Discrete modeling of ultra-high-performance concrete with...

20
Discrete modeling of ultra-high-performance concrete with application to projectile penetration Jovanca Smith a , Gianluca Cusatis b, * , Daniele Pelessone c , Eric Landis d , James ODaniel e , James Baylot e a Department of Civil and Environmental Engineering, Northwestern University, Tech Building Room A323, Evanston, IL 60208, USA b Department of Civil and Environmental Engineering, Northwestern University, Tech Building Room A125, Evanston, IL 60208, USA c ES3, San Diego, CA 92101, USA d Department of Civil and Environmental Engineering, University of Maine, 303 Boardman Hall, Orono, ME 04469, USA e Geotechnical and Structures Laboratory, U.S. Army Engineer Research and Development Center, Vicksburg, MS 39180, USA article info Article history: Received 21 April 2013 Received in revised form 15 October 2013 Accepted 22 October 2013 Available online 1 November 2013 Keywords: Ultra-high-performance concrete Lattice Discrete Particle Model Fracture Fragmentation Penetration abstract In this paper, the Lattice Discrete Particle Model for ber reinforced concrete (LDPM-F) is calibrated and validated with reference to a new high-strength, ultra-high-performance concrete (UHPC) named CORTUF and applied to the simulation of projectile penetration. LDPM-F is a three-dimensional model that simulates concrete at the length scale of coarse aggregate pieces (meso-scale) through the adoption of a discrete modeling framework for both ber reinforcement and embedding matrix heterogeneity. In this study, CORTUF parameter identication is performed using basic laboratory ber pull-out experi- ments and experiments relevant to a CORTUF mix without ber reinforcement. Extensive comparisons of the numerical predictions against experimental data that were not used during the calibration phase (relevant to both plain CORTUF and CORTUF with ber reinforcement) are used to validate the calibrated model and to provide solid evidence of its predictive capabilities. Simulations are then carried out to investigate the behavior of protective CORTUF panels subjected to projectile penetration, and the numerical results are discussed with reference to available experimental data obtained at the Engi- neering Research and Development Center (ERDC). Ó 2013 Elsevier Ltd. All rights reserved. 1. Introduction The design exibility and natural durability of concrete help to make it the most used man-made material in the world. As opposed to other building materials, such as brick and stone, concrete has revolutionized previous construction methods with its intrinsic versatility and high compressive strength. Within the last century, a tremendous effort has been made to increase concrete strength. Over time, each new generation of the material became stronger and more durable, birthing the title we know today as ultra-high- performance concrete (UHPC). The material is sought for many uses including its adoption in high-rise structures, long-span bridges, offshore structures, and structural rehabilitation [1,2]. There are various methods used today to develop higher compressive strength in cementitious composites. One method is strengthening the interfacial transition zone (ITZ), the weak area surrounding the aggregate particle, which is often accomplished through a reduction of the aggregate size; this enhances homoge- neity and reduces stress concentrations between the aggregate and mortar under loading. Another method to enhance the ITZ me- chanical properties is reducing the water-to-binder ratio [3]. This process can be counter productive if not balanced, since the high rates of hydration often results in autogeneous shrinkage and can lead to premature cracking during the setting phase [4]. Many high- performance concretes use silica fume, which acts like a microller and can also react with calcium hydroxide to increase the nal strength of the composite [5]. Other methods reported in the literature that enhance the compressive strength of concrete are particle-size gradation for optimum packing of aggregates [6] and heat treatment [7e10]. * Corresponding author. McCormick School of Engineering and Applied Science, NorthwesternUniversity, Tech Building Room A125, 2145 N Sheridan Rd, Evanston, IL 60208-3109, USA. Tel.: þ1 847 491 4027. E-mail addresses: [email protected] (J. Smith), g-cusatis@ northwestern.edu (G. Cusatis), [email protected] (D. Pelessone), landis@maine. edu (E. Landis), James.L.O[email protected] (J. ODaniel), James.T.Baylot@ usace.army.mil (J. Baylot). Contents lists available at ScienceDirect International Journal of Impact Engineering journal homepage: www.elsevier.com/locate/ijimpeng 0734-743X/$ e see front matter Ó 2013 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.ijimpeng.2013.10.008 International Journal of Impact Engineering 65 (2014) 13e32

Transcript of Discrete modeling of ultra-high-performance concrete with...

Page 1: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

lable at ScienceDirect

International Journal of Impact Engineering 65 (2014) 13e32

Contents lists avai

International Journal of Impact Engineering

journal homepage: www.elsevier .com/locate/ i j impeng

Discrete modeling of ultra-high-performance concretewith application to projectile penetration

Jovanca Smith a, Gianluca Cusatis b,*, Daniele Pelessone c, Eric Landis d, James O’Daniel e,James Baylot e

aDepartment of Civil and Environmental Engineering, Northwestern University, Tech Building Room A323, Evanston, IL 60208, USAbDepartment of Civil and Environmental Engineering, Northwestern University, Tech Building Room A125, Evanston, IL 60208, USAc ES3, San Diego, CA 92101, USAdDepartment of Civil and Environmental Engineering, University of Maine, 303 Boardman Hall, Orono, ME 04469, USAeGeotechnical and Structures Laboratory, U.S. Army Engineer Research and Development Center, Vicksburg, MS 39180, USA

a r t i c l e i n f o

Article history:Received 21 April 2013Received in revised form15 October 2013Accepted 22 October 2013Available online 1 November 2013

Keywords:Ultra-high-performance concreteLattice Discrete Particle ModelFractureFragmentationPenetration

* Corresponding author. McCormick School of EngiNorthwestern University, Tech Building Room A125, 2IL 60208-3109, USA. Tel.: þ1 847 491 4027.

E-mail addresses: [email protected] (G. Cusatis), [email protected] (Dedu (E. Landis), James.L.O’[email protected] (Jusace.army.mil (J. Baylot).

0734-743X/$ e see front matter � 2013 Elsevier Ltd.http://dx.doi.org/10.1016/j.ijimpeng.2013.10.008

a b s t r a c t

In this paper, the Lattice Discrete Particle Model for fiber reinforced concrete (LDPM-F) is calibrated andvalidated with reference to a new high-strength, ultra-high-performance concrete (UHPC) namedCORTUF and applied to the simulation of projectile penetration. LDPM-F is a three-dimensional modelthat simulates concrete at the length scale of coarse aggregate pieces (meso-scale) through the adoptionof a discrete modeling framework for both fiber reinforcement and embedding matrix heterogeneity. Inthis study, CORTUF parameter identification is performed using basic laboratory fiber pull-out experi-ments and experiments relevant to a CORTUF mix without fiber reinforcement. Extensive comparisons ofthe numerical predictions against experimental data that were not used during the calibration phase(relevant to both plain CORTUF and CORTUF with fiber reinforcement) are used to validate the calibratedmodel and to provide solid evidence of its predictive capabilities. Simulations are then carried out toinvestigate the behavior of protective CORTUF panels subjected to projectile penetration, and thenumerical results are discussed with reference to available experimental data obtained at the Engi-neering Research and Development Center (ERDC).

� 2013 Elsevier Ltd. All rights reserved.

1. Introduction

The design flexibility and natural durability of concrete help tomake it themost usedman-madematerial in theworld. As opposedto other building materials, such as brick and stone, concrete hasrevolutionized previous construction methods with its intrinsicversatility and high compressive strength.Within the last century, atremendous effort has been made to increase concrete strength.Over time, each new generation of the material became strongerand more durable, birthing the title we know today as ultra-high-performance concrete (UHPC). The material is sought for many

neering and Applied Science,145 N Sheridan Rd, Evanston,

du (J. Smith), g-cusatis@. Pelessone), landis@maine.. O’Daniel), James.T.Baylot@

All rights reserved.

uses including its adoption in high-rise structures, long-spanbridges, offshore structures, and structural rehabilitation [1,2].

There are various methods used today to develop highercompressive strength in cementitious composites. One method isstrengthening the interfacial transition zone (ITZ), the weak areasurrounding the aggregate particle, which is often accomplishedthrough a reduction of the aggregate size; this enhances homoge-neity and reduces stress concentrations between the aggregate andmortar under loading. Another method to enhance the ITZ me-chanical properties is reducing the water-to-binder ratio [3]. Thisprocess can be counter productive if not balanced, since the highrates of hydration often results in autogeneous shrinkage and canlead to premature cracking during the setting phase [4]. Many high-performance concretes use silica fume, which acts like a microfillerand can also react with calcium hydroxide to increase the finalstrength of the composite [5]. Other methods reported in theliterature that enhance the compressive strength of concrete areparticle-size gradation for optimum packing of aggregates [6] andheat treatment [7e10].

Page 2: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e3214

Although there are many methods focused on increasing thecompressive strength of concrete, techniques to increase tensileresistance and toughness are still in their infancy. One methodtypically used for this purpose is the addition of fibers in the matrix[5,11e14]. Multi-scale fiber reinforcing, in which fibers of variouslengths and cross sectional sizes are used in the same mix, has alsobeen explored in recent times [15]. When fiber reinforcing is used,the type and amount of fibers need to be optimized in order tomaximize tensile strength and toughness while preserving work-ability and efficiency in terms of cost-to-performance ratio.

Enhancement of material toughness is particularly important inthe case of the UHPC materials that are being tested to determinetheir efficiency against blast impact and penetration and for theiruse in particular areas susceptible to man-made and natural haz-ards [16,17]. As it pertains to projectile impact testing, manyresearchers found the penetration depth to increase withincreasing impact velocities, and to decrease with increasingcompressive strength. The perforation ballistic limit (the minimumvelocity at which the bullet completely perforates the specimen) istypically reported to increase with higher compressive strength[18]. Aggregate size also plays an important role in material impactresponse. Stronger and larger sized aggregate particles improveimpact resistance as long as enough workability is maintained[3,12,19]. For UHPC without fiber reinforcing, penetration eventscause dynamic propagation of many micro-cracks and often lead toshattering of the target. On the contrary, inclusion of fibers confinescracks in the crater region in both front and rear faces and creates amore localized damage around the crater. However, for increasingfiber volume fraction, a saturation limit seems to exist after whichno significant change is observed in both crater area and penetra-tion depth [20e22].

When dealing with the effect of dynamic events on concretestructures it is also important to consider the strain-rate depen-dence of concrete response. Typical experimental observations onregular strength concrete report an increase of macroscopicmechanical properties such as, Young’s modulus, compressivestrength, tensile strength, and fracture energy, for increasing strainrate (see, among many others, [23e25, 27e32]).

Evidences of strain-rate dependence of UHPC behavior are morescarce and contradictory compared to regular strength concrete.A few studies have been performed on the effect of strain rate onUHPCs that investigated the effect of fiber reinforcement andincreasing strain-rate under tensile and compressive loadings. Someresearchers found UHPCs with fibers to be less sensitive to strain-rate effects than similar strength unreinforced material [33,34].Some other researchers [35] found no rate sensitivity for concretereinforced with hooked steel fibers. Caverzan et al. [36] investigatedthe effect of temperature on strain rates of high-performance fiber-reinforced concrete (HPFRC) and found an increase in strength forincreased strain rate up to a rate of 150 s�1 for temperatures up to200 �C, while material strength decreased for strain-rates above150 s�1 for higher temperatures due to fiber failure.

In most cases, rate effect is quantified through the DynamicIncrease Factor (DIF) defined as the ratio between the dynamicproperty of interest and the associated quasi-static value. Theevolution of DIF with strain-rate is often approximated through abilinear curve inwhich the second linear branch, starting for strain-rates around 1e10 s�1, has a much steeper slope than the firstbranch. In many cases, the DIF curves are obtained by interpolatingbetween experimental data showing huge scatter, especially forhigh strain-rates. Furthermore, it is common modeling procedureto equip concrete constitutive equations with rate dependence byusing the DIF curve as a multiplicative factor for various modelparameters. As some authors have shown recently [37e39], thisapproach is not correct in general because it does not distinguish

between “intrinsic” phenomena, which should be included in theconstitutive equations, and “apparent” phenomena, which arestructural features of the response and should or should not beincluded in the constitutive equations depending on the spatial andtemporal resolution of the adopted numerical model.

At least three intrinsic rate-dependent phenomena can beidentified to affect concrete mechanical behavior: 1) creep, 2)Arrhenius-type behavior of fracture processes, and 3) contributionto the load carrying capacity of capillary and adsorbed water.

Concrete creep, the increase in time of deformation underconstant applied load, is a complex phenomenon whose funda-mental mechanism resides in the cement nano-structure, and it ishypothesized to be the result of shear slips among Calcium SilicateHydrate (CSH) platelets [40,41]. Since the nano-structure of cementevolves as result of cement hydration and other chemical reactions,creep behavior changes over time: in general, young concrete fea-tures a much more pronounced viscous behavior compared to oldconcrete. In addition, concrete creep is affected bymoisture contentand temperature and their time rates [42]. In particular, the lowerthe relative humidity the less concrete creep deformation is typi-cally observed. Wittmann [43] obtained an 87% decrease in creepdeformation reducing the relative humidity from 100% to 10%. Fromthe mathematical point of view, concrete creep is typically simu-lated by aging, temperature, and relative humidity dependentvisco-elastic constitutive equations; this naturally leads to adependence of the stress state on strain-rates.

However, creep cannot fully explain the fracturing behavior ofconcrete under dynamic loading. For example, creep cannotsimulate experimental data on three point bending specimensshowing almost instantaneous rehardening when subjected to asudden increase of loading rate in the softening regime [44]. Tomodel this phenomenon, Ba�zant analyzed the fracture processwithin the classical theory of thermally activated processes, and byrepresenting the frequency of bond rupture at the nano-scale withan Arrhenius equation, he derived a dependence of the cohesivestress (stress across a propagating cohesive crack) upon crackopening rate. This formulation was successfully used inRefs. [37,45,46] to simulate the dynamic behavior of concrete underlow to moderate strain rates.

