Sunteți pe pagina 1din 25

Powder Technology H.Hofmann, P.

Bowen
Part 2: Compaction Literature used for this chapter: Randall M. GERMAN, Powder Metallurgy Science, second edition, Metal Powder Industries Federation, 105 College Road East Princeton, New Jersey 08540-6692 U.S.A. ISBN 1-878954-42-3 (1994) Violaine GUERIN, EPFL, THESE N 3021 (2004) J.C Cunnigham, I.C. Sinka, A.Zavaliangos, J. of PHARMACEUTICAL SCIENCES, 93, (2004) 2022 - 2039 I.C. Sinka, J.C Cunnigham, A.Zavaliangos, J. of PHARMACEUTICAL SCIENCES, 93, (2004) 2040 - 2053 Olle Skrinjar, Per-Lennart Larsson Acta Materialia 52 (2004) 1871-1884 D. Poquillon, J. Lemaitre, V. Baco-Carles, Ph. Tailhades, J. Lacaze, Powder Technology 126 (2002) 65-74 D. Poquillon , V. Baco-Carles, Ph. Tailhades, E. Andrieu, Powder Technology 126 (2002) 75-84

Most applications for powder metallurgy dictate that high densities be attained in the final product. Powder densification can be achieved through one of three methods: i) sintering to densify a low density perform, ii) press to a high density followed by sintering, or iii) simultaneously press and sinter using a full density technique. Powders that exhibit good sintering densification can be shaped using low pressures, often with the aid of an organic binder. Injection moulding is an example of that process. Compaction relies on an external source for deforming the powders into a high density component that approaches the final geometry. The means of delivering that pressure to the powder, the mechanical constraints, and the rate of pressurization are the main process parameters which determine the resulting density. Densification by compaction processes involves particle deformation and is covered with attention to pressure transmission and friction effects. From a mechanical point of view, the compaction process can be roughly divided into three stages: The first stage is the packing process, referred to as Stage 0 compaction. The next stage is referred to as Stage I compaction, and is characterized by the surrounding of the particles by connected pores with a relative density range from around 0.60 up to approximately 0.800.90. Stage II compaction is characterized by the sealing off of the pores between the particles, and the material behaving more or less like a porous solid. In general, there are three ways of analysing the mechanical behaviour at cold and/or hot compaction: Powder Technology, Part II: Compaction 1

Phenomenological description based on experimental observation. Macroscopic constitutive equation for a porous solid, often based on phenomenological assumptions. Micromechanical approach to analyse the deformation of individual particles in detail.

The four major mechanisms controlling densification are rearrangement, plastic deformation, power law creep and diffusional flow. Almost without exception, theoretical analyses have concentrated on the last three mechanisms while neglecting rearrangement, 1. Phenomenology of Compaction To achieve greater densities as the tap density (after filling and vibration) requires an external pressure. As shown in Figure 1 the initial rate of densification with the application of pressure is high. With continued deformation the slope of the density versus pressure curve declines, reflecting particle work hardening. At the beginning of a compaction cycle, the powder has a density approximately equal to the apparent density. Voids exist between the particles, and even with vibration, the highest obtainable density is only the tap density. For a loose powder there is an excess of void space, no strength, and a low coordination number (number of touching neighbour particles). As pressure is applied, the first response is rearrangement of the particles with filling of large pores, giving a higher packing coordination. That is analogous to vibrating the powder. Rearrangement is aided by smooth, hard particle surfaces. Increasing pressure provides better packing and leads to decreasing porosity with the formation of new particle contacts (Fig. 2). The graphs are for spherical bronze particles and show that porosity decreases as the number of contacts and the contact area between particles increase. The point contacts undergo elastic deformation, and at all points in the compaction cycle a residual elastic energy is stored in the compact.

Fig. 1: A sketch of the density versus compaction pressure during metal powder compaction, showing key stages and declining compressibility as the density increases. [German] High pressures increase density by contact enlargement through plastic deformation. Thus, the pressure causes localized deformation at the contacts, giving work (strain) hardening and allowing new contacts to form as the gaps between particles collapse. The interparticle contact zones take on a flattened appearance with a circular profile. The green density p and diameter of the circular profile X are related as follows:
2 2 X = D 1 ( 0 / ) 3 (1.1) Where D is the particle diameter and 0 is the initial density corresponding to X = 0. The subsequent bond strength depends on the amount of shear at the particle contacts. The maximum shear stress occurs at the center of the contact and is highest when the contact is 1

Powder Technology, Part II: Compaction

Fig. 2: Data for the die compaction of a spherical bronze powder. The upper plot shows the porosity decrease with pressurization, the middle plot shows the average number of contacts per particle (repacking), and the lower plot gives the contact area (deformation). Compaction involves both particle repacking and deformation mechanisms. [German ] small. During deformation, cold welding at the interparticle contacts contributes to the development of strength in the compact. The strength after pressing, but before sintering, is termed the green strength. The large pores are eliminated first and the particle coordination number increases to further distribute the load. Consequently, further gains in pressed density (the green density) require even greater expenditures of energy from the external pressure source. The coordination number typically starts near six to seven in the loose state. During compaction the coordination number Nc varies with fractional porosity as follows: (1.2) N c = 14 10.4 0.32 As shown in Figure 3, at high densities the coordination number approaches 14, corresponding to a tetrakaidecahedron, an efficient space filling polyhedron with a low surface area to volume ratio. Actual coordinations at full density range from 13.4 to 14.2. A sketch of a tetrakaidecahedron1 is shown in Figure 4 and Figure 5 is a scanning electron micrograph of the fracture surface from a full density compact formed from spherical particles. With considerable work hardening or with brittle material, densification can occur by fragmentation. The compact surface area increases due to fragmentation; however, the strength shows little improvement. The hardness and work hardening behaviour both influence compaction. Likewise, a small particle size hinders compaction because of the higher interparticle friction and higher particle work hardening rate. At very high compaction pressures, in excess of 1 GPa, massive deformation occurs, leaving only small pores. Continual pressurization beyond that level is of little benefit, the material response is similar to that of a dense solid.

