bott masters thesis etd

Upload: gustavo-lacerda-dias

Post on 04-Jun-2018

227 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/13/2019 Bott Masters Thesis Etd

    1/144

    HORIZONTAL STIFFNESS OF WOOD DIAPHRAGMS

    by

    James Wescott Bott

    Thesis submitted to the faculty of

    Virginia Polytechnic Institute and State University

    in partial fulfillment of the requirements for the degree of

    MASTER OF SCIENCE

    in

    CIVIL ENGINEERING

    APPROVED:

    J. Daniel Dolan, Co-Chairman W. Samuel Easterling, Co-Chariman

    Joseph R. Loferski

    April 18, 2005

    Blacksburg, Virginia

    Keywords: wood diaphragm, shear stiffness, diaphragm stiffness, diaphragm deflection

  • 8/13/2019 Bott Masters Thesis Etd

    2/144

    HORIZONTAL STIFFNESS OF WOOD DIAPHRAGMS

    by

    James Wescott Bott

    ABSTRACT

    An experimental investigation was conducted to study the stiffness of wood diaphragms.

    Currently there is no method to calculate wood diaphragm stiffness that can reliably account for

    all of the various framing configurations. Diaphragm stiffness is important in the design of wood

    framed structures to calculate the predicted deflection and thereby determine if a diaphragm may

    be classified as rigid or flexible. This classification controls the method by which load is

    transferred from the diaphragm to the supporting structure below.

    Multiple nondestructive experimental tests were performed on six full-scale wood

    diaphragms of varying sizes, aspect ratios, and load-orientations. Each test of each specimen

    involved a different combination of construction parameters. The construction parameters

    investigated were blocking, foam adhesive, presence of designated chord members, corner and

    center sheathing openings, and presence of walls on top of the diaphragm.

    The experimental results are analyzed and compared in terms of equivalent viscous

    damping, global stiffness, shear stiffness, and flexural stiffness in order to evaluate the

    characteristics of each construction parameter and combinations thereof. Recommendations are

    presented at the end of this study as to the next steps toward development of an empirical method

    for calculating wood diaphragm stiffness.

  • 8/13/2019 Bott Masters Thesis Etd

    3/144

  • 8/13/2019 Bott Masters Thesis Etd

    4/144

    iv

    TABLE OF CONTENTS

    ABSTRACT ii

    ACKNOWLEDGEMENTS iii

    TABLE OF CONTENTS iv

    TABLE OF FIGURES vi

    TABLE OF TABLES viii

    CHAPTER

    I. INTRODUCTION 1

    1.1 Introduction 11.2 Objectives and Scope of Research 41.3 Literature Review 7

    1.3.1 Early Testing 71.3.2 Dynamic Testing 101.3.3 Similar Diaphragms 12

    II. EXPERIMENTAL PROCEDURE 15

    2.1 Scope of Testing 152.2 Test Apparatus 162.3 Diaphragm Construction 232.4 Test Parameters 272.5 Instrumentation 382.6 Test Protocol 452.7 Test Data Analysis 46

    2.7.1 Yielding 472.7.2 Global Deformation 492.7.3 Cyclic Stiffness 512.7.4 Shear Deformation 56

    2.7.5 Shear Stiffness 582.7.6 Flexural Deformation 602.7.7 Flexural Stiffness 602.7.8 Hysteretic Energy 612.7.9 Equivalent Viscous Damping 63

    III. RESULTS AND DISCUSSION 67

  • 8/13/2019 Bott Masters Thesis Etd

    5/144

    v

    3.1 Introduction 673.2 Test Conditions 673.3 Nail Bending Test Results 693.4 Moisture Content and Density Results 713.5 Construction Parameter Results 73

    3.6 Diaphragm Stiffening Methods 78

    IV. SUMMARY, CONCLUSIONS, AND RECOMMENDATIONS 82

    4.1 Summary 824.2 Conclusions 83

    REFERENCES 86

    APPENDIX A 88

    APPENDIX B 95

    APPENDIX C 119

    APPENDIX D 120

    APPENDIX E 128

  • 8/13/2019 Bott Masters Thesis Etd

    6/144

    vi

    TABLE OF FIGURES

    Figure Page

    1.1 Cross-Section of a Typical Wood Framed Floor 2

    1.2 Deep Beam Analogy 4

    2.1 Load Frame, Actuator Connection, and Load Distribution Channel 17

    2.2 Triangular Reaction Frame 18

    2.3 Basic Test Apparatus and Configuration 20

    2.4 Triangular Reaction Frame Plan View Schematic 21

    2.4a Elevation View of Triangular Reaction Frame 21

    2.5 Partial Section of Diaphragm Test Apparatus (loading parallel to joists) 22

    2.6 Rim-Joist Splice for 10 x 40 ft. Specimens 24

    2.7 Basic Specimen Sizes / Orientations (16 x 20 ft. and 20 x 16 ft.) 25

    2.7 (Cont.) Basic Specimen Sizes / Orientations (10 x 40 ft.) 26

    2.8 Fully Sheathed 10 x 40 ft. Specimen 27

    2.9 Corner Sheathing Opening 29

    2.10 Center Sheathing Opening 30

    2.11 Test Configuration with Chords (and corner opening) 32

    2.12 Test Configuration without Chords 32

    2.13 Test Configuration with Walls 34

    2.14 Wall-Lifting Davits 35

    2.15 10 x 40 ft. Specimen with Walls and Wall-Braces 36

    2.16 Application of Sprayed Foam Adhesive 37

    2.17 Foam Adhesive Shown After Removal of a Sheathing Panel 38

  • 8/13/2019 Bott Masters Thesis Etd

    7/144

  • 8/13/2019 Bott Masters Thesis Etd

    8/144

    viii

    TABLE OF TABLES

    Table Page

    3.1 Average Moisture Content and Density Results by Specimen 72

    3.2 Average Percent Differences by Construction Parameter 74

    A.1 Specimen 1 Test Results 89

    A.2 Specimen 2 Test Results 90

    A.3 Specimen 3 Test Results 92

    A.4 Specimen 4 Test Results 93

    A.5 Specimen 5 Test Results 94

    A.6 Specimen 6 Test Results 94

    B.1 Test Comparisons for the Effects of Blocking 96

    B.2 Test Comparisons for the Effects of Foam Adhesive 99

    B.3 Test Comparisons for the Effects of Blocking and Foam Adhesive 101

    B.4 Test Comparisons for the Effects of Increased Nail Density 102

    B.5 Test Comparisons for the Effects of Chords 103

    B.6 Test Comparisons for the Effects of Walls 105

    B.7 Test Comparisons for the Effects of Center Sheathing Openings 109

    B.8 Test Comparisons for the Effects of Corner Sheathing Openings 113

    C.1 Instrument Descriptions 119

    D.1 Specimen 1 Joists Moisture Content and Density 121

    D.2 Specimen 1 Sheathing Moisture Content and Density 122

    D.3 Specimen 2 Joists Moisture Content and Density 123

    D.4 Specimen 3 Joists Moisture Content and Density 124

  • 8/13/2019 Bott Masters Thesis Etd

    9/144

    ix

    D.5 Specimen 4 Joists Moisture Content and Density 125

    D.6 Specimen 5 Joists Moisture Content and Density 126

    D.7 Specimen 6 Joists Moisture Content and Density 127

    E.1 Specimen 1 Test Descriptions 129

    E.2 Specimen 2 Test Descriptions 130

    E.3 Specimen 3 Test Descriptions 132

    E.4 Specimen 4 Test Descriptions 133

    E.5 Specimen 5 Test Descriptions 134

    E.6 Specimen 6 Test Descriptions 134

  • 8/13/2019 Bott Masters Thesis Etd

    10/144

    CHAPTER I: INTRODUCTION 1

    CHAPTER I

    INTRODUCTION

    1.1 INTRODUCTION

    Modern structural engineering frequently involves sheathed construction, a load

    resistance method exemplified in various structural elements by many combinations of suitable

    materials. A common form of sheathed construction, the diaphragm, is a thin, usually planar

    system of sheathing and frame members, intended to withstand considerable in-plane forces.

    When referring to residential housing, everyday plywood construction typically comes to mind.

    Most apparent examples of diaphragms are walls, upper-story floors, and roofs of

    everyday structures such as residential houses, office buildings, and warehouses. Though similar

    in function, wall diaphragms, called shear walls, require different consideration for design and

    analysis, and thus fall outside of the scope of this investigation. Roofs and above-grade floors,

    when designed as such, fall into the classification as true diaphragms. Typical combinations of

    materials employed are wood sheathing on wood frame, metal sheathing on wood frame, metal

    sheathing on metal frame, wood sheathing on metal frame, and variations using concrete,

    structural insulation panels, and other construction materials. This research project is limited to

    wood-framed and plywood-sheathed floor diaphragms typical in residential housing.

    The common floor and roof diaphragm serves dual purposes by supporting vertical forces

    (from loads such as furniture, people, snow, uplift, and its own dead load) and by transmitting

    horizontal forces (from wind pressure or earthquake accelerations) to the supporting shear walls.

  • 8/13/2019 Bott Masters Thesis Etd

    11/144

    CHAPTER I: INTRODUCTION 2

    Floors and roofs are inherently able to carry gravitational loads due to the typical design by

    which appropriately spaced framing members are covered with sheathing and fastened together.

    Joists and rafters, the common framing members of floors and roofs, respectively, are oriented to

    maximize the moment of inertia for resistance to flexure. Thus, a 2x10 floor joist would be

    installed such that the nominal ten-inch side is vertical. The sheathing spans the distance

    between and transmits loads to the framing members below. The framing members then

    distribute the loads proportionally to supporting walls or posts. Also, when adequately fastened

    together, the sheathing and framing can produce a flexurally efficient composite section.