Furthermore, concrete is a porous material featuring a complexinternal structure characterized by pore sizes spanning variousorder of magnitudesdfrom few nanometers to several millimeters.Depending upon the level of relative humidity, water is present inconcrete pores in various forms (capillary, absorbed, or hindered)and, according to classical poro-mechanics concepts, it cancertainly contribute to the overall carrying capacity. However, un-der quasi-static loading conditions the amount of load that porewater can carry is negligible because water is much morecompressible than the solid skeleton. More complicated scenarioarises under high confinement quasi-static triaxial loading becausepore collapse occurs and water is squeezed out, even leading to acertain decrease in carrying capacity due to lubrication effects [47].On the contrary, at high strain rates, water can contribute signifi-cantly to concrete strength. As far as tensile loading is concerned,some studies [48] have tried to explain the effect of water at highstrain rates through the so-called Steffan effect [49] according towhich the force needed to pull apart two parallel rigid platescontaining a Newtonian fluid in between is proportional to therelative velocity of the two plates. This approach, althoughchampioned by some authors [50,51], has been criticized becauseits application to absorbed and hindered water (forming themajority of water in real concrete at relative humidity lower that60e70%) is questionable. For triaxial compressive loading at highstrain rates, water has a significant effect because, as pores collapse,water does not have enough time to be released leading to a

Page 3: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e32 15

significant build up of pore pressure. Forquin and coworkers [47]report an increase of 55% of the hydrostatic stress measured insaturated concrete samples at a strain of 0.08 and a strain-rate of120 s�1 compared to dry samples.

While the intrinsic phenomena discussed above should (if rele-vant) be included in the formulation of appropriate constitutiveequations for concrete, apparent phenomena should be filtered outfrom the experimental evidences and not considered in theconstitutive equations unless they occur at a length scale smallerthan the model resolution. The most important, and often mis-interpreted, phenomenon leading to an apparent strain-ratedependence of concrete behavior is due to inertia effects. Duringdynamic events, stresswaves travel through concrete and their pathas well as their magnitude are affected by the presence of externalboundaries aswell as by concrete internal heterogeneitydthis leadsto “overstresses” and fictitious confinement effects resulting inhigher peak stresses erroneously interpreted as higher materialstrength. This aspect was analyzed in detail by the second author inan earlier publication [37]. As long as inertia is correctly accountedfor in the interpretation of experimental results and in the numer-ical simulations, no inertia effects should be included in concreteconstitutive equations, especially when they are formulated at finelength scales and capture material heterogeneity.

Another phenomenon associated with dynamic loading is thechange in crack patterns. On one extreme, under quasi-staticloading, a concrete bar in direct tension always fails with thepropagation of one singlemacro-crack; on the other extreme, underultra-high strain rates in which inertia effects are dominant, thesame bar features a much more distributed crack pattern andcomminution/pulverization might occur [52]. Also the microscopiccharacteristics [50,53] of crack patterns change with increasingstrain-rate. For example, it has been observed that at high strain-rates, cracks initiate at the same time at multiple locations and intheir subsequent propagation follow shorter paths of higher resis-tance (for example fracturing aggregate pieces) as opposed to moretortuous paths of least resistance (for example through the ITZ onthe aggregate boundaries). Also, it is well reported for concrete [54]and othermaterials [55] the phenomenon of crack branching underdynamic crack propagation. Overall, the change in crack patternsleads to higher energy dissipation resulting in apparent strengthand toughness increase for increasing strain rates.

2. Numerical modeling for UHPC materials and the LDPM-Fformulation

Optimization of cementitious composites, with material prop-erties tailored to a specific application, has been traditionally pur-sued through experimental investigations that are quite expensiveand are not sustainable for investigation of a large number ofdifferent variables. A less expensive alternative is to optimize thematerial through numerical simulations.

Typical constitutive modeling for mode I fracture in concreteincludes: (1) the cohesive crack model (CCM) [56,57], which simu-lates damage processes through the occurrence of displacementdiscontinuities, across which the ability to transfer stresses de-creases as function of the discontinuity jump (crack opening); and(2) the crack band model [58,59], where cracks are assumed to bedistributed over a certain band whose width is deemed a materialproperty.

For more general triaxial stress states, strain-softening tensorialconstitutive equationshavebeenproposed in the literature [60e63],the most notable of which are the ones based on the microplanetheory [64e68]. However, these type of models, suffer from patho-logical mesh sensitivity and spurious energy dissipation uponmeshrefinement [69e71]. This problem has been somewhat addressed

through the formulation of non-local models with either gradientformulations or integral formulations, which allow simulation ofmicro-crack interaction and size effect [72e76].

A different class of models providing a better representation ofthe actual cracking phenomena occurring in the internal structureof the material is the one consisting of discrete approaches (latticeand particle models) in which concrete is discretized “a priori”according to its meso-structure. Earlier attempts in this directionare reported in Refs. [77e82], while more recent developments arediscussed in Refs. [83,84].

Cusatis and coworkers formulated, calibrated, and validated acomprehensive discrete model for concrete, entitled LatticeDiscrete Particle Model (LDPM) [85e87] and extended it to includethe effect of fiber reinforcing [88,89]. The resulting model, labeledLDPM-F, demonstrated an unprecedented ability to model fiber-reinforced concrete and to correctly predict the tougheningmechanisms due to the effect of fibers. For this reason, LDPM-F wasselected in this study as the best candidate for the simulation ofUHPC concrete materials in general, and CORTUF in particular.

LDPM represents concrete as a two-phase material composed ofaggregate pieces held together by an embedding cementitiousmatrix. In the numerical representation, polyhedral shaped parti-cles (Fig. 1a) simulate individual aggregate pieces (approximated asspheres) and their surrounding matrix. Geometry and topology ofthe model are obtained through: (1) random generation andplacement of aggregate pieces; (2) delaunay tetrahedralization ofaggregate centers; and (3) dual tessellation defining particle in-terfaces (triangular facets in Fig. 1a) assumed to be the location ofpotential cracks in the meso-structure. Fiber reinforcing is includedin such away to retain the discrete character of LDPM. All the fibersare randomly distributed in the material volume, and each fiber isassociated with all intersecting meso-scale facets (Fig. 1b). Furtherdetails of the adopted fiber generation algorithm are discussed laterin this paper.

LDPM describes meso-structure deformation through theadoption of rigid-body kinematics [85] for each individual particle.Based on this assumption and for given displacements and rota-tions of the particles associated with a given facet, the relativedisplacement EuCF at the centroid of the facet can be used to definethe following measures of strain.

eN ¼ nTEuCF[

; eM ¼ mTEuCF[

; eL ¼ lTEuCF[

(1)

where [ is the length of the straight line connecting two adjacentparticle centers, and n, m, l are unit vectors defining a local systemof reference such that n is orthogonal to the facet, and m, l aremutually orthogonal as well as tangential to the facet. Cusatis andShauffert [90] showed that the strain definitions in Eq. (1) areconsistent with the definition of strain in classical continuummechanics. These strains can then be used to compute the LDPMfacet stress tractions, tc ¼ tcNnþ tcMmþ tcLl, in accordance with theLDPM constitutive law (Appendix A1). In addition, when a tensilefacet deforms beyond the tensile elastic limit, the meso-scale crackopening can be calculated as w ¼ wNn þ wMm þ wLl, where

wN ¼ [�eN � tcN=EN

�; wM ¼ [

�eM � tcM=ET

�;

wL ¼ [�eL � tcL=ET

� (2)

and EN, ET are the elastic normal and tangential LDPM stiffnesses,respectively (see Appendix A1).

Furthermore, at the facet level, the formulation assumes a par-allel coupling between the fibers and the surrounding concretematrix. In this case, the total stress traction on each facet can becalculated as

Page 4: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

0 0.02 0.04 0.060

50

100

150

200

250

300

350

Stre

ss (

MPa

)

Strain

Coarse

Fine

0 0.1 0.2 0.30

0.5

1

1.5

2x 10

4

Forc

e (N

)

Displacement (mm)

Coarse

Fine

Facet

Ls

Lf

nf

n

Fiber

da = 4 mm da = 0.6 mm

20 mm 20 mm

a) b)

d) e)

Tetrahedron Edge

Particle

Triangular Facet

g) h)

c)

f)

i)

c

c c p

p

101

100

0

0.2

0.4

0.6

0.8

1

CD

F fo

r E

dge

Len

gth

Edge Length (mm)

Fine

Coarse

Fig. 1. CORTUF-Plain internal structure at the sub-millimeter scale and particle size comparison. a) Polyhedral cell, b) Fiber-facet interaction, c) Bond splitting mechanism, d)Undamaged, and e) Damaged cylindrical specimen for Split test, f) Coarse and fine scale aggregate distribution, g) Crack distribution length comparison of coarse- and fine-scaledaggregates, h) Uniaxial Compression; and i) Split test comparison for coarse and fine aggregate size distribution on 20 mm cube specimen.

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e3216

t ¼ tcðeN; eM; eLÞ þ1Ac

Xf˛Ac

Pf ðwN;wM;wLÞ (3)

where Ac is the facet area, and Pf is the crack-bridging force for eachfiber crossing the given LDPM facet.

As formally noted in Eq. (3), the fiber contribution, Pf, to the loadcarrying capacity is assumed to be a function of the meso-scalecrack opening, w, and is formulated according to a precise micro-mechanical analysis of failure mechanisms affecting the fiberresistance to pull-out. These mechanisms include debonding, fric-tion pull-out resistance, spalling, change in fiber orientation duringpull-out, snubbing, and fiber rupture. The mathematical descrip-tion involved is well documented in the literature [91], discussed indetail by Ref. [88], and summarized in Appendix A2.

The LDPM formulation also simulates rate effect of concretestrength and toughness at the meso-scale by including creep andArrhenius-type behavior. Details of themathematical description ofthe two mechanical phenomena and their amalgamation withinthe LDPM framework are reported in Refs. [46,92]. By virtue of this

formulation, LDPM (and its predecessor, the CSL model [37]) hasbeen applied successfully to the simulation of a variety of experi-mental data relevant to the dynamic response of regular strengthconcrete [46,92]. In addition, LDPM has been recently adopted forthe simulation of penetration, perforation, blast, and impact ofregular strength concrete and it has shown excellent predictivecapabilities [93e95].

3. LDPM-F modeling of CORTUF concrete

This section presents the numerical modeling of CORTUF, anUHPC recently developed and characterized at the Geotechnicaland Structures Laboratory (GSL), United States Army EngineerResearch and Development Center (ERDC) [98,99].

3.1. Mix-design, basic material properties, and internal structure

CORTUF was designed both with fiber reinforcement (CORTUF-Fiber) and without fiber reinforcement (CORTUF-Plain), and its

Page 5: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

Table 1Mixture composition for CORTUF.

Material Proportion by weight

Cement 1Sand 0.967Silica flour 0.277Silica fume 0.389Superplasticizer 0.0171Water (tap) 0.208Steel fibers (CORTUF-Fiber) 0.31

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e32 17

average composition is summarized in Table 1. The adopted sandhas a maximum grain size of 0.6 mm. In addition, the basic CORTUFmix is characterized by cement content, c, equal to 794 kg/m3,which can be back calculated from the measured wet density(2442.5 kg/m3) assuming 7.5% entrained air.

It is worth observing here that for the sake of constructing theLDPM meso-structure, both sand and silica flour were considered“aggregate”, and both cement and silica fume were considered“binder”. This results in water-to-binder and aggregate-to-binderratios of w/b ¼ 0.15 and a/b ¼ 1.03, respectively.

Fibers embedded in thematrix were hooked steel Dramix ZP305with 0.55 mm diameter, cross sectional area of 0.2376 mm2, 30 mmnominal length, and a reported Young’s modulus and tensilestrength of 210,000 MPa, and 1100 MPa, respectively.

LDPM-F modeling requires, as a first step, to provide anapproximated geometric description of concrete internal structureat themeso-scale, i.e., the scale at which thematerial can be viewedas a two-phase composite in which hard and stiff inclusions areembedded in a somewhat softer and weaker matrix. In addition,LDPM-F provides realistic representation of material failure iffracture processes are associated with paths, the tortuosity ofwhich depends on the size of the major heterogeneities. Thistypically happens if aggregate particles and/or inclusions are strongenough to deviate cracks and make them propagate through theembedding matrix only.

Fig. 1d shows a CT-scan of an undamaged CORTUF cylindricalspecimen. A voxel intensity histogram was used in the X-raymicrotomography to describe the phases of the material [100]. Thevisible circular holes form the pore spaces and surrounding air inthe matrix, the larger darker particles are the aggregates, thecement hydrates are the smaller lighter particles, and the brightwhite inclusions are unhydrated particles. A particulate internalstructure is clearly visible and strongly motivates the adoption ofLDPM-F for CORTUF modeling. Furthermore, Fig. 1e shows a post-mortem CT-scan of the same CORTUF specimen subjected tosplitting test. The crack path is affected by the presence of themajorinclusions, which in most cases, are undamaged (see white arrowsin Fig. 1e) and determined the location and propagation path of thevisible cracks.

3.2. Coarse-grained approximation

In the process of approximating the internal structure of con-crete, one of the most important parameters is the maximumaggregate size, da, which for CORTUF is equal to 0.6 mm. Theminimum aggregate size included in LDPM simulations is typicallyassumed to be d0 ¼ da/2, so if the particle-size distribution is gov-erned by a Fuller curve with exponent nF ¼ 0.5, the percentage ofsimulated aggregate is about 30%. This ensures, in most cases, agood resolution of concrete internal structure and a reasonablerepresentation of meso-scale potential crack paths. However, par-ticle sizes ranging from 0.3 mm to 0.6 mm lead to a very largenumber of particles for typical numerical specimens. For example,for a 50.8 mm (2 in.) diameter and 101.6 mm (4 in.) height cylinder

used for the CORTUF characterization in compression, one wouldhave about 1 million particles; for the CORTUF slabs that were304.8mm� 304.8mm� 50.8mm (12 in.�12 in.� 2 in.) and testedunder projectile impact at ERDC, the particle number would bemore than 20 million.