The tetrakaidecahedron is a polyhedron with 14 faces. It is composed of eight hexagonal faces and six square faces. The number of edges is 36 and the number of corners is 24. If the length of a straight edge segment on the on the polyhedron is L, then the volume V, surface area S and grain size G are related as follow:
V = 128L3 S=

432 + 6 L2

G = 8L

Those relations also prove useful in discussions about the grain geometry during sintering or cold compaction

Powder Technology, Part II: Compaction

Fig. 3: The coordination number (number of contacts per particle) increases with densification by compaction. [German]

Fig. 4: The tetrakaidecahedron is a model polygonal shape representative of the final particle geometry after compaction to full density.

Fig. 5: This fracture surface illustrates the polygonal particle shape associated with full densification by compaction. The initially spherical particles have deformed to a shape similar to the tetrakaidecahedron. [German] Die wall friction with the powder is a main problem in uniaxial powder compaction. The friction causes the applied pressure to decrease with depth in the powder bed. Many important intrinsic characteristics of a powder affect the pressure-density-strength relations in a compact. They include material properties such as hardness, work (strain) hardening rate,

Powder Technology, Part II: Compaction

surface friction, and chemical bonding between particles. Equally important are the extrinsic factors associated with the powder size, shape, lubrication, and the mode of compaction.

1.1 Fundamentals of Compaction Consider a cylindrical compact of diameter D and height H such as drawn in Figure 6. Analyzing a thin section of height dH when there is an external pressing force, shows that the pressure on top of the element P and that transmitted through the element bottom Pb will differ by the normal force acting against friction.

Fig 6: 5. The balance of forces during die compaction [German] On this thin section the balance of forces along the axis can be expressed as follows:

F = 0 = A( P P) + F
b

(1.3)

Where Fn is the normal force, is the coefficient of friction between the powder and the die wall, and A is the cross sectional area. The normal force can be given in terms of the applied pressure with a proportionality constant z that actually varies with the compact density. That factor represents the ratio of the radial stress, thus Fn = zPD dH (1.4) Combining 1.3 and 1.4 gives the pressure difference between the top and bottom of the powder element dP as, F 4 zP dP = P Pb = f = dH (1.5) A D And after integration: 4 zx (1.6) P( x ) = P exp D The equation is applicable to a single action pressing. It shows that pressure decreases with depth in the powder bed. Sample data for the compaction of copper are given in Figure 6, Examples of plots of Equation 1.6 are given in Figure 7. Note the effect of an increase in the term zH/D. The wall friction contributes to a decreased pressure with depth. Double acting compaction will have a simultaneous pressure profile from both the top and bottom punches. For the double acting case, Equation 1.6 is valid, but the distance x is now the distance to the nearest punch. The result is a more homogeneous pressure distribution in the compact. For a single-ended pressing, the average compaction stress is estimated as H (1.7) = P 1 2 z D

Powder Technology, Part II: Compaction

and for a double-ended pressing the average stress is approximately H = P 1 z D

(1.8)

Fig. 6: The transmitted pressure in a copper powder compact as a ratio to the compact thickness, with both parameters normalized for the applied pressure and compact geometry.[German]

Fig. 7: The pressure gradient below the top punch in single-acting die pressing as given by Equation 1.6. [German]. In die pressed powder compacts, density gradients result from the pressure gradients. For a copper powder, the density gradients with both single- and double- action pressing are shown in Figure 8 along radial cross sections. In both compacts, the height to diameter ratio is unity, the coefficient of friction is 0.3 and the pressure ratio is 0.5. In the single action pressing, the lowest density occurs at the compact bottom. Alternatively, the double action pressing has the lowest density in the heart of the compact. Considering that the compact strength is dominated by the green density, it is easy to speculate that the strength and hardness patterns will look quite similar to the density profiles shown in Fig.8.

Powder Technology, Part II: Compaction

Fig.8: Constant density lines in cylinders of compacted copper powder. The contours are in g/cm3. [German] The other important factor is the height to diameter ratio. As the ratio increases, density gradients in a compact increase and the overall compact density decreases. Fig. 9 gives the results for a single-action pressing of copper using a constant compaction pressure of 700 MPa. Plots of the approximate pressure distribution in the compacts are given for height to diameter ratios of 0.42, 0.79, and 1.66.

