Download as pdf or txt
Download as pdf or txt
You are on page 1of 298

Clemson University

TigerPrints
All Dissertations Dissertations

8-2017

Strength and Serviceability Performances of


Southern Yellow Pine Cross-Laminated Timber
(CLT) and CLT-Glulam Composite Beam
Mengzhe Gu
Clemson University, mengzhg@g.clemson.edu

Follow this and additional works at: https://tigerprints.clemson.edu/all_dissertations

Recommended Citation
Gu, Mengzhe, "Strength and Serviceability Performances of Southern Yellow Pine Cross-Laminated Timber (CLT) and CLT-Glulam
Composite Beam" (2017). All Dissertations. 2014.
https://tigerprints.clemson.edu/all_dissertations/2014

This Dissertation is brought to you for free and open access by the Dissertations at TigerPrints. It has been accepted for inclusion in All Dissertations by
an authorized administrator of TigerPrints. For more information, please contact kokeefe@clemson.edu.
STRENGTH AND SERVICEABILITY PERFORMANCES
OF SOUTHERN YELLOW PINE CROSS-LAMINATED
TIMBER (CLT) AND CLT-GLULAM COMPOSITE BEAM

A Dissertation
Presented to
the Graduate School of
Clemson University

In Partial Fulfillment
of the Requirements for the Degree
Doctor of Philosophy
Civil Engineering

by
Mengzhe Gu
August 2017

Accepted by:
Dr. Weichiang Pang, Committee Chair
Dr. Thomas Cousins
Dr. Brandon Ross
Dr. Nadarajah Ravichandran
Abstract

Cross-Laminated Timber (CLT) is a relatively new engineered mass timber construction

material in the North America. While CLT was invented roughly three decades ago in Europe, it

was not popular in the North America until recent years, in which there has been a surging demand

for mass timber construction in the North American market. The advantages from the material itself

and the level of pre-fabrication offers a flexible system suitable for large and tall wood buildings.

However, the design provisions for CLT have not been developed for the building codes of the United

States (US). Thus, the use of CLT is limited to custom projects.

There were two phases in this research. During the inception of this research, there was

no commercial production of structural grade CLT panels in the US using local wood species. The

first phase of the research investigated the manufacturing process of CLT panels in a pilot scale

using Southern Yellow Pine (SYP), a group of tree species native to the Southeast US. Tests were

conducted to evaluate the adhesive and structural performances of CLT panels in flexural and shear.

The test results showed that SYP can be used to produce CLT panels that meet and exceed the

product standard requirements for CLT in the North America.

The results generated from the pilot study were used in the second phase of the research to

design, fabricate and test a full-scale 5 ft. wide and 40 ft. long CLT-Glulam composite floor system

designed for long span applications. The proposed CLT-Glulam composite section is analogous to a

typical precast concrete box girder, which is designed to improve the material efficiency. The CLT

panels and Glulam beams in the composite floor system were fastened by self-tapping screws installed

at an incline angle. To investigate the feasibility of using the CLT-Glulam composite section for floor

system in residential/office construction, a non-destructive vibration (modal) test was conducted to

evaluate the serviceability behavior and destructive structural tests were performed to evaluate

the flexure and shear strengths. The modal test utilized an instrumented heel-drop excitation to

ii
simulate the impact from human footsteps and accelerometers to record the vibration responses.

The vibration test results revealed that the composite floor could produce vibration sensitive to

building occupants. The results of the destructive tests verified that the composite floor system

could safely carry the design load in terms of bending and shear strengths. Based on the findings

of the experimental study, it was found that the design of the CLT-Glulam composite system is

governed by serviceability and not ultimate strengths. It is recommended that, in addition to

designing for the ultimate strengths, the serviceability criteria be considered with equal important

as the strength design for improved occupant comfort.

iii
Dedication

TO THE WILDLY WONDERFUL WORLD THAT CREATED ME,

MY LOVING PARENTS,

GRANDPARENTS AND

MS. LUSI HONG.

iv
Acknowledgments

I would like to sincerely thank my advisor, Dr. Weichiang Pang, for introducing me to the

subject of engineered timber and leading me through my research endeavors. Thank you for your

offer letter (dated May 28, 2013) that resulted in once-in-a-lifetime opportunities and adventures in

PhD study. I came to Clemson to work with you, and I could not have made a better choice that

will forever guide my future.

I would also like to thank my committee members Dr. Cousins, Dr. Ross and Dr. Ravi for

providing suggestion and comments in experimental testing, those were invaluable to my research. I

enjoyed your jokes, or mockery or sometimes both. It was my pleasure to have you as my committee

members and I have learnt a tremendous amount in and out of the class. Special thank goes to

Dr. Layton for granting me access to the woodshop for pilot studies, her generous tool supply and

Rick’s smoke pork rib. In addition, I’d like to thanks Lt Col Anthony Barrett from the US Air Force

Academy for providing invaluable experience and knowledge in experimental modal testing.

Sincere thanks also go to the Civil Engineering lab technicians: Danny Mertz, Scott Black,

Sam Biemann, and Alex Pittman. No single test would be possible without the help from you guys.

