Disipación de ondas de compactación en explosivos granulares de baja densidad

La microestructura de explosivos granulares puede afectar el efecto de disipación del calentamiento que se produce a causa de las ondas de choque dentro de ellos y pueden desencadenar tanto una combustión e iniciar una detonación. Debido a que los fenómenos de iniciación ocurren sobre distancias que son más grandes que el tamaño medio de partícula, la teoría homogeneizada, que trata los fenómenos de la detónica en la macroescala, y es un riguroso método de promediar basado en extensiones de la perturbación en múltiples escalas, a menudo se utilizan para describir estados termodinámicos locales dentro y detrás de la onda de choque, tales que son considerados como la manifestación promedio de campos termodinámicos en la escala de la partícula.

The microstructure of granular explosives can affect dissipative heating within compaction shocks that can trigger combustion and initiate detonation. Because initiation occurs over distances that are much larger than the mean particle size, homogenized (macroscale) theories are often used to describe local thermodynamic states within and behind shocks that are regarded as the average manifestation of thermodynamic fields at the particle scale. In this paper, mesoscale modeling and simulation are used to examine how the initial packing density of granular HMX (CHNO) CHNO having a narrow particle size distribution influences dissipation within resolved, planar compaction shocks. The model tracks the evolution of thermomechanical fields within large ensembles of particles due to pore collapse. Effective shock profiles, obtained by averaging mesoscale fields over space and time, are compared with those given by an independent macroscale compaction theory that predicts the variation in effective thermomechanical fields within shocks due to an imbalance between the solid pressure and a configurational stress. Reducing packing density is shown to reduce the dissipation rate within shocks but increase the integrated dissipated work over shock rise times, which is indicative of enhanced sensitivity. In all cases, dissipated work is related to shock pressure by a density-dependent power law, and shock rise time is related to pressure by a power law having an exponent of negative one.

I. INTRODUCTION

Shock initiation of low density granular explosives (GXs) is an inherently macroscale phenomenon that occurs over distances that are appreciably larger than the mean particle size. Because GXs contain high inter-particle porosity (>25% by volume), dispersed compaction shocks can induce intense thermomechanical fluctuations at the particle scale due to inelastic pore collapse, even for relatively mild shocks. The integrated effect of dissipative heating within compaction shocks is important because it can locally trigger combustion that can spread and subsequently lead to detonation under suitable confinement. Unlike GXs, plastic-bonded explosives (PBXs) contain much less porosity (<5%); therefore, comparatively stronger shocks are required to trigger combustion, and possibly by different mechanisms, such as friction work occurring at deconsolidated explosive-binder surfaces. Mesoscale Modeling and Simulation (M&S) is currently being used to examine the evolution of thermomechanical fluctuations within and behind shocks in GXs and PBXs,1–11 but the relative importance of dissipation mechanisms, their dependence on microstructure (e.g., particle size, shape, and packing distributions, and intra-particle defects) and loading conditions, and their influence on macroscale behavior remain largely unresolved issues. In addition to enhancing fundamental understanding of such phenomena, this information is useful in developing microstructure-dependent theories for shock induced ignition and combustion of these materials at the macroscale. It is tacitly assumed that mesoscale modeling applied to these materials is compatible with commonly used macroscale theories, though it is not often rigorously examined, particularly as it pertains to compaction shocks in GXs.

The primary objective of this study is to computationally examine dissipation within resolved planar compaction shocks in a GX and to establish a functional relationship between dissipated work and shock strength, and its dependency on particle packing density. Dissipated work is important because it is sensitive to the shock loading history of the material and it represents a key macroscale measure of hot-spot formation at the particle scale. Models used to describe compaction of granular materials, including GXs, often relate changes in effective porosity to pressure,12 which can result in significant dissipation and hot-spots provided that the compaction rate is sufficiently high. Such a relation can also facilitate the development of macroscale theories that describe history-dependent phenomena, such as shock desensitization. This paper summarizes predictions given by both mesoscale M&S and an independent macroscale compaction theory applied to HMX (CHNO). The compaction theory is based on principles of mixture theory and is an extension of the theories formulated by Baer and Nunziato13 and Powers .14 The explosive HMX is chosen because its compaction behavior is empirically well-characterized and it is commonly used in applications.15–18 Only simple microstructures are considered to isolate the effects of initial packing density from those due to complex variations in particle size and shape.

A secondary objective of this study is to examine the extent to which mesoscale predictions of compaction shock profiles agree with those given by the macroscale compaction theory. No attempt is made to adjust or fit macroscale model parameters to obtain best agreement with mesoscale predictions; rather, emphasis is placed on validating the macroscale theory based on resolved mesoscale computations for quasi-steady shocks. Such examinations can also provide rigorous mesoscale interpretations of macroscale predictions, particularly as it pertains to the relation between microstructure, shock dissipation, and hot-spot formation. An analysis of hot-spots based on predicted temperature fields within particles (as done in Ref. 9) is beyond the scope of this study; however, it is possible to formulate a comprehensive relation that includes hot-spot information, such as intensity, size, and spatial proximity, for use with macroscale theories. This issue constitutes an ongoing topic of research.