Fig. 9: Pressure distributions in compacted copper powder cylinders (single-action pressing) of three height to diameter ratios. Only the right half of each cylinder is shown. The applied pressure is 700 MPa. [German] 1.2 Parametric Relations Green Density as a Function of Applied Pressure The compressibility of a powder is related to the density attained at a preset compaction pressure. It provides a basis for comparing powders for their ease of compaction. As a first approximation to die compaction, the incremental densification with pressure depends on the remaining porosity as follows: d = dP (1.9) Where P is the applied pressure, is the fractional porosity, and is a proportionality constant that reflects the fundamental material behaviour. Rearranging and integrating Equation 1.9 gives (1.10) ln = P 0 Where o is the apparent porosity at the beginning of compaction. Equation 1.9 ignores the mechanism of compaction; however, it appears to be empirically applicable to metal powder

Powder Technology, Part II: Compaction

compaction in the pressure range from 100 to 700 MPa. It is necessary to account for the initial consolidation rearrangement transients and multiple mechanisms of compaction, thus Equation 1.9 is modified where B accounts for rearrangement: (1.11) ln = B P 0 Other expressions, similar to Equation 1.10 are listed in the literature (for example: German page 221, Table 6.29) The relationship represented by Equation 1.10 fits the observed compaction behaviour of most powders. Example consolidation data on three metal powders are given in Figure 10. At pressures of approximately 300 MPa, the behaviour shifts, indicating a change in the mechanism controlling compaction from yielding to work hardening. Note that copper is the lowest strength material and shows the lowest porosity at all compaction pressures. On the other hand, the stainless steel is a prealloyed powder that exhibits the highest strength and highest porosity. Both the high and low pressure regions behave as predicted by Equation 1.10. At compaction pressures below approximately 300 MPa, the material yield strength dominates, as illustrated by the data in Figure11.a. The plot shows how the slope correlates to the wrought yield strength of the material. At high compaction pressures, the compaction efficiency decreases as full density is approached. In this region, compaction is governed the work hardening behaviour of the material as illustrated by Figure 11b. The rate of work hardening limits compaction when fracture is absent.

Fig. 10: The variation in fractional porosity with applied pressure for sponge iron, spherical copper, and irregular stainless steel powders with a particle size between 106 and 125 m. The data are plotted using a log porosity scale. [German]

Powder Technology, Part II: Compaction

Fig. 11 a and b: Data for various metal powders (Al, Zn, Ni, Fe, W, Cu, and stainless steel) show how the compaction behaviour at pressures below approximately 300 MPa is governed by the material yield strength (a). Work hardening governs the high pressure compaction (b). The slopes correspond to those obtained from Equation 1.10. Green Strength Variation with Density Pores reduce the effective load bearing area in a compact; thus, the bulk strength is reduced by the lower cross-sectional area. Also, pores act as stress concentrators and are effective crack initiation sites. Consequently, a powder compact is expected to have a strength less than that of an equivalently processed wrought (100% dense) material. One of the problems in pressing is a lack of structural integrity in the green compact, manifested as cracks, laminations, large density (strength) gradients, and poor handling characteristics. Consequently the green strength variation is expected to vary with density has considerable practical application. The compaction relations depend on the various powder characteristics like particle size, shape, and interparticle friction. Lubrication and compact dimensions also influence the ease of consolidation. The green strength is expected to vary with the fractional density of the compact as follow = C 0 f ( ) (1.12) where is the strength, C is a constant, o is the wrought strength, and f() is the density dependence. For monosized spheres, the strength is directly linked to the contact zone size, which depends on the pressed density. In most cases, the strength varies with the fractional density as a power function, thus, = C 0 m (1.13) where the exponent m might be about 6.

2. Macroscopic constitutive model The results of the characterisation of the compactibility of powders using phenomenological models are usually plotted as porosity-axial stress functions. In general, the axial stress is determined by measuring the axial force, whereas the porosity is determined from measuring the tablet dimensions and mass or other methods, such as helium pyknometry. Porosity-axial stress functions can be used for the comparison of materials. For example, greater slopes of the Heckel plot (see Fig 11a) suggest a greater degree of plasticity in the material; in other words, it indicates ease of densification, which is related to the hardening characteristics of the powder. This approach is deficient because it considers only the average stress along the direction of compaction, it ignores the radial stress transmission (factor z in equation (1.4) is not defined quantitatively) and friction, and is incapable of addressing the three-dimensional character of the stress field and the density inhomogeneity that is present in shapes other than the low-profile (i.e., low thickness- to-diameter ratio) flat tablets. Over the last 15 years, a

Powder Technology, Part II: Compaction

comprehensive approach has been developed for the analysis of compaction using continuum mechanics principles. This approach is based on the following components: Equilibrium equations (balance of forces transmitted through the material) Continuity equation (conservation of mass) Geometry of the problem Constitutive behavior of the powder (stress strain behavior) Boundary conditions including loading (e.g.,displacement and velocity) and friction between the tooling and the powder Initial conditions (e.g., initial relative density of powder) Because of the significant nonlinearity in material properties and contact stresses, a typical powder compaction problem cannot be solved analytically without major simplifications and thus a numerical approach is required. Today, numerical modeling is possible because of the proliferation of powerful inexpensive computers and the existence of commercial finite element software. Therefore, one of the main task of a researcher/engineer is to develop experimental inputs for such modeling efforts, for example, the constitutive model for the material and die wall friction.
2.1 Powder as a Mechanical Continuum

Regardless of the nature of the powder, the size of the particles is usually more than two orders of magnitude smaller than the dimensions of typical pressed workpieces. A representative volume element of a sufficiently large number of particles can, therefore, be defined such that it i) represents properly the macroscopic response of the material, and ii) its mechanical response is insensitive to the statistical variation present at the particle level. In this case, the powder aggregate can be regarded as a continuum medium. Within this framework, the specifics of the interaction between particles, such as cohesion and interparticulate friction, are combined and included into density-dependent material parameters of a macroscopic constitutive model that relates the average of stresses and strains over the representative volume. Understanding continuum models requires familiarity with fundamental concepts of applied mechanics. In such models, the stresses and strains are considered at the level of the representative volume element. Inside the representative volume element, stresses vary within individual particles and from particle to particle. Averaging out such quantities over the representative volume element essentially smooths out their variation to such a degree that a continuum representation is possible. Within the continuum mechanics framework, powder compaction can be viewed as a forming event during which large irrecoverable deformation takes place as the state of the material changes from loose packing to near full density. From a continuum point of view, the properties of the material evolve; for example, the material stiffness increases during compaction. Therefore, the material properties at any one stage during compaction should be expressed as a function of some measures of its state, that is, one or more state variables.

