You are on page 1of 298

Seismic Behavior of Sheathed

Cold-Formed Steel Stud


Seismic Behavior of Sheathed Cold-Formed Steel Stud Shear Walls: An Experimental Investigation Luigi Fiorino - 2003

Shear Walls:
An Experimental Investigation

Luigi Fiorino

G Sliding-hinge

Rotational-hinge

Dottorato di Ricerca in Ingegneria delle Strutture


Universit degli Studi di Napoli Federico II
Polo delle Scienze e delle Tecnologie

Luigi Fiorino

Seismic Behavior of Sheathed Cold-Formed


Steel Stud Shear Walls:
An Experimental Investigation

Tesi di dottorato
XVI Ciclo

Dottorato di ricerca in Ingegneria delle Strutture


I would like to express my sincere gratitude to my tutor Prof. Federico
M. Mazzolani, who has guided me during the doctoral activity. His
enthusiasm and determination have represented a clear example for my
research activity and not only for it.
I am much thankful to Prof. Raffaele Landolfo who offered and made
possible my studies in the field of cold-formed steel structures. This research
would not have been possible without his supervision and support.
Thanks to Dr. Gaetano Della Corte for his assistance in this study,
particularly in the field of seismic engineering. I appreciate a lot of useful help
that I received from His.
I would like to thank to Dr. Beatrice Faggiano for her aid. She always
somehow found time to help me and all-working group.
The suggestion and constructive criticism of Dr. Gianfranco De Matteis
is also gratefully acknowledged.
I thank my fellow engineers, Ettore de Cesbron de la Grennellais,
Gianmaria Di Lorenzo, Maurizio Manganiello, Simeone Panico, Francesco
Portioli, who shared with me the experiences of the PhD period.
I respectfully acknowledge the financial support given to the research
project by the Italian Ministry for University and Research. Also I extend the
acknowledgements to the BPB Italia for the furniture of gypsum wall board
sheathings and the assembling of the specimens, the GUERRASIO for the
furniture of cold-formed steel profiles, the TECFI s.r.l. for the furniture of
screws, and the HILTI Italia for the furniture of anchors.
I also wish to thank my family for their affection, and encouragement as
well as their appreciation of my studies.
Last, but not least thanks to my wife Stefania for his love, patience,
emotional and material support.
To my grandmother Assuntina
i

Contents

Foreword 1

Chap ter I
Low-rise residential buildings built with cold-formed
lightweight steel members 5
1.1 BASIC CONSTRUCTION SYSTEMS.. 6
1.2 STICK-BUILT CONSTRUCTIONS......... 8
1.3 PROFILES....15
1.4 SHEATHINGS. 18
1.4.1 Wood-based panels... 18
1.4.2 Gypsum-based boards... 20
1.5 CONNECTIONS.. 21
1.5.1 Screws... 22
1.5.2 Welding..... 24
1.5.3 Pins.... 24
1.5.4 Nails.. 26
1.5.5 Bolts.. 26
1.5.6 Anchors..... 26
1.5.7 Adhesives.. 30
1.5.8 Hold-down..... 30
1.6 THE BUILDING PROCESS... 31
1.7 BASIC PROBLEMS OF STRUCTURAL DESIGN. 34
1.7.1 Vertical loads 34
1.7.2 Horizontal loads 35
1.8 ADDITIONAL CONSIDERATIONS 38
1.8.1 Fire resistance 38
1.8.2 Durability.. 39
1.8.3 Thermal and acoustic insulation, air and moisture permeability.. 40
1.9 ADVANTAGES AND LIMITATIONS. 41
REFERENCES 42

Chap ter I I
Design of cold-formed steel stud shear walls 45
2.1 MAIN DESIGN METHODOLOGIES FOR SHEAR STRENGTH OF CFSSSWS... 46
Conten ts ii

2.2 SEMI-ANALYTICAL EVALUATION OF SHEAR STRENGTH . 48


2.2.1 Strength of the sheathing-to-frame connections (S-F).. 50
2.2.2 Strength of the frame (stud buckling strength) (F) 55
2.2.3 Strength of the frame-to-foundation connections (F-F) 67
2.3 UNITS SHEAR DESIGN VALUES FOR CFSSSWS.. 76
2.3.1 UBC and IBC design tables.. 76
2.3.2 IRC provisions on the PSW method utilizations.. 83
REFERENCES 84

Chap ter I II
Te s t i n g o f c o l d - f o r m e d s t e e l s t u d s h e a r w a l l s : r e v i e w o f
existing literature 89
3.1 SUMMARY OF MAIN EXISTING TEST PROGRAMS. 89
3.1.1 McCreless & Tarpy (1978)... 90
3.1.2 Tarpy & Hauenstein (1978).. 92
3.1.3 Tarpy (1980)..92
3.1.4 Tarpy & Girard (1982).. 94
3.1.5 Tissell (1993) 96
3.1.6 Serrette (1994)... 97
3.1.7 Serrette & Ogunfunmi (1996)... 97
3.1.8 Serrette et al. (1996a,b). 99
3.1.9 Serrette et al. (1997a) 101
3.1.10 Serrette et al. (1997b) 104
3.1.11 NAHB Research Center (1997) 107
3.1.12 Gad et al. (1999a, b).. 108
3.1.13 Selenikovich et al. (1999). 111
3.1.14 COLA-UCI (2001) 112
3.1.15 Dubina & Fulop (2002). 114
3.1.16 Branston et al. (2003) 116
3.2 FACTOR AFFECTING CFSSSW BEHAVIOR UNDER LATERAL LOADING: THEY
INDIVIDUATION AND ANALYSIS. 118
3.2.1 Sheathing (type, thickness, orientation) 118
3.2.2 Framing (stud size, thickness and spacing, presence of X-bracing). 120
3.2.3 Fastener (type, size, spacing) 120
3.2.4 Geometry (wall aspect ratio, openings) 121
3.2.5 Type of loading (monotonic or cyclic).. 123
3.2.6 Construction techniques and anchorage details 124
REFERENCES 124

Chap ter I V
Evaluation of seismic capacity: the monotonic test 127
4.1 THE STUDY CASE: BASIC ASSUMPTIONS 128
4.2 GENERAL DESIGN PRINCIPLES... 131
4.3 TEST PROGRAM 133
4.3.1 Description of test specimens 134
4.3.2 Test set-up. 140
4.3.3 Instrumentation. 142
4.3.4 Test procedure for the monotonic test... 143
4.4 TEST RESULTS.. 144
REFERENCES 152
Conten ts iii

Chap ter V
Evaluation of seismic demand 153
5.1 MODEL OF THE HYSTERETIC BEHAVIOR OF CFSSSW SYSTEMS. 154
5.1.1 The loading branch without pinching... 155
5.1.2 The loading branch with pinching 156
5.1.3 The unloading branch ... 158
5.1.4 Calibration of the model 159
5.2 ASSESSMENT OF DEFORMATION DEMAND 166
5.2.1 Data-base of considered earthquake ground motions... 166
5.2.2 Displacement demand evaluation. 171
5.3 DEFINITION OF A LOAD HISTORY FOR A CYCLIC TESTING. 180
5.3.1 Loading histories in quasi-static cyclic loading tests: basic procedures for Multiple Step Test...180
5.3.2 Definition of the deformation history for the cyclic testing.. 185
REFERENCES 191

Chap ter VI
The cyclic test 193
6.1 TEST PROCEDURE 194
6.2 TEST RESULTS.. 197
REFERENCES 205

Conclusions 207

Appendices
Appendix A: Summary of existing experimental results.. 15 pp
Appendix B: Summary of objectives and results of existing
experimental studies.. 6 pp
Appendix C: Evaluation of seismic action and wall shear strength for
the study case.. 14 pp
Appendix D: Construction details of the sub-assembly specimen
. 10 pp
A p p en d ix E : M o n o t o n i c t e s t r e s u l t s . . 11 p p
A p p en d ix F : C y c l i c t e s t r e su l t s . . 11 p p
v

Index of figures

FIGURE 1.1: STICK-BUILT CONSTRUCTION... 6


FIGURE 1.2: PANELIZED CONSTRUCTION.. 7
FIGURE 1.3: MODULAR CONSTRUCTION 8
FIGURE 1.4: STRUCTURE OF TYPICAL STICK-BUILT CONSTRUCTION (NASFA 2000). 9
FIGURE 1.5: WALL FRAMING OF TYPICAL STICK-BUILT CONSTRUCTION (NASFA 2000)
.. 11
FIGURE 1.6: TYPICAL WALL FRAMING DETAILS (NASFA 2000). 11
FIGURE 1.7: FLOOR FRAMING OF TYPICAL STICK-BUILT CONSTRUCTION (NASFA 2000)
.. 12
FIGURE 1.8: ROOF FRAMING OF TYPICAL STICK-BUILT CONSTRUCTION (NASFA 2000)
.. 14
FIGURE 1.9: BASIC FRAMING SYSTEMS IN STICK-BUILT CONSTRUCTIONS (CSSBI 1991)
.. 15
FIGURE 1.10: CHANNEL SECTION DIMENSIONS.. 16
FIGURE 1.11: ON SITE COLD FORMING OF PROFILES 17
FIGURE 1.12: UN-REINFORCED HOLES IN WEBS OF STUDS. 18
FIGURE 1.13: WOOD-BASED PANELS PRIMARILY USED IN THE CONSTRUCTION MARKET
.. 20
FIGURE 1.14: SCREW POINT TYPES.. 23
FIGURE 1.15: SCREW HEAD TYPES... 24
FIGURE 1.16: PNEUMATICALLY DRIVEN AND POWDER-ACTUATED PINS. 25
FIGURE 1.17: SPIRAL SHANK NAILS. 26
FIGURE 1.18: LOAD-TRANSFER MECHANISMS. 27
FIGURE 1.19: CAST-IN-PLACE ANCHORS 27
FIGURE 1.20: POST-INSTALLED ANCHORS - MECHANICAL SYSTEMS. 29
FIGURE 1.21: POST-INSTALLED ANCHORS - ADHESIVE SYSTEMS 30
FIGURE 1.22: HOLD-DOWN (BRANSTON ET AL. 2003) 31
FIGURE 1.23: FOUNDATION. 32
FIGURE 1.24: INSTALL GROUND FLOOR. 32
FIGURE 1.25: ERECT WALLS33
FIGURE 1.26: INSTALL FIRST FLOOR... 33
FIGURE 1.27: INSTALL ROOF.. 33
FIGURE 1.28: DISTRIBUTION OF HORIZONTAL (WIND) LOADS IN A HOUSE CONSTRUCTION
.. 37
Ind ex of fig ures vi

FIGURE 1.29: WALL BRACED WITH STEEL SHEET X-BRACING.. 37


FIGURE 1.30: BLOCKING.. 38
FIGURE 1.31: TYPICAL INSULATION DETAILS. 41

FIGURE 2.1: MODEL OF A SIMPLE SHEAR WALL. 47


FIGURE 2.2: COMPARISON BETWEEN SEGMENT METHOD AND PSW METHOD.... 48
FIGURE 2.3: FACTORS AFFECTING CFSSSWS SHEAR BEHAVIOR.. 49
FIGURE 2.4: FAILURE MODES OF SHEATHING-TO-FRAME CONNECTIONS UNDER SHEAR
LOADING 51
FIGURE 2.5: ASSUMED SHEATHING FASTENER FORCES (EASLEY ET AL. 1982).. 54
FIGURE 2.6: BUCKLING MODES OF AN UNSHEATHED STUD.. 56
FIGURE 2.7: FLANGE BUCKLING MODEL FOR DISTORTIONAL COLUMN BUCKLING
(LANDOLFO ET AL. 2002). 58
FIGURE 2.8: BUCKLING MODES OF A SHEATHED STUD (AISI 1996). 63
FIGURE 2.9: FAILURE MODES OF SCREWED CONNECTIONS UNDER SHEAR LOADING. 68
FIGURE 2.10: FAILURE MODES OF BOLTED CONNECTIONS UNDER SHEAR LOADING.. 70
FIGURE 2.11: FAILURE MODES OF ANCHORS UNDER TENSION LOADING (HILTI 2001) 72
FIGURE 2.12: FAILURE MODES OF ANCHORS UNDER SHEAR LOADING (HILTI 2001) 74
FIGURE 2.13: MONOTONIC SHEAR LOAD-LATERAL DISPLACEMENT RESPONSE 78
FIGURE 2.14: REVERSED-CYCLIC SHEAR LOAD-LATERAL DISPLACEMENT RESPONSE 78

FIGURE 3.1: INFLUENCE OF THE SHEATHING ORIENTATION. 120


FIGURE 3.2: INFLUENCE OF THE ASPECT RATIO. 122
FIGURE 3.3: INFLUENCE OF THE OPENINGS. 122
FIGURE 3.4: INFLUENCE OF CYCLIC LOADING 123
FIGURE 3.5: CYCLIC RESPONSE FOR OSB 4-8 US C-A TEST PERFORMED BY BRANSTON ET
AL. (2003)... 124

FIGURE 4.1: TYPICAL ANALYZED STICK-BUILT HOUSE... 128


FIGURE 4.2: EUROCODE 8 TYPE 1 ELASTIC SPECTRUM ACCELERATION... 133
FIGURE 4.3: GLOBAL 3D VIEW OF THE GENERIC SUB-ASSEMBLY 135
FIGURE 4.4: A CLOSE-UP VIEW OF THE CONNECTION BETWEEN END STUDS AND
FOUNDATION 135
FIGURE 4.5: A CLOSE-UP VIEW OF THE CONNECTION BETWEEN THE FLOOR JOISTS AND
THE WALL STUDS... 135
FIGURE 4.6: CFS PROFILES (MANIFACTURED BY GUERRASIO). 137
FIGURE 4.7: SHEATHINGS 137
FIGURE 4.8: SELF DRILLING SCREWS (MANIFACTURED BY TECFI S.R.L.). 137
FIGURE 4.9: HOLD-DOWN CONNECTOR.. 138
FIGURE 4.10: (A) SHEAR AND (B) HOLD-DOWN ANCHORS (MANUFACTURED BY HILTI
Ind ex of fig ures vii

ITALIA)... 138
FIGURE 4.11: LOAD TRANSFER SYSTEM. 141
FIGURE 4.12: TEST SET-UP.. 142
FIGURE 4.13: INSTRUMENT ARRANGEMENT 143
FIGURE 4.14: UNIT SHEAR RESISTANCE VS. MEAN DISPLACEMENT CURVE.... 145
FIGURE 4.15: SPECIMEN CONDITION AT STEP 1(D=10MM).. 146
FIGURE 4.16: SPECIMEN CONDITION AT STEP 2 (MAXIMUM SHEAR RESISTENCE -
D=36MM) 147
FIGURE 4.17: SPECIMEN CONDITION AT STEP 3 (D=80MM). 148
FIGURE 4.18: SPECIMEN CONDITION AT STEP 4 (D=130MM)... 149
FIGURE 4.19: SHEAR VS. DISPLACEMENT CURVES FOR WALL 1 AND WALL 2. 151
FIGURE 4.20: SHEAR VS. DISPLACEMENT MEASURED BY HORIZONTAL LVDTS ... 151
FIGURE 4.21: SHEAR VS. DISPLACEMENT MEASURED BY VERTICAL LVDTS... 152

FIGURE 5.1: BASIC DEFINITIONS (DELLA CORTE ET AL. 1999).. 155


FIGURE 5.2: THE LOADING BRANCH WITHOUT PINCHING.. 156
FIGURE 5.3: THE LOADING BRANCH WITHOUT PINCHING (DELLA CORTE ET AL. 1999)
.. 157
FIGURE 5.4: THE LOADING BRANCH WITHOUT PINCHING (DELLA CORTE ET AL. 1999)
.. 158
FIGURE 5.5: THE UNLOADING BRANCH (DELLA CORTE ET AL. 1999)..159
FIGURE 5.6: CYCLIC TEST PROTOCOLS.. 163
FIGURE 5.7: CYCLIC RESPONSE. 164
FIGURE 5.8: NUMERICAL VS. EXPERIMENTAL CYCLIC RESPONSE FOR THE C01-4(A)
SPECIMEN.. 164
FIGURE 5.9: NUMERICAL VS. EXPERIMENTAL MONOTONIC RESPONSE. 165
FIGURE 5.10: GEOGRAPHICAL DISTRIBUTION OF THE SELECTED STRONG-MOTION
RECORDS 169
FIGURE 5.11A: ELASTIC SPECTRA OF EARTHQUAKE RECORDS FOR TYPE A SOIL. 169
FIGURE 5.11B: ELASTIC SPECTRA OF EARTHQUAKE RECORDS FOR TYPE B SOIL. 170
FIGURE 5.11C: ELASTIC SPECTRA OF EARTHQUAKE RECORDS FOR TYPE C SOIL. 170
FIGURE 5.11D: ELASTIC SPECTRA OF EARTHQUAKE RECORDS FOR TYPE D SOIL. 170
FIGURE 5.11E: ELASTIC SPECTRA OF EARTHQUAKE RECORDS FOR TYPE E SOIL.. 171
FIGURE 5.12: NUMERIC MODEL SCHEMATIZATION... 171
FIGURE 5.13: IDA PROCEDURE EXAMPLE.. 174
FIGURE 5.14A: IDA CURVES FOR LAZIO-ABRUZZO EARTHQUAKE... 175
FIGURE 5.14B: IDA CURVES FOR UMBRO-MARCHIGIANO EARTHQUAKE.. 176
FIGURE 5.15A: IDA CURVES FOR SOIL TYPE A. 176
FIGURE 5.15B: IDA CURVES FOR SOIL TYPE B. 177
FIGURE 5.15C: IDA CURVES FOR SOIL TYPE C. 177
FIGURE 5.15D: IDA CURVES FOR SOIL TYPE D. 178
FIGURE 5.15E: IDA CURVES FOR SOIL TYPE E..... 178
FIGURE 5.16: DETERIORATION OF A CFSSSW SYSTEM (BRANSTON ET AL. 2003)... 180
FIGURE 5.17: BASIC PARAMETERS IN A TYPICAL CYCLIC OF LOADING (ATC 1992).. 181
Ind ex of fig ures viii

FIGURE 5.18: DEPENDENCE OF MEAN NUMBER OF INELASTIC EXCURSIONS ON NATURAL


PERIOD AND DUCTILITY RATIO (KRAWINKLER 1996).. 183
FIGURE 5.19: DEPENDENCE OF MEAN NUMBER OF THE SUM OF NORMALIZED PLASTIC
DEFORMATION RANGES ON NATURAL PERIOD AND DUCTILITY RATIO
(KRAWINKLER 1996).. 183
FIGURE 5.20: LOADING HISTORY FOR MULTIPLE STEP TEST (KRAWINKLER 1996) 184
FIGURE 5.21: ASSUMED DEFINITIONS OF INDIVIDUAL PLASTIC DEFORMATION RANGE
.. 186
FIGURE 5.22A: RESULTS OF THE STATISTIC CHARACTERIZATION OF THE DEFORMATION
DEMAND IN TERMS OF MAXIMUM NORMALIZED DEFORMATION 188
FIGURE 5.22B: RESULTS OF THE STATISTIC CHARACTERIZATION OF THE DEFORMATION
DEMAND IN TERMS OF NUMBER OF INELASTIC EXCURSION. 188
FIGURE 5.22C: RESULTS OF THE STATISTIC CHARACTERIZATION OF THE DEFORMATION
DEMAND IN TERMS OF SUM OF NORMALIZED PLASTIC DEFORMATION
RANGES.. 189
FIGURE 5.22D: RESULTS OF THE STATISTIC CHARACTERIZATION OF THE DEFORMATION
DEMAND IN TERMS OF RATION BETWEEN THE MEAN VALUE AND THE
MAXIMUM VALUE OF THE PLASTIC DEFORMATION RANGE.. 189

FIGURE 6.1: CYCLIC TEST PROTOCOL. 196


FIGURE 6.2: NUMERICAL CYCLIC RESPONSE... 196
FIGURE 6.3: UNIT SHEAR RESISTANCE VS. MEAN DISPLACEMENT CURVE.. 198
FIGURE 6.4: BEHAVIOR OF OSB SHEATHING-TO-FRAME CONNECTIONS... 200
FIGURE 6.5: BEHAVIOR OF GWB SHEATHING-TO-FRAME CONNECTIONS. 200
FIGURE 6.6: FAILURE OF SHEATHING-TO-FRAME CONNECTIONS DUE TO SCREW HEADS
PULL THROUGH... 201
FIGURE 6.7: FAILURE OF SHEATHING-TO-FRAME CONNECTIONS DUE TO RUPTURE OF
SHEATHING EDGES 201
FIGURE 6.8: UNZIPPING OF SHEATHINGS.. 202
FIGURE 6.9: BUCKLING OF END STUDS.. 202
FIGURE 6.10: DEFORMATION OF WALLS FOR LATERAL DISPLACEMENT MORE THAN 36MM
.. 202
FIGURE 6.11: CYCLIC VS. MONOTONIC BEHAVIOR 204
FIGURE 6.12: SHEAR VS. DISPLACEMENT CURVES FOR WALL 1 AND WALL 2.... 205
ix

Index of tables

TABLE 2.1: FAILURE MODES IN A CFSSSW BRACED LATERALLY WITH WOOD BASED
PANELS.. 50
TABLE 2.2: INDICATIVE SHEAR RESISTANCE OF SCREWS.. 52
TABLE 2.3: 1997 UBC NOMINAL SHEAR STRENGTH VALUES FOR SEISMIC ACTIONS... 79
TABLE 2.4: 2000 IBC NOMINAL SHEAR STRENGTH VALUES FOR SEISMIC ACTIONS .... 80
TABLE 2.5A: 1997 UBC NOMINAL SHEAR STRENGTH VALUES FOR WIND ACTIONS .... 80
TABLE 2.5B: 1997 UBC NOMINAL SHEAR STRENGTH VALUES FOR WIND ACTIONS.. 81
TABLE 2.6A: 2000 IBC NOMINAL SHEAR STRENGTH VALUES FOR WIND ACTIONS... 81
TABLE 2.6B: 2000 IBC NOMINAL SHEAR STRENGTH VALUES FOR WIND ACTIONS... 82
TABLE 2.7: SAFETY AND RESISTANCE FACTORS PER 1997 UBC AND 2000 IBC 82
TABLE 2.8: MAXIMUM UNRESTRAINED OPENING HEIGHT PER 2000 IRC.. 83
TABLE 2.9: ADJUSTMENT FACTORS FOR APPLICATION OF THE PSW METHOD PER 2000
IRC... 84

TABLE 3.1: LISTING OF TESTS OF CFSSSW 90


TABLE 3.2: NORMALIZED SHEAR CAPACITY FOR DIFFERENT SHEATHING TYPE AND
THICKNESS 119

TABLE 4.1: MAIN STRUCTURAL DIMENSION 129


TABLE 4.2: UNIT DEAD LOADS 129
TABLE 4.3: VALUES OF PARAMETERS DESCRIBING THE EUROCODE 8 TYPE 1 ELASTIC
RESPONSE SPECTRUM 130
TABLE 4.4: MAXIMUM ELASTIC SPECTRUM ACCELERATION AND CORRESPONDING
LATERAL FORCE VALUES 133
TABLE 4.5: FULL-SCALE SPECIMEN MATERIALS AND CONSTRUCTION DATA 139

TABLE 5.1: WALL GEOMETRY AND MATERIAL DATA 160


TABLE 5.2: CYCLIC TEST PROTOCOLS 162
Index of tables x

TABLE 5.3: RESULTS OF CALIBRATION OF LOWER BOUND CURVE AND TRANSITION


PARAMETERS OBTAINED ON THE BASIS OF EXISTING CYCLIC TESTS 165
TABLE 5.4: RESULTS OF CALIBRATION OF UPPER AND LOWER BOUND CURVE
PARAMETERS OBTAINED ON THE BASIS OF MONOTONIC TESTS CARRIED-OUT
IN THE CURRENT RESEARCH 165
TABLE 5.5: DATA-BASE OF SELECTED EARTHQUAKE RECORDS 168
TABLE 5.6: IDA RESULTS 179
TABLE 5.7: PREDICTED AND EXPERIMENTAL DEMANDS FOR A BILINEAR SDOF
(KRAWINKLER 1996) 185
TABLE 5.8: RESULTS OF THE STATISTIC CHARACTERIZATION OF THE DEFORMATION
DEMAND 190

TABLE 6.1: CYCLIC TEST PROTOCOL 195


TABLE 6.2: COMPARISON BETWEEN THE DEFORMATION DEMAND PARAMETERS
OBTAINED FROM THE STATISTIC CHARACTERIZATION AND THE ADOPTED
LOADING HISTORY 197
1

Foreword

In recent years, the use of cold-formed steel members in low-rise buildings


has increased significantly in North-America, North-Europe and Australia.
The cold-formed steel products, traditionally used as secondary structures
(profiles and decks) of industrial buildings, have had in such countries a
brilliant and diffused application as main structures of residential (housing)
and commercial constructions, mainly thanks to the expected economic
advantages deriving from the significant reduction of the construction time.
The structural typologies, where cold-formed steel profiles are adopted as
main structural elements, according to the prefabrication level, can be
classified in: stick built (lowest prefabrication degree); panelized
(intermediate prefabrication degree); and modular (highest prefabrication
degree) constructions. They are generally made of steel profiles and
sheathings assembled together by means of mechanical fasteners. The profiles
are generally cold-formed lightweight galvanized steel with C (lipped
channel), U (unlipped channel), Z, and L sections. The most widely used
structural sheathings are wood-based and gypsum-based panels, but also steel
sheet sheathing may be used. Due to the presence of different structural
materials various connection typologies, as screws, welding, pins, nails, bolts,
anchors and adhesives, may be employed.
For the design of these structural systems under both vertical and
horizontal loads two approaches can by used, which are different as respect to
common structures. In the first approach, named all-steel design, the
contribution of the attached sheathings is neglected and the profiles are
considered as free-standing members. In the second, named sheathing
braced design, the interaction between profiles and sheathings is taken in-to
2

account in the evaluation of the load bearing capacity. As a consequence, in


the case of horizontal loads two basic lateral load resisting systems may be
considered. Following the all-steel design approach, the in-plane stability in
roofs, floors and walls should be provided by the use of diagonal steel flat
strap bracings. While, considering the sheathing braced design approach, if
sheathing elements have adequate strength and stiffness, and if there are
adequate connections to the profiles, then roofs, floors and walls can perform
as diaphragms.
Although today the utilization of the cold-formed steel profiles in Italy is
still limited to secondary structure, the aptitude of such constructive systems
to be quickly and easily assembled makes this structural system an interesting
alternative solution to the use of prefabricated modules (containers) for the
efficient management of emergency situations. One important example, for
the national territory, is the management of post-earthquake emergency
situations.
Such an employment of the cold-formed steel housing obviously requires
an accurate study of the seismic behavior of this structural system. The study
must be carried out according to modern and advanced knowledge of
earthquake engineering, using the recently developed performance-based
design and analysis approach. In fact, over the last decade, great emphasis has
been placed on the need to overcome limitations of current codified, force-
based and prescriptive, seismic design procedures and to implement a
probabilistic performance-based approach. The latter is based on the coupling
of multiple performance levels and ground motion intensities, thus generating
performance objectives to be satisfied. While the application of the
performance-based methodology has long been tested for classic reinforced
concrete and steel structures, its application to light cold-formed steel housing
remains largely unexplored.
The design of classic building structures is based on force-reduction factors
that, exploiting the structure own ductility, avoids collapse, safeguards human
lives and allows a relatively less expensive structural design. In case of light-
gauge cold-formed steel framed structures, the seismic weight of the building
and its content is significantly smaller, allowing the design to be carried out
with relatively very low values of the force reduction factors. This event is
particularly favorable, because of the relatively small ductility of the light-
Fo rewo rd 3

gauge structure. Consequently, cold-formed structures are more damage-


tolerant than classic structures under earthquakes having the design intensity
(i.e. earthquakes having a 10% probability of being exceeded in 50 years
according to modern design standards). However, the ability of this type of
structure to survive more violent earthquakes (rare earthquakes, for example,
those having a 2% probability of being exceeded in 50 years according to
current research trends) is currently unexplored and need to be clarified. This
requires the evaluation of seismic performance through the comparison
between the seismic demand and the seismic capacity. In fact, adequate
performance implies that the capacity exceed the imposed demand with an
adequate margin of safety. For a structural component or sub-assembly, this
means that its capacities as well as the demands imposed by earthquakes need
to be quantified.

The main objective of this dissertation is to give a contribution to the


evaluation of seismic performance of low-rise residential buildings built with
cold-formed steel members through the evaluation of the seismic demand, the
seismic capacity and their comparison.
The same reason motivates a joint research effort between the University
Federico II of Naples and the University Gabriele DAnnunzio of Chieti-
Pescara, Italy. Namely, the research project, entitled A theoretical and
experimental study on the feasibility of using cold-formed steel members in
seismic zones is supported by the Italian Ministry of University and Research
(MIUR) as a part of a more comprehensive research program, devoted to
study innovative steel structures for the seismic protection of buildings.

Although the study had general purpose, with regard to low-rise residential
buildings built with cold-formed steel members, the attention has been
focused on the stick-built construction system because it represents the basic
system for the development of more industrialized constructions (panelized
and modular constructions). In particular, considering this construction system
especially advantageous when the members-to-sheathing interaction is takes
in-to account (sheathing braced design), the research has been concerned on
the seismic behavior of cold-formed steel stud shear walls (CFSSSWs)
laterally braced with sheathings (sheathed).
4

The dissertation, as well as the research program, has been articulated in six
main phases.
The first phase of the research has been particularly devoted to study the
construction system, the building process and the basic aspects of structural
design under vertical and horizontal loads (Chapter I). The second part of the
study has been dedicated to analyze the available main design methodologies
for the evaluation of shear capacity for CFSSSW systems (Chapter II). The
review of existing literature focused on the critical examination of previous
tests on CFSSSWs, together with the definition of basic factors that influence
their lateral load response, have been the objects of the third phase (Chapter
III).
The forth phase of the research (Chapter IV) represented the fist step of the
experimental activity. In fact, it has been dedicated to evaluate the seismic
capacity for small building made by a stick-built construction system. In
particular, the design and the monotonic testing of a full-scale CFSSSW sub-
assembly specimen, it representing a realistic model of typical lateral load
resisting systems of stick-built house structures, has been included.
The fifth phase (Chapter V) of the research activity has been a theoretical
one. It has been articulated in the following three main steps: (1) development
of reliable mathematical models calibrated using the experimental results of
the monotonic test, in order to capture the complex hysteretic response of
CFSSSWs; (2) evaluation of the deformation demand to stud shear wall
systems obtained on the basis of a large database of acceleration records; and
(3) definition of a load history for a successive cyclic testing.
Finally, in the sixth phase (Chapter VI) the second step of the experimental
program was performed. In fact, in this stage a specimen nominally identical
to that tested under monotonic loading has been subjected to fully reversing
cyclic horizontal displacements according to the definition of the deformation
history parameters carried out in the theoretical phase.
5

Chapter I
Low-rise residential buildings built with
cold-formed lightweight steel members

The use of light steel framing as a method of low-rise house construction


has increased significantly in industrialized countries in recent years. In some
European countries this is from a small base (Lawson & Ogden 2001). In
other countries, such as USA, Canada, Australia and Japan, there is a well-
developed industry with well-established practices. In fact, in the USA, 15000
steel homes were built in 1993, 75000 were constructed in 1996 and it is
expected that in the year 2002 the steel house constructions will be about
375000 (Yu 2000).
The present generation of steel framed house construction systems has been
derived from traditional timber framed house construction systems without
any significant change in either the architecture or the interior and exterior
finishing materials.
This Chapter describes the construction system (Sections 1.1 through 1.5);
the erection procedure (Section 1.6); the fundamental problems of structural
design (Section 1.7); the additional considerations on fire resistance, durability
and insulation (Section 1.8); the advantages and the limitations of the low-rise
residential buildings built with cold-formed lightweight steel members
(Section 1.9).
Low-rise residential buildings built with co ld -formed lightweight steel me mbers 6

1.1 BASIC CONSTRUCTION SYSTEMS

There are three basic systems of constructions for low-rise house which are
named (Landolfo et al. 2002):
x stick-built constructions;
x panelized constructions;
x modular constructions.

Stick-built constructions
The stick-built construction (see Fig. 1.1) is characterized by the lowest
prefabrication degree and it is the most common method used for built cold-
formed steel (CFS) frames, because it is the same as the familiar stick-built
wood construction method. In fact, in these systems the wood members have
replaced with appropriate CFS members. In general, the steel members may
be cut to length on job-site with a hand-held saw. Similar to wood
construction, steel components are fastened together on the floor surface into
wall sections and tilted into positions. Once the wall sections are structurally
connected together, exterior and interior sheathing materials are applied. The
construction time is short, the steel structure of a typical detached house
requiring between 2 and 4 days.
The advantages of stick constructions are (Grubb and Lawson 1997):
x the system can accommodate larger constructions tolerances;
x the workshop facilities associated with modular constructions are not
necessary;
x simple constructions techniques without heavy lifting equipment may
be used;
x members can be densely packed for transport.

Figure 1.1: Stick-built construction.


Low-rise residential buildings built with co ld -formed lightweight steel me mbers 7

Panelized constructions
In panelized construction (see Fig. 1.2), wall and floor sub-frames and roof
trusses may be prefabricated in the factory. The panels are lifted into position
on the job-site and fastened together generally by bolting to form the required
building geometry. This method of construction is particularly efficient when
there is repetition of panel types and dimensions. In contrast to stick-built
construction, in the panel construction exterior sheathings, thermal insulation
and some of the lining and finishing materials may also be applied to the steel
sub-frames in the factory.
The main advantages of panel construction are (Grubb and Lawson 1997):
x speed of erection;
x factory standards of quality control during fabrication of the units;
x reduction of site labor costs,
x scope for automation in factory production.

Figure 1.2: Panelized construction.

Modular constructions
Prefabrication of wall, floor and roof units is taken a stage further in
modular construction (see Fig. 1.3). Here, lightweight steel boxes, which may
for example be hotel room units, are completely prefabricated in the factory
before being delivered to the site. It is usual for all internal finishes, fixtures
and fittings, and even the carpets, curtains and furniture, to be fitted in the
Low-rise residential buildings built with co ld -formed lightweight steel me mbers 8

factory. On job-site, the units are stacked side-by-side and up to several


storeys high on prepared foundations and service connections made to form
the complete structure. Nowadays, many hotels and motels are built in this
way.

This chapter is particularly focused on the stick-built construction system


that certainly represents the more used structural solution. Besides, this
structural solution is also the basic system for the development of more
industrialized constructions, like the panelized construction and the modular
construction.

Figure 1.3: Modular construction.

1.2 STICK-BUILT CONSTRUCTIONS

Figure 1.4 shows typical stick-built construction (NASFA 2000).


The basic elements of a stick-built construction are:
x foundations;
x walls framing;
x floors framing;
x roofs framing.
Low-rise residential buildings built with co ld -formed lightweight steel me mbers 9

Figure 1.4: Structure of typical stick-built construction (NASFA 2000).

Foundations
Two kinds of foundations are typically used in a stick-built construction:
x poured concrete walls foundations;
x slab-on-grade foundations.

For both foundation types, up-lift loads represent the critical problem. In
fact, when the horizontal loads are large, the overturning moment is high, and
the tendency for the wall to overturn is elevated. Consequently, for
counterbalance the up-lift force the foundation must be capable of distributing
the load and the weight of the foundations itself must be exceeds that of up-lift
force.

Walls framing
Wall framing generally have two and possibly three separate functions.
Their primary function is to carry vertical load from the floors and roof above
(load bearing wall). In external walls, they also have to resist the lateral
pressure from the wind and transmit this to the floor and roof diaphragms and
to the foundations. In addition, certain walls may also form part of the system
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 10

resisting in-plane forces from wind or seismic shear (bracing wall). A typical
wall framing system, component and terminology are shown in Figure 1.5 and
some constructional details are shown in Figure 1.6 (NASFA 2000).
The main structural components of wall framing construction are:
stud: vertical structural element of a wall assembly, which support
vertical loads and/or transfer lateral loads (typically lipped
channel sections).
wall track: horizontal member used such as top and bottom plate for walls
(typically unlipped channel sections).

Wall studs are typically C-shaped galvanized CFS sections placed with
their flanges in contact with the wall sheathing. The profiles are usually on a
module ranging approximately from 300mm to 600mm. The wall sheathing
may be, for example, a gypsum or wood fiber based board, plywood or, in
some cases steel sheet or corrugated sheathings. If this wall sheathing has
adequate strength and stiffness and if there is adequate attachment to the studs,
then the axial load bearing capacity may be substantially increased by the
resulting structural interaction. This is mainly as a consequence of the
resistance provided against lateral buckling modes.
Evidently, wall studs can be designed as free-standing members without
taking advantage of the influence of the other elements of the wall
construction and modern codes of practice allow this on the basis of
calculations alone. Inclusion of the stiffening influence of the walls has to be
semi-empirical based on the interpretation of test results.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 11

Figure 1.5: Wall framing of typical stick-built construction (NASFA 2000).

Figure 1.6: Typical wall framing details (NASFA 2000).


Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 12

Floors framing
Floor framing construction is typically lightweight and dry. However, there
are circumstances where heavier constructions are specified, usually to meet
requirements for fire and/or sound isolations.
Generally, the floor construction is constituted by cold-formed lightweight
steel profiles with a wood-based sheathing. The profiles are usually on a
module that coincides with studs on the supporting elevations. A typical steel
floor system, component and terminology are shown in Figure 1.7 (NASFA
2000).

Figure 1.7: Floor framing of typical stick-built construction (NASFA 2000).

The main structural components of floor framing construction are:


floor joists: horizontal member that supports floor loads (typically
lipped channel sections).
floor track: horizontal member used as band or rim joists for
flooring systems (typically unlipped channel sections).
web stiffener: additional material that is attached to the web to
strengthen the member against web crippling. Also
called bearing stiffener.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 13

Joist designs are rarely based on the bending resistance of the cross-
sections, in fact, normally the serviceability criteria of deflection and vibration
control the design. Steel floors can sometimes resonate with the vibrations
induced by human footsteps and, although this does not usually produce
structural damages, it can be a source of discomforts. This problem tends to
become more acute when lightweight construction is used.
In some cases, the floor can be realized by trapezoidally profiled steel deck
supported on primary beams and carrying a wood-based walking surface or,
more infrequently, it can be realized by composite (steel-concrete) deck
supported on primary beams.

Roofs framing
For the roof framing construction, steel truss with similar proportions to a
conventional gang-nailed timber truss is often about six times as strong as the
timber equivalent and is uneconomic at modest spans and load levels. For this
reason, when a habitable roof space is not required steel framed houses are
often completed with timber roof trusses. Steel trusses can, however, be used
economically when they are spaced at wider centers with purlins spanning
between the trusses. All-steel solutions become much more favorable when a
usable roof space is required. In Figure 1.8 (NASFA 2000) a typical solution
is shows, in which attic frames are connected in a conventional manner. An
alternative system could use sandwich panels either spanning from caves to
ridge or supported on intermediate purlins.
The main structural components of wall framing construction are:
rafter: structural framing member (usually sloped) that
supports roof loads (typically lipped channel sections).
ceiling joist: horizontal structural cold-formed lightweight steel
profiles that supports a ceiling and attic loads (typically
lipped channel sections).
ridge member: horizontal member placed at intersection between the
top edges of two sloping roof surfaces.
fascia: member applied to the rafter ends as an edge member
for attachment of roof sheathing, exterior finishes, or
gutter.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 14

Figure 1.8: Roof framing of typical stick-built construction (NASFA 2000).

The others main structural components of floor, wall and roof framing
constructions are (NASFA 2000):
header: horizontal built-up structural framing member used over wall,
roof or floor openings to transfer loads above opening to
adjacent framing member.
blocking: solid block or piece of material placed between structural
members to provide lateral bracing as in bridging and/or edge
support for sheathings.
bridging: bracing or blocking placed between joist to provide lateral
support.
flat strap: sheet steel cut to a specified width without any bends. Typically
used for bracing and transfer of loads by tension.

There are two basic framing systems in stick-built constructions that are
usually defined as (CSSBI 1991):
x platform system;
x balloon system.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 15

In platform construction, as shown in Figure 1.9a, the structure is built


storey by storey so that each floor can serve as a working platform for the
construction of the floor above. The walls are not structurally continuous and
loads from the walls above are generally transferred through the floor
structure to the walls below. The wall studs are connected to horizontal tracks
top and bottom and the floor joists are seated on the top track of the studs
below. Sufficient stiffening is incorporated in the connection to ensure the
safe transfer of vertical load though the floor construction.
In balloon construction, as shown in Figure 1.9b, the wall panels are
continuous over two or more storeys and the floors are attached to them. It
follows that loads from the floors above pass down the load-bearing studs
without affecting the incoming floor construction.

a: Platform framing system b: Balloon framing system

Figure 1.9: basic framing systems in stick-built constructions (CSSBI 1991).

1.3 PROFILES

Steel members used in residential buildings in the construction of floors,


walls and roof are predominantly cold-formed lightweight galvanized steel
members. The profiles are cold-formed from structural quality steel sheet with
a yield strength ranging from 230MPa to 350MPa, an ultimate tensile strength
ranging from 310MPa to 480MPa. Moreover, the ultimate tensile strength-to-
yield strength ratio shall be at least 1.10 and the total elongation shall be at
least 10%.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 16

As illustrated previously, the main profiles are commonly divided into


lipped and unlipped channel sections. The lipped channel sections are usually
used for joist, stud and rafter profiles, while the unlipped channel sections are
normally utilized for track profiles. These profiles are identified using:
x web depth;
x flange width;
x thickness.

In this dissertation a lipped channel section (see Fig. 1.10a) is identified as


C web depth x flange width x lip size x thickness (all dimensions in mm),
while a unlipped channel section (see Fig. 1.10b) is identified as U web depth
x flange width x thickness (all dimensions in mm).

a: Lipped channel section dimensions. b: Unlipped channel section dimensions.

Figure 1.10: Channel section dimensions.

In North America, stud type members typically range from about 90 to


200mm in web depth, with an about 40mm flange width, and with thicknesses
ranging from 0.8 to 2.5mm. Joist type members have a flange width ranging
from around 40 to 45mm and range in web depth approximately from 140 to
310mm with thicknesses ranging from 0.8 to 2.5mm.
In Europe, dimensions of studs and joists are similar to that used in North
America except for the thickness, which in European profiles ranges from 1.2
to 3.2mm.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 17

The profiles may by cold-formed and cut in the factory or on site.


Generally, it is more convenient to order precut cold-formed members because
in this way it is possible to save time and material. But, if the exact
dimensions of structure are unknown, it is usually preferable cut, and in some
case, cold form the profiles on site, as shown in Figure 1.11.

Figure 1.11: On site cold forming of profiles.

The profiles normally have holes in their webs for water lines, gas pipe,
and electrical wiring, as shown in Figure 12. As far as concern the hole
dimension, it is possible to consider the following classification.
Un-forced holes (small hole): if the depth of the hole, measured across the
web, does not exceed about 40mm, and the length of the hole, measured along
the web, do not exceed about 100mm. In these cases, it is not necessary to
take any precaution.
Reinforced holes (intermediate holes): if web holes violates any of the
requirements reported at the above point. In these cases, the web shall be
strengthened with an appropriate solid steel plate, stud section, or track
section.
Big holes: if the depth of the hole, measured across the web, exceeds about
75% of the depth of the web, and/or, the length of the hole, measured along
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 18

the web, exceeds about 150mm. In these cases, the profiles shall be designed
in accordance with accepted engineering practices taking in to account of
presence of the hole.

Figure 1.12: Un-reinforced holes in webs of studs.

1.4 SHEATHINGS

The most widely used types of sheathing are:


x wood-based panels;
x gypsum-based boards,
x steel sheathings.

Neglecting the well-known steel sheet sheathing, some discussion about on


the main types of wood-based and gypsum-based products used in steel
framed house constructions are given hereafter.

1.4.1 Wood-based panels


All wood-based panels are composite materials made of wood elements
such as veneer, particles, flakes, or fibers and some adhesive. The main types
of structural wood-based panels are (Faherty and Williamson 1999):
x veneer panels;
x particleboard panels.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 19

The most commonly known structural wood-based panels are glued


assembly of thin sheets of wood called veneer products. The main veneer
panels are plywood and Laminated Veneer Lumber (LVL). The type of veneer
panel generally used for structural purposes in building constructions is the
plywood.
Plywood is a panel or sheet product composed of thin sheets of veneer, or
plies (see Fig. 1.13a). Plywood always has an odd number of layers, each
layer consisting of one or more plies. These layers are glued together with the
grain of adjacent layers placed at right angles to one another. The face layers
are oriented with their grains parallel to the long dimension of the panel, with
the crossbands thus having their grains parallel to the short sides of the panel.

The term particleboard is any board that is made from wood particles in
the form of flakes, fibers, strands, wafers, chips, etc. that are combined with
synthetic resins or some other bonding systems. The generic product covers
many types of board, differing in particle size, orientation and position and
includes flakeboard, fibreboard, Oriented Strand Boards (OSB) and
waferboard. Particleboard panels primarily used for wall bracing and flooring
in the construction market are oriented strand boards and fiberbond.
Oriented strand board consists of wood strands bonded with adhesives to
form a mat (see Fig. 1.13b). Like the veneer in plywood, these mats are
layered and oriented perpendicular to each other for maximum strength,
stiffness and stability. The individual strands are typically 7 to 10 centimeters
long. Oriented strand board is widely used as construction sheathing, as the
web material for wood I-joists, as the structural membranes of structural
insulated panels, and in an increasing number of other applications. Its
popularity is reflected in the large number of new oriented strand board mills
being built, and in the plant expansions of several current oriented strand
board major producers. In the USA, the value of oriented strand board has
production increased of 50% over the last 10 years.
Fibreboard panels typically used in building constructions include
hardboard that is used for exterior cladding and insulation boards that are used
for cladding in buildings to insulate for heat or cold (see Fig. 1.13c). Some
boards are produced with a coating of sealant to help protect from water.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 20

Wood-based products are manufactured in sheets or panels with about 1.2 x


2.4m as a generally standard size and with thickness varying from 8 to 30mm.
Other sizes may be available for specific scopes, although widths other than
1.2m are rare.
The main product standards for wood-based panels, which describe
minimum properties that the product must have to be acceptable, are: the
PS1-95 (1983) and the ANSI A208.1-93 (1993) used in the USA, and the
EN 313-1 (1997) and EN 313-2 (2000) for the plywood and the EN 309
(1993) for the particleboard used in Europe.

a. Plywood panel b. Oriented strand board panel c. FibreBoard panel

Figure 1.13: Wood-based panels primarily used in the construction market.

1.4.2 Gypsum-based boards


The gypsum-based boards are a family of panel products consisting of a
non-combustible core, primarily of gypsum, with a paper surfacing on the
face, back and long edges. The main types of gypsum-based sheet products
used in building constructions are standard gypsum wallboard, water-resistant
gypsum board, fire-resistant gypsum board, and foil-backed gypsum board.
The water-resistant gypsum board is a wallboard specially processed for
use as a base for ceramic and other non-absorbent wall tiles in bath and
shower areas. These products consist of a solid set, water-resistant gypsum
core covered with durable water-resistant backing paper and green-colored
face paper.
The fire-resistant gypsum board is a wallboard that provides greater fire
resistance. In fact, in this board type a specially formulated core provides fire
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 21

resistance ratings when used in tested assemblies. Facing paper is generally


colored pink.
The foil-backed gypsum board is a wallboard panel consisting of a gypsum
core enclosed in heavy natural-side and strong liner paper on the backside to
which aluminum foil is laminated.
Gypsum boards are manufactured with 1.2m generally standard widths,
with lengths varying from 2 to 3.5m, and with thicknesses ranging from 6 to
23mm.
The main product standards for gypsum-based boards, which fix minimum
properties for these products, are the ISO 6308 (1980) and the ASTM
C1396/C1396M-02 (2002).

1.5 CONNECTIONS

In a residential CFS structure, taking into account the various structural


materials, it is possible to consider the following connections typology (AISI
1993):
x steel-to-steel connections;
x steel-to-sheathing (wood-based and gypsum-based sheathings)
connections;
x steel-to-concrete connections.

For these connections typology the most common connector systems are:
x screws (steel-to-steel and steel-to-sheathing connections);
x welding (steel-to-steel connections);
x pins (steel-to-steel, steel-to-sheathing and steel-to-concrete
connections),
x nails (steel-to-sheathing connections),
x bolts (steel-to-steel connections);
x anchors (steel-to-concrete connections);
x adhesives (steel-to-sheathing connections).

Moreover, a connection system named hold-down, typically used in the


stick-built wood constructions for transferring the uplift loads, is also
employed in the stick-built CFS constructions for the same scope.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 22

Others relatively new techniques, known a press-joining or clinching


(Pedreschi and Sinha 1996) and rosette-joining (Machelainen et al. 1998) can
advantageously be used for steel-to-steel connections, particularly when the
CFS systems are prefabricated in the factory (in the cases of panelized and
modular constructions).

