4.11. Rheometry

4.11.1. Material Data Parameters

This chapter describes the fluid models available in Fluent Materials Processing, including details about the related material data parameters for each.

4.11.1.1. Overview of Fluid Properties and Flow Characteristics

The proper selection of a fluid model is one of the most important aspects in the simulation of a flow. You need to always consider both the fluid and the flow. A particular constitutive equation is valid for a given fluid in a given flow.

To determine an appropriate model for your problem, you need to first collect as much data as possible about the fluid properties. Typical information includes the following:

  • Steady viscometric properties (shear viscosity and first normal-stress difference ). These data characterize the fluid in the presence of large deformations.

  • Oscillatory viscometric properties (storage and loss moduli and ), also known as linear viscoelastic data because they correspond to small deformations.

  • Elongational viscosity. Although obtaining data on elongation is difficult and not very frequent, knowledge of the elongational viscosity can be of interest if the process involves a visible elongation component, for example, fiber spinning. This may be useful in choosing the appropriate constitutive equation and estimating the values of the various parameters.

These data are not enough to evaluate the relevance of viscoelasticity in a given process. It is also necessary to characterize the flow itself and compare a characteristic time of the material to a characteristic time of the flow. In many situations, the flow can be characterized by a typical shear rate, which can be understood as a wall shear rate in a region of high gradients. For example, in a fiber-spinning process, a typical shear rate will occur at the wall in the vicinity of the die exit. In a contraction or expansion flow (for example, Figure 4.123: Contraction and Expansion Flow), consider the shear rate in the narrow section.

Figure 4.123: Contraction and Expansion Flow

Contraction and Expansion Flow

In a planar flow (Figure 4.123: Contraction and Expansion Flow a),

(4–24)

where is a typical distance.

In an axisymmetric flow (Figure 4.123: Contraction and Expansion Flow b),

(4–25)

where is a typical radius.

You also need to determine the elasticity level of the fluid. This can be accomplished by evaluation of the fluid’s characteristic relaxation time. When the oscillatory functions and are available, their intersection (occurring at a shear rate , as shown in Figure 4.124: Storage and Loss Moduli Curves) is often a reasonable choice for selecting a typical relaxation time. Indeed, flows characterized by a typical shear rate lower than are essentially dominated by viscous forces, while viscoelastic effects may play an important role in flows characterized by a shear rate higher than .

Note that, due to the technological limitations of some rheometry equipment, it is not always possible to obtain viscoelastic data in the range of shear rates (or angular frequencies) where the process operates. In this case, your only option is to extrapolate experimental data for higher shear rates or angular frequencies. The selection of a particular model for such a case will be more qualitative.

Figure 4.124: Storage and Loss Moduli Curves

Storage and Loss Moduli Curves

A typical dimensionless number used to estimate the viscoelastic character of a flow is the Weissenberg number , which is the product of the relaxation time and a typical shear rate :

(4–26)

When is low, generalized Newtonian models are sufficient to describe the flow; only at higher values of are viscoelastic models required to characterize memory effects.

Note that the Weissenberg number is probably not the best indicator for viscoelastic models with several relaxation times or if there is shear thinning in the flow. In such cases, a useful dimensionless number is the recoverable shear , defined as the ratio of the first normal-stress difference to twice the steady shear stress :

(4–27)

The recoverable shear gives the level of elasticity of a flow: if >1, the viscoelastic character of the flow is important, and a viscoelastic model is required.

The different fluid models are studied in different flows described in Rheological Properties.

Plots the rheometric curves for different models and flows. Carefully investigate the scales used for those plots as logarithm scales, linear scales for the logarithm of the quantities or a mixed of linear and logarithm scales can be used.

4.11.1.2. Generalized Newtonian Model

4.11.1.2.1. Introduction
4.11.1.2.1.1. Equations

For a generalized Newtonian fluid, the constitutive equation has the form

(4–28)

where is the extra-stress tensor, is the rate-of-deformation tensor, and is the viscosity, which can depend upon both the second invariant of and the temperature .

The general form for the viscosity is written as

(4–29)

where is the local shear rate. Therefore, and represent the shear-rate and temperature dependence of the viscosity, respectively.

4.11.1.2.1.2. Inputs

To specify a newtonian model, you need to first select Generalized Newtonian model.

 Rheometry Fluid Model Model Type

Specify an appropriate Fluid Name.

Choose the Shear Rate Dependence and Temperature Dependence law.

See Non-Automatic Fitting and Automatic Fitting for information about where and how the material data specification occurs in the non-automatic and automatic fitting procedures, respectively.

See Shear Rate Dependence of Viscosity and Temperature Dependence of Viscosity for details about the parameters and characteristics of each fluid model.

4.11.1.2.2. Shear Rate Dependence of Viscosity

There are currently 10 laws available for .

 Rheometry Fluid Model Viscosity LawShear Rate Dependence

4.11.1.2.2.1. Constant

 Rheometry Fluid Model Viscosity LawShear Rate DependenceConstant

For Newtonian fluids, a constant viscosity

(4–30)

is the default setting. is referred to as the Newtonian or zero-shear-rate viscosity, and its default value is 1.

The units for and its name in the Fluent Materials Processing interface are as follows:

ParameterName in MatPro Rheometry toolMass Length Time
Viscosity [Pa s] 1–1–1

Figure 4.125: Constant (Shear-Rate-Independent) Viscosity shows a plot of a constant .

Figure 4.125: Constant (Shear-Rate-Independent) Viscosity

Constant (Shear-Rate-Independent) Viscosity

4.11.1.2.2.2. Bird-Carreau Law

 Rheometry Fluid Model Viscosity LawShear Rate DependenceBird-Carreau

The Bird-Carreau law for viscosity is

(4–31)

where is the infinite-shear-rate viscosity, is the zero-shear-rate viscosity, is natural time (that is, inverse of the shear rate at which the fluid changes from Newtonian to power-law behavior) and is the power-law index.

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in MatPro Rheometry toolMass Length Time
Zero Shear Viscosity [Pa s] 1 –1–1
Infinite Shear Viscosity [Pa s] 1 –1–1
Time Constant [s] 1
Power Law Index

By default, and are equal to 1, and and are equal to 0. Figure 4.126: Bird-Carreau Law for Viscosity shows a plot of a for the Bird-Carreau law.

Figure 4.126: Bird-Carreau Law for Viscosity

Bird-Carreau Law for Viscosity

The Bird-Carreau law is commonly used when it is necessary to describe the low-shear-rate behavior of the viscosity. It differs from the Cross law primarily in the curvature of the viscosity curve in the vicinity of the transition between the plateau zone and the power law behavior.

4.11.1.2.2.3. Power Law

 Rheometry Fluid Model Viscosity LawShear Rate DependencePower

The power law for viscosity is

(4–32)

where is the consistency factor, is the natural time, and is the power-law index, which is a property of a given material.

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in MatPro Rheometry toolMass Length Time
Consistency Factor [Pa s] 1–1–1
Time Constant [s]   1
Power Law Index

By default, , , and are equal to 1. Figure 4.127: Power Law for Viscosity shows a plot of for the power law.

Figure 4.127: Power Law for Viscosity

Power Law for Viscosity

The power law is commonly used for the algebraic simplicity of its formulation. Although it can be a good candidate for describing the behavior of polyethylene or rubber for shear rates spanning over 2 or 3 decades, it does not predict any plateau for low shear rates. Consequently, it can be a convenient choice for the analysis of internal flow, but it should preferably be avoided when analyzing extrusion flows.

4.11.1.2.2.4. Bingham Law

 Rheometry Fluid Model Viscosity LawShear Rate DependenceBingham

The Bingham law for viscosity is

(4–33)

where is the yield stress and is the critical shear rate, beyond which Bingham’s constitutive equation is applied. For shear rates less than , the behavior of the fluid is normalized in order to mimic as much as possible a solid body and to guarantee appropriate continuity properties in the viscosity curve.

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in MatPro Rheometry toolMass Length Time
Plastic Viscosity [Pa s] 1 –1–1
Yield Stress Threshold [Pa] 1 –1–2
Critical Shear Rate [s^-1] –1

By default, , , and are equal to 1. Figure 4.128: Bingham Law for Viscosity shows a plot of for the Bingham law.

The Bingham law is commonly used to describe materials such as concrete, mud, dough, and toothpaste, for which a constant viscosity after a critical shear stress is a reasonable assumption, typically at rather low shear rates.

Figure 4.128: Bingham Law for Viscosity

Bingham Law for Viscosity

4.11.1.2.2.5. Modified Bingham Law

 Rheometry Fluid Model Viscosity LawShear Rate Dependencemodified Bingham

A modified Bingham law for viscosity is also available:

(4–34)

where .

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in Rheometry Tool Mass Length Time
Plastic Viscosity [Pa s] 1 –1–1
Yield Stress Threshold [Pa] 1 –1–2
Critical Shear Rate [s^-1] –1

By default, , , and are equal to 1. Figure 4.129: Modified Bingham Law for Viscosity shows a plot of for the modified Bingham law.

Figure 4.129: Modified Bingham Law for Viscosity

Modified Bingham Law for Viscosity

Compared to the standard Bingham law, the modified Bingham law is an analytic expression, which means that it may be easier for Fluent Materials Processing to calculate, leading to a more stable solution. The value has been selected so that the standard and modified Bingham laws exhibit the same behavior above the critical shear rate, .

4.11.1.2.2.6. Herschel-Bulkley Law

 Rheometry Fluid Model Viscosity LawShear Rate DependenceHerschel-Bulkley

The Herschel-Bulkley law for viscosity is

(4–35)

where is the yield stress, is the critical shear rate, is the consistency factor, and is the power-law index.

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in Rheometry Tool Mass Length Time
Yield Stress Threshold [Pa] 1 –1–2
Consistency Factor [Pa s] 1 –1–1
Critical Shear Rate [s^-1] –1
Power Law Index

By default, , , , and are equal to 1. Figure 4.130: Herschel-Bulkley Law for Viscosity shows a plot of for the Herschel-Bulkley law.

Like the Bingham law, the Herschel-Bulkley law is commonly used to describe materials such as concrete, mud, dough, and toothpaste, for which a constant viscosity after a critical shear stress is a reasonable assumption. In addition to the transition behavior between a flow and no-flow regime, the Herschel-Bulkley law exhibits a shear-thinning behavior that the Bingham law does not.

Figure 4.130: Herschel-Bulkley Law for Viscosity

Herschel-Bulkley Law for Viscosity


Note:  An oblique arrow on the right of a symbol indicates how the symbol changes, and vertical/horizontal arrows indicate how the curve transforms as the symbol changes.


4.11.1.2.2.7. Modified Herschel-Bulkley Law

 Rheometry Fluid Model Viscosity LawShear Rate Dependencemodified Herschel-Bulkley

A modified Herschel-Bulkley law is also available:

(4–36)

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in Rheometry ToolMass Length Time
Yield Stress Threshold [Pa] 1 –1–2
Consistency Factor [Pa s] 1 –1–1
Critical Shear Rate [s^-1] –1
Power Law Index

By default, , , , and are equal to 1. Figure 4.131: Modified Herschel-Bulkley Law for Viscosity shows a plot of for the modified Herschel-Bulkley law.

Figure 4.131: Modified Herschel-Bulkley Law for Viscosity

Modified Herschel-Bulkley Law for Viscosity

Compared to the standard Herschel-Bulkley law, Figure 4.130: Herschel-Bulkley Law for Viscosity, the modified Herschel-Bulkley law is an analytic expression, which means that it may be easier for Fluent Materials Processing to calculate, leading to a more stable solution. The integer value 3 that appears in the argument of the exponential term has been selected so that the standard and modified Herschel-Bulkley laws exhibit the same behavior above the critical shear rate, .

4.11.1.2.2.8. Cross Law

 Rheometry Fluid Model Viscosity LawShear Rate DependenceCross

The Cross law for viscosity is

(4–37)

where is the zero-shear-rate viscosity, is the natural time (that is, inverse of the shear rate at which the fluid changes from Newtonian to power-law behavior) and is the Cross-law index (= 1– for large shear rates).

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in Fluent Materials Processing Mass Length Time
Zero Shear Viscosity [Pa s] 1 1 –1
Time Constant [s]   1
Cross Law Index

By default, is equal to 1, and and are equal to 0. Figure 4.132: Cross Law for Viscosity shows a plot of for the Cross law.

Figure 4.132: Cross Law for Viscosity

Cross Law for Viscosity

Like the Bird-Carreau law, the Cross law is commonly used when it is necessary to describe the low-shear-rate behavior of the viscosity. It differs from the Bird-Carreau law primarily in the curvature of the viscosity curve in the vicinity of the transition between the plateau zone and the power law behavior.

4.11.1.2.2.9. Modified Cross Law

 Rheometry Fluid Model Viscosity LawShear Rate Dependencemodified Cross

A modified Cross law for viscosity is also available:

(4–38)

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in Fluent Materials Processing Mass Length Time
Zero Shear Viscosity [Pa s] 1 1 –1
Time Constant [s]   1
Cross Law Index

By default, is equal to 1, and and are equal to 0. Figure 4.133: Modified Cross Law for Viscosity shows a plot of for the Cross law.

Figure 4.133: Modified Cross Law for Viscosity

Modified Cross Law for Viscosity

This law can be considered a special case of the Carreau-Yasuda viscosity law (Equation 4–39), where the exponent has a value of 1.

4.11.1.2.2.10. Carreau-Yasuda Law

 Rheometry Fluid Model Viscosity LawShear Rate DependenceCarreau-Yasuda

The Carreau-Yasuda law for viscosity is:

(4–39)

where is the zero-shear-rate viscosity, is the infinite-shear-rate viscosity, is the natural time (that is, inverse of the shear rate at which the fluid changes from Newtonian to power-law behavior), is the index that controls the transition from the Newtonian plateau to the power-law region and is the power-law index.

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in Fluent Materials Processing Mass Length Time
Zero Shear Viscosity [Pa s] 1–1–1
Infinite Shear Viscosity [Pa s] 1–1–1
Time Constant [s]   1
Power Law Index
Plateau Index

By default, , , and are equal to 1, and and are equal to 0. Figure 4.134: Carreau-Yasuda Law for Viscosity shows a plot of for the Carreau-Yasuda law.

Figure 4.134: Carreau-Yasuda Law for Viscosity

Carreau-Yasuda Law for Viscosity

The Carreau-Yasuda law is a slight variation on the Bird-Carreau law (Equation 4–31). The addition of the exponent allows for control of the transition from the Newtonian plateau to the power-law region. A low value ( < 1) lengthens the transition, and a high value (>1) results in an abrupt transition.

4.11.1.2.3. Temperature Dependence of Viscosity

 Rheometry Fluid Model Viscosity LawTemperature Dependence

As discussed in Introduction, the general form for the viscosity can be written as the product of functions of shear rate and temperature. There are two ways in which this relationship can be expressed:

(4–40)

(4–41)

where and represent the shear-rate and temperature dependence of the viscosity, respectively.

In Equation 4–40, the temperature scales the viscosity so there is only a vertical shift on the model curves vs. temperature. Four of the temperature-dependent laws follow this format:

  • Arrhenius approximate law

  • Arrhenius law

  • Fulcher law

  • WLF law

In Equation 4–41, the time-temperature equivalence is introduced by also scaling the shear rate by temperature. Therefore, there is a horizontal shift in addition to the vertical shift on the model curves vs. temperature. Three of the temperature-dependent viscosity laws follow this format:

  • Arrhenius approximate law

  • Arrhenius law

  • WLF law

By default, there is no temperature dependence of the viscosity (that is, ).

4.11.1.2.3.1. Arrhenius Law

 Rheometry Fluid Model Viscosity LawTemperature DependenceArrhenius

The Arrhenius law is given as

(4–42)

where is the ratio of the activation energy to the perfect gas constant and is a reference temperature for which .

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in MatPro Rheometry ToolMass Length Time Temperature
Ea/R [K] 1
Reference Temperature [K] 1

By default, Ea/R is set to 0 and the Reference Temperature to 300 [K]. Figure 4.135: Arrhenius Law for the Temperature Dependence of the Viscosity With Vertical Shift Only shows a plot of for the Arrhenius law.

Figure 4.135: Arrhenius Law for the Temperature Dependence of the Viscosity With Vertical Shift Only

Arrhenius Law for the Temperature Dependence of the Viscosity With Vertical Shift Only

When the Arrhenius law is selected for the temperature dependence of the viscosity, you have the option of selecting the shift that is applied such as a vertical shift only or a combination of both vertical and horizontal shifts. If the vertical shift is selected, the viscosity curve will be shifted vertically, downwards or upwards, subsequently to a temperature increase or decrease, respectively. Figure 4.135: Arrhenius Law for the Temperature Dependence of the Viscosity With Vertical Shift Only suggests a vertical shift only, where Equation 4–40 is used. Figure 4.136: Approximate Arrhenius Law for the Temperature Dependence of the Viscosity With Both Vertical and Horizontal Shifts suggests a combination of both vertical and horizontal shifts, where Equation 4–41 is applied.

4.11.1.2.3.2. Approximate Arrhenius Law

 Rheometry Fluid Model Viscosity LawTemperature DependenceArrhenius approximate

The approximate Arrhenius law is obtained as the first-order Taylor expansion of the Arrhenius law (Equation 4–42):

(4–43)

Where is the ratio of the activation energy to the perfect gas constant and is a reference temperature for which H(T) = 1. The behavior described by Equation 4–43 is similar to that described by Equation 4–42 in the neighborhood of . Equation 4–43 is valid as long as the temperature difference is not too large.

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in MatPro Rheometry ToolMass Length Time Temperature
Ea/R [K] 1
Reference Temperature [K] 1

By default, Ea/R is set to 0 and the Reference Temperature is set to 300 [K]. Figure 4.136: Approximate Arrhenius Law for the Temperature Dependence of the Viscosity With Both Vertical and Horizontal Shifts shows a plot of () for the approximate Arrhenius law.

Figure 4.136: Approximate Arrhenius Law for the Temperature Dependence of the Viscosity With Both Vertical and Horizontal Shifts

Approximate Arrhenius Law for the Temperature Dependence of the Viscosity With Both Vertical and Horizontal Shifts

When the approximate Arrhenius law is selected for the temperature dependence of the viscosity, you have the option of selecting the shift that is applied such as a vertical shift only or a combination of both vertical and horizontal shifts. If the vertical shift is selected, the viscosity curve will be shifted vertically, downwards or upwards, subsequently to a temperature increase or decrease, respectively. Figure 4.135: Arrhenius Law for the Temperature Dependence of the Viscosity With Vertical Shift Only suggests a vertical shift only, where equation (Equation 4–40) is used. Figure 4.136: Approximate Arrhenius Law for the Temperature Dependence of the Viscosity With Both Vertical and Horizontal Shifts suggests a combination of both vertical and horizontal shifts, where equation (Equation 4–41) is applied.

4.11.1.2.3.3. Fulcher Law

 Rheometry Fluid Model Viscosity LawTemperature DependenceFulcher

Another definition for comes from the Fulcher law  5:

(4–44)

where , , and are the Fulcher constants. The Fulcher law is used mainly for glass.

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in MatPro Rheometry ToolMass Length Time Temperature
F1
F2 [K] 1
F3 [K] 1

By default, , and are equal to 0.

Although the Fulcher temperature dependence law can be combined with a non-constant shear-rate dependence, it is originally developed for modelling the temperature dependence of the viscosity for molten glass. Hence, it is usually combined with a constant (unity) viscosity factor. You can give the following interpretation for the three coefficients , and .

As illustrated in Figure 4.137: Typical Viscosity Curve vs. Temperature, very high temperature are encountered. You start with , which corresponds to the temperature where the viscosity is infinite. The law can no longer be applied if the fluid temperature is lower than . For preventing issues, a cut-off has been introduced so that the calculation is not hindered when temperatures less than are encountered.