    Resistance to vertical forces, though a primary consideration in design and construction of floors

    and roofs, is not the subject of this study.

    Sheathing

    Wood Floor Joists

    Figure 1.1 Cross-Section of a Typical Wood Framed Floor

  • 8/13/2019 Bott Masters Thesis Etd

    12/144

    CHAPTER I: INTRODUCTION 3

    Horizontal forces applied to diaphragms are almost exclusively from wind and

    earthquakes. Wind pressure on the exterior walls is transmitted proportionately along the edge

    of a diaphragm as a uniform load. In the case of wind-loaded floors, connections with the top

    plate of the wall below and the bottom plate of the wall above provide paths for load transfer.

    When subjected to earthquake accelerations, its own inertia, or resistance to motion, and that of

    attached walls or partitions causes horizontal loading of a diaphragm. Diaphragms are usually

    more than capable of withstanding these loadings due to high in-plane shear capacity. Sheathing

    material itself exhibits considerable in-plane shear strength. Hence, the reason a sheet of

    plywood is much more rigid when loaded along the thin edge (in-plane) as opposed to the large

    flat surface (out of plane). Accordingly, a low aspect ratio system of sheathing panels, properly

    fastened together end-to-end along the edges, has an effective shear capacity. The fact that it is

    usually so thin makes sheathing an efficient and lightweight means of resisting in-plane loads.

    Resistance to these in-plane loads through diaphragm action may be compared to the

    loading of a deep wide-flange beam, as illustrated in Figure 1.2. The shear walls of a structure

    are analogous to the simple supports of the beam and provide the reaction against the forces

    transmitted through the diaphragm. The inter-connected sheathing panels behave like the web of

    the beam to resist the shear component of in-plane loads. And, the extreme edges of the

    sheathing and/or the boundary members running perpendicular to the direction of the loads

    simulate the flanges of the beam by carrying the tension and compression from the flexural

    reaction to the loads.

  • 8/13/2019 Bott Masters Thesis Etd

    13/144

    CHAPTER I: INTRODUCTION 4

    Flange

    Web

    Figure 1.2 Deep Beam Analogy

    The behavior of floor and roof diaphragms has become an important issue with respect to

    lateral stiffness and deflection. It has been noted that there is seldom a problem with the strength

    of diaphragms, because failures are predominantly associated with the connections between a

    diaphragm and supporting walls. Though the occurrence of actual failures is rare, diaphragms

    have sometimes been a controlling factor in the overall failure of structures during seismic events

    (Dolan 1999). Poor understanding of wood diaphragm behavior has spurred the interest of

    researchers to formulate more accurate methods of analysis and design similar to methods

    already employed in the design of cold-formed steel diaphragms.

    1.2 OBJECTIVES AND SCOPE OF RESEARCH

    The objective of this study is to evaluate the stiffness effects of various diaphragm

    construction parameters for use in the development of an accurate method to determine shear

  • 8/13/2019 Bott Masters Thesis Etd

    14/144

    CHAPTER I: INTRODUCTION 5

    stiffness of wood diaphragms based on the formulas currently used in cold-formed steel design.

    Such a formula would allow an improved method of predicting diaphragm deflections. The

    capability to accurately calculate diaphragm stiffness and deflections will enhance the safety and

    economy of wood diaphragms.

    The current lack of an accurate method to predict diaphragm stiffness prevents designers

    from knowing exactly how much stiffness to expect from any given design. Adequate

    diaphragm stiffness is required in order to allow load sharing among the supporting shear walls.

    In other words, flexible diaphragms resist loads locally (i.e., they can not transfer loads

    horizontally very far). Thus, the loads induced into a flexible diaphragm must be transferred to

    local supports (i.e., walls that are the closest to the location of the induced load). A perfectly

    rigid diaphragm would be the other end of the spectrum where all of the supporting walls share

    in resisting the load according to their relative stiffness. In reality, the diaphragm stiffness falls

    somewhere in between these two extremes. However, the higher the diaphragm stiffness, the

    better the load sharing capability is of the structural system and therefore, the better the expected

    performance.

    Currently, diaphragms must be classified as either flexible or rigid in order to select one

    of two different design methods for the transfer of load to the supporting structure. Wood

    diaphragms have been traditionally assumed as flexible, thereby allowing a load distribution

    design based on the tributary area method. This method, however, is not applicable when

    diaphragms exhibit torsional irregularities such as asymmetric geometry and openings (e.g.

    stairwells) or differences in locations of center of rigidity and center of force. To account for

    such torsional irregularities a designer must assume a rigid diaphragm and transfer the load based

    on relative stiffness of the supporting walls.

  • 8/13/2019 Bott Masters Thesis Etd

    15/144

    CHAPTER I: INTRODUCTION 6

    According to the NEHRP Recommended Provisions for Seismic Regulations for new

    Buildings and Other Structures (1997)and as adopted in the Uniform Building Code (1997)and

    the 2000 International Building Codea diaphragms classification changes from rigid to flexible

    when diaphragm deflection under load is equal to or greater than twice the deflection of the

    supporting walls. Thus, designers are now forced to calculate the stiffness of wood diaphragms,

    (and convert the stiffness into a corresponding deflection) in order to even make the distinction

    between rigid and flexible.

    Predicting diaphragm stiffness (or deflection) is difficult because currently there is no

    simple and accurate method that can account for geometrical irregularities as well as all of

    todays varying construction practices. The one current method for calculating deflection of

    wood diaphragms as developed by APA is complicated and is not able to incorporate many

    factors such as sheathing openings, absence of chords, use of sheathing adhesive, and non-

    rectangular shapes. Designers need a simple and accurate method to determine wood diaphragm

    stiffness if they are expected to even begin to select the proper load distribution method.

    The specific objective of this study is to evaluate several basic diaphragm construction

    details for their individual and combined effects on diaphragm stiffness. These results will then

    be used under another task of the CUREE-Caltech Woodframe Project to develop and calibrate a

    finite element model for diaphragm analysis. The overall goal of the experimental diaphragm

    testing and finite element modeling is the development of an equation to accurately predict wood

    diaphragm stiffness in the form of shear stiffness, G, as already accomplished by the cold-

    formed steel industry.

    This specific task is accomplished by a series of experimental tests on full-scale

    diaphragms followed by careful analysis of the results. The test materials and procedures are

  • 8/13/2019 Bott Masters Thesis Etd

    16/144

    CHAPTER I: INTRODUCTION 7

    discussed in detail in Chapter II. Data collected from the tests is evaluated in Chapter III for

    numerical trends indicating the effects of the various specimen configurations on stiffness.

    These trends are used to make general observations regarding the stiffening characteristics of

    each diaphragm construction parameter. All of the stiffness results for every test of each

    specimen are listed for reference in Appendix A.

    1.3 LITERATURE REVIEW

    The objective of this literature review is to examine the studies in the field of theoretical

    and experimental diaphragm research. A study of research performed in this field is important in

    order to know what characteristics have already been established and what questions of

    diaphragm behavior are still unanswered. Sporadic since the 1950s, most of the testing of wood

    diaphragms has occurred at the facilities of the Douglas-fir Plywood Association (DFPA),

    American Plywood Association (APA), Oregon State University, Oregon Forest Products

    Laboratory, Washington State University, and West Virginia University. The volume of

    literature available is small, therefore, rather than place the review in a separate chapter, the

    reviewed literature is provided as a section within this chapter.

    1.3.1 Early Testing

    The DFPA sponsored some early tests in diaphragm behavior. Countryman (1952)

    describes lateral tests on plywood-sheathed diaphragms. Four specimens, 12 x 40 ft. and 20 x 40

    ft., and six one-quarter scale models, 5 x 10 ft., were tested by monotonic loading at fifth-points.

    The specimens had varying parameters such as blocking, openings, staggered panels, gluing,

    plywood thickness, nail size, and boundary nailing patterns. Stiffness of the diaphragms was

    calculated from measured lateral deflection in the middle of the lower chord and applied load.

  • 8/13/2019 Bott Masters Thesis Etd

    17/144

    CHAPTER I: INTRODUCTION 8

    Shear deformation, and not flexural deformation was determined to be the predominant form of

    deflection. Load versus deflection plots show that the actual deflection was consistently higher

    than calculated values using existing equations. It was found that diaphragms behave like a

    horizontal girder with a shear-resistant web. Due to their location at the extreme edges, chord

    members of a diaphragm act like the flanges of a girder by resisting the flexural tension and

    compression forces. The sheathing serves as the web of the girder to resist the shear. Strength

    and stiffness of the specimens was found to be primarily dependent on the strength of the nailed

    plywood-to-frame connections.

    Due to over conservative design codes, the DFPA pursued further studies in diaphragm

    action. Countryman and Colbenson (1954) report on tests of fifteen full-scale diaphragms,

    conducted to better understand the strength effects from:

    1. Omission of blocking

    2. Panel arrangement

    3. Nailing schedules

    4. Span-thickness combinations

    5. Length-width ratio

    6. Seasoning of frame lumber

    7. Use of three inch lumber

    8. Cut-in blocking for chords

    9. Load application perpendicular to joists

    10.Screwed cleats in lieu of blocking

    All 24 x 24 ft. specimens were monotonically loaded with four equal lateral forces at fifth

    points of the span, and deflections were measured from the middle of the unloaded chord.

    Plywood thickness and nailing schedule, along with blocking to a lesser degree, were found to be

    the predominant factors in determining strength and stiffness. Ultimate applied shears ranged

  • 8/13/2019 Bott Masters Thesis Etd

    18/144

    CHAPTER I: INTRODUCTION 9

    from 733 to 2530 plf, while ultimate deflections occurred from 0.52 to 3.2 in. For blocked

    diaphragms, the measured deflections are consistent with a formula produced as a result of the

    DFPA Report No. 55 (Countryman 1952) with an average error of 15%.