The paradigm for this study is illustrated in Figs. 1 and 2. The mesoscale analysis describes two-dimensional (2D), plane strain loading of a large ensemble of deformable HMX particles. The representative ensemble shown in Fig. 1 consists of approximately 4000 randomly packed, initially circular particles having an average size of d¯=60μm

with a narrow distribution. The average initial solid volume fraction of the ensemble is ϕ¯0=0.84 and all inter-particle pores are massless. Though the assumptions of 2D plane strain and circular particle geometry are restrictive, they enable leading-order effects of packing density on compaction shock dissipation to be examined based on well-resolved computations; they also form a basis for systematically examining the effects of other microstructural features and three-dimensionality in the future. Periodic conditions are imposed along transverse boundaries of the computational domain, and a free condition is imposed at the far-field axial boundary. A rigid piston impacts the material with constant speed causing the rapid development and subsequent propagation of a quasi-steady compaction shock having speed . Shock strength is characterized in terms of effective pressure P¯¯¯ which depends on both ϕ¯0 and . Figure 2(a) illustrates a computed piston supported shock having speed  = 2136 m/s for ϕ¯0=0.84 and   = 500 m/s. The shock quickly develops into a quasi-steady wave following impact, as indicated by its position-time curve in Fig. 2(b), and its pressure P¯¯¯=1.6GPa

is sufficiently high to eliminate porosity. The shock speed is lower than the ambient sound speed of solid HMX (≈2740 m/s) due to energy dissipation and dispersion associated with pore collapse.

Click to view

FIG. 1.Representative computational domain and boundary conditions used for the mesoscale simulations.

Click to view

FIG. 2.Development of a quasi-steady compaction shock corresponding to  = 500 m/s and ϕ0=0.84

: (a) pressure contours at  = 4.5 s and (b) shock position as a function of time indicating a quasi-steady shock speed of ≈ 2136 m/s.

As discussed later in this paper, quasi-steady compaction shock profiles are obtained by first filtering thermomechanical fields at fixed times and then performing ensemble averaging of the resulting spatially filtered profiles, as illustrated in Fig. 3 for the case shown in Fig. 2. This procedure is necessary to obtain the relation between effective packing density, shock pressure, and dissipated work, and to obtain effective profiles that can be directly compared with those given by the macroscale theory. Fluctuations in spatially filtered profiles are relatively small for ϕ¯0=0.84

, but become more pronounced as ϕ¯0

decreases.

Click to view

FIG. 3.Development of a quasi-steady compaction shock corresponding to  = 500 m/s and ϕ0=0.84

: (a) superposed quasi-steady profiles highlighting local fluctuations; and (b) effective quasi-steady pressure profiles at advancing times.

The plan of this paper is as follows. First, the mesoscale and macroscale theories are summarized in Sections II and III, respectively, where their mathematical and physical descriptions of compaction induced dissipation are highlighted. Next, predictions are given in Section IV that illustrate how effective shock profiles vary with ϕ¯0

, , and P¯¯¯

, with emphasis placed on examining the dissipation rate and dissipated work within shocks. Comparisons between profiles and shock rise times computed by the mesoscale and macroscale theories are analyzed and discussed. Last, key conclusions of this study are summarized in Section V.

II. MESOSCALE MODEL

The mesoscale model tracks the temporal and spatial evolution of thermomechanical fields within large ensembles of deformable particles. A detailed exposition of the model equations, contact conditions, and constitutive theory is given by Panchadhar and Gonthier;19 only a brief summary is given here that emphasizes features relevant to this study.

A.  Equations and dissipationEvolution equations for mass, momentum, and energy within particles in their current configuration are given by

ρ̇ =ρu̇ ,
(1) 
ρu¨=σ,
(2) 
ρcvṪ =q+ρr,
(3) 

where is the local density, is the displacement vector, σ is the Cauchy stress tensor, is the deformation induced heating, is the heat flux, is the constant volume specific heat, and ()/x is the spatial gradient operator. The time derivative of a variable is denoted by a “dot” above it. Particles are assumed to be initially stationary and stress free.

Mechanical and thermal contact conditions imposed along particle boundaries Γ are given by

σn=tconΓt,
(4) 
kTn=qf+qconΓt,
(5) 

where is the local unit normal vector to Γ, is the local contact traction vector, is the local frictional heat flux, is the local heat flux needed to impose ideal thermal contact, and is the thermal conductivity. The stress–strain behavior of the explosive material is modeled by a hyperelastic, multiplicative, finite strain constitutive theory with a Von-Mises yield criterion and Perzyna viscoplastic flow rule. Friction is modeled using an Amontons–Coulomb stick-slip theory. Details of the constitutive theory are given in Refs. 6, 19, and 20. Thermomechanical properties for HMX used in the simulations are listed in Table I.

TABLE I.TABLE I.

Click to view

TABLE I.HMX material properties used for the mesoscale simulations.

Because a key focus of this study is to examine dissipation within compaction shocks, the energetics described by the model is briefly highlighted. The total mass-specific deformation power is locally given by P=σ:d/ρ

, where d(Ḟ F1+(Ḟ F1))/2 is the strain rate tensor, F=I+0u is the deformation gradient tensor, is the second order identity tensor, and 0()/X is the Lagrangian gradient operator. By partitioning into elastic and plastic components (i.e., d=de+dp), and partitioning into volumetric and deviatoric components (i.e., σ=PI+τ, where is pressure and τ

is the deviatoric stress tensor), the mass-specific deformation power may be expressed by

P=1ρ(τ:(ddp))+Pρtr(ddp)+1ρ(σ:dp),
(6) 

where ρ=ρ0det(F) and is the initial density in the undeformed configuration. Terms on the right side of this equation represent shear, compression, and plastic work rates, respectively.

Within the context of this model, shear work results in a non-thermal change of stored strain energy, whereas compression and plastic work result in deformation induced heating [represented by in Eq. (3)]. As shown in Refs. 20 and 21, only plastic work is dissipative in that it causes an irreversible production of entropy, with the local dissipation power given by

Tη̇ =1ρ(σ:dp)0.
(7) 

Here, is the mass-specific entropy. Friction work occurring along inter-particle contact surfaces is also dissipative when slip occurs; the frictional dissipation power is locally given by Tη̇ f=tcvr, where vr is the relative slip velocity. Because frictional dissipation is localized along contact surfaces, it does not appreciably contribute to the effective (or bulk) mass-specific dissipated energy. Therefore, dissipated energy by plastic work dominates the effective response, though frictionally induced tangential surface tractions can indirectly contribute to Eq. (7) by enhancing plastic deformation within particles. It is noted that frictional dissipation can, however, produce small mass regions of elevated temperature along inter-particle contact surfaces (hot-spots) that may influence the shock sensitivity of reactive solids. Comprehensive discussions and analysis of hot-spots are given in Refs. 6, 9, and 22.