If increases, the overall viscosity curve decreases. Eventually, an increase of leads to a more visible dependence of the viscosity curve with respect to the temperature.

Figure 4.137: Typical Viscosity Curve vs. Temperature

Typical Viscosity Curve vs. Temperature

In the image above, the blue line is obeying the Fulcher Law and the viscosity can be bounded to prevent numerical issues (orange line).

In Figure 4.137: Typical Viscosity Curve vs. Temperature, you plot the Fulcher viscosity curve vs. temperature. Very high values are obtained at low temperatures. The law as such does not permit considering temperatures below the value of . To prevent numerical difficulties which may originate from very high values, an upper bound of 1014 has been assigned to the viscosity by default. That value is sufficient for mimicking the behavior of a body which is partly solidified.

4.11.1.2.3.4. WLF Law

 Rheometry Fluid Model Viscosity LawTemperature DependenceWLF

The Williams-Landel-Ferry (WLF) equation is a temperature-dependent viscosity law that fits experimental data better than the Arrhenius law for a wide range of temperatures, especially close to the glass transition temperature:

(4–45)

where and are the WLF constants, and and are reference temperatures.

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in MatPro Rheometry ToolMass Length Time Temperature
C1
C2 [K] 1
Reference Temperature [K] 1
Reference Temperature Difference [K] 1

Figure 4.138: Effect of Increasing c2 on the WLF Law for Viscosity and Figure 4.139: Effect of Increasing c1 or Ta on the WLF Law for Viscosity show the impact of each parameter on the viscosity curves. A large value of will enhance the decrease of the viscosity with respect to temperature, while a larger value of will spread out the viscosity dependence with respect to temperature when it is around the reference temperature value.

It is important to note that the quantity must remain positive in a flow simulation.

Figure 4.138: Effect of Increasing c2 on the WLF Law for Viscosity

Effect of Increasing c2 on the WLF Law for Viscosity

Figure 4.139: Effect of Increasing c1 or Ta on the WLF Law for Viscosity

Effect of Increasing c1 or Ta on the WLF Law for Viscosity

When the WLF law is selected for the temperature dependence of the viscosity, you have the option of selecting the shift that is applied such as a vertical shift only or a combination of both vertical and horizontal shifts. If the vertical shift is selected, the viscosity curve will be shifted vertically, downwards, or upwards, subsequently to a temperature increase or decrease, respectively. Figure 4.136: Approximate Arrhenius Law for the Temperature Dependence of the Viscosity With Both Vertical and Horizontal Shifts suggests a vertical shift only, where equation (Equation 4–40) is used.

4.11.1.3. Differential Viscoelastic Model

4.11.1.3.1. Introduction

The differential approach to modeling viscoelastic model is appropriate for most practical applications. Many of the most common constitutive models for viscoelastic model are provided in Fluent Materials Processing, including Maxwell, Oldroyd, Phan-Thien-Tanner, Giesekus, FENE-P, POM-POM, and Leonov models. Appropriate choices for the viscoelastic model and related parameters can yield qualitatively and quantitatively accurate representations of viscoelastic behavior.

Improved accuracy is possible if you use multiple relaxation times to better fit the viscoelastic behavior at different shear rates.


Note:  While differential viscoelastic models are compatible with 2D and 3D models, they are not compatible with the shell model.


4.11.1.3.1.1. Equations

For a differential viscoelastic model, the constitutive equation for the extra-stress tensor is

(4–46)

(the viscoelastic component) is computed differently for each type of viscoelastic model. (the purely viscous component) is an optional component, which is always computed from

(4–47)

where is the rate-of-deformation tensor and is the viscosity factor for the Newtonian (that is, purely viscous) component of the extra-stress-tensor referred to as the additional viscosity.

4.11.1.3.1.2. Inputs

To specify a viscoelastic model, you need to first select Differential Viscoelastic model.

 Rheometry Fluid Model Model Type

Specify an appropriate Fluid Name.

Choose the Model and specify Number of Relaxation Modes and Additional Viscosity [Pa s].

For each Mode, specify Relaxation Time [s], Partial Viscosity [Pa s] and Alpha.

Finally, select the law and parameters for the Thermal Dependency.

See Non-Automatic Fitting and Automatic Fitting for information about where and how the material data specification occurs in the non-automatic and automatic fitting procedures, respectively.

See Differential Viscoelastic Models and Temperature Dependence of Viscosity and Relaxation Time for details about the parameters and characteristics of each fluid model.

4.11.1.3.2. Differential Viscoelastic Models

 Rheometry Fluid Model Differential Viscoelastic Properties Model

There are currently eight differential models:

  • Maxwell and Oldroyd-B:

    These are the simplest viscoelastic constitutive equations, although in many situations they are the most numerically cumbersome. Both models exhibit a constant viscosity and a quadratic first normal-stress difference. They should be selected either when very little information is known about the fluid, or when a qualitative prediction is sufficient. For fluids exhibiting a very high extensional viscosity, the Oldroyd-B model is preferred over the Maxwell model.

  • Phan-Thien-Tanner, Johnson-Segalman, and Giesekus:

    These models are the most realistic. They exhibit shear thinning and a non-quadratic first normal-stress difference at high shear rates. These properties are controlled by their respective material parameters (, , and ), as described in the model description below. Also, the selection of nonzero values for and will lead to a bounded steady extensional viscosity.

    For stability reasons in a simple shear flow, a non-zero additional viscosity must be selected. This is true for the Johnson-Segalman and Phan-Thien-Tanner models when is nonzero, and for the Giesekus model when >0.5. The addition of a purely viscous component to the extra-stress tensor affects the viscosity, but not the first normal-stress difference. Shear thinning is still present, but the viscosity curve also shows a plateau zone at high shear rates.

    Poor control of the shear viscosity is the usual drawback of the Johnson-Segalman and Phan-Thien-Tanner models used with a single relaxation time, especially toward high shear rates.


    Important:  Note that you cannot explicitly select the Johnson-Segalman model in the Fluent Materials Processing interface. It is obtained by selecting the Phan-Thien-Tanner model and setting the value of to 0.


  • FENE-P:

    A single mode of the FENE-P model requires only three parameters (, and the length ratio for the spring), yet it predicts a realistic shear thinning of the fluid and a first normal-stress difference that is quadratic for low shear rates and has a lower slope for high shear rates. It has been observed in practice that viscometric properties of several fluids can often be accurately modeled. The FENE-P model is well suited for simulating the rheological behavior of dilute solutions.

  • POM-POM:

    The pom-pom molecule consists of a backbone to which arms are connected at both extremities. In a flow, the backbone may orient in a Doi-Edwards reptation tube consisting of the neighboring molecules, while the arms may retract into that tube. The concept of the pom-pom macromolecule makes the model suitable for describing the behavior of branched polymers. The approximate differential form of the model is based on the equations of macromolecular orientation and macromolecular stretching in connection with changes in orientation. In this construction, the pom-pom molecule is allowed only a finite extension, which is controlled by the number of dangling arms. In particular, the strain hardening properties are dictated by the number of arms. Beyond that, the model predicts realistic shear thinning behavior, as well as a first and a possible second normal stress difference.

  • Leonov:

    This model has been developed for the simultaneous prediction of the behavior of trapped and free macromolecular chains for filled elastomers with carbon black and/or silicate. From the point of view of morphology, macromolecules at rest are trapped by particles of carbon black, via electrostatic van der Waals forces. Under a deformation field, electrostatic bonds can break, and macromolecules become free, while a reverse mechanism may develop when the deformation ceases. You can therefore be facing a macromolecular system consisting of trapped and free macromolecules, with a reversible transition from one state to the other one.

    This model involves two tensor quantities and a scalar one. The tensors focus respectively on the behavior of the free and trapped macromolecular chains of the elastomer, while the scalar variable quantifies the degree of structural damage (debonding factor). The model exhibits a yielding behavior. It is intrinsically nonlinear, as the nonlinear response develops and is observable at early deformations.

Details about each model are provided below.

4.11.1.3.2.1. Maxwell Model

 Rheometry Fluid Model Differential Viscoelastic Properties Model Maxwell

The Maxwell model is one of the simplest viscoelastic constitutive equations. It exhibits a constant viscosity and a quadratic first normal-stress difference. Due to its simplicity, it is recommended only when little information about the fluid is available, or when a qualitative prediction is sufficient. Even in this case, the Oldroyd-B model, which can include a purely viscous component, is preferable for numerical reasons.

4.11.1.3.2.1.1. Equations

The equation for the Maxwell model is the upper-convected Maxwell model, in which the purely viscous component of the extra-stress tensor () is equal to zero. For single-mode, the viscoelastic component () is computed from

(4–48)

where is a model-specific relaxation time, is the rate-of-deformation tensor, and is a model-specific viscosity factor for the viscoelastic component of . The relaxation time is defined as the time required for the shear stress to be reduced to half of its original equilibrium value when the strain rate vanishes. A high relaxation time indicates that the memory retention of the flow is high. A low relaxation time indicates significant memory loss, gradually approaching Newtonian flow (zero relaxation time).

4.11.1.3.2.1.2. Inputs

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in Fluent Materials Processing Mass Length Time
Partial Viscosity [Pa s] 1 –1–1
Relaxation Time [s] 1

By default, is set to 1 and to 10-16.

Additionally, by default, Number of Relaxation Modes is set to 1. You can select multiple relaxation modes, in which case multiple sets of parameters will have to be specified (, ).

4.11.1.3.2.1.3. Behavior Analysis

Figure 4.140: Maxwell Model for a Shear Flow shows the viscometric functions of the Maxwell model in a simple shear flow. In this example (where =1 s and =1000 Pa-s), is constant, is linear, is quadratic, is zero, is constant, is zero, and is linear, showing non-asymptotic behavior.

Figure 4.140: Maxwell Model for a Shear Flow

Maxwell Model for a Shear Flow

Figure 4.141: Maxwell Model for an Extensional Flow shows the behavior of the Maxwell model in a simple extensional flow.

Figure 4.141: Maxwell Model for an Extensional Flow

Maxwell Model for an Extensional Flow

In this example (where =1 s and =1000 Pa-s), , , and are unbounded for , and

(4–49)

(4–50)

(4–51)

(4–52)

Figure 4.142: Maxwell Model for a Transient Shear Flow shows the behavior of the Maxwell model in a transient shear flow.

Figure 4.142: Maxwell Model for a Transient Shear Flow

Maxwell Model for a Transient Shear Flow

In this example (where =1 s, =1000 Pa-s, and  s-1), there is no stress overshoot and the transient phase depends upon the relaxation time.

4.11.1.3.2.2. Oldroyd-B Model

 Rheometry Fluid Model Differential Viscoelastic Properties Model Oldroyd-B

The Oldroyd-B model, like the Maxwell model, is one of the simplest viscoelastic constitutive equations. It is slightly better than the Maxwell model, because it allows for the inclusion of the purely viscous component of the extra stress, which leads to better behavior of the numerical scheme. Oldroyd-B is a good choice for fluids that exhibit a very high extensional viscosity.

4.11.1.3.2.2.1. Equations

For the Oldroyd-B model, is computed from Equation 4–48, and is computed (optionally) from Equation 4–47. from Equation 4–48 is the partial viscosity, while in Equation 4–47 is the additional viscosity

4.11.1.3.2.2.2. Inputs

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in Fluent Materials Processing Mass Length Time
Partial Viscosity [Pa s] 1 –1–1
Relaxation Time [s] 1
Additional Viscosity [Pa s] 1 –1–1

By default, is set to 1 while is set to 10-16. is set to 0.

Additionally, by default, Number of Relaxation Modes is set to 1. You can select multiple relaxation modes, in which case multiple sets of parameters will have to be specified (, ).

4.11.1.3.2.2.3. Behavior Analysis

Figure 4.143: Oldroyd-B Model for a Shear Flow shows the viscometric functions of the Oldroyd-B model in a simple shear flow. In this example, =1 s and (with the viscosity ratio equal to 0.15) =850 Pa-s and =150 Pa-s. In the resulting curves, is constant, is linear, is quadratic, is zero, is constant, is zero, and is linear, showing non-asymptotic behavior. Notice that the curves are moved down in comparison to the Maxwell model; this is due to the Newtonian part of the model (nonzero value for ), which reduces the viscoelastic effects (, , , and ).

Figure 4.143: Oldroyd-B Model for a Shear Flow

Oldroyd-B Model for a Shear Flow

Figure 4.144: Oldroyd-B Model for a Transient Shear Flow shows the behavior of the Oldroyd-B model in a transient shear flow. In this example, =1 s, =1000 Pa-s, and  s–1. Notice that there is an instantaneous response of the shear stress to the applied shear rate; this is due to the Newtonian part of the model originating from the additional viscosity whose value was set to 150 Pa.s. Otherwise, the Oldroyd-B model exhibits the same behavior as the Maxwell model.

Figure 4.144: Oldroyd-B Model for a Transient Shear Flow

Oldroyd-B Model for a Transient Shear Flow

4.11.1.3.2.3. Phan-Thien-Tanner Model

 Rheometry Fluid Model Differential Viscoelastic Properties Model Phan-Thien-Tanner

The Phan-Thien-Tanner model is one of the most realistic differential viscoelastic models. It exhibits shear thinning and a non-quadratic first normal-stress difference at high shear rates.

4.11.1.3.2.3.1. Equations

The PTT model computes from

(4–53)

and is computed (optionally) from Equation 4–47. in Equation 4–53 is the partial viscosity, while in Equation 4–47 is the additional viscosity.

and are material properties that control, respectively, the shear viscosity and elongational behavior. A nonzero value for leads to a bounded steady extensional viscosity.

4.11.1.3.2.3.2. Inputs

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in Fluent Materials Processing Mass Length Time
Partial Viscosity [Pa s] 1 –1–1
Relaxation Time [s] 1
Additional Viscosity [Pa s] 1 –1–1
Epsilon
Xi

By default, is set to 1 while is set to 10-16. is set to 0. and are also set to 0 by default.

Additionally, by default, Number of Relaxation Modes is set to 1. You can select multiple relaxation modes, in which case multiple sets of parameters will have to be specified (, , , ).

4.11.1.3.2.3.3. Behavior Analysis

In a simple shear flow (Figure 4.145: PTT Model for a Shear Flow), for >0, you can see a shear-thinning effect and a non-quadratic behavior for the first normal-stress difference . You will also notice that, for >0, the elasticity level remains finite for increasing shear rate (asymptotic behavior).

Figure 4.145: PTT Model for a Shear Flow

PTT Model for a Shear Flow

The parameter also affects the extensional viscosities, as shown in Figure 4.146: PTT Model for a Steady Extensional Flow. The steady extensional viscosities are finite, and tend toward the Newtonian component of the extensional viscosity (that is, they are uniaxial) for large extension rates. For small values of , there is extension thickening and thinning; for large values, there is only extension thinning.

Figure 4.146: PTT Model for a Steady Extensional Flow

PTT Model for a Steady Extensional Flow


Important:  For a single-mode PTT model, if the parameter is not zero, then the additional viscosity must be at least 1/8 of the Partial Viscosity in order to ensure the stability of the shear flow. However, this value may decrease when does not vanish. The slope of the shear stress vs. shear rate curve must be positive everywhere, contrary to what is shown on the left in Figure 4.147: Effect of ξ on the PTT Model for a Shear Flow with =0.1.


Figure 4.147: Effect of ξ on the PTT Model for a Shear Flow

Effect of ξ on the PTT Model for a Shear Flow

The parameter has almost no effect on extensional viscosity, as shown in Figure 4.148: Effect of ξ on the PTT Model for a Steady Extensional Flow. The maximum of the extensional viscosities decreases when increases.

Figure 4.148: Effect of ξ on the PTT Model for a Steady Extensional Flow

Effect of ξ on the PTT Model for a Steady Extensional Flow

In a transient shear flow (Figure 4.149: PTT Model in a Transient Shear Flow), a moderate stress overshoot is observed. The stress overshoot increases as shear rate increases. Shear thinning is observed, and the normal stress is non-quadratic. The transient phase is reduced as the shear rate increases.

Figure 4.149: PTT Model in a Transient Shear Flow

PTT Model in a Transient Shear Flow

4.11.1.3.2.4. Giesekus Model

 Rheometry Fluid Model Differential Viscoelastic Properties Model Giesekus

Like the PTT model, the Giesekus model is one of the most realistic differential viscoelastic models. It exhibits shear thinning and a non-quadratic first normal-stress difference at high shear rates.

4.11.1.3.2.4.1. Equations

The Giesekus model computes from

(4–54)

and is computed (optionally) from Equation 4–47. in Equation 4–54 is the partial viscosity, while in Equation 4–47 is the additional viscosity.

is the unit tensor and is a material constant that controls the extensional viscosity and the ratio of the second normal-stress difference to the first. For low values of shear rate,

(4–55)

For the majority of fluids, this ratio is between 0.1 and 0.2.

4.11.1.3.2.4.2. Inputs

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in Fluent Materials Processing Mass Length Time
Partial Viscosity [Pa s] 1 –1–1
Relaxation Time [s] 1
Additional Viscosity [Pa s] 1 –1–1
Alpha

By default, is set to 1 while is set to 10-16. is set to 0. Additionally, is also set to 0 and Number of Relaxation Modes to 1. You can select multiple relaxation modes, in which case multiple sets of parameters will have to be specified (, , ).

4.11.1.3.2.4.3. Behavior Analysis

In a simple shear flow (Figure 4.150: Giesekus Model for a Shear Flow), controls the shear-thinning effect. The first normal-stress difference is non-quadratic, and the cut-off appears earlier if increases. If >0.5, you should specify a non-zero Additional Viscosity [Pa s] in order to avoid instabilities.

Figure 4.150: Giesekus Model for a Shear Flow

Giesekus Model for a Shear Flow

Figure 4.151: Effect of α on the Giesekus Model for an Extensional Flow shows the behavior of the Giesekus fluid in an extensional flow.

Figure 4.151: Effect of α on the Giesekus Model for an Extensional Flow

Effect of α on the Giesekus Model for an Extensional Flow

In this case, the steady extensional viscosities are finite. For small values of extension thickening occurs, and for large values extension thinning occurs.

In a transient shear flow (Figure 4.152: Giesekus Model for a Transient Shear Flow), the stress overshoot is less severe than for the PTT model; there are fewer oscillations.

Figure 4.152: Giesekus Model for a Transient Shear Flow

Giesekus Model for a Transient Shear Flow

The duration of the transient phase depends on the imposed shear rate (the same behavior as for the PTT model). For a high shear rate, you can observe stress overshoots during the transient phase. With increasing shear rate, the overshoot increases while the final value of the displayed properties decreases. The duration of the transient phases decreases as the shear rate increases.

4.11.1.3.2.5. FENE-P Model

 Rheometry Fluid Model Differential Viscoelastic Properties Model Fene-P

The FENE-P model is derived from molecular theories and assumes that the polymer macromolecules are idealized as dumbbells linked with an elastic connector or spring and suspended in a Newtonian solvent of viscosity . Unlike in the Maxwell model, however, the springs are allowed only a finite extension, so that the energy of deformation of the dumbbell becomes infinite for a finite value of the spring elongation. This model predicts a realistic shear thinning of the fluid and a first normal-stress difference that is quadratic for low shear rates and has a lower slope for high shear rates.

4.11.1.3.2.5.1. Equations

The FENE-P model computes from

(4–56)

where the configuration tensor is computed from

(4–57)

and is the ratio of the maximum length of the spring to its length at rest:

(4–58)

is an equilibrium length that corresponds to rigid motion (in this case, =0 and the tension in the connector equals the Brownian forces). is the maximum allowable dumbbell length. Figure 4.153: Dumbbell Definitions for the FENE-P Model shows how the distance between dumbbells is based on the relative position of both ends.

Figure 4.153: Dumbbell Definitions for the FENE-P Model

Dumbbell Definitions for the FENE-P Model

is always greater than 1. As becomes infinite, the FENE-P model reduces to the upper-convected Maxwell model.

is computed (optionally) from Equation 4–47. in Equation 4–56 is the partial viscosity, while in Equation 4–47 is the additional viscosity.

