16
SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES OF STRUCTURAL COMPOSITE LUMBER Laszlo Bejo Senior Research Scientist University of Western Hungary H-9400 Sopron, Ady E. u. 5. Hungary Formerly, Graduate Research Assistant WVU, Division of Forestry and Elemer M. LangAssociate Professor West Virginia University Division of Forestry Morgantown, WV 26506 -6125 (Received September 2003) abstract Structural composite lumber (SCL) products were introduced into the construction practice several de- cades ago. Their apparent advantages over traditional lumber did not generate copious research interests. However, increasing demands for structural materials coupled with the decreasing quality and quantity of raw materials are forcing the industry to introduce short rotation trees or species having unfavorable prop- erties into the manufacturing processes. Consequently, there is a need for research to further enhance the effective use of renewable natural resources. This article describes the development of simulation models that estimate the bending and orthotropic compression modulus of elasticity (MOE) of laminated veneer lumber (LVL) and parallel strand lumber (PSL). The Monte Carlo simulation-based routines use the physical/mechanical properties of primary con- stituting elements, obtained from probability distributions, to calculate a particular property of the com- posite system. Furthermore, the orthotropic behavior of the wood constituents due to their position in the composite is modeled by well-established theoretical/empirical equations. Results and experimental vali- dation regarding the geometric, physical, and mechanical attributes showed reasonably good agreement between simulated and experimental values. Developed models have good potential for predicting the elastic parameters of composites using new raw materials or novel design features. Keywords: Wood composites, elastic properties, simulation modeling. introduction Traditionally, solid wood is the raw material most frequently used for residential construction in the United States. Environmental concerns and the declining quality and quantity of timber re- sources have driven the industry toward alterna- tive solutions. Such pressures triggered the development of structural wood-based composite materials. These are excellent substitutes for solid wood in many aspects: they are lightweight, strong, and durable. Moreover, some of the species/products combinations exhibit higher and more consistent mechanical properties due to the densification during the consolidation process; and they demonstrate better dimensional stability than those of solid wood. Many of their proper- ties can also be engineered to a certain extent. Wood and Fiber Science, 36(3), 2004, pp. 395 – 410 © 2004 by the Society of Wood Science and Technology † Member of SWST.

SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES OFSTRUCTURAL COMPOSITE LUMBER

Laszlo BejoSenior Research Scientist

University of Western HungaryH-9400 Sopron, Ady E. u. 5.

HungaryFormerly,

Graduate Research AssistantWVU, Division of Forestry

and

Elemer M. Lang†Associate Professor

West Virginia UniversityDivision of Forestry

Morgantown, WV 26506-6125

(Received September 2003)

abstract

Structural composite lumber (SCL) products were introduced into the construction practice several de-cades ago. Their apparent advantages over traditional lumber did not generate copious research interests.However, increasing demands for structural materials coupled with the decreasing quality and quantity ofraw materials are forcing the industry to introduce short rotation trees or species having unfavorable prop-erties into the manufacturing processes. Consequently, there is a need for research to further enhance theeffective use of renewable natural resources.

This article describes the development of simulation models that estimate the bending and orthotropiccompression modulus of elasticity (MOE) of laminated veneer lumber (LVL) and parallel strand lumber(PSL). The Monte Carlo simulation-based routines use the physical/mechanical properties of primary con-stituting elements, obtained from probability distributions, to calculate a particular property of the com-posite system. Furthermore, the orthotropic behavior of the wood constituents due to their position in thecomposite is modeled by well-established theoretical/empirical equations. Results and experimental vali-dation regarding the geometric, physical, and mechanical attributes showed reasonably good agreementbetween simulated and experimental values. Developed models have good potential for predicting theelastic parameters of composites using new raw materials or novel design features.

Keywords: Wood composites, elastic properties, simulation modeling.

introduction

Traditionally, solid wood is the raw materialmost frequently used for residential constructionin the United States. Environmental concerns andthe declining quality and quantity of timber re-sources have driven the industry toward alterna-tive solutions. Such pressures triggered the

development of structural wood-based compositematerials. These are excellent substitutes forsolid wood in many aspects: they are lightweight,strong, and durable. Moreover, some of thespecies/products combinations exhibit higher andmore consistent mechanical properties due to thedensification during the consolidation process;and they demonstrate better dimensional stabilitythan those of solid wood. Many of their proper-ties can also be engineered to a certain extent.

Wood and Fiber Science, 36(3), 2004, pp. 395–410© 2004 by the Society of Wood Science and Technology

† Member of SWST.