To reduce the computational cost of the simulation, a coarse-graining (CG) technique exploiting a force matching approach wasadopted [101]. This coarse-graining technique is based on thefollowing observations: (1) particle size does not influence themacroscopic LDPM response as long as the macroscopic behaviorremains strain-hardening anddoes not feature damage localization;(2) particle size does influence the macroscopic LDPM response inthe case of macroscopic strain-softening and damage localization;(3) the main parameter governing LDPM strain-softening behavioris the material characteristic length, defined as [t ¼ 2E0Gt=s2t ,where E0 ¼ normal modulus, Gt ¼ meso-scale tensile fracture en-ergy, and st ¼ meso-scale tensile strength. For a coarse-grainingfactor k ¼ dcoarsea =da ¼ dcoarse0 =d0 and with the same Fuller coeffi-cient for both the coarse and the fine system, the macroscopiccoarse response can be shown to be a good approximation of thefine response if the coarse characteristic length is calculated as

[coarset ¼ [t[coarse

[(4)

where [ ¼ interparticle distance, which is locally variable accordingto the grain size distribution, and the overbar represents the meanvalues of the interparticle distance distributions. It must be observedthat in general, [

coarse=[sk, because the LDPM system is not

composed by close-packed spheres and, consequently, the inter-particle distance distribution does not coincide with the particle sizedistribution. In this paper, a maximum aggregate size equal to 4 mmand a Fuller coefficient equal to 0.5 were adopted for the coarsenedsystem. Fig. 1f shows a 20 mm specimen with visible particles formaximum aggregate sizes of 4 mm (left) and 0.6 mm (right); Fig. 1gshows the associated Cumulative Distribution Function (CDF) of theinter-particle distance. In this case, [coarse=[ ¼ 5:5 is obtained.

The accuracy and effectiveness of the proposed CG technique isinvestigated in Fig.1h and i, which compare the average response ofthree fine (dashed curve) and three coarse (solid curve) LDPMspecimens under unconfined compression tests and splitting ten-sile tests, respectively. The simulated specimens were 20 mm sidecubes comprised of an average of 51,000 and 450 particles for thefine and coarse system, respectively. Simulations were performedon the Diamond system at the ERDC Department of DefenseSupercomputing Resource Center, having 2.8 GHz Intel Xeon X5560Nehalem-EP processors on its login and compute nodes, with 2processors per node, each with 4 cores, for a total of 8 cores pernode. The simulations utilized 1 node and were characterized by arun time of 10 h and 20 min (compression) and 5 h and10 min (splitting) for each run of the fine and coarse systems,respectively. As can be seen from the results obtained, the coarseapproximation does provide very accurate results with a significantsaving in terms of computational time; the difference in the peakresponse is about 7% for the compression test and 6% for thesplitting test. The observation must be made that the analyzedexamples are characterized by a damage zone whose size is largerthan the coarse maximum aggregate size. The same level of accu-racy cannot be expected for cases in which this condition is notsatisfied.

3.3. Fiber reinforcing

Fiber reinforcing is always used in UHPC to increase toughness(energy absorption capability) and ductility. Within LDPM, fibers

Page 6: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e3218

are explicitly modeled by generating spatial distributionsmimicking the ones found in real specimens. From a geometricpoint of view the physical fibers can be described using fewparameters, in the case of straight fibers: volume fraction, Vf,length, Lf, and equivalent diameter, df. Each fiber is modeled using asequence of one ormore segments linked together. Single segmentsare sufficient for generating straight fibers. Multiple segments arenecessary for generating tortuous fibers.

These geometric parameters are used for generating randomfibers inside a control volume. For cast fiber-reinforced concretespecimens, the control volume is assumed to be equal to the volumeof the concrete part. In this case, fiber location and orientation areaffected in thevicinityof theboundary.On the contrary, formachinedor cored specimens, the control volume is assumed to be larger thanthe actual specimen so that fibers intersecting the external surface ofthe specimens are treated as cut. The portion of a cut fiber inside thespecimen is shorter than the length of original fiber, and this affectsthemechanical characteristics of the fibereconcrete interaction nearthe specimen boundaries.

The following algorithm is used for generating random fiberswhose number inside a given control volume V can be computed asN ¼ 4Vf V=ðLfpd2f Þ. The first scenario details the case of uniformlydistributed fibers. For each fiber, a random point with uniformprobability in space inside the control volume is selected as theinitial point of the fiber. An initial random direction is thencomputed by (1) generating, in a unit cubic volume, a series ofadditional random points with a uniform probability in space, (2)rejecting all points outside a unit sphere, and (3) accepting the firstpoint occurring inside the unit sphere. The vector connecting thefiber origin to this point inside the unit sphere identifies an unbi-ased spatial direction. The computed direction is used to initializethe fiber. If the fiber consists of multiple segments to simulate fibertortuosity, then the direction of each additional segment is modi-fied using the following logic.

A random direction, w, is computed as done previously, fol-lowed by the normal n to the current fiber direction u1 asn ¼ m=jmj, where m ¼ w � (w$u1)u1. A proportional componentis then added to the tortuosity factor t by v2 ¼ u1 þ t n, and the newfiber direction u2 is normalized using u2 ¼ v2=jv2j. For t ¼ 0, thefiber continues straight in the initial direction. If the fiber exits fromthe control volume V, the fiber is discarded.

In some cases, fiber orientation is not random but has a biastowards a preferential direction. This bias is obtained by a scalingfactor, s, for a stated preferential direction. For s ¼ 1, there is nopreferential alignment, while a generic value of s makes alignments times more likely for the preferential fiber orientation. The mostappropriate scaling factor must be chosen through comparisonsbetween numerical and experimental specimens. The random fibergeneration algorithm is adjusted to include preferentially oriented

Fig. 2. Effect of high strain rate on compression test of UHPC a) Plot of DIF with

fibers after the spatially random direction is computed. The dotproduct c ¼ n$d of random n and preferential orientation d iscomputed. The random direction is then extended along the pref-erential direction and normalized to compute the random fiberdirection by v ¼ n þ (s � 1)cd and n ¼ v=jvj. Figs. 4a and c showgenerated fiber distributions without and with preferentialorientation.

In the spirit of the discrete multi-scale physical character ofLDPM, the occurrences of fiber-facet intersections are determinedby actually computing the locations where fibers cross inter-cellfacets. This computation can require significant resources forlarge models in terms of both time and memory. For this reason, anefficient bin-sorting algorithm was developed for computing theseintersections. For each intersection, the lengths of the fiber on bothsides of the facet and the angle at which the fiber intersects thefacet are also computed. These parameters are saved in the facetdata structure and used during the simulation for computing thefiber effect.

The modeling of fiber reinforcement in CORTUF is thencompleted with the simulation of fibereconcrete interaction,which requires the formulation of a local force versus slippageconstitutive equation simulating fine-scale failure phenomenaoccurring at the interface between fibers and embedding concretematrix. The LDPM-F formulation [88,89] of such a law, adopted inthis study and summarized in Appendix A2, includes the descrip-tion of the following failure mechanisms, (1) fiber/matrixdebonding, (2) fiber frictional pull-out, (3) local matrix failure dueto stress concentration at the point of exit of fibers from the matrix(micro-spalling), (4) pull-out strength increase due to additionalfrictional effects caused by change of direction of fiber duringcracking (snubbing), (5) fiber rupture, (6) fiber strength reductiondue to localized plastic deformations, and (7) fiber yielding due tolocal fiber bending.

In addition to these failure mechanisms, there is evidence forCORTUF concrete that splitting failure, i.e., a sudden and dynamicpropagation of a crack in a plane containing the fiber axis, mightcontribute significantly to the overall tensile resistance in tensionof the material. Let us consider a straight segment of a fibersubject to tension. At the interface between the fiber and thesurrounding concrete, shear stresses develop, which in turn leadto inclined principal stress directions. When the maximum prin-cipal stress exceeds the tensile strength, radial cracks initiate, andthe concrete matrix surrounding the fiber is forced to transferload through an inclined compression field, see Fig. 1c. Thecomponent of this inclined field acting orthogonally to the axis ofthe fiber results in an expansive pressure on the surroundingconcrete of magnitude p ¼ Pftan a/(pdfLe). The angle a depends onboth the geometric characteristics of the fiber and the remotestress state. From this pressure, the maximum circumferential

varying strain rates; Failure pattern for, b) 10�1 s�1, c) 0.05 s�1, d) 10 s�1.

Page 7: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

Loa

d (N

)

CMOD (mm)0 0.02 0.04 0.06 0.08 0.1

0

200

400

600

800

1000

Experimental

Numerical

0 0.05 0.1 0.15 0.20

200

400

600

800

1000

1200

Axi

al S

tres

s (M

Pa)

Axial Strain (Offset = 0.05)

UniaxialStrain

Triaxial300 MPa

Numerical

Experimental

Triaxial0 MPa

a) b) c)

Loa

d (N

)

Displacement (mm)0 5 10 15

0

50

100

150

200

250

300Embed 12.7 mmEmbed 6.35 mm

Beta = 0

Beta = 0.1

Beta = 0.048

Fig. 3. Tests results for the parametric identification of LDPM-F for a) 3-point bending, b) uniaxial strain and triaxial compression, c) single fiber pull-out.

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e32 19

tensile stress acting at the fiber/matrix interface (see Fig. 1c) canbe computed approximately as

smaxq ¼ p

4c2f =d2f þ 1

4c2f =d2f � 1

(5)

where cf is the distance between the fiber axis and the closestexternal surface of the specimen.

By introducing the expression for p into Eq. (5), setting thefailure condition, smax

q ¼ s�t , and solving for the fiber force, one cancalculate the splitting force

Psplitf ¼ pdf Les

*t

tan a

!4c2f =d

2f � 1

4c2f =d2f þ 1

zpdf Les

*t

tan afor cf =df[1

(6)

where s�t can be assumed to coincide with the tensile strength ofthe matrix or the composite, depending on fiber concentration andlevel of fiber interaction. At this point, a few observations are inorder. (1) The splitting mechanism highlighted below is typicallyobserved in reinforced concrete members [102] when concreterebars are not provided with proper confinement through eitherenough concrete cover or transverse reinforcement. It is worthpointing out that, in regular concrete, rebars have diameter sizes ofthe same order of the aggregate particles; the same situation arises

Fig. 4. a) Random fiber orientation, b) CT scan of fiber orientation, c) preferential fiber orientspecimen on CORTUF-Plain, f) failure mode for 3-point bending specimen on CORTUF-Fibe

in CORTUF concrete where fiber diameters have similar character-istic sizes of material heterogeneities. Based on this analogy, itseems logical to observe in CORTUF failure mechanisms that aresimilar to the ones of reinforced concrete and that are not observedin standard fiber-reinforced concrete. (2) Splitting failure isexpected to be negligible in the case of fiber reinforcement withlow stiffness and high Poisson’s ratio compared to the one of con-crete (e.g., for Polyvinyl Alcohol, PVA, fibers), because such prop-erties prevent the build-up of a significant expansive pressure. (3) Ifbond resistance is limited, bond failure occurs before the inclinedcracks can develop. As such, straight fibers are less prone to inducesplitting failure than hooked and bent fibers.

3.4. Rate effect

Due to the specific curing protocol used to manufactureCORTUF, it can be assumed confidently that capillary and absor-bed water content is very small and actually close to none. Forthis reason, intrinsic phenomena such creep and moisture effectcan be neglected in the simulations. Arguably one could includethe Arrenyus-type effect but this would require identification ofthe relevant parameters, which is possible only through theanalysis of rate-dependent experimental data that is not availableat the moment for CORTUF. Consequently, a choice was made toneglect any intrinsic phenomena in the LDPM-F formulation forCORTUF.

ation, d) validation test results for 3-point bending, e) failure mode for 3-point bendingr.

Page 8: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

Table 2Tests performed for calibration and values of LDPM parameters in simulations.

Testperformed

LDPM parameter UHPC UHPC-L UHPC-U RSC

Uniaxialstrain

1. Compressivestrength

500MPa

425MPa

575MPa

150MPa

2. Initial hardeningmodulus ratio

0.36 0.306 0.414 0.4

3. Densificationratio

2.5 2.5 2.5 1

4. Transitionalstrain ratio

2 2 2 2

Uniaxialunconfinedcompression

5. Shear strengthratio

17 17 17 2.7

6. Initial friction 0.5 0.5 0.5 0.27. Softeningexponent

0.2 0.2 0.2 0.2

Triaxialcompression

8. Deviatoric strainthreshold ratio

2 2 2 1

9. Deviatoricdamage parameter

1 1 1 5

10. Asymptoticfriction

0 0 0 0

11. Transitionalstress

300MPa

300MPa

300MPa

600MPa

3-Point test(Notch)

12. Tensile strength 4 3.4 4.6 4.0313. Tensilecharacteristic length

150 150 150 120

Single fiberpull-out

14. Bond strength 11.5MPa

11.5MPa

11.5MPa

15. Coarseaggregate

4 mm 4 mm 4 mm 8 mm

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e3220

However, the model is still able to capture the rate-effect asso-ciated with apparent mechanisms. Fig. 2a shows the DIF obtainedby simulating a cylindrical (50.8 mm � 101.6 mm) CORTUF spec-imen fixed at the base and subjected, at the top, to various loadrates corresponding to nominal strain rates (velocity of the top sideof the specimen divided by the specimen length) from 0.001 s�1 to10 s�1. As one can see the response shows an increase of themacroscopic compressive strength up to 1.24 times the quasi-staticvalue. This increase results solely from inertia effects and change incrack patterns (Fig. 2bed). Finally, it is worth noting that theadopted formulation cannot capture apparent mechanisms occur-ring at a length scale smaller than the adopted minimum LDPMparticle size (2 mm in this study). However, introduction in theconstitutive equations of these fine scale effects would be only afutile academic exercise without relevant and appropriate experi-mental data needed to calibrate and validate the resulting morecomplicated model.