The motion of the dumbbells is the result of hydrodynamic, Brownian, and spring forces. represents the tension in the spring (spring forces) and the Brownian motion. represents the Newtonian (hydrodynamic) forces.

See 1 for additional information about the FENE-P model.

4.11.1.3.2.5.2. Inputs

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in Fluent Materials Processing Mass Length Time
Partial Viscosity [Pa s] 1 –1–1
Relaxation Time [s] 1
Additional Viscosity [Pa s] 1 –1–1
L^2

By default, is set to 1 while is set to 10-16. is set to 1.1 and to 0. Additionally, by default, Number of Relaxation Modes is set to 1. You can select multiple relaxation modes, in which case multiple sets of parameters will have to be specified (, , ).

4.11.1.3.2.5.3. Behavior Analysis

The behavior of the FENE-P model with small values of for a simple shear flow is illustrated in Figure 4.154: Effect of Small Values of L^2 on the FENE-P Model for Shear Flow. Shear thinning occurs with this model, and for large values of shear rate, the slope is –2/3. Therefore the addition of a Newtonian viscosity component is not required for stability. The first normal-stress difference is non-quadratic, and the second normal-stress difference is 0. The cut-off appears sooner when decreases, down to a value of 3. No asymptotic behavior is observed. For low values of shear rate, decreases as decreases.

Figure 4.154: Effect of Small Values of L^2 on the FENE-P Model for Shear Flow

Effect of Small Values of L^2 on the FENE-P Model for Shear Flow

The behavior of the FENE-P model with large values of for a simple shear flow is illustrated in Figure 4.155: Effect of Large Values of L^2 on the FENE-P Model for Shear Flow.

Figure 4.155: Effect of Large Values of L^2 on the FENE-P Model for Shear Flow

Effect of Large Values of L^2 on the FENE-P Model for Shear Flow

For large values of , the FENE-P model is observed to exhibit Maxwellian behavior: quadratic first normal-stress difference and close to . For close to 1, Newtonian behavior is observed: quadratic but small first normal-stress difference, tends toward 0, cut-off occurs at high shear rates. For low shear rates,

(4–59)

For extensional flows, controls the extensional viscosity. As shown in Figure 4.156: Effect of L^2 on the FENE-P Model for Extensional Flow, the extensional viscosities are finite. For large values of , the FENE-P model is observed to exhibit Maxwellian behavior: the extensional viscosities are very high for . For close to 1, Newtonian behavior is observed: the extensional viscosities are constant.

Figure 4.156: Effect of L^2 on the FENE-P Model for Extensional Flow

Effect of L^2 on the FENE-P Model for Extensional Flow

The behavior of the FENE-P model for a transient shear flow is shown in Figure 4.157: Effect of Large Values of L^2 on the FENE-P Model for Transient Shear Flow and Figure 4.158: Effect of Mid-Range Values of L^2 on the FENE-P Model for Transient Shear Flow. For high shear rates, the stress overshoots in the transient phase. When the shear rate increases, the final value and the transient phase decrease while the overshoot increases. For large values of , the FENE-P model is observed to exhibit Maxwellian behavior: no stress overshoots. For mid-range values of , the stress overshoots increase and the transient phase decreases as decreases. For close to 1, Newtonian behavior is observed: no stress overshoots and a short transient phase even for high values of shear rate.

Figure 4.157: Effect of Large Values of L^2 on the FENE-P Model for Transient Shear Flow

Effect of Large Values of L^2 on the FENE-P Model for Transient Shear Flow

Figure 4.158: Effect of Mid-Range Values of L^2 on the FENE-P Model for Transient Shear Flow

Effect of Mid-Range Values of L^2 on the FENE-P Model for Transient Shear Flow

4.11.1.3.2.6. POMPOM Model [DCPP]

 Rheometry Fluid Model Differential Viscoelastic Properties Model POMPOM

In the POMPOM model, the pom-pom molecule consists of a backbone to which arms are connected at both extremities. In a flow, the backbone may orient in a Doi-Edwards reptation tube consisting of the neighboring molecules, while the arms may retract into that tube. The concept of the pom-pom macromolecule makes the model suitable for describing the behavior of branched polymers. The approximate differential form of the model is based on equations of macromolecular orientation, and macromolecular stretching in relation to changes in orientation.

The model, referred to as DCPP (2, 8), allows for a nonzero second normal stress difference. The DCPP model computes from an orientation tensor, and a stretching scalar (states variables), on the basis of the following algebraic equation:

(4–60)

where is the shear modulus and is a nonlinear material parameter (the nonlinear material parameter will be introduced later on). The state variables and are computed from the following differential equations:

(4–61)

(4–62)

In these equations, and are the relaxation times associated with the orientation and stretching mechanisms respectively. In the last equation, characterizes the number of dangling arms (or priority) at the extremities of the pom-pom molecule or segment. It is an indication of the maximum stretching that the molecule can undergo, and therefore of a possible strain hardening behavior. can be obtained from the elongational behavior. is a nonlinear parameter that has enabled the introduction of a non-vanishing second normal stress difference in the DCPP model.

A multi-mode DCPP model can also be defined. Each contribution will involve an orientation tensor and a stretching variable . A few guidelines are required for the determination of the several linear and nonlinear parameters.

Consider a multi-mode DCPP model characterized by modes sorted with increasing values of relaxation times (increasing seniority). The linear parameters and characterizing the linear viscoelastic behavior of the model can be determined with the usual procedure.

Then the relaxation times () for stretching should be determined. Depending on the average number of entanglements of backbone section, the ratio should be within the range of 2 to 10. For a completely unentangled polymer segment, you may accept the physical limit of =. should also satisfy the constraint , since sets the fundamental diffusion time for the branch point controlling the relaxation of polymer segment ().

The parameter indicating the number of dangling arms (or priority) at the extremities of a pom-pom segment , also indicates the maximum stretching that can be undergone by that segment, and therefore its possible strain hardening behavior. For a multi-mode DCPP model, both seniority and priority are assumed to increase together towards the inner segments; hence should also increase with . The parameter can be obtained from the elongational behavior.

is a fifth set of nonlinear parameters that control the ratio of second to first normal stress differences. The value of parameter should range between 0 and 1. For moderate values, corresponds to twice the ratio of the second to the first normal stress difference, and may decrease with increasing seniority.

As for other viscoelastic models, a purely viscous component can be added to the viscoelastic component , in order to get the total extra-stress tensor:

(4–63)

where

(4–64)

where is the rate-of-deformation tensor and is the additional (Newtonian) viscosity.

4.11.1.3.2.6.1. Inputs

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in Fluent Materials Processing Mass Length Time
Additional Viscosity [Pa s] 1–1–1
Relaxation Time for Orientation [s] - - 1
Shear Modulus [Pa] 1–1–2
Relaxation Time for Stretching [s] - - 1
Number of Arms - - -
Xi - - -

By default, the shear modulus and the relaxation time for orientation are respectively initialized to 1 and 10-16. The relaxation time for stretching and xi are initialized to 0 while Number of Arms is initialized to 2. Additionally, by default, Number of Relaxation Modes is set to 1. You can select multiple relaxation modes, in which case multiple sets of parameters will have to be specified (, , , , ).

4.11.1.3.2.6.2. Behavior Analysis

Figure 4.159: Effect of Parameter ξ for Steady Shear Flow shows the steady viscometric behavior of a single mode DCPP fluid model for various values of the parameter . For the present illustration, the shear modulus equals 1000, while the relaxation times for orientation and stretching have been assigned the values 1 and 0.5, respectively. As can be seen, constant viscosity and quadratic first normal stress difference are obtained at low shear rates. Nonlinear behavior is found beyond . We also find that an increasing value of enforces the nonlinear behavior, while it also generates a non-vanishing second normal stress difference. The other nonlinear parameters and have actually a negligible influence on the viscometric properties.

Figure 4.159: Effect of Parameter ξ for Steady Shear Flow

Effect of Parameter ξ for Steady Shear Flow

In Figure 4.160: Effect of Parameter q on Steady Elongation Viscosity, the steady elongation viscosity of a single mode DCPP fluid model for increasing values of is displayed. For the continuous curves, the shear modulus equals 1000, while the relaxation times for orientation and stretching have been assigned the values 1 and 0.5, respectively. Also, the nonlinear parameter is equal to 0.1. As is known for the DCPP model, and more generally for pom-pom models, the parameter is an indication of branching, and therefore of strain hardening in elongation. As can be seen from Figure 4.160: Effect of Parameter q on Steady Elongation Viscosity, the elongation viscosity increases when the strain rate is larger than , and the strain hardening is enhanced for increasing values of . The figure also shows the steady elongation viscosity obtained for as well as for . As can be seen, the influence of these parameters on the steady elongation viscosity remains moderate as compared to that of parameter .

Figure 4.160: Effect of Parameter q on Steady Elongation Viscosity

Effect of Parameter q on Steady Elongation Viscosity

Finally, Figure 4.161: Effect of Parameter q on Transient Elongation Viscosity for Different Values of the Elongation Rate shows the transient elongation viscosity of various single-mode DCPP fluid model characterized by different branching levels (), at elongation rates successively equal to 0.1, 1 and 10. We find that all curves collapse at low strain rate (0.1), while they markedly differ at high strain rate (10).

Figure 4.161: Effect of Parameter q on Transient Elongation Viscosity for Different Values of the Elongation Rate

Effect of Parameter q on Transient Elongation Viscosity for Different Values of the Elongation Rate

4.11.1.3.2.7. Leonov Model

 Rheometry Fluid Model Differential Viscoelastic Properties Model Leonov

Elastomers are usually filled with carbon black and/or silicate. From the point of view of morphology, macromolecules at rest are trapped by particles of carbon black, via electrostatic van der Waals forces. Under a deformation field, electrostatic bonds can break, and macromolecules become free, while a reverse mechanism may develop when the deformation ceases. You can therefore be facing a macromolecular system consisting of trapped and free macromolecules, with a reversible transition from one state to the other one.

Leonov and Simhambhatla have developed a rheological model (9, 3) for the simultaneous prediction of the behavior for trapped and free macromolecular chains. This model for filled elastomers involves two tensor quantities and a scalar one. These tensor quantities focus respectively on the behavior of the free and trapped macromolecular chains of the elastomer, while the scalar quantity quantifies the degree of structural damage (debonding factor). The model exhibits a yielding behavior. It is intrinsically nonlinear, as the nonlinear response develops and is observable at early deformations.

In a single-mode approach, the total stress tensor can be decomposed as the sum of free and trapped contributions as follows:

(4–65)

As for other viscoelastic models, a purely viscous component is added to the viscoelastic components in order to get the total extra-stress tensor:

(4–66)

where is the rate-of-deformation tensor and is the viscosity.

In Equation 4–65, subscripts and respectively refer to the free and trapped parts. Each of these contributions obeys its own equation. In particular, they invoke their own deformation field described by means of Finger tensors.

An elastic Finger tensor is defined for the free chains, which obeys the following equation:

(4–67)

where is the relaxation time, is the unit tensor, while and are the first invariant of and , respectively, defined as

(4–68)

(4–69)

The implemented material function that appears in Equation 4–67 is written as follows:

(4–70)

The parameter must be and increases slightly the amount of shear thinning.

Similarly, an elastic Finger tensor is defined for the trapped chains, which obeys the following equation:

(4–71)

where and are the first invariant of and , respectively, defined as

(4–72)

(4–73)

In the equation for the trapped chains, the variable quantifies the degree of structural damage (debonding factor), and is the fraction of the initially trapped chains that are debonded from the filler particles during flow. The function is a structural damage dependent scaling factor for the relaxation time and is referred to as the “mobility function".

A phenomenological kinetic equation is suggested for :

(4–74)

In Equation 4–74, is the local shear rate while is the yielding strain. Also, is a dimensionless time factor, which may delay or accelerate debonding.

For the mobility function appearing in Equation 4–71, the following form has been implemented:

(4–75)

The above selection for the mobility function endows the rheological properties with a yielding behavior. When is large (or unbounded), the algebraic term dominates the constitutive equation for (Equation 4–71), and the solution is expected to be . When is vanishing, becomes governed by a purely transport equation; this may lead to numerical troubles when solving a complex steady flow with secondary motions (vortices). This situation can occur if parameter is set to zero and under no-debonding situation (). Therefore, you should impose a small (but nonzero) value for parameter (by default, we suggest the value 0.05, which is a reasonable compromise between rheological properties and solver stability). Based on this, parameter can be understood as the value of the mobility function under no-debonding.

Finally, in order to relate the Finger tensors to the corresponding stress tensor, potential functions are required. For and , the following expressions are suggested:

(4–76)

(4–77)

with and . It is interesting to note that has no effect on the shear viscosity, while it contributes to a decrease of the elongational viscosity. On the other hand, the parameter increases both shear and elongational viscosities. From there, stress contributions from free and trapped chains in Equation 4–65 are respectively given by:

(4–78)

(4–79)

where parameter is the initial ratio of free to trapped chains in the system. A vanishing value of indicates that all chains are trapped at rest, while a large value of indicates a system that essentially consists of free chains.

4.11.1.3.2.7.1. Inputs

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in Fluent Materials Processing Mass Length Time
Additional Viscosity [Pa s] 1–1–1
Relaxation Time [s] - - 1
Shear Modulus [Pa] 1–1–2
Alpha - - -
Beta - - -
N- - -
M - - -
Nu - - -
K - - -
Q - - -
Gamma* - - -

By default, is set to 0, is set to 10-16, and are set to 1, , and are set to 0, is set to 2, is set to 0, is set to 1 and is set to 2. Additionally, by default, Number of Relaxation Modes is set to 1. You can select multiple relaxation modes, in which case multiple sets of parameters will have to be specified (, , , , , , , , , ).

4.11.1.3.2.7.2. Identification of Model Parameters and Functions

From the point of view of rheology and numerical simulation, for single- and multi-mode fluid models, a purely viscous contribution must be added to the total extra-stress tensor. Actually, this is largely motivated by the fact that the matrix of the discretized system can be singular when all fields are initialized to values that correspond to the solution at rest. Hence, the first or only mode will always be accompanied by a Newtonian contribution, whose corresponding viscosity value received a unit default value. This value can be modified you.

Also, as suggested above, a non-vanishing value should be selected for the mobility function under no-debonding.

As can be seen, next to parameters and controlling the linear properties, the model involves two functions and several nonlinear parameters. In a single mode approach, the influence of these parameters on the viscometric and elongational properties can be easily identified, and appropriate values can be selected accordingly. By default, the nonlinear parameters are assigned values that are relevant from the point of view of rheology. In a multi-mode approach, in order to facilitate the definition of a flow case, corresponding nonlinear parameters should preferably be identical for each mode.

4.11.1.3.2.7.3. Behavior Analysis

In simple shear flow, the Leonov model exhibits shear thinning, which is slightly affected by some parameters. Figure 4.162: Shear Viscosity of the Leonov Model with Parameters G=1000, λ=1, q=1, β=0, ν=2, γ*=2, and α=1, k=n=m=0 (continuous lines). shows that an increase of the parameter (initial ratio of free to trapped chains) slightly decreases the shear viscosity at low shear rates. This can be easily understood if you consider, for example, that when =0, the material consists only of trapped chains at rest. The figure also shows that parameter increases the shear viscosity at high shear rates, while parameter has a very limited influence. Finally, as can be seen in Figure 4.162: Shear Viscosity of the Leonov Model with Parameters G=1000, λ=1, q=1, β=0, ν=2, γ*=2, and α=1, k=n=m=0 (continuous lines)., shear viscosity curves do not show a plateau at low shear rates. This is the fingerprint of the yielding behavior of the fluid model, which is controlled by the value of the mobility function under no-debonding (parameter ). If increases, the viscosity curves exhibit a plateau at low shear rates. However, as can be seen in the insert, this does not affect the behavior at high shear rates, while it may improve the stability of the solver.

Figure 4.162: Shear Viscosity of the Leonov Model with Parameters G=1000, λ=1, q=1, β=0, ν=2, γ*=2, and α=1, k=n=m=0 (continuous lines).

Shear Viscosity of the Leonov Model with Parameters G=1000, λ=1, q=1, β=0, ν=2, γ*=2, and α=1, k=n=m=0 (continuous lines).

Dashed and dashed-dotted lines show the viscosity for the value of the parameters as indicated. The insert shows the viscosity curves obtained for various values of the mobility function under no-debonding (parameter k).

Figure 4.163: First Normal Stress Difference of the Leonov Model with Parameters G=1000, λ=1, q=1, β=0, ν=2, γ*=2, and α=1, k=n=m=0 (continuous lines). shows that similar trends are found for the first normal stress difference. Figure 4.163: First Normal Stress Difference of the Leonov Model with Parameters G=1000, λ=1, q=1, β=0, ν=2, γ*=2, and α=1, k=n=m=0 (continuous lines). shows that an increase of the parameter slightly decreases the first normal stress difference at all shear rates. The figure also shows that parameter increases the first normal stress difference at all shear rates, while parameter decreases it at high shear rates. Finally, as can be seen, the first normal stress difference shows a plateau at low shear rates; this is a counterpart of the yielding behavior of the fluid model, which is also controlled by the value of the mobility function under no-debonding (parameter ). If increases, the first normal stress difference exhibit a quadratic behavior at low shear rates; however, as can be seen in the insert, this does not affect the behavior at high shear rates, while it may improve the stability of the solver.

Figure 4.163: First Normal Stress Difference of the Leonov Model with Parameters G=1000, λ=1, q=1, β=0, ν=2, γ*=2, and α=1, k=n=m=0 (continuous lines).

First Normal Stress Difference of the Leonov Model with Parameters G=1000, λ=1, q=1, β=0, ν=2, γ*=2, and α=1, k=n=m=0 (continuous lines).

Dashed, dashed-dotted and dotted lines show the first normal stress difference for the value of the parameters as indicated. The insert shows the curves of first normal stress difference obtained for various values of the mobility function under no-debonding (parameter ).

In simple elongation flow, the Leonov model exhibits marked strain thinning at low strain rates; it is slightly affected by some parameters. Figure 4.164: Elongation Viscosity of the Leonov Model with Parameters G=1000, λ=1, q=1, n=1, ν=2, γ*=2, and α=1, β=k=m=0 (continuous lines). shows that an increase of the parameter (initial ratio of free to trapped chains) slightly decreases the elongation viscosity at low strain rates. This can be easily understood if you consider, for example, that when =0, the material consists only of trapped chains at rest. The figure also shows that parameter increases the elongation viscosity at high strain rates, while parameters and decrease the elongation viscosity. Finally, as can be seen in Figure 4.164: Elongation Viscosity of the Leonov Model with Parameters G=1000, λ=1, q=1, n=1, ν=2, γ*=2, and α=1, β=k=m=0 (continuous lines)., elongation viscosity curves do not show a plateau. This is the fingerprint of the yielding behavior of the fluid model, which is controlled by the value of the mobility function under no-debonding (parameter ). Actually, if increases, the elongation viscosity curves exhibit a plateau at low strain rates. However, as can be seen in the insert of Figure 4.164: Elongation Viscosity of the Leonov Model with Parameters G=1000, λ=1, q=1, n=1, ν=2, γ*=2, and α=1, β=k=m=0 (continuous lines)., this does not really affect the behavior at high strain rates while it may improve the stability of the solver.

Figure 4.164: Elongation Viscosity of the Leonov Model with Parameters G=1000, λ=1, q=1, n=1, ν=2, γ*=2, and α=1, β=k=m=0 (continuous lines).

Elongation Viscosity of the Leonov Model with Parameters G=1000, λ=1, q=1, n=1, ν=2, γ*=2, and α=1, β=k=m=0 (continuous lines).

Dashed, dashed-dotted and dotted lines show the elongation viscosity for the value of the parameters as indicated. The insert shows the curves of the steady elongation viscosity obtained for various values of the mobility function under no-debonding (parameter ).


