You are on page 1of 33

Accepted Manuscript

Compaction and solid-state sintering of tungsten powders: MPFEM simulation and


experimental verification

Qian Jia, Xizhong An, Haiyang Zhao, Haitao Fu, Hao Zhang, Xiaohong Yang

PII: S0925-8388(18)31266-0
DOI: 10.1016/j.jallcom.2018.03.387
Reference: JALCOM 45617

To appear in: Journal of Alloys and Compounds

Received Date: 30 December 2017


Revised Date: 28 March 2018
Accepted Date: 29 March 2018

Please cite this article as: Q. Jia, X. An, H. Zhao, H. Fu, H. Zhang, X. Yang, Compaction and solid-state
sintering of tungsten powders: MPFEM simulation and experimental verification, Journal of Alloys and
Compounds (2018), doi: 10.1016/j.jallcom.2018.03.387.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

Compaction and Solid-State Sintering of Tungsten Powders: MPFEM


Simulation and Experimental Verification
Qian Jia, Xizhong An∗, Haiyang Zhao, Haitao Fu, Hao Zhang, Xiaohong Yang
School of Metallurgy, Northeastern University, Shenyang 110004, PR China

Abstract

PT
The uniaxial die compaction and solid-state sintering of different sized tungsten powders

in powder metallurgy (PM) process was numerically reproduced in multi-particle FEM

RI
(MPFEM) modeling from particulate scale. The effects of particle size, initial packing

SC
structure, compaction pressure and sintering temperature on the relative density of the

tungsten powder component were systematically studied and discussed. Various macroscopic

U
and microscopic properties of the powder mass with different initial packing structures during
AN
compaction and sintering were characterized and analyzed. These properties include the

overall relative density, local stress distributions, force structure and transmission, particle
M

rearrangement and deformation, void filling behavior, as well as the densification dynamics

and mechanism. The results show that by properly controlling the operating conditions such
D

as initial packing structure, compaction pressure and sintering temperature, high performance
TE

PM component with high relative density, uniform density, stress, and void distributions can

be obtained. Meanwhile, physical experiments were implemented to verify the simulation


EP

results. It is indicated that MPFEM modeling used in current work can provide the researchers

with an effective method to simulate the whole PM process for refractory tungsten powders
C

from particulate scale, especially the coupling of different stages can reduce the assumptions
AC

in the modeling and make the simulation results more accurate and much closer to the actual

process.

Keywords: Tungsten powders; Solid-state sintering; Uniaxial die compaction; Powder

Metallurgy; Particulate scale MPFEM simulation


Corresponding author. Tel: +86 24 83689032; Email: anxz@mail.neu.edu.cn

1
ACCEPTED MANUSCRIPT

1. Introduction

Tungsten metals are widely used in many key areas of industry like aerospace, military,

electronics and chemistry etc. due to their superior performance such as high strength and

hardness, high elastic modulus, small thermal expansion coefficient, and excellent corrosion

resistance properties [1,2]. However, because of the high melting point of tungsten (3410 °C),

PT
casting and other traditional forming methods are not applicable in the production of tungsten

metals and components [3-5]. In comparison, powder metallurgy (PM) has recently become a

RI
promising method in manufacturing refractory metals like tungsten and molybdenum etc. due

SC
to its advantages such as low cost, materials saving, accurate structure control, easy operation,

mass production, and most importantly the environment friendly and net shape or near net

U
shape forming. Therefore, PM has nowadays been widely used in this regards. Among PM
AN
procedure, powder compaction stage and solid-state sintering stage play a key role. However,

feedback from the production practice indicates that there are still some bottlenecks yet to be
M

solved. For example, how to obtain the tungsten components with high relative density,

uniform local density and stress distributions and excellent mechanical properties? What are
D

the dynamics and mechanisms during compaction and sintering? How to precisely control the
TE

microstructure? All these problems still restrict its further developments. Under this

circumstance, much work in the past decades was carried out both physically and numerically,
EP

and quick progresses have been made.

Physically, using different PM forming methods and sintering processes, numerous


C

experiments have been carried out to get high performance pure tungsten and tungsten alloy
AC

parts. In forming stage, most studies were focusing on the effect of compaction pressure,

tungsten powder size and pressing rate on the properties of the tungsten compacts [6-17]. For

example, Poster [6] and Bewlay [7] discussed the effects of the pressure on the relative

density using die compaction and dynamic compaction. Bewlay [8] and Kyu [9] studied the

effects of tungsten particle diameters and compaction rates during isostatic pressing and

plasma pressure compaction, respectively. Bhalla [15], Kim [16] and Cho et al. [17] fabricated

2
ACCEPTED MANUSCRIPT

tungsten composites and tungsten alloy composites by various routes including uniaxial

compaction, triaxial compaction, isostatic pressing, explosive compaction, and made

comparison with each forming method. In sintering stage, much work has been conducted to

study the effects of sintering temperature, sintering time, heating rate, tungsten powder sizes

on the properties of the final sintered components including microstructure of tungsten grains,

PT
various defects, strength , hardness, toughness and so on [18-24]. For instance, Kthari [18]

and Abu-Oqail [22] researched the effect of sintering time and temperature. German et al. [19]

RI
built a predictive model to describe the relationship between sintered strength, sintered

hardness, density and tungsten powder size. Krian [20] and Hyesook [21] decreased tungsten

SC
crystallite size and found that fine tungsten powder could get high density tungsten sintered

component at the reduced sintering temperature. Even though corresponding physical

U
experiments could reveal the impact of different factors on the PM forming process and
AN
sintering process from macroscopic perspective, they could not intuitively and effectively

characterize the powder behavior during compaction and sintering behavior from particulate
M

scale. I.e. the particle rearrangement and deformation, stress distributions and transfer, as well

as the densification dynamics and mechanisms are hard to be identified in-situ in physical
D

experiments. Satisfactorily, these barriers in physical experiments can be overcome by