Particle fracture is not modeled in this study despite evidence indicating that it can cause changes in microstructure during quasi-static compaction of granular HMX.17,23 Because HMX has a low fracture surface energy (approximately  ≈ 0.05 J/m2),24 it is energetically inconsequential compared with the effective dissipated work induced by shocks. For example, if the average particle size of HMX decreases from approximately  = 60 m to  = 10 m due to fracture of spherical particles, then the change in particle surface area is approximately given by ΔA6(1/d21/d1)/ρ0=263m2/kg

. The mass-specific fracture energy is approximately efγfΔA=13J/kg

which is much less than the plastic work associated with shock induced pore collapse (30–100 kJ/kg). Therefore, fracture will minimally influence shock profiles analyzed in this study, though it may substantially alter hot-spot formation at the particle-scale.

The mesoscale model equations are computationally integrated subject to the prescribed boundary conditions using a combined finite and discrete element technique.19 The finite-element method (FEM) is used to numerically integrate the unsteady 2D conservation equations and the viscoplastic flow rule governing inelastic deformation, and the discrete element method (DEM) is used to account for interactions between particles. The DEM utilizes a conservative potential based penalty method in which the normal contact traction between particles is estimated by penalizing their penetration, and frictional tractions are estimated using a penalty regularized Amontons–Coulomb law. The initial particle ensemble is generated using a pseudo-gravity drop method. Particles are individually discretized using constant strain, triangular finite elements, where each particle contains approximately 400–800 elements depending on its size. The combined FEM/DEM algorithm has been verified and is nominally second-order accurate in both space and time.19

B.  Averaging techniqueIt is necessary to average (filter) mesoscale field predictions to obtain effective shock profiles that can be examined and directly compared with those predicted by the macroscale theory. Time averaging is particularly important for low density materials having locally large spatial fluctuations in porosity that can result in large temporal fluctuations in spatial shock structure.

The averaging procedure used in this study is as follows. First, one-dimensional (1D) effective shock profiles are obtained at fixed time by spatially averaging mesoscale fields over suitably chosen windows; a typical window is illustrated in Fig. 4. To estimate the average of a generic variable Φ at an axial location x=x¯

, a window having area is locally positioned at x¯ which represents a rectangle of length and width , and an appropriate average is computed based on all finite elements contained within the window. Mass weighted averages are computed for variables that represent mass specific quantities, such as velocity (mass-specific momentum) and mass-specific power and dissipation rate [given by Eqs. (6) and (7)]. Variables that represent volume specific quantities, such as mass density (volume specific mass) and stress (volume specific energy), are computed using area weighted averages. For a variable Φ, at an axial position x=x¯i

, mass and area weighted averages are, respectively, defined by

Φ¯¯m(x¯i)=AsρΦdAAsρdA,Φ¯¯a(x¯i)=AsΦdAAs,
(8) 

where is the total area of explosive mass contained within the averaging window. Discrete representations of these averages are given by

Φ¯¯m⎛⎝x¯i⎞⎠=Na(i)j=1ρ(j)Φ(j)A(j)Na(i)j=1ρ(j)A(j),
(9) 
Φ¯¯a⎛⎝x¯i⎞⎠=Na(i)j=1Φ(j)A(j)Na(i)j=1A(j),
(10) 

where Φ() is the value of the variable at the centroid of a finite element, () is its current area, () is its current density, and Na(i) is the number of finite elements within the window. The smoothness of average profiles depends on both the number of axial positions for which the running average is computed and the window width , and is most sensitive to the choice of . High-frequency fluctuations in variables result if is too small, whereas excessively smeared profiles result if it is too large. Therefore, it is necessary to establish a rationale for determining an appropriate window size that will provide meaningful comparisons with the macroscale theory, as discussed below.

Click to view

FIG. 4.Schematic of the transverse averaging window used to obtain effective 1-D profiles from mesoscale fields.

Second, spatially averaged quasi-steady profiles obtained at different times for a given simulation were superposed and mathematically averaged to obtain an effective (mean) profile; this process is referred to as ensemble averaging in this study. Approximately 50 quasi-steady profiles were averaged for each simulation. Figure 5 illustrates bounds on the computed variation in quasi-steady profiles for solid pressure and mass-specific dissipation rate [given by Eq. (7)] for ϕ¯0=0.68

and ϕ¯0=0.84

at   = 300 m/s. Corresponding ensemble average profiles are indicated in red in the figure. Larger spatial fluctuations are predicted for materials having lower packing density. The peak dissipation rate within a quasi-steady shock can vary by nearly 100% which may significantly affect the material’s local heating response resulting in stochastic ignition behavior.9 Because this study is concerned with effective dissipation within compaction shocks, profiles presented in this paper represent averages in space and time.

Click to view

FIG. 5.Ensemble averages used to obtain quasi-steady shock pressure and dissipation rate profiles predicted by mesoscale simulations for  = 300 m/s: (a) ϕ¯0=0.68

and (b) ϕ¯0=0.84

.