1.5.1 Screws
The most common fastener for steel framing is the self-tapping screw. In
one operation, it can drill the hole and fasten the material to steel framing.
These screws come in a variety of styles to fit a vast range of requirements.
For exterior applications, screws are available finished with zinc, cadmium or
co-polymer coatings.
Screws are available in diameters ranging from 3.5mm to 6.5mm. Lengths
typically vary from 12mm to as much as 75mm depending on the application.
Screws are generally 10mm to 12mm longer than the thickness of the
connected materials so that a minimum of three threads shall extend beyond
the connected material. It is important that the drill point be as long as the
material thickness to achieve an efficient drilling.
Two specific point types are commonly used, as shown in Figure 1.14.
x Self-Drilling Screws: Externally threaded fasteners with the ability to
drill their own hole and form their own internal threads without
deforming their own thread and without breaking during assembly.
These screws are used with 0.8mm steel or thicker.
x Self-Piercing Screws: Externally threaded fasteners with the ability to
pierce relatively thin steel material. They are commonly used to attach
rigid materials, such as gypsum wallboard, to 0.8mm or thinner steel.

Self-drilling screws are manufactured in a variety of head configurations to


meet specific installation needs. The main head types, as shown in Figure
1.15, are:
x pan head: This common head configuration generally fastens studs to
track, connects steel strapping or furring channels to studs or joists,
and steel door frames to studs;
x modified truss head: An extremely low profile head commonly used
for attaching metal lath to steel framing (also referred to as lath head);
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 23

x hex washer head: This is also a common style for penetrating steel and
is more commonly used on thicker steel materials. The washer face
provides a bearing surface for the driver socket, assuring greater
stability during driving;
x oval head: Used when an accessory that will be attached to the framing
has oversized holes or requires a low profile appearance;
x flat head: Designed to countersink without causing splintering or
splitting of wood flooring and finishes;
x trim head: Used for fastening wood trim or other thick dense finish
materials to steel studs. The small head pierces into the trim material,
allowing easy finishing with minimal disturbance of the material
surface;
x bugle head: Designed to countersink slightly in gypsum wallboard or
sheathing, plywood or finish material without crushing the material or
tearing the surface. Leaves a flat, smooth surface for easy finishing;
x low-profile head: Maintains a pleasing appearance at fastening points
where the application does not require jut heads. This style should be
used when rigid sheathings or finish material is to be installed over top
of the screw head;
x wafer head: Larger than the flat or bugle head, the wafer head is used
for connecting soft materials to studs. The large head provides an
ample bearing surface to achieve a flat, clean, finished appearance.

Figure 1.14: Screw point types.


Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 24

Figure 1.15: Screw head types.

1.5.2 Welding
Welding can be used to prefabricate steel framing components into panel
assemblies and trusses. Welding is also commonly used in the attachment of
site-fabricated or erected panels and for connecting shelf angles to steel
framing.
For helping assure quality installation, all work should be completed by
welders qualified in the welding of sheet steel. Where the studs, joists and
framing accessories have been fabricated from galvanized or painted steel, the
coating will normally be burned away by the welding processes. Where
required, weld areas should be re-touched with the appropriate paint or cold
galvanizing to maintain corrosion resistance.

1.5.3 Pins
Two specific types are commonly used, as shown in Figure 1.16:
x pneumatically driven pins;
x powder-actuated pins.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 25

Pneumatically driven pins are quite new to the CFS framing industry.
They use techniques similar to nail guns for wood, and are commonly used for
nailing plywood to steel. They are also available for fastening thicker cold-
formed steel to concrete.
These fasteners are available with mechanical or electro-zinc plating, or co-
polymer coatings, depending on corrosion resistance requirements. Head
diameters range from 6mm to 10mm, shank diameters vary from 2.5mm to
8mm, and lengths from 12.5 to 200mm.
The shanks may be step-down, knurled, or smooth, as shown in Figure
1.16a. Step-down shanks are generally used on thicker materials for greater
load carrying capacities. Pins are typically supplied in bulk, in collated strips,
or in coils, depending on tool requirements.

Powder-actuated pins are frequently used to fasten CFS framing members


to concrete or to hot-rolled steel. For example, they are widely used to attach
steel track to concrete slabs and foundations. Powder-actuated pins systems
can be threaded, headed with a wisher or standard headed, as shown in Figure
1.16b.
The holding strength of fastener in concrete depends on the compressive
strength of the concrete, sank diameter, depth of penetration, and spacing and
edge conditions. Headed or threaded drive pins are available with knurled
shanks to increase holding power in structural steel. Fasteners loaded in
tension may require washers to prevent the fastener from prematurely pulling
through the sheet steel.

headed with a standard


step-down knurled smooth threaded
wisher headed
a: Pneumatically driven pins. b: Powder-actuated pins.

Figure 1.16: Pneumatically driven and powder-actuated pins.


Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 26

1.5.4 Nails
CFS framing members may be also fastened to sheathing with nails.
Generally, nails are spiral shank nails (see Fig. 1.17). They are case hardened
and phosphate treated to etch their surface for increased holding power. They
can be used to fasten plywood subfloors and underlayment to steel.

Figure 1.17: Spiral shank nails.

1.5.5 Bolts
Bolts commonly used for connecting hot-rolled steel members can be also
used to fasten CFS framing to other steel component. Generally, pre-drilling
of holes is necessary for this fastening system. Washers should be provided
for oversized or slotted holes. When the bolt is loaded in tension, washers may
be required to prevent premature pull-through of the anchor.

1.5.6 Anchors
Anchor commonly used for connecting concrete to steel can be separated in
two major categories (Hilti 2001):
x cast-in-place anchors, those are placed before the concrete is cast;
x post-installed or drilled-in anchors, those are installed into hardened
concrete.

Each of these categories is composed of a variety of different anchors, all


of which transfer loads from the attachment to the concrete in a variety of
ways. The primary load-transfer mechanisms (see Fig. 1.18) under tension
loading are friction, keying, and bonding as well as combinations. For shear
loading, it is keying or direct bearing.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 27

Figure 1.18: Load-transfer mechanisms.

Generally, design engineers specify cast-in-place anchors (see Fig. 1.19) if


they know beforehand where anchors are to be installed. The main cast-in-
place anchors are the standard fasteners (headed bolts, J bolts and L bolts,
stud-welded plates), proprietary shapes (threaded inserts, proprietary anchors
and proprietary shapes), through bolts (usually sleeved), and special shapes
(shear lugs, channel bars). All these anchors use keying as load transfer
mechanism. Many of these types of anchors have special uses. Stud-welded
plates and shear lugs provide large shear resistance, while channel bars give
specific attachment capability. J and L bolts are typically used for anchoring
sill plates to foundations, but have a tendency to straighten and pull-out under
higher tension loading.

Figure 1.19: Cast-in-place anchors.

With the development and improvements of rotary hammer drills and


carbide-tipped bits, the user has the capability to install many different kinds
of post-installed anchors (see Fig. 1.20) in hardened concrete virtually
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 28

anywhere that is accessible to the drill. Post-installed anchors can be divided


into two major types, depending on the method of transferring load into the
concrete. They are mechanical systems and bonded or adhesive systems.
Anchors can be also be cross-classified according to their load carrying
capability into heavy-duty, medium duty, and light-duty.
The mechanical anchoring systems cover the range from heavy-duty to
light-duty capacities. The main kinds of mechanical anchors are undercut
anchors and expansion anchors. The undercut anchors have been on the
market for about 20 years, they are excellent for use under both static and
dynamic loads. They obtain their holding capacity through keying, that is,
direct bearing on the concrete, and, under proper installation, can withstand
very high load without slipping out of the drilled hole. They are the preferred
anchors for use where cracks in tension zones of the concrete can be expected
to occur. The expansion anchors have been available for at least 30 years.
There are two basic types that are distinguished by their operating principles.
The first, torque-controlled expansion anchors, are installed by inserting the
anchor into the drilled hole, and applying the prescribed setting torque to the
head or nut. A cone at the bottom of the anchor is pulled up into the concrete
with local crushing, and providing both friction and localized keying as load-
transfer mechanism. There are several types available that vary significantly in
their ability to resist static and dynamic loads. The heavy-duty sleeve anchor
can resist dynamic loads as well as function well in expected cracks in
concrete. The second major type of expansion anchor is the displacement-
controlled expansion anchor. Two primary examples are the drop-in and the
self-driller. Drop-in anchors are installed in the predrilled hole by use of a
setting tool that drives a plug into the expansion portion of the anchor. The
lower section of the anchor is expanded into concrete, which experiences local
crushing. The second type has cutting teeth on the lower end and drills its own
hole. The anchor is driven onto an expansion plug that expands the lower
portion of the anchor into the concrete. These anchors derive their holding
capacity from friction and keying.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 29

Figure 1.20: Post-installed anchors - mechanical systems.

Bonded resin or adhesive anchors (see Fig. 1.21) were generally introduced
into the construction market about 20 years ago. Bonded systems use a
combination of adhesive bond and micro keying into the pores of the concrete.
Early systems used polyester resin, epoxies, and later, vinyl ester resins. In
recent years, a larger variety of resins have been developed that have
individual advantages, such as use in high temperatures, low temperatures,
damp and wet holes, etc. For two component epoxy systems, the ratio to resin
is critical. Prepackaged cartridge systems assure that the proper mixing is
obtained.
While a variety of installation methods are used, most are two component
resin systems that anchor threaded rod into predrilled holes. Most will resists
dynamic loads, both seismic and fatigue, but documentation in the form of test
reports should be obtained. Bonded or adhesive anchoring systems are not
well suited for cracked tensile zones of concrete since about the bonding is
lost.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 30

Figure 1.21: Post-installed anchors - adhesive systems.

1.5.7 Adhesives
Adhesives are generally optional when screws are used in attaching
subfloors or underlayment to steel joists, and are necessary when nails are
employed. The use of adhesives can provide for long-term bonding capacity
that contributes to a stable assembly. Adhesives are also utilized in attaching
drywall materials and paneling to steel studs or when laminating panels
together, serving to eliminate some, but not all, of the mechanical fasteners.

1.5.8 Hold-down
Uplift forces exist on shear walls because the horizontal forces are applied
to the top of the wall. These uplift forces try to lift up one end of the wall and
push the other end down. In some cases, the uplift force is large enough to tip
the wall over. Uplift forces are greater on tall short walls and less on low long
walls. Bearing walls have less uplift than non-bearing walls because gravity
loads on shear walls help them resist uplift. Shear walls need hold-down
devices at each end when the gravity loads cannot resist all of the uplift. The
hold-down device then provides the necessary uplift resistance. In fact, hold-
downs transfer uplift forces from the end of the wall through a floor to a wall
or foundation below.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 31

Hold-downs that connect a shear wall to the foundation are bolted or


screwed to the end stud and they are anchored to the footings using cast-in-
place or post-installed anchors that are connected to the hold-down. This
anchors transfer the uplift force from the wall down into the foundation. This
minimizes the tendency of the wall to overturn (Fig. 1.22).
Hold-downs that connect two walls through a floor come in pairs, one
above and one below. The hold-downs are bolted or screwed their respective
end stud. The uplift forces are then transferred from the wall above to the wall
below through a threaded rod that is bolted to the hold-downs.
Similar to bolted hold-downs, metal straps can be used as hold-downs. The
strap must be long enough to pass through the floor framing and be attached to
the end studs so that the required number of screws or bolts is provided
between the strap and the end stud. The strap should also be stiff and straight
to reduce slippage.

Figure 1.22: Hold-down (Branston et al. 2003).

1.6 THE BUILDING PROCESS

The main phases of the building process for a stick-built structure as that
shown in Figure 1.4 are:
x foundation;
x install ground floor;
x erect ground walls;
x install first floor;
x erect first floor walls;
x install roof.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 32

Some of these building phases are shown in Figures 1.23 through 1.27.
These are considered to be self-explanatory.

Slab-on-grade foundation. Poured concrete walls.

Figure 1.23: Foundation.

Floor joists are fastened to the top track of a steel- Joist track is installed over the ends of the floor
framed wall with screws driven through the joist joists. The track is fastened to each joist with self-
flanges. Additional web stiffeners fastened with drilling screws and is anchored to foundation with
self-drilling screws are added to each joist-to-track anchor bolt or other connection as required.
connections as required.

Figure 1.24: Install ground floor.


Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 33

Stud tracks are installed over the ends of the Sheathings are screwed to both studs and stud track
studs. Each top and bottom tracks are fastened for a secure connection between sheathing and cold-
to each studs with self-drilling screws. formed frame.

Figure 1.25: Erect walls.

Floor joists are fastened to the track with the Floor decking is fastened to the floor joists and tracks
same procedure reported in Figure 1.24. generally using self-drilling screws.

Figure 1.26: Install first floor.

In roof framing structures pre-assembled steel truss can be used. Alternately, the profiles can be screwed
with the same procedure used for floor joists.

Figure 1.27: Install roof.


Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 34

1.7 BASIC PROBLEMS OF STRUCTURAL DESIGN

As well as in a common structure, the basic requirements in a stick-built


structure are two:
x the transmission of vertical loads to the foundation;
x the transmission of horizontal loads to the foundation.

1.7.1 Vertical loads


The design of a stick-built structure under vertical loads is relatively less
complicated. In fact, rafters and joists carry essentially bending loads, whereas
studs carry basically compression axial load.
Rafters, joists and studs can be designed by using one of two different
approaches, as follows:
x all-steel design;
x sheathing braced design.

In the all-steel design, the contribution of the attached sheathings is


neglected and the profile is considered as free-standing members without
tacking in to account of the influence of the interactions with the sheathing
elements. In this way, the calculated resistance of the profile is based solely on
the condition of lateral bracing of the member. The advantage of this approach
is that the load bearing capacity, being compression or bending, is determined
without having to rely on the structural bracing capability of the sheathing
material.
In the sheathing braced design, the presence of attached sheathings is
taken in-to account in the valuation of the compressive resistance calculation.
In fact, if these elements have adequate strength and stiffness, and if there are
adequate connections to the members, then the load bearing capacity may be
significantly improved. This is principally a result of the resistance provided
against global buckling modes. In modern code (ENV 1993-1-3 1996, AISI
1996, AS/NZS 4600 1996), it is allowed to take in-to account the members-to-
sheathings interaction with semi-empirical calculations based on the
interpretation of test results.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 35

1.7.2 Horizontal loads


The design of a stick-built structure under horizontal loads (wind or
seismic loads) is more problematic. The components of this problem are two
(Davies 1998), as shown in Figure 1.28 for wind load. (1) The floor and roof
frames must be able to act as diaphragms in plan to transmit horizontal loads
on the wall frames parallel to the direction of the loads. (2) The wall frames
must be able to transmit the in-plane forces from the floor and roof frames
down to the foundations. It is found that significant uplift forces may arise at
the leeward end of the wind-frames. Providing an adequate tie to adequate
foundations is another aspect of this problem.

There are two main methods existing for providing in-plane stability in
floor, roof and wall:
x the use of diagonal bracings (X- and K-bracings);
x the use of floor, roof, or wall structure as diaphragms.

When X-bracings are used (see Fig. 1.29), flat straps of thin steel pass
under the bottom flange of the joists or rafters in the floor or roof and over the
external faces of the studs in the walls. Due to their high slenderness, these
straps act only in tension. They are connected to the primary structure at their
ends in order to transfer the calculated tensile force and at their intersection in
order to reduce their tendency to sag.
An alternative system is K-bracing which takes the form of C-sections
fixed within the depth of the primary structure. These members act in either
tension or compression and, together with the adjacent members form a truss.
As a result, adjacent members with a larger section than the norm and
appropriate connections are required.
Evidently, diagonal bracing should be used wherever possible and current
practice is to use this method for the floor and roof and for walls where there
are few door and window openings. However, the currently popular housing
style results in elevations that contain much opening. In such cases, diagonal
bracing may not be possible and it is necessary to search for another solutions.

Floor, roof or walls constructions often are made of panels supported by


steel joists, rafters or studs respectively. Only when the connections between
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 36

sheathings and supporting members are properly designed, these structures


can perform as a diaphragm and the structural system is termed a box
system. In this way, it is possible to take advantage of suitable sheathings
materials.
Floors and roofs can be considered horizontal, simple supported
diaphragms, whereas walls can be regarded as vertical, cantilevered
diaphragms. For full diaphragms design, it is necessary to analyze sheathings,
connections, diaphragm edge members, and tie-down behavior. In fact, a
diaphragm acts in a manner analogous to a deep beam, where the sheathings
act as a web, resisting shear, while the diaphragm edge members perform the
function of flange, resisting bending. The diaphragm edge members are
commonly called chords, and may be joists, trusses, studs, etc.
Due to the great depth of most diaphragms in the direction parallel to that
of load, and to their way of assembly, the behavior differs slightly from that of
the usual beam.
Shear stresses have been proven essentially uniform across the depth of the
diaphragm, rather than showing significant parabolic distribution as in the web
of a beam.
Chords in a diaphragm carry all flange" stresses acting in a simple tension
and compression, rather than devising these stresses significantly with the
web.
As in any beam, consideration must be given to bearing stiffeners,
continuity of webs and chords, and to web buckling, which is normally
resisted by the framing members.
Diaphragms vary considerably in load-carrying capacity, depending on
whether they are blocked or unblocked. Blocking consists generally of lipped
or unlipped channel sections placed between the joists or other primary
structural supports for the specific purpose of connecting the edges of the
panels (see Fig. 1.30). Systems that provide support framing at all sheathing
edges are also considered blocked. The reason of blocking in diaphragms is to
allow connection of sheathings at all edges for better shear transfer. Buckling
of unsupported sheathing edges, with consequent reduced load bearing
capacity, controls unblocked diaphragm behavior.
The three main part of a diaphragm are the web, the chords, and the
connections. Since the individual pieces of web must be connected to form a
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 37

unit, since the chord members in all probability are not a single pieces, since
web and chords must be fixed so that they act together, and since the loads
must have a path to other elements or to the foundation, connections are
critical to good diaphragm action. Their choice actually becomes a major part
of the design procedure.

Figure 1.28: Distribution of horizontal (wind) loads in a house construction.

Figure 1.29: Wall braced with steel sheet X-bracing.


Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 38

sheathing joist

blocking
flat strap

Figure 1.30: Blocking.

1.8 ADDITIONAL CONSIDERATIONS

1.8.1 Fire resistance


Due to several aspects, the issues of CFS structures in fire are much
different in comparison to those for hot-rolled structures.
Cold-formed members are generally used in low-rise constructions where
the fire resistance requirements may be less onerous.
In CFS profiles, the thin material produce more limited thermal capacity in
comparison to that for hot-rolled members.
For hot-rolled members, the critical steel temperature, at which structural
failure may by anticipated, is in the region 450-700C, depending on the
structural system and the load level in the vicinity of the fire. In absence of
any calculation model, a conservative approach for CFS profiles is to design
the fire protection systems to achieve a critical temperature of 350C.
In conventional hot-rolled construction, the fire resistance may be provided
by protection applied to the members in the form of boards or sprayed on
materials which delay the temperature increase in the steel by providing
thermal insulation. Otherwise, CFS profiles in residential buildings are more
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 39

often part of a wall or floor structure such that they do not require specially
applied fire protection of this sort. In fact, walls and floors are usually built up
of layers of materials such as mineral wool and gypsum board in order to meet
the requirements of sound and thermal insulation and these materials generally
also have favorable properties of fire insulation. Thus, these elements of the
construction can be built up in such a way that they meet all of the
performance requirements with regard to resistance to fire, sound and thermal
loss in a consistent manner.

1.8.2 Durability
With good quality detailing, the durability of a CFS structure should be the
same of traditional construction. There are two characteristics for
guaranteeing good durability:
x avoid the accumulation of moisture in the external envelope;
x avoid bringing incompatible materials into contact with each other.

These aspects are related because most types of material incompatibility


are aggravated in the presence of moisture. Main forms of material
incompatibility to avoid are the followings.
x Bimetallic contact: this is only likely if the cladding or fasteners are of
a different material to the main structure.
x Cementitious materials-to-steel contact: galvanized steel reacts with
fresh cementitious materials with the formation of hydrogen bubbles
on the surface of the zinc which reduce the bond with the steel. This
reaction can be controlled by the addition of soluble chromates to the
cementitious material but a more reliable solution is to chromate the
surface of the galvanized steel.
x Gypsum plaster-to-steel contact: moist gypsum plaster attacks
galvanized steel both in the fresh condition and if it dries out and
subsequently becomes wet. The G275 coating of normally galvanized
steel will prevent corrosion of the metal.
x Wood-to-steel contact: wood is a corrosive material that can be made
more corrosive by the protective treatments given to it. Corrosion of
metals in contact with timber is a function of the moisture content of
the timber. Moreover, there is less risk with soft woods than with hard
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 40

woods. Correct seasoning of the wood significantly reduces the


problem.

To limit the corrosion of steel and to improve the durability there are
general precautions which can be followed.
x To ensure that the cladding is sufficiently impermeable and to detail it
in such a way that rainwater does not penetrate the structural frame of
the building.
x To ensure that the external envelope performs thermally in such a way
that condensation of moisture from humid air is not possible.

1.8.3 Thermal and acoustic insulation, air and moisture permeability


In the design of walls and roofs of lightweight construction thermal and
acoustic insulation as well as air and moisture permeability are most
important.
The need to achieve an adequate level of thermal insulation is now well
understood and the requirements seem to become more onerous year by year.
A fundamental aspect is the need to avoid thermal bridging. It is less well
understood that this theoretical thermal performance can be reduced by air
infiltration, particularly at the roof space of a building. It is also well
understood that it is necessary to avoid condensation of moisture within the
structure and that a primary requirement is an appropriately placed vapor
barrier.
Control of thermal insulation, air and moisture permeability is largely a
matter of good detailing. For most climates, the first essential is a high level of
thermal insulation and this generally involves the provision of an appropriate
thickness of insulation material as well as a cavity. The insulation material is
generally mineral wool although a wide variety of suitable alternatives is
available. It is also essential to avoid thermal bridges and moisture
penetration. Some insulation details of wall and roof construction are shown
in Figure 1.31.
Sound transmission and attenuation is an other most significant problem
because the occupants of buildings are rightly sensitive to noise intrusion from
other occupants of the same building or from outside. Fiberglass or cellulose
insulation installed in floor assemblies can significantly reduce the sound
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 41

transmission. Moreover, good construction practices can also result in


significant reduction in sound transmission through CFS floors.

Insulation of a wall. Insulation of a roof.

Figure 1.31: Typical insulation details.

1.9 ADVANTAGES AND LIMITATIONS

There are numerous advantages that can be obtained by using CFS


members in residential construction. The main benefits are the followings.
x High structural performance: The CFS members offer one of the
highest load capacity-to-weight ratio of any building component
currently on the market. The strength of steel permits the use of greater
spans than would be possible with wood components, often requiring
less material. For example, joists and studs can be positioned 600mm
on center and support the same load as wood framing 400mm on
center.
x Lightweight: The lightweight of framing components reduces total
building and seismic loads. Moreover, it is easy to handle, contributing
to reduced labor costs and worker fatigue.
x High quality: The steel profiles are produced in factory, consequently
they are more straight, uniform and consistent in quality than wood,
hand-laid masonry and in-situ concrete elements.
x Design flexibility: The variety of size and thicknesses of steel
contribute to flexibility. For example, to obtain a desiderate design it
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 42

can reduce the width of a joist but compensate with a heavier gauge
steel and not change the spacing of members.
x Construction speed. It is possible to erect a CFS structure in much less
time than a traditional structure.
x Dry constructions: Except for the foundation, the use of the CFS
structures allow dry constructions, therefore, dirt, dust and general lack
of precision associated with hand-laid masonry, in-situ concrete are
avoided.
x Ease of wiring installation: Holes are normally preformed simplifying
the installation of the wirings.
x Recyclables: Cold-formed steel members are easily recycled.
x Common appearance: Once exterior and interior finishes are installed,
a traditional and a CFS house are indistinguishable from each other.
x Incombustibility: Steel does not burn and prevents the spread of fire.

The principal limitations that restraining the use of CFS framing in


residential house are the followings.
x Thermal performance: The thermal performance of steel using
standard framing techniques is below that of traditional framing due to
thermal conductivity of steel. Remedies, such as applying external
insulation to the frame, add to the cost and the difficulty of fastening
with screws.
x Fastening: The installation labor cost of CFS framing is higher than
for wood framing. The primary labor component where productivity
lags is the fastening of steel members using self-drilling screws.
Development of pneumatic and clinching technology holds promise of
improving fastening.

REFERENCES

AISI (1993) Fasteners for residential steel framing. AISI (American iron and Steel
Institute). Washington DC.
AISI (1996) Cold-formed steel design manual. AISI (American Iron and Steel
Institute). Washington DC.
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 43

ANSI A208.1-93 (1993) Mat-formed wood particleboard. ANSI (American National


Standards Institute). New York.
ASTM C1396/C1396M-02 (2002) Standard specification for gypsum board. ASTM
(American Society for Testing and Materials). West Conshohocken, PA, USA.
AS/NZS 4600 (1996) Cold-formed steel structures. AS/NZS (Australian
Standards/New Zealand Standards). Sydney.
Branston, A., Boudreault, F., Rogers, C.A. (2003) Testing on steel frame / wood
panels shear walls. Progress Report, Departement of Civil Engineering and Applied
Mechanics, McGill University. Montreal.
CSSBI (1991) Canadian Steel framing design manual, CSSBI (Canadian Sheet Steel
Building Institute). Cambridge, Ontario.
Davies J.M. (1998) Light gauge steel framing for housing construction. In
Proceedings of the 2nd International conference on thin-walled structures.
Singapore.
EN 309 (1993) Wood particleboards Definition and classification. CEN (European
Committee for Standardization). Bruxelles.
EN 313-1 (1997) Plywood. Classification and terminology - Classification. CEN
(European Committee for Standardization). Bruxelles.
EN 313-2 (2000) Plywood - Classification and terminology - Terminology. CEN
(European Committee for Standardization). Bruxelles.
ENV 1993-1-3 (1996) Eurocode 3: Design of steel structures Part 1-3: General
rules - Supplementary rules for cold formed thin gauge members and sheeting. CEN
(European Committee for Standardization). Bruxelles.
Faherty, K.F. and Williamson, T.G. (1999) Wood Engineering and construction
handbook (Third edition). McGraw-Hill.
Grubb, P.J. & Lawson, R.M. (1997) Building design using cold formed steel sections:
construction detailing and practice. SCI Publication P165. The Steel Construction
Institute. Ascot, Berkshire, United Kingdom.
Hilti (2001) Hilti North America product technical guide.
ISO (1980) Gypsum plasterboard Specification. ISO (International Organization for
Standardization). Geneva.
Landolfo, R., Di Lorenzo, G, Fiorino, L. (2002) Attualit e prospettive dei sistemi
costruttivi cold-formed. Costruzioni metalliche No.1
Lo w-rise residen tial b uildings built with co ld -formed lightweight steel me mbers 44

Lawson, R.M. & Ogden, R.G. (2001) Recent developments in the light steel housing
in the UK. In Proceedings of the 9th North steel conference of construction institute.
Helsinki.
Mkkelkinen, P., Kesti, J., Kaitila, O. (1998) Advanced method for light-weight steel
truss joining. In Proceedings of the Nordic steel construction conference '98. Bergen,
Norway.
NASFA (2000) Prescriptive Method For Residential Cold-Formed Steel Framing
(Year 2000 Edition). NASFA (North American Steel Framing Alliance). Lexington,
KY, USA.
Pedreschi, R.F. & Sinha, B.P. (1996) The potential of press-joining in cold-formed
steel structures. Construction and building materials, Vol.10.
PS1-95 (1983) Voluntary product standard PS1-95 for commercial and industrial
plywood, U.S. Department of commerce, National institute of standard and
technology. Gaithersburg, MD, USA.
Schuster, R.M. (1996) Residential applications of cold-formed steel members in
North America. In Proceedings of the 5th International Structural Stability Research
Concilium SSRC. Chicago.
Yu, W.W. (2000) Cold Formed Design (3rd Edition). John Wiley & Sons. New
York.
45

Chapter II
Design of cold-formed steel stud shear
walls

Cold-formed/light gauge steel (CFS/LGS) buildings typically use wood


structural panel sheathing fastened to repetitive member framing to provide an
adequate lateral force resisting system (LFRS) to resist seismic and wind
loads. Therefore, the evaluation of the shear strength values used in the design
of these systems are critical to the accuracy and efficiency of an engineering
analysis and design for the cold-formed steel stud shear walls (CFSSSW).
The objective of this Chapter is to provide a concise basic design reference
for CFS/LGS LFRSs. The Chapter is organized in three main parts. In the first
part (Section 2.1) the main shear wall design methodologies are illustrated.
The semi-analytical evaluation of shear strength for CFSSSWs is examined in
the second part (Section 2.2). In the third part (Section 2.3) the units shear
design values for CFSSSWs reported in the main Codes (UBC 1997, IBC
2000) are examined.
46 Chapter II

2.1 MAIN DESIGN METHODOLOGIES FOR SHEAR


STRENGTH OF CFSSSWs

Two main design methodologies for the evaluation of the shear capacity of
a CFSSSW exist, following methods similar to that used for wood-framed
walls:
x segment method;
x perforated shear wall method.

The segment method is a traditional shear wall design methodology. The


basic principles are demonstrated in the simple shear wall segment model of
Figure 2.1.
This model is commonly used in current engineering practice to design
wall segments in a wood frame building to resist lateral loads. In this simple
approach, the effects of contributions from various connections and wall
portions that are not part of the designed LFRS are neglected, as shown in
Figure 2.2a. Therefore, while the method provides a simple analysis, it
requires greater amounts of connection for each shear wall segment that
decreases the construction speed and increases the costs. This method of shear
wall design is most appropriate for heavily loaded walls and those with many
large openings that effectively break a wall line into segments. An example
would be the design of shear wall segments to either side of a garage opening
that supports more than a roof load above. In other less demanding situations
this process will tend to provide a conservative design with the largest
amounts of hold-down connectors.
This method is described in numerous technical resources on wood-framed
shear wall design (Hoyle & Woeste 1986, Ambrose & Dimitry 1987, Beyer 1993,
APA 1996, SSTD 10-97 1997).
Design of co ld -fo rmed steel stud shear walls 47

unit shear v

shear
force V
Basic relations:

h V=vw
typical T=C=Vh/w
overturning
anchor

v
T C
w

Figure 2.1: Model of a simple shear wall.

In the perforated shear wall method the resistance of shear walls with
un-restrained openings can be determined with reasonable accuracy (about
5%) by an empirical method known as the Perforated Shear Wall (PSW)
design method (Dolan & Heine 1997a, b, Dolan & Johnson 1997a, b, NAHB
Research Center 1997, 1998).The PSW design method is very simple, and
only requires that a fully-sheathed wall line with perforations for windows and
doors be restrained at the ends with a hold-down bracket or adequate corner
framing in lower capacity shear walls (Dolan & Heine 1997c), as shown in
Figure 2.2b. For determining the shear wall capacity, all that is needed is the
unit shear value for the shear wall construction, the area of wall openings, the
length of full-height wall segments, and the overall length of the wall. These
values are inputs to a simple two-step equation that gives the overall wall
capacity without the use of internal connection detailing or hold-downs.
In particular, the PSW design method for wood-frame shear walls
appearing in the IBC is based on the Sugiyama & Matsumotos (1994)
equation. This equation gives the ratio of the strength of a shear wall with
openings to the strength of a fully sheathed shear wall without openings (F):
48 Chapter II

r
F (2.1)
3  2r
where r is the sheathing area ratio as defined in the following equation:
1
r (2.2)
Ao
1
h Li
where Ao is the total area of openings, h is the height of the wall and Li is the
length of the full height wall segment.

Wall 1 Wall 2 Wall 3 Wall

w1 w2 w3 w

(a) V = v (w1 + w2 + w3) (b) V=Fvw

Figure 2.2: Comparison between Segment method (a) and PSW method (b).

For both the Segment and the PSW methodology the evaluation of shear
wall capacity of a fully sheathed shear wall without openings is required. This
evaluation may be achieved by a semi-analytical approach based on the model
of a simple shear wall or by the nominal shear value tables for specific wall
configurations obtained on the basis of experimental results.

2.2 SEMI-ANALYTICAL EVALUATION OF SHEAR


STRENGTH

The shear capacity of a CFSSSW is quite complex and depends by many


factors. If a shear wall is laterally braced by sheathings this factors may by
grouped as follows (Fig. 2.3):
x strength of the sheathings;
x strength of the sheathing-to-frame connections;
Design of co ld -fo rmed steel stud shear walls 49

x strength of the frame (stud buckling strength);


x strength of the frame-to-foundation connections.

Frame

Sheathing-to-frame
connections Frame-to-foundation
connections

Figure 2.3: Factors affecting CFSSSWs shear behavior.

A different failure mode is associated to each factor reported above.


Following the limit states design philosophy, the smallest value obtained from
all of these possible failure modes will control the shear capacity of the wall
(v), as defined by the following relation:
v min^v S ; v S  F ; v F ; v F  F ` (2.3)
in which vS, vS-F, vF and vF-F are the shear strengths associate to sheathing,
sheathing-to-frame connections, frame and frame-to-foundation connections
failure, respectively.
In most cases, a shear wall fails due to the sheathing-to-frame collapse. In
fact, all the CFSSSW components are designed according to capacity design
principles, in such a way to promote the development of the full shear strength
of sheathing-to-frame connections. For this reason, the studs (frame) are
designed to avoid failure due to buckling in end studs. Analogously, the hold-
down and shear anchors (frame-to-foundation connections) are designed to
prevent either shear failure at the base of the walls or failure due to
overturning. Moreover, the types of sheathings generally used in the
50 Chapter II

CFSSSWs (wood-based and gypsum-based panels) are able to avoid the


sheathing collapse. Consequently the evaluation of strength for the sheathing
failure usually is not performed (vS >> max {vS-F ; vF ; vF-F}).
Strength calculations for the different failure modes of a CFSSSW laterally
braced with wood-based sheathings, as schematised in Table 2.1, are detailed
below.

CFSSSW Components Failure modes


Sheathings Neglected
Sheathing-to-frame connections (S-F) Bearing in the steel frame (bs)
Tilting of the screw (ts)
Screw shear (ss)
Bearing in the wood panels (bp)
Frame (F) Stud buckling (sb)
Tension frame-to-foundation connections (F-F)N Hold-down (hd)
Shear in hold-down-to-frame connection (hd-fr)
Tension in the hold-down-to-foundation connection (hd-fo)
Shear frame-to-foundation connections (F-F)V Anchor-to-frame connection (a-fr)
Anchor-to-foundation connection (a-fo)

Table 2.1: Failure modes in a CFSSSW braced laterally with wood based panels.

2.2.1 Strength of the sheathing-to-frame connections (S-F)


The wood-based sheathing-to-steel frame screw connections are subjected
to shear forces consequently their failure limit modes include, as reported in
Figure 2.4:
x bearing failure in the steel frame;
x tilting failure of the screw;
x screw shear failure;
x bearing failure in the wood panels.

Therefore, the shear strength of a sheathing-to-frame connection (FS-F) is


equal to the minimum of the shear strengths associated to each failure mode,
as defined by the following relation:
FS  F min^FS  F ,bs ; FS  F ,ts ; FS  F ,ss ; FS  F ,bp ` (2.4)
where:
FS-F,bs: is the bearing strength of the screws for steel members;
Design of co ld -fo rmed steel stud shear walls 51

FS-F,ts: is the tilting strength of the screws;


FS-F,ss: is the shear strength of the screws;
FS-F,bp: is the bearing strength of the screws for wood panels.

tilting failure of the screw


Bearing failure in the steel frame

screw shear failure


bearing failure in the wood panels

Figure 2.4: Failure modes of sheathing-to-frame connections under shear loading.

Bearing failure in the steel frame (FS-F,bs)


Because cold formed steel frame are relatively thin (0.61.5mm) compared
with the thickness of wood-based sheathings (1020mm) normally used in
CFSSSW structures, it is possible to calculate the bearing strength of the
screws for steel members (FS-F,bs) with the following equation:
FS  F ,bs D t f d f u ,s (2.5)
where:
D: is a coefficient defined as follows:
D = 2.1 in Eurocode 3 - Part 1.3 (ENV 1993-1-3 1996)
D = 2.7 in AISI (1999)
t f: is the thickness of the cold-formed steel member;
d: is the nominal diameter of the screw;
fu,s: is the tensile strength of the cold-formed steel member.

Tilting failure of the screw (FS-F,ts)


The tilting strength of the screws (FS-F,ts) is defined as follows:
3.2 t 3f d f u ,s
0.5
FS  F ,ts for ENV 1993-1-3 (1996) (2.6)
4.2 t 3f d f u ,s
0.5
FS  F ,ts for AISI (1996) (2.7)
52 Chapter II

Screw shear failure (FS-F,ss)


The shear resistance of screws must be documented by testing and
provided by the manufacturer. For avoiding a brittle and sudden fracture of a
screw subjected to a shear force, the shear strength of the screws (FS-F,ss) must
be larger than the bearing strength (FS-F,bs) and tilting strength (FS-F,ts) of the
screws (FS-F,ss 1.2 FS-F,bs and FS-F,ss 1.2 FS-F,ts in ENV 1993-1-3 (1996) or
FS-F,ss 1.25 FS-F,bs and FS-F,ss 1.25 FS-F,ts in AISI (1996)). Some indicative
values of the characteristic shear resistance of screws made of hardened or
stainless steel are reported in Table 2.2.

Nominal diameter (mm) Hardened steel Stainless steel


4.8 5.2 4.6
5.5 7.2 6.5
6.3 9.8 8.5
8.0 16.3 14.3

Table 2.2: Indicative shear resistance of screws (kN/screw).

Bearing failure in the wood panels (FS-F,bp)


The bearing strength of the screws for wood panels (FS-F,bp) is defined by
the following equation (Faherty & Williamson 1999):
t d f
FS  F ,bp 3.5 C D s u b,w (2.8)
KD
where:
CD: is the duration factor, which is equal to 1.6 for wind and
earthquake loads;
KD: is a coefficient that depends on the diameter of the screw:
KD = 2.2 for D 4.3mm;
KD = 10 D + 0.5 for 4.3mm < D < 6.4mm;
KD = 3.0 for D > 6.4mm;
ts : is the thickness of the wood-based sheathing;
du: is the unthreaded shank diameter of the screw;
fb,w: is the dowel bearing strength of the wood-based sheathing.
Design of co ld -fo rmed steel stud shear walls 53

Shear strength of the wall associated to sheathing-to-frame connections


strength (vS-F)
Two methods exist to calculate the shear strength associated to sheathing-
to-frame connections:
x Easley et al.s method;
x simplified method.

The Easley et al.s method (Easley et al. 1982) is a design procedure that
relates the shear capacity per unit length (vS-F) with the shear strength of the
sheathing-to-frame connections (FS-F), based on equilibrium. Based on the
results of observation tests performed on wood-frame shear walls with
plywood panels, the fasteners forces in a typical sheathing of a loaded shear
wall were assumed to be directed as shown in Figure 2.5. In particular:
x The fastener forces in the sheathing ends were assumed to have both x-
and y-components. The x-components (Fex) were assumed uniform in
the x-direction. The y-components (Feyi) were assumed proportional to
the distances of the fasteners from the sheathing center line (xei).
x The fastener forces in the sheathing sides (Fs) were assumed to be
uniform and to act only in the y-direction (along the stud). The Fs
forces were assumed to be proportional to their distances from the
sheathing center line.
x The fastener forces in the interior studs (Fsi) were assumed to act only
in the y-direction (along the stud) and to be proportional to the
distances xsi.

Indicating with:
a: the length of the sheathing;
h: the height of the wall;
ne: the number of the end fasteners;
ns: the number of the side fasteners, excluding those at the end;
nsi: the number of the fasteners in each interior stud, excluding those
at the end;
m: the number of the interior studs.
54 Chapter II

Interior
Fasteners

Figure 2.5: Assumed sheathing fastener forces (Easley et al. 1982).

the relations between the side fastener forces (Fs), the end fastener forces (Fei)
and the shear force per unit length acting on the shear wall (va) are the
followings:
Fs h
Ds (2.9)
va E
0.5
Fei a 2 2
h
D ei  2 x ei (2.10)
va ne aE

where:
4 I e  2n si I s
E ns 
a2
ne
2
Ie x
i 1
ei

m
2
Is x
i 1
si
Design of co ld -fo rmed steel stud shear walls 55

Defining with Dmax the largest value between that obtained from Ds and Dei,
the shear strength per unit length (vS-F) is obtained as follow:
1
vS F FS  F (2.11)
D max
in which FS-F is defined in Equation (2.4).
With the simplified method it is assumed that only the end screws resist
horizontal load, and consequently, the contribution of side and interior
fasteners is ignored. According to this hypothesis the shear strength per unit
length (vS-F) is obtained from the value of sheathing-to-frame connections
shear strength (FS-F), multiplied by the number of the end fasteners per unit
length (ne):
v S  F n' e FS  F (2.12)

2.2.2 Strength of the frame (stud buckling strength) (F)


Considering that the end studs are subjected to high compression force due
to overturning, it is possible that these members will fail due to stud buckling.
The sheathing, usually connected to studs by screws, significantly restrains
the buckling behavior of the studs. Therefore, the strength of the sheathed stud
wall system is significantly larger than that of unsheathed studs.

Unsheathed stud
In general, considering an unsheathed stud it can present three different
types of buckling: local, distortional and global (flexural and/or torsional)
buckling or their coupling.
When global buckling occurs, any cross-section of the stud moves as a
rigid body with a half-wavelength comparable to that of the stud. This type of
buckling is governed by the global slenderness of the stud and normally
implies the collapse of the structure.
Local buckling is characterized by a relatively short half-wavelength of the
order of magnitude of individual plate elements, while the fold lines remain
straight.
Distortional buckling implies the rotation and the translation of the multiple
elements of the cross-section. This bucking mode occurs at a half-wavelength
56 Chapter II

intermediate to local and global mode buckling. Also, the cross section gets
much more distorted than in the local one.
Both local and distortional modes do not imply the collapse of the
structure. In fact, in these cases post-buckling strength is developed.
Figure 2.6 illustrates these three modes of buckling for an unsheathed stud.

Figure 2.6: Buckling modes of an unsheathed stud.

The behavior of cold-formed steel members is characterized by high non-


linearity; consequently, its accurate prediction is feasible only through a finite
element analysis that takes in to account both mechanical and geometrical
non-linearity (Della Corte et al. 2003, Fiorino et al. 2003). As a result of this
difficulty, the design methodologies currently assumed in many Codes adopt
semi-empirical approaches that predict the ultimate strength through the
determination of the elastic buckling load and considering a series of ultimate
strength curves (Fiorino & Landolfo 2001).

LOCAL BUCKLING
Local buckling is typically treated by ignoring any interaction between
elements (flanges, web and lips). Each element is considered independent
from the others and classical plate buckling equations based on isolated
simply supported plates are generally used. In this approach, termed element
Design of co ld -fo rmed steel stud shear walls 57

model, each element of the section is predicted to buckle at a different stress.


The critical stress of each element is obtained by the well-known Euler
formula:
2
S 2E t
V cr kV (2.13)
12(1  Q 2 ) b
where b and t are respectively the width and the thickness of the considered
element, E and Q are respectively the Youngs modulus and the Poissons ratio
of the material and kV is the buckling factor which depends on the type of
element (double or simply supported element) and on the stress distribution
along the element. The buckling factor kV is equal to 4 for a stiffened element
(double supported element) and to 0.43 for an unstiffened element (simply
supported element) under uniform compression.
In order to better predict the actual local buckling stress, semi-empirical
methods including elements interaction have been proposed by Batista (1988)
and Shafer & Pekz (2001). In these approaches the expressions for kV are
determined by empirical close-fit solutions to finite strip analysis results.
For the determination of the axial strength under local buckling (Nb,l)
effective widths and effective cross-section properties are generally used.
According to the semi-empirical Von Karmans approach, the non-uniform
distribution of stresses, which crop up in the post-buckling range, can be
replaced by an equivalent uniform stress distribution equal to the yield limit
stress (fy) acting on the effective width of the plate (beff). The effective width
formula usually adopted is the well-known semi-empirical formula proposed
by Winter:
beff U b (2.14)
where:
U 1 if O p d 0.673
0.22 1
1 
U if O p ! 0.673 (2.15)
O
p Op

in which the normalized plate slenderness O p is given by:


fy
Op (2.16)
V cr
58 Chapter II

The reduced properties of effective plate elements in compression may be


combined with the full width of plate element in tension to give an effective
cross-section for use in strength calculations.

DISTORTIONAL BUCKLING
Distortional buckling usually involves rotation of each flange and lip about
the flange-web junction in opposite directions as shown in Figure 2.6. The
web undergoes flexure at the same half-wavelength as the flange buckle and
the whole section may translate in a direction normal to the web also at the
same half-wavelength as the flange and web buckling deformations. The web
buckle involves single curvature transverse bending of the web.
A general model for the determination of the elastic distortional buckling
stress under axial compression has been originally developed by Lau &
Hancock (1987). Figure 2.7 shows this analytical model that is based on a
flange buckling where the flange is treated as a compression member
restrained by a rotational and a transational spring. The rotational spring
stiffness kI, represents the torsional restraint from the web and the transational
spring stuffiness kx, represents the torsional restraint to transational movement
of the cross-section. Lau & Hancock showed that the transational spring
stiffness kx, does not have much significance and the value of kx was assumed
to be zero. The key to evaluating this model is to consider the rotational spring
stiffness kI, and the half buckling wavelength O, while taking account of
symmetry.

Figure 2.7: Flange buckling model for distortional column buckling (Landolfo et al. 2002).
Design of co ld -fo rmed steel stud shear walls 59

The Authors gave a detailed analysis in which the effect of the local
buckling stress in the web and of shear and flange distortion were taken into
account in determining expressions for kI and O. This gives rise to a rather
long and detailed series of explicit equations for the distortional buckling
stress that, not-withstanding their cumbersome nature, are included in the
Australian/New Zealand Code (AS/NZS 4600 1996).
A similar set of explicit equations bas also proposed by Schafer (2001) and
will be incorporated in future AISI Specifications.
For the determination of the axial strength under distortional buckling
(Nb,d), Hancock (1985) developed expressions on the basis of effective widths
for distortional buckling.

COUPLED BUCKLING
For taking in-to account the interaction between local and global buckling,
the axial strength (Nb) has to be based upon the effective cross-section,
calculated for uniform compression.
In Eurocode 3 - Part 1.3 (ENV 1993-1-3 1996) a well-known Ayrton-Perry
formula is used for the calculation of the axial strength:
N b F Aeff f y (2.17)
1
F d1 (2.18)
I  I  O 2
2 0.5

I 0.5>1  D O  0.2  O 2 @ (2.19)


where:
Aeff: is the area of effective cross-section at uniform
compression at yield limit stress;
Aeff f y
O is the normalized slenderness;
N cr
Ncr: is the minimum elastic critical axial force for global
(flexural and/or torsional) buckling for the gross cross-
section;
D: is the imperfection factor corresponding to the
appropriate buckling curve:
D= 0.34 for lipped channel section (curve b);
D= 0.49 for unlipped channel section (curve c).
60 Chapter II

In AISI (1996) the axial strength (Nb) for concentrically loaded


compression members is calculated with the following equation:
N b Ae f n d Aeff f y (2.20)
where:
Ae: is the area of effective cross-section at uniform
compression at stress fn;
fn 0.658 f
O2c
y if Oc 1.5
0.877
fn fy if Oc > 1.5
O2c
fy
Oc is the normalized slenderness;
V cr
Vcr: is the minimum elastic critical stress for global
(flexural and/or torsional) buckling considering the
gross cross-section.