    In conjunction with the research described above, the DFPA also sponsored tests at the

    Oregon Forest Products Laboratory. The two 20 x 60 ft. roof diaphragms tested, were

    constructed with the lightest framing and plywood thickness permissible at that time for a roof of

    this size (Stillinger and Countryman 1953). The 2 x 10 joists were framed at 24 in. o.c. and

    sheathed with 3/8-in. thick plywood. One of diaphragms was blocked along the panel edges.

    The diaphragms were loaded monotonically by hydraulic jacks at the fifth points. The 3/8-in.

    thick plywood was found to be adequate, though not as strong as specimens with thicker

    plywood. The lightweight framing system performed adequately. Lastly, it was found that for

    unblocked diaphragms, no special boundary nailing detail was required regardless of the reduced

    strength.

    The APA became interested in lateral shear testing of diaphragms not composed of

    Douglas-fir plywood. Tissell (1966) validated the DFPA tests from 1955 as well as going on to

    test diaphragms of other various species of wood that were becoming popular in construction.

    Nineteen full-scale 16 x 48 ft. diaphragms were tested. Plywood characteristics, sheathing-to-

    framing connections, nail types, and framing member types were varied in the tests. Monotonic

    loading from 16 hydraulic jacks at 3 ft. on center was used to approximate a uniform lateral load.

    Lateral deflections were measured with dial gages at mid span of the tension chord. The design

    shear values were found to be very conservative, with the average ultimate load being 1545 plf

    and the average allowable design load being only 420 plf (includes factors of safety). Sheathing

    of different species of wood was found to have a small but accountable change in shear strength

  • 8/13/2019 Bott Masters Thesis Etd

    19/144

    CHAPTER I: INTRODUCTION 10

    and stiffness. However, effects from plywood grade and quality were found to be negligible.

    Tissell concluded that shear strength equivalent to that of blocked diaphragms is possible by

    stapling tongue-and-groove 2-4-1 plywood. Further, shorter ring-shank nails are permissible as

    long as a minimum penetration is attained. Open-web steel joist-framed diaphragms were

    slightly stronger than the lumber framed diaphragms. The DFPA design values determined from

    the tests previously discussed (Countryman and Colbenson 1954) were found to adequately

    conservative.

    1.3.2 Dynamic Testing

    GangaRao and Luttrell (1980) explain the efforts at West Virginia University to quantify

    shear response of diaphragms with the ultimate goal being the preparation of accurate analysis

    models for future design purposes. Since diaphragms had been mainly studied under static

    loading conditions, they propose that stiffness characteristics are an equally critical issue in a

    correct estimation of behavior under real-life dynamic loading. Preliminary dynamic results

    from tests at West Virginia University were used to derive joint slip and shear deformation

    response equations based on dynamic loading. They predicted that damping characteristics with

    respect to joint slip are the critical factors needed to appropriately describe diaphragm behavior

    under dynamic loads.

    At the time, Polensek (1979) was the only researcher making attempts at quantifying

    damping characteristics for horizontal dynamic loading. His tests of plywood sheathed

    diaphragms with six or ten inch joists yielded average equivalent viscous damping ratios

    between 0.07 and 0.11. However, he considered that the data accumulated had been too varied

    for an accurate estimation of the damping ratio. It was apparent, however, that an increase in the

    damping effect is directly proportional to floor span.

  • 8/13/2019 Bott Masters Thesis Etd

    20/144

    CHAPTER I: INTRODUCTION 11

    At West Virginia University, Jewell (1981) performed experimental tests on partial

    (cantilever) diaphragms in order to analyze a range of different parameters such as nail spacing,

    boundary conditions, connection details, load type, and damping capacity. Three 16 x 24 ft.

    diaphragms and six 16 x 16 ft. diaphragms were tested under monotonic, cyclic, and impact

    loads in the directions perpendicular and parallel to the joists. Replica diaphragms were also

    modeled in the same configurations as flexible composite members in a finite element analysis to

    determine any inaccuracy in this theoretical approach. Based on a comparison of the theoretical

    and experimental test results, Jewell was able to analyze relationships of plywood behavior, nail

    slip, effect of loading, effect of joist hangers, and damping to the stiffness of diaphragms. In

    most cases, the finite element approach yielded slightly less conservative results for stiffness

    (i.e., predicted deflections were lower than actual), based on the parameters listed above.

    Corda (1982) and Roberts (1983) performed additional cantilever diaphragms tests at

    West Virginia University in another codependent study involving laboratory testing and finite

    element modeling. Corda tested six 16 x 24 ft. specimens cyclically and statically to failure in

    order to study local and global in-plane shear stiffness response to variations of blocking,

    openings, plywood thickness, corner stiffeners, and framing nail sizes. It is noteworthy that nail

    softening after loads up to 9 kips on some specimens caused a large decrease in stiffness.

    Increased plywood thickness (without using longer nails) and corner openings reduced strength

    but had little effect on stiffness. Roberts theoretical analysis of equivalent models of the

    diaphragms tested by Corda, showed some evident discrepancies. Problems with the finite

    element analysis program included the limitation to monotonic loading, inaccurate predictions of

    panel slip, and the iterative processes of calculation of diaphragm deflection with respect to nail

    slip, a bilinear relationship, demanding the modification of results to a nonlinear solution using

  • 8/13/2019 Bott Masters Thesis Etd

    21/144

    CHAPTER I: INTRODUCTION 12

    the tangent stiffness method. Based on the problems encountered, Roberts suggested that the

    limitations imposed on the program user in modeling plywood diaphragms need to be eliminated

    by further experimental research into stiffness characteristics of panel slip, plywood layout and

    connection, diaphragm openings, and nail slip under cyclic loading.

    A recent APA report by Tissell and Elliott (1997) describes diaphragm testing for high

    load conditions equivalent to earthquakes accelerations. The primary intent was to formulate

    design and construction approaches for these high-load diaphragms, which may incorporate

    use of two layers of plywood, thicker plywood, or stronger fastener conditions. Ten of the

    diaphragms tested were 16 x 48 ft., and the dimensions of an eleventh specimen were changed to

    10 x 50 ft. Hydraulic jacks at a spacing of 24 in. o.c. were used to apply a cyclic uniform load

    along the long side of the diaphragms. Results show that it is possible to increase shear strength

    by increasing the number of fasteners or adding another layer of sheathing in areas of high shear.

    This report also notes that plywood panel shear capacity must be checked for high-load

    diaphragms. Staples were found to be adequate fasteners in lieu of nailed sheathing-to-framing

    connections. Along the same lines, field glued joints and a reduced number of nails are adequate

    for these diaphragms.

    1.3.3 Similar Diaphragms

    The abundant studies of floors comprised of materials other than wood are important in

    order to understand general behavior of diaphragms. It is possible that wood diaphragm design

    methods may be simplified and accurately rationalized in terms of methods already in use for

    other materials. A theoretical study of the behavior of composite steel beam and concrete deck

    diaphragms was made by Widjaja (1993) at Virginia Polytechnic Institute and State University.

    Similar to many efforts currently in progress for wood diaphragms, the purpose of this study was

  • 8/13/2019 Bott Masters Thesis Etd

    22/144

    CHAPTER I: INTRODUCTION 13

    to develop an accurate finite element analysis model that predicts diaphragm behavior,

    incorporates possible variations of design parameters, and derives design strength equations.

    Similarly, experimental cantilever tests on cold-formed steel diaphragms are important in the

    design of many steel-frame building roofs and composite floors (during construction, before

    concrete). Post-frame diaphragm testing (wood frame and metal sheathing) is also significant in

    terms of the more agricultural or shed type buildings.

    With sponsorship from NUCOR Research and Development, Hankins et al. (1992)

    performed eighteen cantilever diaphragm tests at Virginia Polytechnic Institute and State

    University to determine strength and stiffness of cold-formed steel sheathing, 20 and 22 gage

    thickness (Vulcraft 1.5B1 deck), welded or bolted to a steel frame. The 16 x 16 ft. specimens

    were subjected to monotonic loads. Thirteen diaphragms utilized an 8 ft. span, requiring only

    one filler beam. The other five specimens had a filler beam spacing of 4 ft., requiring three filler

    beams. Bolt and puddle weld arrangement, used to secure the sheathing, was varied to determine

    its effects on diaphragm behavior. Results from the tests indicate that specimens with thicker

    gage sheathing have more strength and stiffness. However, even though specimens with smaller

    filler beam spacing (three filler beams as opposed to one) had more strength, the diaphragm

    stiffness was less in some cases.

    Hausmann and Esmay (1977) report the results of tests on twenty-six full size post-frame,

    metal-clad diaphragm panels. All specimens were 8 x 16 ft. with rafters at 4 ft. o.c. along the 16

    ft. side and purlins at 2 ft. o.c. along the 8 ft. side. Loaded monotonically at the ends of the three

    interior rafters, the panels were analyzed for strength and stiffness based on varying parameters

    such as framing arrangement, type, number, and metal of fasteners, aluminum or steel sheathing,

    and with or without insulation. It was determined that purlins laid flat to the rafters was the more

  • 8/13/2019 Bott Masters Thesis Etd

    23/144

    CHAPTER I: INTRODUCTION 14

    suitable method of framing. Screw fasteners in the panel valleys increased diaphragm stiffness

    and strength, especially for the steel clad specimens. Aluminum panels were more suitable with

    nailing, due to a larger cover width for each sheet. Placing insulation between wood-framing

    and metal cladding not only reduces diaphragm strength, but also seriously affects stiffness.