TE

numerical simulations.
EP

Numerically, there are mainly two branches in modeling the PM process of the tungsten

powder, i.e. forming and sintering. In the forming process, large amount of simulation work
C

was carried out by using continuous finite element method (FEM) from macroscopic scale
AC

[25-33]. For example, John [25] and Zohoor and Mehdipoor [26,27] simulated the explosive

compaction process of tungsten powders and studied the effects of the wall thickness of the

hollow cylinder and detonation velocity on cracks and grain size of the obtained tungsten

parts, respectively. Zhou [28] and Almanstötter et al. [29] simulated the die compaction

process of doped tungsten powders using modified Drucker-Prager Cap model [31-33], where

axial and radial stresses as well as axial strain were discussed. In the sintering process, many

numerical simulations have been done based on different methods such as Monte Carlo (MC)
3
ACCEPTED MANUSCRIPT

[34-38], master sintering curve (MSC) method [39-40] and FEM [45-48]. For instance,

Aladazabal and coworkers [34,35] proposed a MC model to simulate the liquid phase

sintering process of tungsten. Park et al. [39] analyzed the phase transformation during

sintering using MSC concepts. Jin et al. [45] modeled the liquid-phase sintering of tungsten

heavy alloys with FEM, in which the distortion and shrinkage of the alloys were predicted.

PT
Even though previous studies can to a certain extent solve the problem and shed lights in

compaction and sintering, while the results are limited. For example, MC modeling can

RI
describe the growth of crystals while it is in molecular scale including only several hundreds

of atoms or molecules. In MSC and FEM simulations, most current studies are focusing on

SC
macroscopic properties in forming and sintering processes with continuum models. It is

difficult to concurrently consider the compaction and sintering together within a same model,

U
each stage is independent with many assumptions. Most importantly, the simulations using
AN
these methods are failed to characterize the microscopic properties such as stress and

distributions, force structure and transmission, particle rearrangement and deformation, void
M

filling behavior, as well as the densification dynamics and mechanisms during compaction

and sintering. These deficiencies can be conquered by MPFEM modeling, which can provide
D

the researchers or engineers with more intensive understanding on the compaction and
TE

sintering of tungsten powders in PM production from particulate scale.


EP

In this paper, the uniaxial die compaction and solid-state sintering of different sized

tungsten powders in PM process were numerically reproduced using MPFEM modeling from
C

particulate scale. The utilized MPFEM method can combine the advantages of continuous
AC

FEM and discrete DEM (discrete element method) modeling and has been successfully

applied in our previous simulations on the compaction of metals and alloys with low melting

point [49-52]. The effects of particle size, initial packing structure, pressing pressure and

sintering temperature on the properties of tungsten powder components were systematically

studied and discussed. Meanwhile, various macroscopic and microscopic properties of the

powder mass with different initial packing structures during compaction and sintering were

characterized and analyzed, these properties include the evolution of overall relative density,
4
ACCEPTED MANUSCRIPT

local relative density and stress distributions, force structure and transmission, particle

rearrangement and deformation, void filling behavior, as well as the densification dynamics

and mechanisms.

2. Simulation method and conditions

PT
DEM and FEM were coupled from particulate scale to investigate the behavior of

individual tungsten particle including its rearrangement, deformation, interactions with

RI
neighboring particles during compaction and sintering. To numerically reproduce the whole

PM process in the actual production, the tungsten powder was first randomly generated in the

SC
die and then falling downward to form the stable static packing for die filling by using DEM

simulation. As a mature and effective dynamic simulation method, DEM has been

U
successfully applied in modeling the packing of particles by researchers since 1980s.
AN
Therefore, the governing equations and force models in DEM are not given here. Interested

readers please refer to our previous studies for details [53-56]. Once the die filling is finished,
M

the initial packing structure is imported into FEM model to complete initialization which

includes materials property and boundary condition setting, contact definition, loading mode
D

and temperature setting etc. Then the compacting followed by solid-state sintering is
TE

simulated by using MPFEM method from particulate scale through commercialized

MSC.Marc software. In this duration, the effects of initial packing structures, operating
EP

parameters, macroscopic and microscopic property characterization and analyses, as well as

the identification of densification dynamics and mechanism are systematically conducted. The
C

flowchart of the whole PM process is shown in Fig. 1 (a).


AC

In the simulation, different initial packing structures are constructed by tungsten powders

with different particle sizes (the diameter d is respectively 1 µm, 5 µm, 10 µm, 15 µm, 20 µm,

50 µm and 100 µm) by DEM. To unify the die wall effects, same ratio of particle diameter

versus die diameter (here cylindrical die was used) is applied in DEM. Here, equal numbers

of particles are generated in DEM for each case to easily compare initial packing structures

with various particle sizes. In addition, to show more clearly the effects of initial packing

5
ACCEPTED MANUSCRIPT

structures on compaction and sintering, the images of particles with different sizes are

enlarged to the same diameter (here is 1 mm) when the DEM generated data are imported into

FEM model. The advantage of the MPFEM model lies in that the particles in the powder mass

during PM processing can be individually discretized for mesh division and analysis. To

guarantee the accuracy in the whole simulation process and detect the evolution particulate

PT
characteristics both internally and externally, homogeneous meshes are designed. The mesh

division for individual particle is shown in Fig. 1 (b), where each particle contains 200 nodes

RI
and 173 elements. Fig. 1 (c) gives the DEM generated initial packing structure in the die. In

the simulation, the die and the punches are set to be rigid and the tungsten powder is set to be

SC
elasto-plastic. The compaction is performed from the upper punch with the lower punch and

the die being fixed.