NEW METHODOLOGIES: DIRECT STRENGTH METHOD AND WHOLE SECTION


APPROACH

Schafer & Pekz (1998) have recently proposed a new procedure which
works only with the gross properties of a member and can take into account
the interaction between local and global buckling and also the interaction
between distortional and global buckling.
Use of the Direct Strength Method for columns requires (1) the
determination of the elastic buckling axial load (Ncr) of the member and (2)
using that information in a series of ultimate strength curves to predict the
strength (Nb) (Fiorino 2000).
For the determination of the elastic buckling load, as an alternative to the
traditional analytical solutions, numerical solutions based on the whole
section approach may be used to accurately calculate the elastic buckling
behavior necessary for step (1). In fact, besides to the traditional finite element
method (FEM) some friendly computer programs, mainly based on the finite
strip method (FSM) (CU-FSM 2003, THIN-WALL 2003) or the generalized
Design of co ld -fo rmed steel stud shear walls 61

beam theory (GBT) (Davies & Leach 1994a, b) have been purposely written
for the determination of the local, distortional and global elastic critical loads.
The procedure employed to calculate the axial strength is the same
underlying empirical assumptions as the effective width method used in the
main Specification. In fact, the axial strength (Nb) is a function of the elastic
buckling load (Ncr) and the yield load (Ny).
In particular, the axial strength (Nb) is the minimum of axial strength for
global (flexural and/or torsional) buckling (Nb,g), local buckling (Nb,l) and
distortional buckling (Nb,d):
N b min^N b, g ; N b,l ; N b,d ` (2.21)
The axial strength for global buckling (Nb,g) is:
N b, g U g N y (2.22)
with:
Og2 0.877
Ug 0.658 if Og d 1.5 or U g if Og ! 1.5 (2.23)
Og2
Ny
in which Og is the normalized global slenderness and Ncr,g is the
N cr , g
minimum of elastic critical axial loads for flexural, torsional and flexural-
torsional buckling.
The axial strength for local buckling (Nb,l) is:
N b,l U l N b, g (2.24)
with:
0.15 1
Ul 1 if Ol d 0.776 or U l 1  if Ol ! 0.776 (2.25)
Ol Ol
0.4
N
in which Ol b,l is the normalized local slenderness and Ncr,l is the
N cr ,l
elastic critical axial load for local buckling.

The axial strength for distortional buckling (Nb,d) is:


N b ,d U d N y (2.26)
with:
62 Chapter II

0.25 1
Ud 1 if Od d 0.561 or U d 1  if Od ! 0.561 (2.27)
Od Od
0.6
N
in which Od y is the normalized distortional slenderness and Ncr,d is
N cr ,d
the elastic critical axial load for distortional buckling.

Sheathed stud
In a sheathed stud the bending strength and diaphragm action due to the
presence of the sheathings increase the load-carrying capacity considerably
(Miller and Pekz 1993, 1994).
In Eurocode 3 - Part 1.3 (ENV 1993-1-3 1996) only the diaphragms-
bracing effect on buckling of beams is considered, while the AISI (1996)
Specification considers the effect of sheathing material on buckling of
columns and beams. In particular, for the columns the AISI (1996) design
requirements are limited only to those studs that have identical wall material
attached to both flanges.
In the evaluation of the axial strength (Nb) of a sheathed stud the AISI
(1996) Specification follows the Diaphragm stiffness mechanical model.
According to this approach the load-carrying capacity is governed by:
x column buckling of studs between fasteners in the plane of the wall
(Fig. 2.4a); or
x overall column buckling of studs (Fig. 2.4b); or
x shear failure of the sheathing.

Therefore, the axial strength (Nb) is equal to the minimum axial strength
associated to the first two failure modes listed above, as defined by the
following relation:
N b min^N bf ; N bo ` (2.28)
where:
Nbf: is the axial strength associated to the column buckling of the stud
between fasteners;
Nbo: is the axial strength associated to the overall column buckling of
the studs.
Design of co ld -fo rmed steel stud shear walls 63

Moreover, it is needed to avoid the shear failure of the sheathing.

(a) column buckling of studs between fasteners (b) overall column buckling of studs

Figure 2.8: Buckling modes of a sheathed stud (AISI 1996).

COLUMN BUCKLING OF STUDS BETWEEN FASTENERS


In case of column buckling of studs between fasteners in the plane of the
wall, the failure mode can present local, distortional, global (flexural and/or
torsional) buckling or their coupling depending on the geometric configuration
of the cross section and the spacing of fasteners. Consequently, in this case the
evaluation of the axial strength (Nbf) is based on the same methodologies
illustrated for unsheathed studs without considering any interaction with the
sheathing material in accordance with Equation 2.20 and considering an
unbraced length equal to two times the distance between fasteners.

OVERALL COLUMN BUCKLING OF STUDS


The overall column buckling of studs braced by sheathing material has
been mainly studied at Cornell University (Yu 2000). In particular, the earlier
AISI provisions were developed on the basis of the Pincus & Fischer (1966),
Errera et al. (1967), Apparao et al. (1969), Errera & Apparao (1976) and
Simaan & Pekz (1976) research. In these studies, the sheathings were
attached either on both flanges or on one flange only of the studs. In addition,
64 Chapter II

both the shear rigidity and the rotational restraint of the sheathings were
considered.
For the purpose of simplicity, in the AISI Specification only the case of
identical sheathings attached on both flanges is considered. Moreover, the
rotational restraint provided by sheathings is neglected.
According to these hypotheses, in AISI (1996) the axial strength (Nbo) for
concentrically loaded compression members is calculated in accordance with
Equation 2.20 with the elastic critical stress (Vcr) value obtained as follows:
x for singly symmetric C-sections, Vcr is the smallest of the following
elastic critical stresses:
V cr V cr ,ip  Qa (2.29)
V V  V cr ,t ,Q  4 E V cr ,op V cr ,t ,Q
2
cr ,op  V cr ,t ,Q  cr ,op
(2.30)
V cr
2E

x for doubly symmetric sections, Vcr is the smallest of the following


elastic critical stresses:
V cr V cr ,ip  Q a (2.31)
V cr V cr ,op (2.32)
In the above formulas:
Vcr,ip is the elastic critical stress for flexural buckling in the
plane of the wall considering the gross cross-section
and an unbraced length equal to the length of the stud;
Vcr,op is the elastic critical stress for flexural buckling in the
out of plane of the wall considering the gross cross-
section and an unbraced length equal to the length of
the stud;
V cr ,t ,Q V cr ,t  Qt
Vcr,t is the elastic critical stress for torsional buckling
considering the gross cross-section and an unbraced
length equal to the length of the stud;
Qa Q/A

Qt Q d 2

4 Ar
0
2
Design of co ld -fo rmed steel stud shear walls 65

s
Q Q0 2 
s'
where:
A: is the area of the gross cross-section;
d: is the depth of the section;
r0 : is the polar radius of gyration of the cross section
about its shear center;
s: is the fastener spacing (152mm s 305mm);
s = 305mm
Q0 : is a parameter provided by AISI (1996) as function of
the sheathing type;
2
x
E 1  0
r0
x0 : is the distance from centroid to shear center.

SHEAR FAILURE OF THE SHEATHING


To prevent shear failure of the sheathing, a value of nominal strength
(fbo=Nbo / A) shall be used in the following equations so that the shear strain of
the sheathing (J) does not exceed the permissible shear strain ( J ) that is
provided by AISI (1996) as function of the sheathing type:
S Ed
J C1  1 d J (2.33)
L 2
where:
x for singly symmetric C-sections:
f bC0
C1 (2.34)
V cr ,ip  f b  Qa

E1
>
f b V cr ,op  f b r02 E 0  x 0 D0  f b x 0 D0  x 0 E 0 @ (2.35)
V cr ,op  f b r V cr ,t ,Q  f b  f b x 0
0
2 2

x for doubly symmetric sections:


f bC0
C1 (2.36)
V cr ,ip  f b  Qa
66 Chapter II

E1 = 0 (2.37)

In the above formulas C0, E0, D0 are initial column imperfections which
shall be assumed to be at least:
C0 = L/350 in a direction parallel to the wall;
E0 = L/700 in a direction perpendicular to the wall;
D0 = L/(10000 d) rad as measure of initial twist of the stud from
the initial ideal, unbuckled shape.

Moreover, if fb>0.5fy, then in the definitions for Vcr,ip, Vcr,op and Vcr,t, the
Youngs modulus (E) and the shear modulus (G) shall be replaced by E and
G, respectively, as defined below:
4 fb f y  fb
E' E (2.38)
f ys
E'
G' G (2.39)
E

Shear strength of the wall associated to stud buckling strength (vF)


Assuming FF = Nb as the strength associated to stud buckling failure, the
shear strength per unit length of the wall (vF) associated to this failure mode is
obtained as follows:
FF
vF (2.40)
h
with h height of the wall.

RESULTS OF CURRENT STUDIES


Recent studies have been dedicated to study cold formed steel stud walls
sheathed with gypsum board panels (Miller & Pekz 1993, 1994; Lee &
Miller 2001a, b; Talue & Mahendran 20001; Schafer & Hiriyur 2002).
From these studies result that the diaphragm stiffness mechanical model
adopted by AISI (1996) may be inaccurate at least for the gypsum sheathings.
In particular, Lee & Miller (2001a) report that the axial strength is
independent of stud spacing, reflecting the localized nature of the wallboard
deformations rather than the shear diaphragm behavior assumed in the current
Design of co ld -fo rmed steel stud shear walls 67

AISI (1996) Specification. Moreover, using the differential equation of


equilibrium the Authors derive an approach for determining the axial strength
of a cold-formed steel stud wall sheathed with gypsum sheathings. In this
approach, it is assumed that the axial load is applied to the centroid of the
gross cross section for each stud and the bracing from the wallboard
connected by screws is represented by elastic springs.
Shafer & Hiriyur (2002) summarize the deficiencies of models currently
adopted by AISI (1996, 2001) and of the approach proposed by Lee & Miller
(2001a). In particular, Shafer & Hiriyur carried out numerical analyses of
sheathed wall systems using a plane stress FEM with imposed displacement,
and a FSM with a discretized sheathing modeled. The Authors concluded
from the numerical results that: (a) the diaphragm stiffness can be yet used
as basic mechanical model; (b) the diaphragm stiffness, per stud, is non-
uniform and is not solely depended from stud spacing (as assumed in AISI
1986) nor it is independent of stud spacing (as assumed in AISI 1996, 2001);
(c) the sheathing resists against the weak axis buckling of the stud by means
of shear stiffness, either locally of the material or globally for the diaphragm;
(d) the sheathing does not have much influence on local buckling of studs; (e)
in cases of highly dissimilar sheathing, such as in case of one-side sheathing,
distortional buckling of studs influences the behavior (moreover both codified
current (AISI 1996, 2001) and novel (Lee & Miller 2001a) approaches do not
take in-to account this aspect; (f) in cases of sheathing on both sides the axial
strength significantly increases in comparison with one-side sheathing even if
sheathings are dissimilar or relatively weak (gypsum board); (g) a simple
approach to evaluate the buckling capacity of unperforated studs with one-side
and/or dissimilar sheathing may be the FSM.
However, additional experimental and theoretical research is needed for
increasing the knowledge of the stud-to-sheathing interaction and to provide
design methodologies that can easily and correctly evaluate the axial strength
of cold-formed steel stud walls sheathed with panels.

2.2.3 Strength of frame-to-foundation connections (F-F)


The frame-to-foundation connections may by grouped in two types: tension
connections and shear connections.
68 Chapter II

Strength of frame-to-foundation tension connections (F(F-F)N)


Possible failure modes of tension connections are:
x failure in the hold-down (F(F-F)N,hd);
x shear failure in the hold-down-to-frame connection (F(F-F)N,hd-fr);
x tension failure in the hold-down-to-foundation connection
(F(F-F)N,hd-fo).

Therefore, the strength of the tension frame-to-foundation connection


(F(F-F)N) is obtained by the minimum strength associated to each failure mode:
F( F  F ) N min^F( F  F ) N ,hd ; F( F  F ) N ,hd  fr ; F( F  F ) N ,hd  fo ` (2.41)

FAILURE IN THE HOLD-DOWN


The strength corresponding to the failure in the hold-down (F(F-F)N,hd)
regularly is not performed because the hold-down is normally purposely
designed by the manufacturer to avoid its failure.

SHEAR FAILURE IN THE HOLD-DOWN-TO-FRAME CONNECTION


The strength corresponding to the shear failure in the hold-down-to-frame
connection (F(F-F)N,hd-fr) depends on the type of connection used which is
generally a bolted or screwed connection.
In case of screwed connections, F(F-F)N,hd-fr is obtained by the minimum of
the following values:
x the net section strength (Fn);
x the bearing/tilting strength (Fb/t);
x the screw shear strength (Fs).

Figure 2.9 shows the failure mode associated to net section, bearing/tilting
and screw shear strengths.

screw shear failure


net section failure bearing / tilting failure

Figure 2.9: Failure modes of screwed connections under shear loading.


Design of co ld -fo rmed steel stud shear walls 69

Because the hold down is thicker and stronger than the cold-formed steel
stud, it is possible to calculate the strength by the equations reported below.

The net section strength (Fn) is defined by the following equation:


Fn Anet f u ,s (2.42)
where:
Anet is the net area of the cold-formed steel member;
fu,s: is the tensile strength of the cold-formed steel member.

The bearing/tilting strength (Fb/t) is defined as follows:


Fb / t n s D b / t t f d f u ,s (2.43)
where:
ns: is the number of screws;
Db/t: is a coefficient defined as follows:
if thd = tf (tilting strength)
D b / t 3.2 t f / d 0.5 d 2.1 in ENV 1993-1-3 (1996)
4.2 t f / d
0.5
Db/t in AISI (1996)
if thd 2.5 tf (bearing strength)
D b/t = 2.1 in ENV 1993-1-3 (1996)
D b/t = 2.7 in AISI (1996)
if tf < thd < 2.5 tf
D b/t is obtained by linear interpolation
t f: is the thickness of the cold-formed steel member;
d: is the nominal diameter of the screw.

The manufacturer usually provides the shear resistance (Fs). However,


some indicative values of Fs are given in Table 2.2 (see Section 2.2.1). The
shear strength of the screws (Fs) must be about 1.2 greater than the
bearing/tilting strength (Fb/t) of the screws to avoid a brittle and sudden
fracture of a screw.

In case of bolted connections, F(F-F)N,hd-fr is obtained by the minimum of the


following values:
x the longitudinal shear strength of the sheet shear strength (Fl);
70 Chapter II

x the net section strength (Fn);


x the bearing strength (Fb);
x the bolt shear strength (Fv).

Figure 2.10 shows the above-introduced failure modes.


Because the hold down is thicker and stronger than the cold-formed steel
fame, it is possible to calculate the strength by the equations reported below.

If the bolted part has a reduced edge distance in the direction of the applied
force the connection can fails due to longitudinal shear collapse of the sheet.
For this failure type, the nominal strength (Fl) is defined as follows:
Fl e1 t f f u ,s (2.44)
where e1 is the end distance from the center of the bolt to the adjacent end of
the connected part, in the direction of load transfer (in ENV 1993-1-3 (1996)
and AISI (1996)), or the distance from the center of the bolt to the nearest
edge of an adjacent hole, in the direction of load transfer (in AISI (1996)
only).

longitudinal shear failure of the sheet shear


bearing failure

bolt shear failure


net section failure

Figure 2.10: Failure modes of bolted connections under shear loading.

For the net section strength (Fn) the ENV 1993-1-3 (1996) gives the
following formula:
d
Fn 1  3r Anet f u ,s d Anet f u ,s (2.45)
u
While, in case of channel sections having two or more bolts in the line of
the force, the AISI (1996) defines Fn as follows:
Design of co ld -fo rmed steel stud shear walls 71

x
Fn 1.0  0.36 L Anet f u ,s (2.46)

but Fn 0.5 Anet fu,s and Fn < 0.9 Anet fu,s


where:
r: is the ratio between the number of bolts at the cross
section and the total number of bolts in the connection;
u = 2e2 p2 with e2, the edge distance from the center of the bolt to
the adjacent edge of the connected part, in the
direction perpendicular to the load transfer and p2, the
spacing center-to-center of bolts in the direction
perpendicular to the direction of load transfer;
x: is the distance from the shear plane to the centroid of
the cross section;
L: is the length of the connection.

The bearing strength (Fb) is defined as follows:


Fb nb D be t f d f u ,s (2.47)
where:
nb: is the number of bolts;
Dbe: is a coefficient defined as follows:
Dbe = 2.5, but Dbe e1/1.2d in ENV 1993-1-3 (1996), where e1 is
the end distance from the center of the bolt to the adjacent end of
the connected part;
in AISI (1996) Dbe depends on the use of washers, thickness of
connected part (tf), type of joint, and fu,s/fy; In particular, in case of
tf < 4.76mm, for connections with washers under both bolt head
and nut Dbe ranges from 3.00 to 3.33, while for connections
without washers under both head and nut or with only one washer
Dbe ranges from 2.22 to 3.00.

In ENV 1993-1-3 (1996), the bolt shear strength (Fv) is defined through the
following relation:
Fv nb D bo As f u ,bo (2.48)
in which:
nb: is the number of bolts;
72 Chapter II

Dbo: is a coefficient that depends on the steel strength grade of the bolt:
Dbo = 0.6 for strength grade 4.6, 5.6 and 8.8;
Dbo = 0.5 for strength grade 4.8, 5.8, 6.8 and 10.9;
As: is the tensile stress area of the bolt;
fu,bo: is the ultimate tensile strength of the anchor steel.

For the bolt shear strength (Fv) the AISI (1996) gives the following
formula:
Fv nb Ab f v ,bo (2.49)
in which:
nb: is the number of bolts;
Ab: is the gross cross-sectional area of the bolt;
fv,bo: is the nominal unit shear strength of the bolt, provided by AISI
(1996) as function of steel strength grade and diameter of bolt.

TENSION FAILURE IN THE HOLD-DOWN-TO-FOUNDATION CONNECTION


The failure mode in the connections between hold-down and foundation
depends on the type of anchor, depth of embedment, concrete strength, edge
distances and spacing between anchors (Hilti 2001).
For both mechanical and adhesive-bonded anchors the main failure modes
under tension loading are concrete cone failure, pull-out (including any
expansion sleeve), tension steel breakage. These different failure modes are
shown in Figure 2.11.

pull-out failure steel breakage failure


concrete cone failure

Figure 2.11: Failure modes of anchors under tension loading (Hilti 2001).

The tension strength of the hold-down-to-foundation connection


(F(F-F)N,hd-fo) is equal to the minimum strength associated to each different
failure mode, as defined by the following relation:
Design of co ld -fo rmed steel stud shear walls 73

F( F  F ) N ,hd  fo min^N c ; N p ; N s ` (2.50)


where:
Nc: is the resistance against concrete cone failure;
Np: is the resistance against pull-out failure;
Ns: is the tensile resistance of steel.

The resistance against concrete cone failure (Nc) may be generally defined
by the following relation:
N c N c ,o rN ,s rN ,e (2.51)
in which:
Nc,o: is the basic value of resistance against concrete cone failure, that
depends on the type (AT) and diameter (da) of anchor, depth of
embedment (de) and concrete strength (fck):
Nc,o=f(AT, da, de, fck)
rN,s: is the anchor spacing reduction factor, that depends on type (AT)
and diameter (da) of anchor, depth of embedment (de) and anchor
spacing (s):
rN,s=f(AT, da, de, s)1
rN,e: is the edge distance reduction factor, that depends on type (AT)
and diameter (da) of anchor, depth of embedment (de) and anchor
edge distance (e):
rN,s=f(AT, da, de, e)1

The resistance against pull-out failure (Np) depends on the type of anchor
(AT), its diameter (da) and depth of embedment (de):
Np=f(AT, da, de) (2.52)
The following relation may generally be used to obtain the tensile
resistance against steel failure (Ns):
N s As f u ,a (2.53)
in which:
As: is the tensile stress area of the anchor;
fu,a: is the ultimate tensile strength of the anchor steel.
74 Chapter II

Strength of frame-to-foundation shear connections (F(F-F)V)


The failure modes of shear connections include:
x shear failure in the anchor-to-frame connection (F(F-F)V,a-fr);
x shear failure in the anchor-to-foundation connection (F(F-F)V,a-fo).

The strength of the shear frame-to-foundation connection (F(F-F)V) is equal


to the minimum shear strength related to the failure mode previously defined:
F( F  F )V min^F( F  F )V ,a  fr ; F( F  F )V ,a  fo ` (2.54)

FAILURE IN THE ANCHOR-TO-FRAME CONNECTION


The shear strength of the anchor-to-frame connection (F(F-F)V,a-fr) may by
calculate following the approach illustrated for the determination of the shear
failure in the hold-down-to-frame connection.

SHEAR FAILURE IN THE ANCHOR-TO-FOUNDATION CONNECTION.


As for the tension failure in the hold-down-to-foundation connection, also
the failure mode in the connections between anchor and foundation depends
on the type of anchor, depth of embedment, concrete strength, edge distances
and spacing between anchors (Hilti 2001).
For both mechanical and adhesive-bonded anchors the main failure modes
under shear loading are concrete edge failure and shear steel breakage, as
shown in Figure 2.12.

concrete edge breakout steel breakage failure

Figure 2.12: Failure modes of anchors under shear loading (Hilti 2001).

The shear strength of the anchor-to-foundation connection (F(F-F)V,a-fo) is


equal to the minimum shear strength associated to failure modes introduced,
as defined by the following relation:
F( F  F )V ,a  fo min^Vc ;Vs ` (2.55)
Design of co ld -fo rmed steel stud shear walls 75

where:
Vc: is the resistance against concrete edge failure;
Vs: is the shear resistance of steel.

The following relation may generally define the resistance against concrete
edge failure (Vc):
Vc Vc ,o rV ,se fV ,d (2.56)
in which:
Vc,o: is the basic value of resistance against concrete edge failure, that
depends on the type (AT) and diameter (da) of anchor, depth of
embedment (de) and concrete strength (fck):
Vc,o=f(AT, da, de, fck)
rV,se: is the anchor spacing and edge reduction factor, that depends on
type (AT) and diameter (da) of anchor, depth of embedment (de),
anchor spacing (s) and edge distance (e):
rN,s=f(AT, da, de, s, e)1
fV,d: is the loading direction factor, that depends on the direction of
shear force (E):
fV,d=f(E)

The shear resistance against of steel failure (Vs) may be generally obtained
by the following relation:
Vs D s As f u ,a (2.57)
in which:
Ds: is a coefficient that depends on the steel strength grade of the
anchor (D = 0.5 0.6);
As: is the tensile stress area of the anchor;
fu,a: is the ultimate tensile strength of the anchor steel.

Shear strength of wall associated to frame-to-foundation connections


strength (vF-F)
The shear strength per unit length associated to frame-to-foundation
connections strength (vF-F) is the minimum value between shear strength per
76 Chapter II

unit length associated to tension (v(F-F)N) and shear (v(F-F)V) connections


strength:
v F  F min^v( F  F ) N ; v( F  F )V ` (2.58)
in which (v(F-F)N) is obtained form strength of the frame-to-foundation tension
connections (F(F-F)N) as follows:
F( F  F ) N
v( F  F ) N (2.59)
h
with h height of the wall and (v(F-F)V) is obtained form strength of the frame-
to-foundation shear connections (F(F-F)V):
v( F  F )V n' F( F  F )V (2.60)
with n number of the end fasteners per unit length.

2.3 UNITS SHEAR DESIGN VALUES FOR CFSSSWs

The basic LFRSs used in CFSSSWs currently recognized by main building


Codes are: shear sheathing and diagonal bracings.
For the sheathed CFSSSWs the Codes include wall sheathed with wood-
based structural panels (plywood and oriented strand board (OSB)), sheet
steel, gypsum wallboard (GWB) and gypsum sheathing board (GSB). In this
case design values are provided in the Codes.
For diagonally braced CFSSSWs a rational set of design procedures with
specific force and drift limitations imposed for seismic event are presented in
building Codes.
The nominal shear design values given by the main building Codes used in
the United States, such as 1997 Uniform Building Code (UBC 1997) and 2000
International Building Code (IBC 2000), are illustrated and commented in
Section 2.3.1. Furthermore, the provisions reported by the 2000 International
Residential Code for One- and Two-Family Dwelling (IRC 2000), for the
application of the PSW method, are synthesized in the Section 2.3.2.

2.3.1 UBC and IBC design tables


Building Codes make a distinction between design for wind and seismic
loading. This distinction is based on the type of testing that has been used to
develop the design values for the shear strength. In fact, for all sheathed
Design of co ld -fo rmed steel stud shear walls 77

CFSSSWs currently permitted in the 1997 UBC and 2000 IBC (plywood,
OSB, sheet steel, GWB, GSB), tabulated design values are based exclusively
on physical testing which included monotonic loading for wind design values
and reversed cyclic loading for seismic design values.
Testing for development of Code design values occurred in two phases
(Serrette et al. 1996a, b and Serrette et al. 1997a, b). The results of the first
phase of testing were incorporated in the 1997 UBC and the results of the
second phase of testing (combined with the results of the first phase) were
incorporated in the 2000 IBC.
The typical test assembly in both phases of testing comprised 8ft.
(2440mm) high walls having a width of 2ft. (610mm), 4ft. (1220mm) or 8ft.
(2440mm) corresponding to aspect ratios (ratio of the wall height to width) of
4:1, 2:1 and 1:1. For the 1997 UBC (the first phase of testing), all walls were
4ft. (1220mm) wide except for gypsum sheathed walls that were 8ft.
(2440mm) wide. The 2ft. (610mm) wide walls were part of the second phase
of testing. For each test, the wall was anchored at both ends for overturning
and between the ends for shear transfer. In the shear wall tests, the sheathing
was oriented parallel to framing (all edges blocked), except for the gypsum
wallboard sheathing tests where the panels were installed perpendicular to
framing with strap blocking the abutting edges (see Chapter 3, Section 3.2.5
for details).
The typical results for the monotonic and reversed cyclic load tests, and the
interpretation of these test results, are illustrated in Figures 2.13 and 2.14,
respectively.
Using the measured wall response, as illustrated in Figures 2.13 and 2.14,
the Code tabulated nominal design strengths for the wall configurations tested
were determined as follows.

For wind design


The nominal wall strength was taken as the smallest of (see Fig. 2.13):
x 3.0 times the strength (averaging positive and negative values) at 0.5in.
(12.7mm) of lateral displacement.
x the maximum strength (averaging positive and negative values) of the
wall.
78 Chapter II

For seismic design


The nominal wall strength was taken as the smallest of (see Fig. 2.14):
x 2.5 times the second loop envelope strength (averaging positive and
negative values) at 0.5in. (12.7mm) of displacement
x the maximum second loop envelope strength (averaging positive and
negative values).

maximum load

load at 12.7mm lateral displacement

12.7mm lateral displacement

Figure 2.13: Monotonic shear load-lateral displacement response.

+ max. stable load + 2nd lood envelope

stable load at +12.7mm

-12.7mm

+12.7mm
stable load at -12.7mm

- 2nd lood envelope - max. stable load

Figure 2.14: Reversed-cyclic shear load-lateral displacement response.


Design of co ld -fo rmed steel stud shear walls 79

In both the 1997 UBC and 2000 IBC, design values are tabulated in terms
of a nominal capacity (Rn). Allowable Stress Design (ASD) and Load and
Resistance Factor Design (LRFD) capacities are computed as:
Rn / : for ASD (2.61)
I Rn for LRFD (2.62)
where : is a safety factor and I is a resistance factor.

The tabulated design values for seismic design under the 1997 UBC and
2000 IBC are reproduced in Tables 2.3 and 2.4, respectively. The tabulated
design values for wind action are reproduced in Tables 2.5 and 2.6 for design
under the 1997 UBC and 2000 IBC, respectively. The safety and resistance
factors for the UBC and IBC are summarized in Table 2.7.
Comparing the 2000 IBC tables with the 1997 UBC tables, it is evident that
a larger number of design options is permitted by IBC. This difference, as
indicated earlier, is the result of additional testing (second phase) that was
completed after publication of the 1997 UBC.
The seismic and wind design values in Tables 2.5 and 2.6 assume that
walls are fully sheathed and hold down anchors are provided at each end of
the wall as required by calculation. All panel edges are assumed to be blocked
for the shear wall applications. Some monotonic test data has shown that one
unblocked edge may reduce the fully-blocked capacity by about 50%, as
shown in Chapter 3 (Section 3.2.1).

No. 8 screws = 4.2 mm nominal diameter screws

Table 2.3: 1997 UBC nominal shear strength (Rn) values for seismic actions (lbs./foot).
80 Chapter II

No. 8 screws = 4.2 mm nominal diameter screws

Table 2.4: 2000 IBC nominal shear strength (Rn) values for seismic actions (lbs./foot).

No. 8 screws = 4.2 mm nominal diameter screws

Table 2.5a: 1997 UBC nominal shear strength (Rn) values for wind actions (lbs./foot).
Design of co ld -fo rmed steel stud shear walls 81

Table 2.5b: 1997 UBC nominal shear strength (Rn) values for wind actions (lbs./foot).

No. 8 screws = 4.2 mm nominal diameter screws

Table 2.6a: 2000 IBC nominal shear strength (Rn) values for wind actions (lbs./foot).
82 Chapter II

Table 2.6b: 2000 IBC nominal shear strength (Rn) values for wind actions (lbs./foot).

Table 2.7: Safety and resistance factors per 1997 UBC and 2000 IBC.

Following the Serrettes comments regarding the tabulated design values it


is possible to extrapolate of the Code design values in a rational manner that
meets the intent and overall safety implied by the Code.
Serrette specifies that some unpublished data show that for walls with the
same material on both sides (and identically attached) the shear values listed
in the tables may be doubled. Although these conclusions are encouraging,
additional testing is needed to verify performance. As a provisional measure,
the Author suggests that it may be adequate to use only 90% of the doubled
value with special attention given to the design of the chords and anchorage of
the wall.
Both the 1997 UBC and 2000 IBC Codes for gypsum board suggest that
sheathings should be applied perpendicular to framing with strap blocking
adjacent to panel edges and solid block in the end bays. This limitation is
based on the configuration used in the testing for development of the Code
Design of co ld -fo rmed steel stud shear walls 83

values. Serrette indicates that the behavior of a shear wall suggests that it
would be rational and safe to allow the use of the Code values for gypsum
board applied parallel to framing provided that all edges are, as required,
blocked.
For gypsum board shear walls, the Codes require the use of both flat strap
and solid end bay blocks at abutting panel edges (when the abutting edge does
not occur at a stud). Serrette reports that, because of the mechanism of load
transfer at the abutting panel edge, solid-blocked end bays are not structurally
necessary for the shear wall application. Moreover He notes that solid
blocking may be required to avoid global buckling of studs.

2.3.2 IRC provisions on the PSW method utilization


In addition to the design data provide in the UBC and IBC, some basic
research has been conducted on perforated CFSSSWs. Based on this research,
the 2000 IRC (IRC 2000) permits the seismic capacity of long walls with
openings to be computed on the basis of PSW method. The IRC provision is
limited to structures located in zones characterized by relatively low seismic
intensity. For applying the PSW method, the maximum aspect ratio of all full
height sheathed segments between openings must be 2:1, including the end
segments of the wall. Therefore, the maximum unrestrained opening height
must respect the limits indicated in IRC provision (see Table 2.8). The values
of the ratio of the strength of a shear wall with openings to the strength of a
fully sheathed shear wall without openings (F) reported by IRC are shown in
Table 2.9. According to the PSW approach, the value of the adjustment factor
(F) depends on sheathing area ratio (r).

1 feet = 305 mm

Table 2.8: Maximum unrestrained opening height (H) per 2000 IRC.
84 Chapter II

Table 2.9: Adjustment factors (F) for application of the PSW method per 2000 IRC.

REFERENCES

AISI (1986) AISI Specification for the design of Cold-Formed Steel structural
members. AISI (American Iron and Steel Institute). Washington DC.
AISI (1996) Cold-Formed Steel Design manual. AISI (American Iron and Steel
Institute). Washington DC.
AISI (2001) North American Specification for the design of Cold-Formed Steel
structural members (November 9, 2001 draft edition). AISI (American Iron and Steel
Institute), Washington DC.
Ambrose, J. & Dimitry, V. (1987) Design for Lateral Forces. John Wiley & Sons,
Inc. New York.
APA (1996) Use Panel Supplement. APA (The Engineered Wood Association).
Tacoma, WA, USA.
AS/NZS 4600 (1996) Cold-formed steel structures. AS/NZS (Australian
Standards/New Zealand Standards). Sydney.
Apparao, T.V.S., Errera, S.J., Fischer, G.P. (1969) Columns braced by girts and
diaphragm. Journal of structural Division. ASCE, Vol.95:965-990
Batista, E.M. (1988) Etude de la stabilit des profiles parois minces et sections
ouverte de type U et C. Ph.D. Thesis, University of Liege. Liege.
Beyer, D.E. (1993) Design of Wood Structures, Third Edition. McGraw-Hill, Inc.
New York.
CU-FSM (2003) http://www.ce.jhu.edu/bschafer/cufsm/index.htm. (Upgraded
August 2003)
Davies, J.M. & Leach, P. (1994a) First-order generalised beam theory. Journal of
construction steel research. Elsevier, Vol.31.
Design of co ld -fo rmed steel stud shear walls 85

Davies, J.M. & Leach, P. (1994b) Second-order generalised beam theory. Journal of
construction steel research. Elsevier, Vol.31.
Della Corte, G., De Martino, A., Fiorino, L., Landolfo, R. (2003) Numerical
modeling of thin-walled cold-formed steel C-sections in bending. In Proceedings of
the Advances in Structures Steel, Concrete, Composite and Aluminum (ASSCCA'03).
Sydney.
Dolan, J.D. & Heine, C.P. (1997a) Monotonic Tests of Wood-frame Shear Walls with
Various Opening and Base Restraint Conditions. Report TE-1997-001, Brooks Forest
Products Research Center, Virginia Polytechnic Institute and State University.
Blacksburg, VA, USA.
Dolan, J.D. & Heine, C.P. (1997b) Sequential Phased Displacement Cyclic Tests of
Wood-frame Shear Walls with Various Opening and Base Restraint Conditions.
Report TE-1997-002, Brooks Forest Products Research Center, Virginia Polytechnic
Institute and State University. Blacksburg, VA, USA.
Dolan, J.D. & Heine, C.P. (1997.) Sequential Phased Displacement Tests of Wood-
framed Shear Walls with Corners. Report TE-1997-003, Brooks Forest Products
Research Center, Virginia Polytechnic Institute and State University. Blacksburg,
VA, USA.
Dolan, J.D. & Johnson, A.C. (1997a) Monotonic Tests of Long Shear Walls with
Openings. Report TE-1996-001, Brooks Forest Products Research Center, Virginia
Polytechnic Institute and State University. Blacksburg, VA, USA.
Dolan, J.D. & Johnson, A.C. (1997b) Cyclic Tests of Long Shear Walls with
Openings. Report TE-1996-002, Brooks Forest Products Research Center, Virginia
Polytechnic Institute and State University. Blacksburg, VA, USA.
Easley, J.T., Foomani, M., Dodds, R.H. (1982) Formulas for wood shear walls.
Journal of structural Division. ASCE, Vol.105, No.11:2460-2478.
Errera, S.J. & Apparao, T.V.S.R. (1976) Design of I-shaped columns with diaphragm
bracing. Journal of structural Division. ASCE, Vol.102.
Errera, S.J., Pincus, G., Fischer, G.P. (1967) Columns and beams braced by
diaphragms. Journal of Structural Division. ASCE, Vol.93:295-318.
ENV 1993-1-3 (1996) Eurocode 3: Design of steel structures Part 1-3: General
rules - Supplementary rules for cold formed thin gauge members and sheeting. CEN
(European Committee for Standardization). Bruxelles.
Faherty, K.F. & Williamson, T.G. (1999) Wood Engineering and construction
handbook (Third edition). McGraw-Hill.
86 Chapter II

Fiorino, L. (2000) Il metodo direct strength per il progetto di profili cold-


formed in acciaio (Tesi di Laurea). Dipartimento di Analisi e Progettazione
Strutturale, Facolt di Ingegneria, Universit di Napoli Federico II. Napoli.
Fiorino, L. & Landolfo, R. (2001). Il metodo direct strength per la progettazione di
membrature in parete sottile formate a freddo in acciaio. In Proceedings of the XVIII
Congresso CTA (CTA 2001), Vol.2. Venezia.
Fiorino, L. & Landolfo, R. (2003) Il ruolo delle imperfezioni geometriche nella
calibrazione di modelli agli elementi finiti per profili sottili inflessi con sezione a C
formata a freddo. In Proceedings of the XIX Congresso CTA (CTA 2003). Genova.
Hancock, G,J. (1985) Distortional buckling of storage rack columns. Journal of
structural engineering. ASCE, Vol.111, No.12:2770-2783.
Hilti (2001) Hilti North America product technical guide.
Hoyle, R.J., Jr. & Woeste, F.E. (1986) Wood Technology in the Design of Structures,
Fifth Edition. Iowa State University Press. Ames, IA, USA.
IBC (2000) International Building Code: 2000. International Code Council, Inc. Falls
Church, VA, USA.
IRC (2000) International Residential Code for One- and Two-Family Dwelling:
2000. International Code Council, Inc. Falls Church, VA, USA.
Landolfo, R., Di Lorenzo, G, Fiorino, L. (2002) Attualit e prospettive dei sistemi
costruttivi cold-formed. Costruzioni metalliche No.1
Lau, S.C.W. & Hancock, G.J. (1987) Distortional buckling formulas for channel
columns. Journal of structural engineering. ASCE, Vol.113, No.5.
Lee, Y. & Miller, T.H. (2001a) Axial strength determination for gypsum-sheathed,
cold-formed steel wall stud composite panels. Journal of structural engineering.
ASCE, Vol.127, No.6:608-615.
Lee, Y. & Miller, T.H. (2001b) Limiting heights for gypsum-sheathed, cold-formed
steel wall studs. Practice periodical on structural design and construction. ASCE,
Vol.6, No.2:83-89.
Miller, T.H. & Pekz, T. (1993) Behavior of cold-formed steel wall stud assemblies.
Journal of structural engineering. ASCE, Vol.119, No.2:641-651.
Miller, T.H. & Pekz, T. (1994) Behavior of gypsum-sheathed cold-formed steel wall
studs. Journal of structural engineering. ASCE, Vol.120, No.5:1644-1650.
NAHB Research Center (1997) Monotonic Tests of cold-formed steel shear walls
with openings. NAHB (National Association of Home Builders). Upper Marlboro,
MD, USA.
Design of co ld -fo rmed steel stud shear walls 87

NAHB Research Center (1998) The Performance of Perforated Shear Walls with
Narrow Wall Segments, Reduced Base Restraint, and Alternative Framing Methods.
NAHB (National Association of Home Builders). Upper Marlboro, MD, USA.
Pincus, G. & Fischer, G.P. (1966) Behavior of diaphragm-braced columns and
beams. Journal of Structural Division. ASCE, Vol.92:323-350.
prEN 1998-1 (2001) Eurocode 8: Design of structures for earthquake resistance
Part 1: General rules Seismic actions and rules for buildings. CEN (European
Committee for Standardization). Bruxelles.
Schafer, B.W. (2001) Thin-walled column design considering local, distorsional and
Euler buckling. Proceedings of the Structural Stability Research Council, Annual
Technical Session and Meeting.
Schafer B.W. & Pekz T. (1998). Direct strength prediction of cold-formed steel
members using numerical elastic buckling solution. In Proceedings of the 2th
International conference on thin-walled structures. Singapore.
Schafer, B.W. & Pekz T. (1999) Local distortional buckling of cold-formed steel
members with edge stiffened flanges. In Proceedings of the 4th International
conference on steel and aluminium structures (ICSAS 99). Helsinki.
Schafer, B.W. & Hiriyur, B. (2002) Analysis of sheathed cold-formed steel wall
studs. In Proceedings of the 16th International Specialty Conference on Cold-formed
Steel Structures. St. Louis, MO, USA:501-513.
Serrette, R., Nguyen, H., Hall, G. (1996a) Shear wall values for light weight steel
framing. Report No. LGSRG-3-96, Light Gauge Steel Research Group, Department
of Civil Engineering, Santa Clara University. Santa Clara, CA, USA.
Serrette, R., Hall, G., Nguyen, H. (1996b) Dynamic performance of light gauge steel
framed shear walls. In Proceedings of the 13th International Specialty Conference on
Cold-formed Steel Structures. St. Louis, MO, USA:487-498.
Serrette, R.L., Encalada, J., Juadines, M., Nguyen, H. (1997a) Static racking behavior
of plywood, OSB, gypsum, and fiberboard walls with metal framing. Journal of
Structural Engineering. ASCE, Vol.123, No.8:1079-1086.
Serrette, R., Encalada, J., Matchen, B., Nguyen, H., Williams, A. (1997b) Additional
shear wall values for light weight steel framing. Report No. LGSRG-1-97, Light
Gauge Steel Research Group, Department of Civil Engineering, Santa Clara
University. Santa Clara, CA, USA.
Simaan, A. & Pekz, T. (1976) Diaphragm braced members and design of wall stud.
Journal of structural Division. ASCE, Vol.102.
88 Chapter II

SSTD 10-97 (1997) Standard for Hurricane Resistant Residential Construction.


Southern Building Code Congress International, Inc. Birmingham, AL, USA.
Sugiyama, H & Matsumoto, T. (1984) Empirical equations for the estimation of
racking strength of plywood-sheathed shear walls with openings. Summaries of
Technical Papers of Annual Meeting, Transactions of the Architectural Institute of
Japan, No.338.
Talue, Y. & Mahendran, M. (2000) Behaviour of cold-formed steel wall frames lined
with plasterboard. Journal of constructional steel research. Elsevier. Vol.57:435-452.
THIN-WALL (2003) http://www.civil.usyd.edu.au/case/thinwall.php. (Upgraded
August 2003)
UBC (1997) Uniform Building Code: Volume 2. International Conference of Building
Officials. Whittier, CA. USA.
Yu, W.W. (2000) Cold Formed Design (3rd Edition), John Wiley & Sons. New
York.
89

Chapter III
Testing of cold-formed steel stud shear
walls: review of existing literature

A large number of experimental research programs have been performed to


study the structural behavior of cold-formed steel stud shear walls (CFSSSW)
laterally braced with sheathings and/or with diagonal steel straps (McCreless
& Tarpy 1978, Tarpy & Hauenstein 1978, Tarpy 1980, Tarpy & Girard 1982,
Tissell 1993, Serrette 1994, Serrette & Ogunfunmi 1996, Serrette et al. 1996a,
b, 1997a, b, NAHB Research Center 1997, Gad et al. 1999a, b, Selenikovich
et al. 1999, COLA-UCI 2001, Dubina & Fulop 2002, Branston et al. 2003).
This Chapter contains a review of the main tests concerning the structural
behavior of CFSSSW carried out in Northern America, Europe and Australia.
The main goal of the review has been to investigate the influence of the basic
parameters on the lateral behavior of CFSSSW.
The Chapter is organized in two main parts. In the first part (Section 3.1) a
summary of the main existing test results is presented. The individuation of
basic factors that influence the lateral load response of CFSSSWs and the
analysis of their influence are illustrated in the second part (Section 3.2).

3.1 SUMMARY OF MAIN EXISTING TEST PROGRAMS

A summary of the main research programs regarding the testing of


CFSSSW carried out in Northern America, Europe and Australia is described
in this Section. A general listing of the tests presented in this Section is
90 Chapter III

reported in Table 3.1. A more complete listing of the test data is given in
Appendix A.

Author No of tests Bracing systems


McCreless & Tarpy (1978)NA 16M GWB
Tarpy & Hauenstein (1978)NA 18M GWB
NA
Tarpy (1980) 4M, 8C GWB, CP
Tarpy & Girard (1982)NA 14M GWB, PLY
NA
Tissell (1993) 8M OSB, PLY
Serrette (1994)NA 12M GWB, GSB, OSB, PLY, X-B
Serrette & Ogunfunmi (1996)NA 13M GSB, GWB, X-B
NA
Serrette et al. (1996a, b) 24M, 16C GWB, OSB, PLY
Serrette et al. (1997a)NA 33M GWB, FBW, OSB, PLY
Serrette et al. (1997b)NA 16M, 28C OSB, PLY, SSS, X-B
NAHB Research Center (1997)NA 4M OSB, GWB
A
Gad et al. (1999a, b) X-B
Selenikovich et al. (1999)NA 6M, 10C OSB, GWB
COLA-UCI (2001)NA 18C OSB, PLY
E
Dubina & Fulop (2002) 7M , 9 C SCS, GWB, OSB
Branston et al. (2003)NA 6M, 6C OSB, PLY
NA
: North American tests; : Australian tests; E: European tests
A
M
: monotonic tests; C: cyclic tests
CP: cement plaster; FBW: FiberBond wallboard; GSB: gypsum sheathing board; GWB: gypsum
wallboard; OSB: oriented strand board; PLY: plywood; SCS: steel corrugated sheet; SSS: steel
sheet sheathing; X-B: steel flat strap X-bracing

Table 3.1: Listing of tests of CFSSSW.

3.1.1 McCreless & Tarpy (1978)


This experimental research consisted of the testing of sixteen shear walls
made of steel studs and gypsum wallboard (GWB) with various aspect ratios
under monotonic loads. The tests were carried out in accordance with ASTM
E 564-76 (1976). The wall construction used was representative of the type of
construction usually used for internal wall partitions.

The objectives of the testing program were:


To determine the variation of the shear strength when the aspect ratio
(height/width) changing from 0.33 (2440/7320mm) to 2 (3660/3660).
To establish the allowable degree of panel distortion before major wall
panel damage.
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 91

To determine if the addition of a single horizontal stiffener located at


mid-height in the plane of the wall could improve the shear behavior.

The wall sizes of specimens were 3660x3660mm, 3660x4880mm,


3660x7320mm, 3050x3660mm, 3050x4880mm, 3050x7320mm,
2440x2440mm, 2440x3660mm, 2440x4880mm and 2440x7320mm (height x
length) with aspect ratios that ranges from 0.33 to 1.00. Each wall was
assembled with 89x0.84mm (web depth x thickness) C (lipped)-sections studs
spaced at 610mm on center. Double back-to-back coupled studs were used at
ends of each wall. The studs were attached to 92x38x0.84mm (web depth x
flange size x thickness) U (unlipped)-sections track with 4.8x13mm (diameter
x length) low profile head screws. GWB, 12.7mm thick, was attached to both
sides of the stud assembly with 3.5x25mm bugle head screws spaced at
305mm at the perimeter and in the field of the panel. Clip angles were used to
anchor the specimens to the load frame.

The Authors observed that for shorter walls bending deformation


dominated. In this case, the bottom track deformed around the exterior corner
tension anchorage point and then the screws in tension corner rotated through
the GWB, followed by cracking separation of the wallboard. The final failure
was due to excessive rotation. For longer walls, where the shear deformation
controlled the behavior, the edge screws rotated through the panel and the
final failure was produced by the stud framing shearing through the GWB
along the top.

McCreless & Tarpy concluded from the test results that:


For the aspect ratio varying between 0.33 and 1.00 the shear strength
was constant but the shear stiffness decreased;
The first noticeable wallboard damage occurred at about 6 to 13mm
total displacement (for the shorter walls) and at about 6mm (for the
longer walls). The real damage occurred at about 13 to 19mm total
displacement (for the shorter walls) and at about 6 to 13mm (for the
longer walls).
92 Chapter III

3.1.2 Tarpy & Hauenstein (1978)


The testing program performed by Tarpy & Hauenstein included eighteen
full-scale walls made of steel studs and GWB with seven different types of
wall panel construction and anchorage details.

The main objectives of the experimental research were:


To determine the effect of different construction and anchorage details
on shear resistance of stud shear walls.
To evaluate the damage thresholds load levels.
To provide a comparison between the performances of wood-framed
and steel-framed shear walls.

One wall type consisted of 51x102mm (width x depth) wood-studs, while


others were constructed using 89x38x0.81mm (web depth x flange size x
thickness) C-sections studs spaced at 610mm on center. GWB, 12.7mm thick,
placed in horizontal position was attached to both sides of the specimens. The
fasteners to connect the panels to the steel frame were 3.5x25mm (diameter x
length) bugle head screws whereas to connect the panels to the wood frame
35mm annular head nails were used.

The Authors observed from the test results that:


For avoiding uplift failure an adequate attachment should be adopted
to connect the track and floor framing systems.
The reduction of the fasteners spacing around the wall perimeter could
provide higher shear strength.
For design purposes, a safety factor of 2.0 was recommended for
design purposes to ensure that the design load level does not exceed
the damage threshold level.

3.1.3 Tarpy (1980)


Nine different types of GWB and cement plaster (CP) attached to light
gauge steel stud were tested under monotonic and cyclic loading protocol by
Tarpy. The tests were carried out in accordance with ASTM E 564-76 (1976)
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 93

The main test objectives were:


To investigate the behavior of the different sheathing types.
To study the effect of GWB fasteners spacing.
To determine the effect of cyclic loading versus monotonic loading on
the shear behavior.
To analyze the behavior of different construction techniques and
anchorage details.
To establish the thresholds for damage of the walls due to lateral
displacement.
To study the contribution to shear capacity of a 45 stud placed at the
bottom corner between the chord members and the adjacent stud.

The wall sizes of specimens were 2440x2440mm or 2440x3660mm (height


x length). The frame was made using 89x38x13x0.84mm (web depth x flange
size x lip size x thickness) C (lipped)-sections studs at 610mm on center. The
studs, with a mean yield strength of 331MPa, were attached to 92x38x0.84mm
U-sections track with 4.8x13mm (diameter x length) low profile head screws.
GWB, 12.7mm thick and 22.2mm thick CP placed in vertical position was
connected to the frame using different fastener type and spacing. In particular,
the following fastener type and spacing were used: 3.5x25mm bugle head
screws spaced at 305mm; 3.5x25mm bugle head screws spaced at 610mm;
3.5x25mm bugle head screws spaced at 152/305mm (perimeter/field);
3.5x13mm pan washer head screws spaced at 197mm; 3.5x10mm bugle head
screws spaced at 305mm.

