FRICTIONALLY-EXCITED THERMOELASTIC INSTABILITY (TEI)

J.R.Barber

Introduction

When two bodies slide against each other, frictional heat is generated and the resulting thermoelastic deformation alters the contact pressure distribution. This coupled thermomechanical process is susceptible to thermoelastic instability (TEI). Above a certain critical speed, a nominally uniform pressure distribution is unstable, giving way to localization of load and heat generation and hence to hot spots at the sliding interface (Barber, 1969, Kennedy & Ling, 1974, Floquet & Dubourg 1994, Bryant et al., 1995, Kao et al., 2000). The problem is particularly prevalent in energy dissipation systems such as brakes and clutches. Hot spots can cause material damage and wear and are also a source of undesirable frictional vibrations, known in the automotive disk brake community as ``hot roughness'' or ``hot judder'' (Kreitlow et al., 1985, Inoue, 1986, Zagrodzki, 1990, Anderson & Knapp, 1990, Lee & Dinwiddie, 1998). For an animated view of infrared experimental observations of the development of hot spots during drag braking of an automotive disk brake, click here. A windows-based software package for estimating the susceptibility of brake and clutch systems to TEI is available for purchase from the University of Michigan. For more information, including sample input and output and a demonstration that can be downloaded, click here.


evidence of hot spotting on a clutch disk

Figure 1 shows one of the plates of a typical multidisk wet clutch after a period of normal service. The dark areas correspond to regions in which high local temperatures have been experienced. Evidence of surface melting can be found in extreme cases. In addition, transfer of friction material components and the products of overheated transmission fluid may be involved. The pattern seen in Figure 1 represents the effect of multiple engagements of the clutch and shows several series of hot spots at different locations overlayed on each other.

Figure 1: Evidence of hot spotting on a clutch disk


clutch disk after a single engagement A better picture of the phenomenon is obtained if the clutch is examined after a single engagement, in which case the regular pattern shown in Figure 2 is obtained. The complete disk in this particular case exhibits 12 equally spaced hot spots on each side and they are arranged antisymmetrically. In other words, the hot spots on the opposite side of the disk are located in the gaps between those shown in the figure.

Figure 2: Clutch disk after a single engagement



Stability analysis

Burton et al. (1973) used a perturbation method to investigate the stability of contact between two sliding half planes. The system is linearized about the uniform pressure state and perturbations are sought which can grow exponentially with time. Their results provided useful insight into the nature of the phenomenon, but there is no inherent length scale in the problem as defined and it was found that sufficiently long wavelengths are always unstable. A length scale can be artificially introduced into the analysis by restricting attention to perturbations below a certain wavelength, estimated as being comparable with the linear dimensions of the practical system, but the resulting predictions for critical speed do not generally show good agreement with those observed experimentally (Dow & Stockwell, 1977, Banerjee & Burton, 1979).

The effect of geometry

The first solution of a TEI problem involving a geometric length scale was given by Lee & Barber (1993), who used Burton's method to analyze the stability of a layer sliding between two half planes. This geometry provided a first step towards that of the typical disk brake assembly, where a disk slides between two pads of a friction material. Using typical material properties from automotive applications, it was found that stability is governed by a deformation mode that is antisymmetric with respect to the mid-plane of the layer and that has a wavelength proportional to the layer thickness.

Despite the considerable idealizations involved in Lee's theory, it provides plausible predictions for the critical speed and the mode shape in typical brake assemblies and is therefore quite widely used in the brake and clutch industry for TEI analysis. However, there is a clear need for a method that will account for other features of the system geometry, such as the finite width of the sliding surface, the axisymmetric geometry of the disk and the `hat' section used to attach the disk to its support. One approach is to use the finite element method to solve the coupled transient thermoelastic contact problem in time (Zagrodzki 1990, Johannson 1993, Zagrodzki et al. 1999). This method is extremely flexible, in that it can accommodate non-linear or temperature-dependent constitutive behaviour, more realistic friction laws and practical loading cycles. However, it is also extremely computer-intensive and appears unlikely to be a practical design tool for three-dimensional problems in the foreseeable future.

Numerical implementation of Burton's method

A promising alternative approach is to implement Burton's perturbation method numerically, leading to an eigenvalue problem to determine the stability boundary. If the exponential growth rate of the dominant perturbation can be assumed to be real, the critical sliding speed is defined by the condition that there exists a steady-state equilibrium perturbation --- i.e. one with zero growth rate. Du et al. (1997) used the finite element method to develop the matrix defining this eigenvalue problem. Yi et al. (1999) used Du's method to explore the effect of disk geometry in an idealized disk brake in which the brake pads are assumed to be rigid and non-conducting. Their results showed that the critical speed is in many cases quite close to that predicted by the considerably simpler analysis of Lee & Barber (1993), which probably explains the success of that analysis in practical applications.