U
AN
Fig. 1 (a) Flowchart in the numerical modeling on the PM process of tungsten powders; (b) mesh division
for individual particle and (c) DEM generated initial packing structure in the die before compaction.
M

To comprehensively consider the pressure and temperature achieved during compaction

and sintering process, the Johnson-Cook model [57] is used to describe the properties of
D

tungsten materials. In this model the flow yield stress is expressed as:
TE

σ y = (A + Bε n )(1 + C ln ε&* )(1 − T * ) m (1)


EP

where A is the yield strength of the material under quasi-static state, B and n are strain

hardening impact indices, C is the strain rate sensitivity index, m is the temperature softening
C

coefficient, ε is the effective plastic strain, T * = (T − Troom ) / (Tmelt − Troom ) is the homologous
AC

temperature; ε&* = ε& / ε&0 , ε& is the strain rate, ε&0 stands for the reference strain rate, ε&* is a

description of the effect of the strain rate enhancement. The effect of the strain rate

enhancement is often considered during high velocity compaction when ε&0 is not equal to 1

s-1. However, since the ordinary die compaction is used in this simulation and the effect of

strain rate enhancement is ignored, ε&0 is set to 1 s-1.


6
ACCEPTED MANUSCRIPT

The strain rate ߝሶ consists of two parts in compaction: elastic strain rate ߝ௘ሶ and plastic
strain rate ߝ௣ሶ ,

ε& = ε&e + ε& p (2)

During sintering, the thermal strain rate ߝ௧௛ሶ is added with ߝ௘ሶ and ߝ௣ሶ to form ߝሶ,

PT
ε& = ε&e + ε& p + ε&th (3)

RI
Here, the elastic strain rate is assumed to be linear and isotropic, as expressed by:

SC
ε&e = Ceσ& (4)

where ۱‫ ܍‬is the elastic compliance matrix. Eq. (4) can also be expressed in the rate from of

Hooke’s law:
U
AN
σ& = Deε&e (5)
M

where ۲‫ ܍‬is the elastic stiffness matrix. For isotropic materials, ۲‫ ܍‬is defined as:

1−v v v 0 0 0
D

‫ۍ‬ v 1−v v 0 0 0 ‫ې‬


‫ێ‬ ‫ۑ‬
୉ v v 1−v 0 0 0 ‫ۑ‬
۲‫( = ܍‬ଵା୴)(ଵିଶ୴) ‫ێ‬
TE

(6)
‫ێ‬ 0 0 0 1 − 2v 0 0 ‫ۑ‬
‫ێ‬ 0 0 0 0 1 − 2v 0 ‫ۑ‬
‫ۏ‬ 0 0 0 0 0 1 − 2v‫ے‬
EP

where E is the Young’s modulus, ν is the Poisson’s ratio. The thermal strains are mainly due to

the thermal expansion as expressed by:


C
AC

ε&th = α T&I (7)

Where α is the thermal expansion coefficient, ܶሶ is the incremental temperature rate, ۷ is the

second order identity tensor. The value of α varies with the temperature for tungsten materials

[59], which can be represented by:

α = −8.69 × 10 −3 + 3.83 × 10 −4 T + 7.92 × 10 −8 T 2 (8)

7
ACCEPTED MANUSCRIPT

When defining the contact, all particles in the model are considered as deformable, while

the die and the punches are set as rigid bodies. Since the friction coefficient between tungsten

particles is difficult to measure, therefore the friction coefficient of 0.2 between particles,

which has been widely used in previous researches [49-51], is also utilized in current

simulations. The contacts between the particles and the rigid bodies are considered as smooth.

PT
During compaction, the pressure from the upper punch will increase from 0 MPa to 1000 MPa,

and then it will be kept constant for a while and be released. During sintering, the temperature

RI
will gradually rise from 25 °C (room temperature) to approximately 2300 °C. In the

simulation, 24.75 s is adopted to describe compaction and sintering processes. Here, 0-1.5 s

SC
corresponds to the loading time; 1.5-1.7 s is the holding time; 1.7-2 s is the unloading time;

and 2-24.75 s is used for sintering. Table 1 lists the material parameters of tungsten powder

U
and the modeling conditions used in current work.
AN
Table 1 Material parameters and modeling conditions used in the simulation.
M

3. Results and discussion


D

3.1 Model validation


TE

To verify the numerical model, corresponding physical experiments were conducted.

Here, the commercial tungsten powder (with the average particle size of 5.48 µm as measured)
EP

with nearly spherical shape was purchased and its composition is listed in Table 2. It can be

seen that the tungsten powder is with relatively high purity. With this powder, uniaxial die
C

compaction was performed in a cylindrical die with an inner diameter of 20 mm. During
AC

physical experiments, the compaction pressure is ranging from 0 to 1000 MPa. Since the

particle size distribution follows the normal distribution and the probability of the particle

diameter with 5.48 µm accounts for about 90% of the powder, implying that most particles are

with the size around 5.48 µm. So 5.48 µm is used as the uniform diameter to simulate

compaction and solid-state sintering processes. Both the compaction curve and the

microstructure obtained in current simulation are compared with our physical results. For

8
ACCEPTED MANUSCRIPT

sintering process, since the sintering temperature is too high to be reached in our physical

experimental conditions, the numerical results will be compared with those from others’ work

[60] for model validation.

Table 2 Composition of the tungsten powder used for experiments.

PT
Fig. 2 (a) gives the relative density evolution of the tungsten powder compact with the

pressure during compaction (compaction curve) in both numerical and physical experiments.

