Sunteți pe pagina 1din 74

Thermodynamics of Nonisothermal Polymer Flows:

Experiment, Theory, and Simulation

Brian J. Edwards
Department of Chemical and Biomolecular Engineering
University of Tennessee-Knoxville

University of Kentucky
Lexington, Kentucky
February 18, 2009
Collaborators and Funding

 Tudor Ionescu: Graduate student, UTK


 Vlasis Mavrantzas, Professor, University of Patras

 Grant #41000-AC7, The Petroleum Research Fund,


American Chemical Society
Outline

 Part I: Introduction and Background


• Introduction to Viscoelastic Fluids
• Definition of the concept of Purely Entropic Elasticity
• Objective
 Part II: Experiment and Theory
• Experimental Approach
• Theoretical Approach
 Part III: Molecular Simulations
• Equilibrium Simulations
• Nonequilibrium Simulations
 Conclusions
Part I: Introduction and Background

 The phenomenon described in this presentation is one manifestation of


viscoelastic fluid mechanics

 Viscoelastic fluids display complex non-Newtonian flow properties under


the application of an external force:

» Pressure gradient
» Shear stress
» Extensional strain (stretching)
Examples of Viscoelastic Fluids

 Paint (&)  Toothpaste


 Crude oil  Grease
 Asphalt  Foodstuffs
 Cosmetics • Ketchup
 Biological fluids • Dough
• Blood • Salad dressing
• Protein solutions  Plastics
 Pulp and coal slurries • Polymer melts
• Rubbers
• Polymer solutions
Newtonian Fluid Dynamics

 The dynamics of an incompressible Newtonian fluid can be


described completely with three equations:

 The Cauchy momentum equation:


Dv
   p    
Dt
 The divergence-free condition:
 v  0
 The Newtonian constitutive equation:

     v    v 
Newtonian Flow Equations Are
Remarkably Robust:

 Simple, low-molecular-weight, structureless fluids are well described in


three dimensions:
 Laminar shear and extensional flows
 Turbulent pipe and channel flows
 Free-surface flows

• The simple, structureless fluid:


Viscoelastic Fluid Dynamics

 A viscoelastic fluid has a complex internal microstructure


 Today’s topic: Polymer melts
 A high-molecular-weight polymer is dissolved in a simple Newtonian
fluid
 At equilibrium, the polymer molecules assume their statistically most
probable conformations, random coils:

Polymer solution
Viscoelastic Flow Behavior

 These conformational rearrangements produce very bizarre


“non-Newtonian” flow phenomena!

 Viscoelastic fluids have very long relaxation times:

Viscoelastic fluid

 12 Newtonian fluid

t Flow off
Viscoelastic Flow Behavior

 Viscoelastic fluids typically display shear-rate dependent viscosities:

Shear-thinning fluid

 Newtonian fluid


Viscoelastic Flow Behavior

 Viscoelastic fluids develop very large normal stresses:

Viscoelastic fluid
 11   22
1 
 2

Newtonian fluid



 Example: Paint
Nonisothermal Flows of VEs

 Nonisothermal flow problems defined by a set of four PDE’s:

 p      g
Dv
• 1) Equation of motion: 
Dt


   v 
• 2) Equation of continuity: t

 Incompressible fluid:   v  0

DUˆ
    q   p  v    : v 
• 3) Internal energy equation: Dt

     G   v     v 
• 4) An appropriate constitutive equation:  1
 Upper-Convected Maxwell Model (UCMM)  
The concept of
Purely Entropic Elasticity

 For simplicity, the internal energy of a viscoelastic liquid is considered as


a unique function of temperature (i.e. not a function of deformation) [1,2]:

Uˆ  Uˆ T 
 This let us define the constant volume heat capacity as:

dUˆ (T )
cˆv 
dT

 For an incompressible fluid with PEE, the heat equation becomes:

   q    : v 
DT
cˆv
Dt
 PEE is always assumed in flow calculations!!!
1. Sarti, G.C. and N. Esposito, Journal of Non-Newtonian Fluid Mechanics, 1977. 3(1): p. 65-76.
2. Astarita, G. and G.C. Sarti, Journal of Non-Newtonian Fluid Mechanics, 1976. 1(1): p. 39-50.
Implications of PEE

 What happens to the energy equation if one does not assume PEE?
• First, the internal energy is taken as a function of temperature and an
appropriate internal structural variable (conformation tensor):
Uˆ  Uˆ T , c 
Uˆ
• Next, the heat capacity is defined as: cˆv 
T c

• Then, the substantial time derivative of the internal energy becomes:

DUˆ DT Uˆ Dc DT Uˆ  c 


 cˆv  :  cˆv  :   v  c 
Dt Dt c T ,V
Dt Dt c T ,V
 t 

• The complete form of the heat equation becomes:


Uˆ  c 
:   v  c     q    : v 
DT
cˆv 
Dt c T ,V  t 
Objective

 Test the validity of PEE under a wide range of processing


conditions using experimental measurements, theory and
molecular simulation
• Experimental approach
 Solve the temperature equation numerically using a finite element
modeling method (FEM)
 Measure the temperature increase due to viscous heating, and compare
the results to the FEM predictions
• Theoretical approach
 Identify all possible causes for the deviations from the FEM predictions
observed in the experimental measurements
 Use a theoretical model to propose a more accurate form of the
temperature equation and test it through the FEM analysis
• Molecular simulation approach
 Use a molecular simulation technique to evaluate the energy balances
under non-equilibrium conditions for compounds chemically similar to
the ones used in the experiments
Part II: Experiment and Theory

 Experimental Approach
• Identify a flow situation in which high degrees of orientation are
developed
 Uniaxial elongational flow generated using the semi-hyperbolically converging
dies (Hencky dies)
 The analysis is not possible in capillary shear flow

• Find numerical solutions to the temperature equation at steady state


using the PEE assumption for this particular flow situation

cˆv v  T   k 2T   : v 
 The solution to this equation will yield the spatial temperature distribution profiles
inside the die channel
 Compute the average temperature value for the exit axial cross-section of the die
• Under the same conditions used in the FEM calculations, measure the
temperature increase due to viscous heating
Experimental Approach

 The semi-hyperbolically converging die (Hencky die)


• Proven to generate a uniaxial elongational flow field under special
conditions 2
 A0   D0 
 
 H  ln    ln  
 Ae   De 

r z  
2 A
zB
R02 Re2 Re2
A L 2 BL 2
R0  Re2 R0  Re2

Hencky 6 Die:
D0  19.96 mm
De  0.9937 mm
H  6
Experimental Approach

 Materials used in this study

Material Grade MI Density Thermal MW PI


(g/10min) (g/cm3) conductivity
(Wm-1K-1)
LDPE Exact 3139 7.5 0.901 0.3 56,950 1.99

HDPE Paxxon 0.3 0.943 0.5 105,200 9.74


AB40003
Experimental Approach

 Calculation of the steady-state spatial temperature distribution profiles


• Used a FEM method to find numerical solutions to the temperature equation

cˆv v  T   k 2T   : v 
• First, elongational viscosity measurements are needed in order to evaluate
the viscous heating term:
3
 : v   zz  e 2
2
• The elongational viscosity is identifiable with the “effective elongational
viscosity” [1] which can be measured using the Hencky dies and the
Advanced Capillary Extrusion Rheometer (ACER)

P
 e   ef 
 H

1. Feigl, K., F. Tanner, B.J. Edwards, and J.R. Collier, Journal of Non-Newtonian Fluid Mechanics, 2003. 115(2-3): p. 191-215.
Experimental Approach

 Advanced Capillary Extrusion Rheometer (ACER 2000)