Note:  These curves are not obtained from Fluent Materials Processing Rheometry tool. They result from semi-analytical calculations.


Figure 4.165: Transient Shear Viscosity of the Leonov Model Versus Time, at Shear Rates Ranging from 10^-2 to 10, With Parameters G=1000, λ=1, q=1, n=1,, ν=2, γ*=2, and α=1, β=k=m=n=0, (continuous lines). shows the transient shear viscosity versus time at shear rates ranging from 10-2 to 10, for various values of parameters and . At first, as can be seen, the transient shear viscosity exhibits an overshoot before reaching the steady value. It is also interesting to note that the response time decreases when the shear rate increases. This actually results from the increasing mobility function under increasing shear rates. Eventually, we find that parameter decreases the elongation viscosity, while the other parameters have a somewhat less marked influence.

Figure 4.165: Transient Shear Viscosity of the Leonov Model Versus Time, at Shear Rates Ranging from 10^-2 to 10, With Parameters G=1000, λ=1, q=1, n=1,, ν=2, γ*=2, and α=1, β=k=m=n=0, (continuous lines).

Transient Shear Viscosity of the Leonov Model Versus Time, at Shear Rates Ranging from 10^-2 to 10, With Parameters G=1000, λ=1, q=1, n=1,, ν=2, γ*=2, and α=1, β=k=m=n=0, (continuous lines).

Dashed and dotted lines show the viscosity for the value of the parameters as indicated.

4.11.1.3.3. Temperature Dependence of Viscosity and Relaxation Time

The viscosity in a non-isothermal differential viscoelastic model can be temperature-dependent. As described in Introduction, the viscosity will be multiplied by a temperature shift function . For non-isothermal differential viscoelastic models, the relaxation time is multiplied by the same temperature shift function. Temperature-dependent functions available for non-isothermal differential viscoelastic models are the Arrhenius law, the Arrhenius approximate law, and the WLF law, all described in Temperature Dependence of Viscosity.

4.11.1.3.4. Multiple Relaxation Times for Differential Viscoelastic Models

If you define multiple relaxation modes for your differential viscoelastic fluid, all modes will obey the same constitutive law. Non-linear parameters can be identical or independently specified for each mode, there is no strong argument for one or the other option. For multiple relaxation modes, the viscoelastic stress is written as follows:

(4–80)

A purely viscous component can also be added.

Consider, for example, a fluid being modeled with two relaxation modes as follows:

  • mode 1: PTT model, =0.1 s,  Pa-s, =0.2,

  • mode 2: PTT model, =1 s,  Pa-s, =0.2,

Figure 4.166: Simple Shear Flow with Multiple Relaxation Times shows the viscometric behavior for this fluid in a simple shear flow. Here, the cut-off is controlled by the larger relaxation time, and the slopes of the curves (for shear rates between 1/ and 1/) are affected.

Figure 4.166: Simple Shear Flow with Multiple Relaxation Times

Simple Shear Flow with Multiple Relaxation Times

Figure 4.167: Extensional Flow with Multiple Relaxation Times shows the viscometric behavior for this fluid in an extensional flow. The slopes of the curves (for extension rates between 1/ and ) are affected.

Figure 4.167: Extensional Flow with Multiple Relaxation Times

Extensional Flow with Multiple Relaxation Times

Figure 4.168: Transient Shear Flow with Multiple Relaxation Times shows the viscometric behavior for this fluid in a transient shear flow. The multiple relaxation modes have an effect on the transient phase. In this case, the length of the transient phase depends upon the larger relaxation time. The multiple relaxation modes also affect the overshoot magnitude.

Figure 4.168: Transient Shear Flow with Multiple Relaxation Times

Transient Shear Flow with Multiple Relaxation Times

4.11.1.4. Integral Viscoelastic Model

4.11.1.4.1. Introduction

While the differential models are reasonably suited for 2D and 3D applications, integral constitutive models can be invoked for describing the behavior of melts in applications where the shell model can be used, such as blow molding and thermoforming. These processes are characterized by an elongation kinematics with relatively limited extension deformation.


Note:  The integral approach to modeling viscoelastic model is limited to shell models.


4.11.1.4.1.1. Equations

For an integral viscoelastic constitutive equation, the extra-stress tensor is computed at time from the following equation:

(4–81)

where is the model-specific memory (kernel) function , is the Cauchy-Green strain tensor , is the current time and is the metric for time integrals.

For non-isothermal flows, can be computed from the isothermal constitutive equation (Equation 4–81), provided that a modified time scale is used for evaluating the strain history:

(4–82)

The modified time scale is related to through the following equation:

(4–83)

where is the shift function, which can be obtained from steady-state shear-viscosity curves at different temperatures. This is the principle of time-temperature equivalence.

4.11.1.4.1.2. Inputs

To specify an integral viscoelastic model, select Integral viscoelastic model.

 Rheometry Fluid Model Model Type

Specify an appropriate Fluid Name.

Specify Number of Relaxation Modes and Additional Viscosity [Pa s].

Specify Relaxation Time [s] and the Partial Viscosity [Pa s].

Finally, select the law and parameters for the Thermal Dependency.

See Non-Automatic Fitting and Automatic Fitting for information about where and how the material data specification occurs in the non-automatic and automatic fitting procedures, respectively.

See Integral Viscoelastic Models and Temperature Dependence of Viscosity for details about the parameters and characteristics of each fluid model.

4.11.1.4.2. Integral Viscoelastic Models

The Integral viscoelastic model is primarily dedicated for the simulation of processes like blow molding and thermoforming. These processes involve a kinematics characterized by extension with a bounded deformation amplitude, while shear is absent. Therefore, damping functions are discarded. Linear properties are often sufficient for characterizing the behavior of the melt in such processes. The linear spectrum already allows the prediction of strain hardening and since deformation involved in the shaping processes are often bounded to Hencky strain up to 1 or 2, it is not necessary to mimic the behavior beyond these values.

4.11.1.4.3. Temperature Dependence of Viscosity

 Rheometry Fluid Model Thermal Dependency Temperature Dependence

Three models are available for the temperature shift function in Equation 4–83: the Arrhenius law, the Arrhenius approximate law, and the WLF law, all described in Temperature Dependence of Viscosity. It is also possible to eliminate the temperature dependence using a temperature shift function equal to 1.

4.11.1.5. Simplified Viscoelastic Model

One of the interesting features of viscoelastic model simulations is the prediction of extrudate swelling, which can be larger than their Newtonian counterparts. Running complex 3D flows with a rheologically sophisticated model, however, can be computationally expensive. Therefore, a more simplified approach is needed: one in which you can qualitatively predict the extrudate free surface. This approach is referred to as the “simplified viscoelastic model" or the “light viscoelastic model".

4.11.1.5.1. Equations

It is known that the first normal stress difference is mainly responsible for enhanced extrudate swell in extrusion flow. This is typically a viscoelastic property. With respect to this, the simplified viscoelastic model is an extension of existing Newtonian fluid models, where a normal stress difference has been incorporated into the force balance. That is, in simple shear flow along the first axis and with a shear rate , the total extra-stress tensor is given by:

(4–84)

In this tensor, is the shear stress component, which involves the shear rate dependent viscosity . Several laws are available for describing the shear viscosity (see Generalized Newtonian Model for more details), for instance, the constant law (Equation 4–30), the Bird-Carreau law (Equation 4–31), the Power law (Equation 4–32), the Cross law (Equation 4–37), the modified Cross law (Equation 4–38), and the Carreau-Yasuda law (Equation 4–39).

The first normal stress is given by . This quantity involves the viscoelastic variable , a quantity that can be referred to as the first normal viscosity, and a weighting coefficient .

The viscoelastic variable obeys a transport equation that involves a characteristic or relaxation time and is given by:

(4–85)

The equation is such that you recover the solution in simple shear flow. The first normal viscosity found in Equation 4–84 is described by means of functions similar to those available for the shear viscosity , where is presently replaced by . In order to facilitate the set up of a flow simulation involving the simplified viscoelastic model, identical dependences for and are considered by default. However, it is important to note that different functions can be selected for the shear and first normal viscosities.

Three algebraic models are available for the relaxation time function:

  • Constant relaxation

  • Bird-Carreau law

  • Power law

Finally, for non-isothermal flows, temperature dependence laws can be selected for the shear and first normal viscosities (see Temperature Dependence of Viscosity for more details). For instance, there is the Arrhenius law Equation 4–42), the approximate Arrhenius law (Equation 4–43), and the WLF law (Equation 4–45).

When defining a non-isothermal case, a single function is used to describe the temperature dependence of the material functions , , , and optionally of .

4.11.1.5.2. Identification of Model Parameters and Functions

The simplified viscoelastic model is mainly an empirical construction. The key ingredient is the normal stress property that is introduced for the prediction of swelling. Although it is possible to qualitatively relate the swelling and the first normal stress difference, a quantitative relationship is not obvious. Methodologies have to be identified and developed for the determination of material functions and parameters. A stepwise technique is recommended for this purpose.

Note that the simplified viscoelastic model has been developed and implemented mainly for the simulation of 3D extrusion flows, therefore including the prediction of extrudate swelling. Therefore, it is acceptable to use cylindrical extrudate swelling data for the identification of the specific model properties.

As seen above, the simplified viscoelastic model involves three material functions and a parameter: the shear viscosity , the first normal viscosity , the relaxation time , and a weighting coefficient . Typically, usual viscosity data should be used for identifying the shear viscosity function. In most situations, shear thinning is experimentally observed, and algebraic relationships such as power law, Bird-Carreau, or Cross laws will be good candidates. However, it is recommended that you consider a law that exhibits a zero-shear plateau if regions of no-deformation are expected over the flow domain.


Important:  The parameters of the shear viscosity can be fitted automatically in Fluent Materials Processing based on experimental steady shear viscosity curve(s), as for a generalized Newtonian model. The other parameters of the model cannot be fitted in Fluent Materials Processing. Note that if rheometric curves are drawn in the chart, only the Newtonian part of the model is seen.


Next, a function and material parameters should be selected for the first normal viscosity . By default, a relationship identical to the selected shear viscosity is considered, as this appears to be a reasonable choice, at least at first. Of course, this default selection can be revised subsequently. The power law, which exhibits unbounded values under zero deformation, should be avoided if large regions of no deformation are expected. Instead, functions that exhibit a plateau, such as the Bird-Carreau laws, should be preferred.

Eventually, for the relaxation time and the weighting coefficient , it is suggested to perform a fast 2D simulation of axisymmetric extrudate swelling, where the effects of the remaining degrees of freedom are examined. Typically, the weighting coefficient will control the swelling intensity versus the flow rate, while the relaxation time function will control the development of the extrudate diameter along the jet, and may have a possible influence also on the developed extrudate geometric attributes. Usually, a constant value or a Bird-Carreau law can be selected for the relaxation time; the value or zero-shear value should preferably be in agreement with the typical times involved in the flow. On the other hand, a series of calculations should be performed with various values of the weighting coefficient , where the development of extrudate versus the flow rate is examined, via an evolution scheme. A comparison with experimental data on swelling should enable the selection of an appropriate numerical value for the weighting coefficient .

4.11.1.5.3. Inputs

To specify an simplified viscoelastic model, select Simplified viscoelastic model.

 Rheometry Fluid Model Model Type

The units for the parameters and their names in the Fluent Materials Processing interface are as follows:

ParameterName in Fluent Materials Processing Mass Length Time
Viscosity Law [Pa s] 1 –1–1
First Normal Viscosity Law [Pa s] 1 –1–1
Relaxation Time Law [s] - - 1
Weighting Coefficient - - -

By default, the viscosity, the first normal viscosity, and the relaxation time functions are constant and set to 1. The weighting coefficient is also set to 1.

4.11.1.5.4. Behavior Analysis

The simplified viscoelastic model involves three material functions and a parameter: the shear viscosity , the normal viscosity , the relaxation time , and a weighting coefficient . Considering the empirical construction of the simplified viscoelastic model, it is probably more relevant to inspect the behavior from the point of view of extrusion flow, in particular from the point of extrudate swelling. As will be seen, the various ingredients may have opposite effects, and enter in competition with each other.

As is known, shear thinning decreases the extrudate swelling. However, the other ingredients will usually enhance the swelling. Consider that the normal and shear viscosities are the same function of their respective dependence variable; the weighting coefficient and the relaxation time are to be considered. Actually, selecting a normal viscosity independently with respect to the shear viscosity will mainly make the analysis more complex without significantly affecting the general conclusions. The weighting coefficient adjusts the intensity of the first normal stress difference, and will essentially enhance the extrudate swelling. This is shown in Figure 4.169: Example of axisymmetric extrusion simulation for the simplified viscoelastic fluid model through a cylindrical tube with a unit radius., which plots the curve of swelling versus the flow rate for various simplified viscoelastic fluid models. The continuous lines show the swelling for various values of the weighting coefficient. Note that the amount of swelling can significantly be affected. In Figure 4.170: Example of axisymmetric extrusion simulation for the simplified viscoelastic fluid model through a cylindrical tube with a unit radius., note the development of the free surface versus the axial distance, for various values of . Note that the development versus the position is not significantly affected be a specific choice of . The relaxation time appears in the transport equation for the viscoelastic variable; consequently it will at first affect the development of the swelling along the flow direction. This is visible in Figure 4.170: Example of axisymmetric extrusion simulation for the simplified viscoelastic fluid model through a cylindrical tube with a unit radius., where the dashed lines indicate the development of swelling versus the axial distance: the development distance increases with the relaxation time, while actually the amount of swelling is less affected by the relaxation time. As can be seen in Figure 4.169: Example of axisymmetric extrusion simulation for the simplified viscoelastic fluid model through a cylindrical tube with a unit radius., the overall swelling is less affected by the value of the relaxation time.

Figure 4.169: Example of axisymmetric extrusion simulation for the simplified viscoelastic fluid model through a cylindrical tube with a unit radius.

Example of axisymmetric extrusion simulation for the simplified viscoelastic fluid model through a cylindrical tube with a unit radius.

Swelling of an extrudate versus the flow rate, for various values of (continuous lines) at =0.5 and for various values of (dashed lines) at . Note that these curves are not obtained from Fluent Materials Processing's Rheometry tool but rather from Fluent Materials Processing calculations.

Figure 4.170: Example of axisymmetric extrusion simulation for the simplified viscoelastic fluid model through a cylindrical tube with a unit radius.

Example of axisymmetric extrusion simulation for the simplified viscoelastic fluid model through a cylindrical tube with a unit radius.

Development of an axisymmetric extrudate versus the axial distance at a flow rate of 10 (see also Figure 4.169: Example of axisymmetric extrusion simulation for the simplified viscoelastic fluid model through a cylindrical tube with a unit radius.), for various values of (continuous lines) at =0.5 and for various values of (dashed lines) at =1.2. Note that these curves are not obtained from Fluent Materials Processing's Rheometry tool but rather from Fluent Materials Processing calculations.

4.11.2. Rheological Properties

Depending on the kinematics of the system you are modeling (steady shear flow, steady extensional flow, oscillatory shear flow, and so on), it is possible to compute and plot several viscometric properties in Fluent Materials Processing as they appear under:

 Rheometry Rheometric Curves Steady and Oscillatory Curves

and

 Rheometry Rheometric Curves Transient Curves

which enables you to mimic several rheometry experiments:

The following sections describe various kinematics properties, along with the name of each property as it appears in Fluent Materials Processing.

4.11.2.1. Steady Simple Shear Flow

Steady simple shear flow is characterized by a horizontal velocity field, illustrated in Figure 4.171: Steady Simple Shear Flow and defined as follows:

(4–86)

where , , and are the velocity components in the , , and directions, respectively, and is the constant shear rate, which is equal to .

Figure 4.171: Steady Simple Shear Flow

Steady Simple Shear Flow

On the basis of this flow field, the following properties can be computed:

  • steady shear stress:

    (4–87)

     Rheometry Rheometric Curves Steady and Oscillatory Curves

    Click Shear Stress.

  • steady shear viscosity:

    (4–88)

     Rheometry Rheometric Curves Steady and Oscillatory Curves

    Click Shear Viscosity.

  • first normal-stress difference:

    (4–89)

     Rheometry Rheometric Curves Steady and Oscillatory Curves

    Click First Normal Stress Difference.

  • second normal-stress difference:

    (4–90)

     Rheometry Rheometric Curves Steady and Oscillatory Curves

    Click Second Normal Stress Difference.

  • first normal-stress coefficient:

    (4–91)

     Rheometry Rheometric Curves Steady and Oscillatory Curves

    Click First Normal Stress Coefficient.

  • second normal-stress coefficient:

    (4–92)

     Rheometry Rheometric Curves Steady and Oscillatory Curves

    Click Second Normal Stress Coefficient.

  • recoverable stress:

    (4–93)

     Rheometry Rheometric Curves Steady and Oscillatory Curves

    Click Stress Ratio Sr.

  • estimated relaxation time:

    (4–94)

     Rheometry Rheometric Curves Steady and Oscillatory Curves

    Click Lambda = Sr / Shear-rate.


Note:  Properties given by Equation 4–87 to Equation 4–94 have nonzero values only for viscoelastic fluids. For this reason, these properties are not available in Fluent Materials Processing for generalized Newtonian fluids.


To compute each of these curves, you will need to specify a minimum and maximum shear rate ( and ), and the number and distribution of sampling points. By default, 100 points are selected between shear rates of 0.001 and 1000, with a logarithmic distribution. When thermal dependence is considered, you need to select the temperature. Note that several temperatures can be selected. These curves parameters can be specified under:

 Rheometry Rheometric Curves Curves parameters

See Defining Numerical Parameters for information about specifying numerical parameters for viscometric property curves. See Specifying the Curves to be Calculated for information about specifying which curves you want to compute and plot. (Note that, if you use the automatic fitting method, Fluent Materials Processing will automatically compute and plot the curves for all properties for which experimental data curves have been defined.)

4.11.2.2. Steady Extensional Flow

Steady extensional flow can be uniaxial, biaxial, or planar. Uniaxial extensional flow is illustrated in Figure 4.172: Uniaxial Extensional Flow and defined as follows:

(4–95)

where is a constant elongational strain rate.

The corresponding stress distribution can be written as

(4–96)

(4–97)

where is the uniaxial extensional viscosity.

Figure 4.172: Uniaxial Extensional Flow

Uniaxial Extensional Flow

Biaxial extensional flow is illustrated in Figure 4.173: Biaxial Extensional Flow and defined as follows:

(4–98)

where is a constant elongational strain rate.

Figure 4.173: Biaxial Extensional Flow

Biaxial Extensional Flow

The corresponding stress distribution can be written as

(4–99)

(4–100)

where is the biaxial extensional viscosity.

Planar extensional flow is illustrated in Figure 4.174: Planar Extensional Flow and defined as follows:

(4–101)

where is a constant elongational strain rate.

Figure 4.174: Planar Extensional Flow

Planar Extensional Flow

The corresponding stress distribution can be written as

(4–102)

where is the planar extensional viscosity.

For extensional flow fields, the uniaxial, biaxial, and planar extensional viscosity curves (, , and ) can be computed. Select Uniaxial Extensional Viscosity, Biaxial Extensional Viscosity and/or Planar Extensional Viscosity, if you want one or more of these curves to be calculated. These options can be found under:

 Rheometry Rheometric Curves Steady and Oscillatory Curves

To compute each of these curves, you will need to specify a minimum and maximum extensional strain rate ( and ), and the number and distribution of sampling points.

By default, 100 points are selected between strain rates of 0.001 and 1000, with a logarithmic distribution. When thermal dependence is considered, you need to select the temperature. Note that several temperatures can be selected. These curves parameters can be specified under:

 Rheometry Rheometric Curves Curves parameters

See Defining and Plotting Curves for information about specifying numerical parameters for viscometric property curves. See Specifying the Curves to be Calculated for information about specifying which curves you want to compute and plot. (Note that, if you use the automatic fitting method, Fluent Materials Processing will automatically compute and plot the curves for all properties for which experimental data curves have been defined.)