3.5. Model parameter identification

The LDPM-F constitutive equations at the facet level depend ona number of model parameters (see Appendix A2) that need to beidentified by fitting experimental data. Extensive discussion ofLDPM and LDPM-F calibration procedures is reported inRefs. [86,89].

In this study, first, the elastic parameters were calibrated bymatching the macroscopic elastic modulus obtained from theaverage slope of the CORTUF-Plain and CORTUF-Fiber unconfinedcompression curves (51,591MPa) and Poisson’s ratio as determinedby Ref. [103] having a value of 0.115. By using the macro-to-mesoformulas reported in Ref. [85], one obtains the normal elasticmodulus, EN ¼ 67,000 MPa, and the shear-normal couplingparameter, a ¼ 0.484. Experimental data from both CORTUF-Plainand CORTUF-Fiber were used, because the LDPM-F formulationneglects the effect of fiber reinforcement on the elastic behavior.This is a reasonable approximation for low-fiber-volume fractionsas the ones typically used in concrete (<10%). Next, identification ofthe other concrete static parameters was performed by matchingthe experimental results of CORTUF-Plain in four different tests,namely, uniaxial strain (UX), uniaxial unconfined compression(UC), triaxial compression (TXC) with 300 MPa confinement, and3-point bending (3PBT) on notched specimens. Finally, the LDPM-Fparameters related to concreteefiber interaction were identifiedby the best-fit of single pull-out tests relevant to two fiberembedment lengths, 6.35 mm (0.25 in.) and 12.7 mm (0.5 in.).Table 2 summarizes the parameters specific to each test as well asthe resulting identified values. Model parameters whose calibra-tion require experimental data not available for CORTUF wereassumed on the basis of standard values available in the literature[86] and reported in Appendix A.

In order to account for the variation that concrete exhibits inexperiments, the meso-structure for each simulated test wasgenerated three times, and simulations were performed on eachmesh generation. Similarly, fibers in the fiber-reinforced CORTUFspecimens were generated three times to reproduce the effect offiber distribution. The average of the numerical results is plotted forthe numerical data discussed hereinafter. Experimental data arecompared with the numerical ones either with average curves,circles points in the figures, or with curve ranges, shaded gray areasin the figures, identified between the minimum and maximumenvelopes of the experimental curves. In addition, unless otherwisespecified, the comparison is done on the basis of simulations per-formed on the coarse-grained model. The numerical simulationswere performed with the Modeling and Analysis of the Response ofStructures (MARS) software [104].

Tensile strength and toughness parameters of LDPM wereidentified by simulating 3PBTs, which were performed on speci-mens that had a 50.8 mm (2 in.) width, a 25.4 mm (1 in.) depth, a203.2 mm (8 in.) span, and a notch dimension of 15.24 mm (0.6 in.)depth and 5.08 mm (0.2 in.) wide. Load versus CMOD (Crack MouthOpening Displacement) curve comparison for the experimental andnumerical results are shown in Fig. 3a.

Fig. 3b shows the best fitting obtained for the Uniaxial Strain(UX) test, characterized by zero lateral expansion, as well as TriaxialCompression (TXC) tests at 0 MPa (Unconfined Compression, UC)and 300 MPa lateral confinement (TXC300), respectively. Thesetests were used to identify the compression-related LDPM param-eters. In Fig. 3b, the stress versus strain curves are offset by 0.05 forclarity of representation, and the unloading branches of the UX andTXC300 are included. Compression tests were all conducted oncored cylindrical specimens having lengths of 101.6 mm (4 in.) anddiameters of 50.8 mm (2 in.) and were loaded through loadingplatens treated with friction reducing materials in order to mini-mize the effect of friction between the platens and the specimenends. Additional details of the experimental procedure can befound in Refs. [98,99]. As can be seen in Fig. 3b, LDPM-F canaccurately reproduce the behaviors of CORTUF at various levels ofconfinement, i.e., very brittle strain-softening behavior at noconfinement and ductile and strain-hardening behavior at highconfinement. The model also nicely captures CORTUF unloadingbehavior.

Finally, the calibration of LDPM-F requires the identification ofthe material parameters governing the fiberematrix interactionmodel. The basic test used for this task is the single fiber pull-out(SPO) test. When a fiber is pulled out of concrete, the initial resis-tance is due to both debonding fracture energy as well as frictionalresistance [105]. Once the fiber debonds completely from thematrix, the resistance shifts to purely frictional, which can be eitherslip-hardening, increasing pull-out force for increasing slippage, orslip-softening, decreasing pull-out force for increasing slippage.This feature of the pull-out behavior can be simulated through the

Page 9: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e32 21

coefficient b (see Ref. [88] and Appendix A) that is >0 in the case ofslip-hardening and <0 in the case of slip-softening.

The shaded portions of Fig. 3c represent the data obtained fromexperimental tests with two distinct embedment lengths, 12.7 mmand 6.35 mm (0.5 in. and 0.25 in.), and direction of pulling coin-ciding with the fiber axis. The shorter embedment length showsslip-hardening, whereas the longer embedment length displaysmore of a slip-softening response, and they can only be fitted withtwo distinct values of the parameter b (see dashed lines in Fig. 3c).A decision was made to adopt b ¼ 0 (see solid lines in Fig. 3c) thatprovides an overestimate of the pull-out capacity of the longerembedment length and an underestimation of the shorter. Theinability of the model to capture the differences in behavior fordifferent embedment lengths is due to the presence of hooks atfiber ends that are not accounted for explicitly in the analyticalmodel [91].

However, this limitation is not crucial for the present study,since only a small portion of the frictional pull-out resistanceaffects the behavior of the overall composite, which is actuallygoverned by other failure mechanisms. One of these mechanisms isthe splitting failure discussed earlier in this paper (see Eq. (6)). Inabsence of specific experimental data, the following assumptionsare employed, s�t ¼ st ¼ 4 MPa, Le z Lf/2 ¼ 15 mm, and a z p/6.This leads to an average splitting resistance of Psplitf ¼ 179 N. Notethat the value a z p/6 is motivated by the likely occurrence ofcompressive stresses parallel to the fiber axis induced by the fiberhooks and leading to a reduction of the inclined crack orientation,which would be p/4 in the case of pure tangential stresses (notethat similar behavior is observed for shear cracks in prestressedconcrete members).

Other failure mechanisms include fiber rupture with inclinedpulling,micro-spalling, and snubbing, whose governing parameters(reported in Appendix A2) were assumed on the basis of Ref. [88].

3.6. Model validation and discussion

This section presents the validation of the formulated model.Validation simulations are necessary to prove that LDPM-F servesits intended purpose of predicting the mechanical responses ofCORTUF-Plain and CORTUF-Fiber. All prediction simulations wereexecuted with no further change to the previous calibrated LDPM-Fparameters, and their results are compared to experimental datathat were not utilized during the parameter identification.

3.7. Fracture tests

The same notched 3PBT used for the fracture characterization ofCORTUF-Plain was also used for CORTUF-Fiber. The experimentswere intended to be performedwith a random distribution of fibers(Fig. 4a) and with a fiber volume fraction equal to Vf ¼ 3% corre-sponding to the fiber-to-binder ratio in mass equal to 0.31 reportedearlier in Table 1.

Numerical results of the corresponding numerical simulationsshowed lower peaks compared to the experimental data as shownin Fig. 4d. However, computer tomography (CT) scans [100]revealed Vf to be between 3.5 and 4% in the vicinity of the notchtip as opposed to the nominal 3%. In addition, the scans providedinsight into the fiber placement and distribution; fibers were notdistributed randomly (Fig. 4a) as intended but rather had a pref-erential orientation in the direction orthogonal to the notch (Fig. 4band c).

Fig. 4d shows the load versus mid-span-deflection curve forCORTUF-Fiber with random (thin solid curves) and preferential(thick solid curves) fiber orientation for Vf equal to 3, 4, and 5%.Simulations with random fiber orientation are consistently lower

than the experimental results with the curve relevant to the highestvolume fraction barely reaching the experimental lower bound. Onthe contrary and in agreement with the CT scans, the predictionswith preferential fiber orientation coincides with the experimentalresults very well. Comparison of the 3PBT CORTUF-Plain result (seeFig. 3a) with the 3PBT CORTUF-Fiber result (see Fig. 4d) demon-strates that the effect of fibers leads to an increase of the carryingcapacity of more than four times, from about 0.9 kN to 3.8 kN (for4% fiber volume fraction). Furthermore, the capability of thematerial to dissipate energy increases by orders of magnitude.Numerical fracture energy values computed through the work-of-fracture concept [71] give estimates of 30 N/m and 9800 N/m forCORTUF-Plain and CORTUF-Fiber (Vf ¼ 4%), respectively, at 40% lossof load carrying capacity.

Figs. 4e and f report the crack distributions at 0.1 mm and5.9 mm deflection for the CORTUF-Plain and CORTUF-Fiber,respectively. Unlike the unreinforced specimen, the fiber additioncreates a more distributed crack pattern with a significant increaseof the fracture process zone (FPZ) size.

The numerical response of CORTUF-Fiber and the associatedfailure mechanisms are further investigated in Fig. 5a. In this figure,the results of the numerical simulations for Vf ¼ 4% obtained bypreventing both fiber rupture and bond splitting (curves P3 and R3),and preventing bond splitting only (curves P2 and R2) are comparedwith the response accounting for all failure mechanisms (curves P1and R1). The prefix “P” indicates preferential fiber orientationwhereas “R” indicates random fiber orientation. Additionally, theCORTUF-Plain results are included for comparison as Plain. As onecan see, the dominating failure mechanism is bond splitting. Whenthis failure mechanism is prevented, the load carrying capacitymore than doubles, i.e., at 3 mm deflection, the ratio between theload curves P3:P1, and R3:R1 is 2.4 and 2.3, respectively. On thecontrary, fiber rupture does not have a significant role as demon-strated by the comparison between P2 and P3 as well as R2 and R3.Moreover, results show a slightly more important effect of fiberrupture for the random orientation case. This is due to the fact that,in this case, there is a higher likelihood for fibers to bridge cracks atan angle leading to fiber bending with associated localized damageat the bent location [91,106e109].

The conclusion that fiber bond splitting might be the dominantmechanism is also confirmed by performed CT-scan analyses,which did not show any significant occurrence of fiber rupture oreven necking and, at the same time, provides evidence of possiblecracks propagating along the fiber axis (see Fig. 6a).

Furthermore, post-mortem evaluation of tested specimensshowed the presence of undamaged fibers (with even intact hooksand bents) across the crack surface (see Fig. 6b). This is notconsistent with fiber rupture, leading to the separation of the fibersin two pieces, or pull-out, leading to plastic deformation andstraightening of the fiber hooks, but it is completely consistent withthe bond splitting mechanism hypothesized in this paper.

For the case with fiber preferential orientation, the onset ofbond splitting failure occurs for a load level of about 3.5 kN asshown in Fig. 5a by the point at which the P1 and P3 curves start todiverge. To further investigate this phenomenon, crack openingdistribution with cumulative distribution function plots are shownin Fig. 5b and c for a centerline displacement of 1.2 mm (dotted linein Fig. 5a), which is about 75% of the maximum stress. Fig. 5b showsthe comparison of the random fiber distributions, while the resultsof fibers with preferential orientation are shown in 5c. The relativecrack distribution of the plain specimen at 75% of its maximumstress is also included in the plots. At the onset of failure, P1 and R1exhibit slightly higher crack distributions than P2, P3 and R2, R3,respectively. As the crack openings become greater than about100 mm for the random distribution, there is a change in response

Page 10: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

Fig. 5. 3-Point bending test results for a) stressestrain curve fiber failure mechanisms; crack distribution function for b) random fiber orientation, c) preferential fiber orientation.

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e3222

as the coupled R2 and R3 are able to bridgemore of the larger cracks.Lastly, comparisons between the plain and the reinforced speci-mens clearly show the crack-bridging effect of the fibers. For theplain specimens, the cracks governing failure are magnitudes lessthan the reinforced specimens. The inability of cracks to propagatebeyond a few mm results in localized failure.

3.8. Tensile strength tests

The tensile strength of the CORTUF-fiber was investigatedthrough four-point bending tests and splitting (Brazilian) tests.

Three different specimen sizes were used for the 4-point (4pt)testsdSize A, Size B and Size C (see Fig. 7a). The width for allspecimens was 102 mm, and the other dimensions in units of mmare as follows: Size A (l1 ¼ 101, l2 ¼ 305, l3 ¼ 356, h ¼ 25), Size B(l1 ¼101, l2 ¼ 305, l3 ¼ 356, h ¼ 102), and Size C (l1 ¼ 304, l2 ¼ 914,l3 ¼ 1016, h ¼ 102). In the 4pt test, a concrete block was supportedon two supports, and a load was applied through a loading plate atthe top of the specimen. To imitate the loading plate used in theexperimental setup, an additional elastic plate was used to applyload on top the specimen in LDPM. Unlike the cylindrical andlargest 4pt specimens that were cast, ERDC reported that the

Fig. 6. a) CT scan of fiber bond failure on 3-point bending specimen, and b) 4-pointbending specimen showing split fiber failure.

smaller size 4pt specimens (A and B) to be cut, and this feature wasalso reproduced in the LDPM simulations.