L
v ram 
exp  H   1
Experimental Approach

 Effective elongational viscosity results


• HDPE
1.00E+07
190ºC
210ºC
Effecive Elongational Viscosity (Pa•s)

230ºC

1.00E+06

1.00E+05
1 10 100
Strain Rate (1/s)
Experimental Approach

 Effective elongational viscosity results


• LDPE
1.00E+07
150ºC
170ºC
Effecive Elongational Viscosity (Pa•s)

190ºC

1.00E+06

1.00E+05
1 10 100
Strain Rate (1/s)
Experimental Approach

 FEM calculations
 cˆv v  T   k2T ef  2
• The heat capacity is considered a function of temperature
 the tabulated values for generic polyethylene are used from [1]
• The thermal conductivity is considered isotropic, and taken as a constant with respect
to temperature and position [1]
• The input velocity field corresponds to a uniaxial elongational flow field in cylindrical
coordinates [2]
1
vr    r vz   z
2
• The effective elongational viscosity is taken as a function of temperature [3],
according to our own experimental measurements

 A0 
 T    0 exp  
 k BT 

1. Polymer Handbook. 1999, New York: Wiley Interscience.


2. Feigl, K., F. Tanner, B.J. Edwards, and J.R. Collier, Journal of Non-Newtonian Fluid Mechanics, 2003. 115(2-3): p. 191-215.
3. Dressler, M., B.J. Edwards, and H.C. Ottinger, Rheologica Acta, 1999. 38(2): p. 117-136.
Experimental Approach

 Sample FEM calculation results


• HDPE, Tin = Twall = 190oC

  2 s 1
  10s
 1
  50 s
 1
Experimental Approach

 Sample FEM calculation results


• Axial temperature profiles
• HDPE, Tin = Twall = 190oC
16

14

12
ΔT = T(r=0,z)-Tin (K)

10

0
0 5 10 15 20 25 30
z (mm)
Experimental Approach

 Sample FEM calculation results


• Radial temperature profiles
• HDPE, Tin = Twall = 190oC
16

14
ΔT = T(r,z=25mm)-Tin (K)

12

10

0
-0.1 -0.05 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45
r(mm)
Experimental Approach

 Complete FEM calculation results


• Average exit cross-section temperature increases with respect to the inlet
• HDPE 10
Tin=190ºC
9
Tin=210ºC
T r , z  L rdr
R
 8
ΔT=<T(r,z=25mm)>-Tin (K)

Tin=230ºC
T  0
R
 rdr
exit
7
0

0
1 10 100
Strain Rate (1/s)
Experimental Approach

 Complete FEM calculation results


• Average exit cross-section temperature increases with respect to the inlet
• LDPE
14
Tin=150ºC
T r , z  L rdr
R
 12 Tin=170ºC
ΔT=<T(r,z=25mm)>-Tin (K)

T  0
R Tin=190ºC
 rdr
exit

0
10

0
1 10 100
Strain Rate (1/s)
Experimental Approach

 Experimental design for the temperature measurements


Experimental Approach

 Complete temperature measurement results


• HDPE
16
Tin=190ºC Measured
14 Tin=210ºC Measured
Tin=230ºC Measured
ΔT=<T(r,z=25mm)>-Tin (K)

12 Tin=190ºC FEMLAB
Tin=210ºC FEMLAB
10 Tin=230ºC FEMLAB

0
1 10 100
Strain Rate (1/s)
Theoretical Approach

 Identify all the factors that may be responsible for the deviations
observed at high strain rates
 Key assumptions made for the derivation of the temperature equation
used in the FEM analysis
DUˆ
• Started with the general heat equation     q   p  v    : v 
Dt