2.2 Elastoplastic Constitutive Models In the case of materials that undergo microstructural changes during the deformation process, for example, powder under die compaction, there are three components of the constitutive model that need to be defined:

i)

the yield criterion, which defines the transition of elastic to plastic deformation;

Powder Technology, Part II: Compaction

10

ii) iii)

the plastic flow potential, which dictates the relative amounts of each component of the plastic flow; the evolution of the microstructure, which in turn defines the resistance to further deformation.

The total strain increment resulting from the application of stress consists of two components, a reversible and an irreversible one. It is assumed here that their decomposition is additive [a more complex multiplicative decomposition of the strain has been developed and is descript in Literature]: e d ij = d ij + d ijpl (2.1) where subscripts i and j vary from 1 to 3 with reference to the coordinate system axes and indicate the three-dimensional nature of the strain. Upon release of the external loads, the reversible (elastic) part of the strain is recovered. The relationship between macroscopic stress and elastic strain and their increments is approximated to the first order by a linear Hookes law:
d ij =
k ,l =1..3

e Lijkl d kl

i = 1..3, j = 1..3

(2.2)

This is the generalization of the one-dimensional form = E to a three-dimensional form. In a simple isotropic case, the 81 (3x3x3x3) components of the elastic stiffness Lijkl contain only two independent parameters (e.g., Youngs modulus and Poissons ratio). The second term on the right-hand side of eq. (2.1) represents the irreversible (plastic) part of the deformation and reflects a variety of underlying irreversible mechanisms such as particle rearrangement, fragmentation, and plastic deformation at the particle level. The basis of a continuum model for compaction is the yield locus, that is, a stress based criterion that defines the limits of elastic deformation in the form of an inequality:

if

F ( ij ,km ,...)< 0, then deformation is elastic F ( ij ,km ,...) =0, then deformation is plastic
F ( ij , k m , ...)

(2.3)

In the general case, the yield function

which can be geometrically described as a

surface in stress space, may be a function not only of the stress state but also internal state variables (km, m = 1 . . . n), that characterize all aspects of the material that affect yielding. When the yield condition is satisfied, a flow potential is postulated to exist and determines the three dimensional character of plastic deformation: G ( ij , km ) d ijpl = d (2.4) ij
G = plastic flow potential; d = plastic strain multiplier The plastic flow potential can also be a function of the stress state, internal state variables, strain rates, deformation history, and temperature. In principle, a large number of parameters may be required to describe faithfully all physical and geometrical aspects of the material and its microstructure; in the case of powder densification a single internal state variable, the relative density, is enough for a first-order approximation of the constitutive model. For isotropic materials, the representation of the stress state can be simplified by use of stress invariants. Here, we will consider functional forms of the yield locus and the plastic potential that depend on two stress invariants, the hydrostatic pressure (p) and Mises equivalent or

Powder Technology, Part II: Compaction

11

effective stress (q), which are associated with volumetric plastic strains and distortion, respectively: 1 p = ( 11 + 22 + 33 ) 3 (2.5) 1/ 2 2 2 2 2 2 2 2 q = ( 11 + p ) + ( 22 + p ) + ( 33 + p ) + 2 12 + 2 23 + 2 13 3 For an isotropic material, that is, for a material with properties that do not vary with direction, it is logical to represent the yield locus and flow potential as functions of these stress invariants. In addition, the yield criterion is often affected by the changes in the structure of the powder compact. Plastic (irreversible) deformation leads to a permanent change in the structure of the compacted powder. A simple scalar quantity is often used, namely, the relative density (RD) of the material, defined as the ratio between the density of the material at a given state and the density of the fully dense material. This selection implies that the elastic properties, the yield condition, and the flow rule are function of the relative density: Lijkl ( RD ); F(q,p,RD) = 0 and G(q,p,RD) = 0 (2.6) Equations (2.1), (2.2), and (2.4) are written in an incremental form because the state of the material evolves during compaction. The state of the microstructure of the compact determines the material response and must be tracked during the process. Its evolution is given on the basis of geometrical terms: d ( RD ) = RD d iipl (2.7)
1=1..3

The complete continuum mechanics constitutive model consists of eqs. (2.1), (2.2), and (2.4). The material properties (2.6) that are determined experimentally are: (a) the elastic constants, (b) the yield surface, and (c) the plastic potential. The model is then completed with the evolution of the state variable according to eq. (2.7), which represents the hardening rule. Several types of constitutive models have been developed and applied in the analysis of compaction of metal and ceramic powders, including Gurson2, Cam-Clay3, and DPC4. The latter model has gained more widespread use because it captures important aspects of the physics of compaction, such as i) the fact that a compact is stronger in compression than shear and tension, ii) it includes commonly used material parameters such as cohesion, angle of internal friction, and yield pressure, and their dependence with internal variables such as RD.
2.2.1 The Cam-Clay Model