Figure 6 illustrates how the effective pressure and dissipation rate profiles vary with for ϕ¯0=0.68

and  = 300 m/s, where 10 ≤  ≤ 200 m. The dissipation rate is expressed by ẇ p¯=Tη̇ ¯ using Eq. (7). High frequency fluctuations result for approximately  < 30 m, whereas smeared profiles result for  > 140 m. Pressure profiles are less sensitive to than dissipation rate profiles. Only minor variations in shock width result for  < 50 m, which is easily seen in the dissipation rate profiles which vanish in the equilibrium states surrounding the shock. In this case,  = 50 m is chosen for purposes of comparison with the macroscale theory because it results in a nearly converged shock width (≈0.4 mm) and minimal fluctuations. A similar rationale is used to identify appropriate window sizes for other values of ϕ¯0

and .

Click to view

FIG. 6.Mesoscale predictions for the variation in shock profiles with averaging window width for ϕ¯=0.68

and  = 300 m/s: (a) effective pressure and (b) effective dissipation rate.

Last, the axial length of the computational domain associated with quasi-steady shock propagation was identified using shock position-time and shock speed-time data obtained by post-processing meso-scale velocity fields. For each simulation, the meso-scale velocity field was spatially filtered to obtain effective profiles with increasing time, with shock position defined by the axial location where the effective velocity within the shock is 50% of the prescribed piston speed value . The shock position-time data were then differentiated to obtain shock speed data, and the position-speed data were combined to identify the axial location at which the shock speed was approximately constant. Predictions indicate that quasi-steady shocks develop rather quickly following impact.

III. MACROSCALE MODEL

The macroscale model tracks the evolution of effective thermomechanical fields based on a hydrodynamic theory that accounts for both elastic and inelastic compaction; a detailed exposition of the theory is given by Gonthier.25 The model is based on principles of mixture theory and is representative of a class of models that have been used to describe the dynamic compaction of granular explosives.13,14,26–28 A brief summary of the model is given below that emphasizes features relevant to this study.

A.  Equations and dissipationThe vector form of the unsteady, 1-D governing equations is given in Cartesian coordinates by

qt+fx=s,
(11) 

where the conserved variable, flux, and source vectors are given by

q=[ρsϕ,ρsϕu,ρsϕE,ρsϕ2,ρsϕϕ˜],
(12) 
f=[ρsϕu,ρsϕu2+Psϕ,ρsϕu(E+Psρs),ρsϕ2u,ρsϕϕ˜u],
(13) 
s=[0,0,0,ρsϕ2(1ϕ)μ(Psβ),ρsϕΛ],
(14) 

and Λ is defined by

Λ={1μ˜(f(ϕ)ϕ˜)0iff(ϕ)>ϕ˜otherwise.
(15) 

Vector components in Eqs. (11)–(15) describe the evolution of mass, momentum, total energy, solid volume fraction, and inelastic volume fraction of the material, respectively. Independent variables are time and position . Dependent variables, which implicitly represent average quantities, are: solid phase density ; solid volume fraction ϕ; no-load or inelastic component of solid volume fraction ϕ˜; velocity ; solid phase pressure ; and mass-specific total energy E=e+(u2/2), where e=es+B, is the mass-specific internal energy of the solid phase, and B=(ϕϕ˜)0βρsϕd(ϕϕ˜) is the mass-specific compaction potential energy that depends on the elastic component of solid volume fraction ϕϕ˜. The configurational (or interparticle) stress is empirically determined and implicitly accounts for contact forces between particles that collectively resist compaction. In Eq. (15), f(ϕ) is the compaction yield function. Compaction is driven by an imbalance between and . The parameters and μ˜ in the evolution equations of ϕ and ϕ˜ establish the rates of stress equilibration ( Psβ) and relaxation to the yield surface ( ϕ˜f). The viscosity controls the compaction shock thickness and rise time, as discussed in Section IV.

The system of equations is mathematically closed by prescribing the relations Ps(ρs,es),β(ρs,ϕ)

, and f(ϕ)

. To this end, a Mie–Gruneisen equation of state is used to describe the thermodynamics of HMX

PsPH=Γνs0(eˆseH),
(16) 

where

PH=[ωνs0s(νs0νs)](νs0νs),eH=12[ω(νs0νs)νs0s(νs0νs)]2.

Here, νs=1/ρs is the solid phase specific volume and Γ is the Gruneisen coefficient. This equation of state is compatible with the Hugoniot relation D=ω+su, where is the shock speed, and and are empirically determined constants.29 The configurational stress and compaction yield function are described by

βϕρs=Λ(ϕϕ0)ln[k(ϕϕ0)]k(ϕϕ0),
(17) 

and

f=ϕf+c(ϕϕf),
(18) 

where Λ, , and are constants, and ϕf is the free-pour solid volume fraction of the virgin material. Because compaction experiments indicate fast relaxation to the yield surface, it is assumed that μ˜0 in Eq. (15) resulting in the equilibrium condition ϕ˜=f(ϕ). Parameter values used for HMX are listed in Table II. These constitutive relations are discussed in more detail in Ref. 25.

TABLE II.TABLE II.

Click to view

TABLE II.HMX material properties used for the macroscale compaction theory.

Of particular importance are the effective heating rates induced by compaction shocks. The mass-specific deformation power is given by

P=Ḃ +ė ρ+ė ϕ+ė ϕ˜=βρsϕddt(ϕϕ˜)+Psρs2dρsdt+(Psβ)ρsϕdϕdt+βρsϕdϕ˜dt,
(19) 

where

dϕdt=ϕ(1ϕ)μ(Psβ),
(20) 

and

dϕ˜dt={f(ϕ)dϕdt0iff(ϕ)>ϕ˜otherwise.
(21) 