    Fastener configurations have important and measurable effects on diaphragm behavior.

    Anderson and Bundy (1990) performed additional post-frame diaphragm tests to outline

    the effects of openings in the sheathing. Fifteen cantilever specimens, 7-2/3 x 12 ft. with two

    interior rafters and seven purlins were tested monotonically with varying amounts of sheathing

    missing. Diaphragms were constructed with SPF lumber, screw fasteners, and steel sheathing.

    Fastener configurations were found to be extremely important for diaphragm stiffness.

    Openings in the sheathing at normal intervals caused the specimens to be ineffective as

    diaphragms. It was also found that spacing of purlins has little impact on strength or stiffness of

    the diaphragms.

    In addition to the physical testing, there has been a great deal of computer modeling of post-

    frame diaphragms for scientific purposes in order to aid designers and validate experimental

    results. For example, Wright and Manbeck (1993), among many others, conducted finite

    element analyses of post-framed diaphragm panels. Following procedures provided by Woeste

    and Townsend (1991), they modeled full size 8 x 12 ft. diaphragms with 2x4 in. purlins at 2 ft.

    o.c., 2x6 in. rafters at 3 ft. o.c., and steel cladding secured with 16d nails. The finite element

    model was compared to three identical experimental diaphragm tests. The finite element model

    closely predicted diaphragm shear strength, but under-estimated shear stiffness by 28%. Results

    show that discrepancies arise due to difficulties in modeling nonlinear behavior of fasteners and

    intricate load paths between the wood frame and steel sheathing.

  • 8/13/2019 Bott Masters Thesis Etd

    24/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 15

    CHAPTER II

    EXPERIMENTAL PROCEDURE

    2.1 SCOPE OF TESTING

    Multiple stiffness tests and one test to failure were performed on each of six full-scale

    diaphragms at the Thomas M. Brooks Forest Products Center of the Virginia Polytechnic

    Institute and State University located in Blacksburg, Virginia. Consortium of Universities for

    Research in Earthquake Engineering (CUREE) sponsored the research under its Wood Frame

    Project, Task 1.4.2 Diaphragm Studies.

    Diaphragm dimensions for four specimens were 20 x 16 ft. with varying orientations,

    while two high-aspect ratio specimens were 10 x 40 ft. Multiple tests on each specimen were

    possible due to the non-damaging deflections being imposed, allowing an economical means of

    incorporating multiple test parameters. Test parameters investigated for effects on diaphragms

    stiffness were: 1) corner opening, 2) center opening, 3) fully sheathed, 4) 6-12 nail pattern, 5) 3-

    12 nail pattern, 6) with/without chords, 7) with/without walls, 8) with/without blocking, and 9)

    with/without foam adhesive. Specimens were subjected to non-destructive, low-amplitude

    dynamic-cyclic loading by a computer-controlled hydraulic actuator, while load and deflection

    values were being recorded by a computer data acquisition system. The final test on each

    specimen, though not a primary focus of this study, was an attempt to cause diaphragm failure.

  • 8/13/2019 Bott Masters Thesis Etd

    25/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 16

    2.2 TEST APPARATUS

    Diaphragm testing was conducted on a 22 X 50 ft. concrete pad with 42 in. wide by 27 in.

    tall concrete back-walls along two adjacent sides. The heavily reinforced back-walls have two

    7/8 in. diameter anchor bolts embedded in the concrete at 2 ft. on center with a 400,000 lb point-

    load capacity at a minimum spacing of 6 feet.

    A computer controlled hydraulic actuator was mounted horizontally at the midpoint of

    the 50 ft. back-wall. The actuator had a 55 kip capacity with a 6 in. stroke, and included a 50

    kip Interface load cell, screwed onto the end of the hydraulic cylinder. Load was transferred

    from a ball joint at the end of the actuator, through a pin-connected gusset to a 20 ft. long

    C6x10.5 steel channel. The channel, as shown in Figure 2.1, was fastened along the entire width

    in the center of the specimen span with 5/8 in. diameter lag screws. In cases where a joist did not

    fall in the center of the diaphragm, 4x4 blocks are placed under the sheathing to provide a

    backing for the lag screws used to attach the steel load distribution channel.

  • 8/13/2019 Bott Masters Thesis Etd

    26/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 17

    Figure 2.1 Load Frame, Actuator Connection, and Load Distribution Channel

    Offset equal distances (based on dimensions of diaphragm specimens) from the centerline

    of the actuator were triangular reaction frames, as shown in the photograph of Figure 2.2. The

    frames were constructed of 4 X 6 in. steel tubes welded together. Each reaction frame was

  • 8/13/2019 Bott Masters Thesis Etd

    27/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 18

    connected to the concrete back-wall with four 7/8 in. diameter anchor bolts. A one-inch thick

    end plate was welded to the end of the steel tube of the reaction frame and was drilled and tapped

    for a 1 in. threaded rod. A two-foot piece of threaded rod was screwed through the plate into

    the steel tube leaving the desired length exposed. A shop-fabricated, full-bridge load cell made

    of 2 in. diameter steel rod and strain gauges screwed onto the opposite end of the threaded rod.

    The load cell had a large hex-nut welded to one end to connect to the threaded rod protruding

    from the triangular reaction frame. Gusset plates welded to the opposite end of the load cell

    provided a pinned connection to the diaphragm support frame.

    Figure 2.2 Triangular Reaction Frame

    Each end of the diaphragm was attached to a 20 ft.-L2x2x steel angle, which was

    welded intermittently to the side of a 20 ft.-3 X 5 in. steel tube using lag screws. One end of

    each steel tube was pin-connected to the gusset plates of the reaction load cells. This support

  • 8/13/2019 Bott Masters Thesis Etd

    28/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 19

    frame served the same purpose as shear walls by transmitting loads out of the diaphragm at each

    end. The reaction frame load cells then measured these loads. The support frame served a

    secondary purpose, to hold the diaphragm at the proper elevation for concentric loading from the

    actuator. Several one-inch diameter PVC pipes were also placed on the concrete under the

    specimen to help hold the interior of the specimen at the proper elevation, and to act as

    frictionless rollers under the joists as load was applied. The schematic drawings of Figures

    2.3, 2.4, and 2.5 illustrate all of the elements of the test apparatus including the specimen support

    frame, the triangular reaction frame, and the load distribution channel.

  • 8/13/2019 Bott Masters Thesis Etd

    29/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 20

    Figure 2.3 Basic Test Apparatus and Configuration

    Steel Channel(C6 x 10.5)

    Lag Screws

    Concrete Back Wall

    See Figure 2.4

    for Detail

    Fig.2.5

    Load Cell

    Actuator

    Diaphragm Sizeand SheathingLayout Varies

    Support Frame3" x 5" Steel Tube

    Triangular ReactionFrame

    Load Cell

  • 8/13/2019 Bott Masters Thesis Etd

    30/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 21

    Load Cell

    Steel TubeFrame

    Fig.2.4a

    Figure 2.4 Triangular Reaction Frame Plan View Schematic

    Pinned Connection

    ConcreteBackwall

    Diaphragm Support Frame

    Figure 2.4a Elevation View of Triangular Reaction Frame

  • 8/13/2019 Bott Masters Thesis Etd

    31/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 22

    Figure 2.5 Partial Section of Diaphragm Test Apparatus(for specimens loaded parallel to the joists)

    Steel Angle:-Welded to Steel Tube-Lag Screwed to End Joist

    2 x 12 Joists @ 16"

    (Douglas Fir)

    2332" T&G Plywood

    Sheathing

    PVC Pipe Rollers

    114"

    Steel Channel (C6X10.5) - LagScrewed to non-structural blocks

    Steel PipeRoller

    3" x 5" x 3 8" Steel Tube

  • 8/13/2019 Bott Masters Thesis Etd

    32/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 23

    2.3 DIAPHRAGM CONSTRUCTION

    Of the six, full-scale diaphragm specimens, two were 16 x 20 ft. in dimension and were

    loaded parallel to the direction of the joists on the 20 ft. side. Two specimens were 20 x 16 ft. in

    dimension, loaded perpendicular to the joists on the 16 ft. side. The last two specimens were 10

    x 40 ft. in dimension, loaded parallel to the joists on the 40 ft. side. Resembling the size and

    shape of one side of a roof of a typical residential home, these 10 x 40 ft. specimens were

    intended to test the envelope of diaphragm performance with respect to aspect ratio.

    The diaphragm specimens were framed with Douglas-fir 2 x 12 joists spaced at 16 in. o.c.

    and nailed with three 16d nails at each end to a 2 x 12 Douglas-fir rim joist. In the case of

    specimens loaded parallel to the direction of the joists, the bottom of each end joist was attached

    to the diaphragm support frame using lag screws as shown in Figure 2.5. Conversely, when

    loading was applied perpendicular to the joists, the rim-joists were connected to the support

    frames. Since the lumber used was 20 ft. in length, the rim joists of the 40 ft. long specimens

    had to be spliced in the center with steel plates and bolts as shown in Figure 2.6 (while such a

    splice may not be feasible in real-life construction due to interference with finish materials,

    effective transfer of compression and tension forces in the chords was essential for valid test

    results). The three main specimen configurations, including loading and reaction locations, are

    schematically illustrated in Figures 2.7a-c. A photograph of a 10 x 40 ft diaphragm specimen is

    presented in Figure 2.8.

    A wood sample was taken from each joist of every specimen for moisture content and

    density analysis. This information was recorded for possible use when evaluating test results,

    since moisture content changes in lumber affects fastener performance.