Du's method rests on the assumption that the dominant perturbation has a real growth rate. The limited range of problems that have been solved analytically suggest that this assumption is justified if one of the two sliding bodies is a thermal insulator, or if the dominant perturbation is independent of the coordinate in the sliding direction, as in `banding' instabilities in axisymmetric systems, where the frictional heating is concentrated in an axisymmetric annular band within the contact area. However, a rigorous proof of this result has never been advanced. When both materials are thermally conducting, the stability boundary is generally determined by a disturbance that migrates with respect to both bodies in or opposed to the direction of sliding (Burton et al. 1973). In a stationary frame of reference, the perturbation will then appear to oscillate in time, corresponding to a complex exponential growth rate. The migration speed is smaller relative to the better thermal conductor and this relative motion approaches zero when the other body tends to the limit of thermal insulation. Practical systems such as brakes and clutches usually involve a steel or cast iron disk sliding against a composite friction material of significantly lower conductivity (typically two orders of magnitude lower than that of steel). As a result, the dominant perturbation migrates only very slowly relative to the metal disk. However, this migration plays an important part in the process, because it reduces the thermal expansion due to a given perturbation in heat input and hence increases the critical speed.

Eigenvalue formulation for the exponential growth rate

Burton's method can be implemented numerically for systems of two thermal conductors by defining the eigenvalue problem for the exponential growth rate. This method was first suggested by Yeo & Barber (1996), who developed it in the context of the related static thermoelastic contact problem, where instability results from the pressure dependence of an interfacial contact resistance.

We first assume that the temperature, stress and displacement fields can be written as the product of a function of the spatial coordinates and an exponential function of time When these expressions are substituted into the governing equations and boundary conditions of the problem, the exponential factor cancels and we are left with a modified system of equations in which the growth rate appears as a linear parameter. A finite element discretization of this modified problem then yields a linear eigenvalue problem for the growth rate.

In order to adapt this method to the sliding contact problem, we need to choose a suitable frame of reference, relative to which at least one of the bodies will necessarily be moving. This introduces convective terms into the heat conduction equation and can present numerical problems when the convective term is large (Christie et al. 1976). The relative magnitude of convective and diffusive terms can be assessed by calculating the Peclet number Pe = Va/k, where V is the convective velocity, k is the thermal diffusivity and a is a representative length scale. Peclet numbers in tribological applications are typically very large. For example, a steel clutch disk of mean diameter 0.2 m rotating at 2000 rpm corresponds to a Peclet number of about 35,000 using the mean diameter for a and even the element Peclet number will be large compared with unity for a realistic discretization. Thus, the convective terms will tend to dominate the finite element solution.

Fortunately, difficulties with convective terms can be avoided by using Fourier reduction in the sliding direction as long as (i) no material points on either sliding body experience intermittent contact and (ii) periodic boundary conditions apply in the sliding direction. These conditions are satisfied for multidisk brakes and clutches, which have an axisymmetric geometry, but which often exhibit signs of damage attributable to TEI with a non-axisymmetric eigenmode, as shown in the figures above. In this case, orthogonality arguments show that all the eigenmodes must have Fourier form in the circumferential direction and the sinusoidal function can be factored out of the equations, leading to an eigenvalue problem in radial and axial coordinates only, for given values of wavenumber and rotational speed. A finite element description of this problem leads to a linear eigenvalue problem for the growth rate.


Results

Typical multi-disk clutch system. All 
dimensions are in mm This method has been implemented in a windows-based software package that is available for purchase from the University of Michigan. Here we apply the method to the practical clutch design shown in Figure 3. The design parameters and material properties were chosen to be consistent with the clutch whose experimentally damaged disks were shown above. The clutch has three steel stators and two composite rotors. The rotors each have a steel core and a friction material layer bonded onto each side. During operation, sliding occurs between the friction material layers and the adjacent stators. A hydraulic pressure, P is applied to the upper piston, causing the stack of disks to be compressed against the lower reaction plate. The corresponding boundary conditions on the (homogeneous) perturbation problem are therefore that the upper surface of the piston be traction-free and the lower surface of the end plate be restrained from axial motion. All exposed surfaces were assumed to be thermally insulated, since practical heat transfer coefficients are so small that they hardly affect thermoelastic instability.

Figure 3: Typical multi-disk clutch system. All dimensions are in mm


Critical speed as a function of 
wavenumber, n, for the five-disk clutch The critical speed for the five disk clutch system is shown in Figure 4 as a function of wavenumber n. Instability first occurs in the banding mode (n=0) at a rotational speed of 218 rad/s, but there is also a local minimum of 237 rad/s at n=10. This clutch is designed for initial engagement speeds of 628 rad/s, so we also calculated the exponential growth rate at this speed. The maximum value of b=39.6 (1/s) corresponds to the Fourier mode with 10 hot spots per revolution, whilst the banding mode grows only at a rate of 12.2 (1/s). Thus, the non-axisymmetric mode would be expected to be dominant in this application.