• Assumption 1: Incompressible fluid  v  0

Uˆ DUˆ
• Assumption 2: Flow is steady  0 and  v  Uˆ
t Dt
dUˆ T 
• Assumption 3: Fluid is Purely Entropic Uˆ  Uˆ T  and cˆv 
dT
• Obtained the temperature equation solved using FEM

cˆv v  T   k2T ef  2


Theoretical Approach

 Furthermore
• As a consequence of Assumption 3, the heat capacity is a
function of temperature only
• Assumption 4: the thermal conductivity is isotropic
• Assumption 5: the velocity flow field corresponds to uniaxial
elongational stretching (with full-slip boundary conditions)
1
vr   r v z  z
2

 Identified Assumptions 3, 4, and 5 as possible candidates


responsible for the deviations mentioned earlier
Theoretical Approach

 Elimination of Assumptions 4 and 5


• Considered anisotropy into the thermal conductivity
 Increased k|| by 20%
 Decreased k┴ by 10%
• Axial temperature profile calculated for HDPE at Tin = 190oC and a strain
rate of 34s-1 14
k_isotropic
12 k_anisotropic
ΔT = T(r=0,z)-Tin (K)

10

0
0 5 10 15 20 25
z (mm)
Theoretical Approach

 Clearly, the PEE assumption seems to be the only remaining factor that
is potentially responsible for the deviations observed at high strain rates

 How do we eliminate it?


• Start with the complete form of the temperature equation for an
incompressible fluid defined earlier

Uˆ  c 
:   v  c     q    : v  (*)
DT
cˆv 
Dt c T ,V
 t 
• First correction: introduce conformation information into the heat capacity
[1,2]  2 K T 
c  c0  T tr c  tr c 0 
1
2 T 2
• Second correction: introduce the second term on the left side of equation (*)

1. Dressler, M., B.J. Edwards, and H.C. Ottinger, Rheologica Acta, 1999. 38(2): p. 117-136.
2. Dressler, M., The Dynamical Theory of Non-Isothermal Polymeric Materials. 2000, ETH: Zurich.
Theoretical Approach

 Both corrections mentioned above require knowledge of the


conformation tensor
• We can use the UCMM to evaluate the conformation tensor
components inside the die channel
 1 k T
c   c  B  
 K T 
• In Cartesian coordinates, the diagonal components of the
normalized conformation tensor work out to be:
 
 1
1 
 
~ ~   z      t   1
c xx  c yy     exp   R   H  
  1  L      1  
 
  1 
 2  
~ 2  z       tR   1
c zz     exp    2 H  

2  1  L 
     2  1
 
Theoretical Approach

 Relaxation time measurements


• Complete results for HDPE and LDPE
1
LDPE
HDPE
Exponential Fit LDPE
Exponential Fit HDPE
0.1
λ (s)

0.01

0.001
0.0019 0.002 0.0021 0.0022 0.0023 0.0024 0.0025 0.0026 0.0027
1/T (K^-1)
Theoretical Approach

 Conformation tensor predictions using the UCMM


• HDPE, Tin = 190oC

1 1
  2 s
tr ~c  / tr ~c z 25mm

1
s
  3

1
s
4 .5
 
1
10s 1
  s
 50


0
0 5 10 15 20 25
z(mm)
Theoretical Approach

 Conformation tensor predictions using the UCMM


• HDPE, all temperatures
1.0E+05

T = 190°C
T = 210°C
T = 230°C
c z 25mm

5.0E+04
tr ~

0.0E+00
1 10 100
Strain Rate (1/s)
Theoretical Approach

 Correlation between the conformation at the exit cross-section and the


difference between the measured and calculated ΔT

1.0E+05
tr(c)(z=25mm) at Tin = 190°C
tr(c)(z=25mm) at Tin = 210°C 6

ΔT_measured-ΔT_calculated (K)
tr(c)(z=25mm) at Tin = 230°C
ΔT_measured-ΔT_calculated at Tin = 190°C
ΔT_measured-ΔT_calculated at Tin = 210°C 5
ΔT_measured-ΔTcalculated at Tin = 230°C
tr ~c z 25mm