  • 8/13/2019 Bott Masters Thesis Etd

    33/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 24

    Figure 2.6 Rim-Joist Splice for 10 x 40 ft. Specimens (typical both sides)

  • 8/13/2019 Bott Masters Thesis Etd

    34/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 25

    (a) 16 x 20 ft. Specimen

    (b) 20 x 16 ft. Specimen

    Figure 2.7 Basic Specimen Sizes / Orientations

    20'

    Load Applied4'x8'x23 32" T&G

    Plywood Sheathing

    16'

    Cut-out showsframing layout below

    2 x 12 Joists @ 16"(Douglas Fir)

    2 x 4 Blocking(on flat)

    2 x 12 Rim-Joist(Douglas Fir)

    20'

    Load Applied

    16'

  • 8/13/2019 Bott Masters Thesis Etd

    35/144

    Load Applied4'x8'x2332" T&G P

    Sheathing

    40'

    (c) 10 x 40 ft. Specimen

    Figure 2.7 (Continued) Basic Specimen Sizes / Orientations

  • 8/13/2019 Bott Masters Thesis Etd

    36/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 27

    Figure 2.8 Fully Sheathed 10 x 40 ft. Specimen

    The specimens were sheathed with nominal 4 x 8 ft. sheets of 23/32 in. tongue-and-

    groove plywood in a staggered panel configuration. Sheathing was cut as required to complete

    the desired panel configuration. Sheets were attached to the framing with 10d nails in a 6/12 nail

    pattern, meaning nails are spaced at 6 in. around the perimeter and at 12 in. on the interior

    supports of each sheathing panel. Typical sub-flooring construction adhesive was not used

    between the joists and plywood sheathing; however some tests involved the use of sprayed foam

    adhesive.

    2.4 TEST PARAMETERS

    Specimens were subjected to a number of different construction variations, including the

    multiple combinations thereof. The variations tested were:

    1. Sheathing openings fully sheathed, corner opening, center opening

  • 8/13/2019 Bott Masters Thesis Etd

    37/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 28

    2. Chord members with / without rim joist

    3. Blocking with / without 2 x 4 blocking

    4. Walls with / without 4 ft. tall stud-framed walls

    5. Sprayed foam adhesive & nails versus nailed only

    6. Sheathing nail density 6/12 versus 3/12 nail pattern

    Variations to the basic specimen listed above, are individually discussed in detail in the

    following paragraphs.

    Openings in the plywood sheathing were intended to simulate common openings in floors

    of residential homes for stairways, atriums, and vaulted ceilings. These openings weaken and/or

    cause torsional irregularities that can dramatically affect the stiffness of diaphragms. Duplex

    (double-headed) 10d nails were used to fasten the sheathing panels that were to be removed from

    the specimens in order to simulate openings. The corner opening in all sizes and orientations of

    specimens was easily achieved by removing one full 4 x 8 ft sheet of plywood from a corner.

    However, the center opening presented more challenges due to the staggered sheathing

    configuration and tongue-and-groove plywood. For both orientations of the 16 x 20 ft.

    specimens, an 8 x 12 ft. rectangular opening was made by prying the unfastened sheets up in the

    center along the tongue-and-groove seam like an army tent and lifting them out. Some sheets

    had to be cut in half to achieve a rectangular opening. Due to the high aspect ratio of the 10 x 40

    ft. specimens, a proportional rectangular opening in the center was not reasonable, since typical

    roof and floor diaphragms would not have such an opening. A schematic drawing and

    accompanying photograph of the two opening types used in the tests are presented in Figures 2.9

    and 2.10.

  • 8/13/2019 Bott Masters Thesis Etd

    38/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 29

    Figure 2.9 Corner Sheathing Opening

    Load Applied

  • 8/13/2019 Bott Masters Thesis Etd

    39/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 30

    Figure 2.10 Center Sheathing Opening

    Load Applied

  • 8/13/2019 Bott Masters Thesis Etd

    40/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 31

    The chords of a diaphragm are the exterior framing members that are oriented

    perpendicular to the direction of loading. They serve to resist bending moment induced in a

    diaphragm while also supporting the extreme edges of the sheathing. In the case of floor

    diaphragms, the chords may either be the rim joist or simply the last joist at each end of the floor,

    depending on the orientation and direction of loading. Residential roof diaphragms typically do

    not have a true rim joist, either at the lower edge along the fascia or at the ridge (unless the fascia

    board or ridge beam is considered to be effective). The absence of an effective chord is

    especially prevalent for roof systems utilizing metal plate connected trusses.

    Though all testing was performed on floor diaphragm specimens, chord effects should be

    similar for roof-like specimens. The effectiveness of chords was quantified by running tests with

    and without the designated chord members (rim joists) in place. Only those specimens having

    rim joists as the chords (specimens loaded parallel to the direction of the joists) could be tested in

    this manner. The rim joists were nailed to the diaphragm at each joist with three 16d duplex

    nails. Plywood edges were nailed to the rim joist with 10d duplex nails at 6 in. o.c. Duplex nails

    were used for easy removal of the rim joists between different test specimen configurations. One

    of the diaphragm specimens is shown in both configurations of having the rim joist acting as the

    chord in place and removed in Figures 2.11 and 2.12 respectively.

  • 8/13/2019 Bott Masters Thesis Etd

    41/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 32

    Figure 2.11 Test Configuration with Chords (and corner opening)

    Figure 2.12 Test Configuration without Chords

  • 8/13/2019 Bott Masters Thesis Etd

    42/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 33

    Blocking is the term used for the short framing members that span between joists and

    serve to interlock the unsupported joints between sheathing panels. They can have the same

    cross-sectional dimensions as the joists (full-depth blocking improves noise and vibration

    dampening), or blocks can simply be smaller lumber laid on flat. In this investigation, specimen

    configurations with blocking used 2 x 4s laid on flat installed between joists where each line of

    unsupported sheathing panel joints would fall prior to installing the sheathing. The blocks were

    fastened to the joists on each side with two 16d common toe nails. Plywood panel edges that fell

    over the blocks were nailed every 6 in. with 10d duplex nails for easy removal to simulate

    blocked and unblocked conditions. To reconfigure the specimen without blocking, the sheathing

    nails were extracted, the diaphragm was tilted on-end with a forklift, and blocks were removed.

    Likewise, replacing blocking involved tilting the diaphragm up to install new 2 x 4 blocks from

    below, and then re-nailing the sheathing to the blocking.

    A potentially significant unknown in diaphragm design is the effect of walls on the

    horizontal stiffness. Walls of a structure transfer wind loads to the floors to which they are

    connected. Also, the mass of the walls themselves present added lateral loads to diaphragms

    during earthquakes. These walls, especially the flexural stiffness of their own bottom plate, may

    also benefit a floor diaphragm by helping to resist these same lateral loads.

    For testing purposes, four-foot high walls were installed along the two chord edges of

    specimens. As shown in Figure 2.13 these walls were constructed of 2 x 4 studs at 16 in. o.c.

    and 7/16 in. OSB sheathing on the outside. The 2 x 4 bottom plate of the walls was fastened

    through the plywood sheathing to the floor joists below with 3 x in. self-tapping Simpson

    screws for a strong connection yet easy removal. For the 16 x 20 ft. specimens, regardless of

    orientation, the walls were set in place and removed with a long, cable-supported, boom attached

  • 8/13/2019 Bott Masters Thesis Etd

    43/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 34

    to a large forklift. Starting with the second specimen, lever-action davits, as shown by the

    photographs of Figure 2.14, were welded to the side support frames at each end of the walls to

    more quickly and safely facilitate raising and lowering for configurations with and without walls.

    As shown in Figure 2.15 for the 40 ft. long specimens, the walls on each side were built in 20 ft.

    sections, set in place with the boom, and connected in the center. From then on, the davits at

    each end of the walls accompanied by braces in the center allowed for repeated installation and

    removal of walls. Braces were required to laterally stabilize the wall segments for the 10 x 40 ft.

    diaphragm specimens.

    Figure 2.13 Test Configuration with Walls

  • 8/13/2019 Bott Masters Thesis Etd

    44/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 35

    (a) Walls Lowered and Attached (b) Walls Unfastened and Raised

    Figure 2.14 Wall-Lifting Davits

  • 8/13/2019 Bott Masters Thesis Etd

    45/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 36

    Figure 2.15 10 x 40 ft. Specimen with Walls and Wall-Braces

    Current trends in construction involve the use of adhesives for an ever-widening range of

    applications. In this case the method of fastening sheathing panels to framing members was

    varied between nailed only and nailed plus sprayed adhesive. The adhesive material used was a

    sprayed, two-part, self-expanding, poly-isocyanurate foam adhesive manufactured by ITW

    Foamseal. The foam adhesive was tested using coupon tests to quantify its stiffness as a

    connection. The connection stiffness was determined to be equivalent to that obtained by using

    elastomeric adhesives typically used in wood floor construction. While elastomeric adhesive

    would have been more representative of traditional construction, it would have prevented the

    possibility of removing sheathing without damage once fastened down, thereby making it costly

    and difficult to alter specimen sheathing configurations. Sheathing fastened down with the foam

    adhesive could be removed with minimal damage by cutting the adhesive at the joints with a

  • 8/13/2019 Bott Masters Thesis Etd

    46/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 37

    knife. Also, testing of the foam adhesive will provide useful information for its effectiveness in

    roof retrofit applications. After having tested all of the nailed-only configurations, the foam

    adhesive was applied to the underside of fully-constructed specimens that could be safely tilted

    on-end (i.e. the two 16 x 20 ft. and the two 20 x 16 ft. specimens). Specifically, the adhesive was

    sprayed along each side of every joist at the interface with the sheathing. Adhesives were not

    used on the 10 x 40 ft. specimens due to the specimens flexibility, which made tilting the

    specimens without damage impossible. A photograph of the foam adhesive being applied is

    shown in Figure 2.16, and a photograph of a sheathing panel removed is shown in figure 2.17.