Your working attitude has constantly inspired me of achieving higher quality of everything. I’m

grateful for all the friends that voluntarily helped me, who were wing-mans as we travelled through

the trial and tribulations of the PhD program: Nathan Schneider, Graham Montgomery, Michael

Stoner, Ross Philips, Will Granbery, Tylor Hood, Fanfu Fan, Yanshijie Wang, Weiyi Yu, Haotian

Li, Zexu Qian, Shang Yu, Mattson Wiksell, Lancelot Reres. You’re all amazing!

Special thanks for all the material suppliers: Columns Lumber for their generous supply of

2x Southern Pine lumbers for pilot scale study; Anthony Forest Products for their donation of the

glulam beams used in the full size composite beam; Structurlam for manufacturing the full-scale

Southern Pine CLT panels; MyTiCon for supplying the incline screws.. Thanks!

v
This research was partially supported by Forest Products Research Initiative Competitive

Grant No. 2012-34638-20236 from the USDA (National Institute of Food and Agriculture) and the

Clemson Wood Utilization and Design (WUD) Institute. Any opinions, findings, and conclusions or

recommendations expressed in this material are those of the author and do not necessarily reflect

the views of the USDA and WUD.

Finally, I would like to thank my parents, and my lovely fiancé Lusi Hong. Thanks for

supporting me during my studies and constantly urging me going forward.

To all of you, thanks for being there for the moment that we encountered.

vi
Table of Contents

Title Page . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii

Dedication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Definition of CLT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Scope of Research and Potential Contributions . . . . . . . . . . . . . . . . . . . . . 4
1.3 Organization of the Dissertation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 Literature Review and Background . . . . . . . . . . . . . . . . . . . . . . . . . . . 6


2.1 Status of the CLT Industry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2 Application of CLT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 Joints and Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.4 Fire performance of CLT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.5 Wood Adhesive . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3 Commercial CLT Manufacturing, Clemson Pilot Production and Specimen


Preparations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.1 CLT production—European Experience . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.2 Clemson pilot CLT manufacturing . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.3 Specimens preparation for structural evaluations . . . . . . . . . . . . . . . . . . . . 56

4 Rolling Shear of Southern Pine CLT . . . . . . . . . . . . . . . . . . . . . . . . . . 59


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
4.2 Background and literature review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.3 Adhesives and specimens prepared . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.4 Test method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.5 Result and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.6 Characteristic rolling shear strength . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

5 Bending Strength and Stiffness of Southern Pine CLT . . . . . . . . . . . . . . . 75


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
5.2 Background and literature review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

vii
5.3 Test method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
5.4 Test result and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.5 Monte Carlo method for bending strength estimation . . . . . . . . . . . . . . . . . . 92
5.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

6 Bending and Shear Strengths of Composite CLT-Glulam Beam . . . . . . . . . .103


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
6.2 Background and motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
6.3 Self Tapping Inclined Screw and Small Scale Tests . . . . . . . . . . . . . . . . . . . 112
6.4 Design and Assembly of CLT-Glulam Composite Beam . . . . . . . . . . . . . . . . . 127
6.5 Assembling of the Composite Beam . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
6.6 Full Scale Destructive Tests of CLT-Glulam Composite Beam . . . . . . . . . . . . . 138
6.7 Conclusion and Future Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159

7 Vibration Properties of Composite CLT-Glulam Beam . . . . . . . . . . . . . . .162


7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
7.2 Motivation and Scope of Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
7.3 Literature Review and Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
7.4 Initial Vibration Assessment Using DG11 Guidelines . . . . . . . . . . . . . . . . . . 177
7.5 Basic Structural Dynamics of Floors . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
7.6 Modal Testing Theory and Equipment Preparation . . . . . . . . . . . . . . . . . . . 181
7.7 Experimental Testing Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
7.8 Presentation of Measured FRFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
7.9 System identification from measured FRF data set . . . . . . . . . . . . . . . . . . . 212
7.10 Finite Element Modeling of Composite Floor System . . . . . . . . . . . . . . . . . . 230
7.11 Vibration Serviceability Assessment and Recommendations . . . . . . . . . . . . . . 241

8 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .248
8.1 Summary of Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
8.2 Summary and Findings of Research Tasks . . . . . . . . . . . . . . . . . . . . . . . . 249
8.3 Limitations and Recommendations for Future Research . . . . . . . . . . . . . . . . 250

Appendices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .252
A Pictures of Major Strength Direction Finger Joint and Lumber Knot Failures . . . . 253
B Screw Joint Capacity Estimation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
C Composite Beam Capacity Estimation . . . . . . . . . . . . . . . . . . . . . . . . . . 261
D Frequency Domain Response of a Multi-Degree-of-Freedom System under Harmonic
Excitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269

viii
List of Tables

3.1 Working properties of adhesives used in this study . . . . . . . . . . . . . . . . . . . 48


3.2 Number of specimens produced for major and minor axes rolling shear test . . . . . 57
3.3 Number of specimens produced for major and minor axes bending test . . . . . . . . 58