Figure 4: Critical speed as a function of wavenumber, n, for the five-disk clutch


Eigenmode

Dominant eigenmode for the temperature in the stator surface The eigenfunction for temperature in the stator surface is shown on the left for the n=10 mode. Comparison of this figure with the experimental observations of Figure 2 above shows that the perturbation analysis correctly predicts an antisymmetric mode with focal hot spots, but the dominant wavenumber is predicted to be 10 in contrast to the 12 hot spots observed experimentally. Various explanations might be advanced for this relatively small discrepancy. The initial speed for clutch engagement is well above the predicted critical value and all wavenumbers between 4 and 14 are unstable at the beginning of the engagement. However, the mode with wavenumber 10 has the highest growth rate and would be expected to dominate the transient process. A more plausible explanation is that clutch friction materials exhibit quite complex constitutive behaviour and it is difficult to select an appropriate incremental elastic modulus for the analysis. The modulus used in the finite element analysis was the incremental modulus obtained in compression tests at the mean engagement pressure, but significant stiffening may occur under service conditions. The critical speed and the dominant eigenmode are both quite sensitive to the modulus of the friction material and plausible values could have been chosen to `fit' the theoretical predictions to a wavenumber of 12. This highlights the fact that the principal difficulty remaining in obtaining reliable theoretical predictions for TEI performance lies in the accurate characterization of the properties of the complex friction materials used.

Notice that the axisymmetric mode in this example has the lowest critical speed but that a non-axisymmetric mode becomes dominant at larger speeds. The general solution of the transient thermoelastic contact problem at constant sliding speed can be written down as an eigenfunction expansion. Furthermore, the fact that some of these terms grow exponentially but that most decay suggests that a severely truncated series (a reduced order model) might give a highly efficient numerical approximation to the transient behaviour, whilst retaining adequate accuracy. This is the subject of an ongoing investigation.

Dominant eigenmode for temperature in 
the cross-sectional plane The corresponding temperature contours in the (r,z) plane predicted for the dominant eigenmode are shown on the right. Notice that the eigenmode is antisymmetric with respect to the central stator, in which the greatest temperature perturbations are recorded. The temperatures in the rotors are close to zero except in a thermal boundary layer that is too thin to be visible in the figure. The two rotors exhibit a `qualitative' symmetry in the sense that hot regions occur at the same locations on the two sides, but the maximum temperatures are lower on those surfaces nearest to the piston and the end plate. Both these predictions were confirmed by the experimental observations. The most severe damage was observed on the central steel disk and the location of hot spots on the other two stators indicated a mode symmetric with respect to the rotors. The attenuation of the disturbance near the ends of the disk stack is probably attributable to the extra rigidity provided to the terminal stators by the piston and end plate. In fact, a simpler model in which the piston and end plate were replaced by rigid non-conducting surfaces predicted a critical speed within 1% of the more exact value.

This explanation also suggests that a more exact sequence of antisymmetric and symmetric perturbations in the stators and rotors respectively would be observed in a clutch with a larger number of disks. This was confirmed by additional finite element calculations for clutches of the same form, but with odd numbers of disks between 3 and 13. The critical speed decreases towards a limit as the number of disks increases and the dominant mode approaches a state in which the perturbation is strictly antisymmetric in the steel disks and symmetric in the composite disks, except for those near to the ends of the stack. This limiting condition was also obtained independently by modelling half of one rotor and one stator, using symmetric/antisymmetric boundary conditions at the respective mid-planes. The number of hot spots in the dominant eigenmode increases slightly with the number of disks, but the solution has essentially converged on the limit for clutches with 11 disks or more.

Conclusions

The Fourier reduction method described here permits a remarkably efficient solution of the frictional thermoelastic stability problem for systems in which the geometry is axisymmetric. The power of the method is demonstrated by the multidisk clutch example, direct numerical simulation of which would represent an extremely challenging computational problem. Values are obtained for the critical sliding or rotational speed and also for the exponential growth rate of each mode when operating above the critical speed.

In conventional clutch systems with alternating steel and composite disks, the dominant unstable mode is usually antisymmetric with respect to the steel disks, symmetric with respect to the composite disks and involves an integer number of focal hot spots around the circumference. This prediction is confirmed by experimental observations of thermal damage in a 5 disk clutch.

The method is easily applied to other examples and can therefore be used to assess the effect of design modifications such as changes in geometry and material properties on the thermoelastic stability of multidisk brakes and clutches.

References

Back to J.R.Barber's homepage