Simulations with both random and preferential orientation ofthe fibers were performed and, similarly to the 3PBT, results showthat the fibers were preferentially oriented in the specimenin the longitudinal direction. Numerical and experimentalnominal stress, sN ¼ 3F(l2 � l1)/bh2, versus nominal strain,3N ¼ 12hD=ð3l22 � 4ðl2 � l1Þ2Þ, curves are reported in Fig. 7bed, forthe specimen sizes A, B, and C, respectively. For the 4% fiber volumefraction, the peak nominal stress was 55.8, 51.7, and 51.7 MPa forthe three different sizes that corresponds to 24.8% and 23% of thecompressive strength. These values are in very good agreementwith the respective experimental values of as shown in Fig. 7bed,but it must be noted that both experimental and numerical dataprovide a significant overestimation of the actual tensile strengthbecause, due to the fiber crack bridging effect, the specimensresponses show a pronounced nonlinear behavior prior to the peaksuggesting a stable crack propagation after crack initiation andbefore reaching the load carrying capacity. The numerical resultsand the experimental data are also compared in Fig. 7e as far as thecrack distribution is concerned, showing the excellent ability ofLDPM in reproducing realistic crack patterns.

The cylindrical specimens used for the splitting tests werecharacterized by a radius of 51 mm (2 in.) and length of 204 mm(8 in.). The test was performed according to the ASTM standard, anda compressive load was applied along the specimen’s length withrectangular rods that prevented crushing of the cylinder under-neath the applied load. The rod had dimensions of 10 mm width,226 mm length, and 5 mm height. In a splitting test, the maximumstress occurs at the center of the specimen in the directionorthogonal to the applied load and decreases closer to the supportsuntil eventually stresses at the support become negative(compression). This leads to a longitudinal crack that splits thecylinder into two pieces as Fig. 7f demonstrates by comparing post-mortem images of both an experiment and a simulation.

Peak stress numerical results obtained for the CORTUF-Fiberspecimens containing Vf values of 3, 4 and 5 were 10.8 MPa,12.3 MPa, and 13.8 MPa respectively, and the peak stress for theCORTUF-Plain specimen was 7.2 MPa. The average experimentalsplitting stress for CORTUF-Fiber was 25.3 MPa, and for CORTUF-Plain was 10.8 MPa. The discrepancy in these values, especiallyfor CORTUF-Fiber, can be explained by the fact that a compressivestress field exists in the direction parallel to the crack path, andconsequently, transverse to the crack bridging fiber during splittingfailure. Such transverse confinement can significantly influence thepull-out resistance by preventing certain failure mechanisms (e.g.,splitting and spalling) and by increasing the frictional resistance.This phenomenon is not currently handled in LDPM-F, because thefibereconcrete interactionmodel does not depend on the confiningstress acting transversely to the fiber axis.

Page 11: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

Fig. 7. Validation test results for 4-point bending test on specimens. a) Pictorial size description; stressestrain curve for specimen b) size A, c) size B, and d) size C; failure modecomparison for e) 4-point bending, and f) splitting tensile test for CORTUF-Fiber specimens.

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e32 23

3.9. Uniaxial, triaxial, and hydrostatic compressive tests

Fig. 8 shows comparisons between experimental data andnumerical simulation results with reference to triaxial tests onCORTUF-Plain and CORTUF-Fiber that were not used in the cali-bration phase. For completeness, both experimental and numericaldata, previously simulated for the purpose of parameter identifi-cation are reported in light gray.

Figs. 8a and d report CORTUF-plain and CORTUF-fiber re-sponses, respectively, under uniaxial strain compression and

a) b)

d) e)

Prin

cipa

l Str

ess

Dif

fere

nce

(MPa

)

Volumetric0 200

0

100

200

300

400

500

600

Vol

umet

ric

Stre

ss (

MPa

)

Volumetric Strain0 0.02 0.04 0.06 0.08

0

100

200

300

400

500

600

700 Uniaxial

Hydrostatic

Experimental

Numerical

Prin

cipa

l Str

ess

Dif

fere

nce

(MPa

)

Volumetric0 200 4

0

100

200

300

400

500

600

Vol

umet

ric

Stre

ss (

MPa

)

Volumetric Strain0 0.02 0.04 0.06 0.08

0

100

200

300

400

500

600

700 Uniaxial

Hydrostatic

Numerical

Experimental

Fig. 8. Tests results for the validation of LDPM-F for a) uniaxial strain and hydrostatic comprCORTUF-Plain, d) uniaxial strain and hydrostatic compression on CORTUF-Fiber, e) uniaxial

hydrostatic compression. The results are plotted in terms of volu-metric stress versus volumetric strain, and consequently theyfeature the same elastic portion of the curve represented by thevolumetric elastic modulus, EV ¼ E/(1 � 2n) ¼ 67000 MPa. In theinelastic regimes, the two curves are distinctly different. Resultsshow that LDPM-F can capture such different behavior, unlike mostother constitutive equations available in the literature. LDPM-F canalso capture the behavior for unloading very well. It is worthnoting that the difficulties of most models to capture the differencebetween the volumetric stress versus volumetric strain curves

c)

f) Stress (MPa)

400 600 800

Experimental

Numerical

0 0.1 0.2 0.3 0.40

200

400

600

800

1000

Axi

al S

tres

s (M

Pa)

Axial Strain (Offset = 0.05)

200

100

20

0

Experimental

Numerical

50

10

300 MPa

Stress (MPa)00 600 800

Experimental

Numerical

0 0.1 0.2 0.3 0.40

200

400

600

800

1000

Axi

al S

tres

s (M

Pa)

Axial Strain (Offset = 0.05)

300 MPa

200

100

Numerical

Experimental

0

5020

10

ession on CORTUF-Plain, b) uniaxial strain on CORTUF-Plain, c) triaxial compression onstrain on CORTUF-Fiber, f) triaxial compression on CORTUF-Fiber.

Page 12: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e3224

obtained under uniaxial strain and hydrostatic compression areassociated with the interpretation that these curves are an“equation of state”. As many other experiments have alreadyshown, this interpretation cannot be used in the case of concrete,and frictional materials in general.

For the uniaxial strain tests, the numerical versus experimentalcomparison is further investigated in Fig. 8b and e with plots ofprincipal stress difference versus volumetric stress. Once again,LDPM-F is able to reasonably simulate the CORTUF response duringboth the loading and unloading phases of the test.

Simulation of the stress versus strain curves from the TXC testson CORTUF-Plain with varying transverse confinement (10, 20, 50,100, 200 MPa) not used in the parameter identification phase isshown in Fig. 8c where the curves are offset by 0.05 for clarity ofrepresentation. Similarly, Fig. 8f shows the CORTUF-Fiber TXC testresults with varying transverse confinements (0, 10, 20, 50, 100,200, 300 MPa). As one can see, the experimental data suggests avery small effect of fibers on the overall compressive response ofCORTUF. This is consistent with the LDPM-F assumption that fibersare engaged only for cracks displaying tensile normal opening andthat they do not contribute significantly to the resistance of thecomposites under combined meso-scale compression andshearing.

Furthermore, CORTUF shows a significant increase of strengthupon confinement. For example, for a transverse confinement of100 MPa, the strength for both CORTUF-plain and CORTUF-fiber isin excess of 550 MPa, corresponding to well over two times theunconfined compressive strength. LDPM-F simulates the strengthincrease for all levels of confinement considered with excellentaccuracy. In addition to strength, the presence of confinementsignificantly affects the ductility of the material. The experimentaldata shows a transition from a brittle strain-softening response toa ductile strain-hardening response for increasing confinement,and the transition seems to occur at a level of confinement be-tween 100 and 200 MPa. LDPM-F simulates the brittle-to-ductiletransition but predicts the transition confinement to be at about100 MPa. In addition, for 20 MPa and 50 MPa confining pressure,the numerical responses seem to overestimate the materialductility, even though this cannot be known with certainty sincethe experimental data does not have the post peak portion of thecurve.

The transition from a brittle-to-ductile behavior in compressionis also clearly seen in Fig. 9aed that display crack patterns at failurewith varying confinement. For increasing confinement, the failureevolves from a brittle split crack, to one that resembles the occur-rence of shear bands, and then a complete, uniformly distributedcrushing that leads to the observed increased ductility.

Fig. 9. Triaxial compression crack pattern comparison for CORTUF-Plain

4. Numerical simulation of projectile penetration intoCORTUF slabs

This section discusses the use of LDPM for the simulation ofpenetration and perforation experiments.

In the recent past, various researchers have attempted this typeof simulations by using other computational technologies. TheFinite Element Method with element deletion [110] is the simplestway to simulate arbitrary propagating cracks and it simulatescracks by removing elements from the mesh when a certain cri-terion is met. This method requires a very fine mesh to providereasonably accurate results and it cannot be used to simulate theeffect of secondary fragments since most of the failed materialdisappears from the simulation.

Meshfree methods, such as the Smoothed Particle Hydrody-namics (SPH) method [111,112], have gained some popularity in thesimulation of penetration mechanics for their ability of allowingvery large deformations. Successful examples of application of SPHto the simulation of perforation through concrete can be found inRefs. [113e115]. However, the main problem of classical mesh-freemethods is the underlying assumption that concrete can be treatedas an isotropic continuum even during failure and fragmentation.Consequently, the transition from continuum to discrete is handledthrough numerical artifacts as opposed to an actual loss of conti-nuity of the governing mathematical formulation.

A solution to this problem is to describe displacement discon-tinuities explicitly. Meshless methods enhanced with the descrip-tion of displacement discontinuities were used successfully inRefs. [116,117] to simulate dynamic failure of concrete. However,this method still neglects the heterogeneous character of concreteand requires, unlike LDPM, additional degrees of freedom torepresent the evolving discontinuities making the computationalcost to increase significantly during fracture and fragmentationcompared to the unfractured state.

The simulated perforation experiments were performed atERDC on three square 304.8 mm (12 in.) CORTUF-Fiber prismaticslabs. The slabs, labeled A, B, and C, had widths 50.8 mm (2.0 in.),63.5 mm (2.5 in.), and 76.2 mm (3.0 in.), respectively. A fragmentsimulating penetrator (FSP) projectile was used in the experiments.Although one of each specimen size was tested experimentally,LDPM allowed for numerous simulations in a cost-effective andtime-efficient manner for a wide range in velocities below, at, andabove the ballistic limit of each panel, as well as various fiber-volume fractions (3, 4, and 5%).

The projectile was modeled as a plastic high-strength steelwith an elastic modulus of 200 GPa. The classical FlanaganBelytschko Single Integration Point finite element algorithm with

for a) 0 MPa, b) 10 MPa, c) 50 MPa, and d) 200 MPa confinements.

Page 13: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e32 25

hour-glass stabilization was used for its speed and accuracy with ahexahedral mesh. Additionally, there was no constraint on the FSPfor either translations or rotations. Fig. 10a shows the side, front,top, and back views of the FSP meshing. For the large contactforces exerted on the projectile, greater deformation was observedfor the FSP in the simulation compared to experimental results.However, due to the caliber having a relatively flat nose as seen inFig. 10a side and top views, the deformations were not excessiveenough to arrest the simulation. A penalty [104] contact algorithmwas adopted between the FSP and the panel, which utilized therelative displacement of the two nodes lists, i.e., FSP node list andpanel node list. The dimensionless nodes for the panel and bulletwere assumed to be spherical with 1 mm and 2 mm thicknessesrespectively for the contact, which occurs when the two sphericalexternal surfaces are in contact. Contact conditions between thenodes of the same list were omitted in the model.

The schematic of the test setup portrayed the panel in a holderheld only at the outer edges. With no definite knowledge about theexact boundary conditions of the panel during the experimenta-tion, all surface nodes for the panel width zone were held fixedwith no possibility of displacement or rotation, and further inves-tigation of the effects of varied boundary conditions were per-formed. Results of this boundary investigation are discussed later inthe paper.

The average ballistic limit for the projectile impacting threedifferent unreinforced specimens of Type A, the smallest panel, was1706 m/s. When a target is impacted, a compressive shock wave

Fig. 10. a) Side, front, top and back view of fsp projectile hexahedron mesh; b) Velocityetimefor size A panel above and fsp above ballistic limit (2743 m/s).

passes through the specimen and is reflected at the rear face as atensile wave. Ultimately, expulsion of the projectile from the frontface causes spalling. In addition, when the reflected tensile stresswave exceeds the tensile capacity of the material, scabbing occurson the rear face [19]. Figs. 10bed display the complete evolution ofthe phenomenon for an impact velocity greater than the ballisticlimit. The change in velocity of the FSP with time is presented inFig. 10b and c shows the penetration depth of the projectile withtime, and Fig. 10d displays the failure patterns observed for thepanel at 0.005ms time increments in LDPM for each stage of impactfor the side-panel view section. Figs.10b and c separate three stagesby vertical dotted lines. The initial decrease in velocity slope is dueto the compressive behavior of the impact, the first change invelocity slope, and is the first stage; the second stage, governed bycompression/shear corresponds to a tunneling effect in the panel;and finally, the rear-face scabbing coincides with the final change invelocity slope, the third stage before the curve plateaus upon pro-jectile exit. Since the width of Panel A is only 50.8 mm (2.0 in.), thetunneling effect is restricted, and the failure observed has limitedtunneling effect, as opposed to Sizes B and C.