The Cam-clay model is a classical plasticity model and is the basis for the inelastic constitutive theory for modelling cohesionless material. Many authors have used the Camclay plasticity model for modelling the powder consolidation process. This is because it is a basic soil model upon which other constitutive models are based. It has a proven record of accomplishment for predicting soil behaviour in conventional triaxial tests and consolidation
Gurson AL. 1977. Continuum theory of ductile rupture by void nucleation and growth. Part I. Yield criteria and flow rules for porous ductile media. Transactions of the ASME. J Eng Mat Technol Jan: 2-15. 3 Schofield AN, Wroth CP. 1968. Critical state solid mechanics. New York: McGraw Hill. 4 Drucker DC, Gibson RE, Henkel DJ. 1957 Soil mechanics and work hardening theories of plasticity. ASCE Trans 122:338-346.
2

Powder Technology, Part II: Compaction

12

analysis. In addition, the model parameters have a well-defined physical meaning. The Camclay theory uses a strain rate decomposition in which the rate of mechanical deformation of the powder is decomposed into an elastic and a plastic part. The main features of the Camclay model are the use of either a linear elasticity or a porous elasticity model, which exhibits an increasing bulk elastic stiffness as the material undergoes compression. The clay plasticity model describes the inelastic behaviour of the material by a yield function that depends on the three stress invariants. An associated flow assumption defines the plastic strain rate, and a strain hardening theory changes the size of the yield surface according to the inelastic volumetric strain. The volume change is

J = J g J el J pl
where J is the ratio of current volume to original volume, Jg is the ratio of current to original volume of the soil grain particles, Jel is elastic part of the ratio of current to original volume of the soil volume, Jpl is the plastic part of the ratio of current to original volume of the soil volume. Volumetric strains are defined as

(2.8)

vol = ln J
el vol = ln J el pl vol

= ln J pl

(2.9)

Thus, the usual additive strain rate decomposition for volumetric strain rates is given by, g el pl d vol = d vol + d vol + d vol (2.10) The elastic volume change of a material sample is given by 1 + e el J el = 1 + e0

(2.11)

Where e0 is the initial void ratio, eel is the void ratio during elastic draining. The void ratio, e, is defined as the ratio of the volume occupied by the voids, Vv, to the volume occupied by the solid part, Vs.

e=
and is related to relative density, by

VV VS

(2.12)

e=

(2.13)

For the Cam-clay model, the elastic behavior is modeled using the porous elasticity model, typically with a zero tensile strength. The porous elasticity model is based on the fact that in porous and granular media, elastic properties are strongly dependent on volumetric strain. For soils, volumetric elastic strain is proportional to the logarithm of equivalent pressure stress. The porous elasticity model is based on the experimental observation that in porous materials during elastic straining, the change in void ratio, e and the change in logarithm of the equivalent pressure stress, p are linearly related, i.e. Powder Technology, Part II: Compaction 13

de el = d (ln p )

(2.14)

where is the logarithmic elastic bulk modulus (a material parameter) In this form, the material has zero tensile strength. If the tensile strength is nonzero, the equivalent relation is de el = d (ln p + ptpl ) (2.15)

Figure 12 shows the virgin consolidation curves comprising of the hydrostatic swelling line and the hydrostatic compression lines. The slopes of these lines give the elastic and plastic bulk moduli respectively. These parameters are needed to define the hardening characteristics of the Cam-clay model.

Fig. 12: Compression behavior for the Cam-clay model Integrating the linear relation from Equation (2.14) and using the relation expressed in (2.11) gives the following volumetric elasticity relationship: p (2.16) ln = 1 J el 1 + e0 p0 where p0 is the initial pressure, prescribed by initial conditions. For zero tensile strength material, it is required that p0 > 0. Hence, the equation yields, 1 + e0 el (1 exp vol ) p = p0 exp (2.17)
el The above relationship between the equivalent stress, p, and volumetric strain, vol can be illustrated in Figure 13. The intercept, p0, is the initial value of the equivalent hydrostatic stress and is used to characterize the hardening characteristics of the Cam-clay model.

Fig. 13: The Cam-Clay Porous elastic volumetric behavior

Powder Technology, Part II: Compaction

14

Yield surfaces are used to express the relation between various stresses. They show the boundaries of certain type of material response. Let us examine a yield surface in the plane of equivalent stress (the hydrostatic pressure (p) and Mises equivalent or effective stress (q), see equ. 2.5). In the p-q plane, the Cam-clay yield surface is made up of two elliptic arcs as shown in Figure 14. One arc passes through the origin where it is tangent to the equivalent stress axis. It intersects the critical state line where its tangent is parallel to the pressure stress axis. The other arc is a smooth continuation of the first arc through the critical state line and intersects the pressure axis at some nonzero value of pressure stress, again with its tangent parallel to the equivalent stress axis.

Fig. 14: Cam-clay yield surface in the p-q plane The model is based on the yield surface 1 p t 1 + 1 = 0 2 q Ma
2 2

(2.18)

where p is the equivalent pressure stress t is a deviatoric stress measure M is a constant that defines the slope of the critical state line is a constant equal to 1.0 on the dry side of the critical state line but may be different from 1.0 on the wet side of the critical state line. a is a hardening parameter that defines the size of the yield surface The deviatoric stress measure, t is given by 3 1 1 1 r t = q 1 + 1 2 K K q where q is the Mises equivalent stress, K is the ratio of flow stress in triaxial tension to the flow stress in triaxial compression. r is the third stress invariant The hardening/softening law controls the size of the yield surface in effective stress space. The size of the yield surface at any time is determined in terms of the hardening parameter, a. It depends on the volumetric plastic strain component. When the volumetric plastic strain is Powder Technology, Part II: Compaction 15

(2.19)

compressive (i.e. when the powder is compacted), the yield surface grows in size. An inelastic increase in the volume causes the yield surface to shrink. The hardening law can have an exponential or a piecewise linear form. Figure 15 illustrates the hardening behavior for the Cam-clay model. Consider the stress path of a sample at A subjected to compression. As the load is applied, the load path progresses from A to B with only elastic changes to the void ratio. Further loading produces an elastic-plastic response and subsequent yield surfaces are reached (C), each with the same shape. Eventually the load path arrives at the critical state line (D) where failure occurs.