Used Mac Distiller 5.0.x Job Options
This report was created automatically with help of the Adobe Acrobat Distiller addition "Distiller Secrets v1.0.5" from IMPRESSED GmbH. You can download this startup file for Distiller versions 4.0.5 and 5.0.x for free from http://www.impressed.de. GENERAL ---------------------------------------- File Options: Compatibility: PDF 1.4 Optimize For Fast Web View: No Embed Thumbnails: Yes Auto-Rotate Pages: No Distill From Page: 1 Distill To Page: All Pages Binding: Left Resolution: [ 1200 1200 ] dpi Paper Size: [ 495 720 ] Point COMPRESSION ---------------------------------------- Color Images: Downsampling: No Compression: Yes Automatic Selection of Compression Type: Yes JPEG Quality: << /VSamples [ 1 1 1 1 ] /HSamples [ 1 1 1 1 ] /Blend 1 /QFactor 0.25 /ColorTransform 1 >> Bits Per Pixel: As Original Bit Grayscale Images: Downsampling: No Compression: Yes Automatic Selection of Compression Type: Yes JPEG Quality: << /VSamples [ 1 1 1 1 ] /HSamples [ 1 1 1 1 ] /Blend 1 /QFactor 0.25 /ColorTransform 1 >> Bits Per Pixel: As Original Bit Monochrome Images: Downsampling: No Compression: Yes Compression Type: CCITT CCITT Group: 4 Anti-Alias To Gray: No Compress Text and Line Art: Yes FONTS ---------------------------------------- Embed All Fonts: Yes Subset Embedded Fonts: No When Embedding Fails: Cancel Job Embedding: Always Embed: [ ] Never Embed: [ ] COLOR ---------------------------------------- Color Management Policies: Color Conversion Strategy: Leave Color Unchanged Intent: Default Device-Dependent Data: Preserve Overprint Settings: Yes Preserve Under Color Removal and Black Generation: No Transfer Functions: Remove Preserve Halftone Information: No ADVANCED ---------------------------------------- Options: Use Prologue.ps and Epilogue.ps: No Allow PostScript File To Override Job Options: No Preserve Level 2 copypage Semantics: Yes Save Portable Job Ticket Inside PDF File: No Illustrator Overprint Mode: Yes Convert Gradients To Smooth Shades: Yes ASCII Format: No Document Structuring Conventions (DSC): Process DSC Comments: Yes Log DSC Warnings: No Resize Page and Center Artwork for EPS Files: No Preserve EPS Information From DSC: No Preserve OPI Comments: No Preserve Document Information From DSC: Yes OTHERS ---------------------------------------- Distiller Core Version: 5000 Use ZIP Compression: Yes Deactivate Optimization: No Image Memory: 524288 Byte Anti-Alias Color Images: No Anti-Alias Grayscale Images: No Convert Images (< 257 Colors) To Indexed Color Space: Yes sRGB ICC Profile: sRGB IEC61966-2.1 END OF REPORT ---------------------------------------- IMPRESSED GmbH Bahrenfelder Chaussee 49 22761 Hamburg, Germany Tel. +49 40 897189-0 Fax +49 40 897189-71 Email: [email protected] Web: www.impressed.de
Adobe Acrobat Distiller 5.0.x Job Option File
<< /ColorSettingsFile () /LockDistillerParams true /DetectBlends true /DoThumbnails true /AntiAliasMonoImages false /MonoImageDownsampleType /Bicubic /GrayImageDownsampleType /Bicubic /MaxSubsetPct 100 /MonoImageFilter /CCITTFaxEncode /ColorImageDownsampleThreshold 1.5 /GrayImageFilter /DCTEncode /ColorConversionStrategy /LeaveColorUnchanged /CalGrayProfile (Gray Gamma 2.2) /ColorImageResolution 350 /UsePrologue false /MonoImageResolution 1200 /ColorImageDepth -1 /sRGBProfile (sRGB IEC61966-2.1) /PreserveOverprintSettings true /CompatibilityLevel 1.4 /UCRandBGInfo /Remove /EmitDSCWarnings false /CreateJobTicket false /DownsampleMonoImages false /DownsampleColorImages false /MonoImageDict << /K -1 >> /ColorImageDownsampleType /Bicubic /GrayImageDict << /HSamples [ 2 1 1 2 ] /VSamples [ 2 1 1 2 ] /Blend 1 /QFactor 0.9 >> /CalCMYKProfile (U.S. Web Coated (SWOP) v2) /ParseDSCComments true /PreserveEPSInfo false /MonoImageDepth -1 /AutoFilterGrayImages true /SubsetFonts false /GrayACSImageDict << /VSamples [ 1 1 1 1 ] /HSamples [ 1 1 1 1 ] /Blend 1 /QFactor 0.25 /ColorTransform 1 >> /ColorImageFilter /DCTEncode /AutoRotatePages /None /PreserveCopyPage true /EncodeMonoImages true /ASCII85EncodePages false /PreserveOPIComments false /NeverEmbed [ ] /ColorImageDict << /HSamples [ 2 1 1 2 ] /VSamples [ 2 1 1 2 ] /Blend 1 /QFactor 0.9 >> /AntiAliasGrayImages false /GrayImageDepth -1 /CannotEmbedFontPolicy /Error /EndPage -1 /TransferFunctionInfo /Remove /CalRGBProfile (sRGB IEC61966-2.1) /EncodeColorImages true /EncodeGrayImages true /ColorACSImageDict << /VSamples [ 1 1 1 1 ] /HSamples [ 1 1 1 1 ] /Blend 1 /QFactor 0.25 /ColorTransform 1 >> /Optimize false /ParseDSCCommentsForDocInfo true /GrayImageDownsampleThreshold 1.5 /MonoImageDownsampleThreshold 1.5 /AutoPositionEPSFiles false /GrayImageResolution 300 /AutoFilterColorImages true /AlwaysEmbed [ ] /ImageMemory 524288 /OPM 1 /DefaultRenderingIntent /Default /EmbedAllFonts true /StartPage 1 /DownsampleGrayImages false /AntiAliasColorImages false /ConvertImagesToIndexed true /PreserveHalftoneInfo false /CompressPages true /Binding /Left >> setdistillerparams << /PageSize [ 576.0 792.0 ] /HWResolution [ 1200 1200 ] >> setpagedevice
Page 2: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

Manufacturers of wood-based structural com-posites are constantly seeking to improve themechanical properties of their products. The tradi-tional trial-and-error approach is costly, time-consuming, and often not feasible. A betteralternative is using simulation modeling to opti-mize the various manufacturing parameters (Lawand Kelton 1991). Using a reliable probability-based simulation model, one can demonstrate theeffects of changing properties or relative positionof the constituting materials on the final perfor-mance of the products without having to run ac-tual manufacturing trials.

A long-term research explored the orthotropicmechanical properties of five hardwood species(Lang et al. 2000, 2002, 2003) at WVU, Divisionof Forestry and University of Western Hungary.Species included: red oak (Quercus rubra), turkeyoak (Quercus cerris), quaking aspen (Populustremuloides), true poplar (Populus x. Euroameri-cana cv. Pannonia), and yellow-poplar (Lirioden-dron tulipifera) as potential raw materials forstructural composite manufacture.

As the concluding part of the comprehensiveproject, the objectives of this study were to de-velop and validate probability-based simulationmodels that can describe the bending and or-thotropic compression MOE of parallel strandlumber (PSL) and laminated veneer lumber(LVL), based on the geometry and orthotropicelastic parameters of their constituents. The sto-chastic parameters of three North Americanspecies were incorporated into simulation mod-els. Such simulation routines may provide an in-expensive tool for evaluating the effect of novelraw materials or other modifications on the ex-pected performance of these structural compositeproducts. Moreover, the models were validatedby comparing their predictions to properties mea-sured on commercial LVL and PSL producedfrom yellow-poplar structural veneers.

literature review

Modeling physical and mechanical propertiesrequires a thorough understanding of spatialstructure of the composites. An early simulationmodel described the structure of paper as consist-

ing of several layers of fibers and interfibrillarspaces or pores (Kallmes and Corte 1960;Kallmes et al. 1961). This work provided a basisfor developing a mathematical model that depictsrandomly packed, short-fiber-type wood com-posites (Steiner and Dai 1993; Dai and Steiner1994a, b.) Results of this investigation were usedin a Monte Carlo simulation program that canmodel different types of mats, and analyze themfor various important geometric characteristics(Lu et al. 1998). The program can also determinethe effect of sampling zone size on the measureddensity distribution. Harris and Johnson (1982)dealt with the characterization of flake orienta-tion in flakeboards. They pointed out that un-bounded distributions are not appropriate for thispurpose and suggested a bounded distribution toprovide angles between 0 and � radians.

