rigid during motion, which now corresponds to the "solitonic" mechanism.

 


Conclusion goto top

  • Soft lubricant film (3 ≤ Nl ≤ 5):

    • for the static frictional force  fs,  Amontons law is explained by the interactions within the lubricant, that leads to  μ ~ 0.5;  the value of  fs  is determined by a particular structure of the frozen metastable lubricant film (i.e., by defects that should be broken at the sliding onset);

    • fs  slowly grows with the time of stationary contact (up to in four times of magnitude) and reaches its maximum for the annealed configuration (aging of the contact);

    • the hysteresis of  vtop(f)  (when the driving force increases/decreases adiabatically) as well as the stick-slip motion at low driving velocities, is due to the melting-freezing mechanism; the transition from stick-slip to smooth sliding takes place at  vc~ 0.1;

    • there are two steady-state sliding regimes, the LoLS sliding at low driving forces/velocities and the LS regime at larger ones. In the LoLS regime the sliding stops when the force/velocity decrease and  v  becomes lower than  vtop~ 0.1; when the force/velocity increases, the lubricant melts and the LS regime begins to operate. The LoLS regime is more stable for narrower films, than for thicker ones;

    • at smooth sliding, the kinetic frictional force is approximately independent on the driving velocity, and is lower (or much lower) than the static frictional force. However, the lubricant effective temperature (the heating of the lubricant due to sliding) as well as the thickness of the film are proportional to the driving velocity;

  • Very thin soft lubricant films (Nl=2 and 1):

    • for two-layer  lubricant film, both its layers are glued to the corresponding substrates in the annealed and stick configurations, and the film structure is crystalline. As a result, the static frictional force is high, the lubricant always melts at the onset of sliding, so that only the LS regime operates. The hysteresis of  vtop(f),  as well as the stick-slip, always correspond to the melting-freezing mechanism, and the threshold velocity of the transition from stick-slip to the smooth sliding  (vc~ 0.6)  is also higher than for thicker films;

    • for one-layer  lubricant film, the static friction is the largest,  fs~ 0.1. The film typically is not melted during sliding, and the stick-slip motion is due to inertia mechanism similarly to the hard lubricant system (because of too high melting temperature of the one-layer film);

  • Perfect sliding (hard lubricant):

    • the solid lubricant film with ideal crystalline structure provides the lowest frictional forces,  μ ~ 10-3 to 10-2;

    • the kinetic frictional force depends on the sliding velocity  v  (Amonton’s laws does not operate for  fkinetic);

    • the transition from stick-slip motion to smooth sliding occurs at   vc~ 0.03 to 0.1;

  • Lubricant with amorphous structure (hard lubricant):

    • the solid lubricant film with a metastable (amorphous) structure leads to a rather high frictional forces,  μ ~ 0.1 to 1;

    • the kinetic frictional force depends on  v  as  fkineticBv  with  B ~ 0.015 to 0.035;

    • the transition from stick-slip motion to smooth sliding occurs at  vc~ 0.1 to 0.5;

  • Role of substrate temperature (hard lubricant):

    • if one starts from the perfect-sliding state, then the velocity decreases as  T  grows until the film melts. After that, the mobility grows with  T,  although it remains lower than the perfect-sliding value. For amorphous lubricant, the mobility increases with  T ;

    • in the molten state, the kinetic frictional force is proportional to the sliding velocity  v, e.g., as  fkinetic≈ 0.017vtop;

    • fkinetic  has peculiarities at  T = Tm  and  T = Tfreeze, where the mobility slows down.

Next: Self-ordering of the lubricant

goto main Back to main page or tribology page


Last updated on November 26, 2009 by O.Braun.    Copyright © by O.Braun.

Translated from LATEX by TTH,  configurations produced by RasTop,  movies by VMD