    Figure 2.16 Application of Sprayed Foam Adhesive

  • 8/13/2019 Bott Masters Thesis Etd

    47/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 38

    Figure 2.17 Foam Adhesive Shown After Removal of a Sheathing Panel

    Sheathing nail density was varied for a few tests on the third and fourth specimens, both

    of which were 20 x 16 ft. loaded perpendicular to the joists. The nail pattern was changed from

    6/12 to 3/12 on the fully sheathed and nailed only configurations. In other words, nail spacing

    around the perimeter of each sheathing panel (where supported by joists or blocking) was

    decreased from 6 in. to 3 in. o.c. using easily removable 10d duplex nails. The 3-12 nail pattern

    was tested while the walls and blocking parameters remain variable.

    2.5 INSTRUMENTATION

    Movements, deflections, and loads were measured at multiple locations on the diaphragm

    specimens using electronic sensors of various types in conjunction with a computer controlled

    Data Acquisition system (DAQ). The DAQ used for this project LABTECH, a Windows PC-

  • 8/13/2019 Bott Masters Thesis Etd

    48/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 39

    based program. Prior to testing, all instruments used in this research project were carefully

    calibrated for accurate results. Before any series of tests in a day, all instruments were checked

    to make sure they were correctly mounted, functioning properly, and zeroed. Due to the exposed

    conditions of the outside testing facility, all instruments were demounted and taken inside or

    covered with plastic daily to protect from inclement weather, dew, and frost.

    An internal Linear Variable Displacement Transducer (LVDT) measured the deflections

    caused by the hydraulic actuator. Signals from this highly sensitive device were transmitted to

    the computer controller, which in turn used the information as the feedback channel to control

    the actuator. Loads measured by the 50 kip Interface load cell were recorded by the DAQ and

    had no effect on the displacement-controlled actuator.

    The custom-built load cells at each end of the diaphragm measured the reaction loads,

    both in tension and compression due to the cyclic loads from the actuator. These reactions

    simulated the shear loads that supporting walls of an actual structure must withstand. Prior to

    use, these load cells, as described in Section 2.2, were separately calibrated in tension only on a

    universal testing machine with an excitation of 10 Volts (compression was not feasible due to a

    pinned-pinned condition when using special calibration fixtures). Both were loaded

    incrementally to 40 kips tension, and in each case the linear calibration plot proved that no

    yielding within the load cell occurred. The slopes of these lines were used in the DAQ as

    multipliers to convert the output voltage signals from the load cells into equivalent values of

    load.

    Horizontal movement of the plywood sheathing relative to the framing members below

    was measured at two locations with external LVDTs. An aluminum bracket mounted to the

    end-joist at each rear corner held the barrels of a pair of LVDTs in place horizontally. The

  • 8/13/2019 Bott Masters Thesis Etd

    49/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 40

    plungers of these instruments react against metal tabs, which were screwed and glued to the

    plywood at the rear corners of the diaphragm. The LVDTs of each pair pointed in orthogonal

    directions to account for biaxial sheathing movement. A photogragh of the LVDT mounting

    setup is shown in Figure 2.18.

    Figure 2.18 LVDTs and Mounting Bracket

  • 8/13/2019 Bott Masters Thesis Etd

    50/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 41

    String Potentiometers (abbreviated, string-pot) were used at multiple locations to

    determine diaphragm movement, deformation, and slippage relative to the test frame. Seven

    string-pots were set along the front face of the specimens to measure the global deflection. A

    string-pot was attached to each steel side support frame to determine the slip in the side load cell

    connections and between the steel frame and the diaphragm itself. Likewise, a string-pot was

    mounted to the steel load channel in the center to determine any slip its lag screw connection to

    the specimen. Two string-pots were mounted diagonally on each side of the diaphragm

    centerline to record the deformation caused by shear deflection during testing. Schematics

    illustrating the positions of each displacement sensor for each specimen configuration are

    presented in Figures 2.19 through 2.21. A list describing each instrument is presented in

    Appendix C.

  • 8/13/2019 Bott Masters Thesis Etd

    51/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 42

    GR1

    Concrete Back Wall

    LVDTL-NS

    LVDTL-EW

    GL1SlipL

    DL1

    SlipC

    DL2 DR1

    GL2 GL3 GC

    LVDT

    R-NS

    LVDTR-EW

    DR2

    3" x 5"Steel Tube

    GR2 SlipRGR3

    Figure 2.19 Instrumentation Plan for 16 x 20 ft. Specimen

  • 8/13/2019 Bott Masters Thesis Etd

    52/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 43

    Concrete Back Wall

    GC GR1 GR2 GR3 SlipRGL1 GL2 GL3SlipL

    SlipC

    DL1 DL2 DR1 DR2

    LVDTL-NS

    LVDTL-EW

    LVDTR-EW

    LVDTR-NS

    Figure 2.20 Instrumentation Plan for 20 x 16 ft. Specimen

  • 8/13/2019 Bott Masters Thesis Etd

    53/144

    Concrete Back Wall

    GC GR1 GR2GL1 GL2 GL3SlipL

    SlipC

    DL1 DL2 DR1

    LVDT

    L-NS

    LVDT

    L-EW

    Figure 2.21 Instrumentation Plan for 10 x 40 ft. Specimen

  • 8/13/2019 Bott Masters Thesis Etd

    54/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 45

    2.6 TEST PROTOCOL

    In most cases of experimental research, how a specimen is stressed and to what extent, is

    equally as important as the specimens characteristics. In this project great care was taken to

    NOT apply deflections so large that specimens reached or exceeded their yield point and were

    damaged. On the other hand, it is critical to apply sufficient deflection to obtain valid test data.

    This limit was found for each specimen size and orientation by loading monotonically in small

    increasing increments until signs of diaphragm damage are seen or heard, or until the slope of the

    load/deflection curve, shown in real-time on the DAQ computer screen, appeared to be

    decreasing. The deflection amount used for all tests was slightly lower than the largest

    monotonic deflection. Additionally, deflections for this project followed a cyclic pattern that

    somewhat simulates the cyclic loading of earthquakes, only not nearly as rapid, since this

    apparatus is not intended or equipped to perform shake-table testing.

    The load protocol for all tests, except those to failure, was five sinusoidal cycles at the

    predetermined deflection, 0.25 for specimens one and two (20 ft. wide), 0.20 for specimens

    three and four (16 ft. wide), and 0.80 for specimens five and six (40 ft. wide). The frequency

    of these cycles was set in the actuator controller at 0.0833 Hz for test durations of 60 seconds.

    Five cycles, or even possibly less, are adequate since this project does not incorporate the effects

    of load fatigue.

    While not a primary focus of this project, the last test of each specimen was an attempt to

    cause failure. The CUREE protocol (Krawinkler et al. 2000) used for these tests is a deflection-

    controlled quasi-static cyclic load history. This protocol is based on a finite series of cycles with

    plateaus and peaks of increasing amplitude. Yield deflection, , was estimated for each

  • 8/13/2019 Bott Masters Thesis Etd

    55/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 46

    specimen and used as the reference deflection from which amplitudes of other cycles were

    determined. If failure did not occur by the end of the series, the CUREE protocol allows for

    additional cycles at higher amplitudes until the specimen fails.

    2.7 TEST DATA ANALYSIS

    As with any physical experiment, manipulation of raw test data (in this case, load and

    deflection values) is required for a logical comparison of the different specimen configurations.

    In this study, several specific variables are calculated in order to weigh the benefits and

    detriments caused by changing test parameters as described in Section 2.4. These variables and

    the methods used for their calculation are presented in the following sections.

    Test data from each deflection and load measuring instrument was recorded by the DAQ

    computer and entered into a spreadsheet format. Each column of data in the spreadsheet

    corresponds to one of the twenty-four channels (instruments) being used, and is ordered

    chronologically with time from the start of each test. A table listing each of these instruments,

    its model and serial number, and calibration coefficient is in Appendix C.

    These text-format spreadsheet files were later imported individually into a Microsoft

    Excel calculation template. This template was programmed with the calculations necessary for

    automatic computation of stiffness results and other important variables. The template also

    provided instant load-deformation graphs for the data.

    The first step of the calculations was to take the raw deflection data and convert it into

    tared values by subtracting out the initial reading. For example, a string potentiometer with a

    range of 10 in. is drawn out 5.25 in. and attached to the specimen. The first data entry for this

    channel will indicate a deflection of 5.25 in. Therefore each of the data points in that column

  • 8/13/2019 Bott Masters Thesis Etd

    56/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 47

    must be tared by either subtracting or adding 5.25 in. depending on the direction of deflection in

    order to attain the actual change in deflection.

    Because the DAQ system had to be manually started and stopped in conjunction with the

    independently controlled actuator, a section of data from the beginning and end of each test was

    invalid. Therefore, the template was also programmed to shorten the data columns to include

    only the meaningful data acquired during testing.

    2.7.1 Yielding

    While not a desired outcome of small-deformation stiffness testing, yielding is an

    important concept to understand in terms of elastic versus plastic behavior. A material subjected

    to a static load will undoubtedly undergo some deformation, though potentially immeasurably

    small depending on its physical properties. If once unloaded, the material returns to its original

    state, it is considered to have behaved elastically. However, if the material is loaded beyond its

    elastic range causing permanent deformation even after being unloaded, then it has experienced

    yielding. The force, fy, required to cause yielding is referred to as the yield point. Further

    yielding caused by continued loading past this point, but before failure, is called plastic

    deformation. Idealized elastoplastic response of a material subjected to force, f, causing

    deformation, , is illustrated in Figure 2.22.