Some researchers attempted to provide a de-tailed explanation and to simulate certain aspectsof particle mat behavior during consolidation.Suchsland (1967) summarized the mat formation,heat- and moisture movement and stress-behaviorof particleboard mats. He provided an explanationfor the formation of horizontal and vertical densitydistributions, and showed how pressing parametersinfluence the latter. Humprey and Bolton (1989)made an in-depth analysis of the multidimensionalunsteady-state heat and moisture transfer duringhot pressing. They built a model based on a modi-fied finite difference approach that could predicttemperature, moisture content, vapor pressure, andrelative humidity in different layers of a mat.

Several works dealt with the compression be-havior of strand mats throughout the pressurecycle. Dai and Steiner (1993) developed a theoret-ical model to describe the compression responseof randomly formed wood flake mats. Their pre-dictions agreed with experimental results reason-ably well. Two further models, using somewhatdifferent approaches to reconstruct the OSB matstructure and to predict the stress-strain relation-ship during consolidation, provided improved es-timation of stress development (Lang and Wolcott1996a,b; Lenth and Kamke 1996a,b.) It has beenproposed that a combination of these two modelscharacterizes the entire stress-strain curve some-what better than the earlier approaches.

396 WOOD AND FIBER SCIENCE, JULY 2004, V. 36(3)

Page 3: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

Physical and mechanical properties of wood-based composites are closely related to density.Vertical and horizontal density distribution(VDD and HDD) generated much research inter-est. Suchsland and Xu (1989, 1991) developedphysical models to examine the effect of HDDon thickness swelling and internal bond strength.They concluded that the durability of flakeboardis substantially affected by the severity of thehorizontal density distribution. Xu and Steiner(1995) presented a mathematical concept forquantifying the HDD. Another study (Wang andLam 1998) linked a simulation program with anexperimental mat through a robot system that de-posited flakes in the simulated positions. Pre-dicted and actual HDD showed good agreement.

Harless et al. (1987) created a very compre-hensive simulation model that can regenerate theVDD of particleboard as a function of the manu-facturing process. Other research in this area in-cluded the characterization of VDD using atrigonometric density function (Xu and Winis-torfer 1996), and a simplified physical model toexamine how the number of flakes, face flakemoisture content, and press closing time affectVDD (Song and Ellis 1997).

Zombori (2001) formed a series of linked sim-ulation and finite element models that could, inturn, recreate the geometric structure, compres-sion behavior, and heat and mass transfer of ori-ented strand board. These models could predictthe inelastic stress-strain response, environmen-tal conditions, moisture content, and density atdifferent points within the panel. His results—some of which are applicable to other compos-ites, like particleboard, too—were in reasonablequalitative agreement with laboratory test data.

Xu and Suchsland created most of the modelssimulating physical and mechanical propertiesof wood composites. They dealt with the linearexpansion of particleboard (1997), followed upby a study discussing the effect of out-of-planeorientation (1998a). In these works, they madeuse of the off-axis MOE, determined by the Han-kinson’s formula. Results of these studies helpedto develop another model (Xu and Suchsland1998b) to simulate the uniaxial MOE of compos-ites with uniform VDD, based on the total volu-

metric work. Xu (1999) modified this simulationto investigate the effect of VDD on the bendingMOE of composites, using the summation ofstiffness values of discrete layers. This modelwas applied to evaluate the effect of percentalignment and shelling ratio on the MOE of OSB(Xu 2000). Simulation results agreed well withexperimental and literature data.

Triche and Hunt (1993) modeled parallel-aligned wood strand composites using finite ele-ment analysis. They created small scale,parallel-aligned strand composites that can beconsidered as physical models for LVL or PSL.The applied model accounted for the effect ofdensification, adhesive penetration, and crush-lap joints, and estimated the tensile strength andMOE of the specimens with excellent accuracy.In a recent study, Barnes (2001) modeled thestrength properties of oriented strand products.He introduced the concept of stress transferangle to assess the effect of strand length andthickness on the mechanical properties of com-posites. He found good agreement between ex-perimental and model-predicted MOE, MOR,and tensile strength values.

Wood-based composite lumbers, such as LSL,LVL, or PSL, are relatively new products thatgenerated less research interest than did compos-ite panels. Results of research works on compos-ite panels can be applied to these products withcare. Meanwhile, available literature does notseem to contain simulation studies that are di-rected specifically towards modeling the geo-metric structure and mechanical properties ofthese structural composites.

theoretical background

Although the structure of the two compositetypes involved in this study is largely different,the fundamental principles used in the simula-tion of their elastic properties are the same. Theeffective bending modulus of composites is de-fined by the stiffness summation of discrete lay-ers (Bodig and Jayne 1982) as follows:

(1)EE I

Ii i= ∑

Bejo and Lang—MODELING THE ELASTIC PROPERTIES OF COMPOSITES 397

Page 4: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

where: E, I – composite MOE and 2nd

order moment of inertia of thecross-section, respectively;

Ei, Ii – layer MOE and moment of inertia of the ith layer with respect to the composite’sneutral axis, respectively.

This theory assumes equal compression andtension stiffness and pertains directly to layeredcomposites like glued-laminated beams, LVL,and plywood. However, it can be applied to non-layered systems, such as PSL, as well. In thiscase, the above summation involves individualstrands, rather than layers. The compression andtension modulus of elasticity may be differentfor some wood species depending on the mois-ture content as reported by Conners and Med-vecz 1992. Therefore, the MOE of a constituentdepends on whether it is located in the compres-sion or the tension zone. Some elements are sub-jected to both compression and tension stresses,and the orthotropic response of woody materialsdepends on their orientation relative to the prin-cipal material coordinates of the composite. Fur-thermore, the inherent densification and resinpenetration may significantly alter the modulusof elasticity, thus impairing the computed re-sults. These points are important to consider dur-ing any model development.

The determination of compression MOE isbased on the calculation of the external work ap-plied to a body and the internal energy storedtherein. These quantities are equal in the linearelastic region. Xu and Suchsland (1998b) de-rived the following equation from this equality:

(2)

where: E, V – compression MOE and totalvolume of the composite;

Ei, Vi – compression MOE and vol-ume of the ith constituent.

The above-discussed simple theories govern thebehavior of SCL materials in many loading appli-cations. However, the stochastic nature of spatialarrangement of the constituents, the natural vari-ability of wood properties, and the unpredictableeffects of manufacturing processes represent real

EE V

Vi i= ∑

challenges in forecasting mechanical properties ofthe final products. Without pursuing absolute pre-cision, our goal was to reasonably model the elas-tic behavior of LVL and PSL on the basis of theproperties of their constituents.

model development

In general, models developed during thisstudy include three major modules that provideelectronically stored information for validationand further computations. First the spatial geo-metric structure of the composite in question ischaracterized using empirical probability distri-butions and/or deterministic variables. In thesecond step, an appropriate routine allocates sto-chastic MOE values to each constituent in theprincipal anatomical directions as input vari-ables for the orthotropic models. The computeddirection-defined stiffness quantities are thenmodified by randomly assigned densificationvalues. Finally, the modulus of elasticity inbending or in compression can be calculatedusing Eqs. (1) or (2).

