Você está na página 1de 7

Applied Rheology

Richard R. Eley, Ph.D.


Senior Scientist
ICI Paints
16651 Sprague Road
Strongsville, OH 44136

The science of Rheology by definition is the study of the deformation and flow behavior of
materials. Rheological science seeks to understand the relationship between applied stress and
the resulting deformation or flow, particularly for materials showing non-simple responses.
Applied rheology endeavors to connect fundamental properties and industrial performance for
complex fluids or viscoelastic solids. Successful performance in a wide range of commercial
products and industrial processes depends on meeting specific flow requirements. Paints and
industrial coatings, molded plastics, adhesives, personal care products and cosmetics, inks,
cement, drilling muds, ceramic slips, solder pastes, foodstuffs and medicines are examples of the
range of industrial materials whose commercial viability depends on having the “right” rheology.
For each of these materials, the necessary rheological properties must be defined with due regard
to the flow conditions which prevail during processing and application. However, it is often
difficult to link fundamental properties with real-world performance. Grounding in the
principles and practices of rheology is essential to industrial scientists, engineers, chemists, and
formulators who need to design products or processes involving non-Newtonian materials.

Introduction
The discussion in this brief article is about protective and decorative coatings, but applies
as well to adhesives, sealants, inks, and a host of commercial fluids of complex rheology. Many
of these products share a common task: they must be applied to substrates and function as a thin
layer. Industrial importance of thin-film fluid flow is illustrated by a few examples:

• Deposition of films of controlled thickness and uniformity by blade coating, slot-die


coating, direct/reverse rollcoating, spray
• Uniform coating of diffusive layer on controlled-release delivery systems
• Controlled, precise application of solder pastes and adhesives on microcircuitry boards
• Liquid permeation of porous materials
• Foam stability

The application, film formation, and defect remediation of thin, fluid layers is an important issue
affecting a variety of industrial processes and products. In this presentation, we give a brief
survey of some newer aspects of coatings rheology characterization. Emphasis is on
interpretation of flow curve data in the analysis of coating flows associated with key application
and film formation processes. We will also touch on the value of viscoelastic properties to the
foregoing, viscoelastic characterization by creep analysis, and the potential for more detailed
understanding offered by computer modeling of coating flows.
Relevance of Rheology to Coatings Formulation
To a degree matched by few other materials, rheology determines success for coatings.
Though all other properties be acceptable, a coating will usually not meet with success if the
rheology is not. Experienced formulators say that more than half the cost of new product
development is consumed in “getting the rheology right”. Moreover, apparently “minor”
changes in a raw material or process can cause significant and unexpected variability in product
rheology, a problem that will obviously require urgent resolution. For all these reasons,
rheological analysis is a vital and cost-effective tool for the coatings industry.

In the times when the majority of paints and industrial coatings were solvent-borne, the array of
solvents from which to choose was large, and provided great latitude in formulating for
performance. Control of rheology, pigment dispersion stability, substrate wetting, open time,
and film formation were relatively straightforward to achieve by solvent selection and blending.
The large-scale move toward environmentally compliant coatings (waterborne, higher solids,
reduced- or zero-VOC) has in general resulted in more complex rheology, while reducing the
number of formulating options and at the same time generating a host of performance/application
problems. In fact, the achievement of solventborne-like flow and appearance in a waterborne
system has remained an elusive target. Reduced-VOC aqueous coatings have consequent higher
surface tension and stronger Marangoni effects, which make it more difficult to coat substrates of
lower quality (lower or less uniform in surface energy), or substrates having sharp edges or small
radii of curvature (holes or sharp corners). Difficult pigment wetting and stable foam are other
consequences of reduced solvent content.

Rheology and Performance


Generally, a prime success criterion is the achievement of a uniform coating layer during
the film formation and solidification process. A final film of uniform thickness implies good
flow and leveling and the avoidance of defect development in the course of film formation.
Achievement of this goal is complicated by complex rheology, complexities of substrate
geometry, geometric and rate factors of the application process, surface tension gradients, and
environmental factors.