4.1 Density and moisture content (MC) of major and minor axis specimens . . . . . . . 63
4.2 Summary of rolling shear strengths for major and minor Axes . . . . . . . . . . . . . 65
4.3 Glue bond strength estimation for major axis specimens . . . . . . . . . . . . . . . . 68
4.4 Characteristic rolling shear strengths . . . . . . . . . . . . . . . . . . . . . . . . . . 73

5.1 Minimum required characteristic test value per PRG-320: CLT designation V31 . . . 79
5.2 Density and MC of tested bending specimens . . . . . . . . . . . . . . . . . . . . . . 82
5.3 PRG-320 suggested ratios of E and G based on E0 . . . . . . . . . . . . . . . . . . . 83
5.4 Section property of 3-ply SP CLT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.5 Mean values of MOE and MOR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
5.6 Characteristic bending test values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
5.7 Lognormal distribution parameters for IGTP No.2 2x6 SP lumbers at 12% moisture
content . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
5.8 Test statistics for in-house bending and IGTP . . . . . . . . . . . . . . . . . . . . . . 99
5.9 Test statistics for CLT MOR and Monte Carlo simulations . . . . . . . . . . . . . . . 102

6.1 Shear Test Results Montgomery (2014) . . . . . . . . . . . . . . . . . . . . . . . . . . 121


6.2 Inclined screw stiffness values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
6.3 Section properties of the composite beam (assume full composite) . . . . . . . . . . . 130
6.4 Bending and shear capacities of the composite beam . . . . . . . . . . . . . . . . . . 130
6.5 Live load and long term deflections . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
6.6 Mid-span deflection and stiffness of the 1st bending test . . . . . . . . . . . . . . . . 142
6.7 Deflection with fixed and pin boundary condition assumptions . . . . . . . . . . . . 145
6.8 Load increments used in 2nd bending test . . . . . . . . . . . . . . . . . . . . . . . . 149
6.9 Shear Test Failure Estimations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
6.10 Load increments used in the shear test . . . . . . . . . . . . . . . . . . . . . . . . . . 157

7.1 Estimation of fundamental natural frequencies based on DG11 guideline . . . . . . . 177


7.2 Digital signal process settings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
7.3 Damping estimates for composite beam with fixed boundary condition . . . . . . . . 220
7.4 Damping estimates for composite beam with pinned boundary condition . . . . . . . 220
7.5 System poles identified for the composite beam during modal testing . . . . . . . . . 225
7.6 Vibration modes for composite CLT-Glulam beam . . . . . . . . . . . . . . . . . . . 225
7.7 Benchmark results of the FE model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
7.8 Composite floor modeling parameters . . . . . . . . . . . . . . . . . . . . . . . . . . 238
7.9 Comparison of experimental and computed natural frequencies . . . . . . . . . . . . 240

ix
List of Figures

1.1 A typical 3-ply CLT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2


1.2 Carbon storage of wood. Adapted from Kolb (2008) . . . . . . . . . . . . . . . . . . 3

2.1 European CLT manufactures as of 2014. Source Grasser (2015) . . . . . . . . . . . . 7


2.2 8-Story Wood Innovation Design Center (Archdaily) . . . . . . . . . . . . . . . . . . 9
2.3 7-story mass timber office building T3 (Archdaily) . . . . . . . . . . . . . . . . . . . 9
2.4 10-story Forte apartment building (Lendlease) . . . . . . . . . . . . . . . . . . . . . . 10
2.5 17-story Brock Commons student residence in UBC(Archdaily) . . . . . . . . . . . . 10
2.6 Practice of platform construction at Redstone Project (Lendlease) . . . . . . . . . . 11
2.7 Types of gravity framings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.8 Post & beam framing: Carbon 12 in Portland OR. . . . . . . . . . . . . . . . . . . . 13
2.9 Two-way slab roof: Chicago Pavilion in Chicago IL (Hedrich Blessing) . . . . . . . . 14
2.10 Honeycomb framing in Murray Grove apartment. Hackney, London (Waugh Thistle-
ton Architects) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.11 Lateral force resisting system of mass timber . . . . . . . . . . . . . . . . . . . . . . 16
2.12 Buckling restrained braces in Carbon 12 . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.13 Steel moment frame in 120 SE Clay Street . . . . . . . . . . . . . . . . . . . . . . . . 17
2.14 Reinforced concrete columns in Roosevelt High School dining hall . . . . . . . . . . . 18
2.15 Illustration of CLT connections (Gagnon et al. 2011) . . . . . . . . . . . . . . . . . . 20
2.16 Inclined self-tapping screw installed with washer guide (Rothoblaas) . . . . . . . . . 21
2.17 Exposed (left) and concealed (right) beam hanger (Rothoblaas) . . . . . . . . . . . . 21
2.18 Concealed hook connector connecting a glulam to a ledger beam (Rothoblaas) . . . . 22
2.19 Hold down and shear bracket(Rothoblaas) . . . . . . . . . . . . . . . . . . . . . . . . 23
2.20 Perforated plate installed at diaphragm joint in Carbon 12 project . . . . . . . . . . 24
2.21 Fire-resistant rating requirements for building elements. IBC 2015, Table 601. . . . . 26
2.22 CLT compartment fire test (left) and post-fire compartment (right). American Wood
Council . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.23 Microscopic pictures of hardwood (balsa) and softwood (pine)(Andersson et al. 2006) 30