Although the mechanical and physical proper-ties of structural composites are more consistentthan those of solid wood, still a vast array of in-teracting variables governs their properties andbehavior under load. To maintain the mathemati-cal tractability of the models, several simplifyingassumptions had to be made. The major hypoth-esis of this research was that a single cross-section might represent the entire beam orcolumn reasonably well. This assumption al-lowed the reduction of the three-dimensionalproblem to a much simpler in-plane modeling.Further simplifications used in the models in-cluded the following:● Composites are treated as prismatic beams;

i.e., their cross-section does not change alongthe length; and the constituents are repre-sented by their individual, true size sectionscorresponding to their relative orientation.

● Veneers used for layers and strands are peeledperfectly tangential to the annual ring (i.e.,their plane is LT).

● A continuous glueline provides perfect adhe-sion between the layers or strands.

398 WOOD AND FIBER SCIENCE, JULY 2004, V. 36(3)

Page 5: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

● The applied glue does not alter the MOE ofthe layers or strands significantly.

● The grain orientation of an LVL layer is al-ways parallel with the longitudinal axis of thebeam; the grain orientation of a PSL strand isalways parallel with the longitudinal axis ofthe strand.

● The layup is random; the MOE values of thelayers or strands are independent of their po-sition. In practice for LVL, the stress wavesorting for face and core designation is visu-ally overridden because of defects. Thus, thisassumption may be justified.

● The thickness of a layer in LVL is constant.Irregularities, such as crushed-lap joints thatconnect two veneer sheets and the inherentvoids have counteracting effects on strengthand stiffness, and are disregarded.

● PSL strand cross-sections are rectangular inshape; strands are not bent or distorted.

● PSL strand width is constant (25 mm).● The densification of a PSL strand is indepen-

dent from its position within the billet.● The neutral plane in bending coincides with

the symmetry axis of the cross-section.

simulating the geometry of composites

Figure 1 shows the geometric features of LVLand PSL. The composites are placed in an orthog-onal coordinate system, where x is the longitudi-nal axis of the beams, y is the main cross-sectionalorientation of the constituents, and z is perpendi-cular to both x and y. One can easily realize that x,y, and z are correspondingly equivalent to the L, T,and R principal anatomical directions in solidwood. Using these axes, it is possible to defineload orientation (�′) and strand/layer orientation(�′) relative to normal stresses acting in the beam.These angles are analogous to grain and ringangle in solid wood, respectively.

For LVL, the simulation of the geometricproperties included the following steps:● Establishing the number of layers. This is a

deterministic variable that depends on theparticular material being simulated. In thepresent case, the number of layers was 15.

● Assigning original and final thickness valuesto each layer from their respective probability

distributions. The two outer layers (face) ateach side and the rest (core) were treated dif-ferently because of the different densificationduring the pressing process.

● Calculating cross-section (Ai), 2nd order mo-ment of inertia around the symmetry axis ofthe composite (Ii), and volume fraction (Vi)for each layer.Figure 2a shows the comparison of actual and

simulated cross-sections for LVL. The stochasticparameters of the spatial structure are the thick-ness of the layers and their locations within thecross-section.

To electronically recreate the spatial structureof PSL, we developed an inverse, iterative simu-lation routine. The method is inverse, because itgenerates the projected-densified thickness ofstrands first. It randomly assigns original thick-ness values in the second step; then it computesthe projected width and projected cross-sectional

Bejo and Lang—MODELING THE ELASTIC PROPERTIES OF COMPOSITES 399

Fig. 1. The geometric structure of LVL (a) and PSL (b).

Page 6: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

area according to the orientation and preservesthe densification of the strands. The process isrepeated until a pre-established area coveragecriterion is satisfied. Note that the densification,affecting the MOE of the strand, is defined bythe original and non-projected densified thick-ness.

Steps of the simulation process included thefollowing:● Choosing the y and z cross-sectional dimen-

sions.● Simulating the number of strands. In PSL, the

number of constituents is a random variable,established by generating the number of

strands per in2 (u), and multiplying this num-ber by the cross-sectional area of the beam.

● Assigning projected-densified and originalstrand thickness (t and to), longitudinal strandangle (�), and a cross-sectional strand devia-tion (�) values to each strand. The model neg-lects the deviation of the strands’ longitudinalaxis from the x-y plane. The original width ofthe strands is 25 mm, except for one strand,which has a smaller width. This strand reflectsthe fractional part of the strand number (u).

● Calculating densifications from t and t0;● Summation of projected-densified strand

areas and checking the area coverage crite-rion;

● Arranging strands in both the y and the z di-rections. To achieve an even coverage, strandcentroids are distributed uniformly in a sys-tematic way, rather than randomly.

● Calculating Ai ,Vi and Ii (flatwise and edge-wise), for each constituent. Strands that pro-trude beyond the boundary of the beamcross-section are handled according to thetorus convention (Hall 1988).The strand deposition process results in over-

lapping strand areas and holes as demonstratedon Figure 2b. These overlapping areas and holesrepresent distorted strands that fill the sectionalarea of PSL in reality. Note, that the horizontaldislocation of strands has no effect on the stiff-ness of the composite or a particular individualstrand in bending. In case of compression, the lo-cation of a strand within the cross-section is in-different regarding the composite MOE. In thesimulated cross-section, a projected area, twoangles, and the position of the strand’s centroidrepresent each constituent.

Stochastic parameters, mentioned above, weregenerated using probability density functions fit-ted to experimental databases. Table 1 summa-rizes the type and parameters of each function.Figure 3 shows two sample histograms, createdfrom experimental data. The diagrams includethe overlaid probability density functions, alongwith a histogram of 1000 simulated random de-viates. Visual appraisal and standard statisticaltests (Kilmogorov-Smirnov; �2) helped to iden-tify the best fitting probability density functions.

400 WOOD AND FIBER SCIENCE, JULY 2004, V. 36(3)

Fig. 2. Actual and simulated cross-sections of LVL (a)and PSL (b).