Here, d()/dt()/t+v() is the Lagrangian derivative. Terms on the left of the second equality in Eq. (19) are, respectively, defined by corresponding terms on the right. Equation (21) is obtained by taking μ˜0 in Eq. (15), with fdf/dϕ=c prescribed by the compaction yield function. In Eq. (19), Ḃ  implicitly accounts for the rate of stored compaction energy at the particle scale due to shear, ė ρ accounts for the compression work rate, and ė c=ė ϕ+ė ϕ˜ accounts for the compaction work rate. The term ė c consists of both a rate-dependent component ė ϕ, which vanishes in the slow compaction limit (i.e., Psβ), and an inelastic component ė ϕ˜. These contributions to the deformation power are analogous to the integrated contributions of those given in Eq. (6) for the mesoscale description. Combining Eqs. (11)–(15) with the second law of thermodynamics gives the strong form of the dissipation inequality

Tsη̇ =ė ϕ+ė ϕ˜0,
(22) 

where η=ηs is the mass-specific granular solid entropy. The first term on the right side of this equation is non-negative, as is the second term because β0 and dϕ˜/dt0 by construction. This thermodynamic description is compatible with a Helmholtz free energy of the form ψ(ρs,Ts,ϕ,ϕ˜)=ψs(ρs,Ts)+B(ϕϕ˜), where is the mass-specific solid phase free energy. Here, is the thermal component of free energy. Rate-dependent compaction is the dominant dissipation mechanism for compaction shocks because ė ϕė ϕ˜. A detailed discussion about the role of ϕ˜ in the mechanics of strain hardened granular material is given in Ref. 25.

The macroscale model is posed as an Initial Value Problem (IVP) that can be integrated to obtain steady spatial shock structures. To this end, Eqs. (11)–(15) are expressed in a right propagating shock-attached frame using the transformation ζ=xDt

and w=uD, where and are position and velocity measured in this frame. Substituting the derivatives t∣∣x=Dζ∣∣t+t∣∣ζ and x∣∣t=ζ∣∣t into these equations, requiring t0, and simplifying the result, gives a system of Ordinary Differential Equations (ODEs) of the form dfdζ=s, with particle velocity given by . Conditions at  = 0 in front of the shock represent the stress-free ambient state: ϕ(0)=ϕ˜(0)=ϕ0,Ps(0)=0,ρs(0)=ρs0,w(0)=D

. The homogeneous ODEs corresponding to the mass, momentum, and energy equations can be directly integrated, with initial conditions applied, and combined to obtain

w=(ρs0ϕ0D)/(ρsϕ),
(23) 
Ps=[(ρs0ϕ0D)2(1ρs0ϕ01ρsϕ)]/ϕ,
(24) 
es+B=12Psϕ(1ρs0ϕ01ρsϕ).
(25) 

Equations (24) and (25), combined with the equation of state, define the Rayleigh line R and Hugoniot curve H for the granular solid. Equation (21) can be directly integrated with ϕ˜0=ϕ0 to give

ϕ˜ϕ0=ϕϕ0f(z)dz.
(26) 

The remaining ODE for ϕ in terms of the wave-coordinate is given by

dϕdζ=ϕ(1ϕ)wμ(Psβ).
(27) 

Compaction shock profiles are obtained by numerically integrating Eq. (27) using an explicit solver contained in the package LSODE, subject to Eqs. (23)–(25), the initial condition ϕ(0)=ϕ0

, and the constitutive relations given by Eqs. (16)–(18). The integration procedure is started by taking the initial value of solid volume fraction to be ϕ=ϕ0+Δϕ, where Δϕ106, which is sufficient to perturb the solution off the ambient equilibrium state with minimal error. Compaction shock equilibrium end states and shock speed can be directly determined by combining the equilibrium conditions Ps=β and ϕ˜=ϕ0 for ϕ0ϕf1(ϕ0), or ϕ˜=f(ϕ) for ϕ>f1(ϕ0), with Eqs. (23)–(25) and Eqs. (16)–(18). The ambient state of the material is unstressed and at standard temperature. The initial solid volume fraction ϕ0 is taken equal to ϕ¯0

in the mesoscale description.

Representative shock profiles for compaction dissipation power, given by Eq. (22), are shown in Fig. 7 for   = 300 m/s. A comparison of these profiles with those predicted by the mesoscale theory is given in the following section. Highlighted in the figure is the dissipation power induced by inelastic compaction ė ϕ˜

and the total dissipation power ė c. Dissipation by inelastic compaction represents a significant fraction of total dissipation (≈15%), particularly for low strength shocks, and its relative significance marginally increases with ϕ0. Most conventional macroscale theories applied to granular explosives do not account for dissipation by inelastic compaction;13,30 consequently, these theories are only formally predictive for strong shocks. Dissipation power increases with ϕ0 for fixed due to an increase in shock impedance of the material which causes faster shocks, whereas spatial shock width is largely insensitive to variations in ϕ0

.

Click to view

FIG. 7.Shock profiles for the (a) inelastic dissipation power ( ė ϕ˜

) and (b) total dissipation power ( ė c) predicted by the compaction theory for  = 300, and ϕ0=0.68

, 0.77, and 0.84.

IV. RESULTS

Results are given in this section that illustrate the variation in compaction shock end states and profiles with shock strength and initial packing density. Emphasis is placed on examining effective dissipation power and dissipated work within shocks because they represent a primary heating mechanism that can trigger combustion. Effective dissipation power for the mesoscale theory is given by ẇ d¯=Tη̇ ¯

and is computed by averaging Eq. (7) using Eq. (9); dissipated work w¯¯¯d

is computed by integrating Eq. (7) in time and averaging the result. Dissipation power for the macroscale theory is directly computed using Eq. (22), and dissipated work is computed by integrating this result in time. Dissipation power also provides a good indication of shock width because it vanishes in both the ambient and shock end state.