3.1 Horizontal finger joint bit (Infinity Cutting Tools) . . . . . . . . . . . . . . . . . . . 38


3.2 Horizontal (left) and vertical (right) finger joint . . . . . . . . . . . . . . . . . . . . . 38
3.3 Vertical finger joint machined with glue applied . . . . . . . . . . . . . . . . . . . . . 39
3.4 Two pieces of lumber ready to be mated in a crowder . . . . . . . . . . . . . . . . . 40
3.5 Swelling and shrinkage checks in CLT with edge bonded layers (left) and without edge
bonded layers (right)(Brandner et al. 2013) . . . . . . . . . . . . . . . . . . . . . . . 41
3.6 Line-wise adhesive application(Ledinek Engineering) . . . . . . . . . . . . . . . . . . 41
3.7 Hydraulic CLT press(D.R.Johnson) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.8 Pneumatic press (Ledinek Engineering) . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.9 Hydraulic CLT press (Ledinek Engineering) . . . . . . . . . . . . . . . . . . . . . . . 44
3.10 CNC milling of CLT panels (Hans Hundegger AG) . . . . . . . . . . . . . . . . . . . 45
3.11 CNC Machining Center (IHB International) . . . . . . . . . . . . . . . . . . . . . . . 45

x
3.12 No.2 & No.3 2x6 Southern Yellow Pine being conditioned at woodshop . . . . . . . . 49
3.13 DELMHORST moisture meter with hammer electrode . . . . . . . . . . . . . . . . . 50
3.14 Lumbers trimmed into 4ft. segments . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.15 Lumber planing in progress. Newman press standby on the right . . . . . . . . . . . 51
3.16 Weighed resins and hardeners ready to be mixed . . . . . . . . . . . . . . . . . . . . 53
3.17 Adhesive application on a 16 f t2 area. Picture show malemin formaldehyde (MF)
adhesive . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.18 Spread the adhesive with a spatula . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.19 A 3-ply CLT being pressed with pressure shown . . . . . . . . . . . . . . . . . . . . . 55
3.20 Sizing a CLT panel with a circular saw . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.21 Pilot panel staged press . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

4.1 The principal axes of wood with respect to grain direction and growth rings . . . . . 60
4.2 Three-point short-span bending test setup for rolling shear (left: major axis ; right:
minor axis). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.3 Typical experimental setup for major axis rolling shear test . . . . . . . . . . . . . . 64
4.4 Schematic view of 3-layer ‘V3’ cross section for calculating rolling shear . . . . . . . 65
4.5 Load displacement curves of major axis rolling shear specimens . . . . . . . . . . . . 67
4.6 Typical rolling shear cracks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.7 Extension of diagonal shear crack into glue lines . . . . . . . . . . . . . . . . . . . . 69
4.8 Glue bond delamination (observed in EPI specimens) and bending failure . . . . . . 69
4.9 Load displacement curves of minor axis rolling shear tests . . . . . . . . . . . . . . . 70
4.10 Failure of minor axis specimen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.11 CDF of major axis rolling shear strength . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.12 CDF of minor axis rolling shear strength . . . . . . . . . . . . . . . . . . . . . . . . . 72

5.1 Schematic third-point bending test setup (top:major axis; bottom:minor axis) . . . . 80
5.2 Typical experimental setup for CLT bending test . . . . . . . . . . . . . . . . . . . . 81
5.3 Bending test load-displacement history for major (left) and minor (right) strength
direction specimens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
5.4 Typical failure of major strength direction Pilot CLT . . . . . . . . . . . . . . . . . . 87
5.5 Wood fiber tension failure at lumber knot . . . . . . . . . . . . . . . . . . . . . . . . 88
5.6 Finger joint failure in Structurlam major strength specimens . . . . . . . . . . . . . 89
5.7 Edgewise finger joint (left); flatwise finger joint (right) . . . . . . . . . . . . . . . . . 89
5.8 Structurlam lumber finger joint process . . . . . . . . . . . . . . . . . . . . . . . . . 90
5.9 Failure mechanism of minor axis specimens . . . . . . . . . . . . . . . . . . . . . . . 91
5.10 Normal distribution for bending strength . . . . . . . . . . . . . . . . . . . . . . . . 92
5.11 Beam modeling using shear analogy method (adapted from Gagnon et al. (2011)) . . 93
5.12 Flow chart of ASTM D1990 adjustment process. 1 M.C. was adjusted to 12% in the
in-house tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
5.13 In-house No. 2 2x6 Southern Pine bending test . . . . . . . . . . . . . . . . . . . . . 98
5.14 Lognormal distributions for IGTP and in-house bending test . . . . . . . . . . . . . 98
5.15 Correlation of MOE and MOR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
5.16 Lognormal distributions of IGTP UTS . . . . . . . . . . . . . . . . . . . . . . . . . . 100
5.17 Typical cross section of major (left) and minor (right) strength direction 3-ply CLT
specimen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
5.18 Results of simulation on bending strength of 3-ply CLT panel . . . . . . . . . . . . . 101