RI
As indicated, the compaction curves in both cases follow the similar trend. While one can also

SC
find that there is a gap between numerical and physical results, which can be ascribed to the

differences between 2D numerical model and 3D physical model, the assumption of particle

U
shape and uniform particle size in the numerical model. Here, the difference between 2D
AN
numerical model and 3D physical model is the main reason because of the loss a degree of

freedom during compaction in 2D. Even though the 2D results cannot fully reproduce the
M

results obtained in 3D, yet the properties such as particle deformation and growth, force

structure and transmission, void filling behavior, as well as the densification dynamics and
D

mechanism all can be identified in 2D models during compaction and solid-state sintering,
TE

which can provide valuable information and a good starting point for the study in 3D. In

addition, the simulation results obtained under the pressure of 0 to 600 MPa are fitted with the
EP

double logarithmic equation [50,51,63]:

( ρm - ρ0 ) ρ
m lg ln[ ] = lg P − lg M (9)
C

( ρm - ρ ) ρ0
AC

where ρm is the theoretical density of the fully dense metal, ρ0 is the apparent density of the

initial packing of the compact, ρ is the apparent density of the compact, P is the applied

pressure, m represents the hardening index and M stands for compaction modulus. Here, each

kind of density in double logarithmic equation is with the unit of kg/m3. As shown in Fig. 2 (b)

that the fitting results exhibit high confidence (high R2 value), which implies the robustness

and reliability of our 2D numerical compaction models. For the verification of the sintering

9
ACCEPTED MANUSCRIPT

process, results from previous work [60] are used for comparison as shown in Fig. 3. One can

see that qualitative agreement of the morphologies of sintered parts obtained from physical

and our numerical work can be identified. Above validations from physical experiments and

analytical equation all provide evidences for the effectiveness of the utilized numerical

models.

PT
Fig. 2 (a) Relative density evolution of tungsten powder compact with the pressure in both numerical and

RI
physical experiments, where the inset figures indicate the packing morphologies (SEM images) of the
compacts physically obtained under different pressures; (b) fitting of the simulation results with double

SC
logarithmic equation [50,51].

Fig. 3 Comparison of morphologies of the sintered parts obtained from both physical experiments [60] (a)
and our numerical simulations (b).

U
AN
3.2 Compaction
M

During die filling, the initial packing density as a function of the particle size is shown in

Fig. 4 (a), where the inset figures indicate the corresponding packing morphologies. It can be
D

seen that with the increase of the particle size, the initial relative density increases, and the
TE

packing structure becomes denser. However, the relative density increment is less significant

when d > 20 µm. Once the stable initial packing is formed, the external pressure from the
EP

upper punch is exerted on the top of the packing for compaction. To identify the universality

of the simulation results, three initial different packing structures formed by the same sized
C

particles were generated for compaction and sintering. Fig. 4 (b) gives the morphologies of
AC

the compacts and the compaction curves for the three cases when the size of the tungsten

powder is 1 µm. One can find that only some minor differences exist, implying that the effects

of the initial packing structures are not significant for powders with the same particle size. Fig.

4 (c) illustrates the evolution of the relative density (defined as the volume occupied by

particles divided by the volume of the die) with the pressure during compaction on tungsten

powder mass with different particle sizes. As can be seen that for each sized powder mass,

three stages during densification can be identified, which is consistent with the results from
10
ACCEPTED MANUSCRIPT

physical experiment [10] and continuum FEM simulations [28,29]. In Stage I (0-320 MPa),

the compaction pressure is relatively low, within which the relative density sharply increases

with the pressure due to the particle rearrangement of sliding and rolling. With the further

increase of the pressure, the compaction steps into Stage II (320-520 MPa) where the relative

density continuously increases, while the increasing rate decreases. In this stage, small

PT
rearrangement and elastic deformation aid the packing densification. When the pressure is

very high (e.g. > 520 MPa), the compaction reaches Stage III. In this stage, the relative

RI
density of the compact does not change much with further increase of the compaction

pressure. This is because tungsten powders have high elastic modulus, even the pressure is

SC
very high, plastic deformation is still very hard to occur at room temperature. In addition,

through comparison one can find that the compaction curves for coarse powders are higher in

U
position than those for fine powders. However, when the particle size is more than 20 µm, this
AN
difference becomes less obvious as indicated in stage III. This trend is in good accordance

with that in Fig. 4 (a), indicating that the denser the initial packing is, the much easier is the
M

compaction densification, and the more superior is the performance of the compact or even

the subsequently sintered part, which will be further discussed in the following section.
D
TE

Fig. 4 (a) Relative density of the initial packing structure as a function of the size of tungsten particles,
where the inset figures indicate the corresponding packing morphologies; (b) relative density-pressure
EP

relationships during compaction on tungsten powders with different initial packing structures, where the
particle size is 1 µm and the inset figures indicate the initial packing structures of three cases; (c) relative
density-pressure relationships during compaction on tungsten powders with different particle sizes.
C

In addition to the macroscopic property in Fig. 4, the microscopic properties, i.e. the
AC

stress/force distributions of the compacts with different particle sizes (here we take d = 1 µm

and d = 20 µm as an example for analysis) were characterized to explore the densification

mechanism during compaction. Fig. 5 shows the equivalent von Mises stress distributions and

normal contact force chains (force structure and transmission) in the compacts formed by

small ((a), (b)) and large particles ((c), (d)), respectively. It can be seen that the force chains

formed between large particles are more evenly distributed than those formed between small

11
ACCEPTED MANUSCRIPT

particles. Hence, the forces in the compact of coarse particles are easier to be transmitted than

those in the compact of fine particles.

Fig. 5 Equivalent von Mises stress distributions ((a), (c)) and Normal contact force structures (force

PT
chains)/transmissions ((b), (d)) in the compacts formed by small (d = 1 µm, shown in (a), (b)) and large (d
= 20 µm, shown in (c), (d)) particles, respectively.