4.11.2.3. Oscillatory Shear Flow

It is often interesting to examine the response of a viscoelastic material to a small-amplitude oscillatory shear rate. This flow allows you to investigate the linear viscoelastic behavior of the material, and yields the storage and loss moduli, and . Since small amplitude deformations are considered, storage and loss moduli for differential and integral models are evaluated as:

(4–103)

For oscillatory shear flow fields, the property curves for the storage and loss moduli ( and ) can be computed. Select Storage Modulus G1 and Loss Modulus G2 if you want Fluent Materials Processing to compute these curves. These options can be found under:

 Rheometry Rheometric Curves Steady and Oscillatory Curves

To compute each of these curves, you will need to specify a minimum and maximum angular frequency ( and ), and the number and distribution of sampling points.

By default, 100 points are selected between angular frequencies of 0.001 and 1000, with a logarithmic distribution. When thermal dependence is considered, you need to select the temperature. Note that several temperatures can be selected. These curves parameters can be specified under:

 Rheometry Rheometric Curves Curves parameters

See Defining and Plotting Curves for information about specifying numerical parameters for viscometric property curves. See Specifying the Curves to be Calculated for information about specifying which curves you want to compute and plot. (Note that, if you use the automatic fitting method, Fluent Materials Processing will automatically compute and plot the curves for all properties for which experimental data curves have been defined.)

4.11.2.4. Transient Shear Flow

You can calculate the response of a viscoelastic material to one or more instantaneous variations of shear rate. Two types of transient shear flows can be considered: one can perform an experiment with multiple constant shear rate per interval, or one can simultaneously perform multiple experiments at constant shear rate. The first type allows you to define several types of transient shear flows:

  • start-up (two time intervals):

    (4–104)

  • stop (two time intervals):

    (4–105)

  • start-up and stop (three time intervals):

    (4–106)

  • double-step (five time intervals):

    (4–107)

    The double-step flow is used to assess the irreversible character of the viscoelastic material.

Figure 4.175: Transient Shear Flows

Transient Shear Flows

The second type of transient shear flow allows you to define one or multiple transient shear experiments, each being performed with constant shear rate.

For transient shear flow fields, the transient property curves for the properties defined by Equation 4–87Equation 4–94 can be computed.

Select Transient Shear Rate, Transient Shear Stress, Transient Shear Viscosity, Transient 1st Normal Stress Difference, Transient 2nd Normal Stress Difference, Transient 1st Normal Stress Coefficient, and/or Transient 2nd Normal Stress Coefficient if you want Fluent Materials Processing to compute one (or more) of these curves.

 Rheometry Rheometric Curves Transient Curves

To compute each of these curves for experiments with multiple constant shear rate per interval, you will need to define the desired number of time intervals during which a constant shear rate is applied. The time interval is bounded by the time values and . All times must be included between the specified minimum and maximum times, ( and ). To compute these curves for multiple simultaneous transient shear experiments with constant shear rate, you only need to specify the number of experiments and their corresponding shear rate. The number of sampling points per time interval must also be specified.

 Rheometry Rheometric Curves Curves parameters

See Defining and Plotting Curves for information about specifying numerical parameters for viscometric property curves. See Specifying the Curves to be Calculated for information about specifying which curves you want to compute and plot. (Note that, if you use the automatic fitting method, Fluent Materials Processing will automatically compute and plot the curves for all properties for which experimental data curves have been defined.)

4.11.2.5. Transient Extensional Flow

You may also want to calculate the response of a viscoelastic material to one or more instantaneous variations of strain rate. Three types of transient extensional flows can be considered: one can perform an extensional experiment with a constant stretch velocity, one can perform an extensional experiment with multiple constant strain rate per time interval, or one can simultaneously perform multiple experiments at constant strain rate.

The first type corresponds to a startup elongation experiment where a sample of initial length L0 is stretched at a constant velocity . The only required input is the ratio /.

The second and third types of transient extensional flows correspond to the types of transient shear flows described in Transient Shear Flow and illustrated by Figure 4.175: Transient Shear Flows (with the extensional strain rate substituted for the shear rate ). The velocity field for a transient extensional flow obeys the same definition as for the corresponding steady extensional flow given in Steady Extensional Flow.

In a transient uniaxial extensional flow, the stress difference is

(4–108)

In a biaxial or planar extensional flow, the stress difference is

(4–109)

For transient extensional flow fields, the transient property curves for the properties defined in Steady Extensional Flow can be computed. Select Transient Extensional Rate, Uniaxial Extensional Stress vs. Strain, Uniaxial Extensional Stress vs. Time, Uniaxial Extensional Viscosity vs. Time, Biaxial Extensional Stress vs. Strain, Biaxial Extensional Stress vs. Time, Biaxial Extensional Viscosity vs. Time, and/or Planar Extensional Stress vs. Strain , Planar Extensional Stress vs. Time, Planar Extensional Viscosity vs. Time if you want Fluent Materials Processing to compute these curves.

 Rheometry Rheometric Curves Transient Curves

To compute each of these curves, you will need to define the desired number of time intervals during which a constant extensional strain rate is applied. The time interval is bounded by the time values and . All times must be included between the specified minimum and maximum times, ( and ). The number of sampling points per time interval must also be specified. See Defining Numerical Parameters for information about specifying numerical parameters for viscometric property curves. See Specifying the Curves to be Calculated for information about specifying which curves you want to compute and plot. (Note that, if you use the automatic fitting method, Fluent Materials Processing will automatically compute and plot the curves for all properties for which experimental data curves have been defined.)

4.11.3. Fitting Material Parameters

This section explains how to use the automatic and non-automatic fitting methods available in Fluent Materials Processing.

4.11.3.1. Introduction

Fluent Materials Processing offers two types of fitting for material data: an automatic method and a non-automatic method. The automatic method is useful for cases where you are primarily interested in directly obtaining the material parameters for a given fluid model. The non-automatic method allows you to perform an in-depth analysis of the properties of a fluid model. If you use the non-automatic method, you can evaluate the sensitivity of the basic viscometric and elongational properties with respect to model parameters. It can also be used when experimental data are obtained on the basis of measurements which are not necessarily accessible via the automatic method (such as transient shear or elongation involving multiple interval with constant shear or strain, respectively). On occasion, the non-automatic method is invoked for fine tuning of automatically fitted parameters. Fluent Materials Processing will try to best fit the parameters of a model without taking solver issues into account. Therefore, it is sometimes necessary to adjust some parameters and fine-tune the fitting to obtain a model which simultaneously mimics the experimental data while being suitable for a complex flow simulation.

Both methods are available for almost all types of models: generalized Newtonian, differential viscoelastic, integral viscoelastic, and simplified viscoelastic. However, for the simplified viscoelastic model, only the viscous part can be fitted. The rheometric curves evaluated by Fluent Materials Processing do not take the additional viscoelastic term of the model into account.

The procedures you need to follow to use the non-automatic and automatic methods are presented in Non-Automatic Fitting and Automatic Fitting, respectively. Details about the inputs for individual fluid models (for example, Bird-Carreau law) are provided in Material Data Parameters, descriptions of the rheological properties for which you can perform fitting are provided in Rheological Properties, and information about controlling the graphical display of the data curves is provided in Defining and Plotting Curves.

4.11.3.2. Reading and Writing Files

4.11.3.2.1. Files Written or Read by Fluent Materials Processing

During an Fluent Materials Processing session, you will generally need to read and write several kinds of files. Table 4.11: Files Written and Read by Fluent Materials Processing lists the files that Fluent Materials Processing can read and/or write. You can use this table to get an overview of the files you may be using, to find out which codes write a particular file, and to see where to look for more information on each file.

Table 4.11: Files Written and Read by Fluent Materials Processing

File TypeCreated ByUsed ByDefault Name or SuffixReferences

Experimental data curve

External sourceFluent Materials Processing (Rheometry)

.csv

(.crv)

Reading Experimental Data

Calculated property curve

Fluent Materials Processing (Rheometry)Fluent Materials Processing (Rheometry).csvSaving Calculated Rheometric Curves
Material data Fluent Materials Processing (Setup – possibly with export from Rheometry)Fluent Materials Processing (Setup – Rheometry via import from Setup).mprmatReading and Writing Material Data

4.11.3.2.2. Reading Experimental Data
4.11.3.2.2.1. Reading Experimental Data Curves for the Non-Automatic Fitting Method

If you use the non-automatic fitting approach in Fluent Materials Processing, you can still read files containing experimental or other data. With the non-automatic fitting approach, you may want to manually and visually search for the optimal parameters to match one or more experimental curves that cannot be fitted automatically. Such experimental data can include transient shear viscosity or transient first normal stress difference vs. time, as well as more complex studies such as stress relaxation, reverse shear flow, and so on.

In this case, the number of accessible experimental curves is greater than for the automatic fitting approach. The approach is identical to that of the automatic fitting method.

 Rheometry Fitting Experimental Curves  New...

In the Properties - Experimental Curves panel, specify a Name, Filename, Curve Type which corresponds to the experiment and Temperature [K].

The experimental curve types are divided into two categories, those that can be used for both non-automatic and automatic fitting methods, and those that can only be utilized in non-automatic fitting methods. They are separated by an explanatory sentence that explains the specific usage. When a curve can be used for non-automatic fitting method only, a warning is written in the property panel for that curve. However, all of the curves listed above can be employed with the non-automatic fitting method.

Eventually, a title is also generated in the Properties panel, on the basis of the data entered (curve type, temperature, etc.) and which is used for easily identifying the curves in the plots window.

Refer to Files Written or Read by Fluent Materials Processing for more details.

4.11.3.2.2.2. Reading Experimental Data Curves for the Automatic Fitting Method

You will need to read your experimental curve data directly into Fluent Materials Processing when using the automatic fitting procedure. The format of the curve file is provided in Reading Experimental Data Curves for the Non-Automatic Fitting Method.

 Rheometry Fitting Experimental Curves  New...

In the Properties - Experimental Curves panel, specify a Name, Filename, Curve Type and the Temperature [K].


Note:  You will need to define at least three types of curves for a viscoelastic model.

  1. Steady Shear Viscosity

  2. Storage Modulus

  3. Loss Modulus

You can also define an optional transient uniaxial extensional viscosity curve as well as steady first normal stress difference. You can specify the curves in any order and the type of the curve you import is defined in the Properties panel.


If the file imported contains the viscosity vs. shear-rate curve, click Steady Shear Viscosity. If it contains the storage modulus vs. angular frequency curve, click Storage Modulus G1. If it contains the loss modulus vs. angular frequency, click Loss Modulus. If it contains the transient extensional viscosity curve, click Transient Extensional Flow. Lastly, if it contains the first normal stress difference vs. shear rate curve, click 1st Normal Stress Difference.

A title is then generated in the Properties panel, on the basis of the data entered (curve type, temperature, etc.) and which is used for easily identifying the curves in the plots window.

For a Transient Extensional Flow, you must specify some flow characteristics:

  • The Mode (Uniaxial, Biaxial, Planar).

  • The Type, for example, whether the curve is extensional viscosity vs. time or stress vs. strain (), where is the initial length of the sample and is its current length.

  • The Regime, for example, whether the sample is stretched at a constant extensional strain rate or constant extensional velocity.

  • The Constant Strain Rate or the Initial Strain Rate is required when the sample is stretched at a constant extensional strain rate or at constant extensional velocity. The initial strain rate is defined as the ratio of the constant extensional velocity to the initial sample length .

Here too a title is then generated in the Properties panel, on the basis of the data entered (curve type, temperature, etc.) and which is used for easily identifying the curves in the plots window.

4.11.3.2.3. Saving Calculated Rheometric Curves

Calculated rheometric curves are saved in a Rheometry subfolder located in the Fluent Materials Processing project folder. The filename explicitly refers to the type of property as well as the evaluation ID when Charts Display is set to addition mode.

4.11.3.2.4. Reading and Writing Material Data

The Rheometry tool primarily communicates outside of Fluent Materials Processing via  Setup Materials.

In input mode, it can import a rheological model from a material data file read or created via  Setup Materials. It can also directly use data entered via  Rheometry Fluid Model. In output mode, it can export the created, fitted or modified rheological model into  Setup Materials. To accomplish this, select:

 Rheometry  Import/Export

When importing, you can specify the Source Material and Fluid Model to Import.

When exporting, you can specify the Target Material and Exported Fluid Model.

Once the rheological model is configured, you can save the corresponding material data file on disk. For this, right-click the material source in:

 Setup Materials  Export

A Select File dialog will open allowing you to specify the Material File name and location where it will be written.

4.11.3.3. Non-Automatic Fitting

4.11.3.3.1. Steps for Non-Automatic Fitting

The steps for non-automatic fitting are as follows:

  1. Fluent Materials Processing allows you to read or define experimental data curves. See Working with Curves. See Reading Experimental Data Curves for the Non-Automatic Fitting Method for a list of available experimental curve types when using the non-automatic fitting method.

  2. Specify which curves you want to calculate (shear viscosity, shear stress, and so on). See Specifying the Curves to be Calculated for details. As mentioned, the non-automatic fitting allows considering experimental data obtained on the basis of measurements which are not necessarily accessible via the automatic method (such as transient shear or elongation involving multiple interval with constant shear or strain, respectively).

  3. Define the numerical parameters for the curve calculation. See Defining Numerical Parameters for details.

  4. Select the fluid model that matches your experimental data best, and define initial values for the associated material parameters. See Selecting the Type of Fluid Model for details.

  5. Modify the value of one material parameter at a time and draw the resulting data curve, until you find the combination that results in a curve most similar to your experimental curve. See Performing the Fitting Analysis for details.

  6. When you are satisfied with the curve fitting, export the result of the fitting to the setup. From there, it can be used for the definition of the simulation and exported into a material data file as well. See Exporting the Results of the Fitting for details.

4.11.3.3.2. Specifying the Curves to be Calculated

After you have read or defined your experimental data curve(s), you will need to specify which curves you want Fluent Materials Processing to calculate during the fitting. See Rheological Properties for details about the types of curves that are available. To select the required properties to be calculated and select each of the curves you want:

 RheometryRheometric Curves Steady And Oscillatory Curves

or

 RheometryRheometric Curves Transient Curves

4.11.3.3.3. Defining Curves Parameters

Next, you will need to set some parameters that will be used by Fluent Materials Processing when it calculates the specified curves during the fitting process. For this, select:

 RheometryRheometric CurvesCurve Parameters

In the Properties - Curves Parameters panel, you can specify the number of points used to represent a curve, whether or not to use a linear or logarithmic distribution of points, the range of shear rates for steady properties, the range of angular frequency for the oscillatory properties, the temperature(s), as well as the type of transient shear and elongational flows and the corresponding parameters, as described below.

  • Number of Points

    The rheometric curves are discretized into a set of points. Using more points results in a better representation of the curves. The default number of points is 100, which is acceptable for most cases.

  • Point Distribution For Steady Curves (this also applies to oscillatory curves) - logarithmic or linear

    These parameters indicate how the calculated points are distributed along the X axis in the equidistant way between successive x values in linear mode, or equidistant between successive log x values in the logarithmic mode.

  • Point Distribution For Transient Curves - linear or logarithmic scale

    These parameters indicate how the calculated points are distributed along the X axis in the equidistant way between successive x values in linear mode, or equidistant between successive log x values in the logarithmic mode.

  • Shear Rate Minimum and Shear Rate Maximum [s^-1] (Range of Shear Rates)

    If you want to compare rheological curves obtained for a steady shear flow, you will need to specify the minimum and maximum values of shear rate for the curves so that they will all be consistent. In the Properties - Curves Parameters panel, you can specify the values for the Shear Rate Minimum [s^-1] and Shear Rate Maximum [s^-1].

  • Extensional Strain Rate Minimum and Extensional Strain Rate Maximum [s^-1]

    If you want to compare rheological curves obtained for a steady extensional flow, you will need to specify the minimum and maximum values of extensional strain rate for the curves so that they will all be consistent. In the Properties - Curves Parameters panel, you can specify the values for the Extensional Strain Rate Minimum and Extensional Strain Rate Maximum [s^-1].

  • Time Minimum and Time Maximum [s]

    If you want to compare rheological curves obtained for a transient flow, you will need to specify the minimum and maximum values of time for the curves so that they will all be consistent. In the Properties - Curves Parameters panel, you can specify the values for the Time Minimum [s] and Time Maximum [s].

  • Angular Frequency Minimum and Angular Frequency Maximum [rad/s]

    If you want to compare rheological curves obtained for an oscillatory shear flow, you will need to specify the minimum and maximum angular frequencies for the curves so that they will all be consistent. In the Properties - Curves Parameters panel, you can specify the values for the Angular Frequency Minimum [rad/s] and Angular Frequency Maximum [rad/s].

  • Temperatures

    For a fluid model involving thermal dependence, you will need to provide at least one temperature. You can also specify that the selected rheological properties be calculated for a series of temperatures. In the Properties - Curves Parameters panel, select the Number of Temperatures (up to 5) and specify the values of the temperatures (First Temperature [K], Second Temperature [K], etc...). For each temperature, the rheological curves will be calculated for comparison with experimental data.

  • Transient Shear Flow

    Two modes of transient shear flows can be defined:

    1. A mode involving multiple intervals with a constant shear rate per interval

      • Constant Shear Rate per Interval

        You have to specify the number of intervals and the start time of each interval and the corresponding shear rate.

    2. A mode involving multiple transient shear flows with constant shear rate.

      • Multiple Experiments at Constant Shear Rate

        You have to specify the number of independent experiments and the corresponding shear rate.

    When you request a curve of the transient shear viscosity, Fluent Materials Processing actually computes the shear stress. When the shear rate is constant, the shear viscosity can be obtained by dividing the shear stress by the shear rate; when the shear rate is not constant, this calculation is not always valid. For being able to calculate the transient shear viscosity obtained in a mode involving constant shear rate per interval, such as in a transient shear start-and-stop experiment, we consider the ratio of the shear stress to the shear rate when the latter one is non vanishing, and we arbitrarily select a vanishing viscosity when the shear rate vanishes. For such an experiment, instead of the transient shear viscosity, it is probably more practical to examine the transient shear stress which is no affected by any specific treatment.

  • Transient Extensional Flow

    Three modes of transient extensional flows can be defined:

    1. A mode involving constant stretch velocity

      • Constant Stretch Velocity

        In this mode, a sample of initial length Lo is stretched at a constant stretching velocity V. You have to specify the initial stretch rate given by V/Lo.

    2. A mode involving multiple intervals with a constant strain rate per interval

      • Constant extension rate per interval

        In this mode, you specify the number of intervals and the start time of each interval and the corresponding strain rate.

    3. A mode involving multiple transient strain flows with constant strain rate

      • Multiple experiments at constant strain rate

        You have to specify the number of independent experiments and the corresponding strain rate.

    When you request a curve of the transient extensional viscosity, Fluent Materials Processing actually computes the extensional stress. When the extensional strain rate is constant, the extensional viscosity can be obtained by dividing the extensional stress by the extensional strain rate; when the extensional strain rate is not constant, this calculation is not always valid. For being able to calculate the transient extension viscosity obtained in a mode involving constant strain rate per interval, such as in a transient strain start-and-stop experiment, we consider the ratio of the extension stress to the strain rate when the latter one is non vanishing, and we arbitrarily select a vanishing viscosity when the strain rate vanishes. For such an experiment, instead of the transient extension viscosity, it is probably more practical to examine the transient extension stress which is no affected by any specific treatment

4.11.3.3.4. Selecting the Type of Fluid Model

To specify the type of fluid model, select:

 RheometryFluid Model

In the Properties - Fluid Model panel, you can choose generalized Newtonian, simplified viscoelastic, differential viscoelastic or integral viscoelastic model. Once you have selected the model type to which you can assign a name, you can select the viscosity law or the constitutive law, and specify a temperature dependence or not. You can then enter initial values for the several material parameters.