6.1 Timber concrete composite with mesh steel plate (Equilibrium Consulting Inc.) . . . 107
6.2 Dowel type anchorage used in timber-concrete composite (Steinberg et al. 2003) . . . 107
6.3 Timber concrete composite system in Life Cycle Tower (Buildipedia) . . . . . . . . . 107

xi
6.4 Cross section of composite glulam-CLT beam with double sided nail nail plate and
screw connection (Jacquier and Girhammar 2015) . . . . . . . . . . . . . . . . . . . . 108
6.5 CLT-LVL composite beam test setup (Masoudnia et al. 2016) . . . . . . . . . . . . . 109
6.6 Farfield bridge with FRP reinforced glulam girders (Advanced Structures & Compos-
ite Center) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
6.7 Proposed concept of hollow mass timber composite beam . . . . . . . . . . . . . . . 111
6.8 Geometric parameters for different wood screws (Jockwer et al. 2014) . . . . . . . . 114
6.9 Force interaction in a inclined screw connection . . . . . . . . . . . . . . . . . . . . . 115
6.10 Common timber connection joints. Left to right: shear tension, shear compression
and crossed joint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
6.11 Forces and stresses for European Yield Mode 3 for a inclined screw connection (Bejtka
and Blass 2002) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
6.12 ASSY and SFS Intec screws studied in Montgomery (2014) . . . . . . . . . . . . . . 120
6.13 Specimen for inclined screw joint Test 3 (Montgomery 2014) . . . . . . . . . . . . . . 121
6.14 ASSY VG Cylindrical 3/8 in. screw . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
6.15 Specimen dimensions for the inclined screw joint shear test . . . . . . . . . . . . . . 123
6.16 Drilling pilot holes using an angle guide . . . . . . . . . . . . . . . . . . . . . . . . . 124
6.17 Inclined screw installation with hammer drill . . . . . . . . . . . . . . . . . . . . . . 124
6.18 Test setup for small scale shear test . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
6.19 Results for two static shear tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
6.20 Screws sheared off in a inclined screw joint . . . . . . . . . . . . . . . . . . . . . . . 127
6.21 Cross section for the CLT-Glulam composite beam . . . . . . . . . . . . . . . . . . . 129
6.22 Maximum screw spacing envelop . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
6.23 Exaggerated deflection shape for a non-composite beam . . . . . . . . . . . . . . . . 134
6.24 Shop drawing for inclined screw . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
6.25 CLT unloading from a flatbed trailer . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
6.26 Placement of the material for bottom flange pilot drilling . . . . . . . . . . . . . . . 136
6.27 Temporary erection of glulam beams and lifeted away top assembly . . . . . . . . . . 137
6.28 CLT panel topping the temporary assembly . . . . . . . . . . . . . . . . . . . . . . . 137
6.29 Installation of inclined screw on the composite beam . . . . . . . . . . . . . . . . . . 138
6.30 Tools used in inclined screw installation . . . . . . . . . . . . . . . . . . . . . . . . . 138
6.31 Support of the long span composite beam . . . . . . . . . . . . . . . . . . . . . . . . 139
6.32 Attachment of string potentiometers on the composite beam . . . . . . . . . . . . . . 140
6.33 USB-1208FS Data acquisition aystem . . . . . . . . . . . . . . . . . . . . . . . . . . 140
6.34 Third-point bending setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
6.35 Moment and shear diagrams for bending test 1 . . . . . . . . . . . . . . . . . . . . . 142
6.36 Load displacement curve for the 1st bending test . . . . . . . . . . . . . . . . . . . . 143
6.37 Deformed beam during a displacement controlled loading procedure . . . . . . . . . 144
6.38 Interface slippages during the displacement controlled loading–1st bending . . . . . . 145
6.39 Boundary conditions for a third-point bending setup . . . . . . . . . . . . . . . . . . 146
6.40 Stiffness reduction in bending test . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
6.41 Three point flexural test setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
6.42 Moment and shear diagrams for bending test 2 . . . . . . . . . . . . . . . . . . . . . 148
6.43 Load displacement curve for the 2nd bending test . . . . . . . . . . . . . . . . . . . . 149
6.44 Rotation in the beam support . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
6.45 Interface slippage during the displacement controlled loading–2rd bending . . . . . . 150
6.46 CLT tension side failures. Red circle: finger joint failure; Rectangle: lumber knot
failure; Green rhombus: survived finger joint . . . . . . . . . . . . . . . . . . . . . . 152
6.47 Tension CLT delamination and rolling shear failure . . . . . . . . . . . . . . . . . . . 152
6.48 Glulam rolling shear and bending failure . . . . . . . . . . . . . . . . . . . . . . . . . 153
6.49 Cutting the CLT-glulam beam for shear test . . . . . . . . . . . . . . . . . . . . . . . 154

xii
6.50 Shear test setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
6.51 Moment and shear diagrams of the shear test . . . . . . . . . . . . . . . . . . . . . . 156
6.52 Load displacement curve for the shear test . . . . . . . . . . . . . . . . . . . . . . . . 157
6.53 Interface slippage during the load controlled shear test . . . . . . . . . . . . . . . . . 158
6.54 Delamination failure at 79.6 kip . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
6.55 Interface slippages after the shear test . . . . . . . . . . . . . . . . . . . . . . . . . . 160
6.56 Opened gap at connection interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