  • 8/13/2019 Bott Masters Thesis Etd

    57/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 48

    Figure 2.22 Idealized Elastoplastic Force-Deformation Curve

    force (f)

    deformation ()

    fy

    y

    failure

    plastic behaviorelastic behavior

  • 8/13/2019 Bott Masters Thesis Etd

    58/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 49

    2.7.2 Global Deformation

    Maximums and minimums were established from deflection or load readings of each

    instrument. The rigid body motion determined from string potentiometers measuring slip in test

    frame connections was subtracted from maximum positive and negative global deformations.

    The resulting adjusted maximum global deformations were plotted against instrument location

    distances along the length of the specimen. This curve represents a diaphragms shape at

    maximum deformation, and also aids in visualizing effects of torsional irregularity. The sign

    convention used for the purposes of this study is illustrated in Figure 2.23. Outward deformation

    caused when the hydraulic actuator pushes out is considered positive. An actual diaphragm

    deformation curve from Specimen 2, Test 8 (16 x 20 ft., no chords, with walls, corner opening,

    blocked, nailed only) is presented in Figure 2.24. Note, the lop-sidedness (towards the right)

    caused by a torsional irregularity due to the corner sheathing opening on that side.

  • 8/13/2019 Bott Masters Thesis Etd

    59/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 50

    Figure 2.23 Simplified Diaphragm Deformation Curve with Sign Convention

    Figure 2.24 Diaphragm Deformation Curve - Specimen 2, Test 8

    -

    +

    -

    g

    +

    g

    Distance AlongDiaphragm Edge

    GlobalDeformation

    (g)

    0.230.19

    -0.12

    -0.16

    -0.21-0.20

    -0.15

    0.13

    0.19

    0.24

    -0.3

    -0.2

    -0.1

    0

    0.1

    0.2

    0.3

    0 2 4 6 8 10 12 14 16 18 20

    Distance, L to R (ft)

    GlobalDeformation,

    RBMS

    ubtracted

    (in)

  • 8/13/2019 Bott Masters Thesis Etd

    60/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 51

    2.7.3 Cyclic Stiffness

    Stiffness is the most basic and yet also the most useful of all the variables determined for

    this study. In lay terms stiffness is simply a measurement of a structures capacity for resisting

    deformation. Although stiffness can be expressed in several different forms, in general it is the

    amount of force required to cause a known unit of elastic deformation. Thus, if a system is

    linearly elastic, then its stiffness, k, can be described as the slope of the force-deformation curve

    as illustrated in Figure 2.25.

    Figure 2.25 Stiffness of a Linearly Elastic System

    Cyclic loading requires a different approach to calculating stiffness. Cyclic loads change

    the force-deformation plot from a straight line to an elliptical loop known as a hysteresis as

    shown in Figure 2.26. Because the slope changes along the hysteresis, cyclic stiffness must be

    approximated. A visual comparison of two common methods of determining cyclic stiffness

    and their associated equations is presented in Figure 2.26. The peak-to-peak method shown in

    Figure 2.26a approximates stiffness as the slope of an imaginary line between the points of

    maximum positive and maximum negative deflection. Figure 2.26b shows the origin-to-peak

    f

    1

    k

  • 8/13/2019 Bott Masters Thesis Etd

    61/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 52

    method where stiffness is an average of slopes of imaginary lines from the origin out to the

    points of maximum positive and negative deflection.

  • 8/13/2019 Bott Masters Thesis Etd

    62/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 53

    (a) Peak-to-Peak Method

    (b) Origin-to-Peak Method

    Figure 2.26 Cyclic Stiffness Calculation Methods

    f

    max

    +

    max-

    k+

    +

    +

    +

    =maxmax

    maxmax ff

    k

    f

    max

    +

    max-

    k+

    k-

    2

    + +=

    kkk

    fmax+

    fmax-

  • 8/13/2019 Bott Masters Thesis Etd

    63/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 54

    The spreadsheet program determined cyclic stiffness values using both the origin-to-

    peak method and the peak-to-peak method. The first value calculated was cyclic stiffness

    based on the force required from the actuator to cause a unit amount of deformation (excluding

    rigid body motion) at the center of the diaphragm specimen. For tests in which the load-

    deformation hysteresis is not centered on the origin due to uneven loading of the specimen, the

    origin-to-peak method may not seem appropriate. While the peak-to-peak method may

    seem better suited for such cases of uneven loading, a comparison of the two methods shows that

    the cyclic stiffness results never varied by more than 3% in all 132 tests. Therefore, both

    methods are assumed to be valid for test data exhibiting uneven loading. Due to the similar

    results under either method and for brevity, this study will focus on analysis using the peak-to-

    peak method from this point forward. (Note: the cause of uneven loading is assumed to be

    residual stresses stored in the specimen from the previous test due to friction.)

    As shown by Figure 2.27, either method produces similar results for a load hysteresis that

    is well centered on the origin. Likewise, for hysteresis loops that are not centered on the origin

    as shown by Figure 2.28, the two methods still yield similar results.

  • 8/13/2019 Bott Masters Thesis Etd

    64/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 55

    -20

    0

    20

    -0.25 0 0.25

    Deformation (in)

    Load(kips)

    Figure 2.27 Cyclic Load-Deformation Hysteresis, Specimen 3, Test 3(Loops symmetric about the origin.)

    -20

    0

    20

    -0.25 0 0.25

    Deformation (in)

    Load(kips)

    Figure 2.28 Cyclic Load-Deformation Hysteresis, Specimen 3, Test 24

    (Loops not symmetric about the origin due to uneven loading.)

    -9.45 kips

    9.46 kips

    -0.17 in

    0.18 in

    -15.18 kips

    9.01 kips

    0.15 in

    -0.20 in

    Origin-to-Peak:

    kip/ft54.12

    0.17-

    9.45-

    0.18

    9.46

    kcyclic =

    +

    =

    Peak-to-Peak:

    kip/ft0.5417.00.18

    45.99.46k

    cyclic=

    +

    +=

    0.2% Difference of 0.1 kip/ft.

    Origin-to-Peak:

    kip/ft68.02

    20.0

    15.18-

    15.0

    01.9

    kcyclic

    =+

    =

    Peak-to-Peak:

    kip/ft1.6920.00.15

    18.159.01k

    cyclic=

    +

    +=

    1.6% Difference of 1.1 kip/ft.

  • 8/13/2019 Bott Masters Thesis Etd

    65/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 56

    2.7.4 Shear Deformation

    Global deformation is the sum of diaphragm shear deformation and flexural deformation

    as illustrated in Figure 2.29. Since the load is applied at the center of diaphragm specimens, then

    theoretically the shear deformation for each half of the diaphragm is equal. Diaphragm shear

    deformation, visually detailed in Figure 2.30, can be determined from a geometric manipulation

    of deflection results from the diagonal string potentiometers.

    FLEXURAL

    DEFORMATION

    GLOBAL DEFORMATION

    (Neglecting Rigid Body Motion)

    SHEAR

    DEFORMATION

    (g)

    (f) (s)

    Figure 2.29 Diaphragm Deformation Theory

  • 8/13/2019 Bott Masters Thesis Etd

    66/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 57

    F

    F2

    F2

    bb

    d

    s

    L2

    L2

    Diagonal String Pot.

    (at zero deflection)

    Diagonal String Pot.

    (at deflection L)

    Figure 2.30 Diaphragm Shear Deformation

    First, the maximum diagonal deformation values in each direction for each half of the

    diaphragm are determined, then averaged together, giving an average diagonal deformation, L,

    for each half of the diaphragm, and for both positive and negative deformation. Using small

    angle assumptions, shear deformation, s, can be expressed as:

    2

    Ls = (2.7.1)

  • 8/13/2019 Bott Masters Thesis Etd

    67/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 58

    where is shear strain of the diaphragm and is calculated from the diagonal string potentiometer

    deflections and geometric diaphragm properties by:

    bd

    dbL

    22 +

    = (2.7.2)

    The data spreadsheet program calculated a s value for both halves of the specimen as

    well as for maximum positive and negative diaphragm deformation. In some cases, shear

    deformation for the left side of the diaphragm did not equal that of the right side and the positive

    maximum did not equal the negative maximum. Though uneven loading can be attributed to

    differences in positive versus negative shear deformation results, the differences in left side

    versus right side shear deformation are due to different shear stiffness of each side.

    2.7.5 Shear Stiffness

    Shear stiffness is equal to shear force divided by shear deformation. Using the shear

    deformations just determined, the spreadsheet program calculates shear stiffness for both sides of

    the diaphragm using:

    +

    +

    +

    +=

    SS

    shear

    VVk (2.7.3)

    where V- and V+ are the maximum positive and negative shear forces applied to a side of the

    diaphragm.

    Under symmetric loading conditions and torsionally regular construction, the shear force

    resisted by either side of the diaphragm is theoretically half of the load applied at the center by

    the actuator. Thus, theoretically for torsionally regular test configurations, Vequals the reaction

    force F/2 as shown by Figure 2.30, and can be verified by the load readings from the reaction

    load cells at each side. However, asymmetric configurations such as a corner sheathing opening

  • 8/13/2019 Bott Masters Thesis Etd

    68/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 59

    may cause the diaphragm to resist more of the actuator-applied load on the stiff side of the

    diaphragm and less on the soft side. In such instances, Vshould be determined independently

    for the left and right sides using the maximum readings from the left and right side load cells,

    respectively. Accordingly, for torsionally irregular configurations (corner opening being the

    only case for this study) the left and right side shear stiffness values must be kept separate.