Results depicting the change in velocity with time for impactvelocities ranging from 15 m/s to 3353 m/s, below and above theballistic limit, are shown for a CORTUF-Plain specimen A in Fig. 11aand b, respectively. For impact velocities below the ballistic limit,there was 100% attenuation of the exit velocity until the ballisticlimit, as shown in Fig. 11a. Fig. 11b shows a decrease in velocityattenuation for increases in projectile impact velocity. Additionally,

plot, c) perforation depthetime plot, and d) Failure patterns at 0.005 ms time intervals

Page 14: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

a) b) c)

d) e) f)

g) h) i)

0 0.05 0.1 0.15 0.20

200

400

600

800

1000

1200

1400

1600

Vel

ocity

(m

/s)

Time (ms)

15 m/s

30

152

305

610

914

1219

1524

1554

0 0.05 0.1 0.15 0.20

0.2

0.4

0.6

0.8

1

Nor

mal

ized

Ene

rgy

Time (ms)

KE

IE

IEC

IEB

DE

0 0.05 0.1 0.15 0.20

0.2

0.4

0.6

0.8

1

Nor

mal

ized

Ene

rgy

Time (ms)

KE

IE

IEC

IEB

DE

0 0.05 0.1 0.15 0.20

500

1000

1500

2000

2500

3000

3500

Vel

ocity

(m

/s)

Time (ms)

1554 m/s

1676

1829

2134

2438

2743

3048

3353

0 0.05 0.1 0.15 0.20

0.2

0.4

0.6

0.8

1

Nor

mal

ized

Ene

rgy

Time (ms)

KE

IE

IEC

IEB

DE

DEC

DEF

0 500 1000 1500 20000

20

40

60

80

100

Pene

trat

ion

Dep

th (

%)

Velocity (m/s)

UHPC

UHPC L

UHPC U

RSC

UHPC F3

UHPC F5

0 0.05 0.1 0.15 0.20

200

400

600

800

1000

1200

1400

1600

Vel

ocity

(m

/s)

Time (ms)

PLasticElasticRigid

0 0.05 0.1 0.15 0.20

0.2

0.4

0.6

0.8

1

Time (ms)

Nor

mal

ized

Ene

rgy

KE

IE

IEC

IEB

DE

600 1200 1800 2400 3000 36000

200

400

600

800

1000

1200

1400

Strike Velocity (m/s)

Exi

t vel

ocity

(m

/s)

UHPC

UHPC L

UHPC U

RSC

UHPC F3

UHPC F5

Fig. 11. Velocityetime curve for a) velocities below the ballistic limit, b) velocities above the ballistic limit; c) energy change of system with time for CORTUF-Plain at ballistic limitfor plastic fsp, d) velocityetime curve for varying fsp at ballistic limit; energy change of system with time at ballistic limit for e) elastic fsp, and f) rigid fsp g) CORTUF-Fiber (fiberballistic limit); parametric study plot with varied composites of h) penetration depthevelocity, and i) exitestrike velocity.

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e3226

the variation in energy (normalized) observed throughout theexperiment for the ballistic limit of CORTUF-Plain, is shown inFig. 11c. The kinetic energy (KE), the energy given to the system,decreases as the velocity of the bullet decreases with time since theenergy of the projectile governs and is directly proportional to thesquare of the projectile velocity. The internal energy (IE) in the plotsreference the work done by internal forces, which results inrecoverable elastic energy and dissipated energy. The internalenergy in the concrete (IEC) and the bullet (IEB) combine to formthe total internal energy in the system. The dissipated energy in theconcrete (DE) is the unrecoverable energy; it is less than the in-ternal energy in the concrete.

Further investigation was performed on the FSP considering theeffect of using elastic and rigid projectiles for the numerical sim-ulations. Fig. 11d shows a comparison of velocity change with timefor an impact velocity above the ballistic limit for plastic, elastic,and rigid projectiles. The results show the disparity betweenvelocities of the elastic and rigid projectiles, and the plastic pro-jectile that absorbs energy. The deformation of the plastic pene-trator is more than that in the elastic and the rigid, which results inan additional 100 m/s residual velocity for the elastic and rigidprojectiles. These values can be explained by comparing theenergies of the three systems. Figs. 11e and f reveal the energy in

the elastic and rigid systems, respectively. Compared to the plasticFSP (11c), the internal energy the projectile contributes to thesystem for the elastic FSP is reduced by 86.6%, and for the rigidprojectile it decreases to zero. Therefore, the internal energy of theconcrete imparts more or all internal energy to the system than theprojectile. For less energy dissipated by the projectile, the residualvelocity of the FSP is increased, and a lower velocity is needed forcomplete perforation. Hence, using a projectile that has a greaterdeformation results in greater ballistic limit.

The experimental penetration tests were conducted on rein-forced CORTUF; therefore, numerical simulations with fibers werealso performed to determine the effects for varied fiber volumetricfractions (VF) of 3%, 4% and 5%. Whereas the average ballistic limitof the plain concrete was 1706 m/s, that of the fiber reinforcedspecimen was 1783 m/s. The energyetime plot for an impact ve-locity on a 3% VF fiber-reinforced specimen is shown in Fig. 11g. Theenergies show a very similar trend to the unreinforced specimensFig. 11c with the clear difference being the energy dissipated in thefiber (DEF). However, most of the total dissipated energy (DE) is stillin the concrete panel as the fiber dissipates only about 3% of thetotal dissipated energy.

Additionally, a parametric analysis was performed on SpecimenA where three LDPM parameters were altered, tensile strength,

Page 15: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e32 27

compressive strength, and the initial hardening modulus ratio, byplus and minus 15% to capture the variations exhibited in the ex-periments for a composite 15% stronger and weaker than the onecalibrated. This value was selected because it closely captured thepost-peak variation of the 3-point tests, and the other parameterswere altered by the same quantity for consistency. Fig. 11h showsthe effect of the varied LDPM parameters in terms of penetrationdepth versus velocity plots for the ultra-high-performance CORTUFspecimen (UHPC), CORTUF 15% stronger (UHPC-U), CORTUF 15%weaker (UHPC-L), CORTUF with 3% and 5% VF (UHPC-F3, UHPC-F5),and a regular strength concrete (RSC). A typical RSC specimen wasused with an experimental compressive strength of 38.35 MPa, anexperimental tensile strength of 3.85 MPa, and a coarse aggregatesize of 8 mm. The parameters used for these specimens are given inTable 2.

Fig. 11h demonstrates the variations in penetration depth forvelocities below the ballistic limit of UHPC-F5 (1798 m/s). For thevarious CORTUF compositions, the penetration depth was onlysignificantly altered for velocities close to the ballistic limit. Theaddition of fibers did not significantly affect the penetration depthof the specimen as the reinforced CORTUF consistently had apenetration depth within the scatter of the unreinforced CORTUFwith strength 15% higher and lower than calibrated. Additionally,the weaker specimen was more affected by the increase in ve-locity, and the RSC had a greater penetration depth than theCORTUF. Moreover, a plot of exit versus strike velocities is shownin Fig. 11i for velocities above the ballistic limit. This plot follows asimilar trend with greatest effect on the different CORTUF com-posites around the ballistic limit. We can conclude that, accordingto these simulations, varying material strength by plus/minus 15%has only an appreciable effect at about the ballistic limit, and theaddition of fibers reduces the exit velocity. However, going from3% to 5% fibers does not incur significant retardation on the exitvelocity.

To determine which configuration best captured the failurepatterns observed as well as to investigate the sensitivity of thepenetration depth and perforation velocity to the applied con-straints, the slabs were modeled using various boundary condi-tions. Three states were studied for specimen size A.

1. Fixed Edge d only nodes along the edges of the panel wereselected, and they were fully constrained for displacement androtation in all directions (see Fig. 12a).

2. Fixed Side d all nodes along the width of the panel sides werefixed to prevent any displacement or rotation (see Fig. 12b). Thisis the boundary condition applied to the panel for the previoussimulations performed.

3. Free d no constraint was applied on the panel.

Fig. 12. Pictorial description of boundary conditions on panel for a) fixed edge, b) fix

Fig. 12c shows the depth of penetration with time plot for animpact velocity of 1676 m/s. Penetration is first completed by theprojectile impacting the free panel, whereas it takes the longest forthe fixed-side. The fixed-edge condition, most similar to the givenboundary conditions, closely follows the fixed side. However, forthe time given, the difference in penetration depth of the free- andthe fixed-sided panel is only about 5%. The reason the penetrationdepth is least in the fixed-side panel is due to the confinement theboundary provides for the panel, which cause resistance to theprojectile. In reality, the actual effects on the boundary are non-uniform, and the fixed-side and completely free scenarios repre-sent the limits bounding the true values of the boundaryconditions.

The crack pattern for the RSC, UHPC, and UHPC with increasingfiber content were also explored using LDPM-F, and the results aredisplayed in Fig. 13. The first row of results shows the scabbing forthe fixed-side boundary, the second row is the scabbing for thecompletely free-boundary condition, and the third row shows theside view for the fixed condition. Additionally, the first columnrepresents the views for the RSC, the second column shows theUHPC, and the third and fourth columns show the UHPC 3% and 5%VF, respectively. The crater size increased for the UHPC in com-parison to the RSC since the higher strengthmaterial with a smalleraggregate size has less inclusions to prevent the propagation of thecrack through the specimen. Also, the scabbing effects are greatlyreduced by the addition of fibers, but there is not much change asthe VF increases from 3 to 5%. For the free-edged panels, splittingradial cracks, which propagate from the main damage zone out-ward as expected, are intensified for the more brittle UHPC and arereduced by the effect of the fibers.

Comparison of the panel failure with the actual experimentaldata is shown in Fig. 14. Experimental and numerical results of theimpact face for the RSC is displayed in Fig. 14a and b, respectively.The red zone of the numerical results represents complete disin-tegration of the panel as seen in the experimental result. Figs. 14cand d show the experimental and numerical results for theUHPC-F3 impact face as well as the FSP penetrator damage. Thecrater size is reduced by the addition of the fibers in the numericalprediction and experimental results. However, the FSP penetratorobtained more damage than observed in the experimental resultsby modeling it as plastic. As previously investigated, this additionaldamage accrued to the projectile results in an overestimate of theballistic limit and a decrease in residual velocity.

This information is also useful in evaluating the disparitybetween the numerical and experimental ballistic limit. The nu-merical ballistic limit of the plain concrete Size A was 1706 m/s,while the ballistic limit of the fiber reinforced specimen, tested byERDC, was 1100 m/s and had a percent difference of 47% when

ed side; c) penetration depthetime plots for varied panel boundary conditions.

Page 16: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

Fig. 13. Comparison of scabbing for fixed-side boundary, free boundary, and side crater view of (a) RSC (b) UHPC (c)UHPC-VF3 (d)UHPC-5.

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e3228

compared to the numerical ballistic limit for UHPC-F3 of 1783 m/s.Experiments performed on the panel resulted in perforation of thethinnest specimen, Size A, with 92.1% attenuation of the velocity,and 100% attenuation for the larger slabs, B and C, for an impactvelocity of 1100 m/s. Based on experimental data on only onespecimen, it is not possible to judge whether or not the differencebetween the numerical and experimental values are within thealways-existing statistical scatter of experimental results. Addi-tionally, many other effects contribute to the disparity of experi-mental and numerical ballistic limits, some of which werepreviously discussed; among them are the effects of a plasticprojectile, boundary conditions of the panel, and the appreciable

Fig. 14. Projectile perforation comparison on impact face of NSC for a) experimental, and b) nnumerical tests.

difference in penetration depth observed for impact velocities closeto the ballistic limit for variations of CORTUF (see Fig.11h and i). Theeffect of coarse graining the panels and modeling a coarse-grainsize greater than 0.6 mm can also be a factor in determining theballistic limit.

Further investigation was performed on the CORTUF-Plainspecimen Sizes B and C to determine the effect on increasingpanel depth, and average ballistic limits of 2042 m/s and 2469 m/swere obtained for these specimens, respectively. Compared toSpecimen A with an average ballistic limit of 1706 m/s, increasingthe depth of the panel by 25% resulted in a 20% increase in ballisticlimit, and a panel with 50% increase inwidth leads to a 45% increase

umerical tests; CORTUF specimens with deformed projectile for c) experimental, and d)

Page 17: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

Fig. 15. Parametric analysis of panels A, B, and C showing a) penetration depthevelocity plot b) exitestrike velocity plot; comparison of c) scabbing, and d) side crater view of panelB for fixed-side boundary; comparison of e) scabbing, and f) side crater view of panel C for fixed-side boundary.

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e32 29

in ballistic limit. Fig. 15 shows a comparison of the three sizes, A, B,and C, for penetration depth versus velocity plots. For very lowimpact velocities, the penetration depth of each specimen is thesame. However, as the velocities increase to the ballistic limit, thethinner specimens, are able to perforate with a lower impactvelocity. Additionally, Fig. 15b shows that the exit velocity for thethinnest specimen is greater than the larger specimens and have asimilar trend as Panel B and C. Figs. 15cef show the failure patternsobserved for Panels B and C at their respective ballistic limits.

5. Conclusions

1. This paper successfully used LDPM-F to simulate the responsesof the novel ultra-high-performance concrete, CORTUF, by firstidentifying the parameters through quasi-static experiments,single pull-out, uniaxial compression, uniaxial strain, triaxialcompression and 3-point bending tests on CORTUF-Plain.

2. Without any change to previously identified parameters, LDPM-F was able to predict the following validation experiments forboth plain and fiber-reinforced CORTUF: uniaxial compression,uniaxial strain, hydrostatic compression, triaxial compression(confinements 10, 20, 50, 100, 200 and 300 MPa), 4-pointbending, and 3-point bending (notched).

3. LDPM-F was also able to simulate vital testing conditionsnecessary to accurately reproduce the experimental resultsincluding the ability to capture both random and preferentiallyoriented fibers and cast or cored specimens.

4. The model remarkably captures the intrinsic transition ofcompressive behavior from a very brittle failure for zero or lowconfinement to a ductile failure for high confinement in thetriaxial tests. Additionally, the model has the essential ability tocapture localized cracks and characteristic crack patterns forplain concrete and a crack-bridging effect through fiber rein-forcement for the CORTUF-Fiber model.

5. In-depth analysis of the failure mechanisms affecting CORTUFbehavior revealed the importance of accounting for micro-

splitting failure induced by the presence of hooks at fiber endsand the brittleness of the embedding matrix.

6. LDPM-F can be used confidently to investigate the behavior ofCORTUF under impact loading, penetration and perforationevents.

7. Simulations of projectile perforation showed the effect ofincreasing fiber content in the target specimens. Fibers added tothe concrete reduce brittleness and hence the area of the crater,but it has only a secondary effect on reducing the exit velocityassociatedwith strike velocities above the ballistic limit. Also, anincrease in fiber content in excess of 3% does not seem to pro-duce an appreciable improvement on the panel performance.