If you plan to perform fitting for the temperature dependence of the rheological properties, you will need to supply several experimental data curves at different temperatures.

4.11.3.3.5. Specifying the Chart Parameters

Some global properties of the chart can be modified. Select:  RheometryRheometric CurvesCharts Parameters. In the Properties - Charts Parameters panel, you can modify Charts Display and flip between replacement and addition mode. For example, you can specify whether newly calculated rheometric curves will replace the previous ones or will be added to the chart. The latter is set by default. You can also modify the axes attributes and specify whether a linear or a logarithmic scale is used for the display. By default, steady and oscillatory curves are displayed with logarithmic scales, while transient properties are displayed on the basis of a linear scale.

4.11.3.3.6. Performing the Fitting Analysis

Now that you have defined initial values for all the relevant material parameters for your model, you can begin the process of determining the values that best fit your experimental data. To accomplish this, follow the steps below.

  1. Fluent Materials Processing will use your initial values to compute the curves you selected and will draw them in the chart, select:

     Rheometry Rheometric Curves  Draw

    Alternatively, you can select  Rheometry Rheometric Curves and press Draw within the Properties - Rheometric Curves panel.

  2. Fill out the necessary values within the Properties - Experimental Curve panel.

     Rheometry Fitting Experimental Curves  New...

    Fill out the necessary values within the Properties - Fitting panel. Repeat the procedure for reading additional experimental curves. See Reading Experimental Data Curves for the Non-Automatic Fitting Method for a list of available experimental curve types when using the non-automatic fitting method.

  3. Return to  Rheometry Fluid Model and change the values of the material parameters in the Properties - Fluid Model panel. See Material Data Parameters for more information on fluid models.


    Note:  It is often easier to vary the material parameters one at a time, so that you can analyze the effects of each of them before trying to actually fit the model.


  4. Update the plot in the chart to show the new curve (as well as the old one and the experimental curve).

     Rheometry Rheometric Curves  Draw

    To compare with previously calculated curves, you should make sure that Charts Display is set to addition mode.

  5. Repeat the previous two steps until the computed curve(s) are close enough to the experimental curves.

  6. If you are varying one parameter at a time, return to  Rheometry Fluid Model and change the value of one of the other material parameters. If not, skip to the end of this procedure.

  7. Update the plot with the newly computed curve.

     Rheometry Rheometric Curves  Draw

  8. Continue to change the value of the second parameter and update the plot until you find the best value for this parameter as well.

  9. Repeat the previous three steps until you have found the best values for all parameters.

4.11.3.3.7. Exporting the Results of the Fitting

Once the fitting is complete, you can export the values for all the relevant rheological parameters into the setup.

 Rheometry  Export

The most recent values of the parameters will be exported into the setup. From there it is then also possible to export the complete material data into a file.

4.11.3.4. Automatic Fitting

4.11.3.4.1. Steps for Automatic Fitting

The steps for automatic fitting are as follows:

  1. Select the fluid model that matches your experimental data best.

  2. Fix values for any of the associated material parameters that you want to remain constant during the fitting calculation. See Selecting the Type of Fluid Model and Fixing Values for Selected Material Parameters for details.

  3. Read the experimental data curve(s) into Fluent Materials Processing. See Reading Experimental Curves for information about reading a data curve. The data must be provided in a .csv or a .crv file format

  4. Define the numerical parameters for the curves calculation. See Defining Numerical Parameters for details.

  5. Draw the experimental curve(s). See Drawing the Experimental Curves for details.

  6. Run the automatic fitting, as described in Performing the Automatic Fitting Analysis.

  7. Export the results of the automatic fitting and subsequently saving it to disk to be used in a calculation.

4.11.3.4.2. Selecting the Type of Fluid Model

To specify the type of fluid model, select:

 RheometryFluid Model

In the Properties - Fluid Model panel, you can choose generalized Newtonian, simplified viscoelastic, differential viscoelastic or integral viscoelastic model. Once you have selected the model type to which you may assign a name, you can select the viscosity law or the constitutive law, and you may specify whether you consider a temperature dependence. Next you enter initial values for the several material parameters.

4.11.3.4.3. Fixing Values for Selected Material Parameters

In some cases, you may want to fix the values of some of the material parameters so that they do not vary during the fitting calculation.

Once you have specified the type of fluid model you want to fit (as described in Selecting the Type of Fluid Model), you can simply fix the value of any material parameters that you want to keep constant during the fitting calculation. For example, the value of a parameter may be fixed due to physical requirements. The reference temperature for the Arrhenius law is a parameter that is commonly fixed.

In the Properties - Fluid Model panel, locate the parameter you want to fix, enter the required value, and click the Fitting button next to the parameter. A dialog box opens where you can modify the attribute of the parameter, as shown in the figure below.

  1. Once done, click Apply. A similar procedure can be used to initialize a parameter to a specific value.

  2. Repeat the previous three steps to define fixed values for any other parameters that you do not want changed during the fitting calculation.

  3. For verification purposes, a summary of the parameters being fixed or (user-) initialized will be displayed in the Properties - Fitting panel. A typical example is shown below.

4.11.3.4.4. Reading Experimental Curves

Within the context of the automatic fitting procedure in Fluent Materials Processing, you will need to read your experimental curve data directly into Fluent Materials Processing. Both .crv and .csv formats can be read.

To load an experimental curve, select:

 Rheometry Fitting Experimental Curves  New...

Alternatively, you can select  Rheometry Fitting Experimental Curves and press New... within the Properties - Experimental Curves panel.

In the Properties - Experimental Curves panel:

You enter a Name, which will be used for you to identify the curve in the data set.

Using Browse... under Filename, locate the experimental curve file and click Open to apply the file.

Specify the Curve Type of experimental curve. For a generalized Newtonian model, only steady shear viscosity curves can be loaded. For a viscoelastic fluid model, experimental curves can contain data on the steady shear viscosity, storage modulus G1 or loss modulus G2 vs angular frequency, transient extensional flow, or first normal stress difference N1 vs. shear rate.

Specify the corresponding temperature for thermal rheological fluid models under Temperature [K].


Note:  If you plan to perform fitting for the temperature dependence of the rheological properties, you will need to supply several experimental data curves at different temperatures.


For the transient extensional flow, additional attributes need to be specified. The mode can be uniaxial, biaxial or planar. The type of data can be the viscosity vs time or the stress vs. (Hencky) strain . The regime can be a constant extensional strain rate and the corresponding strain rate is required, or constant extensional velocity and the initial strain rate is required.

Repeat these steps if you have additional experimental curves to load.

To delete an experimental curve, right-click it in the Outline View and select Delete.

4.11.3.4.5. Defining Numerical Parameters

You will need to set some parameters that will be used by Fluent Materials Processing when it calculates the curve(s) during the fitting process. To access the Fitting Parameters, select:

 Rheometry Fitting Fitting Parameters  New...

In the Properties - Fitting Parameters panel, you can specify the range of shear rates and several other parameters, as described below. The list of Fitting Parameters depends on the selected model.

  • Maximum Number of Iterations

    In general, 50 iterations (the default) are enough to get converged results. In some cases, more iterations are needed, especially for fitting viscoelastic models when large number of modes are involved.

  • Update Graphics After Each Fitting Run

    This is activated by default and Ansys suggest to keep it unchanged. This allows you to check whether further improvements are reported after a new series of iterations.

  • Enable Fitting of Relaxation Times

    By default, relaxation times are not fitted. Rather, they are distributed at a regular interval in the range specified (logarithmic scale) for the relaxation times (see below). However, it is possible to optimize the distribution of the relaxation times. This sometimes helps to obtain a better fit.

    When activating the fitting of relaxation times, the range of relaxation times defined can be affected. The acceptable range of relaxation times is from (0.999*min_relax_time) to (1.001*max_relax_time). When this option is activated, the fitting can become more complex, requiring a few more iterations and increasing the calculation time to achieve a converged solution. By default, you do not fit the relaxation times.

  • Link Non-linear Parameters

    When a multi-mode differential viscoelastic model is selected, all modes obey the same constitutive equation. Despite this, a multi-mode model may involve a long series of linear and non-linear parameters, which need to be identified on the basis of a few data only. You can activate the option to Link Non-linear Parameters, so that parameters of a given type will be assigned the same value for all modes. This may improve the fitting procedure by increasing the sensitivity of the model properties with respect to parameters.

  • Relaxation Times Range

    By default, relaxation times are not fitted but are distributed at regular intervals in the specified range (logarithmic scale). You can enter the values respectively in the boxes next to Relaxation Time Minimum and Relaxation Time Maximum.

  • Window of Shear Rates

    If you want to compare rheological curves obtained for a steady shear flow, you will need to specify the minimum and maximum values of shear rate for the curves so that they will all be consistent. You can enter the values respectively in the boxes next to Shear Rate Minimum and Shear Rate Maximum [s^-1].

  • Window of Times

    If you want to compare rheological curves obtained for a transient flow, you will need to specify the minimum and maximum values of time for the curves so that they will all be consistent. You can enter the values respectively in the boxes next to Time Minimum and Time Maximum [s]

  • Window of Angular Frequencies

    If you want to compare rheological curves obtained for an oscillatory shear flow, you need to specify the minimum and maximum angular frequencies for the curves so that they are all consistent. You can enter the values respectively in the boxes next to Angular Frequency Minimum and Angular Frequency Maximum [rad/s]

  • Weight of the Shear Viscosity Curves

    The weighting allows you to assign more importance to one or more curves compared to others. For example, in the fitting of a viscoelastic model, if you are not interested in fitting the shear viscosity, you could set the weighting for the shear viscosity curves to a much lower value than the others.

    To set the weighting for the shear viscosity curves, enter the desired value in the box next to the item Shear Viscosity Curves. The default value is 1.

  • Weight of the Moduli G1 and G2 Curves

    The weighting allows you to assign more importance to one or more curves compared to others. To set the weighting for the storage and loss moduli curves, enter the desired value under Moduli G1 and G2 Curves. The default value is 1.

  • Weight of the Extensional Viscosity Curves

    The weighting allows you to assign more importance to one or more curves compared to others. For example, in the fitting of a viscoelastic model for a flow that is mainly extensional (fiber spinning), the fitting of the extensional viscosity must be better than the fitting of the other curves, so you should set the weighting for the extensional viscosity curves to a higher value than the others.

    To set the weighting for the extensional viscosity curves, enter the desired value in Extensional Viscosity Curves. The default value is 1.

  • Weight of the First Normal Stress Difference N1 Curves

    The weighting allows you to assign more importance to one or more curves compared to others. To set weighting for the first normal stress difference curves, enter the desired value in First Normal Stress Difference N1 Curves. The default value is 1..

4.11.3.4.6. Drawing the Experimental Curves

To draw the experimental data curves in the chart, select:

 Rheometry Fitting Experimental Curves  Draw

See Modifying Curve Display Attributes for information about changing the appearance of the curves.

4.11.3.4.7. Performing the Automatic Fitting Analysis

When you have completed the setup of your fitting problem, begin the calculation by selecting:

 Rheometry Fitting  Run Fitting

Fluent Materials Processing will automatically compute curves corresponding to all of the specified experimental data curves. When the automatic fitting calculation is complete, Fluent Materials Processing will save the results. You can export the rheological model to the setup at any time. See Exporting the Results of the Fitting.

Fluent Materials Processing will automatically update the chart with the computed curve(s) and the experimental curve(s). The steady shear viscosity, storage and loss moduli, and the first normal stress difference will be plotted on the same log-log graph. A second chart will display the extensional viscosity curve(s), using a linear scale for both the extensional viscosity and the time, unless otherwise specified. See Modifying Curve Display Attributes for information about changing the appearance of the curves.

If you want to plot curves other than those that were directly calculated, you can select them by clicking  Rheometry Rheometric Curves and specifying the required curves in the Properties - Rheometric Curves panel.

In an automatic fitting for a viscoelastic model, steady shear viscosity, storage modulus, and loss modulus are the mandatory curves whereas, first normal stress difference and extensional viscosity are the optional curves. For a generalized Newtonian model, the steady shear viscosity is a mandatory curve. After an automatic fitting, it is possible to pursue fitting, if necessary.

Click  Rheometry Fitting and select Pursue Fitting.

This option is available when at least one fitting is complete. Fitting is continued with the values of the parameters of the initialized model using the values obtained at the previous fitting step. The fixed parameters will not change. You do not have to change any flag associated to the parameter under  Setup Materials as Fluent Materials Processing automatically makes these changes.

4.11.3.4.7.1. Evaluating the Automatic Fitting

Calculations are stopped when the maximum number of iterations is reached, or if the following distances are adequately reduced:

  • The distance between two successive solutions.

  • The distances between the solution and the experimental points.

4.11.3.4.7.1.1. Evaluating the Distance Between Two Successive Solutions

Given the assumption that a fluid model has n parameters to fit:

(4–110)

If you compare the value of these parameters for two successive iterations i and i+1, respectively:

(4–111)

(4–112)

The distance for parameter j becomes:

(4–113)

where = 10-6.

The global distance D is calculated as:

(4–114)

4.11.3.4.7.1.2. Evaluating the Distance Between Solution and Experimental Points

In this case, assume you are comparing a single experimental curve to its calculated counterpart, at iteration i.

The experimental curve is composed of a set of P points:

(4–115)

If is the shear rate and if the shear rate range is defined as , the subset of points having in that interval is taken.

At iteration i, you can evaluate the model curve knowing the value of the parameters at that iteration:

(4–116)

For the set of abscissas , the model curve is composed of a set of P points:

(4–117)

The distance for point j can be defined as:

(4–118)

such that

(4–119)


Note:  When there are several experimental curves, the distance printed in the listing will be the weighted sum of the distances of each curve.

Subsequently to an automatic fitting, you are encouraged to check the physical and numerical relevance of the calculated parameters. In addition, you always have the possibility to perform a subsequent manual parameter fitting.


4.11.3.4.8. Exporting the Results of the Fitting

Once the fitting completed, you can export the values for all the relevant rheological parameters into the setup. This can be done by selecting  Rheometry  Export. The most recent values of the parameters will be exported into the setup. From there it is then also possible to export the complete material data into a file.

4.11.4. Defining and Plotting Curves

This chapter describes how to visualize the rheological properties of various fluid models and to fit them to experimental data, by defining or reading experimental data curves and plotting curves graphically using Fluent Materials Processing.

4.11.4.1. Overview

You can calculate property curves in Fluent Materials Processing or load .csv files previously created. For information about calculating and loading curves, see Rheological Properties and Working with Curves respectively. These curves are automatically displayed in the chart and become accessible when at least one curve is calculated. You can manipulate the display using your mouse and the various graphical user interface controls. Each curve can have different attributes (color, line). Curves that are calculated are saved in a sub-folder of the Fluent Materials Processing project folder.

When you click the Draw button under the Properties - Rheometric Curves panel,

 Rheometry Rheometric Curves

The curves that were selected will be updated and added to the chart. This allows you to compare several models or fine tune a fitting. See Non-Automatic Fitting for details.


Note:  You will not need to click Draw and Rheometry if you use  Rheometry Fitting as Fluent Materials Processing will automatically update the chart for you after it completes the fitting calculation.


4.11.4.1.1. Definitions of Terms

The following terms are used throughout this chapter:

axis

is a segment of a chart. Each axis has a range of values, and an interval (called a gap). You add a name to the axis, modify the format and precision of the printed numbers, and modify the range and the type of scaling (linear or logarithmic). See Modifying the Axis Attributes for details about modifying axis parameters.

chart

is a window where curves are displayed. The chart is automatically created upon the first inquiry of a curve display. Multiple curves can be plotted on the same chart.

curve

is a set of (x,y) pairs of values. The parameters of a curve include whether the data points are connected by a line, the color and thickness of the line, and the markers for the data points. See Working with Curves for information about selecting and modifying curves.

plots

is a tab with a window where curves are displayed. Curve names are also listed next or under the chart. Right-clicking the Plots window provides access to the properties of axes, curves and charts.

4.11.4.2. Working with Curves

4.11.4.2.1. Format of Experimental Data Curves

The first format is .csv, and has the following structure:

The first line displays the separator used, the second line is mainly intended for you, the user, and the subsequent lines contain the data. It is important that the last data line ends with <CR>.

The second format is .crv, a legacy file format with the following structure:

The first five lines do not contain any rheological data. Instead, the second line contains useful information. The actual data is read from the sixth line. Again, it is important that the last data line ends with <CR>.

At least two data points are needed to draw a curve. Fluent Materials Processing can read files that contain many data points.

4.11.4.2.2. Reading Curve Files

For information on reading curve files, see Reading Experimental Data Curves for the Automatic Fitting Method.

4.11.4.2.3. Deleting a Curve

If you want to delete an experimental data curve, select the curve in question under Rheometry FittingExperimental Curves and select Delete. If you want to delete a curve calculated by Fluent Materials Processing, unselect it from the list of rheometric curves

 Rheometry Rheometric Curves

and click Clean plots followed by Draw.

4.11.4.2.4. Saving a Curve

Curves are automatically saved in a Rheometry subfolder located under C:\Users\myaccount\matpro-yyyy-mm-dd-hhmmss\Rheometry.

4.11.4.2.5. Modifying Curve Display Attributes

You can modify how a curve is displayed in the chart by selecting it in the curve list and using the curve settings.

To change the color of a curve, right-click the Plots window and select Set curve color...:

A dialog box opens where you can select the targeted curve:

Curves are identified by a title selected by Fluent Materials Processing, on the basis of your entered data for experimental curves, and on the basis of the selected rheometric curves for calculated properties. These types of curves are easily identified from each other via the letter 'x' or 'r' appearing at the beginning of the title, a letter respectively standing for experimental and rheometry.

When selecting a curve, a new dialog box opens where you can specify the color by selecting in the Basic colors section. You can choose to make a more general selection via the palette, the HSV/RGB values or via its HTML code.

Click OK. The chart is automatically updated.

To change the line of a curve, right-click the Plots window and select Set curve type...:

A dialog box opens where you can select the targeted curve:

Upon curve selection, a new dialog box opens where you can select the type of line:

Click OK. The chart is automatically updated.

4.11.4.3. Cleaning and Modifying Chart

Information about cleaning and modifying charts is presented in the following subsections.

4.11.4.3.1. Cleaning a Chart

A chart may contain many curves, resulting, for example, from the successive evaluation of model properties. To clean a chart:

Select

 Rheometry Rheometric Curves

in the Outline View and click Clean plots in the Properties - Rheometric Curves panel, or

Select

 Fitting Experimental Curves

in the Outline View and click Clean plots in the Properties - Experimental Curves panel.

Experimental curves will still be available for display, while rheometric curves corresponding to the most recent set of model parameters will be reevaluated upon request.

4.11.4.3.2. Modifying the Title and Legend

To modify the chart title, right-click the Plots window and select Set Title. Enter the title in the dialog box that opens and click OK.

Drawn curves are identified by a color and are accompanied by a legend. By default legends are displayed to the left of the chart. Legends can also be displayed below the chart. You can specify the location by right-clicking the Plots window and selecting Legend Left/Bottom.

4.11.4.3.3. Modifying the Axis Attributes

There are several attributes of the axes that you can modify such as the range, the precision of the numbers attached to axis markers and the display of grids. All of these attributes are independently accessible for both vertical and horizontal axes (respectively Y- and X-axis) by right-clicking the Plots window.

To modify the range of values along the x-axis, right-click the Plots window, and select X AxisCustom range.

A dialog box will open allowing you to specify the lower and upper bound of the X-axis, along with the syntax min, max:

The chart is automatically updated.

Right-click the Plots window and select Auto range to restore the calculated min and max values.

For changing the number of gridlines along the x-axis, right-click the Plots window and select X AxisGridlines....

A dialog box opens where you can specify the number of gridlines:

The chart is automatically updated.


Note:  The usual number of gridlines is 9 for a log scale.