Tarpy concluded from the test results that:


The use of CP increased the shear strength and the stiffness.
The use of two layer of GWB increased the shear capacity, while
decreasing the shear stiffness, in comparison with single layer.
The reduction of the fastener spacing increased the shear strength and
stiffness.
Cyclic load decreases shear strength and damage threshold level.
The corner anchorage influenced the shear behavior dramatically.
Hold-down exhibited higher shear strength than bolt and washer
94 Chapter III

anchors. Densely spaced powder actuated fasteners (connected to a


supporting concrete beam) provided similar restraint to the hold-down.
The shear resistance did not vary extensively when using different
types of interior shear anchorage.
The use of a 45 stud placed at the bottom corner between the chord
members and the adjacent stud had little effect on the shear capacity.

3.1.4 Tarpy & Girard (1982)


The study of Tarpy & Girard was performed in response to a need to
develop design criteria for steel stud sear wall panels with different
construction details and sheathing materials. The materials used for the
sheathings were GWB, gypsum sheathing board (GSB) and plywood (PLY).
The experimental program was based on the testing of fourteen specimens
under monotonic load following the requirements of ASTM E 564-76 (1976).

The objectives of the research were:


To determine the effect of different construction techniques and
anchorage details on the shear resistance of stud shear walls with
different types of sheathing. In particular, the parameters examined
were:
The effect of using light gage clip angles and powder-actuated
fasteners in place of bolts and washers to anchor the base of the
wall and the effect of anchoring the wall through transverse floor
joists.
The effect of PLY or gypsum exterior sheathing in place of GWB.
The effect of using fillet welds instead of self drilling screws to
attach the studs to the tracks.
The effect of using a 406mm rather a 610mm stud spacing.
To establish the thresholds for damage of the walls due to lateral
displacement.

All wall sizes were 2440x2440mm (height x length), except for one
specimen that had a wall size of 2440x3660mm. The stud wall systems were
constructed using 89x25x13x0.84mm C-sections studs attached to
92x38x0.84mm U-sections track. Double back-to-back coupled studs were
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 95

employed at the ends of each wall. GWB, 12.7mm thick, was attached to both
sides of the specimens, except for two specimens, that had GWB as internal
panel and PLY (12.7mm thick) or GSB (12.7mm thick), respectively, as
exterior panels. The sheathing orientation was always horizontal. All panels
were connected to the frame with 3.5x25mm (diameter x length) bugle head
screws spaced at 305mm at the perimeter and in the field of the panel. Clip
angles fixed with bolts or powder-actuated fasteners were used to connect the
base of some walls.

Tarpy & Girard noted that all wall types had the same basic failure mode.
The bottom track deformed around the anchorage device at the uplift corner of
the wall and the cracking of the GWB happened at the same locations from
the corner fasteners to the edge of the panel.

They concluded that:


When the bolt and washer anchorage details were used without clip
angles the shear capacity decreased. Moreover, the use of closely
spaced powder actuated fasteners negligibly increased the shear
behavior in comparison to using corner clips. When the wall was
anchored through the floor joist it showed lower shear capacity than
when it was connected directly to the test frame. Therefore, it was
suggested a rigid attachment to connect the wall panel to the floor or
roof framing systems and it was recommended the use of clip angles.
The use of PLY increased the shear strength whereas the use of GSB
decreased it in comparison with the employ of GWB.
The use of welded stud to track connections provided the same shear
strength as screw connections.
The reduction of the stud spacing did not increase significantly the
shear strength and stiffness.
For design purposes, a safety factor of 2.0 was recommended to
determine the design shear strength from the ultimate shear strength
for the type of stud shear walls examined.
96 Chapter III

3.1.5 Tissell (1993)


Tissell conducted for the American Plywood Association eight monotonic
loading tests on walls that were sheathed with either oriented strand board
(OSB) or PLY and that had various frame thicknesses.

The tests were carried out to achieve the principal following goals:
To examine the effect of fasteners size and spacing.
To study the influence of frame thickness.

All wall sizes were 2440x2440mm (height x length). The frame was made
using 64x41x1.88mm (web depth x flange size x thickness) C-sections studs
at 610mm on center attached to 64x41x1.88mm U-sections track for the
specimens 1, 7 and 8. The frame used for other specimens were made using
89x41x1.50 studs at 610mm on center attached to 89x41x1.50 track
(specimens 2 and 3) and 89x41x1.19 studs at 610mm on center attached to
89x41x1.19 track (specimens 4, 5 and 6).
Fasteners types were 4.2mm diameter screws for the 1.19mm frame
thickness and 4.8mm diameter screws or 3.7mm diameter pins for the 1.50
and 1.88mm frame thickness. Different fasteners schedule was used. In fact,
the edge fasteners spacing varied form 76 to 152mm while the field fasteners
spacing was 305mm.
The different sheathing types for tests included 9.5mm thick PLY
(specimens 1, 2, 4 and 5), 15.9mm thick PLY (specimen 8), 11.1mm thick
OSB (specimens 3 and 6) and 15.1mm thick OSB (specimen 7). The sheathing
orientation was vertical in each case.

The Author reported that in most cases failure occurred due to buckling of
the single end studs or the bottom track at the anchor bolts. These premature
collapses prevented the full development of the shear capacity of the panels.
For this reason the test results did not offer a correct valuation of the behavior
of the sheathing panels.
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 97

3.1.6 Serrette (1994)


This experimental program consisted of the monotonic load testing of
twelve CFSSSW. The walls were laterally braced with 51x0.84mm (width x
thickness) X-bracing steel flat strap (X-B) on one side only (specimen 1),
12.7mm thick GWB on both sides (specimen 2), 12.7mm thick GWB and
12.7mm thick GSB in combination with X-B (specimen 3), 11.9mm thick
PLY (specimens from 4 to 8) or 11.1mm thick OSB (specimens from 9 to 12)
on one side only. All sheathings were vertically oriented, except for specimens
6 and 7 that had sheathings horizontally oriented.
The 2440mm by 2440mm steel frame was identical for all walls. The frame
was made using 152x41x0.84mm (web depth x flange size x thickness) C-
sections studs at 610mm on center attached to 152x32x0.84mm U-sections
track. Double back-to-back coupled studs were used at ends of walls.
The panels to frame connections were 3.5mm diameter screws for
specimens 2, 3, 4, and 12; 4.2mm diameter screws for the specimens 6, 7, 8
and 11; and 2.9mm diameter pins for the specimens 5 and 9. All fasteners
were spaced at 152mm at the perimeter and at 305mm in the field.

From the tests results, it is possible to conclude that:


The PLY panels carried slightly higher loads in comparison with OSB
panels.
The use of 2.9mm diameter pins decreased the maximum shear
strength in comparison with 3.5 and 4.2mm diameter screws.
The walls with panels oriented vertically and with blocked panels
oriented horizontally provided essentially the same shear strength.
When blocking is omitted for the walls with horizontal panels, the
shear capacity of the wall was reduced by more than 50%.

3.1.7 Serrette & Ogunfunmi (1996)


A series of monotonic loading tests were conducted by Serrette &
Ogunfunmi on 13 walls. These tests included three different shear resisting
systems: framed walls with X-B (type A); framed walls with GSB on the face
and GWB on the back (type B); and framed walls with GSB and X-B on the
face and GWB on the back (type C).
98 Chapter III

The main objectives of the investigation were:


To study the contribution of X-B, GSB and GWB, and the
combination of X-B, GSB, GWB to the in-plane shear resistance of
steel stud walls.

The 2440mm by 2440mm steel frame was identical for all walls. The frame
was made using 152x32x0.84mm (web depth x flange size x thickness) C-
sections studs at 610mm on center attached to 152x0.84mm (web depth x
thickness) U-sections track. At the ends of each wall, double back-to-back
coupled studs were used to avoid chord buckling. The fasteners used for the
frame connections were 4.2x13mm (diameter x length) wafer-head screws.
Clip angles were used to connect each wall specimen to the base of the test
frame.
The different shear resisting systems included:
x Type A: framed walls with 51x0.84mm (width x thickness) X-B on the
face. The straps were attached to the frame using gusset connections
that were designed for the yield strength of the strap.
x Type B: framed walls with 12.7mm thick GSB on the face and
12.7mm thick GWB on the back. The gypsum panels were orientated
vertically and were attached to the frame using 3.5x25mm bugle head
screws spaced at 152mm at the perimeter and at 305mm in the field.
x Type C: framed walls with both Type A and B shear resisting systems.
All specimens were raised with GSB and X-B on the face and GWB
on the back except for one specimen that was built with X-B on both
sides.

The Authors reported that for type A walls, failure resulted from excessive
lateral deflections that followed the yielding of the tension X-B. For type B
and C walls, at approximately half of the maximum load, screw rotation
occurred at the perimeter edge, and at the maximum load, the paper along the
edges of the panels broke at the locations of the screws. Type B walls
provided about 2.1 times the maximum load of type A walls, type C walls
with one X-B on one side increased the maximum load by approximately 1.3
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 99

times over type B walls, and type C wall with X-B on both sides amplified the
maximum load by about 1.8 times in comparison with type B walls.

Serrette & Ogunfunmi concluded that:


The use of X-B plus gypsum board reduced the permanent deflection
of the wall and increased the shear strength without decreasing the
stiffness.
The use of X-B plus gypsum board is not practical due to the need to
pretension the straps and the need for additional screws to connect the
straps.

3.1.8 Serrette et al. (1996a,b)


Serrette conducted twenty-four monotonic loading tests and sixteen cyclic
loading tests on walls with various sheathing types. All specimens were
sheathed with OSB or PLY on a side, except for 10 specimens of the
monotonic tests that were sheathed with OSB on a side and GWB on the other
side or with GWB on both sides. The tests were planned to generate design
data for specific wall constructions and also answer certain fundamental
design questions.

The monotonic test program was divided in two phases with different
objectives.
The first phase included walls sheathing on one side only and was
addressed to the following issues:
To compare the differences in the behavior of OSB and PLY.
To examine the effect on the shear strength of a variation of the aspect
ratio (height/width) from 1 (2440/2440) to 2 (2440/1220).
To examine the effect of dense fasteners schedules.
The second phase consisted of walls sheathing on both sides and was
addressed to the following questions:
To study the behavior of walls with OSB on a side and GWB on the
other.
To study the behavior of walls with GWB on both sides.

The objectives of cyclic test program were the followings:


100 Chapter III

To determine the relative strength of walls with OSB and with PLY.
To study the effect of dense fasteners schedules.
To determine the relative strength of walls in cyclic and in monotonic
tests.

The wall sizes of specimens were 2440x2440mm or 2440x1220mm (height


x length) for monotonic tests and 2440x1220mm for cyclic tests. The frame
was made using 89x41x10x0.84mm C-sections studs at 610mm on center
attached to 89x32x0.84mm U-sections track. Double back-to-back coupled
studs were employed at ends of each wall to avoid studs buckling. Hold-
downs (tie-downs) were used to connect each wall specimen to the base of the
test frame.

The different sheathing types, sheathing orientations and screws spacing


for monotonic tests included:
x 2440x2440mm framed walls with 15.9mm thick PLY vertically
orientated with 152/305mm (perimeter/field) screw spacing.
x 2440x2440mm or 2440x1220mm framed walls with 11.1mm thick
orientated strand board vertically or horizontally orientated with
fasteners spacing varying between 152/305mm and 51/305mm.
x 2440x1220mm framed walls with 11.1mm thick orientated strand
board on one side with fasteners spacing varying between 152/305mm
and 51/305mm and 12.7mm thick gypsum wallboard on the other side
with 178/178mm screw spacing, both external and internal sheathings
were vertically orientated.
x 2440x2440mm framed walls with 12.7mm thick GWB on both side
with fasteners spacing varying between 178/178mm and 102/102mm,
both external and internal sheathings were horizontally orientated.
For cyclic tests the sheathings used were 11.1mm thick orientated strand
board or 15.9mm thick PLY. The panels were always vertically orientated and
the screws spacing varying between 152/305mm and 51/305mm.
For both monotonic and cyclic tests, the fasteners used were: 4.2x13mm
(diameter x length) wafer-head screws for the frame connections; 4.2x25mm
flat head screws for the PLY and OSB panel and 4.2x32mm bugle head
screws for gypsum wallboard.
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 101

From the monotonic tests results, the Authors concluded that:


The overall behavior of the PLY and OSB panel assemblies was
practically identical. In particular the PLY walls carried slightly higher
loads.
The sheathings oriented horizontally exhibited slightly higher shear
strength than the panels oriented vertically.
The shear strength was practically the same for the aspect ratio varying
between 1 and 2.
The reduction of the fastener spacing increased significantly the shear
strength.
The walls with OSB panels on one side and GWB on the other side
exhibited similar failure behavior, but degraded more gradually than
the walls with OSB on one side alone in terms of shear capacity.
The walls with GWB panels on both sides had much lower shear
strength than walls with OSB sheathing.

Serrette et al. concluded form the results of the cyclic tests that:
PLY and OSB panel assemblies sowed a little difference in the cyclic
shear strength.
As in the monotonic tests, the wall shear strength increased
significantly with the reduction of the fastener spacing.
The shear behavior of walls observed in cyclic tests was somewhat
lower than the shear behavior exhibited in monotonic tests for walls of
similar construction details.

3.1.9 Serrette et al. (1997a)


In this paper, the Authors presented the results of thirty-three full-scale
monotonic racking tests and twenty small-scale lateral shear tests on PLY,
OSB, GWB and FiberBond wallboard (FBW) attached to light gauge steel
stud.

The main test objectives were:


To investigate the behavior of the different sheathings material.
To examine the effect of fasteners type and size.
102 Chapter III

To examine the effect of fasteners spacing for gypsum and FBW.


To compare the behavior of the panels oriented vertically and panels
oriented horizontally with steel flat strap and solid blocking attached
across the mid-height.
To compare the full-scale and small-scale test results.

Two different frames were used in the full-scale tests: type A and type B.
The type A frame was 2440x2440mm (height x length) in size constructed
with ASTM A446 Grade A galvanized steel members. The studs were
152x41x10x0.84 C-sections spaced at 610mm on center and the tracks were
152x25x0.84mm U-sections. The chord studs were two back-to-back coupled
profiles. The type B frame was identical to type A except that 51x0.84mm
(width x thickness) steel flat strap was attached across the mid-height of the
wall and that solid blocking was installed above the strap in the and bays. The
type A and B frames were sheathed with: 11.9mm thick PLY; 11.1mm thick
OSB; 12.7mm thick FBW; 12.7mm thick GWB. The panels were placed
either one side or both sides and they were oriented either vertically or
horizontally depending on the test. The frame fasteners were 4.2x13mm
(diameter x length) wafer-head screws. Panels were attached to the frame
using either 3.5x25mm bugle-head screws, 4.2x32mm flat-head screws with
countersinking nibs, or 3.7mm diameter steel pins depending on the test.
Hold-downs were used to connect each wall specimen to the base of test
frame.

The small-scale specimens consisting of single 610x610mm panels


attached to opposite flanges of the studs that were oriented horizontally with a
single stud on top and double studs placed back-to-back on the bottom. The
specimens were sheathed with panels including 11.9mm thick PLY; 11.1mm
thick OSB; 12.7mm thick FBW; 12.7mm thick GWB. The panels were
connected with three 3.5x25mm bugle-head screws (152mm on center
spacing) to the top stud and with two rows of six screws (102mm on center
spacing) to the bottom studs.

The full-scale tests failed initially by fasteners rotation (tilting) about the
plane of the stud flange. In some tests the chord studs were subjected to local
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 103

crushing at the bearing end. Otherwise, generally the specimens failed either
when the edges of the panels broke off at the screw fastener or when the panel
pulled over the head of the screws. In some cases where 3.5mm diameter
screws were used in PLY and OSB panels, fracture of the screws occurred.

The Authors reported for the full-scale tests the following conclusions:
The behavior of the PLY and OSB panels was comparable, whereas
the strength of gypsum and FBW walls was relatively low. The use of
GWB on the interior of the wall and PLY on the exterior produced a
higher shear capacity, by approximately 18%, in comparison with the
PLY wall.
The use of 3.7mm diameter nails decreased the maximum shear
strength in comparison with 3.5 and 4.2mm diameters screws. The
maximum shear strength was not influenced by the size of the fastener.
The failure mode for the specimens with 3.5mm diameter screws was
fracture of fasteners and thus walls with 3.5mm diameter screws may
fail at lower loads than walls with 4.2mm diameter screws.
For gypsum and FBW the reduction of the fastener spacing increased
significantly the shear strength.
The walls with blocked panels oriented horizontally provided
essentially the same shear capacity but higher stiffness than a
comparable wall with panels oriented vertically. When blocking was
omitted from the walls with horizontal panels, the shear capacity of the
wall was reduced by more than 50%.

For the small-scale tests the conclusions of the Authors were:


Both the strength and stiffness of the PLY specimens were more than
that of the OSB specimens. The strength of gypsum and FBW walls
was less than that of PLY and OSB specimens.
The normalized shear strength for small-scale tests was similar as
those for full-scale tests, thus the small-scale tests were considered to
be useful in an evaluation of the relative resistance of different wall
assemblies.
104 Chapter III

3.1.10 Serrette et al. (1997b)


The Authors illustrated in this work the results of sixteen monotonic and
twenty eight cyclic shear tests conducted on walls assembled with PLY, OSB,
steel sheet sheathing (SSS) and X-B attached to light gauge steel frame on
only one side.

The objectives of the test program were the followings:


To obtain design data for walls with high aspect ratios (varying from 2
to 4).
To define limits for framing thickness for sheathing attached with
4.2mm diameter screws.
To study the behavior of different types of sheathings (PLY, OSB,
SSS).
To evaluated the performance of X-B.
To examine the effect of various fastener schedules.

Both monotonic and cyclic tests were conducted on walls with


89x43x13mm (web depth x flange size x lip size) C-sections studs at 610mm
on center and with 89x32 (web depth x flange size) U-sections track. Double
studs (back-to-back) were used at the ends of the walls. All PLY and OSB
panels were orientated vertically. The fastener types used were 4.2x13mm
modified truss head screws for the frame connections, 4.2x25mm (diameter x
length) flat head screws for the PLY and OSB panels, and 4.2x13mm
modified truss head screws for the X-B and SSS. Hold-downs were used to
connect each wall specimen to the base of the test frame.

For monotonic tests, the studs were 0.84mm thick except for one case
where 1.09mm thick studs were used (tests 3-4). The following assemblies
were tested:
x 2440x1220mm walls with 114x0.84mm (width x thickness) and
191x0.84mm X-B (tests 1-4).
x 2440x610mm walls with 11.1mm thick OSB attached to the frame
using screws spaced at 305mm in the field and at 152, 102 and 51mm
at the perimeter (tests 5-10).
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 105

x 2440x610mm walls with 0.46 and 0.69mm thick SSS attached to the
frame using screws spaced at 305mm in the field and at 152 and 102 at
the perimeter (tests 11-14).
x 2440x1220mm walls with 0.46mm thick SSS attached to the frame
using screws spaced at 305mm in the field and at 152 at the perimeter
(tests 15-16).

For the cyclic tests, the studs were 0.84mm thick except for some cases. In
fact, for A1-A8 tests the end studs were 1.09mm thick, while for B1-B2 and
B3-B4 tests all studs were 1.09 and 1.37mm thick respectively. The different
shear resisting systems included:
x 2440x1220mm walls with 11.9mm thick PLY and 11.1mm thick OSB
attached to the frame using screws spaced at 305mm in the field and at
76 and 51mm at the perimeter (tests A1-A8).
x 2440x1220mm walls with 11.9mm thick PLY attached to the frame
using screws spaced at 305mm in the field and at 152mm at the
perimeter (tests B1-B4).
x 2440x1220mm walls with 114x0.84mm (width x thickness) and
191x0.84mm X-B (tests C1-C4).
x 2440x1220mm walls with 0.46mm thick SSS attached to the frame
using screws spaced at 305mm in the field and at 152 at the perimeter
(tests D1-D2).
x 2440x1220mm walls with 0.46mm thick SSS attached to the frame
using screws spaced at 305mm in the field and at 152mm at the
perimeter (tests D1-D2).
x 2440x610mm walls with 11.1mm thick OSB attached to the frame
using spaced at 305mm in the field and at 152, 102 and 51mm at the
perimeter (tests E1-E6).
x 2440x610mm walls with 0.69mm thick SSS attached to the frame
using screws spaced at 305mm in the field and at 102 and 51mm at the
perimeter (tests F1-F4).

For the monotonic tests, Serette et al. reported that the walls with
114x0.84mm X-B failed due to local buckling of the end studs, differently, the
106 Chapter III

walls with 191x0.84mm X-B failed due to local buckling in the top track and
end studs, aggravated by bending due to the eccentricity of the strap force.
For the walls with OSB panels with a 51mm edge screw spacing, the end
studs buckled just above the track. With a 102mm edge screw spacing
displacements became excessive. With a 152mm edge screw spacing failure
was initiated by buckling of the end stud at the hold-down.
Failure of walls with SSS resulted from rupture of the steel sheet along the
line of the screws at the edges. Diagonal tension field patterns were not
observed.

In the case of cyclic tests, for orientated strand board and PLY sheathings
with aspect ratio of 2, failure was initiated by screw heads pulling through the
sheathing or screw pulling out of the framing in 1.09mm thick framing (tests
A1-A8 and B1, B2), but in 1.37mm thick framing (tests B3, B4) some screws
failed in shear.
Failure modes in walls with X-B (C1-C4) were similar to those observed in
monotonic tests of similar specimens.
Analogously, for orientated strand board sheathings with an aspect ratio of
4 (tests E1-E6) failure modes were similar to those observed in monotonic
tests of similar specimens.
Finally, failure modes for walls with SSS (D1, D2, F1-F4) resulted from a
combination of screws pulling out of the framing, rupture of the steel sheet
along the line of screws at the edges, and in some cases local buckling of the
end studs.

From the tests results, the Authors concluded that:


The shear strength appreciably decreased when the aspect ratio
increased from 2 to 4.
The use of thicker and back-to-back coupled end studs for the walls
with PLY and OSB panels allowed to fully develop the shear strength
of the panels-to-frame connections also in the cases in which dense
fasteners schedules were used. 4.2mm screws should be limited to
1.09mm thick framing. In fact, this screw behaved well in the 0.84 and
1.09mm thick frames (screws pull-out or pull-through failure) but
fractured in shear when 1.37mm thick frames were used.
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 107

The walls sheathed with SSS had a ductile behavior without sudden
decreases in shear load capacity; moreover, the use of thick sheathings
increased the shear resistance, but the failure mode moved from
rupture at the edges of the sheathing to screw pull-out from the
framing.
In the design of walls with X-B, the designer must consider that the
force in the strap may be larger than that corresponding to the nominal
yield strength; also, if X-Bs are installed on one side of the wall only,
the effect of eccentricity should be considered.
Decreasing the screws spacing result in the increased maximum shear
load.

3.1.11 NAHB Research Center (1997)


In this paper, the National Association of Home Builders (NAHB)
Research Center presented the results of four monotonic load tests of
12190mm long, cold-formed steel (CFS) framed shear walls with openings.

The goals of the experimental program were the followings:


To study the capacity of steel-framed shear walls as influenced by the
presence of openings and, in particular, to investigate the suitability of
using the Perforated Shear Wall (PSW) design method, originally
developed for wood structures (Sugiyama & Matsumoto, 1993), also in
case of light-gauge steel-framed shear walls.
To briefly investigate the effects of reduced anchoring constraints in
view of future research to account for restraint provided by corners
without including hold-down brackets.
To provide a direct comparison between the performances of wood-
framed and steel-framed shear walls.

All shear walls specimens were 2440x12190mm (height x length) in size.


The frames were constructed of 89x38x0.84mm (web depth x flange size x
thickness) C-sections studs spaced at 610mm on center. Exterior sheathings
consisted of 11.1mm thick OSB oriented vertically and fastened to the frames
with 4.2mm diameter screws spaced 152mm along the perimeter and 305mm
in the field of the panels. Interior sheathings were 12.7mm thick GWB
108 Chapter III

oriented vertically and attached to the frames with 3.5mm diameter screws
spaced 178mm along the perimeter and 254mm in the field.
The wall 1 was fully sheathed, while the walls 2A and 2B had one
2032x1219mm (height x length) door and one 1727x2400mm window.
Besides the same openings present in the wall 2A and 2B, the wall 4 had one
2032x3658mm door. The specimens 1, 2A and 4 were constructed with
typical details. In particular two hold-down anchors were used on each of
these walls (one at each end). On the contrary, the specimen 2B was
constructed without hold-down anchors.

It was reported in the NAHB document that similar modes of failure were
observed in the specimens with hold-down anchors. The initial loading was
linear until the screws began to pull through the GWB, then; it was observed a
slight reduction in stiffness. As the load approached ultimate capacity the
OSB panels cracked at the perimeter screw connections.
The wall without hold-down anchors also showed failure of the interior
panels, although the exterior panels and the bottom track were unable to
distribute the uplift forces at the end of the wall. In fact, the bottom track
failed in bending due to uplift at the location of the first anchor bolt.

The NAHB presented the following conclusions:


The calculation of the shear capacity using the PSW design method
appeared valid, but revealed a conservative prediction of ultimate
shear strength.
Hold-down reduced uplift and increased the ultimate shear capacity by
allowing more sheathing-to-frame screws to resist shear.
The lateral load resisting mechanism for both wood-framed and steel-
framed shear walls appeared to be similar.

3.1.12 Gad et al. (1999a, b)


The Authors illustrated in this work an important investigation in the
seismic performance of residential structures with CFS frames. The research
involved an extensive racking and dynamic testing program on the both two-
and tree-dimensional framing assemblies with X-B bracing.
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 109

The objectives of the experimental program were the followings:


To study the interaction between the different components of a typical
wall assemblage.
To quantify the lateral stiffness and strength contributions of
plasterboard and determine its reliability under cyclic loads.
To investigate the inertial loading effect of brick veneer walls on the
framing assembly.
To develop guidelines that enable prediction of the behavior of
complete structures.

The experimental program was divided into two steps: testing of two-
dimensional unlined frames with different frame connection types and testing
of a one-room-house at various stages of construction.

The two-dimensional specimens measured 2400x2400mm (height x length)


and were constructed from CFS channel sections with steel grade of G300.
The section of the studs was 75x32x1.2mm (web depth x flange size x
thickness), for the plates (stud track) the section was 78x31x1.2mm and for
the noggings (blocking placed between all studs) was 72x34x1.2mm. Each X-
B was 1.0 thick and 25mm wide with steel grade of G250. For simulating the
mass of a steel sheet roof, a 350kg concrete beam was bolted to the top of the
frame.
The different connection types used were:
x Tab-in-slot connections, which are essentially pinned connections.
These connections represent the lower bound response.
x Welded connections, which represent the upper bound response.
On the walls with tab-in-slot connections were conducted slow cyclic
racking and dynamic tests. In particular, the dynamic tests included pluck
(applying a hammer blow on the concrete mass on the wall panel), swept sine
wave and simulated earthquakes.
The tests conducted on the welded frames were swept sine wave and
simulated earthquakes only.

The one-room-house specimen measured 2300x2400x2400mm high and


was constructed from full-scale components. It represents a section of a
110 Chapter III

rectangular house with plan dimensions of 11000x16000mm. The mass,


applied on the test house through a concrete slab fixed to the joists, was
2300kg. This mass corresponded to that due to roof tiles, battens, insulation,
ceiling lining and trusses for a plan area of 11000x2400mm. The two walls in
the east-west direction were framing assemblies with X-B bracing while the
two walls in the north-south direction were non-load bearing and had standard
900x2100mm door openings.
All construction details were standard except for the hold-down that was
over-designed to eliminate a potential uplift failure. All the framing members
were 75x35x1.0mm CFS channel sections with steel grade of G550. Tab-in-
slot connections were used for to connect the stud, plates and noggings.
Plasterboard lining, 10mm thick, was used for the ceiling and the walls, and
connected to the frame with 4.1x25mm (diameter x length) bugle head self
drilling screws spaced 200mm along the wall vertical edges, 600mm along the
top and bottom plates, 300mm in the field of the ceiling and 400mm in the
field of the walls. Skirting boards, 55mm ceiling cornices and set corner joints
were used in conjunction with the plasterboard lining. The brick veneer walls
were connected to the studs by metal clip-on brick ties.
The house was tested in the east-west and north-south directions at
different stages of construction. The tests conducted were mainly slow cyclic
racking, swept sine wave and simulated earthquakes.

From the two-dimensional tests results, the Authors concluded that the
important component in unlined single frames is the strap bracing system. In
particular they reported that:
The failure of the frame was governed by the failure of the X-B.
The dynamic characteristics of the frame were governed by the initial
tension in the straps. In fact, the welded frame had a lower natural
frequency (7.0Hz) than the tab-in-slot frame (6.3Hz) and this
contradicted the expectation that the welded frame would be stiffer
than the frame with tab-in-slot.
The type of connections between the framing members did not seem to
have an influence on the structural response of the braced frames.
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 111

Gad et al. concluded from the three-dimensional test results that residential
steel frames used in the tests performed well under racking and earthquake
loads. In particular, the Authors reported their conclusions for each different
stage of construction.
For unlined frames:
The behavior was governed by the X-B system.
The initial tension in the X-B increased the frame stiffness of the
frame and when the X-B yield then the true stiffness of the X-B system
(strap and its connections) defined the stiffness of the frame.
The type of strap bracing-to-plate connections governed the failure
load and mechanism.
For the lined frames:
Plasterboard fixed as non-structural component provided higher
stiffness, load carrying capacity and damping than X-B.
When X-B and plasterboard were combined, the overall stiffness and
strength of the system is simple addition of individual contributions
from X-B and plasterboard.
The plasterboard combined with ceiling cornices, skirting board and
set corner joints, resisted about 60-70% of the applied racking load
whereas the X-B resisted 30-40%.
For brick veneer walls:
In-plane brick veneer walls attached to the frame via clip-on ties did
not contribute to the stiffness of the system.
Different displacements between the frame and the out-of-plane brick
veneer walls were mainly accommodated by deformation of the stud
flanges rather than deformation of brick ties.

3.1.13 Selenikovich et al. (1999)


Selenikovich et al. presented the results of monotonic and cyclic tests on
six-teen full-size shear walls with and without openings.

The objectives of the study were:


To establish the effect of size of openings on shear wall performance.
To compare the shear strength of the walls with previsions of the PSW
design method (Sugiyama & Matsumoto, 1993).
112 Chapter III

To determine the effect of cyclic loading on shear behavior.


To examine the effect of the addition of internal GWB.

The specimens were built in accordance with the Builders Steel-Stud


Guide (AISI, 1996). All walls were 2440x12192mm (height x length) in size
with the same type of framing, sheathing, fasteners, and fasteners schedules.
The frames consisted in 89x38x0.84mm (web depth x flange size x thickness)
C-sections studs at 610mm on center. Double studs (back-to-back) were used
at the ends of the walls and around doors and windows. Exterior sheathings
were 11.1mm thick OSB sheathing oriented vertically. Interior 12.7mm thick
GWB sheathings oriented vertically were on an additional monotonic test. The
fasteners to connect the steel frame members were 4.2mm (diameter) low
profile head screws whereas to connect the sheathing to the frame were used
4.2mm bugle head screws. Sheathing screws were spaced 152mm on
perimeter and 305mm in field to attach OSB panels, and 178mm on perimeter
and 254mm in field for GWB. Hold-downs were used to connect each wall
specimen to the base of test frame.

The Authors observed that the predominant failure mode was head pull-
through of sheathing screws and bending of frame elements. They concluded
the following:
Long, fully sheathed walls were significantly stiffer and stronger but
less ductile than walls with openings.
The predictions of the PSW design method were conservative at all
levels of monotonic and cyclic loading.
Cyclic loading did not influence the elastic behavior of the walls but
reduced their deformation capacity.
The strength of fully sheathed walls was affected more significantly by
cyclic loading than walls with openings.
Adding of GWB panels increased the shear strength and stiffness of
fully sheathed walls under monotonic load.

3.1.14 COLA-UCI (2001)


The City of Los Angeles University of California (COLA UCI) tests
were carried out with the aim to develop an understanding of the probable
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 113

dynamic behavior of PLY and OSB panels attached to wood frame walls or
light-gauge steel frame by fasteners such as nails or screws.

The main goals of the study were:


To determine whether the shear wall system could be modeled in their
linear displacement range as principally shear deforming systems.
To determine the cyclic force-displacement relationships for
commonly used light-framed walls with shear panels.
To provide data for the improvement of the design of lateral-load
resisting shear wall systems.
To make recommendations to the City of Los Angeles if code changes
are warranted by the data and data analysis.

A total of thirty-six groups of 2440x2440mm (height x length) shear wall


with three specimens per group were tested under cyclic load. The groups
from 1 to 13 and from 20 to 36 had wood frame walls, whereas the groups
from 14 to 19 had light-gauge steel frame.
The specimens with steel frame included 11.1mm thick OSB or 11.9 thick
PLY, which were attached to frame with 4.2x25mm (diameter x length) bugle
head screws spaced 305mm in the field and 51, 102, or 152 at the perimeter.
The 89x41x10x0.84 C-sections studs were spaced at 610mm on center and
were connected to the 89x38mm (web depth x thickness) U-sections track.
Double studs were coupled back-to-back at the end of the wall to prevent local
and flexural buckling in the chords. Strap hold-down anchors were used to
connect each wall specimen to the base of test frame.

Tests results showed that:


A reduction of the fastener spacing nonlinearly increased the shear
strength and stiffness for both the light-gauge steel framed and wood
framed stud walls.
With the same sheathing types and fastener spacing, steel-framed walls
exhibited somewhat higher shear strength and ductility but less
hysteretic damping than wood-framed walls.
114 Chapter III

3.1.15 Dubina & Fulop (2002)


The experimental program presented by Dubina & Fulop was based on six
series of full-scale tests with different cladding arrangements. Each series
consisted of identical wall panels, tested statically both monotonic and cyclic.

The objectives of test program were the followings:


To compare monotonic and cyclic behavior of wall-stud shear walls.
To confirm the previous results about the effect of interior gypsum
cladding.
To study the effect of openings.
To compare the behavior of different cladding materials and cross
bracing.
To provide experimental information for the calibration of finite
element models.

All shear walls specimens were 2440x3600mm (height x length) in size.


The CFS frames consisted in 600S175-62 studs (150x1.5mm (web depth x
thickness) C-sections) at 600mm on center, while the tracks were 600T225-62
(154x1.5 (web depth x thickness) U-sections). Double studs (back-to-back)
were used at the ends of the walls and around openings.

The different wall specimen series included:


x Series 0: constituted by one wall specimen consisting of CFS frames
without some shear resisting system.
x Series I: represented by three wall specimens with exterior cladding
consisting of steel corrugated sheet (SCS) LTP20/0.5 placed in
horizontal position and fixed to the frame using SL2-T-A14
(4.8x22mm (diameter x length)) self-tapping screws spaced at 114mm
at sheet ends and at 229mm in the field.
x Series II: composed by three wall specimens with exterior cladding
similar to those of the series I and with interior cladding consisting of
12.5mm thick GWB placed in vertical position and fixed to the frame
at 250mm in the field and at the perimeter of the panel.
x Series III: constituted by two wall specimens braced by means of
110x1.5mm (width x thickness) steel straps on both sides of the frame.
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 115

The straps were fixed to the wall structures using SPEDEC SL4-F-
4.8x16 (4.8x16mm (diameter x length)) and SD6-T16-6.3x25
(6.3x25mm (diameter x length)) self-drilling screws. The numbers of
screws were designed for the yield strength of the strap.
x Series IV: composed by three wall specimens similar to those of the
series II except for the presence of the one 1200mm wide door.
x Series OSB I: composed by two wall specimens with exterior
consisting of 10mm thick OSB placed in vertical position and fixed to
the frame using bugle head self-drilling screws of 4.2mm diameter
spaced at 250mm in the field and at 105mm at the perimeter of the
panel.
x Series OSB II: composed by three wall specimens similar to those of
the series OSB I except for the presence of the one 1200mm wide
door.

The Authors reported that for Series I and II local deformation in the
uplifted corners was followed by profile-end distortion, gradual deformation
in connections and failure occurred in seam lines. The behavior of Series IV
was very similar to the ones in Series I and II, with much stronger corners
uplift. For the Series III after buckling of compressed straps in the early stage,
the local deformation of the lower track followed. Important plastic elongation
of the straps was observed, but because of this unexpected failure of the
corner occurred. In the case of Series OSB I the failure mechanism of the
specimens was different from SCS specimens due to different sheeting
orientation. Failure of the specimens was sudden when one vertical row of
screws unzipped from the stud and both pull over the screw head, and failure
of OSB margin was observed. For the Series OSB II important inclination of
the screws developed in the OSB panels-to-lower track connection, followed
by sudden rupture of this connection line.

The Authors concluded that:


Very significant pinching, and reduced energy dissipation characterize
the hysteretic behavior.
116 Chapter III

The behavior of gypsum panels was satisfactory. In fact they could


follow even extreme deformation of the wall without significant
damage.

3.1.16 Branston et al. (2003)


In an attempt to draw a link between the existing database of steel frame /
wood panel shear tests carried out in the USA and any future experiment with
Canadian products, a series of 12 match tests was conducted by Branston et
al..

The goal of the research project was to reproduce the results of some tests
completed by Serrette et al. (1996a) and COLA-UCI (2001) using steel and
wood panel products purchased in the USA and following the procedures set
up by previous researchers. In particular, the overall long-term objective of the
shear wall research is to investigate the performance of, and to provide
guidelines for the design of the light gauge steel structures when subjected to
earthquake loading.

In this research program, twelve full-scale steel frame shear walls, with
either OSB or PLY sheathing, were tested (6 monotonic load tests and 6 cyclic
load tests).
The frame was made using 89x41x10x0.84 C-sections studs spaced at
610mm on center attached to the 89x38mm (web depth x thickness) U-
sections track. Double studs were coupled back-to-back at the end of the wall
to prevent local and flexural buckling in the chords. The fasteners used for the
frame connections were 4.2x13mm wafer-head screws. The panels were
connected to the frame with 4.2x38mm bugle head screws. Wall specimens
were connected to the base of test frame with hold-downs (tie-downs) for
specimens OSB 4-8 US M-, OSB 4-8 US C- and PLY 8-8 US M- and with
strap hold-down anchors for specimens PLY 8-8 US C-.
The different tests included:
x OSB 4-8 US M- A, B, C: Monotonic tests of 1220x2440mm (height x
length) walls with 0.84mm thick steel framing sheathed on one side
with 11mm thick OSB. Self-drilling sheathing screws were spaced at
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 117

102mm at the perimeter and at 305mm in the field. (Previous tests:


Serrette et al. (1996a) OSB 1D3, 1D4).
x OSB 4-8 US C- A, B, C: Cyclic tests of 1220x2440mm (height x
length) walls with 0.84mm thick steel framing sheathed on one side
with 11mm thick OSB. Self-drilling sheathing screws were spaced at
102mm at the perimeter and at 305mm in the field. (Previous tests:
Serrette et al. (1996a) AISI OSB 3, 4).
x PLY 8-8 US M- A, B, C: Monotonic tests of 2440x2440mm (height x
length) walls with 0.84mm thick steel framing sheathed on one side
with 12mm thick PLY. Self-drilling sheathing screws were spaced at
152mm at the perimeter and at 305mm in the field. (Previous tests:
Serrette et al. (1996a) PLY 1A6, 1A7).
x PLY 8-8 US C- A, B, C: Cyclic tests of 2440x2440mm (height x
length) walls with 0.84mm thick steel framing sheathed on one side
with 12mm thick PLY. Self-drilling sheathing screws were spaced at
152mm at the perimeter and at 305mm in the field. (Previous tests:
COLA-UCI (2001) 14A. 14B, 14C).

In general, the failure of all wall specimens was restricted to the wood to
steel connections and the studs did not fail in compression, nor did the hold-
downs suffer any type of failure.

From the monotonic tests results, the Authors concluded that a variation in
the performance of the walls from the different research programs was
observed. In a number of instances the shear capacity and stiffness of the tests
specimens differed by a substantial amount. The variation in strength of the
walls can be attributed in general to the different material properties that may
have existed in both the steel and wood panel members. With respect to the
PLY 8-8 US C- A, B, C walls, the higher ultimate strength and stiffness
obtained for the COLA-UCI tests can most likely be traced to the use of
sheathing structural 1 grade panels in the original tests, whereas regular
sheathing grade PLY was used in the construction of the match tests.
The addition of a single horizontal stiffener located at the mid-height of the
wall did not increase the shear capacity and was not recommended since it
increased the construction difficulties and the cost of installation.
118 Chapter III

3.2 FACTORS AFFECTING CFSSSW BEHAVIOR UNDER


LATERAL LOADING: THEY INDIVIDUATION AND
ANALYSIS

The performance of house structures braced with cold-formed/light gauge


steel (CFS/LGS) framing under horizontal load is determined by the
horizontal (diaphragms) and vertical (stud walls) lateral force resisting
systems (LFRSs) behavior. If both horizontal diaphragms and vertical stud
walls are frames braced with structural panels and/or with diagonal bracing,
the same basic structural parameters influence their shear behavior. For
investigating this influence, the rather large number of experimental programs
conducted on CFSSSW with different assemblies has been illustrated in the
previous Sections. From the examination of these experimental programs, it is
possible to individuate the following basic factors that influence the lateral
behavior of CFSSSW:

sheathing (type, thickness, orientation);


framing (stud size, thickness and spacing, presence of X-bracing);
fastener (type, size, spacing);
geometry (wall aspect ratio, openings);
type of loading (monotonic or cyclic);
construction techniques and anchorage details.

Their effects on shear wall behavior are synthesized in the following. A


more complete summary of the conclusions obtained by several researchers is
given in Appendix B.

3.2.1 Sheathing (type, thickness, orientation)


The basic behavior of different sheathing types and thicknesses has been
studied by Tarpy (1980), Tarpy & Girard (1982), Serrette et al. (1996a, b,
1997a, b), Salenikovich et al. (1999), Dubina & Fulop (2002).
Those studies included walls constructed with PLY, OSB, GWB, GSB,
FBW, CP, SSS, and SCS. For understanding the different structural capacity
associated to different sheathing types and thicknesses, a comparison between
the average values of shear strength derived from the existing experimental
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 119

results is reported in Table 3.2. These values are normalized with respect to
average values of shear strength for walls braced with 11.1mm thick OSB.
From these comparisons, it is possible to conclude that the PLY provided a
little (about 10%) higher shear strength in comparison with the OSB; whereas
the capacity of the GWB and GSB is much (about 60%) lower than that of the
OSB. Moreover, the adding of GWB to the opposite side of the wall assembly
increased the shear strength by about 40%. Finally, the SSS presented a
relatively smaller (about 10%) shear strength than those of OSB.

Sheathing type and thickness No of specimens Average Standard deviation


OSB t=11.1mm 1.00
PLY t=11.9mm 13 1.10 0.09
GWB t=12.7mm + GWB t=12.7mm 2 0.83 0.01
OSB t=11.1mm+GWB t=12.7mm 4 1.39 0.58
SSS t=0.46mm 1 0.91
SSS t=0.69mm 3 0.86 0.12
GWB t=12.7mm+GSB t=12.7mm 2 0.78 0.36
+:double sheathing (internal and external sheathings)
t: thickness

Table 3.2: Normalized shear capacity for different sheathing type and thickness.

Serrette & Ogunfunmi (1996) and Serrette et al. (1994, 1997a) examined
the influence of sheathing orientation and the contribution of horizontal steel
straps installed at mid-height of the walls and fastened to the blocking. A
comparison between the shear strength for walls with sheathing oriented
vertically and horizontally, with and without solid blocking, is reported in
Figure 3.1. In particular, this comparison demonstrates that the panels oriented
parallel to the studs provided the same shear strength than those applied
perpendicular to the studs with strap blocking. In addition, when in the case of
panels perpendicular to studs blocking was omitted, the shear capacity of the
wall, was reduced by more than 50%.
120 Chapter III

30
VERTICAL
Ultimate shear
HORIZONTAL + SOLID BLOCK
strength HORIZONTAL
25 [kN/m]

Serrette et al . Serrette et al . Serrette et al . Serrette et al . Serrette et al .


(1996a) (1997a) (1994) (1997a) (1997a)
20

PLY-T7
PLY-T4
1A5,1A6

OSB-T5
OSB-T4
1A2,1A3

GYP-T2
15

8
6

GYP-T1
10

PLY-T6
7
5

0
OSB OSB PLY PLY GWB+GWB

Figure 3.1: Influence of the sheathing orientation.

3.2.2 Framing (stud size, thickness and spacing, presence of X-bracing)


The effects of framing stud size, thickness and spacing have been studied
by Tarpy & Girard (1982), Tissel (1993) and Serrette et al. (1997b). Those
studies indicated that the use of thicker (1.09mm thickness) and multiple
chord studs (back-to-back coupled C-sections) allowed the development of the
full shear strength of the wall panels. In addition, the reduction of the stud
spacing did not increase significantly the shear performance.
The contribution of steel flat strap tension X-bracing has been investigated
by Serrette & Ogunfunmi (1996), Serrette et al. (1997b) and Gad et al.
(1999a). In particular, Serrette et al. concluded that, although the adding of
flat strap improved the shear behavior, their employment is not practical.

3.2.3 Fastener (type, size, spacing)


An extensive experimental study on the effect of fastener type, size and
spacing has been performed by Tarpy & Hauenstein (1978), Tarpy (1980),
Tarpy & Girard (1982), Tissel (1993), Serrette et al. (1996a, b, 1997a, b), Gad
et al. (1999a), COLA-UCI (2001). Those studies showed that the reduction of
the fastener spacing at the panel edges improved significantly the shear
performance. Furthermore, they proved that for PLY (11.9mm thick) or OSB
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 121

(11.1mm thick) panels fastened to members with a thickness from 0.84mm to


1.09mm, 4.2x25mm (diameter x length) flat-head or bugle-head self-drilling
screws should be used, while for gypsum panels (12.7mm thick) 3.5x25.4mm
bugle-head self-drilling screws should be adopted.

3.2.4 Geometry (wall aspect ratio, openings)


McCreless & Tarpy (1978), Tarpy & Girard (1982) and Serrette et al.
(1996a, 1997b) studied the effect of the wall aspect ratio (height/length)
variation. The shear strength values obtained from the cited test results are
reported as a function of aspect ratio (H/L) in Figure 3.2. The Figure
illustrates, according to the McCreless & Tarpy (1978) and Serrette et al.
(1996a) investigations, how the shear strength is practically insensitive to the
aspect ratio, when this is in the range [0.3 , 2]. Moreover, according to results
given in Serrette et al. (1997b), the shear strength appreciably decreases when
the aspect ratio increases from 2 to 4.
The NAHB Research Center (1997), Salenikovich et al. (1999) and Dubina
& Fulop (2002) examined the influence of the opening size. The shear
strength values obtained from their tests are reported as a function of the
sheathing area ratio (r) in Figure 3.3, where r is defined in following equation:
1
r (3.1)
Ao
1
h Li
with Ao: the total area of openings; h: the height of the wall; and Li: the length
of the full height wall segment.
The Figure shows that a reduction of r increases the shear capacity. The
NAHB Research Center and Salenikovich et al. also investigated the
suitability of using the PSW design method, originally developed for wood
structures (Sugiyama & Matsumoto 1994), also in case of CFSSSW. In

particular, they concluded that the PSW method reveals a conservative


prediction of shear strengths.
122 Chapter III

M78: McCreless & Tarpy (1978); TG82: Tarpy & Girard (1982); S+96a:
Serrette et al. (1996a); S+97a: Serrette et al. (1997a)
H: height of the wall; L: length of the wall

Figure 3.2: Influence of the aspect ratio.

N97: NAHB Research Center (1997); S+99: et al. (1999); FD02: Dubina &
Fulop (2002)
-m: monotonic test; -c: cyclic test
U=1/(1+Ao/H6Li): sheathing area ratio
Ao: total area of openings; H: height of the wall; Li: length of the full height
wall segment

Figure 3.3: Influence of the openings.


Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 123

3.2.5 Type of loading (monotonic or cyclic)


The effect of cyclic loading on the lateral strength, stiffness and ductility of
steel stud shear walls systems has been investigated by Serrette et al. (1996a,
b and 1997b), Salenikovich et al. (1999), COLA-UCI (2001), Dubina & Fulop
(2002) and Branston et al. (2003). The results of those studies showed that
cyclic loadings reduce the shear performance of steel stud shear walls, in
comparison with monotonic loads.
The ultimate shear strengths obtained under both monotonic and cyclic
loading tests, using nominally identical specimens are shown in Figure 3.4.
The comparison between monotonic and cyclic test results reveals that the
shear strength under cyclic loads is about 10% in average lower than that
under monotonic load.
Moreover, a typical cyclic response of CFSSSW is characterized by
significant pinching of the hysteresis response, with reduced energy
dissipation capacity, as shown in Figure 3.5 for a test performed by Branston
et al. (2003).