Fig 15: Hardening behavior of a Cam-clay material In the exponential form, the hardening parameter, a, is determined by an initial value of the hardening parameter, a0, and the amount of inelastic volume change that occurs, Jpl. 1 J pl (2.20) a = a0 exp (1 + e0 ) J pl The initial size of yield surface is expressed by specifying the initial value of the hardening parameter, a0. This parameter is dependent only on the initial conditions and does not vary with temperature or any other field variable. It can be expressed in terms of the intercept of the virgin consolidation line with the void ratio axis, e1, in the plot of void ratio versus logarithm of the effective pressure stress (refer Figure 12.). 2.2.2 Drucker-Prager-Cap Model (DPC) TheDPC Model at low hydrostatic pressures is a shear failure model, similar to those used in granular flow like the Cam-Clay Model, and reflect the dependence of the strength on the confining pressure. It predicts that the strength in tension is smaller than that in compression, a concept that is common for rocks, brittle materials, and pressed powder compacts. In its simplest form, it is represented by a straight line in the pq plane, which is also known as the Mohr-Coulomb shear failure line FS: Fs ( q, p ) = q d p tan( ) = 0 (2.21) The two parameters are termed cohesion d5 and internal friction angle b.6 If the stress state is such that the corresponding Mises equivalent stress and hydrostatic pressure result in a value of F(q, p)<0, then the stress causes only elastic deformation. If the stresses are such that eq. (2.21) is satisfied, the material fails in shearing.

5 6

d = 0 in the Cam-Clay Model = M in the Cam-Clay model

Powder Technology, Part II: Compaction

16

2.2.3 Example

Dans l'article [ D. Poquillon], les auteurs cherchent modliser le plus ralistement possible la densification de 2 poudres de fer lors de leur compactage. Ces 2 poudres sont presses uniaxialement dans une matrice rectangulaire (32.36x13.28mm) ou cylindriques (D=10mm) Poudre S: les grains sont plutt sphriques et correspondent des agglomrats de taille moyenne 15m de petites cristallites de 0.5 1 m. Poudre E: les grains irrguliers sont rugueux et spongieux et sont des agglomrats de taille moyenne 30 m de cristallites mtalliques aciculaires enchevtres d'1 2 m . Les conditions du compactage sont les suivantes: la pression applique sous air varie de 100 350MPa et la pression finale est maintenue pendant 2min du starate de zinc dans une solution actate est ajout pour amliorer le compactage Ainsi, 2 types de corps verts sont obtenus pour chaque poudre utilise: des cylindres appels PS et PE, et des barres appeles S et E. Description de la densification lors du processus de compactage: 1. Cet tat de dformation lastique est rversible. 2. Les particules se rarrangent et le glissement des interparticules. 3. Dformation plastique due entrelacement des particules qui se localisent en premier au niveau des contacts entre les particules. Les tapes 1 et 2 du compactage sont dcrites par le modle de Cam-Clay (Figure 12) tandis que la troisime tape est modlise par la loi de puissance.
Modle de Cam-Clay: ce modle a initialement t dvelopp pour simuler le comportement des argiles en tenant compte de la dformation plastique et du glissement des grains les uns sur les autres. Cette approche peut aisment tre transpose au compactage des poudres sphriques commerciales de fer. Comme seule la compression uniaxiale est considre ici, les quations mcaniques de ce modle peuvent tre simplifies: la seule dformation possible a lieu le long de l'axe de compression.

zz =

0 1

(2.22)

zz = deformation le long de l'axe de z


o 0 et sont respectivement les densits relatives initiale et actuelle du corps cru. D'aprs ce modle, la rponse du matriau dpend de 5 paramtres: deux dcrivent l'tat initial: le taux de vide initial e0 et la pression d'coulement plastique Pi0 trois sont des caractristiques intrinsques: la pente de la droite d'tat critique M = qm / pm o qm et pm sont les maxima respectifs de la contrainte dviatrice et de la contrainte moyenne, la pente de charge et de dcharge et la pente de charge sur la surface de rendement. La Figure 16 permet de visualiser graphiquement les paramtres et . Dans le modle de Cam-Clay, M est constante et indpendante de Pi0. Comme seuls les tests uniaxiaux sont
Powder Technology, Part II: Compaction 17

effectus, M n'est pas value ici. La Figure 17 reprsente schmatiquement la rponse du matriau selon le modle de Cam-Clay : Etape 1: seule la dformation lastique des grains est implique: la courbe de dcharge est donc similaire la courbe de charge. La rugosit de surface prvient le glissement des particules. Les tensions rversibles proviennent de la dformation lastique du squelette lors du rarrangement des particules. = 0 log 10 ( Pc ) (2.23) o = 0.2 et Pc la pression de compactage.

Fig. 16 : Reprsentation schmatique de l'volution du taux de vide avec la pression de compactage selon le modle de Cam-Clay. Dfinition des pentes et [ D. Poquillon].