Materials used in study have an initial solid volume fraction within the range 0.68ϕ¯s0.84

and a mean particle size of d¯=60μm with a narrow distribution. Figure 8 shows particle ensembles having ϕ¯s=0.68

and 0.84. Significant local fluctuations in packing density exist within low density material that can influence shock propagation and dissipation, as discussed below. Simulations are performed for piston speeds within the range 100 ≤   ≤ 500 m/s because of their relevance to deflagration and shock-to-detonation transition. The mesoscale algorithm used in this study, implemented using FORTRAN 90, was executed on 64, 2.66 GHz Dual Core Xeon 64-bit workstations having 4 GB RAM each. One-dimensional parallelization was accomplished by dividing the spatial domain into transverse sectors containing 50–60 particles each, and by assigning each sector to a specific processor. Message Passing Interface (MPI) libraries were used to exchange information between processors at sector boundaries. A typical run time for a single mesoscale simulation was approximately 3 days.

Click to view

FIG. 8.Representative initial particle ensembles generated by the pseudo-gravity settling algorithm.

A.  Shock end statesFigure 9 compares shock end states in the and UpP¯¯¯

Hugoniot planes, where is the steady shock speed and P¯¯¯=P¯¯¯ϕ¯=Psϕ is the effective pressure behind the shock predicted by both mesoscale simulations and macroscale compaction theory. Effective pressure is a commonly used measure of shock strength. Also shown in the figure are data for granular HMX having ϕ¯0=0.6615 and ϕ¯0=0.99.31 For fixed , both and P¯¯¯ increase with increasing ϕ¯0

due to higher acoustic impedance of the material and enhanced stress bridging between particles within shocks.

Click to view

FIG. 9.Effective shock end states (Hugoniots) predicted by the mesoscale simulations and compaction theory for different initial packing densities ( 0.68ϕ¯s0.84

): (a) plane and (b) plane. Mesoscale simulation predictions are connected by solid lines, compaction theory predictions are connected by dashed lines, and experimental data are connected by dotted lines. Red, black, and green markers correspond to ϕ¯0=0.68

, 0.77, and 0.84, respectively.

Overall, agreement between predictions and data is consistent and reasonable. At higher packing densities, the compaction theory predicts lower shock speeds for approximately  ≤ 250 m/s, and predicts higher shock pressures for   ≥ 250 m/s, compared with the mesoscale simulations. This critical value of piston speed separates two compaction shock regimes: a low pressure regime referred to as strength dominated because material strength is important in preventing complete pore collapse and a high pressure regime referred to as pressure dominated because the pressure is sufficient to eliminate porosity.32 Material strength plays a less significant role than pressure for pressure dominated shocks. Discrepancies in shock speed for strength dominated waves are due to inter-particle friction in the mesoscale simulations which effectively enhances material rigidity under plane-strain confinement and results in enhanced stress transmission and lower dissipation. Because friction is less consequential for pressure dominated shocks, better agreement between Hugoniots is predicted. Discrepancies in shock pressure, which are most pronounced for pressure dominated shocks, are due to differences in the equation of state used by the two descriptions. Structures analyzed in this study largely represent pressure dominated shocks.

B.  Spatial shock profilesFigures 10–12 illustrate the computed variation in spatial shock profile with ϕ¯0

for   = 300 m/s and 500 m/s. These profiles connect the ambient material state to the quasi-equilibrium state behind the shock. Shown in the figures are profiles for effective pressure, dissipation power, and dissipated work, respectively, expressed in a shock-attached reference frame. Mesoscale simulations fundamentally resolve shock width based on interactions between particles whereas it is modeled by the value of viscosity in the compaction theory.

Click to view

FIG. 10.Shock pressure profiles predicted by the mesoscale simulations and compaction theory: (a)  = 300 m/s and (b)   = 500 m/s. Mesoscale simulation profiles are given by solid lines, and compaction theory profiles are given by dashed lines. Red, black, and green lines correspond to ϕ¯0=0.68

, 0.77, and 0.84, respectively.

Click to view

FIG. 11.Dissipation rate profiles predicted by the mesoscale simulations and compaction theory: (a)  = 300 m/s and (b)   = 500 m/s. Mesoscale simulation profiles are given by solid lines, and compaction theory profiles are given by dashed lines. Red, black, and green lines correspond to ϕ¯0=0.68

, 0.77, and 0.84, respectively. Vertical bars on mesoscale profiles indicate the range of fluctuations in effective peak dissipation rate due to local spatial variations in density.

Click to view

FIG. 12.Dissipated work profiles predicted by the mesoscale simulations and compaction theory: (a)  = 300 m/s and (b)   = 500 m/s. Mesoscale simulation profiles are given by solid lines, and compaction theory profiles are given by dashed lines. Red, black, and green lines correspond to ϕ¯0=0.68

, 0.77, and 0.84, respectively.

Several observations are noteworthy regarding shock width and dissipation. First, shock widths predicted by mesoscale simulations and compaction theory are comparable suggesting that the value of viscosity  = 40 kg/(m s) used with the compaction theory is compatible with a microstructure having a nearly uniform particle size of d¯=60μm

. Dissipation rate profiles indicate that shock width is weakly dependent on initial packing density for ranges considered in this study, and that it appreciably decreases with increasing shock strength.