8. The reduced damaged area observed in CORTUF-Fiber comparedto CORTUF-Plain is an important feature associated with post-impact residual strength of CORTUF panels and their perfor-mance under multiple hit scenarios.

Acknowledgment

This effort was sponsored by the U.S. Army Engineer Researchand Development Center. Permission to publish was granted by theDirector, Geotechnical and Structures Laboratory. The work of thefirst and second authors was also supported under DTRA grantNo. HDTRA1-09-1-0029 to Rensselaer Polytechnic Institute (withsubcontracts to Northwestern University and Sandia National Labs).

A. LDPM-F Constitutive Equations

A.1. LDPM Constitutive Equations

Normal and shear stresses in the elastic regime of LDPM are:sN ¼ EN 3N; sM ¼ ET 3M; sL ¼ ET 3L, where EN ¼ E0, effective normalmodulus, ET ¼ aE0, and a ¼ shear-normal coupling parameter.

For stresses and strains beyond the elastic limit, LDPM meso-scale failure is characterized by three mechanisms as describedbelow.

Page 18: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e3230

Fracture and cohesion due to tension and tension-shear. Fortensile loading ( 3N > 0), the fracturing behavior is formulated

through an effective strain, 3¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi32N þ að 32M þ 32L Þ

q, and stress,

s ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffis2N þ ðsM þ sLÞ2=a

q, which define the normal and shear

stresses as sN ¼ 3N(s/ 3); sM ¼ a 3M(s/ 3); sL ¼ a 3L(s/ 3). The effectivestress s is incrementally elastic ð _s ¼ E0 _3Þ and must satisfy theinequality 0�s�sbt( 3,u) where sbt ¼ s0(u)exp[�H0(u)h 3� 30(u)i/s0(u)], hxi ¼ max{x,0}, and tanðuÞ ¼ 3N=a 3T ¼ sN

ffiffiffia

p=sT . The post

peak softening modulus is defined as H0ðuÞ ¼ Htð2u=pÞnt , whereHt is the softening modulus in pure tension (u ¼ p/2) expressed asHt ¼ 2E0/(lt/le � 1); lt ¼ 2E0Gt=s2t ; le is the length of the tetrahe-dron edge; and Gt is the meso-scale fracture energy. LDPM providesa smooth transition between pure tension and pure shear(u ¼ 0) with parabolic variation for strength given by

s0ðuÞ ¼ str2stð�sinðuÞ þffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffisin2ðuÞ þ 4acos2ðuÞ=r2st

qÞ=½2acos2ðuÞ�,

where rst ¼ ss/st is the ratio of shear strength to tensile strength.Compaction and pore collapse from compression. Normal stresses

for compressive loading ( 3N < 0) are computed through theinequality �sbc( 3D, 3V)�sN�0, where sbc is a strain-dependentboundary function of the volumetric strain, 3V, and the deviatoricstrain, 3D. Beyond the elastic limit, �sbc models pore collapse as alinear evolution of stress for increasing volumetric strain withstiffness Hc for � 3V� 3c1 ¼ kc0 3c0 : sbc ¼ sc0 þ h� 3V � 3c0iHc(rDV);Hc(rDV) ¼ Hc0/(1 þ kc2hrDV � kc1i); sc0 is the meso-scale compressiveyield stress; and kc1, kc2 are material parameters. Compaction andrehardening occur beyond pore collapse (� 3V� 3c1). In this case onehas sbc ¼ sc1(rDV) exp[(� 3V � 3c1)Hc(rDV)/sc1(rDV)] andsc1(rDV) ¼ sc0 þ ( 3c1 � 3c0)Hc(rDV).

Friction due to compression-shear. The incremental shear stressesare computed as _sM ¼ ET ð_3M � _3

pMÞ and _sL ¼ ET ð_3L � _3

pL Þ, where

_3pM ¼ _lv4=vsM , _3pL ¼ _lv4=vsL, and l is the plastic multiplier with

loading-unloading conditions 4 _l� 0 and _l� 0. The plastic potential

is defined as 4 ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffis2M þ s2L

q� sbsðsNÞ, where the nonlinear fric-

tional law for the shear strength is assumed to besbs ¼ ss þ (m0 � mN)sN0[1 � exp(sN/sN0)] � mNsN; sN0 is the tran-sitional normal stress; m0 and mN are the initial and final internalfriction coefficients.

A.2. FibereMatrix Interaction Behavior

The pull-out resistance of the fiber is related to the load, P andthe relative displacement/slippage, v, between concrete and fiber.Full debonding is represented by a critical slippage valuevd ¼ ð2s0L2e=Ef df Þ þ ð8GdL

2e=Ef df Þ1=2 where Le is a generic embed-

ment length, Gd is the bond fracture energy, and s0 is a constantfrictional stress. When v < vd (debonding),PðvÞ ¼ ½ðp2Ef d

3f ðs0vþ GdÞ=2�1=2. After full debonding (v > vd),

P(v) ¼ P0[1 � (v � vd/Le)][1 þ (b(v � vd)/df)] where P0 ¼ pLedfs0 andb ¼ 0;>0;<0 if the interfacial friction is independent, increases,decreases with slippage, respectively.

When the orientation of the embedded segment and the crackbridging segment of fiber are different, localized fracture and frag-mentation can occur reducing the embedded fiber length by a spal-ling length, sf ¼ PfNsin(q/2)/kspstdfcos2(q/2), where Pf is the crack-bridging force, q is the angle between the normals to the embeddedsegment and the crack face, st is themeso-scale tensile strength, 40

f isthe change in fiber orientation, and ksp is a material parameter.

The snubbing model adopted in LDPM-F assumes that the fiber,in a perfectly flexible manner, wraps around the exit of the tunnelcrack that was shortened by spalling. The change in the tensile loadfor the flexible tendon is expressed as Pf ¼ expðksn40

f ÞPðvÞwhere P

is the summation of all slip-friction and debonding forces parallelto the embedment length, and ksn is the snubbing parameter. For aPf leading to fiber rupture, there will be no contribution from thefibers to structural strength and ductility. Single fiber pull-out testson polyvinyl alcohol (PVA) revealed lower rupture loads forincreasing 40

f . This strength reduction (due to bending) is adoptedin LDPM-F, and expressed as sf � suf e

�krup40f where sf is the axial

stress in the crack-bridging segment, krup is a material parameterand suf is the ultimate tensile strength of the fiber.

A detailed description of the LDPM idealization of concretemeso-structure can be found in Cusatis et al. [85] and more infor-mation on the fiber stiffness and unloading adopted in LDPM-F isprovided in Ref. [88].

References

[1] Yazici H, Yardimici MY, Aydin S, Karabulut A. Mechanical properties ofreactive powder concrete containing mineral admixtures under differentcuring regimes. Constr Build Mater 2009;23(3):1223e31.

[2] Ming-Gin L, Yung-Chih W, Chui-Te C. A preliminary study of reactive powderconcrete as a new repair material. Constr Build Mater 2007;21(1):182e9.

[3] Zhang M, Shim V, Lu G, Chew C. Resistance of high-strength concrete toprojectile impact. Int J Impact Eng 2005;31(7):825e41.

[4] Wittmann FH. Crack formation and fracture energy of normal and highstrength concrete. Sadhana 2002;27(4):413e23.

[5] Habel K, Viviani M, Denari E, Bruhwiler E. Development of the mechanicalproperties of an ultra-high performance fiber reinforced concrete (UHPFRC).Cem Concr Res 2006;36(7):1362e70.

[6] Meddah MS, Zitouni S, Belâabes S. Effect of content and particle size distri-bution of coarse aggregate on the compressive strength of concrete. ConstrBuild Mater Apr. 2010;24:505e12.

[7] Corinaldesi V, Moriconi G. Mechanical and thermal evaluation of ultra highperformance fiber reinforced concretes for engineering applications. ConstrBuild Mater 2012;26(1):289e94.

[8] Sajedi F, Razak HA. Effects of thermal and mechanical activation methods oncompressive strength of ordinary Portland cement “slag mortar”. Mater DesFeb. 2011;32:984e95.

[9] Sajedi F. Effect of curing regime and temperature on the compressivestrength of cement-slag mortars. Constr Build Mater Nov. 2012;36:549e56.

[10] Razak HA, Sajedi F. The effect of heat treatment on the compressive strengthof cement-slag mortars. Mater Des Sept. 2011;32:4618e28.

[11] Kang S-T, Lee Y, Park Y-D, Kim J-K. Tensile fracture properties of an ultra highperformance fiber reinforced concrete (UHPFRC) with steel fiber. ComposStruct 2010;92:61e71.

[12] Zhang MH, Sharif MSH, Lu G. Impact resistance of high-strength fibre-rein-forced concrete. Mag Concr Res 2007;9831(3):199e210.

[13] Naaman AE, Reinhardt HW. Strain hardening and deflection hardening fiberreinforced cement composites HPFRCC-4. Mater Struct 2003;30:757e9.

[14 Dancygier AN, Savir Z. Flexural behavior of HSFRC with low reinforcementratios. Eng Struct 2006;28:1503e12.

[15] Liang NH, Liu XR, Sun J. Experimental study of splitting tensile for multi-scalepolypropylene fiber concrete. Adv Mater Res Jan 2012;450e451:168e73.

[16] Yi N-H, Kim J-HJ, Han T-S, Cho Y-G, Lee JH. Blast-resistant characteristics ofultra-high strength concrete and reactive powder concrete. Constr BuildMater 2012;28(1):694e707.

[17] Corbett GG, Reid ISR, Johnson W. Impact loading of plates and shells by free-flying projectiles: a review. Int J Impact Eng 1996;18(2):141e230.

[18] Li QM, Tong DJ. Perforation thickness and ballistic limit of concrete targetsubjected to rigid projectile impact. J Eng Mech 2003;0(September):1083e92.

[19] Tai Y. Flat ended projectile penetrating ultra-high strength concrete platetarget. Theor Appl Fract Mech Apr. 2009;51:117e28.

[20] Nyström U, Gylltoft K. Comparative numerical studies of projectile im-pacts on plain and steel-fibre reinforced concrete. Int J Impact Eng2011;38:95e105.

[21] Dancygier AN, Yankelevsky DZ. High strength concrete response to hardprojectile impact. Int J Impact Eng 1996;18(6):583e99.

[22] Neil EFO, Neeley BD, Cargile JD. Tensile properties of very-high-strengthconcrete for penetration-resistant structures. Shock Vib 1999;6:237e45.

[23] Mainstone R. Properties of materials at high rates of straining or loading.Matr Constr 1975;8(2):102e16.

[24] Dilger WH, Koch R, Kowalczyk R. Ductility of plain and confined concreteunder different strain rates. J Am Concrete Inst 1984;81(1):73e81.

[25] Bischoff PH, Perry SH. Compressive behaviour of concrete at high strain rates.Mater Struct 1991;24:425e50.

[27] Watstein D. Effect of straining rate on the compressive strength and elasticproperties of concrete. ACI 1953;49(4):729e44.

[28] Hughes BP, Gregory R. Concrete subjected to high rates of loading incompression. Mag Contr Res 1972;24(78):25e36.

Page 19: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e32 31

[29] Malvern LE, Tang T, Jenkins DA, Gong JC. Dynamic compressive strength ofcementitious materials. Mater Res Soc 1985;64(1):119e38.

[30] Abrams DA. Effect of rate application of load on the compressive strength ofconcrete. ASTM 1917;17(2):364e77.

[31] Lundeen RL. Dynamic and static tests of plain concrete specimens: report 1.Tech. rep. Vicksburg, Mississippi: US Army Engineers Waterways ExperimentStation; 1963. p. 1e38

[32] Mechtcherine V, Millon O, Butler M, Thoma K. Mechanical behavior of SHCCunder impact loading. In: Parra-Montesinos G, Reinhardt H, Naaman A, ed-itors. High performance fiber reinforced cement composites 6. Vol. 2 ofRILEM state of the art reports. Netherlands: Springer; 2012. p. 297e304.

[33] Wang S, Zhang M-H, Quek ST. Effect of high strain rate loading oncompressive behaviour of fibre-reinforced high-strength concrete. MagConcr Res 2011;63(11):813e27.

[34] Wang Y, Wang Z, Liang X, An M. Experimental and numerical studies ondynamic compressive behavior of reactive powder concretes. Acta Mech SolSin 2008;21(5):420e30.

[35] Kim DJ, El-Tawil S, Naaman AE. Rate-dependent tensile behavior of highperformance fiber reinforced cementitious composites. Mater Struct May2008;42:399e414.

[36] Caverzan A, Cadoni E, di Prisco M. Dynamic tensile behaviour of high per-formance fibre reinforced cementitious composites after high temperatureexposure. Mech Mater Apr. 2013;59:87e109.

[37] Cusatis G. Strain-rate effects on concrete behavior. Int J Impact Eng Apr.2011;38:162e70.

[38] O�zbolt J, Rah KK, Me�strovi&cacute D. Influence of loading rate on concretecone failure. Int J Fract May 2006;139:239e52.

[39] O�zbolt J, Sharma A. Numerical simulation of dynamic fracture of con-crete through uniaxial tension and l-specimen. Eng Fract Mech 2012;85:88e102.

[40] Powers TC. Mechanism of shrinkage and reversible creep of hardenedcement paste. Cem Concr Assoc 1965:319e44.

[41] Powers TC. The thermodynamics of volume change and creep. Mat Struct1968;1(6):487e507.

[42] Ba�zant Z, Cusatis G, Cedolin L. Temperature effect on concrete creep modeledby microprestress-solidification theory. J Eng Mech June 2004;130:691e9.

[43] Wittmann FH. Interaction of hardened cement paste and water. J Am CeramSoc 1973;56(8):409e15.