Page 7: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

modeling the elastic parameters of theconstituents

The MOEs of the constituents were generatedin a three-step process:

1. Determining the orientation of the con-stituents relative to the direction of stresses.

2. Assigning compression and/or tension MOEvalues to the constituents in the given orien-tation. This involves the use of orthotropictheories, which require MOE values in vari-

ous anatomical directions, as input parame-ters. Experimental determination of these pa-rameters and the validation of the orthotropicmodels have been presented in two formerarticles. The model simulates compressionand tension MOE using the three-dimensional Hankinson’s formula (Lang etal. 2002) and a combination of the Hankin-son’s formula and the orthotropic tensor the-ory (Lang et al. 2003), respectively. Theelastic input parameters for these formulasare generated assuming normal distribution.

3. Modifying the constituents’ MOE values toaccount for the effect of densification. Themodel increases the MOE denomination ofthe veneers or strands, due to density in-crease during hot pressing, based on experi-mental, second-order densification curves.

We developed the simulation routines usingthe FORTRAN 90 programming environment.The source code includes four programs, as wellas several different functions to generate randomdata from different distributions, so as to simu-late the orthotropic tensile and compressionMOE along with the densification effect. A sam-ple flowchart is included in Figure 4, demon-strating the simulation process for PSL.

A module stores the species-specific input pa-rameters for quaking aspen (Populus tremu-loides), red oak (Quercus rubra), andyellow-poplar (Liriodendron tulipifera). Theseparameters include the summary statistics andprobability density functions of input E values,density, and sectional dimensions. The routinecan be easily extended to cope with new speciesby adding their parameters to this module.

experimental validation

Materials and Methods

Materials used for validation of the modelswere commercial LVL and PSL. While the LVLhad only yellow-poplar veneers, PSL containedapproximately 25% southern yellow pine (Pinusspp.) strands. Note that the input elastic proper-ties in the principal material directions are gen-erated randomly from probability density

Bejo and Lang—MODELING THE ELASTIC PROPERTIES OF COMPOSITES 401

Fig. 3. Sample histograms of the geometric input pa-rameters with the probability density functions overlaid.

Page 8: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

functions. The obvious overlap of the probabilitydensity functions of MOE for the two speciesmay introduce only a few outliers into themodel. Consequently, the effect of mixture ofspecies was neglected. All testing materials werekept in a temporary environmental chamber atapproximately 21ºC temperature and 65% rela-tive humidity (RH) during the entire duration ofthe project. Moisture contents of the test materi-als were periodically checked on control speci-mens using standard ASTM procedure (ASTM1996c).

In the first phase of the validation, twentysimulated cross sections for both LVL and PSLwere compared to actual sections of the materi-als. The links between the actual and simulatedsections were the number of layers and the num-ber of strands per in2 (u) for LVL and PSL, re-spectively. Assigning unit length to thegenerated cross-sections, the density of the sim-ulated composites could be calculated and com-pared to actual values. Experimental densitieswere measured and calculated using structuralsize composites. Additionally, for LVL the totalsimulated thickness of the beam was comparedto actual thickness; and, for PSL the cumulative,projected strand areas (�Ai) and the cumulativemoment of inertia (�Ii) of the strands were mea-sured up to target quantities.

Twenty structural size beams of LVL and PSL,tested both edgewise and flatwise using a 4-point

bending setup over 2.44-m span, provided ex-perimental bending MOE data. The nominalcross-sections were 45 mm � 95 mm and 75 mm� 140 mm for LVL and PSL, respectively. Theprocedure strictly followed the specification ofthe relevant ASTM standard (ASTM 1996a).The Baldwin type testing apparatus had a loadcell with 180 kN capacity. A linear potentiometerhaving 0.01-mm accuracy measured the rela-tive deflection of the shear free section of thebeam. Load-deflection data were collected inreal time using a computerized acquisition sys-tem.

Compression MOE determination followedthe specification of ASTM standard D 143–94,secondary method. Accordingly, specimen di-mensions were: 25 � 25 � 100 mm. For evalua-tion of the orthotropic nature of the composites,the principal dimensions of the specimens wererotated systematically by 45º increments, whichresulted in six groups of load and strand orienta-tion (�′ and �′). Note that �′ and �′ are analo-gous to grain and ring angle of solid woodspecimens. When a particular combination ofload/strand orientation was required, LVL sec-tions were glued side-to-side to provide 100 mmof specimen length. Ten replications of eachcombination provided statistically reliable ex-perimental compression MOE data. Compres-sion load application and the double-sided strainmeasurements were conducted on an MTS

402 WOOD AND FIBER SCIENCE, JULY 2004, V. 36(3)

Table 1. Type and parameters of the probability density functions of the stochastic input variables.

Input Probability Function parameters a

parameter Unit density function � �

Original veneer thicknessQuaking aspen mm Extreme Value A 3.108 0.069 —Red oak mm Logistic 3.046 0.048Yellow-poplar mm 2-param. Weibull — 3.162 42.761

LVL b

to mm Normal 2.760 0.140 —tc mm Logistic 2.940 0.090 —

PSL

u #/in2 Normal 11.64 0.36 —t mm Normal 2.051 0.391 —� ° Normal 0.030 4.668 —� ° Normal 2.991 14.709 —

a – location; � – scale; � – shape.b to – thickness of the two outside layers on both sides ; tc – thickness of the remaining (core) layers.

Page 9: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

Bejo and Lang—MODELING THE ELASTIC PROPERTIES OF COMPOSITES 403

Fig. 4. Flowchart of the bending MOE simulation of PSL.

Page 10: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

servo-hydraulic universal testing machine,equipped with a 10 kN 1N load cell. Pairs ofstrain data were obtained continuously using twoMTS clip-on gauges. The interested reader canfind a detailed description of the specimenpreparation practice and experimental techniquein a related publication (Lang et al. 2002). Bothfor bending and compression, the collected load-deformation data were within the linear elasticregion.

Bending MOE simulation consisted of gener-ating twenty LVL and PSL beams, and calculat-ing their MOE both in flatwise and edgewiseorientations. The cross-sectional dimensions ofthe simulated beams were the same as the nomi-nal dimensions of the tested specimens. The en-tire Monte Carlo simulation was repeated twentytimes resulting in 400 cross-sections where eachrepresented a particular composite beam edge-or flatwise.

Compression model validation was similar tothe above procedure, but it involved 20 replica-tions of only 10 simulated specimens in each ofsix orientations per composite types. This pro-cess resulted in a total of 2400 simulation runs.

Standard statistical procedures including t-test, Mann-Whitney Rank Sum Tests helped toevaluate the differences between experimentaland simulated results. All tests were conducted

at 95 % confidence level (� �0.05).

results and discussion

To confirm the validity of the first module, thesimulated physical properties, including densityand geometrical characteristics, were compared toexperimental data. Table 2 contains the summarystatistics of the results. Simulated quantities areaverage values of the mean, standard deviation,minimum, maximum, and skewness calculatedfrom the twenty Monte Carlo simulations, con-taining twenty beams for each composite type.