RI
3.3 Solid-state sintering

Previous studies indicate that even the compaction pressure is very high, the relative

SC
density of the tungsten powder compact is still very low, which is mainly determined by the

special materials characteristics of metal tungsten. Therefore, in order to obtain very dense

U
tungsten component, sintering stage in PM process is necessary, while the research in this
AN
aspect from particulate scale is rather sparse in literature. In this section, detailed numerical

work was carried out to simulate the sintering of the tungsten compact and study the effects of
M

particle size and sintering temperature on the macroscopic and microscopic properties of the

final component. Fig. 6 illustrates the evolution of relative density with the temperature for
D

powder mass with different particle sizes during solid-state sintering, where the inset figures
TE

are the morphologies of local structures at different temperatures. One can see that three

stages can also be categorized during sintering process [61,62], whose boundaries are
EP

determined by the number nodes and detailed discussion can been found in next section. The

initial stage (0-300 °C) is the sintering neck formation stage in which the particles are
C

transformed from point contact or surface contact to form crystal binding. The intermediate
AC

stage (300-1650 °C) is the sintering neck growth stage, which is most significant for

densification by reducing porosity; in this stage, our results are agreeable with the results

from the literature [45,46]. The final stage (1650-2300 °C) is characterized by isolated pores

located at the corners of the grain boundaries, the relative density changes a little by closing

pores in this stage. Meanwhile, Fig. 6 also shows that the relative densities of the sintered

parts formed by large particles or small particles are almost the same at the end. Hence, it can

12
ACCEPTED MANUSCRIPT

be concluded that temperature is the dominant factor in the densification process of sintering.

The parts with relative density of 98% are obtained at 2000 °C, which improves the quality,

strength, and hardness of the tungsten components.

PT
Fig. 6 Evolution of relative density with temperature for powder mass with different particle sizes during
solid-state sintering, where the images of local structures in the inset figures are respectively corresponding

RI
to: (a) 300 °C; (b) 1650 °C; (c) 2300 °C.

To further study the densification mechanism from particulate scale during the solid-state

SC
sintering process, the evolution of the node contact number and relative displacement with the

temperature is carefully investigated with particle diameter equal to 20 µm. Fig. 7 (a) gives

U
the evolution of number of contact nodes with the temperature during sintering when d = 20
AN
µm, where the inset figures are the overall contact structures at different temperatures. One

can find that the trend of the node contact number with the temperature is consistent with the
M

relative density, which can also be divided into three stages. In the initial stage, the number of

the contact nodes is almost constant, indicating that the sintering neck is formed and slowly
D

grows. In the intermediate stage, increased number of the contact nodes manifests the neck
TE

growth. While in the final stage, the variation of the node contact number becomes

insignificant, and the pores are gradually spheroidized and closed in this process. In addition,
EP

evolution of the relative displacement with the temperature and the displacement vector

distribution of each node have been researched as respectively shown in Fig. 7 (b) and Fig. 8.
C

Here, the dimensionless relative displacement is defined as the displacement divided by the
AC

initial displacement. It can be seen from Fig. 8 that large displacement is concentrated in the

upper part. Therefore, two particles as marked in Fig. 7 (b) were selected at the top (particle 1)

and bottom (particle 2) for quantitative analysis. Here, the relative displacement is defined as

the displacement of the particle divided by its diameter (d). As indicated, the displacement of

the particles does not change much with the temperature. However, the overall displacement

of particle 2 is less than that of particle 1 due to the dense initial sintered structure around

13
ACCEPTED MANUSCRIPT

particle 2. Through the comparison between Fig. 8 (b)-(e), one can find that particle 1 is

coarsened after sintering, implying the poor performance of the final sintered components. In

addition, it can also be seen from the inset figure of Fig. 7 (b) that particle 1 tends to move

toward the large pores, indicating the pore-filling mechanism for the densification process. In

addition, the particles with regular hexagonal shapes can be obtained in the dense structures

PT
after sintering, which ensure the isotropy of the components.

RI
Fig. 7 Evolution of the number of contact nodes (a) and relative displacement (b) with the temperature
during solid-state sintering when d = 20 µm.

SC
Fig. 8 (a) Displacement vector distribution in the sintered component at 2300 °C when d = 20 µm, where

U
the color indicates the size of the displacement and the vector represents the direction of the displacement;
(b)-(e) the shapes of particle 1 ((b), (c)) and particle 2 ((d), (e)) before ((b), (d)) and after ((c), (e)) sintering.
AN
Although the particle size has comparatively weak impact on the densification process,
M

its effects on stress concentration in the powder mass after solid-state sintering is significant.

As shown in Fig. 9, stresses are mainly concentrated at pore structures and boundaries where
D

cracks are easy to initiate and propagate, which can deteriorate the property of the final
TE

tungsten components, and this result is similar to that in physical experiments [23].

Meanwhile, the stress concentration in the sintered part formed by small sized particles is
EP

even more serious, this is because the structures (refers to the structures before compaction,

after compaction, and after sintering) composed of small sized particles include numerous
C

voids, and this phenomenon has been identified in others’ work [19]. However, the grains of
AC

small particles are finer after sintering which will get better structure property. Therefore, the

best case is selecting relatively large particles with small porosity for sintering. Based on this

consideration, powders with the particle size of 20 µm seems to be the best choice and the

relative density of 98% for the components with finer sintered grains, uniform and small

residual stresses can be obtained by sintering suitable tungsten compacts at the temperature of

about 2000 °C.

14
ACCEPTED MANUSCRIPT

Fig. 9 Residual equivalent von Mises stress distributions in the components after solid-state sintering at
2300 °C, where: (a) d = 1 µm; (b) d = 20 µm.

4. Conclusions

Cold uniaxial die compaction and solid-state sintering of tungsten powders with different

PT
particle sizes were numerically reproduced by using multi-particle finite element method

(MPFEM) from particulate scale. The effects of operating parameters during compaction and

RI
sintering were studied, and various macroscopic and microscopic properties of the compacts

and sintered components were systematically characterized and compared. Finally the

SC
densification dynamics and mechanisms were identified. Following conclusions can be

drawn:

U
AN
(1) The powder composed of relatively large sized particles is eligible to form much denser

initial packings and compacts, more uniform stress distribution, and smaller sintered
M

residual stress. But these enhancements become insignificant when the particle diameter

is greater than 20 µm.