To modify the precision of the numbers along the x-axis, right-click the Plots window and select X AxisLabel format....

A dialog box opens where you can specify the precision along with the syntax %.ne, where n indicates the number of digits. For example %.1e if one digit is required:

You can repeat the same operations for the y-axis by right-clicking the Plots window and selecting the Y-axis. Two additional options will appear allowing you to select a linear or a log scale:

4.11.5. Guidelines for Viscoelastic Models

4.11.5.1. Introduction

There are several viscoelastic models available in Fluent Materials Processing, as described in Material Data Parameters. These models involve linear and nonlinear parameters, which in turn carry viscometric and extensional properties. It can often be a difficult task to select the best constitutive equation with the most appropriate material parameters. The task can be more easily addressed if it is broken into three smaller parts such as:

  • The number of modes.

  • Which constitutive equation are being used.

  • The parameter settings.

These parts implicitly assume that everything is known about the material being modeled and that all properties are equally important, which is usually not the case. Indeed, in actual industrial practice, only some of the information is known and assumptions are therefore necessary.

The purpose of this section is to suggest useful guidelines for the selection of a constitutive model and associated parameters. Two approaches are possible such as evaluating the numerical values of parameters to match some experimental data in a given range, or trying to fit all viscometric (and possibly elongational) measured data over a broad range. These two approaches lead more or less to the selection of a rheological model for a flow and to the selection of a rheological model for a fluid, respectively.

Although you may prefer one of these two approaches, useful guidelines can be found in both. Therefore, the guidelines that follow will be presented on the basis of the flow being simulated. For example, the kinematics involved in profile extrusion is significantly different from that in blow molding or thermoforming. Swelling during extrusion results from a velocity rearrangement and normal stress difference developed in a shear flow, while blow molding involves an elongational component with a strain-hardening or strain-thinning response from the melt.

Recommendations will be given for the four most-commonly encountered types of flow:

4.11.5.2. The Weissenberg Number

Consider a flow whose typical kinematics is given by , and λ is the relaxation time of the fluid. The Weissenberg number is given by

(4–120)

This number indicates the elasticity involved in a flow.

The Weissenberg number is sometimes understood as the ratio between a normal-stress difference and a shear stress. Such an interpretation raises questions about the relevance of Weissenberg numbers as high as 20 or even 100. Indeed, this would involve a very high normal stress, and besides a few biological fluids, the majority of macromolecular fluids would not sustain such a high stress.


Important:  Note that a polymer melt is characterized by a relaxation spectrum, rather than by a single relaxation time. Hence, when evaluating , it is important to consider an adequate relaxation time λ.


4.11.5.3. Viscometric and Rheometric Measurements

In general, you should obtain rheological measurements for the melt in the following order of importance: oscillatory properties (storage and loss moduli), steady-state shear viscosity, first normal-stress difference and/or transient uniaxial elongational viscosity. Of course, it is not always possible to obtain such an extensive range of measurements, which can often be costly.

4.11.5.3.1. Oscillatory Properties

The oscillatory properties (storage and loss moduli, represented by and , respectively) should always be provided. For most polymer melts, the measurement of the oscillatory properties is easy. Storage and loss moduli are usually obtained over several decades. The most typical measurement device is the cone-plate rheometer. Measurements can be taken at one temperature if an isothermal model is being considered, or at various temperatures if a thermal model is being considered.

In general, measurements can be taken at various temperatures in order to obtain master curves of and at a suitably selected temperature. For this, the time-temperature equivalence can be used in order to properly shift the curves. By doing so, it is possible to expand the range of experimental data for the linear properties, provided that the material exhibits visible temperature dependence. See Empirical Rules and Principles for details.

4.11.5.3.2. Steady-State Shear Viscosity

The steady-state shear viscosity is also always needed. When possible, it should also be measured using capillary rheometry, and the data for shear viscosity vs. true shear rate should be obtained. If such an experiment is not feasible, you can instead extract information about shear viscosity from the empirical Cox-Merz rule [4] , assuming that the modulus of the complex viscosity can be obtained from dynamic data. Alternatively, you can extract information about shear viscosity from the empirical mirror relationship of Gleissle [6], which suggests that the steady shear viscosity curve is a mirror image of the transient shear viscosity measured at a low shear rate. This assumes that transient viscosity measurements are indeed possible. See Empirical Rules and Principles for details.

Even though the knowledge of the shear viscosity is not important for all processes, the automatic fitting procedure requires it. Extrapolating from measured quantities can provide additional data, but data should not be extrapolated over more than one decade.

4.11.5.3.3. First Normal-Stress Difference

If available, the steady data of the first normal-stress difference is also useful. Data can be acquired using a cone-plate rheometer or using a technique based on the hole pressure. It is also possible to estimate the first normal-stress difference on the basis of capillary extrudate swell, by using an empirical formula suggested by Tanner [11].

If such data cannot be acquired, you can instead extract information about the steady first normal-stress difference using the empirical Cox-Merz rule [4] (p. 201), assuming that the storage modulus can reveal a good indication about the first normal-stress difference. See Empirical Rules and Principles for details.

Even though the knowledge of the first normal-stress difference is not important for all processes, the automatic fitting procedure can benefit from it. While extrapolating from measured quantities can provide additional data, data should not be extrapolated over more than one decade.

4.11.5.3.4. Transient Uniaxial Elongational Viscosity

If available, transient uniaxial elongational viscosity data are also useful. Data should be obtained by measuring at a constant strain rate or constant stretching velocity, which can be accomplished using equipment involving the EVF technology or of the Münstedt type (for example, filament stretching).

4.11.5.4. General Strategy for Fitting

4.11.5.4.1. Weighting Measured Data

Once you have obtained the data described in Viscometric and Rheometric Measurements, you can weigh their importance within Fluent Materials Processing, as described in Defining Numerical Parameters. This is useful if, for example, there is some uncertainty about the data for a particular property, or if some data have been measured at a different temperature.

4.11.5.4.2. Assigning a Value to a Parameter

Material parameters are specified within a range of acceptable values. For example, viscosity factors must be positive. If the automatic fitting method is used, each material parameter can have three types of status: subject to fitting, initial value, or fixed value. By default, all parameters are subject to fitting, meaning that Fluent Materials Processing will compute the best value based on the experimental data. If you have a priori knowledge of the value of a parameter, you can speed up the fitting calculation by assigning a fixed value to it.

Since there may be many parameters for a model, and there may be only a limited amount of experimental data available, it can be difficult to compute the best set of parameters. In this case, it is preferable to assign a fixed value to one or more parameters, so that they will not change during the fitting calculation.

If the model involves only one relaxation time, its value can be either assigned or computed by Fluent Materials Processing. For a multi-mode model, however, you will need to specify the spectrum of relaxation times yourself, or let it computed by Fluent Materials Processing. In that last case, you have just to specify the minimum and maximum possible relaxation times.

Next, there are nonlinear parameters for the models. For the PTT model (described in Phan-Thien-Tanner Model), the parameters ε and ξ control the elongational viscosity and viscometric properties, respectively: an increasing ε reduces or even cancels the strain hardening, while ξ affects shear-thinning properties as well as the amount of second normal-stress difference. Based on your needs, knowledge, or available experimental data, you may want to fix these values. Typically, strain hardening occurs for to , and disappears for . For example, low values for ε should be specified for a LDPE, while moderate values are appropriate for a LLDPE or a HDPE. On the other hand, shear thinning occurs for nonzero values of ξ, which usually is set to about 0.2. For practical purposes, it can be given a value of 0.5 or even as high as 1.

The Giesekus model (described in Giesekus Model) involves the parameter α, which simultaneously increases shear thinning and the second normal-stress difference while it reduces the strain-hardening property. Again, based on your needs, knowledge, or available experimental data, you can fix the value of α. When viscometric properties are relevant for the flow, values of α ranging from 0.2 to 0.8 are common. If elongational properties are needed, α plays a role similar to ε in the PTT model, and very low values (10-3 to 10-2) should be considered if strain hardening is needed.

For a PTT model with a nonzero value of ξ or a Giesekus model with α>0.5, as well as for the DCPP and the Leonov models, in both single- and multi-mode models, it is important to check whether the shear stress remains a monotonically increasing function of shear rate. A non-increasing shear stress can be corrected by adding a purely Newtonian component to the stress tensor. For single mode PTT model with a zero value of ε and for single mode Giesekus model with α = 1, the viscosity of this component is at least 1/8 of the viscosity factor of the only mode. For single mode PTT model with a non-zero value of ε, for single mode Giesekus model with α less than 1, as well as for multi-mode models, the viscosity of this component can be lower.

The DCPP model (described in Differential Viscoelastic Models) involves the parameter ξ, which simultaneously increases shear thinning and the second normal-stress difference while the parameter q increases the strain-hardening property. You can fix the value of ξ and q, based on your requirements, knowledge, or available experimental data. When viscometric properties are relevant for the flow, values of ξ ranging around 0.2 are reasonable. If elongational properties are needed, and in particular if strain hardening is needed, the parameter q should be increased. It reflects the number of branches of the POM-POM macromolecule and therefore affects the behavior in elongation.

The Leonov model (described in Differential Viscoelastic Models) involves several nonlinear parameters, affecting either the viscometric behavior or the elongation properties. You assign values to some of these nonlinear parameters, based on your requirements, knowledge, or available experimental data. Parameters q and affect the transition from trapped to free configuration of macromolecular chains. When viscometric properties are relevant for the flow, it is interesting to note that enhances the shear thinning property, while increases the viscosity. Parameter has no effect on the shear viscosity, while it contributes to a decrease of the elongational viscosity. If elongational properties are needed it can be noted that increases the strain hardening, while and decrease it.

It is possible that the fitting calculation yields values for nonlinear parameters that are unusual, although within the limits of accuracy. In this case, you should set these parameters to more appropriate fixed values, and rerun the fitting calculation. This will yield another set of parameters with the expected properties.

In general, the fitting calculation will determine parameter values on the basis of the available experimental data. However, the available data do not necessarily include the operating conditions, as measurement techniques do not always allow for reaching the conditions present in the actual process. Fiber spinning is a typical example, where the melt is processed at strain rates much higher than those available for rheometric measurement. For such cases, you can extrapolate from available data.

4.11.5.4.3. Using Identical or Independent Nonlinear Parameters

When a multi-mode differential viscoelastic model is selected, all modes obey the same constitutive equation. Despite this, a multi-mode model may involve a long series of linear and nonlinear parameters, which need to be identified based on some data only. You have the option of specifying that nonlinear parameters of a given type will be assigned the same value for all modes. This may sometimes improve the fitting procedure by increasing the sensitivity of the model properties with respect to nonlinear parameters.

Using identical values for nonlinear parameters of a given type is not a requirement; by default, the nonlinear parameters are independent. It is interesting to note that the time-temperature equivalence is advocated for using identical values for nonlinear parameters of a given type, although that argument does not formally hold for nonlinear properties. See Empirical Rules and Principles for details.

4.11.5.4.4. Relaxation Time vs. Relaxation Spectrum in Extrusion, Fiber Spinning, and Film Casting

When a large amount of data is available, it can be tempting to build a rheological model involving a broad relaxation spectrum, even ranging up to 100 s or beyond. In most cases, this will be impractical and not very useful. Indeed, in a cessation of steady shear flow, measurements reveal that the relaxation mechanism occurs with a time scale on the order of , where is a typical shear rate involved in the experiment.

This observation allows for the identification of a typical time scale for the description of mechanisms occurring in steady flow processes, such as extrusion, fiber spinning, or film casting. An extrusion flow is characterized by a typical wall shear rate , while fiber spinning and film casting are characterized by a typical elongation rate . Consequently, if a single-mode constitutive equation is selected, the corresponding relaxation time should be specified as about or , respectively. For a multi-mode model, the relaxation times should be selected in the vicinity of or , respectively. This is quite important, since it enables the setup of a model that is in agreement with the typical time scales involved in the simulation. Note that the above comments also raise questions about the relevance of Weissenberg numbers as high as 10 or 100.

For most applications, the computational domain is open, with fluid entry and exit. The residence time of fluid particles in the computational domain usually remains moderate, so extremely long relaxation times are not usually effective. Fluid particles trapped in vortices usually do not affect the main flow; they are instead a consequence of it. Finally, in extrusion, the extruded material solidifies long before the effects of these long relaxation times become visible.

4.11.5.4.5. Relaxation Time vs. Relaxation Spectrum in Blow Molding and Thermoforming

Typical time scales for blow molding and thermoforming are rather short: from on the order of a tenth of a second for a milk bottle, to on the order of a few seconds for a gas tank. Hence, starting from a reasonable assumed initial rest state, stresses develop within that time interval. Although the deformation speed (and therefore the strain rate) is usually unknown, deformations remain moderate. Consequently, dynamic or linear measurements alone provide a good characterization of the melt for blow molding and thermoforming.

The selected spectrum of the rheological model may, of course, involve very short times and very long times. Actually, very short time scales (that is, those smaller than the typical process time) can be replaced by a purely Newtonian contribution. Similarly, long times (that is, much longer than the process time) probably do not have the opportunity to develop their own contribution to the stresses, and can be omitted, if necessary.

4.11.5.4.6. Relaxation Time vs. Relaxation Spectrum in Pressing

Pressing applications are very similar to simple squeeze flows. Typical time scales for pressing applications are rather short: they range from on the order of a tenth of a second up to a few seconds. With the exception of rheometric measurements, they depend more or less on the geometrical size of the melt sample. Hence, starting from an initial rest state, stresses develop within that time interval. Although the rate of deformation can be large, deformations remain moderate. Consequently, dynamic or linear measurements alone provide a good characterization of the melt for pressing.

The selected spectrum of the rheological model may, of course, involve very short times and very long times. Actually, very short time scales (that is, those smaller than the typical process time) can be replaced by a purely Newtonian contribution. Similarly, long times (that is, much longer than the process time) probably do not have the opportunity to develop their own contribution to the stresses, and can be omitted if necessary.

4.11.5.5. Guidelines for Extrusion

Transient and steady-state situations can be considered for 2D viscoelastic extrusion simulations, while it is reasonable to essentially consider steady-state flow situations in 3D. Transient extrusion processes are encountered in parison production, while steady-state conditions are met for profile extrusion. The guidelines in this section focus on steady-state cases.

4.11.5.5.1. Important Effects
4.11.5.5.1.1. 2D Extrusion

In 2D extrusion flows, swelling at the die exit is caused by both the velocity rearrangement and the relaxation of the normal stresses. The velocity profile in the channel results mainly from viscous forces, while the normal stress is a viscoelastic effect connected to the shear rate. The material may also exhibit properties such as strain thinning and strain hardening, but these effects in extrusion are negligible.

4.11.5.5.1.2. 3D Extrusion

In 3D extrusion flows, the normal-stress difference also plays a role in the swelling, but velocity rearrangements in 3D generate much more dramatic effects than in 2D. Indeed, a 3D cross-section may be such that the resulting velocity distribution is strongly non-uniform. Typically, low velocities are encountered in narrow cross-sections and tiny details, while high velocities are encountered in wide-open regions. At the die exit,the local lack of flow rate, i.e. of extruded material, is responsible for significant deformations which develop in order to obtain a uniform velocity distribution in the extrudate. Quite obviously, this leads to a further reduction of the previously narrow cross-sections.

Therefore, more so in 3D than in 2D, an appropriate flow balancing inside the die (based on stresses, velocity, pressure, and so on) may improve the flow. Finally, as in 2D, the effects of the elongational properties are negligible, compared to those resulting from velocity rearrangement and normal-stress difference, and they can therefore be neglected.

4.11.5.5.2. Recommended Experimental Data

The linear properties and nonlinear shear viscosity provide good insight into the viscoelastic character of the material being extruded. Also, if the nonlinear shear viscosity is not available, it can often be estimated using the Cox-Merz rule [4] or the Gleissle mirror relationship [6] (see Empirical Rules and Principles for details). When feasible, this set of experimental data can be usefully completed with first normal-stress difference data. For most materials, measurement of linear properties for angular frequencies ranging from 0.01 to 100 rad/s , or perhaps up to 1000 rad/s is achievable. If you consider a typical wall shear rate in the extrusion process, it is usually included within this range of measurements. Otherwise, extrapolation should be considered.

For 3D flows, it is practical to select a “computationally light" rheological model, to reduce the computational cost of the Fluent Materials Processing simulation. Hence, it is useful to identify a typical shear rate for the flow, and obtain viscometric data one decade on each side of this shear rate. Doing this implicitly reinforces the idea of a constitutive equation for a flow (rather than for a fluid). Consequently the fitted model for a given shear-rate decade will probably not be the best one for another shear-rate decade.

4.11.5.5.3. Recommended Models and Parameters

Consider the typical wall shear rate in the extrusion process. If a constant viscosity is observed around , the Maxwell or Oldroyd-B model is recommended. If shear thinning occurs around , the PTT or Giesekus model is recommended. If qualitative information on the macromolecular behavior is required, it can also be interesting to consider using the DCPP model.

For filled materials, such as rubber, the Leonov model can also be considered, but the large number of unknowns involved warns against having unrealistic modeling ambitions.

Both single- and multi-mode models are acceptable for a 2D model, but a single-mode model is strongly recommended for a 3D model.

For a single-mode model, select a relaxation time on the order of . For a two-mode model, select one relaxation time < and one > , with no more than one decade between relaxation times. For a three-mode (or more) model, select relaxation times < and > , preferably with no more than one decade between relaxation times.

For a strain-hardening material (for example, LDPE), a low value can be specified for the PTT model’s ε or the Giesekus model’s α. Values of 10-3 to 10-2 are typical. For strain-thinning or moderate strain-hardening materials (for example, LLDPE or HDPE), a higher value—typically about 0.1—can be specified. Also, for strain hardening materials, the DCPP model can be used with a large enough value of q (number of arms).

For the simulation of the flow of filled materials, the use of the Leonov model can be a good idea. The model involves several parameters, and have received reasonable default values. It is worth mentioning that the Leonov model involves the calculation of several tensors, and that the use of a multi-mode model can be computationally expensive.

Finally, for very large flow simulations, it may be relevant to consider the simplified viscoelastic model suggested in Simplified Viscoelastic Model, suited for extrusion simulation. Here, the identification of parameters is based on rheometric information, such as viscosity and swelling versus the flow rate. Typically, the first normal viscosity involved in the model equals the shear viscosity by default, while a relaxation time value or function and a weighting factor must be identified in order to reproduce the swelling behavior. That is, a 2D axisymmetric flow simulation is required for parameters identification.

In the automatic fitting procedure, it is preferable to consider the data in the range of shear rates of interest, typically one decade above and one below. If data extrapolation is necessary, it should be done over no more than one decade. Also, use appropriate weighting factors (see Weighting Measured Data) if some data are more reliable than others.

The whole shear viscosity curve for the model may differ from measurements at low shear rates, but this can generally be disregarded. Indeed, low shear rates are encountered only in a few areas of the flow, and involve usually a fraction of the total flow rate, such that the total impact on the momentum is negligible.

4.11.5.6. Guidelines for Fiber Spinning

Transient and steady-state situations can be considered for 2D viscoelastic fiber-spinning simulations, while it is reasonable to essentially consider steady-state flow situations in 3D. Fiber spinning is a continuous process, so primarily steady-state simulations are relevant. The guidelines in this section will therefore focus on steady-state cases.

4.11.5.6.1. Important Effects
4.11.5.6.1.1. 2D Fiber Spinning

Several mechanisms are involved in 2D fiber spinning. First, a take-up velocity is assigned at the end of the computational domain. This velocity leads to significant geometric changes and to the development of the free surface. A swelling may develop at the die exit, but it is usually not a critical feature; it is quickly hindered or annihilated by the take-up (pulling) velocity.

The take-up velocity plays a dominant role in the free jet. A transverse velocity gradient exists in the die, while the free jet is endowed with an axial velocity gradient. The occurrence of a significant strain rate is typical for fiber spinning. Many melts involved in fiber spinning exhibit a more-or-less pronounced strain-hardening behavior, as this property is known to enhance the stability of the process.