S+96a: Serrette et al. (1996a); S+96ab: Serrette et al. (1996a, b); S+97b:
Serrette et al. (1997b); S+99: Salenikovich et al. (1999); FD02: Dubina &
Fulop (2002); B03: Branston et al. (2003)

Figure 3.4: Influence of cyclic loading.


124 Chapter III

Figure 3.5: Cyclic response for OSB 4-8 US C-A test performed by Branston et al. (2003).

3.2.6 Construction techniques and anchorage details


The study of the effect of construction techniques and anchorage details
has been performed by Tarpy & Hauenstein (1978), Tarpy (1980), Tarpy &
Girard (1982), NAHB Research Center (1997) and Gad et al. (1999a). Almost
all the cited references highlight the important effect of the corner foundation
anchorage details on the shear response of steel stud walls. In fact, if the uplift
loads are not directly transmitted from the end studs to the foundation by rigid
attachment as such as clip angles or hold-downs, then a significant bending of
the bottom track occurs, up to failure with a reduced shear strength and
stiffness.

REFERENCES

ASTM E 564-76 (1976) Standard Practice for Static Load Test for Shear Resistance
of Framed Walls for Buildings. ASTM (American Society for Testing and Materials).
West Conshohocken, PA, USA.
Testing of co ld-fo rmed steel stud sh ear walls: rev iew of existing literatu re 125

AISI (1996) Builders Steel-Stud Guide. Publication RG-9607, AISI (American Iron
and Steel Institute). Washington DC.
Brantson, A., Boudreault, F., Rogers, C.A. (2003) Testing on steel frame / wood
panels shear walls. Progress Report, Departement of Civil Engineering and Applied
Mechanics, McGill University. Montreal.
COLA-UCI (2001) Report of a testing program of light-framed walls with wood-
sheathed shear panels. Final report to the City of Los Angeles Department of
Building and Safety, Structural Engineers Association of Southern California. Irvine,
CA, USA.
Dubina, D. & Fulop, L.A. (2002) Seismic performance of wall-stud shear walls. In
Proceedings of the 16th International Specialty Conference on Cold-formed Steel
Structures. St. Louis, MO, USA: 483-500.
Gad, E.F., Duffield, C.F., Hutchinson, G.L., Mansell, D.S., Stark, G. (1999a) Lateral
performance of cold-formed steel-framed domestic structures. Engineering
Structures, Elsevier, Vol.21, No.1: 83-95.
Gad, E.F., Chandler, A.M., Duffield, C.F., Stark, G. (1999b) Lateral behaviour of
plasterboard-clad residential steel frames. Journal of Structural Engineering, ASCE,
Vol.125, No.1: 32-39.
McCreless, S. & Tarpy, T.S. (1978) Experimental investigation of steel stud shear
wall diaphragms. In Proceedings of the 4th International Specialty Conference on
Cold-formed Steel Structures. St. Louis, MO, USA: 647-672.
NAHB Research Center (1997) Monotonic Tests of cold-formed steel shear walls
with openings. (NAHB) National Association of Home Builders. Upper Marlboro,
MD, USA.
Salenikovich, A.J., Dolan, J.D., Easterling, W.S. (2000) Racking performance of long
steel-frame shear walls. In Proceedings of the 15th International Specialty Conference
on Cold-formed Steel Structures. St. Louis, MO, USA: 471-480.
Serrette, R. (1994) Light gauge steel shear wall test. Light Gauge Steel Research
Group, Department of Civil Engineering, Santa Clara University. Santa Clara, CA,
USA.
Serrette, R.L. & Ogunfunmi, K. (1996) Shear resistance of gypsum-sheathed light-
gauge steel stud walls. Journal of structural engineering, ASCE, Vol.122, No.4: 386-
389.
Serrette, R., Nguyen, H., Hall, G. (1996a) Shear wall values for light weight steel
framing. Report No.LGSRG-3-96, Light Gauge Steel Research Group, Department of
Civil Engineering, Santa Clara University. Santa Clara, CA, USA.
126 Chapter III

Serrette, R., Hall, G., Nguyen, H. (1996b) Dynamic performance of light gauge steel
framed shear walls. In Proceedings of the 13th International Specialty Conference on
Cold-formed Steel Structures. St. Louis, MO, USA: 487-498.
Serrette, R.L., Encalada, J., Juadines, M., Nguyen, H. (1997a) Static racking behavior
of plywood, OSB, gypsum, and fiberboard walls with metal framing. Journal of
Structural Engineering, ASCE, Vol.123, No.8: 1079-1086.
Serrette, R., Encalada, J., Matchen, B., Nguyen, H., Williams, A. (1997b) Additional
shear wall values for light weight steel framing. Report No.LGSRG-1-97, Light
Gauge Steel Research Group, Department of Civil Engineering, Santa Clara
University. Santa Clara, CA, USA.
Sugiyama, H & Matsumoto, T. 1984. Empirical equations for the estimation of
racking strength of plywood-sheathed shear walls with openings. Summaries of
Technical Papers of Annual Meeting, Transactions of the Architectural Institute of
Japan, No.338.
Tarpy, T.S. (1980) Shear resistance of steel-stud wall panels. In Proceedings of the
5th International Specialty Conference on Cold-formed Steel Structures. St. Louis,
MO, USA: 331-348.
Tarpy, T.S. & Girard, J.D. (1982) Shear resistance of steel-stud wall panels. In
Proceedings of the 6th International Specialty Conference on Cold-formed Steel
Structures. St. Louis, MO, USA: 449-465.
Tarpy, T.S. & Hauenstein, S.F. (1978) Effect of construction details on shear
resistance of steel-stud wall panels. Vanderbilt University. Nashville, TN, USA. A
research project sponsored by American Iron and Steel Institute. Project No.1201-
412.
Tissell, J.R. (1993) Wood structural panel shear walls. Report No.154, APA (The
Engineering Wood Association). Tacoma, WA, USA.
127

Chapter IV
Evaluation of seismic capacity:
the monotonic test

The number of experimental tests, which have been performed to study the
lateral behavior of cold-formed steel stud shear walls (CFSSSWs) is rather
wide. The basic factors that influence the shear response of CFSSSWs have
been objects of previous studies, as presented in Chapter 3. Although the
existing experimental database appears sufficient, there has been little
research on the lateral-load transfer from the horizontal diaphragms to stud
walls. The same consideration applies for the influence of construction details,
such as bearing stiffeners and joist track profiles, anchorage details,
connections between joists, bearing stiffeners, joist track and top stud track.
Moreover, there has not been specific study on the effect of gravity loads on
the shear walls behavior even if it may be anticipated to be small. For this
reason, an experimental phase based on two identical stud shear wall sub-
assemblies, tested under both monotonic and cyclic loading has been
dedicated to evaluate the seismic capacity.
In this Chapter, the following Section (Section 4.1) is dedicated to define
the study case and the relevant assumptions. In the Section 4.2 the general
design principles are presented. The experimental program is illustrated in
Section 4.3. In particular, for both monotonic and cyclic tests the description
of test specimens, the test set-up and instrumentation are presented in Sections
4.3.1 through 4.3.3. While the Section 4.3.4 is devoted to illustrate the test
procedure for the only monotonic test. Finally, the test results for the
monotonic test are presented in Section 4.4.
128 Chap ter IV

4.1 THE STUDY CASE: BASIC ASSUMPTIONS

The study case examined is a typical one-family one-story dwelling as that


shows in Figure 4.1. The plan dimensions of the house are about 7 x 11m,
while its height is about 6m. The assumptions on the main dimensions of the
structure are presented in Table 4.1.
The structure is a stick-built construction in which both horizontal (roof
and floors) and vertical (walls) diaphragms are cold-formed frames sheathed
with structural panels.

L=11.4m

W=7.0m

(a) first floor


N

W=6.4m

(b) east elevation

Figure 4.1: Typical analyzed stick-built house.


Evaluatio n of seismic capacity: th e mo notonic test 129

Number of stories (n) 2


Length (L) 11.4 m
Wide (W) 7.0 m
Height (H) 6.4 m
Roof slope 100%
Area (A) 75 m2

Table 4.1: Main structural dimension.

Element Unit load


roof 0.75 kN/m2
floor 0.75 kN/m2
walls 0.35 kN/m2

Table 4.2: Unit dead loads.

According to the units load reported in Table 4.2, the assumed


characteristic gravity loads, to be considered for computing the lateral seismic
forces, are:
x dead loads: gk = 2.2 kN/m2 (including roof, floor and walls);
x snow loads: qks = 0.5 kN/m2;
x live loads: qkl = 2.0 kN/m2;

As it may be observed by comparison between dead, snow and live load


values, the great advantage of this structural typology is the lightness. In fact
the dead loads represent only about 50% of the total load.
Therefore, according to rules given in Eurocode 1 Part 1 (ENV 1991
1996) for the loads that must be considered acting together with seismic loads,
dead and live loads are combined in this way:
wk g k  \ EI qks  qkl = 3.0 kN/m2 (4.1)
in which:
wk: is the seismic weight;
\EI = 0.3 is the combination coefficient.

For the choice of the seismic area, it is assumed that the house is located in
a medium seismic zone in Central Italy (design value of the peak ground
acceleration ag=0.25g). The soil conditions assumed are, according to
130 Chap ter IV

classification of the Eurocode 8 (prEN 1998-1 2001): (A) rock; (B) stiff soil;
(C) soft soil; (D) very soft soil; and (E) alluvium soil. Finally, the Type 1 (far
field) design elastic spectrum Saed(T) reported in Eurocode 8 is adopted:
T
S aed (T ) a g S 1  K 2.5  1 if 0 T TB
TB
S aed (T ) a g S K 2.5 if TB T TC (4.2)
T
S aed (T ) a g S K 2.5 C if TC T TD
T
T T
S aed (T ) a g S K 2.5 C 2 D if T TD
T
where:
Saed(T): is the design elastic spectrum acceleration;
T: is the first mode vibration period;
S: is the soil factor;
K: is the damping factor (K=1 for a damping ratio Q=0.05)
ag: is the design peak ground acceleration;
TB and TC: are the limits of constant spectral acceleration branch;
TD : is the value that defines the beginning of the constant
displacement response range.

The factors S, TB, TC, and TD, for the horizontal direction are reported in
Table 4.3.
Soil type S TB TC TD
A 1.00 0.15 0.4 2.0
B 1.20 0.15 0.5 2.0
C 1.15 0.20 0.6 2.0
D 1.35 0.20 0.8 2.0
E 1.40 0.15 0.5 2.0

Table 4.3: Values of parameters describing the Eurocode 8 Type 1 elastic response spectrum.

For the purpose of seismic analysis the following hypotheses are adopted:
x the ground motion acting in North-South direction (see Fig. 4.1);
x the floor and roof sheathings act as rigid diaphragms;
x the possible torsional effects are neglected;
Evaluatio n of seismic capacity: th e mo notonic test 131

x the dynamic lateral behavior of the structure is represented by the


dynamic lateral behavior of the walls;
x the Segment method is used for describing the shear behavior of the
CFSSSWs, assuming that the sum of the length of the full height wall
segments (6wi) in North-South direction is 6wi=18.0m;
x the dynamic lateral behavior of the CFSSSWs is described by a SDOF
system;
x the acting lateral force corresponding to design seismic action should
be less than the force corresponding to the elastic limit state of the
CFSSSWs.

According to the listed assumptions, it is possible to define basic


parameters of the SDOF system representing a piece of wall with unit length
(1m). In fact, the seismic weight per unit length (ws) results:
wk A
ws = 12.5 kN/m (4.3)
wi
and, considering the following lateral stiffness per unit length (k) (Gad et al.
1999):
k = 0.5 3.0 kN/mm/m (4.4)
it is possible to compute the natural vibration period (T) of the SDOF:
ws / g
T 2S = 0.1 0.3 s (4.5)
k
Finally, a damping ratio Q=0.05 is adopted.

4.2 GENERAL DESIGN PRINCIPLES

The structure has been designed in accordance with the Prescriptive


Method For Residential Cold-Formed Steel Framing (NASFA 2000).
The wall height of the example house is 2500mm. Studs are made of
100x50x10x1.00mm (web depth x flange size x lip size x thickness) C
(channel lipped)-section, while joists consist of 260x40x10x1.50mm C-
section. The studs and the joists are spaced 600mm on center. The walls are
sheathed with 9.0mm thick oriented strand board (OSB) external panels and
12.5mm thick gypsum wallboard (GWB) internal panels. The floor sheathings
132 Chap ter IV

are made of 18.0mm thick OSB panels. The panel-to-wall framing


connections are 4.2x25mm self-drilling flat head screws for the OSB panels
and 3.5x25mm self-drilling bugle head screws for GWB panels.
All CFSSSW components (steel members, panels and connections) are
designed according to capacity design principles, in such a way to promote the
development of the full shear strength of panel-to-wall framing connections.

For determining the panel-to-wall framing connections spacing, the lateral


force (vS) acting on a wall with unit length may by calculated by the following
formula:
v S S aed (T ) ws (4.6)
in which the weight per unit length is ws=12.5 kN/m (from Eq. 4.3), ag=0.25,
K=1, and the design elastic spectrum acceleration (Saed(T)) may be obtained
from Equation 4.2 for different soil types, as reported in Figure 4.2.
Considering the natural vibration period T ranging from 0.1 to 0.3s (according
to Eq. 4.5), the maximum values of the design elastic spectrum acceleration
(Saed(T=0.1-0.3s)max) and corresponding lateral force values may be obtained,
as reported in Table 4.4.
Finally, considering the maximum value of the acting lateral force
(vS=11kN/m), panel-to-wall framing connections spaced at 150mm at the
perimeter and at 300mm in the field may be adopted. In fact, for this fastener
spacing, the estimated lateral strength is about vR=18kN/m, while the force
corresponding to the elastic limit state is about vE = vR / 1.5 = 12kN/m.
All calculations are shown in detail in Appendix C.
Evaluatio n of seismic capacity: th e mo notonic test 133

1.0
Saed(T) / g
0.9

0.8 E

0.7 C D
0.6

0.5

0.4
A B
0.3

0.2

0.1
T [s]
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2.0

Figure 4.2: Eurocode 8 Type 1 elastic spectrum acceleration.

Soil type Saed(T=0.1-0.3s)max / g vS (kN/m)


A 0.63 7.9
B 0.75 9.4
C 0.72 9.0
D 0.84 10.5
E 0.88 11.0
max 0.88 11.0

Table 4.4: Maximum elastic spectrum acceleration and corresponding lateral force values.

4.3 TEST PROGRAM

Although the existing experimental database is rather wide, there has been
little research on the lateral-load transfer from the horizontal diaphragms to
stud walls. The same consideration applies for the influence of construction
details, such as bearing stiffeners and joist track profiles, anchorage details,
connections between joists, bearing stiffeners, joist track and top stud track.
Moreover, there has not been specific study on the effect of gravity loads on
the shear walls behaviour even if it may be anticipated to be small.
134 Chap ter IV

For this reason, the main part of the experimental program has been based
on two full-scale identical CFSSSW sub-assemblies, tested under both
monotonic and cyclic loading.

4.3.1 Description of test specimens


The generic specimen of the stud shear wall sub-assembly is shown in
Figures 4.3 a and b. In particular, Figure 4.3a provides a drawing of the
generic specimen, while Figure 4.3b shows a picture of its set-up.
The sub-assembly is a satisfactory model of the typical lateral load
resisting systems of house structures braced with CFSSSWs and sheathed with
structural panels. Typical construction details are shown in Figures 4.4 and
4.5. In particular, Figure 4.4a illustrates a drawing of the detail of the
foundation anchorage of the end studs; besides the installation of the hold-
down connector is shown in Figure 4.4b. Figure 4.5 shows, instead, the detail
of the connections between floor joists and wall studs.
The specimen was designed in accordance with general design principles
given in the Section 4.3. In fact, the generic wall framing, which was 2400mm
long and 2500mm height, consisted of single top and bottom tracks made of
100x40x1.00mm (web depth x flange size x thickness) U (unlipped)-sections
and both single intermediate and double back-to-back end studs, made of
100x50x10x1.00mm (web depth x flange size x lip size x thickness) C
(lipped)-sections, spaced 600mm on center. The floor framing consisted of
260x40x10x1.50mm C-section joists, which were spaced 600mm on center,
with single span of 2000mm and of 260x40x1.00mm U-section tracks.
Evaluatio n of seismic capacity: th e mo notonic test 135

(a) Drawing of specimens (b) Set-up of specimens

Figure 4.3: Global 3D view of the generic sub-assembly.

intermediate
end stud stud

hold-down
connector
shear
anchor
hold-down
anchor

stud track

X-bracing

(a) Drawing of detail of foundation anchorages (b) installation of a hold down connector

Figure 4.4: A close-up view of the connection between end studs and foundation.

joist track floor OSB


panel

bearing
stiffener

stud track
joist
wall
OSB panel
end stud
wall
GWB panel

X-bracing

Figure 4.5: A close-up view of the connection between the floor joists and the wall studs.
136 Chap ter IV

All frame members were cold-formed and fabricated from FeE350G


(S350GD+Z/ZF) hot dipped galvanized (zinc coated) (EN 10142 2002) Grade
steel (nominal yield strength fy=350MPa and nominal tensile strength
ft=420MPa). The used cold-formed steel (CFS) profiles were manufactured by
the Company GUERRASIO (Roccapiemonte, Salerno, Italy). They are shown
in Figure 4.6.
Wall and floor external sheathings were made of Type 3 oriented strand
board (OSB/3) (EN 300 1997), which were manufactured by the Company
KRONO FRANCE. The thicknesses of the wall and floor OSB sheathings
were 9.0 and 18.0mm, respectively. The internal sheathings of the wall were
made of 12.5mm thickness gypsum wallboards (GWB) (ISO 6308 1980),
which were manufactured by the Company BPB ITALIA. The sheathing
orientation was always vertical. Both OSB and GWB sheathings are shown in
Figure 4.7 a and b, respectively.
The fasteners adopted for the frame (steel-to-steel) connections were
4.2x13mm (diameter x length) self-drilling modified-truss head screws. For
the stud-to-OSB and stud-to-GWB sheathing connections, 4.2x25mm self-
drilling flat head screws and 3.5x25mm self-drilling bugle head screws were
used, respectively. These screws were spaced at 150mm at the perimeter and
at 300mm in the field. 4.2x38mm self-drilling flat head screws were adopted
for joist-to-OSB sheathing connections. The screws, which were used to
assemble the specimens, are shown in Figure 4.8. They were manufactured by
the Company TECFI s.r.l. (Villaricca, Naples, Italy).
The foundation was simulated by two 280x380mm (depth x width)
rectangular concrete beams. The walls were connected to the foundation by
intermediate shear anchors and purposely-designed welded steel hold-down
connectors were used at the end studs (see Fig. 4.9). In particular, as shear
anchors 8.0mm diameter mechanical anchors were employed (see Fig. 4.10a),
while adhesive-bonded anchors were adopted to connect the hold-down
connectors with the foundation (see Fig. 4.10b). Both the used mechanical and
adhesive-bonded anchors were manufactured by the Company HILTI
ITALIA.
Evaluatio n of seismic capacity: th e mo notonic test 137

(1) (2) (3) (4) (5)


(1) bearing stiffener; (2) stud; (3) stud track;
(4) joist; (5) joist track.

Figure 4.6: CFS profiles (manufactured by GUERRASIO).

(2)

(1)

(a) (1) OSB wall sheathing; (2) OSB floor sheathing


(b) GWB sheathing manifactured by BPB ITALIA
manufactured by KRONO FRANCE

Figure 4.7: Sheathings.

(1) (2) (3) (4)

(1) 3.5x25mm bugle head; (2) 4.2x32mm flat head; (3)


4.2x25mm flat head; (4) 4.2x13mm modified truss head.

Figure 4.8: Self-drilling screws (manufactured by TECFI s.r.l.).


138 Chap ter IV

Figure 4.9: Hold-down connector.

(1)

(2)

Adhesive-bonded anchors
HST M8 mechanical anchors (1) HIT-RE 500; (2) HIS-N(8.8) M20 3.5x25mm bugle head.

Figure 4.10: (a) shear and (b) hold-down anchors (manufactured by HILTI ITALIA).

All the components of the stud shear wall sub-assembly (members,


sheathings and connections) have been designed according to capacity design
principles, in such a way to promote the development of the full shear strength
of sheathing-to-wall framing connections. For this reason, the shear anchors
and the hold-down connectors have been designed to prevent either shear
failure at the base of the walls or failure due to overturning. Besides, double
lipped C-sections at the ends of the walls have been adopted to avoid failure
due to buckling in end studs.
Evaluatio n of seismic capacity: th e mo notonic test 139

Global transverse stability of the sub-assembly was achieved by means of


pre-tensioned steel round bar X-bracing (see Figures 4.3a, 4.4a and 4.5).
Main details of the specimen components are given in Table 4.5, whereas
drawings of the main construction details of the sub-assembly specimen are
reported in Appendix D.

Cold formed steel (CFS) members


FeE350G (S350GD+Z/ZF) hot dipped galvanized (zinc coated) steel
Steel grade
(Nominal yield strength fy=350MPa; nominal tensile strength ft=420MPa)
Studs C 100x50x10x1.00mm (2400mm long) (by GUERRASIO)
Wall members
Tracks U 100x40x1.00mm (2700mm long) (by GUERRASIO)
Joist C 260x40x10x1.50mm (2200mm long) (by GUERRASIO)
Floor
tracks U 260x40x1.00mm (2700mm long) (by GUERRASIO)
members
Bearing stiffeners C 100x50x10x1.00mm (260mm long) (by GUERRASIO)
Sheathings
1200x2500x12.5mm (width x height x thickness) GWB vertically oriented
Interior
Wall (PLACOLAST PLACO by BPB ITALIA)
sheathings 1250x2500x9.0mm Type 3 OSB vertically oriented (KRONOPLY 3 by
Exterior
KRONO FRANCE)
1250x2500x18.0mm Type 3 OSB vertically oriented (KRONOPLY 3 by
Floor sheathing
KRONO FRANCE)
Concrete beams foundation
Concrete type C20/25 (characteristic strength on a cube 150x150x150mm fck,cube=25MPa)
Dimensions 380x280mm rectangular section (3590mm long)
Frame-to-foundation connections
Hold-down
Purposely-designed welded steel hold-down
connector
Hold-down anchors HIT-RE 500 with HIS-N(8.8) M20 adhesive-bonded anchors (by HILTI ITALIA)
Shear anchors HST M8 mechanical anchors (by HILTI ITALIA) spaced at 100mm
Steel-to-steel connections
4.2x13mm (diameter x lenght) modified truss head self drilling screws (by
CFS members
TECFI s.r.l.)
CFS members-to-hold
6mm diameter bolts
down connector
Steel-to-Sheathing connections
3.5x25mm bugle head self drilling screws (by TECFI s.r.l.) spaced at
Interior
150mm at the perimeter and at 300mm in the field
Walls
4.2x25mm flat head self drilling screws (by TECFI s.r.l.) spaced at 150mm
Exterior
at the perimeter and at 300mm in the field
4.2x32mm flat head self drilling screws (by TECFI s.r.l.) spaced at 150mm
Floor for sheathing-to-track connections and at 250mm for sheathing-to-joist
connections

Table 4.5: Full-scale specimen materials and construction data.


140 Chap ter IV

4.3.2 Test set-up


Two types of load were applied: gravity and racking loads.
The CFSSSW sub-assembly simulated a section of a rectangular house with
plan dimension equal to a x b = 3.6x4.2m. The applied total gravity load was
equal to 45kN. This load was equivalent to dead (including roof, floor and
walls), snow and live loads applied to the plan area, it being: wk x a x b =
3.0x3.6x4.2=45kN, where wk is the seismic weight (see Eq. 4.1).
Racking loads were uniformly applied to the floor by means of two
programmable servo-hydraulic actuators (MTS System Corporation), each one
characterised by a range of displacement of 500mm and a load capacity of
500kN. The concentrated actuators loads were transformed into a distributed
load applied to the floor panels by means of a purposely-designed load
transfer system that is shown in Figure 4.11. It consisted of transversal and
longitudinal hot rolled members. In particular, four CPN100 transversal
profiles were positioned over the floor sheathings, while the transversal
members placed lower down the floor sheathings were two CNP100 (external
profiles) and two flat straps (internal profiles). The upper and bottom
transversal members were bolted back-to-back using two 8.0mm diameter
bolts spaced at 250mm on center. In such a way the transversal profile and the
floor sheathings were closely connected together. As longitudinal members
four HEA100 (two for each side) profiles were used. These profiles were
bolted with the transversal profile using four 14mm diameter bolts for each
joint, as illustrated in Figure 4.11.
The load actuators were restrained to keep the horizontal position and a
sliding-hinge is placed between the load actuator and the ends part of the
longitudinal members, in order to act as a filter transferring to the structure
exclusively horizontal racking loads. This testing apparatus allowed the
capacity of the floor to transmit forces to the walls to be checked, up to failure
of the vertical stud-to-sheathing connections. The Figure 4.12 shows the
global view of the test set-up.
Evaluatio n of seismic capacity: th e mo notonic test 141

No. 2 - 8mm
CNP100 diameter bolts
HEA100 spaced at 250mm
on center

No. 4 -
14mm
diameter
bolts

HEA100
CNP100
Flat strap
100x6mm

Figure 4.11: Load transfer system.


142 Chap ter IV

Figure 4.12: Test set-up.

4.3.3 Instrumentation
Fourteen linear variable differential transducers (LVDTs) were used to
measure displacements of the specimens during the tests, as shown in Figures
4.13 a and b. In particular, five LVDTs were installed for each wall (Fig.
4.13a): LVDTs w1 was used to measure the horizontal displacements at the
top of the wall (lateral in-plane displacement); LVDTs w2 and w3 measured
the horizontal displacements at the bottom of the wall (lateral in-plane
sliding); LVDTs w4 and w5 measured the vertical displacements at the bottom
of the wall (uplift and compression). One LVDT (f1) was installed for each
foundation beam (Fig. 4.13a) to measure the horizontal displacement. Finally,
two LVDTs (d1 and d2) were installed on the floor (Fig. 4.13b) to measure the
lateral longitudinal displacements.
The horizontal displacements of the floor (d1 and d2) and walls (w1, w2
and w3) were measured using LVDTs with a range of displacement of 250mm
and an accuracy of 0.02mm. For measuring the horizontal displacements of
the foundations (f1 and f2) LVDTs with a range of displacement of 100mm
Evaluatio n of seismic capacity: th e mo notonic test 143

and an accuracy of 0.01mm were used. Finally, the vertical displacements of


the walls (w4 and w5) were measured using LVDTs with a range of
displacement of 20mm and an accuracy of 0.01mm. All LVDTs used for this
testing program were Penny & Giles Controls.
Only for a wall (wall 2), two clinometers with a range of measuring of 20
and an accuracy of 0.02 (i1 and i2) were used to measure the angle between
the end studs at the bottom of the wall and the horizontal plane.
The load was measured through the actuator load cells.
Data was acquired at 20Hz (one data point every 0.05 seconds).

a
a2 Wall 2
d2
w1

w4 w5

w2 w3
d1
f1
i1 i2 a1 Wall 1
(a) Walls (b) Floor
actuator load (No. 2)
LVDT (No. 14)
inclinometer (No. 2)

Figure 4.13: Instrument arrangement.

4.3.4 Test procedure for the monotonic test


In the monotonic loading regime, the specimen was subjected to progressive
horizontal deflection. In particular, the loading procedure for the monotonic
test was articulated in two phases.
In the first phase, to evaluate the permanent set at 2, 4, 6 and 10mm, the
specimen were unloaded at these displacements. In the second phase, the
specimen was loaded up to a displacement of 150mm without unloaded. This
test protocol involved displacements at rate of 0.10mm/s for displacements
less than 10mm and at rate of 0.20mm/s for displacement exceeded 10mm.
144 Chap ter IV

4.4 TEST RESULTS

The measured responses of all instruments are reported in Appendix E.


The global behavior of the CFSSSW sub-assembly under monotonic
racking loads may be represented through the relationship between the
measured unit shear resistance (v) and the mean displacement (d) of the
specimen. In particular, the value of v and d are defined as:
V1  V2
v (4.7)
Lt
d1  d 2
d (4.8)
2
where:
V1 and V2: are the forces measured by the actuators a1 and a2, respectively;
d1 and d2: are the displacements measured by the actuators a1 and a2,
respectively;
Lt=4.800m: is the total length of the walls.

The v-d response curve is showed in Figure 4.14.


In the same Figure the values of the experimental lateral strength
(vEXP=18.5kN/m), estimated lateral strength (vR=17.9kN/m) and acting seismic
force (vS=11.0kN/m) are also reported. From the comparison between the
experimental and the estimate lateral strength, it is possible to observe that the
calculation of the shear strength using the semi-empirical methodologies
illustrated in the Chapter 2 (developed in detail in Appendix C for the study
case) appears to be valid, it revealing a small conservative prediction of 3%.
Evaluatio n of seismic capacity: th e mo notonic test 145

2 Experimental lateral strength (vEXP = 18.5kN/m)

Estimate lateral strength (vR = 17.9kN/m)

1 Acting seismic force (vS = 11.0kN/m)

Figure 4.14: Unit shear resistance (v) vs. mean displacement (d) curve.

In Figure 4.14 the following significant load steps are represented:


x Step 1: lateral displacement equal to 10mm. At this step, in OSB
sheathings-to-frame connections the tilting of the screws about the
plane of the stud flange started, while for the GWB sheathings-to-
frame connections the bearing of the GWB panels begun. At this
displacement level, the deformations of walls as well as the local
deformation of the sheathing-to-frame connections were not evident,
as shown in Figure 4.15.
x Step 2: maximum shear resistance (lateral displacement equal to
36mm). At this load step the tilting of the screws in the OSB
connections, as well as the bearing in the GWB panels were evident, as
shown in Figures 4.16 b and c, respectively. Also the global
deformation of walls was clearly observable, as shown in Figure 4.16a.
x Step 3: lateral displacement equal to 80mm. At this displacement level,
in both the OSB and GWB-to-frame connections the screw heads
initiated to pull through the sheathings, as shown in Figures 4.17 c and
d. At this point, due to the rotation of the sheathings, the upper sides of
146 Chap ter IV

the GWB panels knocked against the joist, as shown in Figure 4.17b.
Probably this phenomenon produced a residual shear resistance
(constant shear force for displacement varying from 70 to 110mm).
x Step 4: lateral displacement equal to 130mm. At this step, the screw
heads had completely pulled through the sheathings, as shown in
Figures 4.18 c and d. As a consequence, the sheathings were
completely unzipped along the panel edges, as shown in Figure 4.18b.

(a) Deformation of the Wall

(b) Deformation of the OSB connections (c) Deformation of the GWB connections

Figure 4.15: Specimen condition at Step 1(d=10mm).

For all displacement levels, the evolution of the deformation of the


specimen was coherent with the sheathing-to-wall framing connections
failure. In fact, wall framing deformed into a parallelogram and the sheathings
had rigid body rotation, as shown in Figures 4.15a, 4.16a, 4.17a and 4.18a.
Moreover, any buckling phenomenon was not observed for the studs as well
as deformations of the OSB sheathing-to-floor framing connections were not
observed during the testing. On the contrary, for floor tracks, a local buckling
phenomenon occurred for lateral displacement larger than about 30mm.
Evaluatio n of seismic capacity: th e mo notonic test 147

Finally, both the shear and the tension anchors did not suffer any type of
failure.

(a) Deformation of the wall

(b) Deformation of the OSB connections

(c) Deformation of the GWB connections

Figure 4.16: Specimen condition at Step 2 (maximum shear resistance - d=36mm).


148 Chap ter IV

(a) Deformation of the wall (b) GWB panel-to-joist contact

(c) Deformation of the OSB connections

(d) Deformation of the GWB connections

Figure 4.17: Specimen condition at Step 3 (d=80mm).


Evaluatio n of seismic capacity: th e mo notonic test 149

(a) Deformation of the wall (b) Unzipping

(c) Deformation of the OSB connections

(d) Deformation of the GWB connections

Figure 4.18: Specimen condition at Step 4 (d=130mm).


150 Chap ter IV

The Figure 4.19 shows the force (V) vs. displacement (d) response curves
for the two tested walls. In particular, force vs. displacement curves obtained
from measures of the actuators a1 and a2 have been reported for the walls 1
and 2, respectively. From the comparison of lateral responses of two walls
some interesting observations are possible:
x the walls had the same behavior for displacements less than about
30mm (displacement for which the applied load approached the
maximum shear resistance), while they exhibited different response for
larger displacements;
x the maximum shear resistances were 47 and 44kN for wall 1 and 2,
respectively; therefore, the wall 1 was more resistant than wall 2 of
about 7%.

The measure of all horizontal LVDTs placed on the specimen in terms of


force vs. horizontal displacement relationships are illustrated in Figures 4.20 a
and b. In particular, the Figure 4.20a shows the force measured from the
actuator a1 as a function of the displacements recorded by the LVDTs located
on the side of the wall 1 (LVDTs a1; d1; w1,1; w1,2; w1,3 and f1), while the
force measured from the actuator a2 as a function of the displacements
recorded by the LVDTs located on side of the wall 2 (LVDTs a2; d2; w2,1;
w2,2; w2,3; f2) are shown in Figure 4.20b. From the examination of these
Figures, for both walls it is possible to observe that:
x the anchorage between the beam foundations of the specimen and the
laboratory floor was full effective; in fact the displacements measured
by LVDTs f1 and f2 were slight (displacements less than 0.2mm);
x the load transfer system between the load actuators and the floor of the
specimen was effective; in fact the difference between the
displacements recorded by actuators (a1 and a2) and those measured
by the LVDTs placed on the floor (d1 and d2) were slight
(displacements less than 0.9mm);
x the load transfer system between the floor and the walls was effective;
in fact, the floor-to-walls sliding was small; in particular, the
difference between the displacements measured by the LVDTs placed
on the floor (d1 and d2) and those recorded by the LVDTs located on
the top side of the walls (w1,1 and w2,1) were less than 4mm.
Evaluatio n of seismic capacity: th e mo notonic test 151

Figures 4.21 a and b show the forces measured from the actuators (a1 and
a2 for Figures 4.21a and b, respectively) as a function of the vertical
displacements recorded by the LVDTs located at the bottom side of the walls
(w1,4; w1,5 and w2,4; w2,5 for Figures 4.21a and b, respectively). From the
examination of these curves, for both walls it is possible to observe that the
vertical displacements measured on the tension side were larger than those
recorded on the compression side. In particular, this behavior was more
evident for displacements less than those corresponding to the maximum shear
resistance.

Figure 4.19: Shear (V) vs. displacement (d) curves for wall 1 and wall 2.

(a) wall 1 (b) wall 2

Figure 4.20: Shear (V) vs. displacement (d) measured by horizontal LVDTs.
152 Chap ter IV

(a) wall 1
(b) wall 2

Figure 4.21: Shear (V) vs. displacement (d) measured by vertical LVDTs.

4.5 REFERENCES

EN 300 (1997) Oriented Strand Boards (OSB) - Definitions, Classification and


Specifications. CEN (European Committee for Standardization). Bruxelles.
ENV 1991-1 (1996) Eurocode 1: Basis of Design and Actions on Structures Part 1:
Basis of Design. CEN (European Committee for Standardization). Bruxelles.
EN 10142 (2002) Continuously hot-dip zinc coated low carbon steel sheet and strip
for cold forming. Technical delivery conditions. CEN (European Committee for
Standardization). Bruxelles.
Gad, E.F., Duffield, C.F., Hutchinson, G.L., Mansell, D.S., Stark, G. (1999) Lateral
performance of cold-formed steel-framed domestic structures. Engineering
Structures, Elsevier, Vol.21, No.1: 83-95.
ISO (1980) Gypsum plasterboard Specification. ISO (International Organization for
Standardization). Geneva.
NASFA (2000) Prescriptive Method For Residential Cold-Formed Steel Framing
(Year 2000 Edition). NASFA (North American Steel Framing Alliance). Lexington,
KY, USA.
prEN 1998-1 (2001) Eurocode 8: Design of structures for earthquake resistance
Part 1: General rules Seismic actions and rules for buildings. CEN (European
Committee for Standardization). Bruxelles.
153

Chapter V
Evaluation of seismic demands

An experiment can provide only information on capacities, but, because of


the strong interrelation between capacity and demand, due consideration must
be given to seismic demand issues. This requires some significant problems to
be preliminarily solved:
x the development of reliable mathematical models of the hysteresis
behaviour of panel shear walls typical of light-gauge cold-formed
framed constructions;
x the assessment of deformation demands under a sufficiently large
database of earthquake ground motions is needed.

The need to carry out this type of statistical-numerical analysis derives


from the peculiar hysteresis behaviour of cold-formed steel stud shear wall
(CFSSSW) systems, which are characterized by strong pinching and reduced
ductility. In fact, commonly performed studies on the evaluation of seismic
demand have been derived on the basis of a bilinear hysteresis assumption. In
this Chapter, the calibration of a mathematical model, able to well interpreting
the hysteretic behavior of CFSSSWss response, is illustrated in Section 5.1.
The evaluation of the seismic demand using the calibrated mathematical
model and some accelerograms of Central Italian earthquakes is presented the
Section 5.2. Finally, because the applicability of the standard load histories to
CFSSSW systems need to be verified, Section 5.3 is dedicated to define a
deformation history for a cyclic testing.
154 Chap ter V

5.1 MODEL OF THE HYSTERETIC BEHAVIOR OF CFSSSW


SYSTEMS

On the basis of experimental evidence of existing tests on CFSSSW


systems, a mathematical model, able to well interpret several behavioral
aspects of their response, has been used (Della Corte et al. 1999). In
particular, the proposed model is able to account for the following aspects of
the mechanical cyclic behavior:
x non linearity;
x pinching.

In the following, the mathematical formulation of the used model is firstly


described. Then, model parameters are calibrated for the simulation of some
available experimental tests. The comparison between numerical and
experimental results shows the reliability of the model in capturing the stable
part of the hysteretic behavior. A this stage of the research the model is unable
to describe the unstable part of the cyclic response.
For describing the shear behavior of CFSSSWs, the relationship between
shear force (V) and lateral displacement (G) is used.
It is appropriate to preliminarily state some fundamental concepts,
extensively used in the following. The basic parameter for the description of a
generic loading history is the deformation excursion, which is constituted by a
loading branch and by the subsequent unloading branch of the V-G path. The
deformation range of excursion ('Gp) is the deformation range between the
beginning and the peak deformation of the excursion (ATC 1992). As it will
be explained herein after, the beginning of a loading branch and the end of an
unloading branch, always stand on a straight line passing through the origin
with a slope equal to the one concerned with the hardening of the system, as
shown in Figure 5.1.
The adopted mathematical function for describing the shear behavior of
CFSSSWs is introduced in two steps. In fact, the mathematical formulation is
firstly referred to systems without pinching effect. Then, pinching phenomena
typically present in the lateral behavior of these structures are introduced.
Ev alu atio n of seismic d eman ds 155

Figure 5.1: Basic definitions (Della Corte et al. 1999).

5.1.1 The loading branch without pinching


For the description of the loading branch in absence of pinching, a
mathematical formulation proposed by Richard & Abbott (1975) has been
adopted. This formulation, expressed in terms of shear force (V) versus lateral
displacement (G) relationship, can be written as follows:
V
k0  k h G  kh G (5.1)
1
n n
k  k G
1  0 h

V0
where:
k0 : is the initial stiffness of the system;
kh : is the slope of the straight-line asymptote of the V-G curve
(hardening line);
V0: is the intersection between V axis and hardening line;
n: is a shape parameter, which regulates the sharpness of transition
from the elastic to the fully plastic behavior (increasing values of n
correspond to an increasing sharpness).
156 Chap ter V

Figure 5.2 illustrates the graphical representation of the Richard-Abbott


mathematical formulation, also reporting the meaning of adopted symbols.
Besides to the parameters previously specified, the conventional yield and
ultimate limits need to be specified. In particular, the conventional yield
strength is defined as the intersection between the straight-line with a slope
equal to k0 going through the origin and the hardening line. The conventional
ultimate strength is defined as the point corresponding to the maximum value
of the shear force (limit of the stable part of the hysteretic behavior).
Consequently it is possible to define the shear force (Vy) and displacement (Gy)
corresponding to the conventional yield point and the ultimate force (Vu) and
displacement (Gu) corresponding to the conventional ultimate point, as shown
in Figure 5.2.

Numerical upper bound curve

Experimental response

Figure 5.2: The loading branch without pinching.

5.1.2 The loading branch with pinching


For describing pinching, two limit curves are introduced, representing an
upper bound and a lower bound to possible V-G values, respectively. Both
curves have a Richard-Abbot type law, with the following parameters:
k0, V0, kh, n for the upper bound curve;
Ev alu atio n of seismic d eman ds 157

k0p, V0p, khp, np for the lower bound curve.

A point (V, G) of the real path is considered to belong also to a Richard-


Abbott type curve, whose, parameters are defined as follows:
k 0t k 0 p  k 0  k 0 p t (5.2)
V0t V0 p  V0  V0 p t (5.3)
k ht k hp  k h  k hp t (5.4)
n pt n p  n  n p t (5.5)
where the parameter t, ranging in [0, 1], defines the transition law from the
lower bound to the upper bound curve. Its mathematical formulation is to be
defined in such a way to reproduce the shape of the curve as experimentally
observed. In the proposed model, in order to describe a pinching-type
behavior, parameter t is defined by the following relationship:
t2
G / G lim t1
t t1 (5.6)
G / G lim  1
in which, t1, t2 and Glim are three additional parameters to be defined on the
basis of experimental data. Figure 5.3 illustrates, qualitatively, the description
of pinching with reference to one single excursion from the origin.

Figure 5.3: The loading branch without pinching (Della Corte et al. 1999).
158 Chap ter V

In case of a generic deformation history, the parameter Glim must be related


to the maximum experienced deformation in the direction of the loading
branch being described. Therefore, it could be evaluated through the
parameter O, defined by the following relationship:
G lim O G 0  G max (5.7)

where:
G0 : is the absolute value of the deformation corresponding to the
starting point of the current excursion;
Gmax: is the maximum absolute value of deformation experienced, in all
previous loading history, in the direction of loading branch to be
described (Figure 5.4).

5.1.3 The unloading branch


The unloading-branch is assumed to be linear with a slope equal to the
initial stiffness k0 up to the intersection with the straight line obtained drawing
the parallel to the hardening line going through the origin, as showed in Figure
5.5. This allows the Bauschinger effect to be considered.

Figure 5.4: The loading branch without pinching (Della Corte et al. 1999).
Ev alu atio n of seismic d eman ds 159

Figure 5.5: The unloading branch (Della Corte et al. 1999).

5.1.4 Calibration of the model


In order to investigate on the capability of the illustrated model for
interpreting the stable part of the hysteretic behavior of CFSSSWs, the values
of upper bound curve (k0, V0, kh, n), lower bound curve (k0p, V0p, khp, np) and
transition (t1, t2, O) parameters need to be evaluated. Besides to the
experimental results obtained with the monotonic tests carried out in this
research, other existing experimental cyclic tests have been considered. In
particular, monotonic and cyclic test results have been used to evaluate upper
bound curve and transition parameters, respectively. While, the results of both
monotonic and cyclic tests have been used to estimate the lower bound curve
parameters.

All existing selected experimental cyclic tests were tests performed on


CFSSSWs similar to those tested in the current research. In fact, the
considered walls were sheathed with oriented strand board (OSB) panels and
had panel-to-wall framing connections spaced at about 150mm at the
perimeter and at about 300mm in the field. In particular, the selected tests are
listed hereafter:
x S+96ab-1(OSB1): AISI OSB1 Serrette et al. (1996a,b);
x S+96ab-1(OSB2): AISI OSB2 Serrette et al. (1996a,b);
x S+97b-18(E1): AISI E1 Serrette et al. (1997b);
x S+97b-18(E2): AISI E2 Serrette et al. (1997b);
x C01-4(A): 17A - COLAUCI (2001).
160 Chap ter V

A summary of the geometry and material data of specimens is given in


Table 5.1.
All tests have been performed with fully reversing cyclic displacements
following the TCCMAR (Technical Coordinating Committee for Masonry
Research) sequentially phased displacement procedure, suggested by the
Structural Engineers Association of South California (SEAOSC 1997).
The TCCMAR procedure uses the concept of the first major event (FME),
which is defined as the first significant limit state that occurs during the test.
In case of CFSSSWs the FME is defined as the yield limit state (YLS). The
FME can be determined from preliminary load tests on an identical test
specimen.

S+96ab-1 S+97b-18
C01-4 (A)
(OSB1) and (OSB2) (E1) and (E2)
Type of loading Cyclic Cyclic Cyclic
Wall size (1) 2440 x 1220 2440 x 610 2440 x 2440
Aspect ratio (2) 2.00 4.00 1.00
Openings No openings No openings No openings
Steel grade (3) A653 Grade SQ 33 A653 Grade SQ 33 A446 Grade A
Studs (4) C 89x41x10x0.84/610 C 89x43x13x0.84/610 C 89x41x10x0.84/610
Tracks (5) U 89x32x0.84 U 89x32x0.84 U 89x38x0.84
Frame screws (6) 4.2x13 wafer-head 4.2x13 mod. truss head 4.2x13 mod. truss head
Sheathings (7) 11.1mm tick OSB 11.1mm tick OSB 11.1mm tick OSB
Sheathing fasteners (6) 4.2x25 flat head 4.2x25 flat head 4.2x25 bugle head
Spacing sheathing
152/305 152/305 152/305
fasteners (8)
Type of anchorage Tie hold-down Hold-down Strip hold-down
(1) height x length (mm)
(2) height / length
(3) A653 Grade SQ 33: Yield strength = 228 MPa, Ultimate tensile strength = 311 MPa
(3) A446 Grade A: Yield strength = 228 MPa, Ultimate tensile strength = 311 MPa
(4) C: lipped channel section web depth x flange size x lip size x thickness / spacing (mm)
(4) Double back-to-back coupled studs were employed at ends of walls
(5) U: unlipped channel section web depth x flange size x thickness (mm)
(6) diameter x length (mm)
(7) 610 x 2440 mm panel size
(8) perimeter/field (mm)

Table 5.1: Wall geometry and material data.

The TCCMAR procedure consists of applying three cycles of fully


reversing, displacement-controlled load at each wall displacement increment
representing 25%, 50%, and 75% of the FME displacement. Then, the wall
displacement is increased for one load cycle to 100% of the FME
displacement. Next, cycles of displacement for one cycle each at 75%, 50%,
Ev alu atio n of seismic d eman ds 161

and 25% of the maximum displacement (100% of the FME displacement) are
applied. This is followed by three cycles of displacement at maximum
displacement (100% of the FME displacement) to stabilize the load-
displacement response of the wall. Then, the next increment of increased
displacement (125% of the FME displacement) is applied, followed by similar
stabilization cycles of loading. This incremental cyclic load-displacement
sequence is continued to 150%, 175%, 200%, 250%, 300% of the FME
displacement, or until the wall exhibits greatly diminished shear load capacity.
The test protocols used for the selected tests are illustrated in Table 5.2 and
Figure 5.6. Figure 5.7 shows the results of the selected tests in terms of shear
force (V) versus lateral displacement (G) curves.
The results of the calibration of the model in terms of lower bound curve
and transition parameters obtained on the basis of Serrette and COLA-UCI
tests are shown in Table 5.3 for a wall with a length of 1m (wall with unit
length). Figure 5.8 shows the comparison between the experimental response
and the numerical one for the C01-4(A) specimen. In particular, in Table 5.3
are reported the values of transition parameters (t1, t2, O) together to values of
the ratios between the lower and upper bound curve parameters ((k0p/k0)c,
(V0p/V0)c, (khp/kh)c, (np/n)c).