    As a caveat to this approach, a recurring problem during testing of Specimens 3, 4, and 5

    was electrical malfunction of the side load cells, especially on the left side. An alternate method

    by statics had to be used for tests in which there was reaction load cell malfunction. As

    previously indicated, for a torsionally regular and symmetrically loaded specimen, each reaction

    force equals half of the actuator-applied force. Similarly, for torsionally irregular specimens (i.e.

    corner opening) the reaction forces combined, though not necessarily equal, should add up to the

    actuator-applied force, F. If for example, the left side load cell malfunctions, its load can be

    approximated by:

    RL RFR = (2.7.4)

    With the experimental shear stiffness, kshear, calculated for both the left and right sides of

    the diaphragm specimen, the spreadsheet program uses elastic beam theory where:

    s

    sAG

    LV

    2

    = (2.7.5)

    where GAsis a more commonly accepted form of shear stiffness. Using the relationship:

    s

    shear

    Vk

    = (2.7.6)

    Equation 2.7.5 may be solved for GAsin terms of kshearto give:

    2

    LkAG shears = (2.7.7)

  • 8/13/2019 Bott Masters Thesis Etd

    69/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 60

    A shear stiffness value, GAs, is calculated for both the left and right sides of the

    diaphragm. The left and right side values are averaged together for torsionally regular

    specimens, but kept separate for torsionally irregular specimens in order to allow proper

    comparison of results. For such cases it is possible that the shear stiffness on one side may be

    considerably higher than the other.

    2.7.6 Flexural Deformation

    As shown visually by Figure 2.29, flexural deformation can be determined by:

    sgf = (2.7.8)

    The spreadsheet program averaged the maximum svalues for each half of the specimen for both

    positive and negative deflection. These average shear deformation values were then subtracted

    from the corresponding maximum positive and negative global deformations to give a maximum

    positive and a maximum negative flexural deformation.

    2.7.7 Flexural Stiffness

    Experimental flexural stiffness, kf, may be expressed as:

    +

    +

    +

    +=

    ff

    f

    FFk (2.7.9)

    The spreadsheet uses the above peak-to-peak equation in order to arrive at one flexural stiffness

    value.

    Elastic beam theory offers an approach to a more common form of flexural stiffness.

    Theoretical flexural deformation of a beam with a concentrated load applied at the center is:

    EI

    FLf

    48

    3

    = (2.7.10)

  • 8/13/2019 Bott Masters Thesis Etd

    70/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 61

    where EI is the flexural stiffness of the beam, or, as in this case, the diaphragm. Solving

    Equation 2.7.10 forEIin terms of kf (recalling that kf = F / f) gives:

    48

    3LkEI

    f= (2.7.11)

    The spreadsheet program calculated flexural stiffness in the above form for each diaphragm test.

    2.7.8 Hysteretic Energy

    Hysteretic energy is the energy dissipated during one cyclic loading of a structure and

    may be quantified as the area inside a load-deformation hysteresis for one cycle. The area within

    a hysteresis from an experimental test can be approximated by numerical integration. Numerical

    integration involves averaging the load values of two consecutive points along the curve.

    Multiplying this average load by the difference between deformation values of the same two

    consecutive data points gives the area under the curve between those two data points.

    Geometrically, the calculation equates to determining the area of very narrow trapezoid. This

    process must be repeated for every pair of consecutive data points all the way around the loop.

    Depending on the location along the curve with respect to the deformation axis, the area is

    considered either positive or negative. The net total of these incremental areas is the hysteretic

    energy as represented algebraically in Equation 2.7.12:

    ( ) ++

    +

    +=

    n

    n

    nnnnd

    ffE

    1

    11

    2 (2.7.12)

    where the integer nrepresents the individual data points around the entire loop. The spreadsheet

    program follows the same process described above to arrive at a value for hysteretic energy for

    each diaphragm test. However, for greater accuracy, the program calculates a total area for three

    consecutive loops around the hysteresis (using data points only from the middle three loops) and

  • 8/13/2019 Bott Masters Thesis Etd

    71/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 62

    then divides the total by three, yielding an average area inside only one loop. See Figure 2.31 for

    a visual description of numerical integration of a hysteresis.

    Figure 2.31 Numerical Integration of a Load-Deformation Hysteresis

    f

    a

    Area under curve between points

    aandbto be considered positivefa

    b

    fb

    +

    a

    b

    f

    c

    fc

    d

    fdd

    c

    -

    Area under curve between points

    canddto be considered negative

  • 8/13/2019 Bott Masters Thesis Etd

    72/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 63

    2.7.9 Equivalent Viscous Damping

    Damping is the mechanism that causes gradual reduction of vibration in a system and

    thereby a loss of energy. In structures, damping and the resulting energy loss is caused by a

    variety of conditions such as internal friction of materials subjected to repeated deformations,

    friction from movements at connections, opening and closing of cracks, and friction with

    external or nonstructural systems with which the structure is in contact. Since the systems that

    can cause damping are seemingly limitless and difficult to identify, a mathematical model

    capable of predicting actual damping is nearly impossible.

    Thus, a concept called equivalent viscous damping is used to represent all of the damping

    mechanisms for a simplified approach under the assumption that the structure behaves as a

    Kelvin solid viscoelastic element (Fischer and Filiatrault 2000). Equivalent viscous damping is

    defined by equating the energy loss during a vibration cycle (i.e. the hysteretic energy as defined

    in Section 2.7.8 and graphically as shown in Figure 2.32) in an actual structure to that of an

    equivalent viscous system.

  • 8/13/2019 Bott Masters Thesis Etd

    73/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 64

    Figure 2.32 Damping relationship to an equivalent viscous system

    Using Figure 2.32 and setting the hysteretic energy, Ed, from one experimental cycle

    equal to that of the equivalent viscous system gives:

    22 oo

    neqd XkE

    = (2.7.13)

    where eqis the equivalent viscous damping ratio, is the experimental test frequency, nis the

    natural frequency of the test structure. Strain energy,ESo, is equal to the area of triangle OAB

    (or triangle OCD) from Figure 2.32 and can be calculated from experimental stiffness, ko at

    maximum deformationXo:

    2

    2

    oo

    OABSo

    XkAE == (2.7.14)

    Substituting Equation 2.7.14 into Equation 2.7.13 gives:

    OABn

    eqd AE 4= (2.7.15)

    f

    Xo

    +

    Xo-

    ko+

    ko-

    B

    A

    OD

    CEd = Hysteretic Energy(Area enclosed by hysteresis loop)

  • 8/13/2019 Bott Masters Thesis Etd

    74/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 65

    Although for this experiment does not actually equal n, for simplicity the assumption as such

    does allow for an acceptable approximation of eq (Chopra 1995). Assuming = n and

    solving for the equivalent viscous damping ratio gives:

    OAB

    deq

    A

    E

    4= (2.7.16)

    As shown in Figure 2.32, the maximum deformations in both the positive and negative directions

    are not necessarily equal. Therefore, the equivalent viscous damping ratios should be calculated

    for triangle OAB using ko+ and triangle OCD using ko

    - and then averaged. Thus, equivalent

    viscous damping ratio, eq, is commonly expressed as:

    So

    deq

    E

    E

    4= (2.7.17)

    whereESois strain energy (equal to the area of triangle OAB or OCD).

    For the purposes of this study, the spreadsheet program calculates the areas on both the positive

    and negative sides of the curve using cyclic stiffness values determined by the origin-to-peak

    method. The spreadsheet takes an average of the two strain energy values and the hysteretic

    energy previously calculated, and uses a formula based on Equation 2.7.17 to determine the

    equivalent viscous damping ratio for each diaphragm test.

    Although equivalent viscous damping is not technically correct for the tests in this study

    due to some non-linearity in the load-deformation response, the maximum specimen deflections

    were kept low to minimize error. The viscous damping term is a measure of all of the damping

    in the system (hysteretic and material) and is used for modeling the more complex system as a

    simplified mass-spring-dash pot system. The equivalent viscous damping is the value for the

    dash pot. While the design community views the concept of equivalent viscous damping as very

  • 8/13/2019 Bott Masters Thesis Etd

    75/144

    CHAPTER II: EXPERIMENTAL PROCEDURE 66

    inaccurate for calculating damping in an actual structure, it is used in research simply as a means

    of comparing damping capability.

  • 8/13/2019 Bott Masters Thesis Etd

    76/144

    CHAPTER III: RESULTS AND DISCUSSION 67

    CHAPTER III

    RESULTS AND DISCUSSION

    3.1 INTRODUCTION

    The objective of this diaphragm study was to test the stiffness of wood diaphragm

    specimens under varying configurations in order to develop a method for determining shear

    stiffness similar to that already used by the cold formed steel industry. Combined, 132 non-

    destructive stiffness tests were performed on 6 different specimens. Due to the high volume of

    data produced, a test-by-test analysis and comparison of results would be monotonous and not in

    the best interest of the reader. Therefore, the calculations applied to each set of test data will be

    thoroughly explained step-by-step in general terms. The remainder of this chapter will be

    devoted to discussion of trends in test results and relating those trends back to the various

    construction parameters. The individual test results are presented in Appendix A.

    3.2 TEST CONDITIONS

    As a foreword to discussion of results, the reader needs to be aware of the ever-changing

    conditions encountered during testing, namely weather. Testing of diaphragm specimens began

    on January 12, 2001 and continued until July 6, 2001. Throughout those six months the weather

    played a substantial role in the test schedule, specimen moisture content, and periodic

    malfunction of instruments and equipment.

  • 8/13/2019 Bott Masters Thesis Etd

    77/144

    CHAPTER III: RESULTS AND DISCUSSION 68

    As indicated in Chapter 2, the diaphragm testing facility is an outside concrete slab with

    no protection from the weather. A large tarp was always used to cover speci