For LVL, good agreement between actual andpredicted thickness values could be detected. Itmeans that the selected probability density func-tions described the variations in face and coreveneer thickness reasonably well. In contrast,the model significantly underestimated the den-sity when the simulated values were compared toactual densities measured on 2.44-m-long LVLbeams. Our initial hypothesis was that the over-lapping veneer joints and the coupling through-the-width voids have no effect on the mechanicalproperties of LVL. Apparently, this is not true re-garding density because the volume of denseLVL containing an extra layer is significantlybigger than that of void volume in the beam.However, the comparison of model predicted

404 WOOD AND FIBER SCIENCE, JULY 2004, V. 36(3)

Table 2. Simulated and experimental geometric and physical properties of LVL and PSL.

Property Unit Source Mean STD1 Min. Max. Skewness

LVLThickness mm Sim. 43.0 0.50 42.1 43.9 0.005

Exp. 43.6 0.15 43.1 44.0 0.013Density kg/m3 Sim. 511.0 10.00 492.0 530.0 –0.100

Exp. 566.0 11.00 541.0 584.0 0.078Exp.2 509.6 9.90 498.6 529.7 0.097

PSL�Ai cm2 Sim. 103.7 1.4 101.4 106.1 0.029 composite cm2 Target 106.5 ——- ——- ——- ——-�Ii cm4 Sim. 498.3 6.8 486.9 509.7 0.001Icomposite cm4 Target 515.1 ——- ——- ——- ——-Density kg/m3 Sim. 673.0 11.0 652.0 694.0 –0.037

Exp. 673.0 16.0 640.0 708.0 –0.122

Sample size: n � 20,1 – Standard Deviation,2 – Density of overlapping veneer-joint free LVL.

Page 11: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

densities to measured ones on joint-free LVLsections demonstrated excellent agreement.

As it was expected, the model slightly underpre-dicted the simulated cross-sectional area of PSLstrands (�Ai) because of the built-in area coverageconstraint. In reality, PSL contains longitudinalvoids due to the imperfect packing of strands. SeeFig. 2b where the black areas represent through-the-length holes of a 10-mm-long PSL section.These voids have technological advantage in re-leasing internal steam pressure from the billet afterhigh-frequency die consolidation. The approxi-mately 97.4% average area coverage that resultedfrom the simulation is a good approximation of areal PSL cross-section. Due to the area coveragecriterion, the model underpredicted the 2nd orderarea moment of inertia of the beam (�Ii) by ap-proximately 5% on average basis. The target or ac-tual cross-sections contain voids that cannot behandled during the experimental determination ofbending MOE. This is an inherent bias of the test-ing procedure that we have accepted. Conversely,the simulated density, which showed excellentagreement with actual values, encouraged the ac-ceptance of this simulation routine.

Table 3 contains the summary statistics ofsimulated and experimental bending MOE ofLVL and PSL. Note that the data of simulationresults are the average of twenty batches oftwenty runs, and not the standard deviation ofaverages but the average of standard deviationsare listed. We do believe that this method pro-

vided better estimation of expected minimumand maximum values. Figure 5 provides a quickvisual overview of the results. The simulatedbending MOE values of LVL in both directionswere slightly lower compared to observed data.The differences, however, were not significant,especially in flatwise application and resultedmainly from the higher standard deviation of testresults. It does appear that the effect of overlap-ping veneer joints (their presence/absence andlocation) manifested in higher variability.

The model significantly and consistentlyoverestimated the bending MOE of PSL beams.There are several plausible explanations for thisphenomenon related to the manufacturing tech-

Bejo and Lang—MODELING THE ELASTIC PROPERTIES OF COMPOSITES 405

Table 3. Comparison of simulated and experimental bending MOE of LVL and PSL.

Direction Mean STD2 Min. Max.of Bending Source n1 (GPa) (GPa) (GPa) (GPa) Skewness

LVLFlatwise Sim. 20�20 12.91 0.79 11.40 14.64 0.100

Exp. 20 13.36 0.72 12.20 15.16 0.212Edgewise Sim. 20�20 12.70 0.41 11.95 13.49 0.063

Exp. 20 13.22 1.21 11.85 15.90 0.022

PSLFlatwise Sim. 20�20 14.46 0.34 13.83 15.10 0.041

Exp. 20 12.57 0.57 11.64 13.92 –0.087Edgewise Sim. 20�20 14.54 0.40 13.78 15.29 –0.052

Exp. 20 12.82 0.75 11.18 13.81 0.1151 Replications/sample size.2 Standard deviation.

Fig. 5. Experimental and simulated bending MOE ofLVL and PSL.

Page 12: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

nology of the product. However, based on the re-sults of the compression MOE simulation andexperimental data, we believe that the inaccu-racy of the orthotropic model to estimate theMOE of composites, and neglecting the out of x-y plane deviation of strands are the majorreasons for these overestimations. Currently themodel is under expansion by introducing anadditional random variable: �, defined as the de-viation angle from the x-y plane. Some improve-ment in accuracy is expected, though the effectof rolling shear deformation in an off-axis strandstill will not be encountered.

The compression MOE simulation routinewas run 2400 times to provide an adequate pre-dicted database. Table 4 contains the average sta-tistical parameters of the twenty compressionMOE simulations in the six selected combina-tions of orientation, along with the experimental

statistics per product type. Figure 6 offers aquick comparison between observed and simu-lated results in each direction for LVL and PSL.

Excellent fit of predictions to observed valueswas identified for both LVL and PSL at 0º/-;45º/0º; 90º/0º �′/�′ angle combinations. This in-dicated that the orthotropic models described thechanges of MOE in the x-y plane exceptionallywell; and the inherent shear deformations—mostly in the LT anatomical plane of the con-stituents—had no explicit effect on the elasticperformance of the products.

Once the direction of normal stresses deviatedfrom the principal x-y plane of the composites,the model consistently and significantly overes-timated the modulus of elasticity. This fact indi-cated that neither the 3D Hankinson’s formulanor the 3D orthotropic tensor approach couldmodel the true elastic response of three dimen-

406 WOOD AND FIBER SCIENCE, JULY 2004, V. 36(3)

Table 4. Comparison of simulated and experimental compression MOE of LVL and PSL at different load (�′) and strand(�′) orientations.