Second, computed dissipation power profiles shown in Fig. 11 qualitatively agree, but quantitative discrepancies exist. Bars are shown on mesoscale profiles indicating the range of fluctuations in effective peak dissipation power within compaction shocks due to local spatial variations in density. The fluctuations are substantial, and are approximately 50% of their ensemble average values. This sensitivity is more pronounced for dissipation power than other integrated quantities because it represents a time rate of change in dissipated work. Peak dissipation power increases with shock strength, and discrepancies between the compaction theory and mesoscale simulations become more pronounced due to differences in their equations of state. Both descriptions indicate that the effective peak dissipation power increases with ϕo

for fixed . This result is important because experiments on the shock loading of low density explosives indicate a reduction in shock sensitivity with increased packing density; therefore, dissipation power by itself is not an accurate measure of shock sensitivity. Dissipated work profiles, obtained by integrating dissipation power across shocks, are shown in Fig. 12. Though peak dissipation power increases with ϕ0

for fixed , dissipated work decreases. An increase in dissipated work increases the effective temperature rise which is compatible with enhanced sensitivity; consequently, it represents a plausible means to characterize shock induced ignition and burn in that it implicitly accounts for hot-spot formation. Moreover, dissipated work can principally account for shock desensitization because it is influenced by shock loading history.33

Shock impedance increases with packing density resulting in faster and stronger shocks for fixed . It is also insightful to examine how dissipated work is affected by packing density for equal strength shocks. To this end, UpP¯¯¯

Hugoniot relations are used to express in terms of P¯¯¯. Figure 13 shows the variation in effective dissipated work with shock pressure and packing density predicted by both the compaction theory and mesoscale simulations. Predictions indicate that the relation between dissipated work and pressure is well-described by the power-law w¯¯¯d=aP¯¯¯ns, where both the prefactor and the exponent n=logw¯¯¯d/logP¯¯¯s depend on ϕ¯0. The value of is predicted to monotonically increase from  = 1.13 for ϕ¯0=0.68 to  = 1.20 for ϕ¯0=0.84 which is indicative of enhanced pressure sensitivity. A consequence of this prediction is that differences in shock dissipation with packing density decrease, albeit slowly, as shock pressure increases. It is interesting that the measured variation in run distance to detonation with input shock pressure (referred to as a Pop-plot curve) for low density HMX having comparable packing densities to those studied here indicates a similar trend. Measurements show that the run distance lP¯¯¯ms, where the constant exponent m=logl/logP¯¯¯s slightly increases with packing density.34 If it is assumed that lw¯¯¯1d to leading-order, then nm

. Though speculative due to complex hot-spot formation, chemistry, and multiphase transport associated with shock initiation, the similar dependency of shock dissipation and Pop-plot curves on packing density warrants additional study.

Click to view

FIG. 13.Variation in dissipated work with shock pressure predicted by the mesoscale simulations and compaction theory. Mesoscale simulation data are given by square markers, and compaction theory data are given by circular markers. The dashed lines represent best fit power-laws. Red, black, and green markers correspond to ϕ¯0=0.68

, 0.77, and 0.84, respectively.

C.  Shock rise timesGas gun experiments often measure stress and velocity histories following particles (i.e., Lagrangian coordinates) from which shock rise times can be determined.15 Rise time is important because it establishes the time scale of dissipative heating induced by pore collapse within shocks. Data indicate that rise time not only depends on shock strength but also on material microstructure, particularly particle size.

Figure 14 shows particle velocity histories computed by the compaction theory and mesoscale simulations for  = 300 m/s and 500 m/s, respectively; time shifts between histories corresponding to different values of ϕ¯0

are not significant and are included for ease of presentation. For consistency with experiments, rise time is estimated as the time between the 5% and 95% of maximum particle velocity levels.15

Click to view

FIG. 14.Particle velocity histories predicted by the mesoscale simulations and compaction theory: (a)  = 300 m/s and (b)   = 500 m/s. Mesoscale simulation profiles are given by solid lines, and compaction theory profiles are given by dashed lines. Red, black, and green lines correspond to ϕ¯0=0.68

, 0.77, and 0.84, respectively. The horizontal offsets in the plots are used for clarity and are inconsequential.

The variation in computed shock rise time with is shown in Fig. 15 for different values of ϕ¯0

. Also shown in the figure are data reported by Sheffield and co-workers15 for low density HMX. Data are shown that highlight variations due to packing density and particle size: here, “coarse” HMX refers to material having a mean particle size of ≈120 m; and “fine” HMX refers to material having a mean size of ≈10–15 m but also having some larger particles of ≈50 m. No information about the particle size distribution is provided in Ref. 15. The data indicate that rise time decreases with increasing and is more sensitive to particle size than packing density. Measured rise times for fine HMX are 0.1 s or less for the entire range of , whereas rise time increases with decreasing piston speed for coarse HMX reaching a value of ≈0.5 s for ≈ 300 m/s. Because the mesoscale simulations performed in this study are for ensembles having a mean particle size of 60 m with a narrow distribution, they are more representative of coarse HMX than fine HMX. Both the compaction theory and mesoscale simulations give rise time estimates that qualitatively agree with data for coarse HMX, though greater sensitivity to ϕ¯0

is predicted than indicated by the data. Because particle size influences rise time, quantitative discrepancies between computed estimates and the data may be partly due to variations in particle size distribution; a systematic study of such effects represents an ongoing topic of our research.

Click to view

FIG. 15.Variation in shock rise time with predicted by the mesoscale simulations and compaction theory. Also shown in the figure are data reported by Sheffield .15 Mesoscale simulation data are given by square markers and are connected by solid lines, and compaction theory data are given by circular markers and are connected by dashed lines. Red, black, and green markers correspond to ϕ¯0=0.68

, 0.77, and 0.84, respectively.

It is insightful to examine how rise time is affected by packing density for equal strength shocks. Shown in Fig. 16 are the computed estimates and data given in Fig. 15 expressed as a function of effective shock pressure. To this end, UpP¯¯¯

Hugoniot data given in Fig. 9(a) for ϕ¯0=0.65 were used to express in Fig. 16 in terms of P¯¯¯. The compaction theory and mesoscale simulations give similar results that exhibit a power-law relation τr=bP¯¯¯z, where the exponent ≈ 1 and the prefactor has a weak dependence on ϕ¯0

. Agreement in computed rise times suggests that a constant value of viscosity may be reasonably assumed if the material has a narrow particle size distribution.