4
5.0E+04
3

0.0E+00 0
1 10 100
Strain Rate (1/s)
Theoretical Approach

 First correction: the conformation dependent heat capacity


 2 K T 
c  c0  T tr c  tr c 0 
1
2 T 2

• For example, the total heat capacity evaluated at the die axis for HDPE
at Tin = 190oC
38 2 s-12 s 1
  10s 1
c  c0  cconf Jmol 1K 1

36   15s 1

  23s 1

34

  34 s 1
32
  50 s 1
50 s-1
30
0 5 10 15 20 25 30
z (mm)
Theoretical Approach

 Second correction
• Rearranging the complete form of the heat equation and making the
appropriate simplifications, we get:
  Uˆ czz 
cv v  T   k T   : v   vz 
ˆ 2


  czz z 

• The axial gradient of czz is already known from the UCMM


• The derivative of the internal energy with respect to czz can also be
evaluated using the UCMM [1]:
 K T  
  K T   T  tr c 
1
u
2  T 
Uˆ Uˆ
czz  tr c   
czz tr c 

Uˆ 1  K T  
   K T   T 
czz 2  T 

1. Dressler, M., The Dynamical Theory of Non-Isothermal Polymeric Materials. 2000, ETH: Zurich.
Theoretical Approach

 Examining the effect of introducing corrections 1 and 2 detailed above


• HDPE, Tin = 190oC

20

Calculated w/ no correction
Calculated w/ Correction 1 UCMM
Calculated w/ Correction 2 UCMM
Calculated w/ Correction 2 Giesekus (β = 0.0065)
15
ΔT=<T(r,z=25mm)>-Tin (K)

Measured

10

0
1 10 100
Strain Rate (1/s)
Part II: Summary

 Provided experimental evidence that PEE is not universally


valid

 Verified a new form for the temperature equation by


essentially eliminating the PEE assumption

 Using the UCMM, two corrections have been made to the


traditional temperature equation
• 1) The conformational dependent heat capacity
 Was found to have a significant decrease with increasing orientation
 Had a negligible effect on the calculated temperature profiles
• 2) The extra heat generation term
 Quantified the temperature profiles in agreement with the experimental
values
Part III: Molecular Simulations

 Simulation Details
• NEMC scheme developed by Mavrantzas and coworkers was used
• Polydisperse linear alkane systems with average lengths of 24, 36, 50
and 78 carbon atoms were investigated
• Temperature effects were also investigated (300K, 350K, 400K and
450K)
• A uniaxial orienting field was applied
 
 0 0 
 xx 
 xx
α 0  0 
 2 
  
 0 0  xx 
 2 

• Simulations were run at constant temperature and constant pressure


P=1atm
Molecular Simulations

 Background
• The conformation tensor is defined as the second moment of the end-to-
end vector R
c  RR
• The normalized conformation tensor is:
~c  c 3
 2
R
0
• The overall chain spring constant is then defined as:

K T   k BT  2
3k BT
R
0

 
• The “orienting field” α:    1  ~ A  , T , c 
k BT   c N  T ,  ,~c  
Molecular Simulations

 Thermodynamic Considerations
• How do we test the validity of PEE under this framework?

A
 , T , c  A0  , T , I   A  U  TS
N ch N ch N ch N ch N ch

• The steps involved in accomplishing this task include:


 Evaluate ΔA via thermodynamic integration

A A     M 1 1 
 0  k BT   c d   b     k BT : c
N ch N ch    T ,b ,   N A  0 
0

 Evaluate ΔU directly from simulation


U
 U tot  , T , c   U tot  , T , I 
N ch
Molecular Simulations

 Potential Model Details


• Siepmann-Karaborni-Smit (SKS) force field
U  U bond  U angle  U torsion  U nonbonded