D

(2) The relative density of 98% for the components with finer sintered grains, uniform and
TE

small residual stresses can be obtained by sintering suitable tungsten compacts at the

temperature of about 2000 °C. And the suitable tungsten compacts refer to green
EP

compacts obtained by pressing tungsten powder particles with 20 µm under pressure of

500 MPa.
C

(3) During compaction, the large pores formed in the initial packing are mainly filled by the
AC

rearrangement and tiny deformation of tungsten powders. However, large amount of rest

pores are filled by the growth of the sintered neck during sintering at the temperature of

300-1650 °C, induced by tungsten softening phenomenon. And increasing temperature

can make the tungsten particles easier to deform.

(4) MPFEM modeling used in current work can provide an effective method to simulate the

whole PM process for refractory tungsten powders from particulate scale, especially the
15
ACCEPTED MANUSCRIPT

coupling of different stages which can reduce the assumptions in the modeling and make

the simulation results more accurate and much closer to the actual process.

Acknowledgements

The authors are grateful to National Key Development Program of the Ministry of

PT
Science and Technology (2017YFB0305603) of China and Fundamental Research Funds for

the Central Universities of China (N162505001) for the financial support of current work.

RI
References

[1] Kelly, W.R. Tyson, Tensile properties of fibre-reinforced metals: copper/tungsten and

SC
copper/ molybdenum, J. Mech. Phys. Solids 13 (1965) 329-338.
[2] L. Zhang, W. Chen, G. Luo, Low-temperature densification and excellent thermal
properties of W-Cu thermal-management composites prepared from copper-coated

U
tungsten powders, J. Alloy. Compd. 588 (2014) 49-52.
AN
[3] C.M. Mate, G.M. Mcclelland, R. Erlandsson, Atomic-scale friction of a tungsten tip on a
graphite surface, Phys. Rev. Lett. 59 (1987) 1942-1947.
[4] R.M. German, E. Olevsky, Mapping the compaction and sintering response of
M

tungsten-based materials into the nanoscale size range, Int. J. Refract. Met. H 23 (2005)
294-300.
[5] S.H. Yoo, T.S. Sudarshan, Dynamic compression behavior of tungsten powders
D

consolidated by plasma pressure compaction, Powder Metall. 42 (1999) 181-182.


[6] A.R. Poster, Factors affecting the compaction of tungsten powders, Powder Metall. 5
TE

(1962) 301-315.
[7] W.H. Gourdin, S.L. Weinland, Dynamic compaction of a monosized spherical tungsten
powder, first ed., Shock Waves in Condensed Matter, Springer US, 1986.
EP

[8] B.P. Bewlay, Consolidation dynamics of tungsten powder during dry bag cold isostatic
pressing, J. Alloy. Compd. 11 (1992) 165-174.
C

[9] C.C. Kyu, H.W. Robert, Plasma pressure compaction of tungsten powders, Mater. Manuf.
Process 19 (2004) 619-630.
AC

[10] K.T. Kim, J.H. Cho, J.S. Kim, Cold compaction of composite powders, J. Eng. Mater. Tec.
122 (2000) 119-128.
[11] J. Gwen, R. Joanna, Plasma activated sintering (PAS) of tungsten powders, Mater. Manuf.
Process 9 (1994) 1105-1114.
[12] B.P. Bewlay, Consolidation dynamics of tungsten powder during dry bag cold isostatic
pressing, Int. J. Pefract. Met. H 11 (1992) 165-174.
[13] X. Chen, X. Li, H. Yan, Factors affecting explosive compaction-sintering of
tungsten-copper coating on a copper surface, J. Alloy. Compd. 729 (2017) 729-735.
[14] M. Dias, F. Guerreiro, J.B. Correia, Consolidation of W-Ta composites: Hot isostatic

16
ACCEPTED MANUSCRIPT
pressing and spark and pulse plasma sintering, Fusion Eng. Des. 98 (2015) 1950-1955.
[15] A.K. Bhalla, J.D. Williams, A comparative assessment of explosive and other methods of
compaction in the production of tungsten-copper composites, Powder Metall. 19 (1976)
31-37.
[16] J.C. Kim, S.S Ryu, Y.D. Kim, Densification behavior of mechanically alloyed W-Cu
composite powders by the double rearrangement process, Scripta Mater. Tech. 39 (1998)
669-676.
[17] J.H. Cho, K.T Kim, Cold compaction of composite powders, J. Eng. Mater. 122 (2000)

PT
119-128.
[18] N.C. Kothari, Sintering kinetics in tungsten powder, J. Les. Comm. Met. 5 (1963)
140-150.

RI
[19] R.M. German, E. Olevsky, Strength predictions for bulk structures fabricated from
nanoscale tungsten powders, Int. J. Refract. Met. H 23 (2005) 77-84.

SC
[20] U.R. Kiran, M.P. Kumar, M. Sankaranarayana, High energy milling on tungsten powders,
Int. J. Refract. Met. H 48 (2015) 74-81.
[21] J. Hyesook, C. Han, K. Byungmoon, Interface activated sintering of tungsten by

U
nano-particles in the spark plasma sintering, Rev. Adv. Mater. Sci. 28 (2011) 157-168.
[22] A. Abu-Oqail, M .Ghanim, M. El-Sheikh, Effects of processing parameters of
AN
tungsten-copper composites, Int. J. Refract. Met. H 35 (2012) 207-212.
[23] L.J. Kecskes, K.C. Cho, Grain size engineering of bcc refractory metals: Top-down and
bottom-up-application to tungsten, Mat. Sci. Eng. A 467(2007) 33-43.
M

[24] R. Liu, X.P. Wang, Characterization of ODS-tungsten microwave-sintered from sol-gel


prepared nano-powders, J. Nucl. Mater. 450 (2014) 69-74.
D

[25] J.E. Reaugh, Computer simulations to study the explosive consolidation of powders into
rods, J. Appl. Phys. 61 (1987) 962-968.
TE

[26] M. Zohoor, A. Mehdipoor, Compaction of tungsten powder using an explosive


compaction setup and numerical simulation of the process, Wseas Trans. App. Theo.
Mech. 1 (2006) 62-69.
EP

[27] M. Zohoor, A. Mehdipoor, Numerical simulation of underwater explosive compaction


process for compaction of tungsten powder, Mater. Sci. Forum 5 (2008) 77-82.
[28] R. Zhou, L.H. Zhang, Numerical simulation of residual stress field in green power
C

metallurgy compacts by modified Drucker-Prager Cap model, Trans. Nonferrous Met.