Mean STD1 Min. Max.�’ �’ Source (GPa) (GPa) (GPa) GPa) Skewness

LVL0° — Sim. 11.33 0.46 10.61 12.11 0.065

Exp. 11.87 1.06 9.71 13.34 0.14245° 0° Sim. 0.79 0.02 0.75 0.83 0.109

Exp. 0.87 0.07 0.76 1.01 0.00445° 90° Sim. 1.52 0.06 1.43 1.61 0.124

Exp. 0.84 0.06 0.75 0.96 0.54590° 0° Sim. 0.41 0.01 0.39 0.43 0.108

Exp. 0.44 0.03 0.41 0.44 –0.02290° 45° Sim. 0.49 0.02 0.47 0.51 0.170

Exp. 0.29 0.05 0.23 0.35 0.11290° 90° Sim. 0.82 0.03 0.77 0.87 0.146

Exp. 0.37 0.03 0.31 0.42 0.135

PSL0° — Sim. 13.34 0.43 12.63 13.99 –0.139

Exp. 13.20 2.09 10.71 17.28 0.12845° 0° Sim. 1.06 0.04 1.00 1.11 –0.081

Exp. 1.08 0.11 0.86 1.25 –0.00245° 90° Sim. 1.90 0.06 1.80 1.99 –0.035

Exp. 0.64 0.05 0.55 0.73 –0.10090° 0° Sim. 0.51 0.01 0.52 0.56 –0.141

Exp. 0.48 0.07 0.39 0.60 –0.22190° 45° Sim. 0.64 0.02 0.61 0.67 –0.035

Exp. 0.30 0.03 0.25 0.36 0.18690° 90° Sim. 1.02 0.04 0.97 1.08 –0.053

Exp. 0.23 0.03 0.19 0.28 –0.254

Page 13: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

sionally oblique, consolidated wood strands orveneers properly over the entire spatial domain.

Figures 7a and b show orthotropic compres-sion diagrams generated from the average simu-lated and experimental compression MOE ofLVL, respectively, using the orthotropic tensortheory (Szalai 1994). Details c and d on this fig-ure depict similar diagrams for PSL. The simu-lated and actual compression MOE valuesapparently coincide in the x-y plane. However,as the layer or strand orientation (�′ ) increasesover approximately 25º, the trends of the modelpredictions are reversed compared to trends ob-served. No such deviations in trends could beobserved when the orthotropic models were val-idated using solid wood specimens (Lang et al.2002). The veneer manufacturing process mayhave introduced cracks, compression sets that al-tered the rolling shear resistance of the veneers;or the densification affected the out-of-plane de-formation under compression. This phenomenonneeds further investigation.

Nevertheless, these models have practical im-portance because they can predict the expectedelastic response of LVL and PSL with excellentaccuracy from the mechanical properties of theirconstituents in the majority of structural loadingconditions. These include support reactions,columns, edgewise beam, and rafter applicationswhere the direction of normal stresses generated

by external loads usually coincides with the x-yplane of the composites.

Sensitivity analysis was performed to furtherdemonstrate the practicality of the developedsimulation models. The analysis consisted ofsimulating the flatwise-bending MOE of PSL,while progressively reducing the scale parameter(standard deviation) of the distribution of strandangle (�) and strand deviation (�). The exam-ined ranges were between 100% and 0% of thescale parameters, i.e., decreasing the variance of� and � from their original level to no variationat all. Other parameters (e.g., strand thickness,the number of strands and the mean value of �and �) were kept constant.

Figure 8 shows the bending MOE of PSL, as afunction of the standard deviation of � and �,expressed by percents of the original standarddeviations. As this diagram shows, decreasingthe deviation of � from its mean value has littleeffect on the MOE of PSL. On the other hand,changing the variance of � affects the MOE sig-nificantly. Bending MOE improves more than12% (nearly 2 GPa), if the variation is com-pletely eliminated. This is unfortunately not pos-sible for PSL. However, if by some innovativemeans, the standard deviation could be reducedby 50%, it would still increase the bending MOEby 8.5%, provided that the other parameters donot change.

The above analysis required minimal modifi-cation to the simulation model. Other inves-tigations may involve more significantprogramming tasks depending on the complexityof the problem to be analyzed.

summary and conclusions

The bending and orthotropic compressionMOE of LVL and PSL have been modeledthrough simulating the cross-sectional geometryof the composites and the elastic parameters ofthe constituents. In general, good agreementshave been found between simulated and experi-mental characteristics.

Despite their simplicity, the models recon-structed the geometric structure and density ofthe composites reasonably well. Analytical

Bejo and Lang—MODELING THE ELASTIC PROPERTIES OF COMPOSITES 407

Fig. 6. Experimental and simulated compression MOE ofLVL and PSL in the selected six combinations of directions.

Page 14: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

works revealed the difficulties in modeling themodulus of elasticity of spatially oblique woodconstituents. It has been concluded that therolling shear in spatially off-axis strands and ve-neers may significantly influence the apparentmodulus of elasticity in compression or in bend-ing. Further research is needed to enhance exist-ing orthotropic models to address the rollingshear deformation and its effect on the elastic re-sponse of wood based composites. On the other

hand, it has been shown that the elastic perfor-mance of the examined composites can be pre-dicted with excellent accuracy if the normalstresses are parallel to the principal x-y plane ofthe structural elements.

Presumably, the practices described undermodel development, the developed functions,and the module that contains species-specific in-formation provide useful tools for research anddevelopment. The discussed work might aid re-

408 WOOD AND FIBER SCIENCE, JULY 2004, V. 36(3)

Fig. 7. Orthotropic diagrams of the simulated and experimental compression MOE of the composites.

Page 15: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

search targeted to enlarge the raw material basisor introduce novel design in the structural com-posite manufacture.

acknowledgments

This research was partially financed by theMcIntire-Stennis Forestry Research Act, projectNo. 978 at West Virginia University. The authorswish to express their appreciation to Trus JoistMcMillan a Weyerhaeuser Business, LVL andPSL plant at Buckhannon, WV, for donatingSCL materials and providing technical assis-tance in veneer manufacturing. This manuscriptis published with the approval of the Director ofthe West Virginia Agricultural and Forestry Ex-periment Station as Scientific Article No. 2857.

references

American Society for Testing and Materials (ASTM).1996a. Standard methods of static tests of lumber in struc-tural size. ASTM D 198–94. ASTM, West Con-shohocken, PA.

———. 1996b. Standard test methods for specific gravity ofwood and wood-base materials ASTM D 2395–93.ASTM, West Conshohocken, PA.

———. 1996c. Standard test methods for direct moisturecontent measurement of wood and wood-base materialsASTM D 4442–92. ASTM, West Conshohocken, PA.

———. 1996d. Standard methods of testing small clearspecimens of timber. ASTM D 143–94. ASTM, WestConshohocken, PA.

Barnes, D. 2001. A model of the effect of strand length andstrand thickness on the strength properties of orientedwood composites. Forest Prod. J. 51(2):36–46.