  12   6 
Rigid rIJ  r0 U nonbonded  4      
 r   r  

k    0 
1
U angle 
2

2
σ
ε
3
U torsion   a k cos  
k

k 0

U nonbonded  U intra  U inter


Equilibrium Simulations

 The equilibrium mean-squared end-to-end distance R


2
0

• Used in the evaluation of the conformation tensor normalization factor


and the chain spring constant
3k B T

3
K T   2
 k B T
2
R R
0 0

• Can be evaluated for the entire molecular weight distribution interval

• Its molecular weight dependence can be fitted to a polynomial function


proposed by Mavrantzas and Theodorou [1]

R2 1 2 3
CX  0
 0   
 X  1b 2 X 1  X  12  X  13

1. Mavrantzas, V.G. and D.N. Theodorou, Macromolecules, 1998. 31(18): p. 6310-6332.


Equilibrium Simulations

 The equilibrium mean-squared end-to-end distance R


2
0

• All systems at T = 450K


3000
C24
2500 C36
C50
2000 C78
Polynomial Fit
Å 
2
0

1500
R2

1000

500

0
0 20 40 60 80 100 120 140
X (#C Atoms)
Equilibrium Simulations

 The equilibrium mean-squared end-to-end distance R


2
0

• The polynomial fitting constants

Temperature α0 α1 α2 α3
450K 8.8427 -77.9066 521.951 -2141.85
400K 8.6677 -30.5968 -681.681 6030.121
350K 9.219 -9.298 -1573.36 13064.89
300K 11.9351 -183.756 2022.19 -9320.51
450K ref. [1] 9.1312 -75.1865 315.742 -500.3518

• For polyethylene, the measured characteristic ratio at T = 413K [2]

C  7.8  0.4
1. Mavrantzas, V.G. and D.N. Theodorou, Macromolecules, 1998. 31(18): p. 6310-6332.
2. Fetters, L.J., W.W. Graessley, R. Krishnamoorti, and D.J. Lohse, Macromolecules, 1997. 30(17): p. 4973-4977.
Equilibrium Simulations

 The equilibrium mean-squared end-to-end distance R


2
0

• Polynomial fits, all temperatures


3500
300K
3000 350K
400K
2500 450K
Ref. [1]
Å 
2

2000
0
R2

1500

1000

500

0
0 20 40 60 80 100 120 140
X (# C Atoms)

1. Mavrantzas, V.G. and D.N. Theodorou, Macromolecules, 1998. 31(18): p. 6310-6332.


Equilibrium Simulations

 The conformation tensor normalization factor μ


• Usually taken as a constant with respect to temperature (PEE
assumption)
• Gupta and Metzner [1] proposed the following for the temperature
dependence of μ

  T  B 1

• This expression was used to fit our equilibrium simulation data


with great success

1. Gupta, R.K. and A.B. Metzner, Journal of Rheology, 1982. 26(2): p. 181-198.
Equilibrium Simulations

 Theoretical considerations for the behavior of μ with respect to


temperature
• If B= - 1, μ is a constant and K(T) is a linear function of temperature
K T  
  T  B 1   K T   kBT 
 K T   T 0
 T 
 The configurational part of the internal energy density of a fluid particle given by
the UCMM vanishes
 K T  
  K T   T  tr c   0
1
u
2  T 

• If B< - 1, μ increases with temperature and R 0 decreases with


2

temperature
3
  T  B 1
R2
0

 The configurational part of the internal energy density of a fluid particle given by
the UCMM may become important at high degrees of orientation
Equilibrium Simulations

 The temperature exponent B


-0.9

-1 Temperature Exponent B
Extrapolation
-1.1

-1.2
B

-1.3

-1.4

-1.5

-1.6

-1.7
0 20 40 60 80 100 120 140
X (#C Atoms)
Non-equilibrium Simulations

 The applied “orienting field” α:


 
 0 0 
 xx 
 xx
α 0  0 
 2 
  
 0 0  xx 
 2 
• The magnitude of αxx will uniquely describe the “strength” of the
orienting field
 Following the definition of α, the conformation tensor will also have a
diagonal form
  A  1   G 
  