Click to view

FIG. 16.Variation in shock rise time with shock pressure predicted by the mesoscale simulations and compaction theory. Estimates given by Eq. (28) were obtained using  = 40 kg/(m s) for coarse particles and   = 10 kg/(m s) for fine particles. Mesoscale simulation data are given by square markers and are connected by solid lines, and compaction theory data are given by circular markers and are connected by dashed lines. Rise time estimates are given by solid lines. Red, black, and green markers correspond to ϕ¯0=0.68

, 0.77, and 0.84, respectively.

A simple estimate for rise time can be obtained from Eq. (20). Because most shocks considered in this study are pressure dominated resulting in the elimination of porosity, it is reasonable to assume that Psβ

. If it is further assumed that Ps

constant, which is compatible with a strong shock approximation, then Eq. (20) gives

τ=(μPs)ϕϕ¯0dϕϕ(1ϕ)=(μPs)ln(11/ϕ¯011/ϕ),
(28) 

where is time. This expression implies that ϕ1 asymptotically as τ. An estimate for rise time can be obtained by evaluating τ(ϕ=0.99). This estimate indicates that τμP1s and that it weakly depends on ϕ¯0 in agreement with data. The viscosity may be assumed to monotonically decrease with particle size in that smaller particles will cause faster stress equilibration rates due to higher frequency pressure reverberations within and between particles.

Also shown in Fig. 16 are estimates based on Eq. (28) for ϕ¯0=0.68

, 0.77, and 0.84. Two values of viscosity are used:  = 40 kg/(m s), the value used throughout this study, which is representative of coarse HMX, and  = 10 kg/(m s) which is representative of fine HMX. Equation (28) properly characterizes the shock pressure dependence z1

obtained by mesoscale simulations, but is conservative because it ignores the configurational stress that implicitly accounts for material strength resulting in faster rise times. Scatter in the rise time data for fine HMX may again be attributable to the presence of large (∼50 m) particles embedded within the material. In a mixture of fine and large particles, it is possible that fine particles control rise time for weak and spatially dispersed shocks, and that large particles (even in relatively low concentrations) affect rise time for strong and thin shocks causing it to approach that of coarse HMX. This possibility may be computationally examined by mesoscale simulations of materials having prescribed particle size distributions, and the appropriate dependence of on microstructure can be established. Departures from power-law behavior may also result as shock strength increases due to the onset of flow instabilities along pore surfaces that result in significant hydrodynamic jetting and vorticity.

V. CONCLUSION

Mesoscale simulations were performed on large ensembles of deformable particles to computationally characterize how dissipation within quasi-steady compaction shocks is affected by shock strength and packing density. The particles, which are representative of the explosive HMX, had an average size of 60 m with a narrow distribution to isolate effects of packing density from those due to more complex distributions. Effective shock profiles, obtained by averaging mesoscale fields over space and time, explicitly account for resolved particle scale interactions within shocks. These profiles were compared with predictions given by an established macroscale theory that describes shock compaction in terms of state variables that are conventionally interpreted as the average manifestation of particle scale fields. Emphasis was placed on examining dissipation power, dissipated work, and shock rise time because of their relevance to the ignition and burn of granular reactive solids.

Effective shock end states and spatial profiles computed by mesoscale simulations compare favorably with those given by the compaction theory. Shock speed and pressure increase with density for fixed piston speed due to higher acoustic impedance of the material and enhanced stress bridging between particles within shocks. Shock width is relatively insensitive to initial density, but significantly decreases with increasing shock strength. Increasing density is shown to increase the dissipation rate within shocks but decrease the integrated dissipated work over shock profiles, which is indicative of reduced sensitivity. Consequently, dissipation rate by itself cannot properly account for observed phenomena, such as shock induced ignition, burn, and desensitization, whereas dissipated work represents a plausible history-dependent thermodynamic quantity that can principally account for these effects. The development of dissipation-dependent macroscale Ignition and Burn (I&B) models for low density granular explosives is an ongoing topic of our research.35 Such models are analogous to entropy-based I&B models used to describe shock initiation of plastic-bonded explosives.36

Quantitative discrepancies between the mesoscale simulations and compaction theory are largely due to differences in their equations of state and are more appreciable for pressure dominated shocks that are sufficiently strong to eliminate porosity. Despite such discrepancies, both descriptions indicate that the relation between dissipated work and pressure is well-described by a power-law having a density-dependent pressure prefactor and exponent. The value of the exponent is predicted to monotonically increase with density which is indicative of enhanced pressure sensitivity. A similar pressure sensitivity has been observed for the shock induced run distance to detonation for low density HMX. Though possibly fortuitous, these analogous sensitivities suggest that run distance to detonation may also be controlled by shock dissipation to leading-order; this issue warrants additional study.

Mesoscale simulations and the compaction theory indicate that the relation between shock rise time and pressure is well-described by a power law having an inverse pressure dependence with a prefactor that is weakly dependent on initial density and an exponent that has an approximate value of unity. This relation is confirmed by a simple analytical estimate for rise time based on the compaction theory. Comparisons indicate that the assumption of constant viscosity in the compaction theory is reasonable for a material having a nearly uniform particle size distribution. Computed rise times compare favorably with data for coarse HMX (mean particle size ∼120 m), whereas data for fine HMX (mean particle size ∼10 m, with some embedded particles of ∼50 m) indicate faster rise times, particularly for low pressure shocks.15 The fine HMX data also indicate a lower sensitivity of rise time to pressure which may be due, in part, to the bimodal size distribution.

This study demonstrates how mesoscale simulations may be used to examine the effect of microstructure on the thermomechanics of compaction shocks in granular reactive solids. These simulations may also be used to assess the validity of assumptions imposed by macroscale theories and to provide rigorous interpretations of their predictions. In addition to the development of I&B models, another focus of the authors’ ongoing research is on characterizing how variations in particle size distribution influence compaction shock dissipation and rise time.

Fuente: http://dx.doi.org