A moderate take-up velocity is sometimes applied in continuous extrusion processes (for example, for guiding or stabilizing the extrudate). Here, the draw ratio is close to 1, which means that the elongation rate involved is often negligible; such cases should be regarded as extrusion instead of fiber spinning.

4.11.5.6.1.2. 3D Fiber Spinning

To some extent, 3D fiber spinning combines the effects encountered in 2D fiber spinning with some of those seen in 3D extrusion. That is, the flow has a strong elongational component due to the take-up velocity, and is affected by the velocity rearrangement in the 3D geometry. This take-up velocity leads to significant geometric changes. A swelling may develop at the die exit, but it is usually not a critical feature; it is quickly hindered or annihilated by the take-up (pulling) velocity.

The kinematics of 3D fiber spinning involves a transverse velocity gradient in the die, while the fiber itself is endowed with an axial velocity gradient. Consequently, the aspect ratio of some details in a cross-section of the fiber may differ significantly from the corresponding aspect ratio found at the die exit.

4.11.5.6.2. Recommended Experimental Data

The elongational viscosity is important, perhaps even more so than the shear viscosity. This property can be measured for moderate strain rates (usually up to 10 s-1 using an elongational rheometer equipped with the EVF device), although the material is often processed at a much higher strain rate. This apparent difficulty can be overcome by considering the following heuristic argument. For a given melt, all curves of transient uniaxial elongational viscosity follow the same lower linear envelope, up to a Hencky strain of 1 or 2. It is therefore reasonable to believe that a similar behavior occurs at higher strain rates, regardless of whether the melt is strain-hardening or strain-thinning. This is a qualitative extrapolation, which results from a speculative extension of the Gleissle mirror relationship to the transient elongational viscosity.

The minimum experimental data needed for successful fitting are the linear properties and the nonlinear shear viscosity (possibly obtained from the Cox-Merz rule [4] or the Gleissle mirror relationship [6] (see Empirical Rules and Principles for details). If data for the transient elongational viscosity are available, they should also be used. When all the properties are available, you can allocate a low weighting to the shear viscosity and a high weighting to the elongational properties (as described in Defining Numerical Parameters), since the elongational component in the fiber is more important.

4.11.5.6.3. Recommended Models and Parameters

The PTT and Giesekus models recommended for extrusion are also recommended for fiber spinning. The Oldroyd-B and Maxwell models are also good choices, especially for highly strain-hardening materials. If qualitative information on the macromolecular behavior is required, it can also be interesting to consider using the DCPP model.

Both single- and multi-mode models are acceptable for a 2D model, but a single-mode model is strongly recommended for a 3D model. For a 2D model, three modes are recommended, with the relaxation times no more than one decade apart.

The flow involves a typical shear rate in the die and a typical elongation rate in the fiber itself. Two or more orders of magnitude may exist between these quantities. It is worth mentioning that the actual deformation in shear evolves linearly with the shear rate, while the actual deformation in elongation evolves exponentially with the elongation rate. Therefore, although values for the typical shear rate in the die can be larger than the values of the elongation rate in the fiber, the actual corresponding deformation will be significantly larger in the fiber. Consequently, you should focus on the elongation character more carefully.

For a single-mode model, select a relaxation time of about . For a multi-mode model, select one relaxation time < and one > , preferably with no more than one decade between relaxation times.

For a strongly strain-hardening material (for example, LDPE), use the Maxwell or Oldroyd-B model when the strain-rate is moderate. Otherwise, consider a low value of ε (typically 10-3 to 10-2) with the PTT model, or a low value of α with the Giesekus model (typically 10-3 to 10-2). For strain-thinning or moderate strain-hardening materials (for example, LLDPE or HDPE), use the PTT or Giesekus model with a higher value—typically about 0.1 or more—for ε or α. Also, for strain hardening materials, the DCPP model can be used with a large enough value of q (number of arms). Currently, the Leonov model and the “simplified viscoelastic model" are not recommended for fiber spinning simulations.

If data on elongational viscosity are available, they should be used for the fitting. If the resulting values of the fitting for the nonlinear parameters are not in agreement with the expected behavior of the melt, they can be fixed as noted in Assigning a Value to a Parameter.

In the automatic fitting procedure, it is preferable to consider the data in the range of angular frequencies and shear rates of interest, typically one decade below and one above. If data extrapolation is necessary, it should be done over no more than one decade. Also, use appropriate weighting factors (see Weighting Measured Data) if some data are more reliable than others.

The shear viscosity curve for the model may differ from measurements, but this can generally be disregarded, since elongation is the main component of the flow.

4.11.5.7. Guidelines for Film Casting

4.11.5.7.1. Important Effects

Film casting processes exhibit features that are very similar to those encountered in fiber spinning (described in Guidelines for Fiber Spinning). That is, the extension results from a take-up velocity, which is significantly higher than the velocity at the die exit. An appropriate modeling approach for film casting in Fluent Materials Processing involves the use of a flat membrane element. Such an approach focuses on the development of the extruded film only, not on the flow inside the die.

4.11.5.7.2. Recommended Experimental Data

Elongation is the component that dominates the flow, and a proper description of the elongational behavior of the material is needed, although draw ratios are usually lower than in fiber spinning. The comments about experimental data for fiber spinning are valid for film casting as well.

4.11.5.7.3. Recommended Models and Parameters

The PTT and Giesekus models recommended for fiber spinning are also recommended for film casting. The Oldroyd-B model is another acceptable choice, especially for melts characterized by a significant strain hardening behavior. The DCPP model can also be used, especially if a qualitative description of the macromolecular behavior is required. The Leonov model, however, is not available for film casting applications. In addition, the simplified viscoelastic model is not suited for the simulation of film casting.

If no experimental data are available for the elongational viscosity, the nonlinear parameters should be fixed on the basis of known melt properties. Therefore, for a strain-thinning or weakly strain-hardening melt, high values for the nonlinear parameters are selected (typically for a PTT model or for a Giesekus model). For a strain-hardening melt, lower values should be used for the nonlinear parameters. If the DCPP model is used, a relatively large number of branches ( in Equation 4–62) should be considered for strain-hardening materials such as LDPE, while a unit or low number of branches should be selected for strain-thinning or weakly strain-hardening melts such as LLDPE.

4.11.5.8. Guidelines for Blow Molding and Thermoforming

Any prediction of blow molding or thermoforming is based on a transient calculation. The melt undergoes deformations in time, and the process conditions (such as the closing speed of the mold and the inflation pressure) may also vary with time.

In 2D, an axisymmetric geometry is often used in the Fluent Materials Processing calculation, although a planar description can also be used. Such a representation enables calculation through the thickness, and therefore prediction of the possible local shear effects. This is also the case when running an actual 3D calculation. For thin 3D objects in 3D, a membrane element is used, since it allows for calculation of complex shapes at a moderate cost. This representation is suited for elongation-like deformations but is unable to report shear deformations across the thickness.

4.11.5.8.1. Important Effects

In blow molding and thermoforming, extension is the dominating component of the deformation. Extension develops in two main directions, and is accompanied by a reduction in thickness. The magnitudes of the extensions in the main directions may differ, and they depend on the geometry. For example, inflating a cylinder involves an azimuthal extension field, while inflating a sphere involves an isotropic extension field.

Since extension is the main component in this process, knowledge of the extensional response of the material is usually necessary. However, although the extension rates can be high, the overall deformation often remains moderate. The Hencky strains are typically on the order of 1 or 2, which correspond to Cauchy strains up to 7.

4.11.5.8.2. Recommended Experimental Data

Since the processes involve important strain rates and moderate extensions, knowledge of the transient linear extensional response is enough for the description of the melt rheology. That is, knowledge of the linear properties is a good starting point. If extensional data are available, they can be used (either in the automatic fitting calculation, or for checking the results of the automatic fitting). The extensional viscosity can be measured for moderate strain rates (usually up to 10 s-1 using an elongational rheometer equipped with the EVF device), although the material can be processed at a higher strain rate.

Although shear is usually not an important component of the total deformation, shear viscosity is needed in the fitting procedure. This can be obtained from the Cox-Merz rule [4] (see Empirical Rules and Principles), and can be assigned a zero weight (see Weighting Measured Data).

4.11.5.8.3. Recommended Models and Parameters

Several modeling approaches can be considered. The constant-viscosity Newtonian model is the simplest approach, and this is recommended if no rheological data are available. The use of a (shear-thinning) generalized Newtonian fluid model should not be considered for blow molding or thermoforming, since such a model will lead to unrealistically high velocities due to strain thinning.

Further modeling recommendations for 2D and 3D simulations are provided below.

4.11.5.8.3.1. 2D and 3D Blow Molding and Thermoforming

In addition to the constant-viscosity Newtonian model, differential viscoelastic models (Maxwell, Oldroyd-B, PTT, Giesekus, FENE-P, DCPP, and Leonov) are available for blow molding and thermoforming. Although multi-mode simulations are available, for reasons of computational cost in Fluent Materials Processing, single-mode transient viscoelastic calculations are suggested, at least in first instance.

The flow involves a typical time scale τ, corresponding to the inflation time. The relaxation time should be of the same order of magnitude as this time scale.

For the viscosity factor, select a value corresponding to the shear viscosity as obtained for a shear rate of 1/τ.

For a strongly strain-hardening material (for example, LDPE), you can use the Maxwell or Oldroyd-B model when the strain rate remains moderate. Alternatively, you can use a low value of ε (typically 10–3 to 10–2) with the PTT model, a low value of α with the Giesekus model (typically 10–3 to 10–2), or a high value of q for the DCPP model. For strain-thinning or moderate strain-hardening materials (for example, LLDPE or HDPE), use the PTT or Giesekus model with a higher value—typically about 0.1 or more respectively for ε or α or the DCPP model with a low value of q.

If data on elongational viscosity are available, they should be used for the fitting. If the resulting values of the fitting for the nonlinear parameters are not in agreement with the expected behavior of the melt, they can be fixed as noted in Assigning a Value to a Parameter.

In the automatic fitting procedure, it is preferable to consider the data in the range of angular frequencies and deformation rates of interest, typically one decade above and one below the value 1/τ. If data extrapolation is necessary, it should be done over no more than one decade. Also, use appropriate weighting factors (see Weighting Measured Data) if some data are more reliable than others.

A purely Newtonian contribution can be added to the model. This corresponds to that part of the spectrum associated with very short times, and the response of which is shorter than the process time τ itself.

The shear viscosity curve for the model may differ from measurements, but this can generally be disregarded, since elongation is the main component of the flow.

4.11.5.8.3.2. Blow Molding and Thermoforming with Shell Models

For computational reasons, the use of the shell element (membrane element) is recommended for blow molding or thermoforming simulations that involve objects that are thin and have geometrically complex shapes. In addition to the constant-viscosity Newtonian model, the integral viscoelastic KBKZ model with a relaxation spectrum is also a good choice since a Lagrangian representation is used. No damping is considered due to the moderate deformations involved in the process, and the model reduces to the Lodge-Maxwell equation. The Lodge-Maxwell model predicts strain hardening up to a level that is experimentally observed.

Based on the oscillatory properties, a spectrum of 4 to 8 relaxation times is recommended. If you use a low number of relaxation times, they should be selected around the typical process (inflation) time τ. If you use a high number of relaxation times, they can spread to values as short as 10–4 and as long as 104 s.

Oscillatory data should be considered in the range of angular frequencies as covered by the range of reciprocal relaxation times. Viscosity factors should be identified based on linear oscillatory properties and shear viscosity as needed by the fitting tool in Fluent Materials Processing. A very low weighting can be used for the shear viscosity. If available, data on alongational viscosity should be used.

A purely Newtonian contribution can be added to the model. This corresponds to that part of the spectrum associated with very short times, and the response of which is shorter than the process time τ itself.

The shear viscosity curve for the model may differ from measurements, but this can generally be disregarded, since elongation is the main component of the flow.

4.11.5.9. Guidelines for Pressing

Any prediction of pressing is based on a transient calculation. The melt undergoes deformations in time in the same manner as a simple squeeze flow. The process conditions (such as mold or plug motion) may also vary with time.

In terms of geometry, 2D planar, axisymmetric, and 3D modeling can be employed in the Fluent Materials Processing calculation. Such a representation enables a calculation through the thickness, and therefore it is possible to predict the local shear effects, as well as the elongation effects. For the former, 2D modeling can certainly make sense from a computational point of view.

4.11.5.9.1. Important Effects

In pressing, shear is the dominating component of the deformation, but it is often accompanied by a significant level of extension. The balance between shear and extension depends on the geometry, as well as on the boundary conditions. For example, shear will often dominate under the adhesion condition, while elongation may play a significant role under full slipping.

Since shear is often the main deformation component that develops in this process, knowledge of the shear viscosity of the material is necessary. While shear rates can be high, the overall deformation usually remains moderate. A Cauchy strain of the order of ten is a reasonable order of magnitude.

4.11.5.9.2. Recommended Experimental Data

Since the processes are so affected by shear rates, obtaining information about the shear viscosity is the first step. Data acquisition via a direct capillary measurement is a good idea, although other techniques are acceptable. Linear properties and elongation data can also be used (either in the automatic fitting calculation or for checking the results of the automatic fitting), if they are available.

The extensional viscosity can be measured for moderate strain rates, usually up to 10 s-1 using an elongational rheometer equipped with the Extensional Viscosity Fixture (EVF).

4.11.5.9.3. Recommended Models and Parameters

As a first comment, it is worth noting that kinematic constraints imposed in the process do not really enable a full development of viscoelastic effects.

Several modeling approaches can be considered. The constant-viscosity Newtonian and generalized Newtonian models are the simplest approaches, and these are recommended when only the viscosity data is available. The generalized Newtonian fluid model is an obvious choice, because shear is an important component of the flow.

In addition to the Newtonian and generalized Newtonian models, differential viscoelastic models (Maxwell, Oldroyd-B, PTT, Giesekus, FENE-P, DCPP, and Leonov) are all good candidates for pressing simulations. Although multi-mode simulations are available, single mode transient viscoelastic calculations are recommended (at least in first instance) because of the computational cost in Fluent Materials Processing.

The flow involves a typical time scale corresponding to the pressing. The relaxation time should be of the same order of magnitude as this time scale. For the viscosity factor, select a value corresponding to the shear viscosity, as obtained for a shear rate that is typical for the process.

For a strongly strain-hardening material (for example, LDPE), you can use the Maxwell or Oldroyd-B model, as long as the strain rate remains relatively low and therefore prevents the development of high stresses. Alternatively, if you use the PTT or the Giesekus model, you should consider specifying a low value for or , respectively: typically, between 10-3 to 10-2. You can also use the DCPP model with a large number of branches (). For filled materials, the Leonov model could be recommended, although it is computationally expensive.

You can use data on the elongational viscosity if it is available. If the resulting values for the nonlinear parameters are not in agreement with the expected behavior of the melt, they can be fixed as noted in Assigning a Value to a Parameter.

4.11.5.10. Empirical Rules and Principles

It is often helpful during fitting to be able to estimate property values when experimental data is unavailable. For some materials, there are rules that link properties together; such rules are often empirical and must be handled with care, as there is no theoretical proof that validates their results. Frequently, these rules are applicable to polymer melts, such as polyethylenes and polystyrenes.

For information about some empirical rules and principles, see the following sections:

4.11.5.10.1. Cox-Merz Rule

For many materials, linear data and is relatively easy to measure for various angular frequencies (rad/s). When applicable, the shear viscosity can be estimated by applying the Cox-Merz [4] empirical rule. This empirical rule states that the modulus of the complex viscosity matches the nonlinear shear viscosity, as shown in the following equation.

(4–121)

A Cox-Merz rule also exists for the first normal stress difference; it is an approximation for low shear rates and is written as follows:

(4–122)

4.11.5.10.2. Gleissle Mirror Relationships

When performing a shear startup experiment for several values of the shear rate, you measure the shear stress as a function of time; the transient shear viscosity is obtained as the ratio of the shear stress and the (constant) shear rate. In general, the measured curves exhibit an initial development of the transient viscosity that matches the linear behavior , and subsequently departs from the linear response. The departure from the linear response is delayed when a low shear rate is considered.

The Gleissle mirror relationship [6] is empirical and relates the linear transient viscosity, with the nonlinear steady shear viscosity, as follows:

(4–123)

There is a graphical mirror relationship between the transient shear viscosity and the nonlinear steady shear viscosity. This relationship has been observed for oils, as well as polymeric melts such as polyisobutylenes and polyethylenes. It also provides a useful tool for validating measurements obtained via various techniques.

A second Gleissle mirror relationship concerns the first normal stress coefficient, . It is given by:

(4–124)

Again, there is a graphical mirror relationship between the transient first normal stress coefficient and the corresponding nonlinear steady property.

When necessary, it may be possible to extend the Gleissle mirror relationship for obtaining preliminary information on the transient elongation viscosity.

(4–125)

4.11.5.10.3. First Normal Stress Difference Relationships

The Cox-Merz rule for the first normal stress difference is only valid for a limited range of shear rates. Laun [7] has proposed an empirical rule that is valid over a larger interval of shear rates, and is given by the following:

(4–126)

Equation 4–126 is found to be valid for polyethylenes. Similar relationships can be created for other families of polymers.

Using an elastic recoil mechanism, Tanner [11] proposed a simple expression that relates the extrudate diameter to the first normal stress difference, as follows:

(4–127)

where and are the diameters of the extrudate and of the die, respectively, and is the wall shear stress in the capillary die.

4.11.5.10.4. Time-Temperature Equivalence

Time-temperature equivalence (or superposition) can be applied, for example, in the measurements of linear properties and . Usually, devices allow a measurement within a given interval of angular frequencies. However, when performing measurements at different temperatures, it is possible to expand the interval of angular frequencies. If the relationship between linear properties and temperature is unknown, it is reasonable to assume that the dependence is the same for and . Therefore, the ratio , when plotted against the angular frequency, will typically have the same shape for different temperatures, though the curves will shift along the horizontal axis. That is, an experiment performed at a lower temperature corresponds to an experiment at a higher temperature within a higher interval of angular frequencies. You can then combine the data by shifting the curves and thereby expand the interval of angular frequencies, possibly beyond the technological limits of the measurement device. Simultaneously, the identification of the corresponding shift factor will characterize the dependence with respect to temperature for subsequent analysis; this is relevant, for example, for the complex viscosity. It is therefore possible to acquire information on the temperature dependence of the properties, as well as to expand the knowledge of linear properties.

To illustrate how time-temperature equivalence can be used, consider the following figure, which shows the measurement of linear properties (thick lines) and (thin lines) for angular frequencies ranging from 10–2 to 102. The data is measured at four temperatures, represented here in black, blue, magenta, and red.

Figure 4.176: G' and G" vs. Angular Frequency [rad/s]

G' and G" vs. Angular Frequency [rad/s]

The next figure shows the properties , for each temperature ranging from black to red. The blue temperature is the reference temperature that will be used to build the master curves.

Figure 4.177: G"/G' vs. Angular Frequency [rad/s]

G"/G' vs. Angular Frequency [rad/s]

The curves from the previous figure can be combined into a single curve, as shown in the following figure. The black curve is shifted to the right, while the magenta and red curves are shifted to the left; the shift factors are kept for further evaluation of the temperature dependence of the properties. Combining these four curves into the dashed line enables you to expand the interval of angular frequencies.

Figure 4.178: Combining G"/G' Curves

Combining G"/G' Curves

The shift factors obtained in the previous step can be applied to the linear properties and , in order to obtain the master curves in the figure that follows.

Figure 4.179: Master Curves for G' and G" vs. Angular Frequency [rad/s]

Master Curves for G' and G" vs. Angular Frequency [rad/s]

The applicability of the time-temperature equivalence depends on the material considered, and is affected by the actual temperature dependence of the properties. Experience and specialized literature can provide further information.