Soc. China 23 (2008) 2374-2382.
AC

[29] J. Almanstötter, A modified Drucker-Prager Cap model for finite element simulation of
doped tungsten powder compaction, Int. J. Refract. Met. H 50 (2015) 290-297.
[30] O. Coube, H. Riedel, Numerical simulation of metal powder die compaction with special
consideration of cracking, Powder Metall. 43 (2000) 123-131.
[31] L.H. Han, J.A .Elliott, A.C. Bentham, A modified Drucker-Prager Cap model for die
compaction simulation of pharmaceutical powders, Int. J. Solids. Struct. 45 (2008)
3088-3106.
[32] W. Bier, S. Hartmann, A finite strain constitutive model for metal powder compaction
using a unique and convex single surface yield function, Eur. J. Mech. 25 (2006)
17
ACCEPTED MANUSCRIPT
1009-1030.
[33] L. Resende, J.B. Martin, Formulation of Drucker-Prager Cap model, J. Eng. Mech. 111
(1985) 855-881
[34] J. Aldazabal, A. Martinmeizoso, J.M. Martinezesnaola, Simulation of liquid phase
sintering using the Monte Carlo method, Mat. Sci. Eng. A 365 (2002) 151-155.
[35] A. Luque, J. Aldazabal, A. Martínmeizoso, Simulation of the microstructural evolution
during liquid phase sintering using a geometrical Monte Carlo model, Model. Simul.
Mater. Sc. 13 (2005) 1057-1060.

PT
[36] R.M. German, P. Suri, S.J. Park, Review: liquid phase sintering, J. Mater. Sci. 44 (2009)
1-39.
[37] A. Luque, J. Aldazabal, J.M. Martínez-Esnaola, Geometrical Monte Carlo model of

RI
liquid-phase sintering, Math. Comput. Simulat. 80 (2010) 1469-1486.
[38] S.B. Lee, J.M. Rickman, A.D. Rollett, Three-dimensional simulation of isotropic

SC
coarsening in liquid phase sintering I: A model, Acta Mater. 55 (2007) 615-626.
[39] S.J. Park, R.M. German, J.M. Martin, Densification behavior of tungsten heavy alloy
based on master sintering curve concept, Metall. Mater. Trans. A 37 (2006) 2837-2848.

U
[40] S.J. Park, J.L. Johnson, Y. Wu, Analysis of the effect of solubility on the densification
behavior of tungsten heavy alloys using the master sintering curve approach, Int. J.
AN
Refract. Met. H 37 (2013) 52-59.
[41] S.J. Park, P. Suri, E. Olevsky, Master sintering curve formulated from constitutive models,
J. Am. Ceram. Soc. 92 (2009) 1410-1413.
M

[42] S.J. Park, J.M. Martin, J.F. Guo, Grain growth behavior of tungsten heavy alloys based on
the master sintering curve concept, Metall. Mater. Trans. A 37 (2006) 3337-3346.
D

[43] R.M. German, Sintering theory and practice, Second ed., Wiley, New York, 1996.
[44] D.C. Blaine, S.J. Park, R.M. German, Application of work-of-sintering concepts in
TE

powder metals, Metall. Mater. Trans. A 37 (2006) 2827-2835.


[45] P.S. Jin, C.S. Hwan, J.L. Johnson, Finite element simulation of liquid phase sintering with
tungsten heavy alloys, Mater. Trans. 47 (2006) 2745-2752.
EP

[46] B. Mcwilliams, A. Zavaliangos, Multi-phenomena simulation of electric field assisted


sintering, J. Mater. Sci. 43 (2008) 5031-5035.
[47] A. Moitra, S. Kim, S.J. Park, Investigation on sintering mechanism of nanoscale tungsten
C

powder based on atomistic simulation, Acta Mater. 58 (2010) 3939-3951.


[48] A. Upadhyaya, R.M. German, Shape distortion in liquid-phase-sintered tungsten heavy
AC

alloys, Metall. Mater. Trans. 29 (1998) 2631-2638.


[49] F. Huang, X.Z. An, Y.X. Zhang, Multi-particle FEM simulation of 2D compaction on
binary Al/SiC composite powders, Powder Technol. 314(2017) 39-48.
[50] Y.X. Zhang, X.Z. An, Y.L. Zhang, Multi-particle FEM modeling on microscopic behavior
of 2D particle compaction, App. Phy. A 118 (2015) 1015-1021.
[51] P. Han, X.Z. An, Y.X. Zhang, Particulate scale MPFEM modeling on compaction of Fe
and Al composite powders, Powder Technol. 314 (2017) 69-77.
[52] G. Gustafsson, H.Å. Häggblad, P. Jonsén, Multi-particle finite element modelling of the
compression of iron ore pellets with statistically distributed geometric and material data,
18
ACCEPTED MANUSCRIPT
Powder Technol. 239 (2013) 231-238.
[53] X.Z. An, R.Y. Yang, K.J. Dong, R.P. Zou, A.B. Yu, Micromechanical simulation and
analysis of one-dimensional vibratory sphere packing, Phys. Rev. Lett. 95 (2005)
205-502.
[54] X.Z. An, A.B. Yu, Analysis of the forces in ordered FCC packings with different
orientations, Powder Technol. 248 (2013) 121-130.
[55] X.Z. An, R.Y. Yang, R.P. Zou, Effect of vibration condition and inter-particle frictions on
the packing of uniform spheres, Powder Technol. 188 (2008) 102-109.