The comparison between the experimental monotonic response and the


numerical one are illustrated in Figure 5.9.
The results of the calibration of the model in terms of upper and lower
bound curve parameters obtained on the basis of the monotonic test are shown
in Table 5.4. The values of the lower bound curve parameters have been
calculated starting from the ratios between the lower and upper bound curve
parameters ((k0p/k0)c, (V0p/V0)c, (khp/kh)c, (np/n)c), which have been previously
derived from existing cyclic tests and the values of upper bound curve
parameters calibrated on monotonic test results (k0, V0, kh, n). In particular, the
following assumptions have been used:
k 0 p k 0 p / k 0 c k 0 (5.8)
V0 p V / V V
0p 0 c 0 (5.9)
k hp k / k k
hp h c h (5.10)
np n / n n
p c (5.11)
162 Chap ter V

S+96ab-1 S+97b-18
C01-4 (A)
(OSB1) and (OSB2) (E1) and (E2)
No. of cycles Displ. (mm) No. of cycles Displ. (mm) No. of cycles Displ. (mm)
3 5.1 3 5.1 3 5.7
3 10.2 3 10.2 3 11.5
3 15.2 3 15.2 3 17.2
1 20.3 3 20.3 1 22.9
1 15.2 1 25.4 1 17.2
1 10.2 1 19.1 1 11.5
1 5.1 1 12.7 1 5.7
3 20.3 1 6.4 3 22.9
1 25.4 3 25.4 1 28.6
1 19.1 1 30.5 1 21.5
1 12.7 1 22.9 1 14.3
1 6.4 1 15.2 1 7.2
3 25.4 1 7.6 3 28.6
1 30.5 3 30.5 1 34.4
1 22.9 1 40.6 1 25.8
1 15.2 1 30.5 1 17.2
1 7.6 1 20.3 1 8.6
3 30.5 1 10.2 3 34.4
1 40.6 3 40.6 1 40.1
1 30.5 1 50.8 1 30.1
1 20.3 1 38.1 1 20.0
1 10.2 1 25.4 1 10.0
3 40.6 1 12.7 3 40.1
1 50.8 3 50.8 1 45.8
1 38.1 1 61.0 1 34.4
1 25.4 1 45.7 1 22.9
1 12.7 1 30.48 1 11.5
3 50.8 1 15.24 3 45.8
1 61.0 3 61.0 1 57.3
1 45.7 1 71.1 1 42.9
1 30.5 1 53.3 1 28.6
1 15.2 1 35.6 1 14.3
3 61.0 1 17.8 3 57.3
1 71.1 3 71.1 1 68.7
1 53.3 1 51.5
1 35.6 1 34.4
1 17.8 1 17.2
3 71.1 3 68.7
1 80.2
1 60.1
1 40.1
1 20.0
3 80.2
1 91.6
1 68.7
1 45.8
1 22.9
3 91.6

Table 5.2: Cyclic test protocols.


Ev alu atio n of seismic d eman ds 163

100
displacement Cyclic frequency = 0.67Hz
80 (mm)

60

40

20

0
0 5 10 15 20 25 30 35 40 45 50 55
-20

-40

-60
S+96ab-1
-80 (OSB1) and (OSB2)
-100 time (s)

100
displacement Cyclic frequency = 1.00Hz
80 (mm)

60

40

20

0
0 5 10 15 20 25 30 35 40 45 50 55
-20

-40

-60
S+97b-18
-80 (E1) and (E2)
-100 time (s)

100
displacement
Cyclic frequency =
80 (mm)
0.25Hz for dispacements 40.6mm
60 0.50Hz for dispacements > 40.6mm

40

20

0
0 5 10 15 20 25 30 35 40 45 50 55
-20

-40

-60

-80 C01-4 (A)


-100 time (s)

Figure 5.6: Cyclic test protocols.


164 Chap ter V

S+96ab-1 (OSB1) S+96ab-1 (OSB2)

S+97b-18 (E1) S+97b-18 (E2)

Displ. at average
C01-4 (A) Average of maximum
Test of maximum +/-
+/- load
load
S+96ab-1 (OSB1) 685 lb/ft. (10.0 kN/m) 1.8 in. (46 mm)
S+96ab-1 (OSB2) 758 lb/ft. (11.1 kN/m) 1.8 in. (46 mm)
S+97b-18 (E1) 700 lb/ft. (10.2 kN/m) 2.8 in. (71 mm)
S+97b-18 (E2) 700 lb/ft. (10.2 kN/m) 2.8 in. (71 mm)
C01-4 (A) 773 lb/ft. (11.3 kN/m) 1.5 in. (38 mm)

Figure 5.7: Cyclic response.

Experimental response
Numerical response

Figure 5.8: Numerical vs. experimental cyclic response for the C01-4(A) specimen.
Ev alu atio n of seismic d eman ds 165

S+96ab-1 S+96ab-1 S+97b-18 S+97b-18 Assumed Dev.


Parameter C01-4(A) Mean
(OSB1) (OSB2) (E1) (E2) values St.
(k0p / k0)c 0.92 0.99 1.09 1.08 0.96 1.00 1.00 0.61
(V0p / V0)c 0.00 0.01 0.01 0.01 0.01 0.01 0.01 0.05
(khp / kh)c 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
(np / n)c 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00
t1 [-] 8 10 10 10 12 10 10 1.4
t2 [-] 0.50 0.40 0.50 0.60 0.50 0.50 0.50 0.07
O [-] 1.20 1.00 0.50 0.50 0.50 0.75 0.75 0.34

Table 5.3: Results of calibration of lower bound curve and transition parameters obtained on
the basis of existing cyclic tests (values for a 1m long wall).

Parameter Monotonic test Assumed values


k0 [kN/mm] 2.80 2.80
V0 [kN] 16 16
kh [kN/mm] 0.11 0.11
n [-] 1.2 1.2
k0p [kN/mm] 1.00 x 2.80 = 2.80
V0p [kN] 0.01 x 16 = 0.16
khp [kN/mm] 0.00 x 0.11 = 0.00
np [-] 1.00 x 1.2 = 1.2
Vy [kN] 17 17
Gy [mm] 6.0 6.0
Vu [kN] 19 19
Gu [mm] 36 36

Table 5.4: Results of calibration of upper and lower bound curve parameters obtained on the
basis of monotonic tests carried-out in the current research (values for a 1m long wall).

25
V (KN/m) "Numerical upper bound curve"
"Experimental monotonic response"
Serie3
20
Vu
Vy
15

10

G (mm)
0
Gy Gu
0 5 10 15 20 25 30 35 40

Figure 5.9: Numerical vs. experimental monotonic response.


166 Chap ter V

5.2 ASSESSMENT OF DEFORMATION DEMAND

One single test of a seismic-resistant component or sub-assembly must


represent a compromise between the need to correctly characterize the specific
structural component subject to a specific deformation history and the
requirement to obtain general conclusions applicable to the same component
with different loading regimes. The only way to fully achieve this compromise
is to analytically study the probable deformation histories the component
would be subjected to. This characterization requires at least the evaluation of
the following demand parameters:
x maximum level of deformation experienced by the structure;
x number of repetition of several discrete values of ductility demands;
x cumulative plastic deformation engagement of the structure;
x distribution of plastic deformation ranges.

It is well known that the loading histories produced by different ground


motions on a given structure are significantly scattered, consequently, the
estimation of the deformation demand will be carried out under a sufficiently
large data-base of earthquake ground motions.

5.2.1 Data-base of considered earthquake ground motions


The selected earthquakes include 26 far field records from Central Italy.
The considered geographic region (zone in which are located the recording
stations) is identified as medium-high seismic zone by the Italian seismic
design code (Ordinanza PCM 2003). In fact, the value of the design peak
ground acceleration of the seismic zone (PGAzone), having a 10% probability
of exceedance in 50 years, is 0.25g. Consequently, a value of the design peak
ground acceleration (PGAdes) equal to 0.25g it has been adopted.
The considered earthquakes are the Lazio-Abruzzo earthquake, occurred
on March, 7th 1984 at 17h49m; the Umbro-Marchigiano earthquake,
occurred on September 1997, 26th at 00h33m; and the Umbro-Marchigiano
earthquake, occurred on September 1997, 26th at 09h40m. For these
earthquakes the Richter magnitude ranges from 5.6 to 5.8, the focal depth
ranges from 6 to 8km, and the source mechanism is of normal type.
Ev alu atio n of seismic d eman ds 167

The earthquake records have been chosen to cover, whenever possible, all
the soil types classified by Eurocode 8 (prEN 1998-1 2001). In particular, for
each soil type, three accelerograms have been selected in such a way that the
shape of the average elastic response spectrum of these records is close to the
shape of the Eurocode 8 Type 1 elastic spectrum acceleration (with a damping
ratio of 0.05). For the accelerograms selected according to these assumptions
the recorded peak ground acceleration (PGArec) value range from 0.01 to
0.76g, the distance between the epicentre and the recording station ranges
from 11 to 100km, and the record length ranges from 14 to 60s.
Earthquakes, recording stations and waveform parameters of considered
records are reported in Table 5.5. For each station, Table 5.5 also shows the
design ground acceleration according to the Italian seismic design code
(PGAzone). In the same Table, a label has been defined for each earthquake
record. As example, the label UM-B-1 refers to a recording of the Umbro-
Marchigiano (UM), which has been carried out on B soil type (B).

The geographical localization of the epicenter of the selected strong


motions and of the recording stations is presented in Figure 5.10.
In Figures 5.11 a, b, c, d, and e comparisons between the elastic response
spectra (Sae(T,Q=0.05)) of the selected records scaled to a design value of the
ground acceleration equal to its design peak ground value
(PGA=PGAdes=0.25g) and the design elastic spectral acceleration
(Saed(T,Q=0.05)) defined by Eurocode 8 for soil type A, B, C, D, and E,
respectively, are reported.
168 Chap ter V

Earthquake name
Date and time Record Epicentral Fault
Record
Richter Magnitude Station name (orientation) PGAzone / g PGArec / g length distance distance Soil type
label
Focal depth [s] [km] [km]
Fault mechanism
LA-A-01 Ponte Corvo(N-S) 0.25 0.064 32.441 31 -
Rock
LA-A-02 Roccamonfina (N-S) 0.25 0.036 22.740 50 - (A)
LA-A-03 Bussi (E-W) 0.25 0.019 25.000 51 -

LA-B-01 Ripa-Fagnano (E-W) 0.25 0.017 14.910 64 -


Stiff soil
LA-B-02 Poggio-Picenze (E-W) 0.25 0.011 15.530 72 - (B)
LA-B-03 Poggio-Picenze (N-S) 0.25 0.017 15.530 72 -
Lazio Abruzzo
1984/05/07 - 17:49:43 LA-C-01 Ortucchio (N-S) 0.35 0.061 21.220 26 -
5.7 Soft soil
LA-C-02 Taranta Peligna (E-W) 0.35 0.077 32.131 39 - (C)
8 km
Normal LA-C-03 Barisciano (E-W) 0.25 0.012 16.540 71 -
Garigliano-Centrale Nucleare 1
LA-D-01 0.25 0.060 26.211 53 -
(N-S)
Very soft
Garigliano-Centrale Nucleare 1 soil
LA-D-02 0.25 0.059 26.231 53 -
(E-W) (D)
Garigliano-Centrale Nucleare 2
LA-D-03 0.25 0.060 26.101 53 -
(N-S)
LA-E-01 Cassino-Sant' Elia (N-S) 0.25 0.147 27.261 23 17 Alluvium
(E)
LA-E-02 Cassino-Sant' Elia (E-W) 0.25 0.114 27.271 23 17
Umbro-Marchigiano UM-A-01 Cagli (N-S) 0.25 0.013 25.240 59 50
1997/09/26 - 09:40:30 Rock
5.8 UM-A-02 Nocera Umbra (E-W) 0.25 0.760 41.129 11 4
(A)
6 km
Normal UM-A-03 Pennabilli (N-S) 0.25 0.015 28.731 100 92
Umbro-Marchigiano
1997/09/26 - 00:33:16
5.6 UM-B-01 Bevagna (N-S) 0.25 0.034 46.108 25 25
7 km Stiff soil
Normal (B)
UM-B-02 Bevagna (E-W) 0.25 0.079 50.308 23 26

UM-B-03 Senigallia (E-W) 0.25 0.037 26.801 78 71


Umbro-Marchigiano
1997/09/26 - 09:40:30 UM-C-01 Castelnuovo-Assisi (N-S) 0.25 0.159 55.077 22 23
5.8 Soft soil
6 km UM-C-02 Rieti (N-S) 0.25 0.015 59.786 67 66 (C)
Normal
UM-C-03 Rieti (E-W) 0.25 0.018 59.786 67 66

UM-E-01 Norcia-Altavilla (E-W) 0.25 0.046 13.760 32 30


Umbro-Marchigiano
1997/09/26 - 00:33:16 UM-E-02 Norcia-Zona Industriale (N-S) 0.25 0.034 52.197 32 26 Alluvium
5.6 (E)
7 km UM-E-03 Norcia-Zona Industriale (E-W) 0.25 0.037 52.197 32 26
Normal

Table 5.5: Data-base of selected earthquake records.


Ev alu atio n of seismic d eman ds 169

Umbro - Marchigiano

PGA (g) values with a


10% probability of
exceedance in 50 years

Lazio - Abruzzo

Q Epicenter
S Station

Figure 5.10: Geographical distribution of the selected strong-motion records.

1.4 Sae(T;Q=0.05)/g 1.4 Sae(T;Q=0.05)/g


Lazio-Abruzzo 1984 EC8 Umbria-Marche 1999 EC8
1.2
Soil type: A LA-A-01 1.2 Soil type: A UM-A-01
LA-A-02 UM-A-02
1.0 LA-A-03 1.0 UM-A-03
mean mean
0.8 0.8
PGA=PGAdes=0.25g PGA=PGAdes=0.25g
0.6 0.6

0.4 0.4

0.2 0.2

0.0 T (s) 0.0


T (s)
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0

Figure 5.11a: Elastic spectra of earthquake records for type A soil.


170 Chap ter V

1.4 Sae(T;Q=0.05)/g 1.4 Sae(T;Q=0.05)/g


Lazio-Abruzzo 1984 EC8 Umbria-Marche 1999 EC8
1.2 Soil type: B LA-B-01 1.2 Soil type: B UM-B-01
LA-B-02 UM-B-02
1.0 LA-B-03 1.0 UM-B-03
mean mean
0.8 0.8
PGA=PGAdes=0.25g PGA=PGAdes=0.25g
0.6 0.6

0.4 0.4

0.2 0.2

0.0 T (s) 0.0


T (s)
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0

Figure 5.11b: Elastic spectra of earthquake records for type B soil.

3.2 Sae(T;Q=0.05)/g 3.2 Sae(T;Q=0.05)/g


Lazio-Abruzzo 1984 EC8 Umbria-Marche 1999 EC8
2.8 Soil type: C LA-C-01 2.8 Soil type: C UM-C-01
LA-C-02 UM-C-02
2.4 LA-C-03 2.4 UM-C-03
mean mean
2.0 2.0

1.6
PGA=PGAdes=0.25g 1.6
PGA=PGAdes=0.25g

1.2 1.2

0.8 0.8

0.4 0.4

0.0 0.0
T (s) T (s)
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0

Figure 5.11c: Elastic spectra of earthquake records for type C soil.

1.4 Sae(T;Q=0.05)/g
Lazio-Abruzzo 1984 EC8
1.2 Soil type: D LA-D-01
LA-D-02
1.0 LA-D-03
mean
0.8
PGA=PGAdes=0.25g
0.6

0.4

0.2

0.0
T (s)
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0

Figure 5.11d: Elastic spectra of earthquake records for type D soil.


Ev alu atio n of seismic d eman ds 171

1.4 Sae(T;Q=0.05)/g 1.4 Sae(T;Q=0.05)/g


Lazio-Abruzzo 1984 EC8 Umbria-Marche 1999 EC8
1.2 Soil type: E LA-E-01 1.2 Soil type: E mean
LA-E-02 UM-E-01
1.0 mean 1.0 UM-E-02
UM-E-03
0.8 0.8
PGA=PGAdes=0.25g PGA=PGAdes=0.25g
0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
T (s) T (s)
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0

Figure 5.11e: Elastic spectra of earthquake records for type E soil.

5.2.2 Displacement demand evaluation


In order to perform the analysis a purposely developed computer code has
been used (Ghersi & Noce 1999). This program allows non-linear dynamic
analysis of two-dimensional structures to be carried out.
Figure 5.12 shows the representation of the numerical model adopted to
schematize the CFSSSW with unit length as a SDOF system.

F Richard Abbott element v(G)


v
M G

h = 2800 mm Truss
element
EA
a
Ground acceleration a(t)
t

Figure 5.12: Numeric model schematization.

In the model, the hysteretic lateral behavior of the CFSSSW is described by


a Richard Abbott element in which the assumed values of upper bound
172 Chap ter V

curve, lower bound curve and transition parameters are the results of the
calibration explained in Section 5.1.4 (see Tables 5.3 and 5.4). According to
basic assumptions illustrated in Chapter 4 (Section 4.2), the assumed value of
mass (M) is equal to M=1250kg. In order to take in to account the second
order effects, a vertical load (F) corresponding to 100% of the mass value has
been considered. A damping ratio of 0.05 has been also used in the model.
For performing the earthquake analyses, the incremental dynamic analysis
(IDA) or dynamic pushover (DPO) procedure is utilised.
The IDA procedure needs the definition of two parameters. The first is
related to the structural performance and can be linked to the damage level of
the structure after an earthquake. The second is a parameter associated to the
magnitude of the earthquake records. In particular, the most used structural
performance and record intensity parameters are the following (Fulop &
Dubina 2003):
x Structural Performance Parameters (SPP):
x inter-story drift (G/h);
x top story displacement;
x maximum plastic rotation;
x accumulated plastic rotation.
x Record Intensity Parameters (RIP):
x elastic spectral acceleration corresponding to the first mode period
(T0) of the structure (Sae(T0));
x recorded peak ground acceleration.

Fixed a structural system and the related SPP, an earthquake record and the
associated RIP, the IDA procedure consists in determining IRP value
corresponding to each prefixed SPP value. As a result of the IDA, a curve
relating the SPP to the RIP can be drawn.
The IDA parameters used in the current research are the inter-story drift
(G/h) and the elastic spectral acceleration of the SDOF system (Sae). In
particular the accelerograms are scaled from 0.05 to 1.95g.
An example of IDA is reported in Figure 5.13 where the significant steps of
the procedure are shown. This analysis has been carried out on the SDOF
Ev alu atio n of seismic d eman ds 173

system previously defined for which the inter-story drift (G/h) has been
adopted as SPP. In the example, the UM-B-1 accelerogram has been
considered as agent earthquake and the elastic spectral acceleration (Sae) has
been assumed as IRP.
The UM-B-1 record has been scaled form 0.05 to 1.95g. In particular,
Figure 5.13a shows the record scaled to design peck ground acceleration
PGAdes (PGA=0.25g) and Figure 5.13b illustrates the elastic response
acceleration spectra of the record for PGA values of 0.25, 0.95 and 1.45g.
Assuming as natural vibration period T=0.13s (obtained for a mass M=1250kg
and a stiffness k=k0=2.80kN/mm), for each prefixed PGA level it is possible
to identify the Sae value (Sae,0.05=0.63g, Sae,0.25=2.40g, Sae,0.45=3.66g) from
Figure 5.13b.
The non-linear dynamic response of the SDOF system, in terms of shear
force (V) versus inter-story drift (G/h), is reported in Figures 5.13c, d, and e for
each assumed PGA level. The maximum absolute value of G/h ((G/h)max) can
be extracted from the V-G/h curves ((G/h)max,0.25=0.001, (G/h)max,0.95=0.024,
(G/h)max,0.45=0.073). Reporting for each obtained value of the (G/h)max the
correspondent Sae value the IDA curve can be drawn, as shown in Figure
5.13f.

The IDA curves, obtained considering all selected earthquake records, are
reported in Figures 5.14 a and b, grouped for Lazio-Abruzzo and Umbro-
Marchigiano earthquakes, respectively. The same curves are shown in
Figures 5.15 a through e, grouped for the different soil types.
Reporting the inter-story drift corresponding to the yield limit
(Gy/h=6/2800=0.0021) and ultimate limit state (Gu/h=36/2800=0.013) on the
IDA curves, the associated elastic spectral accelerations Saey and Saeu can be
determined. The Saed, Saey and Saeu values and the Saeu/Saey, and Saeu/Saed ratios
computed for each earthquake records are also reported in Table 5.6.
Moreover, the mean and the standard deviation values have been calculated
for both the Lazio-Abruzzo (LA) and Umbro-Marchigiano (UM) record
groups, for all soil type groups and for all accelerograms.
174 Chap ter V

0.5 8.0
a (t)/g PGA=0.25g Sae/g (IRP) PGA=01.45g
0.4 7.0
PGA=0.95g
0.3
6.0
PGA=0.25g
0.2
5.0
0.1
4.0
0.0 Sae,1.45
0 10 20 30 40 50 60 3.0
-0.1
Sae,0.95
-0.2 2.0

-0.3 1.0
Sae,0.25
-0.4
0.0
t (s) T (s)
-0.5 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0

(a) Accelerogram of the UM-B-1 record (b) Elastic spectra of the UM-B-1 record
30 30
V [kN] PGA=0.25g V [kN] PGA=0.95g

20 20

10 10

(G/h)0.25 G/h (SPP) (G/h)0.95 G/h (SPP)


0 0
-0.04 -0.03 -0.02 -0.01 0.00 0.01 0.02 0.03 0.04 -0.04 -0.03 -0.02 -0.01 0.00 0.01 0.02 0.03 0.04

-10 -10

-20 -20

-30 -30

(c) V - d/h response for PGA = 0.25g (d) V - d/h response for PGA = 0.95g
30 4.0
V [kN] PGA=1.45g Sae/g (IRP)
3.5
20
3.0

10 2.5
(G/h)1.45
2.0
0
G/h (SPP)
-0.04 -0.03 -0.02 -0.01 0.00 0.01 0.02 0.03 0.04 1.5
-10
1.0

-20 0.5
G/h (SPP)
0.0
-30 0.000 0.020 0.040 0.060 0.080 0.100 0.120 0.140

(e) V - d/h response for PGA = 1.45g (f) IDA curve

Figure 5.13: IDA procedure example.

The Saeu/Saey ratio can be considered as a measure of seismic toughness of


the structures. For all earthquake records, this parameter ranges from 1.26 to
3.89 with a mean value of 2.27 and a standard deviation of 0.62. Based on
these results a discrete ductility for the CFSSSW systems can be relied upon.
The results are enough scattered, what highlights the dependence of the
Saeu/Saey ratio from the considered earthquake record.
Ev alu atio n of seismic d eman ds 175

From examination of the data reported in Table 5.6 it can also be observed
that the UM earthquake records are more severe than the LA group. In fact,
the mean value of Saeu/Saey is slightly higher (about 12%) in the LA records.
Moreover, the results indicate that the LA data are more scattered in
comparison with UM data.

The comparison between the results obtained for different soil types shows
that the accelerograms recorded on soil type D are the most severe. In fact, in
this case the lowest mean value of Saeu/Saey ((Saeu/Saey)mean=1.52) has been
obtained. The least severe results have been found for soil types A and B, for
which the mean values of Saeu/Saey are 2.17 and 2.13, respectively.
Examining the scattering of data, it may be noted that the maximum
dispersion has been obtained for the soil type C, while in the case of soil type
D the scattering of the results is the smallest one.

7
LA-A-01 LA-A-02 LA-A-03
Sae/g LA-B-01 LA-B-02 LA-B-03
6 LA-C-01 LA-C-02 LA-C-03
LA-D-01 LA-D-02 LA-D-03
LA-E-01 LA-E-02
5

G/h
0
0.000 Gy / h 0.005 0.010 Gu / h 0.015

Figure 5.14a: IDA curves for Lazio-Abruzzo earthquake.


176 Chap ter V

7
UM-A-01 UM-A-02 UM-A-03
Sae/g UM-B-01 UM-B-02 UM-B-03
6 UM-C-01 UM-C-02 UM-C-03
UM-E-01 UM-E-02 UM-E-03

G/h
0
0.000 Gy / h 0.005 0.010 Gu / h 0.015

Figure 5.14b: IDA curves for Umbro-Marchigiano earthquake.

7
Sae/g
6
LA-A-01
LA-A-02
5
LA-A-03
UM-A-01
4 UM-A-02
UM-A-03
3

1
G/h
0
0.000 Gy / h 0.005 0.010 Gu / h 0.015

Figure 5.15a: IDA curves for soil type A.


Ev alu atio n of seismic d eman ds 177

7
Sae/g
6
LA-B-01
LA-B-02
5
LA-B-03
UM-B-01
4 UM-B-02
UM-B-03
3

G/h
0
0.000 Gy / h 0.005 0.010 Gu / h 0.015

Figure 5.15b: IDA curves for soil type B.

7
LA-C-01
LA-C-02
6 LA-C-03
UM-C-01
UM-C-02
UM-C-03
5

1
G/h
0
0.000 Gy / h 0.005 0.010 Gu / h 0.015

Figure 5.15c: IDA curves for soil type C.


178 Chap ter V

7
Sae/g
6
LA-D-01
LA-D-02
5 LA-D-03

1
G/h
0
0.000 Gy / h 0.005 0.010 Gu / h 0.015

Figure 5.15d: IDA curves for soil type D.

7
Sae/g
6
LA-E-01
LA-E-02
5 UM-E-01
UM-E-02
4 UM-E-03

1
G/h
0
0.000 Gy / h 0.005 0.010 Gu / h 0.015

Figure 5.15e: IDA curves for soil type E.


Ev alu atio n of seismic d eman ds 179

Earthquake record Saed / g Saey / g Saeu / g Saeu / Saey Saeu / Saed


LA-A-01 0.63 1.03 2.06 2.00 3.30
LA-A-02 0.63 0.88 2.08 2.35 3.33
LA-A-03 0.63 1.01 2.46 2.44 3.93
LA-B-01 0.68 0.87 1.68 1.94 2.46
LA-B-02 0.68 1.38 2.53 1.83 3.70
LA-B-03 0.68 1.13 3.11 2.75 4.55
LA-C-01 0.72 1.79 6.17 3.44 8.62
LA-C-02 0.72 2.09 5.09 2.43 7.11
LA-C-03 0.72 4.36 16.99 3.89 23.73
LA-D-01 0.68 1.02 1.56 1.53 2.28
LA-D-02 0.68 1.08 1.53 1.42 2.24
LA-D-03 0.68 1.04 1.69 1.62 2.47
LA-E-01 0.87 1.02 3.10 3.05 3.56
LA-E-02 0.87 1.36 3.63 2.68 4.18
UM-A-01 0.63 1.15 3.29 2.86 5.22
UM-A-02 0.63 1.78 3.76 2.11 5.96
UM-A-03 0.63 0.87 1.10 1.26 1.75
UM-B-01 0.58 1.10 1.99 1.81 3.43
UM-B-02 0.58 1.27 2.29 1.81 3.95
UM-B-03 0.58 0.79 2.10 2.66 3.62
UM-C-01 0.62 1.81 4.20 2.33 6.78
UM-C-02 0.62 1.47 2.24 1.52 3.61
UM-C-03 0.62 1.02 2.20 2.16 3.54
UM-E-01 0.74 2.14 5.37 2.51 7.26
UM-E-02 0.74 1.46 3.44 2.35 4.64
UM-E-03 0.74 1.59 3.51 2.20 4.74
Mean (Type A soil records) 1.12 2.46 2.17 3.92
St. Dev. (Type A soil records) 0.34 0.95 0.54 1.50
Mean / St. Dev. (Type A soil records) 0.30 0.39 0.25 0.38
Mean (Type B soil records) 1.09 2.28 2.13 3.62
St. Dev. (Type B soil records) 0.23 0.49 0.45 0.69
Mean / St. Dev. (Type B soil records) 0.21 0.22 0.21 0.19
Mean (Type C soil records) 2.09 6.15 2.63 8.90
St. Dev. (Type C soil records) 1.17 5.54 0.88 7.54
Mean / St. Dev. (Type C soil records) 0.56 0.90 0.33 0.85
Mean (Type D soil records) 1.05 1.59 1.52 2.33
St. Dev. (Type D soil records) 0.03 0.08 0.10 0.12
Mean / St. Dev. (Type D soil records) 0.03 0.05 0.07 0.05
Mean (Type E soil records) 1.51 3.81 2.56 4.88
St. Dev. (Type E soil records) 0.41 0.89 0.33 1.41
Mean / St. Dev. (Type E soil records) 0.27 0.23 0.13 0.29
Mean (LA records) 1.43 3.83 2.38 5.39
St. Dev. (LA records) 0.91 4.03 0.73 5.59
Mean / St. Dev. (LA records) 0.64 1.05 0.31 1.04
Mean (UM records) 1.37 2.96 2.13 4.54
St. Dev. (UM records) 0.41 1.18 0.47 1.57
Mean / St. Dev. (UM records) 0.30 0.40 0.22 0.34
Mean (all records) 1.40 3.43 2.27 5.00
St. Dev. (all records) 0.71 3.04 0.62 4.18
Mean / St. Dev. (all records) 0.51 0.89 0.28 0.84

Table 5.6: IDA results.


180 Chap ter V

5.3 DEFINITION OF A LOAD HISTORY FOR A CYCLIC


TESTING

5.3.1 Loading histories in quasi-static cyclic loading tests: basic


procedures for Multiple Step Test
Seismic capacity measures for a structural component or sub-assembly are
strength, stiffness, inelastic deformation capacity and cumulative capacity
such as energy dissipation capacity. All these parameters are expected to
deteriorate as the number of damaging cycles and the amplitude of cycling
increases. The type of deterioration depends on the failure mode of the
structural element and the adopted loading history. Figure 5.16a shows a
typical example of deterioration for a CFSSSW subject to the loading history
reported in Figure 5.16b. The typical cyclic behavior of CFSSSWs denotes
strong pinching and reduced ductility.

(a) Cyclic response of a CFSSSW (b) Time history

Figure 5.16: Deterioration of a CFSSSW system (Branston et al. 2003).

The choice of a loading history for seismic testing of a structural


component or sub-assembly should be based on the following cumulative
damage concepts, as reported in Krawinkler (1996).
x Every excursion in the inelastic range causes damage in a structure that
brings it closer to failure. Moreover, the damage increases as the
inelastic excursion amplifies. Thus, relative amount of damage caused
Ev alu atio n of seismic d eman ds 181

by an inelastic excursion depends on the individual plastic deformation


range of the excursion ('Gp) (see Fig. 5.17).
x For a given deformation amplitude the damage is largest for a
symmetric excursion.
x The damage due to inelastic excursions is cumulative, thus it depends
on the number of inelastic excursions (Np) and the sum of normalized
plastic deformation ranges (6'Gpi/Gy).
x The damage depends on sequence effects, which means that it depends
on the sequence in which large and small excursions are applied to the
structure. In particular, the importance of sequence effects has not yet
been established.

Figure 5.17: Basic parameters in a typical cyclic of loading (ATC 1992).

Besides to the maximum level of deformation experienced by the structure,


which can be defined by the maximum normalized deformation (P=(G/Gy)max),
the main capacity parameters for characterizing the loading histories are the
number of inelastic excursions (Np), their individual plastic deformation range
('Gp) and the sum of normalized plastic deformation ranges (6'Gpi/Gy), which
182 Chap ter V

may be used as the basic cumulative damage parameter. For improving the
loading history description may be appropriate to add another parameter to
describe the distribution of individual plastic deformation range. The latter can
be represented by the ratio between the mean value (('Gp)av) and the
maximum value (('Gp)max) of the plastic deformation range (('Gp)av/('Gp)max).
This parameter (('Gp)av/('Gp)max) with the ones employed by Krawinkler
will be used in the current research to determine the loading history.
The demands imposed by an earthquake on a structural component or sub-
assembly depend on its configuration in a structure, the strength and elastic as
well inelastic dynamic characteristics of the structure, and the seismic input to
which the structure may be subjected.
For developing loading histories, Nassar and Krawinkler (1991) made
general considerations starting from studies on seismic demands on single
degree of freedom (SDOF) systems. These studies have been based on bilinear
(with 10% strain hardening) and stiffness degrading SDOF systems with
ductilities (P) ranging from 2 to 8 subjected to a set of 15 Western U.S.
earthquake ground motions. The magnitude of the earthquakes was varied
from 5.7 to 7.7, their durations varied significantly, and they represented
ground motions at stiff soil sites.
As a result of these studies, the Authors show the significant dependence of
some demand parameters (Np and 6'Gpi/Gy) on the natural period (T) of the
structures and the ductility ratio (P). In particular:
x The number of inelastic excursions (Np) amplifies with an increase in
the natural period of the structure (T) for T 0.2s, but decreases with
an increase of T for T 0.2s. Moreover, for each T value, Np increases
when the ductility ratio (P) amplifies (see Fig. 5.18).
x Similar to Np, the sum of normalized plastic deformation ranges
(6'Gpi/Gy) increases with an amplification of natural period (T) for T
0.2s but decreases with an increase of T for T 0.2s. Moreover, for
each T value, Np increases when the ductility ratio (P) amplifies (see
Fig. 5.19).
x The magnitudes of individual plastic deformation ranges ('Gp) of the
inelastic excursions can be represented by a lognormal distribution.
Large plastic deformation ranges are less frequent than small ones. In
fact, Hadidi-Tamjed (1987) reports that the mean of the plastic
Ev alu atio n of seismic d eman ds 183

deformation ranges (('Gp)av) in an earthquake is usually less than 15%


of the maximum plastic deformation range (('Gp)max).

On basis of these results, Krawinkler (1996) suggests the choice of testing


program and associated loading history. In particular, if cumulative damage
modeling is not object of the research, the monotonic load-displacement
response can be easily predictable, the deterioration is slow, then the Multiple
Step Test is the recommended test program.

Figure 5.18: Dependence of mean number of inelastic excursions on natural period and
ductility ratio (Krawinkler 1996).

Figure 5.19: Dependence of mean number of the sum of normalized plastic deformation
ranges on natural period and ductility ratio (Krawinkler 1996).
184 Chap ter V

The loading history for a Multiple Step Test consists of a series of


stepwise increasing deformation cycles, as shown in Figure 5.20. In Figure
5.20, G indicates the deformation control parameters, and ' represents
increment in peak deformation. In this history, the cycles should be symmetric
and several cycles should be completed during each step. The firsts load steps
should be performed in the elastic range to obtain stable and reliable values of
stiffness properties. The following steps should be performed at and beyond
yielding.

Figure 5.20: Loading history for Multiple Step Test (Krawinkler 1996).

For satisfying these requirements, the Author reports a table (Table 5.7) in
which shows, for three selected period (T=0.2s, T =0.5s, T =2.0s) and four
ductility ratio levels (P=2, P=4,P=6,P=8), representative values of predicted
and experimental SDOF seismic demands obtained from the loading history
reported in Figure 5.20 in which ' = Gy.
Ev alu atio n of seismic d eman ds 185

Np 6'Gpi/Gy
T P
mean mead + st. dev. experim. mean mead + st. dev. experim.
2 12 19 6 4 7 11
4 28 42 16 28 41 57
0.2
6 36 55 24 54 76 127
8 39 59 32 78 109 229
2 8 12 6 3 5 11
4 19 30 16 23 36 57
0.5
6 24 35 24 41 64 127
8 26 38 32 64 97 229
2 4 7 6 3 4 11
4 7 10 16 13 20 57
2.0
6 9 12 24 25 36 127
8 10 14 32 38 52 229

Table 5.7: Predicted and experimental demands for a bilinear SDOF (Krawinkler 1996).

The values of the total number of inelastic excursions (Np) and the sum of
normalized plastic deformation ranges (6'Gpi/Gy) reported in Table 5.7, as
well as that frequently adopted in loading sequences for CFSSSWs (SEAOSC
1997) testing, have been derived based on a bilinear hysteresis assumption. As
a consequence, a study aiming at characterizing the deformation history to
CFSSSW structural systems is needed.

5.3.2 Definition of the deformation history for the cyclic testing


Using a sufficiently large database of earthquake ground motion records,
the choice of an appropriate loading history for a multiple step test needs (a)
definition of the demand parameters and (b) definition of the magnitude of the
earthquake records.
Selection of the demand parameters has been based on cumulative damage
concepts, as explained in Section 5.3.1. As a result of these considerations the
assumed parameters are the following:
x maximum level of deformation that, from the results of monotonic
tests carried-out in the current research, can be characterized by a
value of ductility ratio equal to P=(G/Gy)max=Gu/Gy=36/6=6;
x number of inelastic excursions (Np);
x sum of normalized plastic deformation ranges (6'Gpi/Gy);
186 Chap ter V

x distribution of individual plastic deformation range (('Gp)av/('Gp)max).

These demand parameters strictly depends on the definition of the


individual plastic excursion. The typical cyclic behavior of CFSSSWs is
characterized by strong pinching; consequently definition of the inelastic
excursion is strictly related to the adopted convention.
In the current research the individual plastic deformation range of the
excursion i ('Gpi) is defined by the difference between the deformation
corresponding to the null value of the shear force achieved on the unloading-
branch of the excursion i ('Gi,0) and the maximum value of deformations
corresponding to the null values of the shear force achieved on the unloading-
branch of the previous excursion ('Gmax,0=max{'G1,0; 'G2,0; 'Gi-1,0}), as
shown in Figure 5.21.
The evaluation of assumed deformation demand parameters under the
selected earthquake ground motions has been carried out by scaling records.
In particular, each accelerogram has been scaled in such way that the elastic
spectral acceleration (Sae) be equal to the spectral acceleration inducing the
ultimate lateral displacement (Saeu) obtained through the IDA. In other words,
each design accelerogram (PGA=PGAdes=0.25g) has been scaled by Sau/Sad.
The Sau/Sad values are reported in Table 5.6.

V (kN) Excursion (i+1)+


Excursion (i)+

'Gp,(i)- 'Gp,(i)-

G (mm)

'Gp,(i)+ 'Gp,(i+1)+

Excursion (i)-
Excursion (i+1)-

Figure 5.21: Assumed definitions of individual plastic deformation range.


Ev alu atio n of seismic d eman ds 187

Results of the statistic characterization of the deformation demand are


synthesized in Figures 5.22 a through d and Table 5.8. In particular, the
maximum normalized deformation ((G/Gy)max) (see Fig. 5.22a), the number of
inelastic excursion (Np) (see Fig. 5.22b), the sum of normalized plastic
deformation ranges (6'Gpi/Gy) (see Fig. 5.22c), and the ratio between the
mean value and the maximum value of the plastic deformation range
(('Gp)av/( 'Gp)max) (see Fig. 5.22d) are reported for each accelerogram. Also
the mean and the standard deviation values have been calculated for all
earthquake records, for all soil types and for both the LA and UM record
groups.
For all accelerograms, the (G/Gy)max values range from 5.1 to 7.5 with a
mean value of 6.1 and a standard deviation of 0.4; Np ranges from 9 to 45 with
a mean and standard deviation values of 24 and 11, respectively; the range of
6'Gpi/Gy is from 5.6 to 9.2 with a mean value of 7.5 and a standard deviation
of 1.0; finally, (('Gp)av/(('Gp)max ranges from 0.09 to 1.21 and its mean and
standard deviation values are 0.28 and 0.22, respectively.
From results reported in Figures 5.22 and Table 5.8, it can also be noted
that the mean values of the parameters associated to the damage level (Np and
6'Gpi/Gy) are similar for both the LA and UM record groups. In particular, the
mean values of Np and 6'Gpi/Gy result moderately higher (about 18 and 10%,
respectively) for UM records. Also the dispersion of data results moderately
higher for UM records in the case of Np and 6'Gpi/Gy. For the parameter
related to the distribution of individual plastic deformation range
(('Gp)av/('Gp)max), both the mean and standard deviation values in LA
earthquake records are sensibly higher in comparison with UM records. In
particular, the mean value results about 50% higher in LA earthquake.
The comparison between the results obtained for different soil types shows
that the highest mean values for the parameters Np and 6'Gpi/Gy have been
obtained for the soil type D, while the mean values of Np and 6'Gpi/Gy result
the lowest for soil types B and E, respectively. Also the scattering of data is
the highest for the soil type D in the case of Np, while for 6'Gpi/Gy the
maximum dispersion has been obtained for the soil type C. For the parameter
(('Gp)av/('Gp)max), the maximum and minimum mean values have been found
for soil types E and D, respectively, while the scattering results larger in the
case of soil type C.
188 Chap ter V

8
mean + st.dev mean: 6.1 - st.dev: 0.4
mean
7
mean st.dev
6

0
LA-A-01
LA-A-02
LA-A-03
LA-B-01
LA-B-02
LA-B-03
LA-C-01
LA-C-02
LA-C-03
LA-D-01
LA-D-02
LA-D-03
LA-E-01
LA-E-02
UM-A-01
UM-A-02
UM-A-03
UM-B-01
UM-B-02
UM-B-03
UM-C-01
UM-C-02
UM-C-03
UM-E-01
UM-E-02
UM-E-03
Figure 5.22a: Results of the statistic characterization of the deformation demand in terms of
maximum normalized deformation.

50
mean: 24 - st.dev: 11
45 mean + st.dev
mean
40
mean st.dev
35
30
25
20
15
10
5
0
LA-A-01
LA-A-02
LA-A-03

LA-E-01
LA-E-02
LA-B-01
LA-B-02
LA-B-03
LA-C-01
LA-C-02
LA-C-03
LA-D-01
LA-D-02
LA-D-03

UM-A-01
UM-A-02
UM-A-03
UM-B-01
UM-B-02
UM-B-03
UM-C-01
UM-C-02
UM-C-03
UM-E-01
UM-E-02
UM-E-03

Figure 5.22b: Results of the statistic characterization of the deformation demand in terms of
number of inelastic excursion.
Ev alu atio n of seismic d eman ds 189

10
mean + st.dev mean: 7.5 - st.dev: 1.0
9 mean
mean st.dev
8
7
6
5
4
3
2
1
0
LA-A-01
LA-A-02
LA-A-03
LA-B-01
LA-B-02
LA-B-03
LA-C-01
LA-C-02
LA-C-03
LA-D-01
LA-D-02
LA-D-03
LA-E-01
LA-E-02
UM-A-01
UM-A-02
UM-A-03
UM-B-01
UM-B-02
UM-B-03
UM-C-01
UM-C-02
UM-C-03
UM-E-01
UM-E-02
UM-E-03
Figure 5.22c: Results of the statistic characterization of the deformation demand in terms of
sum of normalized plastic deformation ranges.

1.3
mean: 0.28- st.dev: 0.22
1.2
1.1
1.0
0.9
0.8
0.7
mean + st.dev
mean
0.6
mean st.dev
0.5
0.4
0.3
0.2
0.1
0.0
LA-A-01
LA-A-02
LA-A-03
LA-B-01
LA-B-02
LA-B-03
LA-C-01
LA-C-02
LA-C-03
LA-D-01
LA-D-02
LA-D-03
LA-E-01
LA-E-02
UM-A-01
UM-A-02
UM-A-03
UM-B-01
UM-B-02
UM-B-03
UM-C-01
UM-C-02
UM-C-03
UM-E-01
UM-E-02
UM-E-03

Figure 5.22d: Results of the statistic characterization of the deformation demand in terms of
ration between the mean value and the maximum value of the plastic deformation range.
190 Chap ter V

Earthquake record (Gp/Gy)max Np 6'Gpi/Gy ('Gp)av/(('Gp)max


LA-A-01 6.1 28 7.6 0.15
LA-A-02 6.1 34 7.1 0.13
LA-A-03 5.9 13 6.9 0.39
LA-B-01 5.9 16 7.2 0.32
LA-B-02 6.2 13 6.0 0.28
LA-B-03 6.0 13 7.8 0.38
LA-C-01 6.3 12 6.6 0.27
LA-C-02 6.4 10 7.2 0.40
LA-C-03 6.4 14 6.7 0.31
LA-D-01 6.1 28 8.5 0.17
LA-D-02 7.5 32 8.7 0.21
LA-D-03 6.0 32 7.2 0.19
LA-E-01 5.6 33 6.7 1.21
LA-E-02 6.0 35 6.6 0.17
UM-A-01 6.3 9 7.4 0.24
UM-A-02 5.1 25 5.6 0.19
UM-A-03 6.7 30 9.2 0.33
UM-B-01 6.3 32 8.1 0.31
UM-B-02 5.8 37 9.2 0.16
UM-B-03 6.2 11 7.7 0.36
UM-C-01 6.3 29 7.8 0.09
UM-C-02 5.8 42 8.6 0.10
UM-C-03 6.4 45 9.0 0.10
UM-E-01 6.0 11 7.1 0.46
UM-E-02 6.0 18 7.3 0.17
UM-E-03 6.5 24 7.9 0.18
Mean (Type A soil records) 6.0 23 7.3 0.24
St. Dev. (Type A soil records) 0.5 10 1.2 0.10
Mean / St. Dev. (Type A soil records) 0.08 0.43 0.16 0.44
Mean (Type B soil records) 6.1 20 7.7 0.30
St. Dev. (Type B soil records) 0.2 11 1.1 0.08
Mean / St. Dev. (Type B soil records) 0.04 0.55 0.14 0.27
Mean (Type C soil records) 6.3 25 7.6 0.21
St. Dev. (Type C soil records) 0.2 16 1.0 0.14
Mean / St. Dev. (Type C soil records) 0.04 0.62 0.13 0.64
Mean (Type D soil records) 6.5 31 8.1 0.19
St. Dev. (Type D soil records) 0.9 2 0.8 0.02
Mean / St. Dev. (Type D soil records) 0.13 0.08 0.10 0.11
Mean (Type E soil records) 6.0 24 7.1 0.44
St. Dev. (Type E soil records) 0.3 10 0.6 0.45
Mean / St. Dev. (Type E soil records) 0.05 0.42 0.08 1.02
Mean (LA records) 6.2 22 7.2 0.33
St. Dev. (LA records) 0.4 10 0.8 0.27
Mean / St. Dev. (LA records) 0.07 0.45 0.11 0.82
Mean (UM records) 6.1 26 7.9 0.22
St. Dev. (UM records) 0.4 12 1.0 0.12
Mean / St. Dev. (UM records) 0.07 0.46 0.13 0.53
Mean (all records) 6.1 24 7.5 0.28
St. Dev. (all records) 0.4 11 1.0 0.22
Mean / St. Dev. (all records) 0.07 0.45 0.13 0.77

Table 5.8: Results of the statistic characterization of the deformation demand.


Ev alu atio n of seismic d eman ds 191

REFERENCES

ATC (1992) Guidelines for cyclic seismic testing of components of steel structures
(ATC-24). ATC (Applied Technology Council). Redwood City, CA, USA.
Branston, A., Boudreault, F., Rogers, C.A. (2003) Testing on steel frame / wood
panels shear walls. Progress Report, Departement of Civil Engineering and Applied
Mechanics, McGill University. Montreal.
COLA-UCI (2001) Report of a testing program of light-framed walls with wood-
sheathed shear panels. Final report to the City of Los Angeles Department of
Building and Safety, Structural Engineers Association of Southern California, Irvine,
CA,USA.
Della Corte, G., De Matteis, G., Landolfo, R. (1999) A mathematical model
interpretino the cyclic behaviour of steel beam-to-column joint. In Proceedings of the
XVII Congresso CTA (CTA 1999). Napoli.
Fulop, L.A. & Dubina, D. (2003). Are the cold-formed wall stud shear walls
dissipative systems in seismic resistant buildings? How much?. In Proceedings of the
4th International Conference on Behavior of Steel Structures in Seismic Areas
(STESSA 2003). Mazzolani F.M. (ed.). A.A. Balchema Publishers.
Ghersi, A. & Noce, (1999) Modalit di utilizzazione del programma DIANA. Istituto
di Scienza delle costruzioni, Universit di Catania. Catania.
Hadidi-Tamjed, H. (1987) Statistical response of inelastic SDOF systems subjected to
earthquake. Ph.D. Dissertation. Department of Civil Engineering, Stanford
University.
Krawinkler, H. (1996) Cycling loading histories for seismic experimentation on
structural components. Earthquake Spectra. Vol. 12, No.1:1-12.
Nassar, A.A. & Krawinkler, H. (1991) Seismic demands for SDOF and MDOF
systems. John A. Blume Earthquake Engineering Center Report No.95, Department
of Civil Engineering, Stanford University.
Ordinanza PCM (2003) Primi elementi in materia di criteri generali per la
classificazione sismica del territorio nazionale e di normative tecniche per le
costruzioni in zona sismica. Ordinanza della Presidenza del Consiglio dei Ministri
No.3274/2003.
prEN 1998-1 (2001) Eurocode 8: Design of structures for earthquake resistance
Part 1: General rules Seismic actions and rules for buildings. CEN (European
Committee for Standardization). Bruxelles.
Richard, R.M. & Abbott, B.J. (1975) Versatile elastic-plastic stress-strain formula.
Journal of mechanical division. ASCE, Vol.101, No.4:511-515.
192 Chap ter V

SEAOSC (1997) Standard method of cyclic (reversed) load tests for shear resistance
of framed walls for buildings. Structural Engineers Association of Southern
California (SEAOSC). Whittier, CA, USA.
Serrette, R., Nguyen, H., Hall, G. (1996a) Shear wall values for light weight steel
framing. Report No. LGSRG-3-96, Light Gauge Steel Research Group, Department
of Civil Engineering, Santa Clara University. Santa Clara, CA, USA.
Serrette, R., Hall, G., Nguyen, H. (1996b) Dynamic performance of light gauge steel
framed shear walls. In Proceedings of the 13th International Specialty Conference on
Cold-formed Steel Structures. St. Louis, MO, USA: 487-498.
Serrette, R., Encalada, J., Matchen, B., Nguyen, H., Williams, A. (1997b) Additional
shear wall values for light weight steel framing. Report No. LGSRG-1-97, Light
Gauge Steel Research Group, Department of Civil Engineering, Santa Clara
University. Santa Clara, CA, USA.
193

Chapter VI
The cyclic test

The second step of the experimental program, which consists in the cyclic
testing of cold-formed steel stud shear wall (CFSSSW) sub-assemblies, is
illustrated in this Chapter.
The specimen tested under cyclic loading was nominally identical to that
tested under monotonic loading. Moreover, also the test set-up and the
instrumentation adopted in the cyclic testing were the same as in the
monotonic one. Consequently, because the description of the test specimen,
test set-up and instrumentation have been previously illustrated in Chapter 4,
only the test procedure and test results are presented in the current Chapter. In
particular, the Section 6.1 is devoted to present the test procedure, while the
test results are presented in Section 6.2.
194 Chap ter VI

6.1 TEST PROCEDURE

In the cyclic test, the specimen was subjected to fully reversing cyclic
horizontal displacements according to the definition of the deformation history
parameters, which have been illustrated in the Chapter 5 (Section 5.3.2). In
particular, the test specimen was subjected to fully reversing cyclic
displacements consisting of a series of stepwise increasing deformation
cycles, by following a procedure similar to the one described in ATC (1992)
for a multiple step test (see Chapter 5 Section 5.3.1).