1
 ~  , T , c ~
c    b, T ,  
k BT 
 
c N  T ,  ,~c   k BT    N ch  T ,b,[  ]

• Therefore, the trace of the conformation tensor may be used as a


unique descriptor for the degree of orientation and extension developed
in the simulations
Non-equilibrium Simulations

 Molecular weight dependence of the degree of orientation


• All systems, T = 450K
16

14 C78

12

10
C50
c

8
tr ~

C36
6
C24
4

 xx
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Non-equilibrium Simulations

 Temperature dependence of the degree of orientation


• C36 system , all temperatures

5
tr ~c 

4
300K
3
350K
2 400K
450K
1

 xx
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Non-equilibrium Simulations

 Energy balances for the oriented systems


• All systems, T = 450K

20
A / N ch
10
Energy Change (J/g)

0 C24 C36 C50 C78

-10

-20

U / N ch
-30

-40
0 0.1 0.2 0.3  xx 0.4 0.5 0.6 0.7 0.8
Non-equilibrium Simulations

 Energy balances for the oriented systems


• C36 system, all temperatures

20 b)
A / N ch
10
Energy Change (J/g)

0 450K 400K 350K 300K

-10

-20
U / N ch
-30

-40
0 0.1 0.2 0.3  xx 0.4 0.5 0.6 0.7 0.8
Non-equilibrium Simulations

 Internal energy broken down into individual components


• C24 system, T = 400K
U total U angle U torsion U intra U inter
   
N ch N ch N ch N ch N ch

5
U intra / N ch
0
U angle / N ch
Energy Change (J/g)

-5

-10
U torsion / N ch
-15
U inter / N ch
-20

-25
U total / N ch
-30
0 0.2
 xx 0.4 0.6 0.8
Non-equilibrium Simulations

 The UCMM prediction for the change in Helmholtz free energy


A 1
 k B T tr ~
c   3  k B T ln det ~
c 
1
N ch 2 2

25

UCMM
20
Integration
A / N ch J / g 

15

10

0
0 0.1 0.2 0.3
 xx 0.4 0.5 0.6 0.7 0.8
Non-equilibrium Simulations

 The conformational part of the heat capacity


• The MW dependence, T = 450K
0.1

-0.1 C24

cconf Jmol 1K 1

-0.2 C36

-0.3
C50
-0.4

-0.5 C78

-0.6
0 0.1 0.2 0.3
 xx 0.4 0.5 0.6 0.7 0.8
Non-equilibrium Simulations

 The conformational part of the heat capacity


• The temperature dependence, C36 system
0.05

450K
0
400K
-0.05 350K
300K
cconf Jmol 1K 1

-0.1

-0.15

-0.2

-0.25

-0.3

-0.35
0 0.1 0.2  xx 0.4
0.3 0.5 0.6 0.7 0.8
Part III: Summary

 Equilibrium simulations
• Revealed a non-linear dependence of K(T) with respect to
temperature
• Improved agreement with experiment in terms of the
characteristic ratio C∞ and temperature exponent B

 Non-equilibrium simulations
• The changes in free energy and internal energy are of similar
magnitude
• The examination of the individual components of the internal
energy provided two useful insights
 The elastic response of single chains is indeed purely entropic
 The inter-molecular contribution to the internal energy of an ensemble of
chains (missing in the isolated chain case) is very important and
explains the trends observed during the experiments
Part IV: Published Research

 “Structure Formation under Steady-State Isothermal Planar Elongational Flow of


n-eicosane: A Comparison between Simulation and Experiment”[1]
• First, we examined the liquid structure predicted by simulation under equilibrium
conditions
 Simulation performed in the NVT ensemble (number of particles N, system volume V and
temperature T are kept constant)
 The state point was chosen the same as in the experiment case (T = 315K and ρ = 0.81 g/cc),
and the experimental data were taken from literature (**)