7.1 Lenzen’s modified Reiher-Meister scale. (Lenzen 1966) . . . . . . . . . . . . . . . . . 168


7.2 Recommended peak acceleration for human comfort for vibrations due to human
activities (Murray et al. 2003b; ISO 1989). The root mean square acceleration is the
square root of the area under the acceleration spectral density (ASD) curve in the
frequency domain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
7.3 Estimated peak accelerations (% gravity) for the CLT-Glulam composite beam . . . 178
7.4 Theoretical and experimental routes of vibration analyses . . . . . . . . . . . . . . . 180
7.5 Handheld spectrum analyzer (TestWorld Inc.) . . . . . . . . . . . . . . . . . . . . . . 183
7.6 Schematic view of the vibration test setup . . . . . . . . . . . . . . . . . . . . . . . . 193
7.7 Accelerometer attachment with hot glue . . . . . . . . . . . . . . . . . . . . . . . . . 194
7.8 In-house instrumented force plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
7.9 Calibration of force plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
7.10 Loadcell noise in decibels (dB) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
7.11 Original and filtered loadcell signal . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
7.12 Modal test signal acquiring system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
7.13 Heel-drop impact on a force plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
7.14 Time history of 7 heel-drop impacts . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
7.15 A single time history of heel-drop impact . . . . . . . . . . . . . . . . . . . . . . . . 203
7.16 Grid lines used in the heel-drop modal testing (only half of the span shown) . . . . . 203
7.17 An array of accelerometers scanning the test piece surface at Column 25 . . . . . . . 204
7.18 Picture of the excitation placement on node 61 . . . . . . . . . . . . . . . . . . . . . 205
7.19 Schematic view of the excitation placement . . . . . . . . . . . . . . . . . . . . . . . 205
7.20 Boundary conditions considered in the experimental modal analyses . . . . . . . . . 207
7.21 Schematic view of the excitation placement . . . . . . . . . . . . . . . . . . . . . . . 208
7.22 Modal test flow chart . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
7.23 Driving point accelerance and coherence for the fixed boundary condition. Force plate
and accelerometer in node 61 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
7.24 Driving point accelerance and coherence for the pinned boundary condition. Force
plate and accelerometer in node 61 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
7.25 Accelerance and coherence plot for all measured 156 nodes for fixed boundary condition214
7.26 Accelerance and coherence plot for all measured 156 nodes for pinned boundary con-
dition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
7.27 Spectrogram of all FRF measurements . . . . . . . . . . . . . . . . . . . . . . . . . . 216
7.28 Composite FRF magnitude created from the heel drop test. Natural frequencies are
identified as red dots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
7.29 Illustration of half power bandwidth method . . . . . . . . . . . . . . . . . . . . . . 219
7.30 Section of frequency range for LSCE method. Figure shows the range selected for the
dominant frequency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
7.31 Example curve fitting at DOF-1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
7.32 2D mode shapes at identified natural frequencies . . . . . . . . . . . . . . . . . . . . 227
7.33 3D mode shapes at identified natural frequencies . . . . . . . . . . . . . . . . . . . . 228
7.34 Coherence function with 10th and 90th percentile values . . . . . . . . . . . . . . . . 229
7.35 Modal assurance criteria values for obtain mode shapes . . . . . . . . . . . . . . . . 230

xiii
7.36 User defined layer shell element in SAP2000 . . . . . . . . . . . . . . . . . . . . . . . 233
7.37 Composite floor model. Red: shell elements; blue: frame elements; green: links and
restraints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
7.38 Centroids of respective elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
7.39 Deflection of a composite beam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
7.40 Computed transverse mode and axial mode . . . . . . . . . . . . . . . . . . . . . . . 237
7.41 Floor model computed mode shapes . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
7.42 Frequency gain in link element stiffness . . . . . . . . . . . . . . . . . . . . . . . . . 242
7.43 2-parameter Weibull distribution fitting for GRM S . . . . . . . . . . . . . . . . . . . 243
7.44 Vibration serviceability assessment for CLT-Glulam composite beam floor system . . 244

1 Specimen1-2 FJ and 1 knot failures . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253


2 Specimen2-1 FJ and 1 knot failures . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
3 Specimen3-1 FJ failure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
4 Specimen4-2 FJ failures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
5 Specimen5-1 FJ and 2 knot failures . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
6 Specimen6-2 knot failures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
7 Specimen7-1 FJ and 2 knot failures . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
8 Specimen8-1 FJ and 2 knot failures . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
9 Specimen9-1 FJ and 2 knot failures . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
10 Specimen10-2 FJ failures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256

xiv
Chapter 1

Introduction

1.1 Definition of CLT

According to the definition given by the American National Standards Institute, Cross-

Laminated Timber (CLT), also known as ‘X-lam’, is a “prefabricated solid engineered wood panel

made of as least three orthogonally bonded layers of solid-sawn lumber of a structural composite

lumber (SCL) that are laminated by gluing of the longitudinal and transverse layers with structural

adhesive to form a solid rectangular-shaped, straight and plane timber, intended for roof, floor or

wall applications” (ANSI 2012b).