For coatings, the process of application and film formation obviously requires not only a large
total deformation, but also a high degree of control of flow, in order to achieve success. Flow
cannot be controlled unless it can be properly measured. The objective for the applied
rheologist, therefore, is to develop methods of rheological characterization that (a) yield accurate
data for complex fluids, and (b) are relatable to the critical processes that paints must undergo.
To meet the latter objective requires characterization methods that cover a wide range of stresses
and time scales. Modern rheological instrumentation is helping to achieve that goal.

The understanding of coating performance in terms of rheology is still far from complete, for
certain reasons. Among these reasons are (a) the sheer complexity of coatings processes, which
complicates the understanding of the role of rheology in process outcomes, and (b) difficulty in
linking measured fundamental properties with real-world performance. One answer to the
dilemma is computer simulation of coating processes, which utilizes the fundamental rheological
data as the required input. A sophisticated fluid dynamics model should take account not only of
coating rheological and physicochemical properties but also of process details, irregular substrate
geometry, environmental factors, body forces (gravity and centrifugal force) and changes in
properties with evaporation and temperature.

However, while computer modeling is perhaps the ideal approach, it is not readily accessible.
Fortunately, many coating problems can be understood and solved from shear viscosity and
viscoelastic data alone, provided the experiments performed are well designed and the results
properly interpreted. The following sections will elaborate on these points, with emphasis on the
understanding and interpretation of steady-shear flow curve data.

Non-Newtonian Behavior
Commercial fluid products comprise a wide variety of materials, with a wide range of
consistencies. In general, polymer solutions and melts (above Mc), emulsions, colloidal
dispersions, and other suspensions of particulate solids at useful concentrations will be non-
Newtonian. For non-Newtonian materials, the viscosity is no longer a material constant, but a
material function— in this case, a “function” of the shear rate (or shear stress). For non-
Newtonian fluids, a viscosity measured at a single shear rate is not an adequate representation of
the rheology of the system. For this reason, and because of other more fundamental
12
shortcomings. , the rugged but simple “single-point” viscometers commonly used in the
industrial laboratory are generally not well suited for the characterization of non-Newtonian
fluids. “Research-quality” rheometers measure some rheological property, or material function,
such as the viscosity as a function of shear rate or shear stress. Normally, a curve will be
produced representing the functional dependence of the measured property. A plot of viscosity
against shear rate or shear stress (normally log-log) may be termed a “flow curve”, and may be
generated using equilibrium or non-equilibrium flow measurement methods.2

We will neglect here the subjects of rheological models, mechanisms of shear thinning and shear
thickening, and yield behavior. Detailed discussions of these important topics, as well as
definitions of rheological terms can be found in [1] and [2] and references therein.

Controlled Stress and Controlled Rate


One has a choice of working principles in modern rotational rheometers: either
“controlled strain” (or controlled strain rate) or “controlled stress”. The difference is in whether
the torque or the angular displacement is the controlled variable. In a controlled-strain
instrument the angular displacement (or the angular velocity) is the independent (controlled)
variable and the viscous drag-torque the dependent (measured) variable. “Controlled-stress”
instruments in actuality control the torque, and measure the resulting angular displacement.
Instruments of the controlled-strain type include, for example, the TA Instruments Weissenberg
Rheogoniometer, Haake CV, Rheometrics ARES, and Bohlin CVO. Instruments of the

1. Eley, R. R., Rheology and viscometry, ASTM Paint and Coatings Testing Manual, 14th ed., ASTM,
Philadelphia, 1995, Chap. 33, 333-368.
2. Eley, R. R., Principles and methods of rheology in coatings, in Encyclopedia of Analytical Chemistry:
Instrumentation and Applications, R. A. Meyers, Ed., J. Wiley & Sons, Ltd., (2000).
controlled-stress type include the TA Instruments AR-1000, Rheometrics SR5, Bohlin CSR,
Haake RS, and Physica MC rheometers.