The load history consisted on applying three cycles of fully reversing,


displacement-controlled load at each wall displacement increment
representing 25% (G= 1.5mm), 50% (G= 3.0mm) and 75% (G= 4.5mm) of the
conventional yield limit state (YLS) displacement (Gy = 6.0mm). Then, the
wall displacement was increased for three load cycles to 100% of the YLS
displacement. Next, cycles of displacement for three cycles each at 150%
(G= 9.0mm), 200% (G= 12.0mm), 300% (G= 12.05mm), 400%
(G= 24.0mm) and 600% (G= 36.0mm) of the YLS displacement were applied.
At this point the load history based on the statistic characterization of
deformation demand, which has been illustrated in Section 5.3.2, was
completed. However, to capture the behavior of the specimen under larger
displacement amplitudes, other cycles of displacement for three cycles each at
700% (G= 42.0mm), 800% (G= 48.0mm), 900% (G= 54.0mm), 1000% (G=
60.0mm), 1100% (G= 66.0mm), 1200% (G= 72.0mm), and 1300% (G=
78.0mm) of the YLS displacement were applied.
This test protocol involved displacements at rate of 2.00mm/s.

The adopted test protocols are illustrated in Table 6.1 and Figure 6.1. The
Figure 6.2 shows the numerical cyclic response in terms of unit shear force (v)
versus lateral displacement (G) curves obtained for the assumed load history
by using the model of the hysteretic behavior of CFSSSW systems, which
have been already illustrated and calibrated in the Section 5.1. As a result, the
representative values of predicted seismic demands, which have been obtained
from the assumed loading history, are:
The cyclic test 195

x maximum level of deformation: (G/Gy)max=6.0;


x number of inelastic excursion: Np=25;
x sum of normalized plastic deformation ranges: 6'Gpi/Gy=6.2;
x ration between the mean value and the maximum value of the plastic
deformation range (('Gp)av/('Gp)max)=0.28.

From the examination of the value of these parameters it can be deduced


that the adopted load history is reasonable. In fact, the mean deformation
demand parameters obtained from the statistic characterization are close to
those characterizing the adopted loading history, as illustrated in Table 6.2.
Moreover, from comparison between the results obtained in this research,
which takes in-to account the cyclic behavior of CFSSSW systems, and those
reported for a bilinear SDOF by Krawinkler (1996) ((Np=30 and 6'Gpi/Gy=36
in average, see Figs. 5.18 and 5.19) a reduction of the parameters associated to
the damage level may be noted in the case of CFSSSW systems. In particular,
the decreasing is of 24% for Np and 79% for 6'Gpi/Gy.

No. of Displacement Displacement / YLS


cycles (mm) displacement (%)
3 1.5 25
3 3.0 50
3 4.5 75
3 6.0 100
3 9.0 150
3 12.0 200
3 18.0 300
3 24.0 400
3 36.0 600
3 42.0 700
3 48.0 800
3 54.0 900
3 60.0 1000
3 66.0 1100
3 72.0 1200
3 78.0 1300

Table 6.1: Cyclic test protocol.


196 Chap ter VI

load history based on the


statistic characterization
of deformation demand

Figure 6.1: Cyclic test protocol.

Figure 6.2: Numerical cyclic response.


The cyclic test 197

(Gp/Gy)max Np 6'Gpi/Gy ('Gp)av/(('Gp)max


Statistic characterization (mean value) 6.1 24 7.5 0.28
Statistic characterization (st. dev. value) 0.4 11 1.0 0.22
Adopted load history 6.0 25 6.2 0.28

Table 6.2: Comparison between the deformation demand parameters obtained from the
statistic characterization (see Table 5.8) and the adopted loading history.

6.2 TEST RESULTS

The measured responses of all instruments are reported in Appendix F.


The global cyclic response may be represented by the unit shear resistance
(v) vs. mean displacement (d) curve, where v and d have been defined in
Chapter 4 (see Equations 4.7 and 4.8). The v-d response curve is showed in
Figure 6.3.
In this Figure the values of the maximum (positive) experimental loads
obtained at the first (vEXP+1) and last (the third one) (vEXP+3) hysteretic loop are
reported together with the minimum (negative) experimental loads
corresponding to the first (vEXP-1) and third (vEXP-3) hysteretic loop. In
particular, the experimental maximum loads were obtained for a lateral
displacement of +36mm. They resulted vEXP+1=+16.4kN/m and
vEXP+3=+10.2kN/m for the first and third cycle, respectively. For the negative
loads the minimum values were found for a lateral displacement of -24mm.
Their values were vEXP-1=-14.8kN/m and vEXP-3=-12.5kN/m for the first and
third cycle, respectively.
From the comparison between the experimental loads obtained at the first
(vEXP+1 and vEXP-1) and third (vEXP+3 and vEXP-3) hysteretic loops, a reduction of
38% and 16% for positive (vEXP+1 vs. vEXP+3) and negative (vEXP-1 vs. vEXP-3)
shear strengths, respectively was noted. These results highlight a clear shear
strength degradation. From the comparison between the positive (vEXP+1 and
vEXP+3) and negative (vEXP-1 and vEXP-3) experimental loads, for the first
hysteretic loop (vEXP+1 vs. vEXP-1) it was observed a higher shear capacity of
11% for positive loads, while for the third hysteretic loop (vEXP+3 vs. vEXP-3) a
higher shear strength of 23% was found for negative loads.
198 Chap ter VI

In Figure 6.3 the estimated lateral strength (vR) is also reported. From the
comparison between the mean value of the experimental lateral strength
obtained at first hysteretic loop (vEXP1 = vEXP+1 - vEXP-1 / 2 = 15.6kN/m) and the
estimated lateral strength (vR=17.9kN/m) it is possible to observe that, in the
case of loads applied cyclically, the semi-analytical calculation illustrated in
the Chapter 2 gives a 15% overestimation of shear strength. Moreover, if the
mean experimental value of experimental lateral strength obtained at third
hysteretic loop (vEXP3 = vEXP+3 - vEXP-3 / 2 = 11.4kN/m) is considered, the semi-
analytical prediction gives a 57% overestimation of the lateral strength.

Estimate lateral strength (vR = 17.9kN/m)

Experimental lateral strength


(vEXP+1 = 16.4kN/m)

Experimental lateral strength


(vEXP+3 = 10.2kN/m)

Experimental lateral strength


(vEXP-1= -14.8kN/m)
Experimental lateral strength
(vEXP-3= -12.5kN/m)

Estimate lateral strength (vR = 17.9kN/m)

Figure 6.3: Unit shear resistance (v) vs. mean displacement (d) curve.

During the test two different behaviors could be identified.


x For lateral displacement less than the one corresponding to maximum
shear resistance the behavior of OSB sheathings-to-frame connections
resulted from a combination of the tilting of the screws about the plane
of the stud flange and the screw heads pulling through the OSB
sheathings, as shown in Figure 6.4 a and b, respectively. The response
of GWB sheathings-to-frame connections was characterized by a
combination of the bearing of the GWB panels and the screws heads
The cyclic test 199

pulling through the GWB panels, as shown in Figure 6.5 a and b,


respectively.
For these displacement levels the skeleton-to-panels deformation was
congruent. In fact, the wall framing deformed into a parallelogram and
the sheathings had rigid body rotation.
x For lateral displacement larger than the one corresponding to the
maximum shear resistance, the heads of the end screws completely
pulled through the sheathings in the bottom half of the walls (see
Figures 6.6 a and b for OSB and GWB sheathings, respectively) or, in
some cases, the screws caused the rupture of the sheathing edges (see
Figs. 6.7 a and b for OSB and GWB sheathings, respectively). As a
result, both the OSB and GWB sheathings became unzipped along the
panel edges in the bottom half of the specimen (see Fig. 6.8).
Moreover, for displacement levels more than 42mm also the
distortional buckling in the end studs of the wall 1 was observed, as
shown in Figure 6.9.
For these displacement levels the deformation of the wall framing still
had the shape of a parallelogram, while due to the rupture sheathing-
to-frame connections, the rotation of the sheathings was limited. As
shown in Figure 6.10.

As in the case of monotonic test, in the cyclic test any deformation of the
OSB sheathing-to-floor framing connections was not observed and the shear
and the tension anchors did not suffer any type of failure.
200 Chap ter VI

(a) Tilting of the screws (b) Screw heads pull through

Figure 6.4: Behavior of OSB sheathing-to-frame connections.

(a) Bearing of the panels (b) Screw heads pull through

Figure 6.5: Behavior of GWB sheathing-to-frame connections.


The cyclic test 201

(a) OSB sheathing-to frame connections (b) GWB sheathing-to frame connections

Figure 6.6: Failure of sheathing-to-frame connections due to screw heads pull through.

(a) OSB sheathing-to frame connections (b) GWB sheathing-to frame connections

Figure 6.7: Failure of sheathing-to-frame connections due to rupture of sheathing edges.


202 Chap ter VI

Figure 6.8: Unzipping of sheathings. Figure 6.9: Buckling of end studs.

Figure 6.10: Deformation of walls for lateral displacement amplitudes more than 36mm.
The cyclic test 203

The comparison between the cyclic and monotonic behavior is shown in


Figure 6.11. In this Figure both cyclic and monotonic responses are
represented through the unit shear resistance (v) vs. mean displacement (d)
curves. From the comparison of these curves it may be observed as the cyclic
loading produced a 11% reduction of the shear strength (by considering the
comparison between the lateral strength found in the monotonic test and the
one obtained from the cyclic test at first hysteretic loop of positive
displacements). On the contrary, by comparing the lateral strength obtained
from the monotonic test with the lateral strength corresponding to the cyclic
test at third hysteretic loop of positive displacements the reduction of shear
capacity increased up to 45%; whereas, considering the shear strength
obtained for negative displacements a 20% and 32% degradation appeared in
the cyclic test at first and third hysteretic loops, respectively.
Moreover, the difference between the cyclic and monotonic response
became significant for the unstable part of the behavior (for displacement
amplitudes more than 36mm).

The force (V) vs. displacement (d) response curves for the two walls is
shown in Figure 6.12. In particular, in this Figure the force vs. displacement
curves obtained from measures of the actuators a1 and a2 have been reported
for the walls 1 and 2, respectively. From the comparison of lateral responses
of two walls observations similar to those deduced for the monotonic test are
possible:
x the walls had a similar behavior for positive displacements less than
+36mm (displacement corresponding to the maximum shear load),
while they exhibited different responses for larger displacements;
x considering the first hysteretic loops, the maximum shear loads (for
positive displacements) were +40 and +39kN for wall 1 and 2,
respectively; therefore, the two walls revealed the same shear capacity;
whereas considering the third hysteretic loops, the maximum shear
loads were +26 and +23kN for wall 1 and 2, respectively; therefore,
the wall 1 appeared to be lightly more resistant than wall 2 of about
13%;
204 Chap ter VI

x the walls sowed similar response for smaller negative displacements


with significant variation of the behavior for displacements less than
-36mm;
x considering the first hysteretic loops, the minimum shear loads (for
negative displacements) were -37 and -34kN for wall 1 and 2,
respectively; consequently, the wall 1 was lightly more resistant than
wall 2 of about 9%; whereas considering the third hysteretic loops, the
minimum shear loads were -26 and -23kN for wall 1 and 2,
respectively; therefore, the difference between walls in terms of shear
resistance was of 13%.

Figure 6.11: Cyclic vs. monotonic behavior.


The cyclic test 205

Figure 6.12: Shear (V) vs. displacement (d) curves for wall 1 and wall 2.

REFERENCES

ATC (1992) Guidelines for cyclic seismic testing of components of steel structures
(ATC-24). ATC (Applied Technology Council). Redwood City, CA, USA.
Krawinkler, H. (1996) Cycling loading histories for seismic experimentation on
structural components. Earthquake Spectra. Vol. 12, No.1:1-12.
207

Conclusions

Seismic performance analysis of low-rise residential buildings built with


cold-formed steel members has been the objective of this research.
The attention has been focused on the seismic behavior of sheathed cold-
formed steel stud shear walls through the evaluation of the seismic capacity
(experimental phase), seismic demand (theoretical phase) and their
comparison.
On the base of observations and results of the experimental and theoretical
investigation, the following conclusions may be drawn:

Global experimental behavior: monotonic response vs. cyclic response


In case of monotonic loading, for all displacement levels, wall framings
deformed into a parallelogram and the sheathings had rigid body rotation, in
such a way that the evolution of the deformation was coherent with the
sheathing-to-wall framing connections failure.
In case of cyclic loading, the global behavior of the specimen was similar to
that observed in the monotonic testing for displacements less than the one
corresponding to the maximum shear capacity, while the skeleton-to-panels
deformation was not congruent for higher displacement levels. Moreover, only
in the case of cyclic testing, the distortional buckling in the end studs of the
wall 1 was observed for displacement amplitudes more than 42mm.
In both the monotonic and cyclic testing any deformation of the oriented
strand board (OSB) sheathing-to-floor framing connections was not observed,
and the shear and the tension anchors did not suffer any type of failure.
208

Local (sheathing-to-wall framing connections) experimental behavior:


monotonic response vs. cyclic response
Sheathing-to-wall framing connections in the case of monotonic test
exhibited similar behavior, but they degraded more gradually than during the
cyclic test.
In particular, for lower displacement levels the behavior of OSB sheathings-
to-frame connections resulted from a combination of the tilting of the screws
about the plane of the stud flange and the screw heads pulling through the
OSB sheathings, while the response of gypsum wallboard (GWB) sheathings-
to-frame connections was characterized by combination of the bearing of the
GWB panels and the screws heads pulling through the GWB panels.
For higher lateral displacements, the heads of the end screws completely
pulled through the sheathings or, in some cases, the screws caused the rupture
of the sheathing edges. As a result, both the OSB and GWB sheathings
became unzipped along the panel edges.

Comparison of shear responses of two walls: monotonic response vs. cyclic


response
For both the monotonic and cyclic loading, the walls had similar behavior
for small displacements, while they exhibited different response for large
displacement levels.
In particular, the wall 1 was more resistant than wall 2 of about 7% in the
case of the monotonic test. In the case of cyclic test, considering the first
hysteretic loops, the two walls revealed the same shear capacity for positive
displacements, while the wall 1 was lightly more resistant than wall 2 of 9%
for the negative ones; whereas, considering the third hysteretic loops the wall
1 was lightly more resistant than wall 2 of 13% for both positive and negative
displacements.

Monotonic response vs. cyclic response in terms of shear strength


By comparing the unit shear strength found in the monotonic test and the
one obtained from cyclic test at the first hysteretic loops of positive
Conclu sion s 209

displacements, the cyclic loading produced a 11% reduction. This reduction


became of 20% for negative displacements; whereas, considering the shear
strength obtained in the cyclic test at third hysteretic loops the degradation
increased up to 45% and 32% for positive and negative displacements,
respectively.
Moreover, the difference between the cyclic and monotonic response
became significant for the unstable part of the behavior.

Shear strength degradation under cyclic loading


From comparison between the experimental loads obtained at the first and
third hysteretic loops a clear shear strength degradation was observed. In fact
the 38% and 16% reduction of shear capacity resulted for positive and
negative shear strengths, respectively.

Effectiveness of lateral-load transfer from horizontal diaphragms to stud walls


The lateral-load transfer from horizontal diaphragms to stud walls was
effective. In fact the floor-to-walls sliding was small in both the monotonic
and cyclic tests.

Reliability of semi-empirical calculation of the shear strength


The calculation of shear strength by using the semi-empirical
methodologies appears valid, it revealing a small conservative 3% prediction
in the case of monotonic loading. On contrary, in the case of cyclic loading,
the results give 15% and 57% overestimation of the shear capacity, if the
highest (first hysteretic loops) and lowest (third hysteretic loops) strength
curves, respectively are used.

Incremental Dynamic Analysis (IDA) results


From the examination of the IDA results, a discrete ductility for the
CFSSSW systems has been found. In fact, considering the Saeu/Saey ratio (in
which Saey and Saeu are the elastic spectral accelerations corresponding to yield
210

limit (Gy) and ultimate limit displacements (Gu), respectively) as a measure of


seismic toughness, for all the examined earthquake records, this parameter
ranges from 1.26 to 3.89 with a mean value of 2.27 and a standard deviation
of 0.62.
The comparison between the results obtained for different soil types has
shown that the accelerograms recorded on soil type D are the most severe and
the least scattered; whereas, the least severe results have been found for soil
types A and B, while the maximum dispersion of data has been obtained for
the soil type C.

Definition of the load history


Once fixed the maximum level of the normalized deformation
(P=(G/Gy)max=Gu/Gy=36/6=6), as result of the statistic characterization of
deformation history, the following parameters have been obtained: number of
inelastic excursions (Np=24 in average); sum of the normalized plastic
deformation ranges (6'Gpi/Gy=7.5 in average); ratio between the mean value
and the maximum value of the plastic deformation range
(('Gp)av/('Gp)max=0.28 in average).
From comparison between the results obtained in this research (that take
in-to account the cyclic behavior of CFSSSW systems) and those reported for
a bilinear SDOF system by Krawinkler (Np=30 and 6'Gpi/Gy=36 in average,
see Figs. 5.18 and 5.19) a reduction of the parameters associated to the
damage level has been observed in the case of CFSSSW systems. In
particular, a decrease of 24% for Np and 79% for 6'Gpi/Gy has been found.
The test procedure, made of fully reversing cyclic displacements it
consisting of a series of stepwise increasing deformation cycles, which has
been derived in this research, results to be reasonable. In fact, the value of
parameters associated to the damage level corresponding to this deformation
history (Np=25, 6'Gpi/Gy=6.2 and ('Gp)av/('Gp)max=0.28) are close to those
obtained from statistic characterization.
A-1

Appendix A

Summary of existing experimental results


A-2

Mc Creless & Tarpy (1978)

Type of aspect horizontal straps (HS) Type of Type of


loading wall size ratio opening frame frame fasteners sheathing sheathing fasteners solid blocking (SB) X-bracing X-bracing fasteners anchorage failure ultimate
Label Test (1) (2) (2) (3) (4) (5) (6) (5) (7) (7) (5) (8) (9) shear
steel grade
sheathing stud size type;
area ratio track size thickness; type; exterior spacing; HB size
hxL h/L r stud spacing type orientation interior spacing SB size size type; number kN/m

M78-1 A - 6.03
3660x3660 1.00
HB ?
M78-2 A* F 5.11
SB ?

M78-3 B 3660x4880 0.75 5.75

M78-4 C 3660x7320 0.50 C (FT) 5.30

M78-5 D 3050x3660 0.83 F 5.47


?
GWB;12.7;? BHSC3.5x25; 305; 305
C89x?x?x0.84*
M78-6 E M 3050x4880 0.63 1.00 LPHSC4.8x13 + - - BCA 5.20
U92x38x0.84
GWB;12.7;? BHSC3.5x25; 305; 305
610
M78-7 F 3050x7320 0.42 - C (FT) 5.66

M78-8 G 2440x2440 1.00 5.84

M78-9 H 2440x3660 0.67 F 5.84

M78-10 I 2440x4880 0.50 6.84

M78-11 J 2440x7320 0.33 C (FT) 5.66

length in mm; ?: not known; -: not present


(1) M: monotonic; C: cyclic
(2) h: heigth of the wall; L: length of the wall
(3) r =1 / [1 + Ao / (h 6Li)] with Ao: area of openings and Li the length of the full height wall segment
(4) C__x__x__x__ (lipped channel section) web depth x flange size x lip size x thickness; U__x__x__ (unlipped channel section) web depth x flange size x thickness; *: double back-to-back coupled end studs
(5) BHSC: bugle-head screws FHSC: flat head screws; LPHSC: low profile head screws; MTHSC: modified truss head screws; WHSC: wafer head screws; __x__: nominal diameter x length; NA: Nails diameter; PI: Pins diameter
(6) GSB: Gypsum sheathing board; GWB: Gypsum wallboard; FB: Fiber board; OSB: Oriented strand board; PLY: Plywood; SCS: Steel corrugated seet; SSS: Steel sheet sheathing;
V: parallel to the frame (vertical); O: orthogonal to the frame (horizontal)
(7) I__x__ flat strap wide x thickness
(8) BCA: bolted clip angles; PCA: clip angles fixed wit powder actuated fasteners; HD: hold-down; SHD: strip-hold down
(9) F: foundation uplift failure; S: stud buckling failure; X: X bracing yielding;
C (FPO): connections failure (fasteners pull out failure); C (FPT): connections failure (fasteners pull throgh failure); C (FS): connections failure (fasteners shear failure); C (FT): connections failure (fasteners tilting failure); C (NS): connections faiO): c
(net section failure);
Appendix A
Tarpy & Girard (1982)

Type of aspect horizontal straps (HS) Type of Type of


loading wall size ratio opening frame frame fasteners sheathing sheathing fasteners solid blocking (SB) X-bracing X-bracing fasteners anchorage failure ultimate
Label Test (1) (2) (2) (3) (4) (5) (6) (5) (7) (7) (5) (8) (9) shear
steel grade
sheathing stud size type;
area ratio track size thickness; type; exterior spacing; HB size
hxL h/L r stud spacing type orientation interior spacing SB size size type; number kN/m

TG82-1 A 2440x2440 1.00 6.03


BCA

TG82-2 B 2440x3660 0.67 5.47

TG82-3 E - 4.57

GWB;12.7;H
TG82-4 G + - 5.66
GWB;12.7;H
A 36
C89x25x13x0.84*
TG82-5 K PCA 3.82
Summary of existing experime ntal results

U92x38x0.84
610
BHSC3.5x25; 305; 305
TG82-6 L M 1.00 ? - - - F 6.20
BHSC3.5x25; 305; 305
GSB;12.7;H
TG82-7 M 2440x2440 1.00 + BCA 3.93
GWB;12.7;H
GWB;12.7;H
TG82-8 N + 7.85
PLY;12.7;H

TG82-9 P 5.28

GSB;12.7;H
TG82-10 Q + 4.38
GWB;12.7;H
A 36
C89x25x13x0.84*
TG82-11 R BCA 6.93
U92x38x0.84
406
length in mm; ?: not known; -: not present
(1) M: monotonic; C: cyclic
(2) h: heigth of the wall; L: length of the wall
(3) r =1 / [1 + Ao / (h 6Li)] with Ao: area of openings and Li the length of the full height wall segment
(4) C__x__x__x__ (lipped channel section) web depth x flange size x lip size x thickness; U__x__x__ (unlipped channel section) web depth x flange size x thickness; *: double back-to-back coupled end studs
(5) BHSC: bugle-head screws FHSC: flat head screws; LPHSC: low profile head screws; MTHSC: modified truss head screws; WHSC: wafer head screws; __x__: nominal diameter x length; NA: Nails diameter; PI: Pins diameter
(6) GSB: Gypsum sheathing board; GWB: Gypsum wallboard; FB: Fiber board; OSB: Oriented strand board; PLY: Plywood; SCS: Steel corrugated seet; SSS: Steel sheet sheathing;
V: parallel to the frame (vertical); O: orthogonal to the frame (horizontal)
(7) I__x__ flat strap wide x thickness
(8) BCA: bolted clip angles; PCA: clip angles fixed wit powder actuated fasteners; HD: hold-down; SHD: strip-hold down
(9) F: foundation uplift failure; S: stud buckling failure; X: X bracing yielding;
C (FPO): connections failure (fasteners pull out failure); C (FPT): connections failure (fasteners pull throgh failure); C (FS): connections failure (fasteners shear failure); C (FT): connections failure (fasteners tilting failure); C (NS): connections faiO): c
(net section failure);
A-3
A-4

Tissel (1993)

Type of aspect horizontal straps (HS) Type of Type of


loading wall size ratio opening frame frame fasteners sheathing sheathing fasteners solid blocking (SB) X-bracing X-bracing fasteners anchorage failure ultimate
Label Test (1) (2) (2) (3) (4) (5) (6) (5) (7) (7) (5) (8) (9) shear
steel grade
sheathing stud size type;
area ratio track size thickness; type; exterior spacing; HB size
hxL h/L r stud spacing type orientation interior spacing SB size size type; number kN/m
?
C64x41x?x1.88
T93-1 1 24.31
U64x88x1.88
610 PLY;9.5;V

T93-2 2 SC4.8x?; 102; 305 15.95


?
C89x41x?x150
U89x41x1.50
610
T93-3 3 OSB;11.1;V 18.21

T93-4 4 M 2440x2440 1.00 1.00 ? SC4.2x?; 152; 305 - - - ? ? 10.92


PLY;9.5;V
?
C89x41x?x1.19
T93-5 5 SC4.2x?; 102; 305 14.01
U89x41x1.19
610

T93-6 6 OSB;11.1;V SC4.2x?; 76; 305 15.98

T93-7 7 OSB;15.1;V PI3.7; 152; 305 15.88


?
C64x41x?x1,88
U64x88x1.88
610
T93-8 8 PLY;15.9;V PI3.7; 102; 305 27.22

length in mm; ?: not known; -: not present


(1) M: monotonic; C: cyclic
(2) h: heigth of the wall; L: length of the wall
(3) r =1 / [1 + Ao / (h 6Li)] with Ao: area of openings and Li the length of the full height wall segment
(4) C__x__x__x__ (lipped channel section) web depth x flange size x lip size x thickness; U__x__x__ (unlipped channel section) web depth x flange size x thickness; *: double back-to-back coupled end studs
(5) BHSC: bugle-head screws FHSC: flat head screws; LPHSC: low profile head screws; MTHSC: modified truss head screws; WHSC: wafer head screws; __x__: nominal diameter x length; NA: Nails diameter; PI: Pins diameter
(6) GSB: Gypsum sheathing board; GWB: Gypsum wallboard; FB: Fiber board; OSB: Oriented strand board; PLY: Plywood; SCS: Steel corrugated seet; SSS: Steel sheet sheathing;
V: parallel to the frame (vertical); O: orthogonal to the frame (horizontal)
(7) I__x__ flat strap wide x thickness
(8) BCA: bolted clip angles; PCA: clip angles fixed wit powder actuated fasteners; HD: hold-down; SHD: strip-hold down
(9) F: foundation uplift failure; S: stud buckling failure; X: X bracing yielding;
C (FPO): connections failure (fasteners pull out failure); C (FPT): connections failure (fasteners pull throgh failure); C (FS): connections failure (fasteners shear failure); C (FT): connections failure (fasteners tilting failure); C (NS): connections faiO): c
(net section failure);
Appendix A
Serrette (1994)

Type of aspect horizontal straps (HS) Type of Type of


loading wall size ratio opening frame frame fasteners sheathing sheathing fasteners solid blocking (SB) X-bracing X-bracing fasteners anchorage failure ultimate
Label Test (1) (2) (2) (3) (4) (5) (6) (5) (7) (7) (5) (8) (9) shear
steel grade
sheathing stud size type;
area ratio track size thickness; type; exterior spacing; HB size
hxL h/L r stud spacing type orientation interior spacing SB size size type; number kN/m

S94-1 1 - - I51x0.84 ? 4.42

S94-2 2 GWB;12.7;V SC3.5x?; 152; 305 - - 10.92


+
GWB;12.7;V SC3.5x?; 152; 305
S94-3 3 - I51x0.84 ? 13.56
Summary of existing experime ntal results

S94-4 4 A653 Grade SQ 33 SC3.5x?; 152; 305 15.31


C152x41x?x0.84*
S94-5 5 M 2440x2440 1.00 1.00 ? PLY;11.9;V PI2.9: 152; 305 ? ? 9.06
U152x32x0.84
S94-6 6 610 14.24
S94-7 7 PLY;11.9;O SC4.2x?; 152; 305 6.14
HB ?
S94-8 8 - - 14.30
SB ?
S94-9 9 PI2.9: 152; 305 - 8.76
S94-10 10 11.50
SC4.2x?; 152; 305 HB ?
S94-11 11 OSB;11.1;V 12.08
SB ?
S94-12 12 SC3.5x?; 152; 305 - 4.63
length in mm; ?: not known; -: not present
(1) M: monotonic; C: cyclic
(2) h: heigth of the wall; L: length of the wall
(3) r =1 / [1 + Ao / (h 6Li)] with Ao: area of openings and Li the length of the full height wall segment
(4) C__x__x__x__ (lipped channel section) web depth x flange size x lip size x thickness; U__x__x__ (unlipped channel section) web depth x flange size x thickness; *: double back-to-back coupled end studs
(5) BHSC: bugle-head screws FHSC: flat head screws; LPHSC: low profile head screws; MTHSC: modified truss head screws; WHSC: wafer head screws; __x__: nominal diameter x length; NA: Nails diameter; PI: Pins diameter
(6) GSB: Gypsum sheathing board; GWB: Gypsum wallboard; FB: Fiber board; OSB: Oriented strand board; PLY: Plywood; SCS: Steel corrugated seet; SSS: Steel sheet sheathing;
V: parallel to the frame (vertical); O: orthogonal to the frame (horizontal)
(7) I__x__ flat strap wide x thickness
(8) BCA: bolted clip angles; PCA: clip angles fixed wit powder actuated fasteners; HD: hold-down; SHD: strip-hold down
(9) F: foundation uplift failure; S: stud buckling failure; X: X bracing yielding;
C (FPO): connections failure (fasteners pull out failure); C (FPT): connections failure (fasteners pull throgh failure); C (FS): connections failure (fasteners shear failure); C (FT): connections failure (fasteners tilting failure); C (NS): connections faiO): c
(net section failure);
A-5
A-6

Serrette & Ogunfunmi (1996)

Type of aspect horizontal straps (HS) Type of Type of


loading wall size ratio opening frame frame fasteners sheathing sheathing fasteners solid blocking (SB) X-bracing X-bracing fasteners anchorage failure ultimate
Label Test (1) (2) (2) (3) (4) (5) (6) (5) (7) (7) (5) (8) (9) shear
steel grade
sheathing stud size type;
area ratio track size thickness; type; exterior spacing; HB size
hxL h/L r stud spacing type orientation interior spacing SB size size type; number kN/m
A-P;
SO96-1 A-2; - - I51x0.84 WHSC4.2x13; 14 X 4.95
A-3
B-P;
A446 Grade A
B-2;
C152x32x?x0.84*
SO96-2 B-3; M 2440x2440 1.00 1.00 WHSC4.2x13 - - - BCA 10.47
U152x?x0.84
B-4;
610
B-5
C-1;
BHSC3.5x25; 152; 305
C-2; GSB;12.7;V
SO96-3 I51x0.84 WHSC4.2x13; 14 C (FT) 14.06
C-3; GWB;12.7;V
BHSC3.5x25; 152; 305
C-4
I51x0.84 WHSC4.2x13; 14
SO96-4 C-5 + 19.02
I51x0.84 WHSC4.2x13; 14
length in mm; ?: not known; -: not present
(1) M: monotonic; C: cyclic
(2) h: heigth of the wall; L: length of the wall
(3) r =1 / [1 + Ao / (h 6Li)] with Ao: area of openings and Li the length of the full height wall segment
(4) C__x__x__x__ (lipped channel section) web depth x flange size x lip size x thickness; U__x__x__ (unlipped channel section) web depth x flange size x thickness; *: double back-to-back coupled end studs
(5) BHSC: bugle-head screws FHSC: flat head screws; LPHSC: low profile head screws; MTHSC: modified truss head screws; WHSC: wafer head screws; __x__: nominal diameter x length; NA: Nails diameter; PI: Pins diameter
(6) GSB: Gypsum sheathing board; GWB: Gypsum wallboard; FB: Fiber board; OSB: Oriented strand board; PLY: Plywood; SCS: Steel corrugated seet; SSS: Steel sheet sheathing;
V: parallel to the frame (vertical); O: orthogonal to the frame (horizontal)
(7) I__x__ flat strap wide x thickness
(8) BCA: bolted clip angles; PCA: clip angles fixed wit powder actuated fasteners; HD: hold-down; SHD: strip-hold down
(9) F: foundation uplift failure; S: stud buckling failure; X: X bracing yielding;
C (FPO): connections failure (fasteners pull out failure); C (FPT): connections failure (fasteners pull throgh failure); C (FS): connections failure (fasteners shear failure); C (FT): connections failure (fasteners tilting failure); C (NS): connections faiO): c
(net section failure);
Appendix A
Serrette et al. (1996a)

Type of aspect horizontal straps (HS) Type of Type of


loading wall size ratio opening frame frame fasteners sheathing sheathing fasteners solid blocking (SB) X-bracing X-bracing fasteners anchorage failure ultimate
Label Test (1) (2) (2) (3) (4) (5) (6) (5) (7) (7) (5) (8) (9) shear
steel grade
sheathing stud size type;
area ratio track size thickness; type; exterior spacing; HB size
hxL h/L r stud spacing type orientation interior spacing SB size size type; number kN/m
1A6,
S+96a-1 PLY;11.9;V 15.50
1A7 -
1A2;
S+96a-2 2440x2440 1.00 OSB;11.1;V 13.29
1A3 FHSC4.2x25; 152; 305
1A5; HB I38x0.84
S+96a-3 14.91
1A6 OSB;11.1;O SB ?
1E1;
S+96a-4 14.96
1E2
1D3;
S+96a-5 FHSC4.2x25; 102; 305 20.61
1D4
1D5;
Summary of existing experime ntal results

S+96a-6 OSB;11.1;V FHSC4.2x25; 76; 305 25.33


1D6
1D7;
S+96a-7 FHSC4.2x25; 51; 305 27.90
1D8 A653 Grade SQ 33
C89x41x10x0.84* BHSC3.5x32; 178; 178
1F1; C89x32x0.84
S+96a-8 M 2440x1220 2.00 1.00 WHSC4.2x13 - - - HD C (FT) 17.75
1F2 610 FHSC4.2x25; 152; 305
GWB;12.7;V BHSC3.5x32; 178; 178
1F3;
S+96a-9 + 22.77
1F4
OSB;11.1;V FHSC4.2x25; 102; 305
BHSC3.5x32; 178; 178
1F5;
S+96a-10 27.49
1F6
FHSC4.2x25; 51; 305

2A1; GWB;12.7;O BHSC3.5x32; 178; 178


S+96a-11 + HB I38x0.84 8.51
1A3
2440x2440 1.00 GWB;12.7;O BHSC3.5x32; 178; 178 SB ?
BHSC3.5x32; 102; 102
2A2;
S+96a-12 12.39
2A4
BHSC3.5x32; 102; 102
length in mm; ?: not known; -: not present
(1) M: monotonic; C: cyclic
(2) h: heigth of the wall; L: length of the wall
(3) r =1 / [1 + Ao / (h 6Li)] with Ao: area of openings and Li the length of the full height wall segment
(4) C__x__x__x__ (lipped channel section) web depth x flange size x lip size x thickness; U__x__x__ (unlipped channel section) web depth x flange size x thickness; *: double back-to-back coupled end studs
(5) BHSC: bugle-head screws FHSC: flat head screws; LPHSC: low profile head screws; MTHSC: modified truss head screws; WHSC: wafer head screws; __x__: nominal diameter x length; NA: Nails diameter; PI: Pins diameter
(6) GSB: Gypsum sheathing board; GWB: Gypsum wallboard; FB: Fiber board; OSB: Oriented strand board; PLY: Plywood; SCS: Steel corrugated seet; SSS: Steel sheet sheathing;
V: parallel to the frame (vertical); O: orthogonal to the frame (horizontal)
(7) I__x__ flat strap wide x thickness
(8) BCA: bolted clip angles; PCA: clip angles fixed wit powder actuated fasteners; HD: hold-down; SHD: strip-hold down
(9) F: foundation uplift failure; S: stud buckling failure; X: X bracing yielding;
C (FPO): connections failure (fasteners pull out failure); C (FPT): connections failure (fasteners pull throgh failure); C (FS): connections failure (fasteners shear failure); C (FT): connections failure (fasteners tilting failure); C (NS): connections faiO): c
(net section failure);
A-7
A-8

Serrette et al. (1996a,b)

Type of aspect horizontal straps (HS) Type of Type of


loading wall size ratio opening frame frame fasteners sheathing sheathing fasteners solid blocking (SB) X-bracing X-bracing fasteners anchorage failure ultimate
Label Test (1) (2) (2) (3) (4) (5) (6) (5) (7) (7) (5) (8) (9) shear
steel grade
sheathing stud size type;
area ratio track size thickness; type; exterior spacing; HB size
hxL h/L r stud spacing type orientation interior spacing SB size size type; number kN/m
OSB1;
S+96ab-1 FHSC4.2x25; 152; 305 10.22
OSB2
OSB3;
S+96ab-2 FHSC4.2x25; 102; 305 13.32
OSB4 OSB;11.1;V
OSB5;
S+96ab-3 FHSC4.2x25; 76; 305 18.61
OSB6
OSB7; A653 Grade SQ 33
S+96ab-4 C 2440x1220 2.00 1.00 WHSC4.2x13 FHSC4.2x25; 51; 305 - - - HD C (FT) 24.81
OSB8 C89x41x10x0.84*
PLY1; C89x32x0.84
S+96ab-5 610 FHSC4.2x25; 152; 305 11.38
PLY2
PLY3;
S+96ab-6 FHSC4.2x25; 102; 305 14.41
PLY4 PLY;11.9;V
PLY5;
S+96ab-7 FHSC4.2x25; 76; 305 21.34
PLY6
PLY7;
S+96ab-8 FHSC4.2x25; 51; 305 23.71
PLY8
length in mm; ?: not known; -: not present
(1) M: monotonic; C: cyclic
(2) h: heigth of the wall; L: length of the wall
(3) r =1 / [1 + Ao / (h 6Li)] with Ao: area of openings and Li the length of the full height wall segment
(4) C__x__x__x__ (lipped channel section) web depth x flange size x lip size x thickness; U__x__x__ (unlipped channel section) web depth x flange size x thickness; *: double back-to-back coupled end studs
(5) BHSC: bugle-head screws FHSC: flat head screws; LPHSC: low profile head screws; MTHSC: modified truss head screws; WHSC: wafer head screws; __x__: nominal diameter x length; NA: Nails diameter; PI: Pins diameter
(6) GSB: Gypsum sheathing board; GWB: Gypsum wallboard; FB: Fiber board; OSB: Oriented strand board; PLY: Plywood; SCS: Steel corrugated seet; SSS: Steel sheet sheathing;
V: parallel to the frame (vertical); O: orthogonal to the frame (horizontal)
(7) I__x__ flat strap wide x thickness
(8) BCA: bolted clip angles; PCA: clip angles fixed wit powder actuated fasteners; HD: hold-down; SHD: strip-hold down
(9) F: foundation uplift failure; S: stud buckling failure; X: X bracing yielding;
C (FPO): connections failure (fasteners pull out failure); C (FPT): connections failure (fasteners pull throgh failure); C (FS): connections failure (fasteners shear failure); C (FT): connections failure (fasteners tilting failure); C (NS): connections faiO): c
(net section failure);
Appendix A
Serrette et al. (1997a)

Type of aspect horizontal straps (HS) Type of Type of


loading wall size ratio opening frame frame fasteners sheathing sheathing fasteners solid blocking (SB) X-bracing X-bracing fasteners anchorage failure ultimate
Label Test (1) (2) (2) (3) (4) (5) (6) (5) (7) (7) (5) (8) (9) shear
steel grade
sheathing stud size type;
area ratio track size thickness; type; exterior spacing; HB size
hxL h/L r stud spacing type orientationinterior spacing SB size size type; number kN/m
S+97a-1 PLY-T1 BHSC3.5x25; 152; 305 C (FS) 15.02
S+97a-2 PLY-T2-N PLY;11.9;V NA3.7; 152; 305 - 9.06
S+97a-3 PLY-T4 15.62
S+97a-4 PLY-T6 BHSC4.2x32; 152 ;305 C (FPO) 6.74
PLY;11.9;O HB I31x0.84
S+97a-5 PLY-T7 15.62
SB ?
PLY;11.9;V NA3.7; 152; 305
S+97a-6 PLYGYP + C (FS) 18.43
GWB;12.7;V BHSC3.5X25; 178; 178 -
S+97a-7 OSB-T1-N OSB;11.1;V NA3.7; 152; 305 8.86
S+97a-8 OSB-T4 BHSC4.2x32; 152 ;305 C (FT) 13.81
S+97a-9 OSB-T5 OSB;11.1;O 14.30
Summary of existing experime ntal results

HB I31x0.84
GWB;12.7;O SB ?
S+97a-10 GYP-T1 M 2440x2440 1.00 1.00 A446 Grade A WHSC4.2x13 + BHSC3.5x25; 152 ;305 - - HD 9.87
C152x41x10x0.84* GWB;12.7;O
U152x25x0.84 BHSC3.5x25; 152 ;305
S+97a-11 GYP-T2 610 11.79

GWB;12.7;V BHSC3.5x25; 178 ;178


S+97a-12 GYP-T3 + 10.68
GWB;12.7;V BHSC3.5x25; 178 ;178
BHSC3.5x25; 102 ;102
S+97a-13 GYP-T4 C (FS) 14.71
BHSC3.5x25; 102 ;102
S+97a-14 FB-T4 FB;12.7;V BHSC3.5x25; 152 ;305 - 5.63
S+97a-15 FB-T5 BHSC3.5x25; 102 ;305 6.04
BHSC3.5x32; 152 ;305
S+97a-16 FB-T6 FB;12.7;V 10.83
+ BHSC3.5x32; 152 ;305
FB;12.7;V BHSC3.5x32; 102 ;305
S+97a-17 FB-T7 15.21
BHSC3.5x32; 102 ;305
S+97a-18 FB-T8 FB;12.7;V BHSC3.5x32; 102 ;305 7.15
length in mm; ?: not known; -: not present
(1) M: monotonic; C: cyclic
(2) h: heigth of the wall; L: length of the wall
(3) r =1 / [1 + Ao / (h 6Li)] with Ao: area of openings and Li the length of the full height wall segment
(4) C__x__x__x__ (lipped channel section) web depth x flange size x lip size x thickness; U__x__x__ (unlipped channel section) web depth x flange size x thickness; *: double back-to-back coupled end studs
(5) BHSC: bugle-head screws FHSC: flat head screws; LPHSC: low profile head screws; MTHSC: modified truss head screws; WHSC: wafer head screws; __x__: nominal diameter x length; NA: Nails diameter; PI: Pins diameter
(6) GSB: Gypsum sheathing board; GWB: Gypsum wallboard; FB: Fiber board; OSB: Oriented strand board; PLY: Plywood; SCS: Steel corrugated seet; SSS: Steel sheet sheathing;
V: parallel to the frame (vertical); O: orthogonal to the frame (horizontal)
(7) I__x__ flat strap wide x thickness
(8) BCA: bolted clip angles; PCA: clip angles fixed wit powder actuated fasteners; HD: hold-down; SHD: strip-hold down
(9) F: foundation uplift failure; S: stud buckling failure; X: X bracing yielding;
C (FPO): connections failure (fasteners pull out failure); C (FPT): connections failure (fasteners pull throgh failure); C (FS): connections failure (fasteners shear failure); C (FT): connections failure (fasteners tilting failure); C (NS): connections faiO): c
(net section failure);
A-9
A-10

Serrette et al. (1997b) - Monotonic tests

Type of aspect horizontal straps (HS) Type of Type of


loading wall size ratio opening frame frame fasteners sheathing sheathing fasteners solid blocking (SB) X-bracing X-bracing fasteners anchorage failure ultimate
Label Test (1) (2) (2) (3) (4) (5) (6) (5) (7) (7) (5) (8) (9) shear
steel grade
sheathing stud size type;
area ratio track size thickness; type; exterior spacing; HB size
hxL h/L r stud spacing type orientation interior spacing SB size size type; number kN/m
A653 Grade SQ 33
1; C89x43x13x0.84*
S+97b-1 I114x0.84 MTHSC4.2x13; 20 9.82
2 U89x32x0.84
2440x1220 2.00 610 - -
A653 Grade SQ 33
3; C89x43x13x1.09*
S+97b-2 I191x0.84 MTHSC4.2x13; 30 12.80
4 U89x32x1.09
610 S
5;
S+97b-3 FHSC4.2x25; 152; 305 7.84
6
7;
S+97b-4 M 1.00 MTHSC4.2x13 OSB;11.1;V FHSC4.2x25; 102; 305 - HD 14.97
8
9; A653 Grade SQ 33
S+97b-5 2440x610 4.00 FHSC4.2x25; 51; 305 26.63
10 C89x43x13x0.84* - -
11; U89x32x0.84
S+97b-6 610 SSS;0.46;? MTHSC4.2x13; 152; 305 7.17
12
13;
S+97b-7 SSS;0.69;? MTHSC4.2x13; 102; 305 C (NS) 14.45
14
15;
S+97b-8 2440x1220 2.00 SSS;0.46;? MTHSC4.2x13; 152; 305 7.05
16
length in mm; ?: not known; -: not present
(1) M: monotonic; C: cyclic
(2) h: heigth of the wall; L: length of the wall
(3) r =1 / [1 + Ao / (h 6Li)] with Ao: area of openings and Li the length of the full height wall segment
(4) C__x__x__x__ (lipped channel section) web depth x flange size x lip size x thickness; U__x__x__ (unlipped channel section) web depth x flange size x thickness; *: double back-to-back coupled end studs
(5) BHSC: bugle-head screws FHSC: flat head screws; LPHSC: low profile head screws; MTHSC: modified truss head screws; WHSC: wafer head screws; __x__: nominal diameter x length; NA: Nails diameter; PI: Pins diameter
(6) GSB: Gypsum sheathing board; GWB: Gypsum wallboard; FB: Fiber board; OSB: Oriented strand board; PLY: Plywood; SCS: Steel corrugated seet; SSS: Steel sheet sheathing;
V: parallel to the frame (vertical); O: orthogonal to the frame (horizontal)
(7) I__x__ flat strap wide x thickness
(8) BCA: bolted clip angles; PCA: clip angles fixed wit powder actuated fasteners; HD: hold-down; SHD: strip-hold down
(9) F: foundation uplift failure; S: stud buckling failure; X: X bracing yielding;
C (FPO): connections failure (fasteners pull out failure); C (FPT): connections failure (fasteners pull throgh failure); C (FS): connections failure (fasteners shear failure); C (FT): connections failure (fasteners tilting failure); C (NS): connections faiO): c
(net section failure);
Appendix A
Serrette et al. (1997b) - Cyclic tests

Type of aspect horizontal straps (HS) Type of Type of


loading wall size ratio opening frame frame fasteners sheathing sheathing fasteners solid blocking (SB) X-bracing X-bracing fasteners anchorage failure ultimate
Label Test (1) (2) (2) (3) (4) (5) (6) (5) (7) (7) (5) (8) (9) shear
steel grade
sheathing stud size type;
area ratio track size thickness; type; exterior spacing; HB size
hxL h/L r stud spacing type orientation interior spacing SB size size type; number kN/m
A1;
S+97b-9 A653 Grade SQ 33 PLY;11.9;V FHSC4.2x25; 76; 305 C 25.90
A2
A3; C89x43x13x0,84 int.
S+97b-10 FHSC4.2x25; 51; 305 31.96
A4 C89x43x13x1.09* ends
A5; U89x32x0.84
S+97b-11 OSB;11.1;V FHSC4.2x25; 76; 305 C (FT) 22.23
A6 610
A7;
S+97b-12 FHSC4.2x25; 51; 305 30.03
A8
A653 Grade SQ 33 - -
B1; C89x43x13x1.09*
S+97b-13 2440x1220 2.00 C (FT) 13.02
B2 U89x32x1.09
Summary of existing experime ntal results