Etape 2: Pi0 est atteint => le taux de vide dcrot rapidement d glissement des grains. Lors de la dcharge, la rponse lastique possde la mme pente que celle de l'tape 1 mais la dformation totale n'est plus nulle: le glissement des grains n'est plus rversible (cf.: Figure 12). Pio = 12 + 0 (2.24)

Powder Technology, Part II: Compaction

18

Fig.17 : Evolution prdite du taux de vide avec la pression de compactage applique pendant les tapes 1 et 2 (courbes avec les petits symboles). Comparaison avec les valeurs exprimentales (courbes avec les gros symboles). La ligne pointille reprsente la limite entre les tapes 1 et 2 [D. Poquillon]. Avec des quations 2.9 et 2.10 on trouve :

= ( Pi 0 ) log10
et

Pc Pi 0

(2.25)

0 3.5
1.48 log10 ( Pi 0 )

(2.26)

Les tapes 1 et 2 semblent parfaitement modlises : les valeurs exprimentales et simules sont en totale adquation. Dformation plastique due entrelacement des particules qui se localisent en premier au niveau des contacts entre les particules est dcrit par une troisime tape : Etape 3: modle de la loi de puissance: pour une pression P1=35MPa. = 1 + ( P P1 ) (2.27) o P1 et 1 sont la pression et la densit la fin de l'tape 2, est la capacit de matriau se dformer (cette valeur est plus forte pour la poudre E que pour la poudre S) et n quivaut au coefficient de durcissement (il est gale 0.55 pour les simulations de fer). Rsultats: Les valeurs calcules par l'quation sont en bonne corrlation avec les rsultats exprimentaux. La loi de puissance (2.13) est bien adapte l'tape 3. Conclusions: Les 3 tapes de la densification d'un compactage uniaxial de poudres de fer sont bien modlises par le modle de Cam-Clay et la loi de puissance. Powder Technology, Part II: Compaction 19

Application of the DCP Model on the compaction of a real ceramic part Source: Numerische Simulation des Pressens und Sinterns mit ABAQUS T. Kraft, O. Coube und H. Riedel, Fraunhofer Institut fr Werkstoffmechanik, Freiburg

Fig 18 : Finite-Elemente-Netz der Pulverschttung der Dichtscheibe

Fig. 19 : Dichteverteilung der Oberseite des Grnlings (Dichte = 0 exp(PEQC4) mit der theoretische Dichte 0 = 3.71 g/cm3)

Fig 20 : Sinterverzug

Powder Technology, Part II: Compaction

20

3. Micromechanical compaction models7

Micromechanical compaction models uses averaging techniques to describe the evolution of a representative reference spherical particle under isotropic densification (Arzt, 19828; Larsson et al., 19969) or arbitrary deformation history (Fleck, 199510; StorDakers et al., 199911).The use of the Radial Distribution Function (RDF) that describes the evolution of the number of particle centres around a given particle for a monodispersed random dense packing is an important tool for these models. It enables one to describe the evolution of contacts during densification starting from the initial random configuration of the particles. In such an approach the kinematics of particles is given by a homogeneous strain field. The homogeneous strain field assumption imposes that each particle motion is given by the macroscopic strain field. Hence, particles indent each other without local rearrangement and no rotation is allowed. For example in the case of isostatic compaction, particles indent each other normally without any tangential shift. The assumption that rearrangement is negligible during the plastic compaction is believed to represent adequately the density regime at which particles are constrained by neighbours because of the large number of contacting particles. From this homogeneous strain field assumption, it is possible to relate the local overlap h between two particles, the contact radius a (Fig. 21), and the average coordination number Z (average number of contacts per particle) to the macroscopic relative density of the packing (D).

Fig. 21: 2-D sketch of the contact geometry between two spherical particles. Particles radius are rp and rq, the overlap is h and the contact radius is a. The angle of obliquity is given by the tangential and normal components of the relative velocity of the centre of one particle with respect to the centre of the other.
Text after C.L. Martina, D.Bouvard , S.Shima Study of particle rearrangement during powder compaction by the Discrete Element Method Journal of the Mechanics and Physics of Solids 51 (2003) 667 693 Arzt, E., 1982. The infuence of an increasing particle coordination on the densification of spherical powders. Acta Metall.Mater.30, 18831890. Larsson, P.L., Biwa, S., StorDakers, B., 1996. Analysis of cold and hot compaction of spherical particles. Acta Mater.44, 36553666.
10
11 9

Fleck, N.A., 1995.On the cold compaction of powders.J.Mech.Phys.Solids 43, 14091431.

Storakers, B., Fleck, N.A, McMeeking, R.M., 1999. The viscoplastic compaction of composite powders. J.Mech.Phys.Solids 47, 785815.