Each of the two main instrument types has characteristic advantages and limitations. The choice
between them depends on the material under test and the intended experiments. In terms of
performance, controlled-stress (CS) instruments can measure much lower angular velocities than
can controlled-strain-rate (CR) instruments, but CS instruments tend to be more limited at the
high angular velocity and oscillation frequency end. CS is better suited to measure long
relaxation times, an advantage of its typically stable torque capability and high angular
resolution.

CR instruments impose a shear-rate sweep while measuring the drag-torque response of


materials. Structured fluids tend to be “shear-sensitive” (more precisely, strain-sensitive).
Consequently, as strain increases exponentially under a linear strain-rate sweep protocol, fluid
structure tends to collapse rapidly, with the result that relatively few data points are obtained to
provide information on structure. In contrast, CS devices use linear (and logarithmic) rates of
stress increase, a test mode which allows materials to “obey their own rules” of stress-strain
response. CS instruments are especially useful for characterizing structured fluids and granular
dispersions such as paint, printing ink, adhesives, ceramic slips, coal slurries, cement, pigment
and colorant dispersions, drilling muds, medicines, foodstuffs, personal care products and
cosmetics, solder pastes, etc. This is particularly true where materials exhibit apparent yield
behavior. CS instruments can, in principle, directly measure the stress at the onset of yield,
avoiding errors associated with extrapolation or curve-fitting methods.

Taking coating flows as a case in point, it is important to realize that the proper variable for
correlating coating rheology to real-world coating processes is the shear stress, not the shear rate.
First of all, coating flows are the outcome of the sum of forces acting on a fluid coating layer.
That is, the rheological response to those forces determines the resulting coating flow.
Therefore, coating flows are not “driven” at a characteristic shear rate, but rather the observed
shear rate is the resultant of the stress driving the process and the corresponding viscosity at that
stress. Using stress as the controlled or independent variable is, in this sense, the more “natural”
way to characterize coatings and other complex fluids.

Rheology of Coating Application and Film Formation


In graphing the viscosity flow curve, the usual practice is to plot with shear rate as the
independent variable. We prefer to use the shear stress as the abscissa for several reasons: i) the
torque (shear stress) is the independent variable for the rheometer used, ii) many industrial flows
are governed by the available shear stress, with the shear rate a rheology-dependent variable, iii)
separation between flow curves is always better as a function of shear stress (more sensitive
variable), and iv) the collapse of particle flocs or gel networks is more obvious.

There is an important point to be made in regard to the role of rheology in coating flow during
application and leveling, as well as in undesirable flows which may lead to film defects (e.g.,
sagging, cratering, crawling, edge withdrawal). It is that these flows are driven by specific shear
stresses, which can be calculated from the forces acting (e.g., gravity and surface tension) and
the geometry of the film [1,2]. However, these processes can occur over a wide range of shear
rates, depending on the coating’s viscosity at the acting shear stress. The shear stress acting on a
coating layer (for a given process) is independent of the rheology. In contrast, the shear rate will
be dependent on paint rheology. For this reason it is preferable to represent flow data as
viscosity vs. shear stress plots, as opposed to viscosity vs. shear rate, which is the more common
practice. The shear rate is a dependent variable, for real processes. The appropriate independent
variable to use to differentiate the performance of paints in terms of their rheology is the shear
stress. Not to do so is wrong in principle and will result in incorrect comparisons of paints with
respect to their relative rates of, e.g., sagging, leveling, pigment settling, or ease of application.
This issue is key to using flow curve data correctly to understand coating performance.

Plotting with shear stress as the independent variable, as in Fig. 1, allows straightforward and
correct comparison of paints A and B at the specific shear stress for a particular process. The
process stresses illustrated in the Figure are for (i) gravity-driven sagging of a 3-mil wet film, (ii)
surface tension-driven leveling of surface roughness, and (iii) application by brush or roller. The
point is that, no matter what the rheology of a paint, the shear stress acting on a coating layer for
a given process is the same (for a given geometry, density, surface tension). The shear rates for
these processes, however, are not the same, as illustrated in Fig. 2. Fig. 2 shows the same two
paints as in Fig. 1, this time with the viscosity plotted as a function of shear rate. Note that
process shear rates are shifted to the left for the higher viscosity paint.