PT
[56] H.Y. Zhao, X.Z. An, DEM modeling on stress profile and behavior in granular matter,
Powder Technol. 323 (2018) 149-154.
[57] G.R. Johnson, W.H. Cook, Fracture characteristics of three metals subjected to various

RI
strains, strain rates, temperatures and pressures, Engin. Frac. Mech. 21 (1985) 31-48.
[58] K.D. Dai, P.W. Chen, Numerical simulation of the shock compaction of W/Cu powders,

SC
Mater. Sci. Forum 673 (2011) 113-118.
[59] B.Y. Jiang, L. Wang, Viscosity model parameters fitting of feed stock in MIM simulation
and analysis, Chin. J. Nonferrous Met. 15 (2005) 429-434.

U
[60] J. Shi, N. Liu, A. Liu, Effect of pressing pressure on microstructure and mechanical
property of pure tungsten, Heat Treat. 30 (2015) 31-35.
AN
[61] R.L. Coble, Sintering crystalline solids. I. intermediate and final state diffusion models, J.
Appl. Phys. 32 (1961) 787-792.
[62] R.M German, Sintering theory and practice, first ed., John Wiley, New York, 1996.
M

[63] J. Hu, Applications of Huang Peiyun’s double logarithmic equation of powder


compacting, Powder Metal. Technol. 3 (1987) 11-20.
D
TE
C EP
AC

19
ACCEPTED MANUSCRIPT

Table 1 Material parameters and modeling conditions used in the simulation.


Material parameters Value Contact definition Load setting Jobs
E/GPa 133
ν 0.22 Particles: deformable
Iteration method: full Updated Lagrange
A/GPa [52] 1510 Die and punch: rigid
Newton-Raphson for large strain
B/GPa [52] 177 Friction between
algorithm Large deformation
n [52] 0.12 particles: 0.2
Segment to

PT
C [52] 0.016 Convergence criterion:
Upper punch: load segment method
Troom/°C 25 relative and residuals
control
Tmelt/°C 3410

RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

Table 2 Composition of the tungsten powder used for experiments.

Composition W Ca Fe Si
Contents (wt. %) 99.8877 0.0723 0.0352 0.0075

PT
RI
U SC
AN
M
D
TE
CEP
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
Fig. 1 (a) Flowchart in the numerical modeling on the PM process of tungsten powders; (b) mesh division
for individual particle and (c) DEM generated initial packing structure in the die before compaction.
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

(a) Simulation (5.48 µm) (b)


Experiment (5.48 µm) 0.0
0.8

lgln[(ρm-ρ0)ρ/(ρm-ρ)ρ0]
Relative Density

-0.4
0.6 1 µm 1 µm R2 = 0.983
5 µm 5 µm R2 = 0.984
-0.8
5.48 µm 5.48 µm R2 = 0.985
10 µm 10 µm R2 = 0.975
0.4 15 µm 15 µm R2 = 0.984
-1.2 20 µm 20 µm R2 = 0.992

PT
50 µm 50 µm R2 = 0.989
100 µm 100 µm R2 = 0.989
0.2 -1.6
0 200 400 600 800 1000 1.2 1.6 2.0 2.4 2.8 3.2
Pressure (MPa) lg P

RI
Fig. 2 (a) Relative density evolution of tungsten powder compact with the pressure in both numerical and
physical experiments, where the inset figures indicate the packing morphologies (SEM images) of the

SC
compacts physically obtained under different pressures; (b) fitting of the simulation results with double
logarithmic equation [50,51].

U
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
Fig. 3 Comparison of morphologies of the sintered parts obtained from both physical experiments [60] (a)
and our numerical simulations (b).

U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
(c)
0.8

SC
III
Relative Density

0.7
1 µm
5 µm

U
II
0.6 10 µm
15 µm
AN
20 µm
I
50 µm
0.5
100 µm

0 200 400 600 800 1000


M

Pressure (MPa)

Fig. 4 (a) Relative density of the initial packing structure as a function of the size of tungsten particles,
D

where the inset figures indicate the corresponding packing morphologies; (b) relative density-pressure
relationships during compaction on tungsten powders with different initial packing structures, where the
TE

particle size is 1 µm and the inset figures indicate the initial packing structures of three cases; (c) relative
density-pressure relationships during compaction on tungsten powders with different particle sizes.
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
Fig. 5 Equivalent von Mises stress distributions ((a), (c)) and Normal contact force structures (force
chains)/transmissions ((b), (d)) in the compacts formed by small (d = 1 µm, shown in (a), (b)) and large (d
= 20 µm, shown in (c), (d)) particles, respectively.
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
Fig. 6 Evolution of relative density with temperature for powder mass with different particle sizes during

SC
solid-state sintering, where the images of local structures in the inset figures are respectively corresponding
to: (a) 300 °C; (b) 1650 °C; (c) 2300 °C.

U
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
Fig. 7 Evolution of the number of contact nodes (a) and relative displacement (b) with the temperature
during solid-state sintering when d = 20 µm.

U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
Fig. 8 (a) Displacement vector distribution in the sintered component at 2300 °C when d = 20 µm, where

SC
the color indicates the size of the displacement and the vector represents the direction of the displacement;
(b)-(e) the shapes of particle 1 ((b), (c)) and particle 2 ((d), (e)) before ((b), (d)) and after ((c), (e)) sintering.

U
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
SC
Fig. 9 Residual equivalent von Mises stress distributions in the components after solid-state sintering at
2300 °C, where: (a) d = 1 µm; (b) d = 20 µm.

U
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

Highlights

• Whole PM process of tungsten powders were simulated by MPFEM from particulate scale.
• Influences of operating parameters on densification behaviour was systematically studied.
• Various macro and micro properties of the powder mass in each PM stage were
characterized and analysed.

PT
• Densification dynamics and mechanisms are identified.

RI
U SC
AN
M
D
TE
C EP
AC

You might also like