610 PLY;11.9;V FHSC4.2x25; 152; 305


A653 Grade SQ 33
B3; C89x43x13x1.37*
S+97b-14 C 1.00 MTHSC4.2x13 - HD C (FS) 13.19
B4 U89x32x1.37
610
C1;
S+97b-15 - - I114x0.84 MTHSC4.2x13; 20 S 11.98
C2
C3;
S+97b-16 I191x0.84 MTHSC4.2x13; 30 12.24
C4
D1;
S+97b-17 SSS;0.46;? MTHSC4.2x13; 152; 305 C (NS) 5.72
D2
E1; A653 Grade SQ 33
S+97b-18 FHSC4.2x25; 152; 305 S 10.52
E2 C89x43x13x0.84*
E3; U89x32x0.84
S+97b-19 OSB;11.1;V FHSC4.2x25; 102; 305 - - 16.43
E4 610
E5;
S+97b-20 2440x610 4.00 FHSC4.2x25; 51; 305 23.70
E6
F1;
S+97b-21 MTHSC4.2x13; 102; 305 14.64
F2 SSS;0.69;? S+C
F3; (FPO+NS)
S+97b-22 MTHSC4.2x13; 51; 305 17.09
F4
length in mm; ?: not known; -: not present
(1) M: monotonic; C: cyclic
(2) h: heigth of the wall; L: length of the wall
(3) r =1 / [1 + Ao / (h 6Li)] with Ao: area of openings and Li the length of the full height wall segment
(4) C__x__x__x__ (lipped channel section) web depth x flange size x lip size x thickness; U__x__x__ (unlipped channel section) web depth x flange size x thickness; *: double back-to-back coupled end studs
(5) BHSC: bugle-head screws FHSC: flat head screws; LPHSC: low profile head screws; MTHSC: modified truss head screws; WHSC: wafer head screws; __x__: nominal diameter x length; NA: Nails diameter; PI: Pins diameter
(6) GSB: Gypsum sheathing board; GWB: Gypsum wallboard; FB: Fiber board; OSB: Oriented strand board; PLY: Plywood; SCS: Steel corrugated seet; SSS: Steel sheet sheathing;
V: parallel to the frame (vertical); O: orthogonal to the frame (horizontal)
(7) I__x__ flat strap wide x thickness
(8) BCA: bolted clip angles; PCA: clip angles fixed wit powder actuated fasteners; HD: hold-down; SHD: strip-hold down
(9) F: foundation uplift failure; S: stud buckling failure; X: X bracing yielding;
C (FPO): connections failure (fasteners pull out failure); C (FPT): connections failure (fasteners pull throgh failure); C (FS): connections failure (fasteners shear failure); C (FT): connections failure (fasteners tilting failure); C (NS): connections faiO): c
(net section failure);
A-11
A-12

NAHB (1997)

Type of aspect horizontal straps (HS) Type of Type of


loading wall size ratio opening frame frame fasteners sheathing sheathing fasteners solid blocking (SB) X-bracing X-bracing fasteners anchorage failure ultimate
Label Test (1) (2) (2) (3) (4) (5) (6) (5) (7) (7) (5) (8) (9) shear
steel grade
sheathing stud size type;
area ratio track size thickness; type; exterior spacing; HB size
hxL h/L r stud spacing type orientation interior spacing SB size size type; number kN/m

N97-1 1 1.00 15.40


HD C (FPT)
? GWB;12.7;V SC3.5x?; 178; 254
N97-2 2A 0.76 9.56
M 2440x12190 0.20 C89x38x?x0.84 SC4.2x? + - - -
U?x?x? OSB;11.1;V SC4.2x?; 152; 305
N97-3 2B 0.76 610 - F 4.56

N97-4 4 0.48 HD C (FPT) 4.78

length in mm; ?: not known; -: not present


(1) M: monotonic; C: cyclic
(2) h: heigth of the wall; L: length of the wall
(3) r =1 / [1 + Ao / (h 6Li)] with Ao: area of openings and Li the length of the full height wall segment
(4) C__x__x__x__ (lipped channel section) web depth x flange size x lip size x thickness; U__x__x__ (unlipped channel section) web depth x flange size x thickness; *: double back-to-back coupled end studs
(5) BHSC: bugle-head screws FHSC: flat head screws; LPHSC: low profile head screws; MTHSC: modified truss head screws; WHSC: wafer head screws; __x__: nominal diameter x length; NA: Nails diameter; PI: Pins diameter
(6) GSB: Gypsum sheathing board; GWB: Gypsum wallboard; FB: Fiber board; OSB: Oriented strand board; PLY: Plywood; SCS: Steel corrugated seet; SSS: Steel sheet sheathing;
V: parallel to the frame (vertical); O: orthogonal to the frame (horizontal)
(7) I__x__ flat strap wide x thickness
(8) BCA: bolted clip angles; PCA: clip angles fixed wit powder actuated fasteners; HD: hold-down; SHD: strip-hold down
(9) F: foundation uplift failure; S: stud buckling failure; X: X bracing yielding;
C (FPO): connections failure (fasteners pull out failure); C (FPT): connections failure (fasteners pull throgh failure); C (FS): connections failure (fasteners shear failure); C (FT): connections failure (fasteners tilting failure); C (NS): connections faiO): c
(net section failure);
Appendix A
Selenikovich et al. (1999)

Type of aspect horizontal straps (HS) Type of Type of


loading wall size ratio opening frame frame fasteners sheathing sheathing fasteners solid blocking (SB) X-bracing X-bracing fasteners anchorage failure ultimate
Label Test (1) (2) (2) (3) (4) (5) (6) (5) (7) (7) (5) (8) (9) shear
steel grade
sheathing stud size type;
area ratio track size thickness; type; exterior spacing; HB size
hxL h/L r stud spacing type orientation interior spacing SB size size type; number kN/m
A monotonic
S+99-1 without M 11.85
GWB
A cyclic OSB;11.1;V BHSC4.2x?; 152 ;305
stabilized
S+99-2 C 1.00 7.91
without
GWB
GWB;12.7;V BHSC4.2x?; 178; 254
Summary of existing experime ntal results

A monotonic
S+99-3 M + 14.69
with GWB
OSB;11.1;V BHSC4.2x?; 152; 305
S+99-4 B monotonic M 7.55
0.76 350S150-33
B cyclic
S+99-5 C 2440x12200 0.20 C89x38x?x0.84* LPHSC4.2x? - - - HD C (FPT) 6.38
stabilized
U89x38x0.84
S+99-6 C monotonic M 5.07
C cyclic 0.56
S+99-7 C 4.27
stabilized
S+99-8 D monotonic M OSB;11.1;V BHSC4.2x?; 152 ;305 4.67
D cyclic 0.48
S+99-9 C 3.68
stabilized
S+99-10 E monotonic M 2.81
E cyclic 0.30
S+99-11 C 2.04
stabilized
length in mm; ?: not known; -: not present
(1) M: monotonic; C: cyclic
(2) h: heigth of the wall; L: length of the wall
(3) r =1 / [1 + Ao / (h 6Li)] with Ao: area of openings and Li the length of the full height wall segment
(4) C__x__x__x__ (lipped channel section) web depth x flange size x lip size x thickness; U__x__x__ (unlipped channel section) web depth x flange size x thickness; *: double back-to-back coupled end studs
(5) BHSC: bugle-head screws FHSC: flat head screws; LPHSC: low profile head screws; MTHSC: modified truss head screws; WHSC: wafer head screws; __x__: nominal diameter x length; NA: Nails diameter; PI: Pins diameter
(6) GSB: Gypsum sheathing board; GWB: Gypsum wallboard; FB: Fiber board; OSB: Oriented strand board; PLY: Plywood; SCS: Steel corrugated seet; SSS: Steel sheet sheathing;
V: parallel to the frame (vertical); O: orthogonal to the frame (horizontal)
(7) I__x__ flat strap wide x thickness
(8) BCA: bolted clip angles; PCA: clip angles fixed wit powder actuated fasteners; HD: hold-down; SHD: strip-hold down
(9) F: foundation uplift failure; S: stud buckling failure; X: X bracing yielding;
C (FPO): connections failure (fasteners pull out failure); C (FPT): connections failure (fasteners pull throgh failure); C (FS): connections failure (fasteners shear failure); C (FT): connections failure (fasteners tilting failure); C (NS): connections faiO): c
(net section failure);
A-13
A-14

Fulop & Dubina (2002)

Type of aspect horizontal straps (HS) Type of Type of


loading wall size ratio opening frame frame fasteners sheathing sheathing fasteners solid blocking (SB) X-bracing X-bracing fasteners anchorage failure ultimate
Label Test (1) (2) (2) (3) (4) (5) (6) (5) (7) (7) (5) (8) (9) shear
steel grade
sheathing stud size type;
area ratio track size thickness; type; exterior spacing; HB size
hxL h/L r stud spacing type orientation interior spacing SB size size type; number kN/m
FD02-1 O M - - - 0.28

FD02-2 I M SCS;0.5;O SC4.8x22; 114; 229 14.69

FD02-3 I C 12.68

SCS;0.5;O
FD02-4 II M SC4.8x22; 114; 229 - - 16.59
GWB;12,T;V
SC4.8x22; 250; 250
SCS;0,5;O
FD02-5 II C 1.00 15.95
GWB;12.5;V

?
C150x?x?x1.5*
FD02-6 III M 2440x3600 0.68 ? - ? 15.32
u154x?x1.5 I110x1,5
600 - - + ?
I110x1,5
FD02-7 III C F 14.87

FD02-8 IV M 0.70 SCS;0.5;O SC4.8x22; 114; 229 11.17

FD02-9 IV C 0.70 10.49

FD02-10 OSB I M 1.00 - - 21.88

FD02-11 OSB I C 1.00 OSB;10;V BHSC4.8x22; 105; 250 19.40

FD02-12 OSB II M 0.70 12.33

FD02-13 OSB II C 0.70 12.76


length in mm; ?: not known; -: not present
(1) M: monotonic; C: cyclic
(2) h: heigth of the wall; L: length of the wall
(3) r =1 / [1 + Ao / (h 6Li)] with Ao: area of openings and Li the length of the full height wall segment
(4) C__x__x__x__ (lipped channel section) web depth x flange size x lip size x thickness; U__x__x__ (unlipped channel section) web depth x flange size x thickness; *: double back-to-back coupled end studs
(5) BHSC: bugle-head screws FHSC: flat head screws; LPHSC: low profile head screws; MTHSC: modified truss head screws; WHSC: wafer head screws; __x__: nominal diameter x length; NA: Nails diameter; PI: Pins diameter
(6) GSB: Gypsum sheathing board; GWB: Gypsum wallboard; FB: Fiber board; OSB: Oriented strand board; PLY: Plywood; SCS: Steel corrugated seet; SSS: Steel sheet sheathing;
V: parallel to the frame (vertical); O: orthogonal to the frame (horizontal)
(7) I__x__ flat strap wide x thickness
(8) BCA: bolted clip angles; PCA: clip angles fixed wit powder actuated fasteners; HD: hold-down; SHD: strip-hold down
(9) F: foundation uplift failure; S: stud buckling failure; X: X bracing yielding;
C (FPO): connections failure (fasteners pull out failure); C (FPT): connections failure (fasteners pull throgh failure); C (FS): connections failure (fasteners shear failure); C (FT): connections failure (fasteners tilting failure); C (NS): connections faiO): c
(net section failure);
Appendix A
Branston et al. (2003)

Type of aspect horizontal straps (HS) Type of Type of


loading wall size ratio opening frame frame fasteners sheathing sheathing fasteners solid blocking (SB) X-bracing X-bracing fasteners anchorage failure ultimate
Label Test (1) (2) (2) (3) (4) (5) (6) (5) (7) (7) (5) (8) (9) shear
steel grade
sheathing stud size type;
area ratio track size thickness; type; exterior spacing; HB size
hxL h/L r stud spacing type orientation interior spacing SB size size type; number kN/m
Summary of existing experime ntal results

OSB 4-8 US A653 Grade SQ 33


M 16.80
B+03-1 M- A, B, C 2440x1220 2.00 C89x41x10x0.84* OSB;11.1;V FHSC4.2x25; 102; 305
OSB 4-8 US U89x38x0.84
C 610 16.00
B+03-2 C- A, B, C 1.00 WHSC4.2x13 - - - HD C
PLY 8-8 US A653 Grade SQ 33
M 16.40
B+03-3 M- A, B, C 2440x2440 1.00 C89x41x10x0.84* PLY;11.9;V FHSC4.2x25; 152; 305
PLY 8-8 US U89x38x0.84
C 610 11.80
B+03-4 C- A, B, C
length in mm; ?: not known; -: not present
(1) M: monotonic; C: cyclic
(2) h: heigth of the wall; L: length of the wall
(3) r =1 / [1 + Ao / (h 6Li)] with Ao: area of openings and Li the length of the full height wall segment
(4) C__x__x__x__ (lipped channel section) web depth x flange size x lip size x thickness; U__x__x__ (unlipped channel section) web depth x flange size x thickness; *: double back-to-back coupled end studs
(5) BHSC: bugle-head screws FHSC: flat head screws; LPHSC: low profile head screws; MTHSC: modified truss head screws; WHSC: wafer head screws; __x__: nominal diameter x length; NA: Nails diameter; PI: Pins diameter
(6) GSB: Gypsum sheathing board; GWB: Gypsum wallboard; FB: Fiber board; OSB: Oriented strand board; PLY: Plywood; SCS: Steel corrugated seet; SSS: Steel sheet sheathing;
V: parallel to the frame (vertical); O: orthogonal to the frame (horizontal)
(7) I__x__ flat strap wide x thickness
(8) BCA: bolted clip angles; PCA: clip angles fixed wit powder actuated fasteners; HD: hold-down; SHD: strip-hold down
(9) F: foundation uplift failure; S: stud buckling failure; X: X bracing yielding;
C (FPO): connections failure (fasteners pull out failure); C (FPT): connections failure (fasteners pull throgh failure); C (FS): connections failure (fasteners shear failure); C (FT): connections failure (fasteners tilting failure); C (NS): connections faiO): c
(net section failure);
A-15
B -1

Appendix B

Summary of objectives and results of


existing experimental studies
B-2 Appendix B

Sheathing
Objectives Authors Conclusions
- the use of cement plaster increases the shear strength and the
stiffness
Tarpy (1980)
- the use of two layer of GWB increases the shear capacity, while
decreasing the shear stiffness, in comparison with single layer
- the use of PLY panels increases the shear strength in comparison
with GWB panels
Tarpy & Girard (1982)
- the use of GWB panels increases the shear strength in comparison
with GSB panels
- the PLY panels carry slightly higher loads in comparison with OSB
panels
- walls with OSB panels on one side and GWS panels on the other
side exhibit similar failure behavior, but degrade more gradually than
Serrette et al. (1996a,b) walls with OSB panels on one side alone
- walls with GWB panels on both sides have much lower shear
strength than walls with OSB panels
- PLY and OSB panels show a small difference in the cyclic shear
strength
- the behavior of the PLY and OSB panels is comparable
- the strength of GWB and FBW is relatively low in comparison with
PLY and OSB panels
Serrette et al. (1997a)
- the use of GWB panels on the interior of the wall and PLY panels on
the exterior produce a higher shear capacity, by approximately 18%,
in comparison with the PLY panels only
- the walls sheathed with SSS have a ductile behavior without sudden
decreases in shear load capacity
behaviour of different
Serrette et al. (1997b) - the use of thick sheathings increases the shear resistance, but the
sheathing type
failure mode move from rupture at the edges of the sheathing to screw
pull-out from the framing
- adding GWB panels increases the shear strength and stiffness of
Selenikovich et al. (1999)
fully sheathed walls under monotonic load
- the behavior of GWB is satisfactory. In fact it could follow even
Fulop & Dubina (2002)
extreme deformation of the wall without significant damage
- the use of X-B plus GWB reduces the permanent deflection and
increases the shear strength without decreasing the stiffness
Serrette & Ogunfunmi (1996) - the use of X-B plus GWB is not practical due to the need to
pretension the straps and the need for additional screws to connect
the straps
- in the design of walls with X-B the designer must consider that the
force in the strap may be larger than that corresponding to the nominal
Serrette et al. (1997b) yield strength
- if X-B are installed on one side of the wall only, the effect of
eccentricity should be considered
For walls laterally braced with X-B only:
- the initial tension in the X-B increases the frame stiffness and when
the X-B yield then the true stiffness of the X-B system defines the
stiffness of the frame
- the type of strap bracing-to-plate connections governs the failure
Gad et al. (1999a)
load and mechanism
For walls laterally braced with X-B and plasterboard:
- when X-B and plasterboard are combined, the overall stiffness and
strength of the system is simple addition of individual contributions
from X-B and plasterboard
- sheathings oriented horizontally exhibit slightly higher shear strength
Serrette & Ogunfunmi (1996)
than panels oriented vertically
effect of sheathing - the walls with blocked panels oriented horizontally provide
orientation and essentially the same shear capacity but higher stiffness than a
contribution of blocking Serrette et al. (1997a) comparable wall with panels oriented vertically
- when blocking is omitted from the walls with horizontal panels, the
shear capacity of the wall is reduced by more than 50%
Summary of objectives and resu lts of ex istin g experimental studies B-3

Framing
Objectives Authors Conclusions
- the reduction of the stud spacing does not increase significantly the
Tarpy & Girard (1982)
shear strength and stiffness
effect of framing stud
- the use of thicker and back-to-back coupled end studs for the walls
size, thickness and
with PLY and OSB panels allows to fully develop the shear strength of
spacing Serrette et al. (1997b)
the panels-to-frame connections also in the cases in which dense
fasteners schedules are used

Fastener
Objectives Authors Conclusions
- reduction of the fastener spacing around the wall perimeter
Tarpy & Hauenstein (1978)
increases the shear strength
- reduction of the fastener spacing increases the shear strength and
Tarpy (1980)
stiffness
- the use of welded stud to track connections provides the same shear
Tarpy & Girard (1982)
strength as screw connections
- reduction of the fastener spacing increases significantly the shear
Serrette et al. (1996a,b)
strength
- the use of 3.7mm diameter nails decreases the maximum shear
strength in comparison with No.6 and No.8 screws
- the maximum shear strength is not influenced by the size of the
fastener
- the failure mode for the specimens with No.6 screws is fracture of
effect of fastener type, Serrette et al. (1997a) fasteners and thus walls with No.6 screws may fail at lower loads than
size and spacing walls with No.8 screws
- for GWB and FB panels the reduction of the fastener spacing
increases significantly the shear strength
- decreasing the screws spacing results in the increased maximum
shear load
- No.8 screws should be limited to 1,09mm thick framing; in fact, this
screw behave well in the 0.84 and 1.09mm thick frames (screws pull-
Serrette et al. (1997b)
out or pull-trhough failure) but fractures in shear when 1.37mm thick
frames are used
- for walls laterally braced with X-B only the type of connections
Gad et al. (1999a) between the framing members dos not seem to have an influence on
the structural response of the braced frames
- a reduction of the fastener spacing nonlinearly increases the shear
COLA-UCI (2001) strength and stiffness for both the light-gauge steel framed and wood
framed stud walls
B-4 Appendix B

Geometry
Objectives Authors Conclusions
- the calculation of the shear capacity using the Perforated Shear
NAHB (1997) Wall Design Method appears valid, but reveals a conservative
prediction of ultimate shear strength
- long, fully sheathed walls are significantly stiffer and stronger but
influence of the opening
less ductile than walls with openings
size
- the predictions of the Perforated Shear Wall Design Method are
Selenikovich et al. (1999)
conservative at all levels of monotonic and cyclic loading
- the strength of fully sheathed walls is affected more significantly by
cyclic loading than walls with openings
- for the aspect ratio varying between 0.33 and 1 shear strength is
Mc Creless & Tarpy (1978) independent of H/L ratio but shear stiffness increases for smaller H/L
ratio
effect of height/length
- the shear strength is practically the same for the H/L ratio varying
(H/L) variation Serrette et al. (1996a,b)
between 1 and 2
- the shear strength appreciably decreases when the aspect ratio
Serrette et al. (1997b)
increases from 2 to 4

Type of loading
Objectives Authors Conclusions
Tarpy (1980) - cyclic loading decreases shear strength and damage threshold level
- PLY and OSB panels show a small difference in the cyclic shear
strength
- the shear behavior of walls observed in cyclic tests is somewhat not
Serrette et al. (1996a,b) as good as the shear behavior exhibited in monotonic tests for walls of
similar constructions
- as in the monotonic tests, in the cyclic tests the wall shear strength
increases significantly with the reduction of the fastener spacing
effect of cyclic loading
- for walls laterally braced with X-B only the dynamic characteristics of
Gad et al. (1999a)
the frame are governed by the initial tension in the straps

- cyclic loading does not influence the elastic behavior of the walls but
reduces their deformation capacity
Selenikovich et al. (1999)
- the strength of fully sheathed walls is affected more significantly by
cyclic loading than walls with openings
- very significant pinching and reduced energy dissipation characterize
Fulop & Dubina (2002)
the hysteretic behavior
Summary of objectives and resu lts of ex istin g experimental studies B-5

Construction techniques and anchorage details


Objectives Authors Conclusions
- to avoid uplift failure it is recommended that a positive attachment
Tarpy & Hauenstein (1978)
should be furnished between the track and floor framing systems

- the corner anchorage influences the shear behavior dramatically


- hold-down exhibits higher shear strength than bolt and washer
anchors
- densely spaced powder actuated fasteners (connected to a
Tarpy (1980) supporting concrete beam) provide similar restraint to the hold-down
- the shear resistance dos not vary extensively when using different
types of interior shear anchorage
- the use of a 45 stud placed at bottom corner between the chord
members and the adjacent stud has little effect on the shear capacity

- when the bolt and washer anchorage details are used without clip
effect of construction angles the shear capacity decreases
techniques and - the use of closely spaced powder actuated fasteners negligibly
anchorage details Tarpy & Girard (1982) increases the shear behavior in comparison to using corner clips
- it is recommended the use of clip angles
- it is suggested a rigid attachment to connect the wall panel to the
floor or roof framing systems

- hold-down reduces uplift and increases the ultimate shear capacity


NAHB (1997)
by allowing more sheathing-to-frame screws to resist shear

- in-plane brick veneer walls attached to the frame via clip-on ties dos
not contribute to the stiffness of the system
- different displacements between the frame and the out-of-plane brick
veneer walls are mainly accommodated by deformation of the stud
Gad et al. (1999a)
flanges rather than deformation of brick ties
- the plasterboard combined with ceiling cornices, skirting board and
set corner joints, resists about 60-70% of the applied racking load
whereas the X-B resists 30-40%
B-6 Appendix B

damage threshold level


Objectives Authors Conclusions
- the first noticeable wallboard damage occurrs:
- at about 6 to 13mm total displacement (for the shorter walls)
- at about 6mm (for the longer walls)
Mc Creless & Tarpy (1978)
- the real damage occurrs:
- at about 13 to 19mm total displacement (for the shorter walls)
determination of the
- at about 6 to 13mm (for the longer walls).
damage threshold level
Tarpy & Hauenstein (1978) - a safety factor of 2.0 is recommended
Tarpy (1980) - cyclic loading decreases the damage threshold level
- a safety factor of 2.0 is recommended to determine the design shear
Tarpy & Girard (1982) strength from the ultimate shear strength for the type of stud shear
walls examined

steel-frame versus wood-frame


Objectives Authors Conclusions
- the lateral load resisting mechanism for both W-F and S-F shear
NAHB (1997)
comparison between walls appears to be similar
steel-frame (S-F) and - with the same sheathing types and fastener spacing, S-F walls
wood-frame (W-F) COLA-UCI (2001) exhibit somewhat higher shear strength and ductility but less
hysteretic damping than W-F walls

full-scale tests versus small-scale test


Objectives Authors Conclusions

- the normalized shear strength for small-scale tests is similar as those


comparison between full-
for full-scale tests, thus the small-scale tests are considered to be
scale and small-scale Serrette et al. (1997a)
useful in a evaluation of the relative resistance of different wall
test results
assemblies
C -1

Appendix C

Evaluation of seismic action and wall shear


strength for the study case
C-2 Appendix C

Evaluation of seismic action

Geometry of the house

Number of stories n= 2
Length L= 11.4 m
Wide W= 7.0 m
Height H= 6.4 m
Roof slope Tan(D)= 100 %
Area A= 75 m2
Length of full height wall segments 6wi= 18.0 m

Unit loads

Dead loads
Roof gkR= 0.75 kN/m2
Floor gkF= 0.75 kN/m2
Walls gkW= 0.35 kN/m2

Snow loads qks= 0.50 kN/m2

Live loads qkl= 2.00 kN/m2

Seismic weight

Total seismic weight


Wk,tot = 6Gk + 6\k Qk

Element gk (kN/m2) qk (kN/m2) Area (m2) Gk (kN) Qk (kN) \EI Wk (kN)


Roof 0.75 0.50 106 80 38 0.30 91
Floor 0.75 2.00 75 56 150 0.30 101
Walls 0.35 0.00 90 32 0 32
Wk,tot = 224

Seismic weight per unit surface


wk = Wk,tot / A = 3.0 kN/m2

Seismic weight per unit lenght of full height wall segments


ws = Wk,tot / 6wi = 12.5 kN/m
Ev alu atio n of seismic actio n and wall sh ear streng th fo r th e stud y case C-3

Type 1 (far field) Eurocode 8 Design elastic spectrum

K=1 damping factor (K=1 for a damping ratio Q=0.05)


ag = 0.25g design peak ground acceleration
S: soil factor
TB and TC: limits of constant spectral acceleration branch
TD : value that defines the beginning of the constant displacement response range

Soil type S TB TC TD
A 1 0.15 0.4 2
B 1.2 0.15 0.5 2
C 1.15 0.2 0.6 2
D 1.35 0.2 0.8 2
E 1.4 0.15 0.5 2
ws / g
T 2S = 0.1 - 0.3 s first mode vibration period
k

k= 0.5 - 3.0 kN/mm/m stiffness per unit lenght of full height wall segments

Maximum design elastic spectrum for first mode vibration period ranging from 0.1 to 0.3s

Soil type Saed / g


A 0.63
B 0.75
C 0.72
D 0.84
E 0.88
max 0.88
C-4 Appendix C

Seismic force per unit lenght of full height wall segments

vS S aed w s

Soil type vS (kN/m)


A 7.9
B 9.4
C 9.0
D 10.5
E 11.0
max 11.0
Ev alu atio n of seismic actio n and wall sh ear streng th fo r th e stud y case C-5

Evaluation of strength of walls

Geometry of the wall

Length L= 2400 mm
Height H= 2500 mm
Studs spacing sts= 600 mm

Materials

Steel grade: FeE350G (S350GD+Z/ZF) hot dipped galvanized (zinc coated) steel (EN10147)
(Nominal yield strength f y =350MPa; nominal tensile strength f t =420MPa)

Wood-based panels: Type 3 OSB (KRONOPLY 3 by KRONO FRANCE)

Gypsum-based panels: GWB (PLACOLAST PLACO by BPB ITALIA)


C-6 Appendix C

Failure modes

Frame

Frame-to-foundation
Sheathing-to-frame connections
connections
Ev alu atio n of seismic actio n and wall sh ear streng th fo r th e stud y case C-7

Sheathing-to-frame connection failure [see Section 2.2.1 in Chapter 2 ]

FS  F ,bs D t f d f u ,s

3 .2 t 3f d
0 .5
FS  F ,ts f u ,s d 2 .1

min FS  F
FS  F , ss

t s d u f b ,w
FS  F ,bp 3 .5 C D
KD vS-F

ts= 9.0 mm thickness of OSB sheathings


tf= 1.0 mm thickness of the frame
d= 4.2 mm nominal diameter of screws
du= 2.8 mm unthreaded diameter of screws
fb,w= 38 MPa dowel bearing strength of OSB sheathings
fu,s= 420 Mpa ultimate tensile strength of members of the frame
FS-F,ss= 4.0 kN shear design resistance of screws

Bearing strength of steel frame


FS-F,bs = 3.70 kN (D=2.1) [see Eq. 2.5 in Chapter 2 ]

Tilting strength of steel studs


FS-F,ts = 2.75 kN [see Eq. 2.6 in Chapter 2 ]

Shear resistance of fastener


FS-F,ss = 4.00 kN

Bearing strength of wood panels


FS-F,bp = 2.44 kN (KD=2.20; CD=1.6) [see Eq. 2.8 in Chapter 2 ]

Sheathing-to-frame connection strength


FS-F = 2.44 kN [see Eq. 2.4 in Chapter 2 ]
C-8 Appendix C

Shear strenght of the wall associated to sheathing-to-frame connection strength


Easley et al.'s method

a= 1250 mm length of the panel


H= 2500 mm heigth of the wall
sts= 600 mm studs spacing
m= 1.0 number of interior stud
efs= 150 mm edge fasteners spacing
ffs= 300 mm field fasteners spacing
ns= 15 number of side fasteners, excluding those at the end
ne= 9 number of end fasteners
nsi= 7 number of fasteners in each interior stud, excluding those at the end

Ie = S xei2 = 1350000 Is = S xsi2 = 0


E = ns + 4Ie + 2 nsi Is / L = 18.46

Force in the side fasteners


Ds = Fs / va = H / E = 135.46 [see Eq. 2.9 in Chapter 2 ]

Force in the end fasteners


Dei = Fei / va = ((a/ne)2 + (2xeimax H/(aE))2)1/2 = 190.26 [see Eq. 2.10 in Chapter 2 ]

Dmax = max {Ds ; Dei} = 190.26

Shear strenght of the wall associated to sheathing-to-frame connection strength


vS-F =1.40(*) 1 / Dmax FS-F= 17.94 KN/m [see Eq. 2.11 in Chapter 2 ]

Shear strenght of the wall associated to sheathing-to-frame connection strength


Simplified method

efs= 150 mm edge fasteners spacing


n'e= 6.67 number of end screws per unit length

vS-F = 1.40(*) n'e FS-F = 22.75 KN/m [see Eq. 2.12 in Chapter 2 ]

* Assuming that the adding of GWB to the opposite side of the wall assembly increases the
shear strength by about 40%
Ev alu atio n of seismic actio n and wall sh ear streng th fo r th e stud y case C-9

Frame (stud buckling) failure [see Section 2.2.2 in Chapter 2 ]

Stud back-to-back coupled C100x50x10x1.00


H= 2500 mm heigth of studs
fys = 350 Mpa yield strength of the studs
fus = 420 Mpa ultimate tensile strength of the studs
Lx = 2500 mm unbraced length (out of plane - x-z plane)
Ly = 300 mm unbraced length (in plane - x-y plane)

Nb,out = 66.96 kN out of plane axial strength [see Eq. 2.17 in Chapter 2 ]
Nb,in = 74.03 kN in plane axial strength [see Eq. 2.17 in Chapter 2 ]

Frame strength
FF = min(Nb,out ; Nb,in)= 66.96 kN

Shear strenght of the wall associated to frame strength


vF = FS-F / H = 26.78 KN/m [see Eq. 2.40 in Chapter 2 ]

The valuation of axial strength of studs has been carried out using
ColdForm computer program (Included in: A., Ghersi, R., Landolfo, F.M.,
Mazzolani (2002) Design of metallic cold-formed thin-walled members. Spon
Press). The results provided by Details window of this program are reported in
the following.

EVALUATION OF THE EFFECTIVE CROSS-SECTION:

Element 1 single edge fold stiffener, connected at end 2


geometrical data: bP = 8.91 mm t = 1.00 mm bP/t = 8.91
end 2 - fr = 1.41 mm
axial stresses: end 1 - sigma = 350.00 MPa
end 2 - sigma = 350.00 MPa
coefficients: k sigma = 0.5000
lambdaP = 0.5414
rho = 1.0000
the element is fully effective

Element 2 doubly supported element


geometrical data: bP = 47.83 mm t = 1.00 mm bP/t = 47.83
end 1 - fr = 1.41 mm
end 2 - fr = 1.41 mm
axial stresses: end 1 - sigma = 350.00 MPa
end 2 - sigma = 350.00 MPa
C -1 0 Appendix C

coefficients: k sigma = 4.0000


lambdaP = 1.0271
rho = 0.7651
beff = 36.59 mm be1 = 18.30 mm be2 = 18.30 mm

Element 3 doubly supported element


geometrical data: bP = 97.83 mm t = 1.00 mm bP/t = 97.83
end 1 - fr = 1.41 mm
end 2 - fr = 1.41 mm
axial stresses: end 1 - sigma = 350.00 MPa
end 2 - sigma = 350.00 MPa
coefficients: k sigma = 4.0000
lambdaP = 2.1008
rho = 0.4262
beff = 41.69 mm be1 = 20.85 mm be2 = 20.85 mm

Element 4 doubly supported element


geometrical data: bP = 47.83 mm t = 1.00 mm bP/t = 47.83
end 1 - fr = 1.41 mm
end 2 - fr = 1.41 mm
axial stresses: end 1 - sigma = 350.00 MPa
end 2 - sigma = 350.00 MPa
coefficients: k sigma = 4.0000
lambdaP = 1.0271
rho = 0.7651
beff = 36.59 mm be1 = 18.30 mm be2 = 18.30 mm

Element 5 single edge fold stiffener, connected at end 1


geometrical data: bP = 8.91 mm t = 1.00 mm bP/t = 8.91
end 1 - fr = 1.41 mm
axial stresses: end 1 - sigma = 350.00 MPa
end 2 - sigma = 350.00 MPa
coefficients: k sigma = 0.5000
lambdaP = 0.5414
rho = 1.0000
the element is fully effective

Element 6 single edge fold stiffener, connected at end 2


geometrical data: bP = 8.91 mm t = 1.00 mm bP/t = 8.91
end 2 - fr = 1.41 mm
axial stresses: end 1 - sigma = 350.00 MPa
end 2 - sigma = 350.00 MPa
coefficients: k sigma = 0.5000
lambdaP = 0.5414
rho = 1.0000
the element is fully effective

Element 7 doubly supported element


geometrical data: bP = 47.83 mm t = 1.00 mm bP/t = 47.83
end 1 - fr = 1.41 mm
end 2 - fr = 1.41 mm
axial stresses: end 1 - sigma = 350.00 MPa
end 2 - sigma = 350.00 MPa
coefficients: k sigma = 4.0000
lambdaP = 1.0271
rho = 0.7651
beff = 36.59 mm be1 = 18.30 mm be2 = 18.30 mm

Element 8 doubly supported element


Ev alu atio n of seismic actio n and wall sh ear streng th fo r th e stud y case C-1 1

geometrical data: bP = 97.83 mm t = 1.00 mm bP/t = 97.83


end 1 - fr = 1.41 mm
end 2 - fr = 1.41 mm
axial stresses: end 1 - sigma = 350.00 MPa
end 2 - sigma = 350.00 MPa
coefficients: k sigma = 4.0000
lambdaP = 2.1008
rho = 0.4262
beff = 41.69 mm be1 = 20.85 mm be2 = 20.85 mm

Element 9 doubly supported element


geometrical data: bP = 47.83 mm t = 1.00 mm bP/t = 47.83
end 1 - fr = 1.41 mm
end 2 - fr = 1.41 mm
axial stresses: end 1 - sigma = 350.00 MPa
end 2 - sigma = 350.00 MPa
coefficients: k sigma = 4.0000
lambdaP = 1.0271
rho = 0.7651
beff = 36.59 mm be1 = 18.30 mm be2 = 18.30 mm

Element 10 single edge fold stiffener, connected at end 1


geometrical data: bP = 8.91 mm t = 1.00 mm bP/t = 8.91
end 1 - fr = 1.41 mm
axial stresses: end 1 - sigma = 350.00 MPa
end 2 - sigma = 350.00 MPa
coefficients: k sigma = 0.5000
lambdaP = 0.5414
rho = 1.0000
the element is fully effective

EFFECTIVENESS OF EDGE STIFFENERS

Edge stiffener for element 2


spring stiffness: k = 0.1296 MPa
geometrical data: Ar = 27.3 mm2 Ir = 212 mm4
yGr = 92.97 mm zGr = 97.84 mm
buckling stress: sig cr,s = 176.24 MPa
reduction: lambda = 1.4092 chi = 0.4411 chiR = 0.4411

Edge stiffener for element 4


spring stiffness: k = 0.1296 MPa
geometrical data: Ar = 27.3 mm2 Ir = 212 mm4
yGr = 92.97 mm zGr = 2.16 mm
buckling stress: sig cr,s = 176.24 MPa
reduction: lambda = 1.4092 chi = 0.4411 chiR = 0.4411

Edge stiffener for element 7


spring stiffness: k = 0.1296 MPa
geometrical data: Ar = 27.3 mm2 Ir = 212 mm4
yGr = 7.03 mm zGr = 97.84 mm
buckling stress: sig cr,s = 176.24 MPa
reduction: lambda = 1.4092 chi = 0.4411 chiR = 0.4411

Edge stiffener for element 9


spring stiffness: k = 0.1296 MPa
geometrical data: Ar = 27.3 mm2 Ir = 212 mm4
yGr = 7.03 mm zGr = 2.16 mm
C -1 2 Appendix C

buckling stress: sig cr,s = 176.24 MPa


reduction: lambda = 1.4092 chi = 0.4411 chiR = 0.4411

gross cross-section effective cross-section

EVALUATION OF DESIGN BUCKLING RESISTANCE

area of the gross section: Ag = 423 mm2


area of the effective section: Aeff = 212 mm2
reduction factor: betaA = 0.4998

x-z plan:
buckling length: l = 2.50 m
radius of gyration of gross cross-section: i = 40.58 mm
slenderness: lambda = 61.61
relative slenderness: lambda bar = 0.5660
imperfection factor: alfa = 0.21
reduction factor: chi = 0.9024

Design buckling resistance: Nb,Rd = 66.96 kN

x-y plan:
buckling length: l = 0.30 m
radius of gyration of gross cross-section: i = 24.02 mm
slenderness: lambda = 12.49
relative slenderness: lambda bar = 0.1147
imperfection factor: alfa = 0.34
reduction factor: chi = 1.0000

Design buckling resistance: Nb,Rd = 74.03 kN


Ev alu atio n of seismic actio n and wall sh ear streng th fo r th e stud y case C-1 3

Frame-to-foundation failure [see Section 2.2.3 in Chapter 2 ]

Tension frame-to-foundation failure

hold-down
(F(F-F)N,hd)
hold-down to - frame min F( F F ) N
(F(F-F)N,hd-fr)
hold-down to foundation
(F(F-F)N,hd-fo) v(F-F)N
F(F-F)N,hd-fr) = 184.05 kN [see Eq. 2.45 in Chapter 2 ]

F(F-F)N,hd-fo) = 258.40 kN [see Eq. 2.51 in Chapter 2 ]

Tension frame-to-foundation connection strength


F(F-F)N = 184.05 kN [see Eq. 2.41 in Chapter 2 ]

Shear strenght of the wall associated to the tension frame-to-foundation connection strength
v(F-F)N = F(F-F)N / H = 73.62 KN/m [see Eq. 2.59 in Chapter 2 ]

Shear frame-to-foundation failure

Shear anchor to frame connection


(F(F-F)V,a-fr)
Shear anchor to foundation connection
min F( F F )V
(F(F-F)V,a-fo)

v(F-F)V

F(F-F)V,a-fr) = 9.00 kN [see Eq. 2.47 in Chapter 2 ]

F(F-F)V,a-fo) = 8.70 kN [see Eq. 2.57 in Chapter 2 ]

Shear frame-to-foundation connection strength


F(F-F)V = 8.70 kN [see Eq. 2.54 in Chapter 2 ]

Shear strenght of the wall associated to the tension frame-to-foundation connection strength
v(F-F)V = n' F(F-F)V = 87.00 KN/m (n'=10) [see Eq. 2.60 in Chapter 2 ]

Shear strenght of the wall associated to the of the frame-to-foundation connection strength
vF-F = 73.62 KN/m [see Eq. 2.58 in Chapter 2 ]
C -1 4 Appendix C

Unit shear strength [see Section 2.2 in Chapter 2 ]

v min ^v S  F ; v F ; v F  F ` [see Eq. 2.3 in Chapter 2 ]

Soil type vS (kN/m)


vS-F 17.9
vF 26.8
vF-F 73.6
min 17.9

Sheathing-to-frame connection failure

frame connection

sheathing
D-1

Appendix D

Construction details of the sub-assembly


specimen
D-2 Appendix D

joist track
(U 260X40X1.00mm)
Joist
(C 260x40x10x1.50mm
bearing stiffener
(C 100x50x10x1.00mm)

stud track
(U 100X40X1.00mm)

end stud
(back-to-back coupled
C 100x50x10x1.00mm)

intermediate stud
(C 100x50x10x1.00mm)

stud track
(U 100X40X1.00mm)

X-bracing

Global 3D view of specimen (without sheathings).


Construction details of th e sub-assemb ly specimen D-3

A close-up view of the tension connection between the wall and the foundation.

A close-up view of the shear connection between the wall and the foundation.
D-4 Appendix D

A close-up view of the back-to-back connection of coupled end studs.

A close-up view of the connection between the intermediate stud and the bottom stud track.
Construction details of th e sub-assemb ly specimen D-5

joist track (U 260X40X1.00mm)

No. 2 4.2x13mm modified truss head self drilling screws

bearing stiffener
(C 100x50x10x1.00mm)

stud track (U 100X40X1.00mm)

stud (C 100x50x10x1.00mm)
joist (C 260x40x10x1.50mm)

A close-up view of the connection between the joist and the joist track.

joist track (U 260X40X1.00mm)

bearing stiffener
(C 100x50x10x1.00mm)

stud track (U 100X40X1.00mm)

joist (C 260x40x10x1.50mm) 4.2x13mm modified truss head self drilling screws


spaced at 75mm on center

A close-up view of the connection between the joist track and the top stud track.
D-6 Appendix D

Close-up views of the connection between the joist, the joist track and the bearing stiffener.

A close-up view of the connection between the joist and the X-bracing.
Construction details of th e sub-assemb ly specimen D-7

1250x1450x18.0mm Type 3 OSB


(KRONOPLY by KRONOFRANCE)
floor sheathing

1250x2500x9.0mm Type 3 OSB


(KRONOPLY by KRONOFRANCE)
exterior wall sheathing
1200x2500x12.5mm GWB
(PLACOLAST by BPB ITALIA)
interior wall sheathing

Global 3D view of specimen (with sheathings)


D-8 Appendix D

Global 3D view of connections between the external sheathing and the wall framing
Construction details of th e sub-assemb ly specimen D-9

3.5x25mm bugle head self drilling screws


spaced at 150mm (at the perimeter)
with edge distance equal to 10mm

stud track (U 100X40X1.00mm)

stud (C 100x50x10x1.00mm)

3.5x25mm bugle head self drilling screws


spaced at 300mm (in the field)

1200x2500x12.5mm GWB
(PLACOLAST by BPB ITALIA)
interior wall sheathing

stud track (U 100X40X1.00mm)

Global 3D view of connections between the internal sheathing and the wall framing
D-10 Appendix D

1250x1450x18.0mm Type 3 OSB


4.2x32mm flat head self drilling screws (KRONOPLY by KRONOFRANCE)
spaced at 150mm floor sheathing
(for sheathing-to-joist track connections)

4.2x32mm flat head self drilling screws


spaced at 250mm
(for sheathing-to-joist connections)

joist track
(U 260X40X1.00mm)

Global 3D view of connections between the floor sheathing and the floor framing
E-1

Appendix E

Monotonic test results


E-2 App end ix E

a2 Wall 2
d2

d1
a1 Wall 1
actuator load (a1; a2)
LVDT (d1; d2)

Instrument arrangement for the floor.

w1

w4 w5

w2 w3

f1
i1 i2
Wall 1 Wall 2
actuator load (a1) actuator load (a2)
LVDT (w1,1; w1,2; w1,3; w1,4; LVDT (w2,1; w2,2; w2,3; w2,4;
w1,5) w2,5)
clinometer (i2,1; i2,2)

Instrument arrangement for the walls.


Monotonic test results E-3

50
kN
45

40

35

30

25

20

15

10

5
mm
0
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150

Force vs. displacement measured by the actuator a1.

50
kN
45

40

35

30

25

20

15

10

5
mm
0
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150

Force measured by the actuator a1 vs. displacement measured by the LVDT w1,1.
E-4 App end ix E

50
kN
45

40

35

30

25

20

15

10

5
mm
0
0 1 2 3 4 5 6 7

Force measured by the actuator a1 vs. displacement measured by the LVDT w1,2.

50
kN
45

40

35

30

25

20

15

10

5
mm
0
0 1 2 3 4 5 6 7

Force measured by the actuator a1 vs. displacement measured by the LVDT w1,3
Monotonic test results E-5

50
kN
45

40

35

30

25

20

15

10

5
mm
0
-15 -14 -13 -12 -11 -10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0

Force measured by the actuator a1 vs. displacement measured by the LVDT w1,4.

50
kN
45

40

35

30

25

20

15

10

5
mm
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

Force measured by the actuator a1 vs. displacement measured by the LVDT w1,5.
E-6 App end ix E

50
kN
45

40

35

30

25

20

15

10

5
mm
0
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150

Force measured by the actuator a1 vs. displacement measured by the LVDT d1.

50
kN
45

40

35

30

25

20

15

10

5
mm
0
0 1 2 3 4 5 6 7

Force measured by the actuator a1 vs. displacement measured by the LVDT f1.
Monotonic test results E-7

50
kN
45

40

35

30

25

20

15

10

5
mm
0
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150

Force vs. displacement measured by the actuator a2.

50
kN
45

40

35

30

25

20

15

10

5
mm
0
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150

Force measured by the actuator a2 vs. displacement measured by the LVDT w2,1.
E-8 App end ix E

50
kN
45

40

35

30

25

20

15

10

5
mm
0
0 1 2 3 4 5 6 7

Force measured by the actuator a2 vs. displacement measured by the LVDT w2,2.

50
kN
45

40

35

30

25

20

15

10

5
mm
0
0 1 2 3 4 5 6 7

Force measured by the actuator a2 vs. displacement measured by the LVDT w2,3.
Monotonic test results E-9

50
kN
45

40

35

30

25

20

15

10

5
mm
0
-15 -14 -13 -12 -11 -10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0

Force measured by the actuator a2 vs. displacement measured by the LVDT w2,4.

50
kN
45

40

35

30

25

20

15

10

5
mm
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

Force measured by the actuator a2 vs. displacement measured by the LVDT w2,5.
E-10 App end ix E

50
kN
45

40

35

30

25

20

15

10

5
mm
0
0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150

Force measured by the actuator a2 vs. displacement measured by the LVDT d2.

50
kN
45

40

35

30

25

20

15

10

5
mm
0
0 1 2 3 4 5 6 7

Force measured by the actuator a2 vs. displacement measured by the LVDT f2.
Monotonic test results E-11

50
deg
45

40

35

30

25

20

15

10

5
mm
0
-5 -4 -3 -2 -1 0

Force measured by the actuator a2 vs. rotation measured by the clinometer i2,1.

50
deg
45

40

35

30

25

20

15

10

5
mm
0
-5 -4 -3 -2 -1 0

Force measured by the actuator a2 vs. rotation measured by the clinometer i2,2.
F-1

Appendix F

Cyclic test results


F-2 Appendix F

actuator load (a1; a2)


LVDT (d1; d2)

Instrument arrangement for the floor.

Wall 1 Wall 2
actuator load (a1) actuator load (a2)
LVDT (w1,1; w1,2; w1,3; w1,4; LVDT (w2,1; w2,2; w2,3; w2,4;
w1,5) w2,5)
clinometer (i2,1; i2,2)

Instrument arrangement for the walls.


Cyclic test results F-3

Force vs. displacement measured by the actuator a1.

Force measured by the actuator a1 vs. displacement measured by the LVDT w1,1.
F-4 Appendix F

Force measured by the actuator a1 vs. displacement measured by the LVDT w1,2.

Force measured by the actuator a1 vs. displacement measured by the LVDT w1,3
Cyclic test results F-5

Force measured by the actuator a1 vs. displacement measured by the LVDT w1,4.

Force measured by the actuator a1 vs. displacement measured by the LVDT w1,5.
F-6 Appendix F

Force measured by the actuator a1 vs. displacement measured by the LVDT d1.

Force measured by the actuator a1 vs. displacement measured by the LVDT f1.
Cyclic test results F-7

Force vs. displacement measured by the actuator a2.

Force measured by the actuator a2 vs. displacement measured by the LVDT w2,1.
F-8 Appendix F

Force measured by the actuator a2 vs. displacement measured by the LVDT w2,2.

Force measured by the actuator a2 vs. displacement measured by the LVDT w2,3.
Cyclic test results F-9

Force measured by the actuator a2 vs. displacement measured by the LVDT w2,4.

Force measured by the actuator a2 vs. displacement measured by the LVDT w2,5.
F-10 Appendix F

Force measured by the actuator a2 vs. displacement measured by the LVDT d2.

Force measured by the actuator a2 vs. displacement measured by the LVDT f2.
Cyclic test results F-11

Force measured by the actuator a2 vs. rotation measured by the clinometer i2,1.

Force measured by the actuator a2 vs. rotation measured by the clinometer i2,2.

You might also like