3
Experimental (*)
2.5 Simulation

2
s(k)

1.5

0.5

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
k (1/Å)
1. Ionescu, T.C., et al.,. Physical Review Letters, 2006. 96(3).
(*) A. Habenschuss and A.H. Narten, J. Chem. Phys., 92, 5692 (1990)
Simulated Elongated Structure

 Next, we examined the structure when the flow field is turned on at steady-state
in terms of the pair correlation function
• The applied velocity gradient is of the form:   0 0 
 
u   0   0 

 0 0 0
 

• Results shown at a reduced elongation rate (m /  ) =1.0


2 1/ 2

• The state point was the same as in the equilibrium case (T=315K and ρ = 0.81
g/cc)

10 1.4
9 Quiescent Melt Quiescent Melt
1.2
8 Elongated Structure Elongated Structure
7 1
6 g_inter(r)
0.8
g(r)

5
4 0.6
3 0.4
2
0.2
1
0 0
0 1 2 3 4 5 6 7 8 9 10 11 12 0 2 4 6 8 10 12 14 16 18
r (Å) r (Å)
Simulated Elongated Structure

 Same structural data, in terms of the static structure factor s(k)

3
Quiescent Melt
2.5
Elongated Structure
2
3
s(k)

1.5 Quiescent Melt


2.5 Elongated Structure
1
2
0.5

0 s_inter(k)
1.5

0 2 4 6 8 10 12 14 16
k (1/Å) 1

0.5

0
0 2 4 6 8 10 12 14 16
k (1/Å)
Comparison with Experiment

 The structure factor s(k) determined via x-ray diffraction from the n-eicosane
crystalline sample
• Identify two regions:
 Inter-molecular region (k<6Å-1), where sharp Bragg peaks are present
 Intra-molecular region (k>6Å-1), where the agreement with simulation is excellent

3.00
Liquid XRD (*)
2.50
Crystal XRD (this work)
3.00
2.00
Simulated Elongated Structure
2.50
s(k)

1.50 Crystal XRD

2.00
1.00

s(k)
0.50 1.50

0.00 1.00
0 2 4 6 8 10 12 14 16
k (1/Å)
0.50

0.00
0 2 4 6 8 10 12 14 16
k (1/Å)

(*) A. Habenschuss and A.H. Narten, J. Chem. Phys., 92, 5692 (1990).
Conclusions and Directions for
Future Research

 Conclusions
• We successfully combined experiment, theory and simulation to
investigate the nature of the free energy stored by polymer melts
subjected to deformation

• First, it was shown that the Theory of Purely Entropic Elasticity is


applicable to polymer melts only at low deformation rates

• Second, molecular theory (the UCMM) was used to propose a


recipe for eliminating the PEE assumption with great results

• In the end, the Molecular Simulation study helped us elucidate


the trends observed in the experimental part
 The simulated conformational dependent heat capacity was found in
good qualitative agreement with experiment
Conclusions and Directions for
Future Research

 Directions for Future Research


• More polymers and processing conditions

 Effects of molecular characteristics

• More accurate viscoelastic models

• Our work in Part IV already led to the development of a constant


pressure version of the NEMD algorithm used

 Longer chain systems and different flow situations also worth


investigating
Acknowledgements

 My advisors, Drs. Brian Edwards and David Keffer

 Dr. Vlasis Mavrantzas

 Dr. Simioan Petrovan

 Doug Fielden

 ORNL – Cheetah and UT SInRG Cluster

 PRF, Grant No. 41000-AC7


Questions?
Experimental Approach

 Same analysis performed for shear flow using a capillary die


• LDPE, Tin = Twall = 170oC, D = 1mm, L = 25mm

  150s 1   4500s 1
Experimental Approach

 Shear flow temperature profile in the measurement device

  4500s 1

S-ar putea să vă placă și