Figure 1.1 shows a typical 5-ply CLT panel. Usually, a CLT panel consists of an odd number

of layers (3,5,7 or more) which are stacked orthogonally to each other and bonded with structural

adhesive or less commonly with dowels or nails. For many structural purposes, consecutive layers

may not be the same thickness. Because of the continuous bonding and consequently, quasi-rigid

composite action between the single layers, a very compact and versatile usable product arises.

Product dimensions allow its application as large-sized wall and floor elements as well as for other

load bearing system. In this way, modular dimensions which typically seen in a light frame wood

construction, can be neglected. This product has opened new dimensions in timber engineering

and allows architects and engineers to design and construct buildings that had been previously

restricted to reinforce concrete, steel or masonry based building materials. In this way, a new

building technique, the so-called solid timber or mass timber construction using CLT is developed,

making it possible to design and construct with engineered timber to previously unknown dimensions

1
and scales.

Figure 1.1: A typical 3-ply CLT

Although a relatively recent development, cross-laminated timber has its origin in the tradi-

tional furniture industry in central Europe. The evolution of the concept of a composite with rigidly

bonded crosswise layers from research through to a fully realized product took place in the 1990s.

Early engineering research occurred first in Switzerland and then in Austria. Meanwhile in Bavaria

(a state in Germany), the first three-story house was erected by the German timber company Merk,

where then the composite was known by the Germans as ‘Dickholz’ translated as ‘thick wood’. There

following a period of 5-6 years during which time the composite was experimented with by small

timber manufactures in the sub Alpine regions of Germany, Austria and Switzerland. In the early

2000s, manufacturing and construction techniques had matured enough for full scale production to

begin. As the popularity of the new timber technology spread, other European countries started

their own manufacturing capacity.

The advantages of CLT, as large sized and panel like solid timber construction elements

for the building sector, are in particularly obvious in its outstanding capability for pre-fabrication,

a dry and clean construction technique. This has allowed for a short erection times on site. The

high dimensional stability underlines accuracy, with the lowest tolerance well known for timber

construction in general. Further decisive criteria which argue for this product are the ability to

transfer the loads two dimensionally and the low mass, which reduce the foundation cost and make

it ideal for retrofit or upgrading of existing structures. Compared to light weight timber structures

(framing, post and beam system), CLT offers many advantages which include air permeability, the

distinctive specific storage capacity for humidity and temperature, the independence of modular

2
dimensions in arranging windows and door openings. The low mass, the stiffness and bearing

capacity against in-plane and out-of-plane stresses makes it ideal in multi-story residential and office

buildings, for schools, single family houses, and for wide-span structures such as bridges. In fact,

rib floors or box beams, as composites of CLT with other engineering timber such as duo, trio or

glue-laminated timber are really advantageous. Worldwide effort in research and development, as

well as processes for erection of small and medium buildings, are ongoing and observable.

Apart from CLT’s excellent applicability as building material, the increasing attention of

green building product in the architecture, engineering and construction communities has resulted in

important environmental advantages over other building materials – wood is the only major building

material derived from a resource that is both sustainable and renewable. The photosynthesis process

of trees removes carbon dioxide from atmosphere and store it in the wood in terms of sugar, and

generates oxygen as a byproduct. As a result, about half of the dry weight of wood is carbon, which

remains sequestered during the life-span of the product and helps offset carbon dioxide emissions–

a major contributor to global warming. One cubic meter of wood stores nearly a metric ton, about

1,900 pounds of CO2 , see Figure 1.2.

Compared to other mineral based construction materials (e.g. concrete, steel, aluminum),

wood requires less energy and water to produce. Life-cycle assessment studies consistently show wood

to be better for the environment in terms of embodies energy, air, water pollution, and greenhouse

gas emissions.

Figure 1.2: Carbon storage of wood. Adapted from Kolb (2008)

Replacing concrete and steel with wood can offset decades of carbon emission from heating,

cooling and powering the building. As advances in technology make it possible to build increasing

sufficient buildings, using wood could result in a negative carbon footprint over the life of the

structure.

3
1.2 Scope of Research and Potential Contributions

CLT has not been fully established in the North America. Compared to Europe markets,

CLT use in the US is small, limited to custom projects where panels are often imported from

Europe or Canada. To translate European successes to the US, CLT performance data is needed

for domestic wood species. Although there are several candidate species for manufacturing CLT

panels in the US, the most likely to be deployed are the softwoods that are now used as structural

building components. Southern Pine, the target species group for this study, is selected attributed

to its superior strength and stiffness, high density and excellent treatability. In addition, Southern

Pine is the most widely planted tree species group in the US and approximately 75% of all seedings

planted each year are Southern Pine (McKeand et al. 2003). Called ‘American’s wood basket’, the

Southern United States produce most of the world’s industrial wood output, providing about 15%

of the world’s industrial round wood and almost 60% of US’s harvest (Smith et al. 2009).