[44] Ba�zant ZP, Gettu R. Rate effects and load relaxation in the static fracture ofconcrete. ACI Mater J 1992;89(5):456e8.

[45] Ba�zant Z, Caner F, Adley M, Akers S. Fracturing rate effect and creep inmicroplane model for dynamics. J Eng Mech 2000;126(9):962e70.

[46] Cusatis G, Ba�zant Z, Cedolin L. 3D Lattice model for dynamic simulations ofcreep, fracturing and rate effect in concrete. In: Ulm FJ, Bazant ZP,Wittmann FH, editors. Creep, shrinkage and durability mechanics ofconcrete and other quasi-brittle materials. Proceedings of the 6th interna-tional conference CONCREEP-6. MIT, Cambridge, MA. USA: Elsevier; Aug2001. p. 113e8.

[47] Forquin P, Safa K, Gary G. Influence of free water on the quasi-static anddynamic strength of concrete in confined compression tests. Cem Concr ResFeb. 2010;40:321e33.

[48] Zheng D, Li Q. An explanation for rate effect of concrete strength based onfracture toughness including free water viscosity. Eng Fract Mech Nov.2004;71:2319e27.

[49] Cottrell A. The mechanical properties of matterIn Wiley series on the scienceand technology of materials. NY: Krieger Publishing Company; 1981. p. 300e430.

[50] Vegt I, Weerheijm J. Influence of moisture content on the dynamic behaviourof concrete. In: Gdoutos EE, editor. Fracture of nano and engineering ma-terials and structures, (Greece). Springer Netherlands. July 2006. p. 501e2.

[51] Cadoni E, Labibes K, Albertini C, Berra M, Giangrasso M. Strain-rate effect onthe tensile behaviour of concrete at different relative humidity levels2001;34(February):21e6.

[52] Ba�zant Zden�ek P, Caner Ferhun C. Comminution of solids caused by kineticenergy of high shear strain rate, with implications for impact, shock, andshale fracturing, PNAS; 2013 [Epub ahead of print November 11, 2013],http://dx.doi.org/10.1073/pnas.1318739110.

[53] Cadoni E, Solomos G, Albertini C. Concrete behaviour in direct tension testsat high strain rates. Mag Concr Res June 2013;65:660e72.

[54] O�zbolt J, Bo�snjak J, Sola E. Dynamic fracture of concrete compact tensionspecimen: experimental and numerical study. Int J Sol Struct December2013;50(25e26):4270e8.

[55] Ramulu M, Kobayashi AS. Mechanics of crack curving and branchingdadynamic fracture analysis. Int J Fract 1985;27:187e201.

[56] Barenblatt GI, Salganik RL, Cherepanov GP. On the nonsteady propagation ofcracks. PMM 1962;26(2):328e34.

[57] Hillerborg A, Modéer M, Petersson PE. Analysis of crack formation and crackgrowth in concrete by means of fracture mechanics and finite elements. CemConcr Res 1976;6.

[58] Rashid Y. Ultimate strength analysis of prestressed concrete pressure vessels.Nucl Eng Des 1968;7(4):334e44.

[59] Ba�zant ZP, Oh BH. Crack band theory for fracture of concrete. Matér ConstrMay 1983;16:155e77.

[60] Bicanic N, Pearce CJ. Computational aspects of a softening plasticity modelfor plain concrete. Mech Cohesive-frictional Mater 1996;1(1):75e94.

[61] Comi C, Perego U. Fracture energy based bi-dissipative damage model forconcrete. Int J Sol Struct 2001;38(3637):6427e54.

[62] Grassl P, Jirásek M. Damage-plastic model for concrete failure. Int J Sol Struct2006;43:7166e96.

[63] Mazars J, Pijaudier-Cabot G. Continuum damage theory e application toconcrete. J Eng Mech 1989;115(2):345e65.

[64] Ba�zant ZP, Caner FC, Carol I, Adley MD, Akers AA. Microplane model M4 forconcrete. I: formulation with work-conjugate deviatoric stress. J Eng Mech2000;126(9).

[65] Di Luzio G, Cusatis G. Solidification-microprestress-microplane (smm) theoryfor concrete at early age. Theory, validation and application. Int J Sol Structmarch 2013;50:957e75.

[66] Caner F, Ba�zant Z. Microplane model M7 for plain concrete: I. Formulation.J Eng Mech 2012:1e32. http://dx.doi.org/10.1061/(ASCE)EM.1943-7889.0000570.

[67] O�zbolt J, Li Y, Kozar I. Microplane model for concrete with relaxed kinematicconstraint. Int J Sol Struct 2001;38(16):2683e711.

[68] Di Luzio G. A symmetric over-nonlocal microplane model m4 for fracture inconcrete. Int J Sol Struct 2007;44(13).

[69] Cedolin L, Ba�zant ZP. Effect of finite element choice in blunt crack bandanalysis. Comput Methods Appl Mech Eng 1980;24:305e16.

[70] De Borst R, Sluys LJ, Muhlhaus HB, Pamin J. Fundamental issues in finiteelement analyses of localization of deformation. Eng Comput 1993;10:99e122.

[71] Ba�zant ZP, Planas J. Fracture and size effect in concrete and other quasibrittlematerials. Boca Raton: CRC Press LLC; 1997. p. 489e560.

[72] Ba�zant ZP, Jirásek M. Nonlocal integral formulations of plasticity and dam-age: survey of progress. J Eng Mech 2002;128(11):1119.

[73] Walsh P. Fracture of plain concrete. Indian Concr J 1972;46.[74] Ba�zant ZP, Kazemi MT. Determination of fracture energy, process zone length

and brittleness number from size effect, with application to rock and con-crete. Int J Fract 1990;44.

[75] Ba�zant ZP, O�zbolt J. Nonlocal microplane model for fracture, damage, andsize effect in structures. J Eng Mech 1990;116(11):2485e505.

[76] Leong HP, Somsak S. Over-nonlocal gradient enhanced plastic-damagemodel for concrete. Int J Sol Struct Dec 2009;46:4369e78.

[77] Herrmann H, Hansen A, Roux S. Fracture of disordered, elastic lattices in twodimensions. Phys Rev B 1989;39(1):637e47.

[78] Ba�zant ZP, Tabbara MR, Kazemi MT, Pijaudier-Cabot G. Random particlemodel for fracture of aggregate or fiber composites. J Eng Mech 1990;116(8):1686.

[79] Schlangen E, Van Mier J. Experimental and numerical analysis of micro-mechanisms of fracture of cement-based composites. Cem Concr Compos1992;14(2):105e18.

[80] Cusatis G, Ba�zant ZP, Cedolin L. Confinement shear lattice model for concretedamage in tension and compression. I: Theory. J Eng Mech 2003;129(12):1439e48.

[81] Cusatis G, Ba�zant ZP, Cedolin L. Confinement-shear lattice model for concretedamage in tension and compression. II: numerical implementation andvalidation. J Eng Mech 2003;129(12):1449e58.

[82] Gianluca C, Luigi C. Two-scale study of concrete fracturing behavior. EngFract Mech 2007;74(1e2):3e17. Fracture of Concrete Materials andStructures.

[83] Eliá�s J, Stang H. Lattice modeling of aggregate interlocking in concrete. Int JFract Mar. 2012;175:1e11.

[84] Cusatis G, Nakamura H. Discrete modeling of concrete materials and struc-tures. Cem Concr Compos 2011;33(9):865e6.

[85] Cusatis G, Pelessone D, Mencarelli A. Lattice discrete particle model (LDPM)for failure behavior of concrete. I: Theory. Cem Concr Compos 2011;33(9):881e90.

[86] Cusatis G, Mencarelli A, Pelessone D, Baylot J. Lattice discrete particle model(LDPM) for failure behavior of concrete. II: Calibration and validation. CemConcr Compos 2011;33(9):891e905.

[87] Alnaggar M, Cusatis G, Di Luzio G. Lattice discrete particle modeling of alkali-silica-reaction (ASR) deterioration of concrete structures. J Eng Mech2013;41:45e59.

[88] Schauffert EA, Cusatis G. Lattice discrete particle model for fiber reinforcedconcrete (LDPM-F): I. Theory. ASCE J Eng Mech July 2012;138:826e33.

[89] Schauffert EA, Cusatis G, Pelessone D, O’Daniel JL, Baylot JT. Latticediscrete particle model for fiber reinforced concrete (LDPM-F): II. Tensilefracture and multiaxial loading behavior. ASCE J Eng Mech 2012;138(7):834e41.

[90] Cusatis G, Schauffert EA. Discontinuous cell method (DCM) for cohesivefracture propagation. In: Oh BH, editor. 7th international conference onfracture mechanics of concrete and concrete structures (FraMCos 7), (Jeju,South Korea). Korea Concrete Institute. May 2010. p. 23e8.

[91] Yang E-H, Wang S, Yang Y, Li VC. Fiber-bridging constitutive law ofengineered cementitious composites. J Adv Concr Technol 2008;6(1):181e93.

[92] Cusatis G, Mencarelli A, Pelessone D, Baylot J. Lattice discrete particle model (LDPM ) for fracture dynamics and rate effect in concrete. ASCE 2008;315(42):1e11.

[93] Mencareli A. Numerical simulation of the effect of blast and penetration onreinforced concrete structures [PhD thesis]. Troy, NY: Rensselaer PolytechnicInstitute; December 2010. p. 1e147.

Page 20: Discrete modeling of ultra-high-performance concrete with ...anil-lab/others/lectures/CMCE/24-LDPM-UHPC-… · Discrete modeling of ultra-high-performance concrete with application

J. Smith et al. / International Journal of Impact Engineering 65 (2014) 13e3232

[94] Cusatis G. The lattice discrete particle model (LDPM) for the numericalsimulation of concrete behavior subject to penetration. Hoboken, NJ, USA:John Wiley and Sons, Inc; 2013. p. 369e87.

[95] Cusatis G, Diluzio G, Cedolin L. Meso-scale simulation of concrete: blastand penetration effects and AAR degradation. In: Cadoni E, di Prisco M,editors. Performance, protection and strengthening of structures underextreme loadingdapplied mechanics and materials. Switzerland: TransTech Publications; 2011. p. 75e80.

[98] Williams EM, Graham SS, Reed PA, Rushing TS. Laboratory characterizationof cortuf concrete with and without steel fibers. Technical Report. Vicksburg,MS: US Army Corps of Engineers, Engineer Research and DevelopmentCenter; May 2009. p. 1e83.

[99] Roth MJ, Rushing TS, Flores OG, Sham DK, Stevens JW. Laboratory investi-gation of the characterization of cortuf flexural and splitting tensile prop-erties. Technical Report. Vicksburg, MS: US Army Corps of Engineers,Engineer Research and Development Center; Sep 2009. p. 1e57.

[100] Landis EN, Keane DT. X-ray microtomography. Mater Charact 2010;61(12):1305e16.

[101] Alnaggar M, Cusatis G. Automatic parameter identification of discretemesoscale models with application to the coarse-grained simulation ofreinforced concrete structures. In: Carrato J, Burns J, editors. 2012 structurescongress, (Chicago, IL). ASCE. 2012. p. 406e17.

[102] Maeda M, Otani S, Aoyama H. Effect of confinement on bond splittingbehavior in reinforced concrete beams. IABSE, Struct Eng Int 1995;5(3):166e71.

[103] Persson B. Poisson’s ratio of high-performance concrete. Cem Concr Res Oct.1999;29:1647e53.

[104] PelessoneD.MARS,modeling and analysis of the response of structuresd user’smanual. San Diego, CA: ES3; 2011. p. 1e237. http://es3inc.com/marssolver/.

[105] Mumm DR, Faber KT. Interfacial debonding and sliding in brittle-matrixcomposites measured using an improved fiber pull out technique. ActaMetall Mater 1995;43(3):1259e70.

[106] Li V, Wang Y, Backer S. Effect of inclining angle, bundling and surfacetreatment on synthetic fibre pull-out from a cement matrix. Composites Mar.1990;21:132e40.

[107] Lee Y, Kang S-T, Kim J-K. Pullout behavior of inclined steel fiber in an ultra-high strength cementitious matrix. Constr Build Mater Oct. 2010;24:2030e41.

[108] Laranjeira F, Molins C, Aguado a. Predicting the pullout response of inclinedhooked steel fibers. Cem Concr Res Oct. 2010;40:1471e87.

[109] Kang S-T, Kim J-K. The relation between fiber orientation and tensilebehavior in an ultra high performance fiber reinforced cementitious com-posites (UHPFRCC). Cem Concr Res Oct. 2011;41:1001e14.

[110] Hibbit, Karlsson, Sorensen. ABAQUS/standard analysis user’s manual. USA:Hibbit, Karlsson, Sorensen Inc; 2007. p. 1e1051.

[111] Lucy L. A numerical approach to the testing of fission hypothesis. Astron J1977;82:1013e24.

[112] Gingold RA, Monaghan JJ. Smoothed particle hydro-dynamics: theory andapplications to non-spherical stars. Mon Not R Astron Soc Nov 1977;181:375e89.

[113] Rabczuk T, Eibl J. Modelling dynamic failure of concrete with meshfreemethods. Int J Impact Eng Nov. 2006;32:1878e97.

[114] Rabczuk T, Xiao SP, Sauer M. Coupling of meshfree methods withfinite elements: basic concepts and test results. Commun Numer MethodsEng. 2006. 1031e65. John Wiley & Sons, Ltd.

[115] Rabczuk T, Eibl J. Simulation of high velocity concrete fragmentation usingSPH/MLSPH. Int J Numer Methods Eng Mar. 2003;56:1421e44.

[116] Rabczuk T, Belytschko T. A three-dimensional large deformation meshfreemethod for arbitrary evolving cracks. Comput Methods Appl Mech Eng May2007;196:2777e99.

[117] Rabczuk T, Belytschko T. Cracking particles: a simplified meshfree methodfor arbitrary evolving cracks. Int J Numer Methods Eng 2004;61(13):2316e43.