Bodig, J., and B. A. Jayne. 1982. Mechanics of wood andwood composites. Van Nostrand Reinhold Co., NewYork, NY.

Conners, T. E., and P. J. Medvecz. 1992. Wood as a bi-modular material. Wood Fiber Sci. 24(4):413–423.

Dai, C., and P. R. Steiner. 1993. Compression behavior ofrandomly formed wood flake mats. Wood Fiber Sci.25(4):349–358.

———, and ———. 1994a. Spatial structure of woodcomposites in relation to processing and performancecharacteristics. Part 2. Modelling and simulation of arandomly-formed flake layer network. Wood Sci. Tech-nol. 28:135–146.

———, and ———. 1994b. Spatial structure of woodcomposites in relation to processing and performancecharacteristics. Part 3. Modelling the formation of multi-layered random flake mats. Wood Sci. Technol.28:229–329.

Hall, P. 1988. Introduction to the theory of coverage pro-cesses. John Wiley and Sons, New York, NY.

Harless, T. E., F. G. Wagner, R. D. Seale, P. H.Mitchell, and D. S. Ladd. 1987. A model to predict thedensity profile of particleboard. Wood Fiber Sci.19(1):81–92.

Harris, R. A., and J. A. Johnson. 1982. Characterizationof flake orientation in flakeboard by the Von Mises prob-ability distribution function. Wood Fiber 14(4):254–266.

Humphrey, P. E., and A. J. Bolton. 1989. The hot pressingof dry-formed wood-based composites. Part II. A simula-tion model for heat and moisture transfer, and typical re-sults. Holzforschung 43:199–206.

Kallmes, O., and H. Corte. 1960. The structure of paper.I. The statistical geometry of an ideal two-dimensionalfiber network. TAPPI 43(9):737–752.

———, ———, and G. Bernier. 1961. The structure ofpaper. II. The statistical geometry of a multiplanar fibernetwork. TAPPI 44(7):519–528.

Lang, E. M., and M. P. Wolcott. 1996a. Model for vis-coelastic consolidation of wood-strand mats. Part I:Structural characterization of the at via Monte Carlo sim-ulation. Wood Fiber Sci. 28(1) 1996, pp. 100–109.

———, and ———. 1996b. Model for viscoelastic consol-idation of wood-strand mats. Part II: Static stress-strainbehavior of the mat. Wood Fiber Sci. 28(3) 1996, pp.369–379.

———, L. Bejo, J. Szalai, and Z. Kovacs. 2000. Or-thotropic strength and elasticity of hardwoods in relationto composite manufacture. Part II: Orthotropy of shearstrength. Wood Fiber Sci. 32(4):502–519.

———, ———, ———, ———, and R. B. Anderson.2002. Orthotropic strength and elasticity of hardwoods inrelation to composite manufacture. Part II: Orthotropy of

Bejo and Lang—MODELING THE ELASTIC PROPERTIES OF COMPOSITES 409

Fig. 8. The effect of decreasing the variation of � and �on the flatwise, bending MOE of PSL.

Page 16: SIMULATION BASED MODELING OF THE ELASTIC PROPERTIES …

compression strength and elasticity. Wood Fiber Sci.34(2):350–365.

———, ———, F. Divos, Z. Kovacs, and R. B. Ander-son. 2003. Orthotropic strength and elasticity of hard-woods in relation to composite manufacture. Part III:Orthotropic elasticity of structural veneers. Wood FiberSci. 35(2):308–320.

Law, A. M., and W. D. Kelton. 1991. Simulation model-ing and analysis. 2nd ed. McGraw-Hill, Inc. New York,NY. 759 pp.

Lenth, C. A., and F. A. Kamke. 1996a. Investigations offlakeboard mat consolidation. Part I. Characterizing thecellular structure. Wood Fiber Sci. 28(2):153–167.

———, and ———. 1996b. Investigations of flakeboardmat consolidation. Part II. Modeling mat consolidationusing theories of cellular materials. Wood Fiber Sci.28(3):309–319.

Lu, C., P. R. Steiner, and F. Lam. 1998. Simulation studyof wood-flake composite mat structures. Forest Prod. J.48(5):89–93.

Song, D., and S. Ellis. 1997. Localized properties inflakeboard: A simulation using stacked flakes. WoodFiber Sci. 29(4):353–363.

Steiner, P. R., and C. Dai. 1993. Spatial structure of woodcomposites in relation to processing and performancecharacteristics. Part 1. Rationale for model development.Wood Sci. Technol. 28:45–51.

Suchsland, O. 1967. Behavior of a particleboard mat dur-ing the press cycle. Forest Prod. J. 17(2):51–57.

———, and H. Xu. 1989. A simulation of the horizontaldensity distribution in a flakeboard. Forest Prod. J.39(5):29–33.

———, and ———. 1991. Model analysis of flakeboardvariables. Forest Prod. J. 41(11/12):55–60.

Szalai, J. 1994. Anisotropic strength and elasticity of woodand wood based composites. Private ed. Sopron, Hun-gary. 398 pp. (in Hungarian.)

Triche, M. H., and M. O. Hunt. 1993. Modeling ofparallel-aligned wood strand composites. Forest Prod. J.43(11/12):33–44.

Wang, K., and F. Lam. 1998. Robot-based research onthree-layer oriented flakeboards. Wood Fiber Sci.30(4):339–347.

Xu, W. 1999. Influence of vertical density distribution onbending modulus of elasticity of wood composite panels:A theoretical consideration. Wood Fiber Sci.31(3):277–282.

———. 2000. Influence of percent alignment and shellingratio on modulus of elasticity of oriented strandboard: Amodel investigation. Forest Prod. J. 50(10):43–47.

Xu, W., and P. R. Steiner. 1995. A statistical characteriza-tion of the horizontal density distribution in flakeboard.Wood Fiber Sci. 27(2):160–167.

———, and O. Suchsland. 1997. Linear expansion ofwood composites: A model. Wood Fiber Sci.29(3):272–281.

———, and ———. 1998a. Influence of out-of-plane ori-entation of particles on linear expansion of particleboard:A simulation study. Forest Prod. J. 48(6):85–87.

———, and ———. 1998b. Modulus of elasticity of woodcomposite panels with uniform vertical density profile: Amodel. Wood Fiber Sci. 30(3):293–300.

———, and P. M. Winistorfer. 1996. Fitting an equationto density profile data using Fourier analysis. Holz Roh-Werkst. 54(1):57–59.

Zombori, B. G. 2001. Modeling the transient effects duringthe hot-pressing of wood-based composites. Ph.D. Dis-sertation, Virginia Polytechnic Inst. and State University,Blacksburg, VA. 212 pp.

410 WOOD AND FIBER SCIENCE, JULY 2004, V. 36(3)