Powder Technology, Part II: Compaction

21

components of the relative velocity of the centre of one particle with respect to the centre of the other. Once the local contact force law is established (with the possibility of rate dependence) and the overall equilibrium at the macroscopic level is written, these models are able to predict the evolution of the relative density as a function of the isostatic pressure or anisotropy effects in the case of more complex loading. These predictions are valid for the early state of compaction (stage I) for which the particles can be described as truncated spheres with mechanical independence of contacts and no geometric impingement between contacts (D >0.90 in any case).The comparison with available experimental data shows adequate results concerning the evolution of the average coordination number and the evolution of the average contact area. The comparison of the calculated pressure is more problematic but still satisfactory.
3.1 The model 3.1.1. Constitutive equations for the contact

The particles are modelled as spheres made of an elasto-plastic material with elastic constants E and and the following hardening relation in the plastic regime: (2.28) = 0 1/ m where 0 is a material constant, m is the hardening coefficient and and are the stress and strain in the uniaxial case. For two contacting particles p and q with radius rp and rq (Fig. 21), we note
r* = rp rq rp + rq

(2.29)

As compaction proceeds, particles will overlap with each other, with h defining the overlap (Fig.21). In the following, n and t denote the unit normal vector to the contact plane and the unit vector parallel to the contact plane in the direction of the relative velocity. The normal force N at the contact is given by the following relations: elastic deformation:
Ne = plastic deformation:
N p = 0 213/ 2 m 311/ m c( m) 2+1/ m ( r * )12 / m h1+1/ 2 m

2 E r * h 3/ 2 2 3 1

(2.30)

(2.31)

where c(m) is a function that is related to the size of the contact area. These expressions are strictly valid only for frictionless contact and for normal indentation. However, it has been shown that friction does not have a large e4ect on the value of the normal force. Concerning the tangential force at the contact, the contact is considered to be either in a sticking state, with negligible relative tangential displacement, or in a state of gross sliding. When the contact is not sliding, a sufficiently large tangential stiffness is assigned to the contact in order to ensure sticking. When the contact is sliding, a simple law of friction is used.

Powder Technology, Part II: Compaction

22

3.1.2 Plastic redistribution of matter at the contact

When plastic deformation occurs, the excess material at the contact between the two overlapping particles is transported away from the contact zone. Depending on the constitutive behaviour of the powder, the material may be deposited far away from the contact zone or stay in the vicinity of the contact. For two particles indenting normally, it has been demonstrated that the hardening exponent m has an important effect on the redistribution profile and on the contact area a2 which is given as 2 a 2 = 2 c ( m ) r *h (2.32)

with the function c(m)2, which appears in Eq.( 2.31), that varies from 0.5 for linear hardening (m=1) to 1.45 for perfect plasticity (m=).When the material is close to perfect plasticity, it can be assumed that the matter at the contact is deposited close to the contact (short-range redistribution) and does not interact with neighbouring particles
3.1.3 2.3. Discrete element method description

The interactions of the spherical particles are accounted for by modelling the evolution of the packing as a dynamic process. At each incremental time step, the contacts between particles are determined and the interparticle forces at the contact are calculated. The total force acting on a particle is then determined and Newtons second law enables the computation of the new accelerations x , velocities x and positions x of each particles using a small time step t. For the typical mass, size and stiffness that are involved in metallic powders applications, the time step is less than 107 s. Such a time step together with the large deformation that is necessary for compaction would result in prohibitive CPU time. Using a technique that has been applied for example for quasi-static deformation simulation of particulate media, the nominal density of the
particles is scaled up by a factor .Under such a scheme, the accelerations and velocities are reduced by orders of magnitude but the forces and strains are unchanged. 3.2 Examples:

Used materials:

The 1rst step before the plastic compaction of the sample is to generate an acceptable random dense packing of particles. In order to create this initial packing, 4000 particles are randomly generated in a cubical cell with no initial contact between the particles. At this stage the relative density is of the order of 0.32 and the packing is in a gas-like condition. From this stage, the preparation of the sample consists in imposing an isostatic densification with the friction set to zero. The zero friction coefficient allows packings to attain a high relative density at the end of the sample preparation.

Powder Technology, Part II: Compaction

23

The cumulative Radial Distribution Function G(r) characterises the number of particle centres within a sphere of radius r for a given packing.Fig. 22 displays the calculated function G(r) from the 4000 particles sample compacted in this study.

Fig. 22: Evolution of the cumulative radial distribution function obtained experimentally by Mason (1968) and Nair et al.(1986) and calculated from the 4000 monosize particles packing compacted isostatically up to the 1rst plastic contact by DEM ().Linear fit to the data of Mason (- - -) used by analytical models. The arrow indicates the value of r = 2R at which full density is obtained with the homogeneous strain field kinematics assumption.

Fig 23: Evolution of the average coordination number Z as a function of relative density D. Comparison of the homogeneous strain field solution from Arzt (1982) or Storakers et al.(1999) and the experimental data of Fischmeister et al.(1978)12 .
Fischmeister, H.F., Arzt, E., Olsson, L.R., 1978. Particle deformation and sliding during compaction of spherical powders: a study by quantitative metallography. Powder Metall.21, 179187.
12

Powder Technology, Part II: Compaction

24

The average contact area provides some more insight into the contact evolution during the isostatic compaction. Figs. 23 and 24 indicate that simulations and/or models that predict low average coordination number are associated with larger contact areas than those that anticipate larger coordination number. This is typically the case when local rearrangement is not taken into account (the homogeneous strain 1eld solution and the model of Storakers et al.(1999) ).As mentioned earlier, when local rearrangement between particles is fully active as for the 0.1 friction coefficient curve and the oblique indentation approximation curve, additional small contacts are anticipated compared to the solutions with no rearrangement. In any case the average contact area associated with simulations with rearrangement is lower than for the simulation with no rearrangement.

Fig. 24: Evolution of the normalised average contact area with the relative density.Comparison of the experimental data from Fischmeister et al.(1978) () and the model of Storakers et al.(1999) with DEM simulations.

Fig. 25:Evolution of the hydrostatic pressure with relative density. Comparison of DEM simulations with the experimental data on an aluminium powder and with the model of Storakers et al.(1999) .

Powder Technology, Part II: Compaction

25

S-ar putea să vă placă și