Sagging is driven by gravitational shear stress σ g whose magnitude depends solely on the wet
film thickness h and density ρ (g is gravitational acceleration):
σ g = ρgh cos θ (1)
(cos θ =1 for a vertical substrate). Predictions of sagging from comparison of viscosities
measured at an arbitrary shear rate will be misleading because paints sag at different shear rates,
depending on their rheology. The proper way to predict relative sagging tendency is to first
select the governing viscosity from a plot of viscosity vs. shear stress, rather than shear rate. The
sagging shear stress calculated from Eq. (1) determines the viscosity controlling sagging (ηsag)
from the flow curve. Fig. 1 compares two paints in this manner, with gravitational shear stress
levels corresponding to 3 mils wet film thickness indicated. Fig. 2 shows flow curves for the
same two paints plotted as a function of shear rate, with the calculated sagging shear rates at 3
mils indicated ( γ&sag = σsag / ηsag ).

The solid vertical line in Fig. 2 approximates the shear rate of the Stormer viscometer. The
viscosities of the two paints at the Stormer shear rate are about the same (Paint B actually
slightly higher). At an arbitrary shear rate of 1 s-1, paint A is about 50% higher than B, whereas
at the actual sagging shear stress or shear rates paint A is 600% higher in viscosity. Clearly,
comparison of viscosities measured at arbitrary shear rates can lead to incorrect predictions of
relative sagging behavior, particularly when flow curves cross over. Comparison of coating
rheology as a function of shear stress both simplifies the process and is more correct in principle.

In a typical paint development laboratory, rheology is characterized by a “two-point” method,


using the Stormer viscometer and the so-called “ICI Cone and Plate Viscometer”. The key
question to ask is whether such a two-point method adequately represents paint rheology. Fig. 2
shows flow curves for two paints, one mildly shear thinning, while the other shows evidence of a
sudden structural collapse at a shear stress of around 250 dyne/cm2. Superimposed on the curves
are points indicating the corresponding measurements from the Stormer and ICI viscometers.
It is apparent that the Stormer and ICI data points are located in the “sheared-out” region of the
paint rheology curves. Also shown in Fig. 2 are vertical dashed lines indicating shear stresses
driving several important film formation processes, namely, leveling of surface waviness, and
gravitational sag at 3, 6, and 12 wet mils film thickness. Clearly, the two-point characterization
method would give no information about the leveling/sagging/pigment settling regimes at all.
The two-point method also misses the region of high structural viscosity due to flocculation.
Presence of the latter will have profound effects on sag, leveling, pigment settling, color, opacity,
etc.

1000 A

Brushing/rolling stress
B
100.0
viscosity (poise)

3-mil sagging stress

10.00
Leveling stress

1.000

0.1000
1.000 10.00 100.0 1000 10000
shear stress (dyne/cm^2)
Figure 1. Viscosity vs. shear stress plot for two paints A ( ) and B ( ). Vertical dashed lines
indicate (i) gravitational shear stress driving sagging for 3-wet-mil paint layer; (ii)
surface-tension-driven leveling stress; (iii) brushing/rolling application.
1000
Leveling
shear rates
100.0
viscosity (poise)

Application
Sagging shear rates
10.00 shear rates

1.000
Stormer B
shear rate
A
0.1000
0.01000 1.000 100.0 100000
shear rate (1/sec)
Figure 2. Viscosity vs. shear rate plot for paints A ( ) and B ( ). Shear rates for sagging, leveling, application are
different for A and B, shifted to left for higher-viscosity paint. Vertical dashed lines indicate shear rates
for each paint and process. Vertical solid line is the approximate shear rate for the Stormer
Viscometer.

Você também pode gostar