The specific objectives and research activities for Southern Pine CLT research presented in

this study are as follows:

1. to evaluate the CLT manufacturing process, adhesive performance, and structural

performance by pilot production and testing of Southern Pine CLT in Clemson University. The

structural performance study includes bending stiffness, bending strength, and rolling shear

strength in both major and minor strength directions;

2. to review the literature on timber-to-timber connections with inclined self-tapping

screws; estimate strength and stiffness of the timber joint and confirm with experimental study;

3. to investigate the feasibility of developing a long span hollow core composite CLT

beam that meets strength, stiffness and serviceability requirements for office occupancy and

commercial buildings. As a part of this task, a first-floor corridor CLT-Glulam composite

beam is designed, assembled and destructively tested to confirm the design philosophy;

4. to evaluate the vibration serviceability of long span CLT beam by performing

dynamic vibration tests. A series of referenced heel drop modal tests are performed to quantify

the vibration characteristics of the constructed beam. Upon completion of the vibration test,

human perception of floor vibration are discussed. In addition, finite element method (FEM)

models are developed and calibrated using the experimental data.

4
Upon completion of the listed tasks, this research (1) provides solid evidence that Southern

Yellow Pine (a group of softwood species) is viable for manufacturing of structural use CLT panels;

(2) deepens the understanding of layered composite wood material through experimental testing and

FEM composite material modeling; (3) presents a solution for a long span composite beam that is

fastened using the state-of-the-art inclined screws; (4) quantifies the vibration characteristics of the

composite floor via experimental data; and (5) presents a validated FEM model that gives researchers

and engineers a method of predicting the fundamental frequency needed for floor vibration design.

1.3 Organization of the Dissertation

A literature review is presented in Chapter 2. This chapter begins with the discussion of

the current status of CLT development in the US, types of construction and mass timber buildings

that are well-known. Then three major topics of CLT are reviewed and discussed– adhesive, fire

performance and connection. In Chapter 3, the manufacturing process of CLT is presented. In the

first part, the technology of commercial presses and CNC (computer numerical control) machines

are reviewed. The second part of Chapter 3 documents the process of a pilot manufacturing of

Southern Pine CLT conducted in the woodshop at Clemson University. The specimen dimensions

and test matrix are outlined for the preparation of experimental studies performed in Chapter 4

and 5. In Chapter 4, the rolling shear strength of 3-ply CLT is tested and confirmed with the CLT

Standard. Chapter 5 reports the experimental studies for bending characteristics of the 3-ply CLT

panels. Monte Carlo simulations are performed and validated against the experimental findings and

these results are presented in Chapter 5. Chapter 6 begins with the literature review and analysis

of timber-to-timber connection. Tension joint is selected among other loading configurations and

three small shear specimens with inclined screws are tested in static and fatigue loading scenarios.

With the information drawn from the shear specimens, a 40-foot-long composite CLT-Glulam beam

is designed, assembled and destructively tested. The experimental results for bending and shear

strengths of the composite beam are presented in Chapter 6. Chapter 7 addresses the vibration

serviceability of the long span composite beam. Chapter 7 also discusses the instrumented heel drop

tests carried out to determine the modal parameters (frequencies, dampings and mode shapes) of the

composite beam and to validate the finite element models. Finally, summary of research outcomes

and conclusions are provided in Chapter 8.

5
Chapter 2

Literature Review and Background

2.1 Status of the CLT Industry

Europe is by far the biggest producer of CLT with its capacity and production output

increased exponentially since year 2000 (Brandner et al. 2013). The main drive for the CLT devel-

opment and production are the green building movements, change in the building codes, consistent

marketing efforts and the development of distribution channels (Crespell and Gagnon 2010). Origi-

nated in central Europe, there are now a large number of CLT productions within Germany, Austria,

Czech Republic and Switzerland. Figure 2.1 provides an pie chart of CLT providers as well as their

production percentage breakdowns in the year of 2014, with a total volume of 472,200 m3 (that’s

almost the volume of 1300 standard swimming pool!). The companies manufacture CLT from all

kinds of domestic conifers specious. Most of them use Polyurethane (PUR) adhesives; however,

Pelamine-Urea Formaldehyde (MUF) is also used in some productions lines. Of the manufactures

listed, Stora Enso, Binderholz, KLH, Mayr-Melnhof are among the larger volume providers with

worldwide supply capabilities.

Interest in CLT as a new engineered wood product in North America is in its early stage

of development. While the classic timber frame construction with OSB or plywood sheathing still

being a sound options for small single/multi-family construction, CLT creates an opportunity for the

North America wood industry to upscale the size of residential and non-residential wood structures.

In the United States, the International Building Code (IBC) limits wood building up to 4 stories.

However, local jurisdictions retain the authority to approve projects on a case-by-case basis (Cain

6
Figure 2.1: European CLT manufactures as of 2014. Source Grasser (2015)

2014). In order to facilitate CLT application in the North America, the building material has to

be implemented into the regulatory system. This is done through a multi-level phases. First, a

CLT standard is needed guiding the production and quality control of the species dependent panels.

Followed by that, the material has to be incorporated in the design manual with the allowable values

obtained in phase one. As a final phase, the standard and design manual need to be adapted to the

building code. For the first phase, United States published a bi-country CLT standard known as

the Standard for Performance Rated Cross-Laminated Timber (ANSI 2012b). This document has

facilitate the CLT production by providing dimensions and tolerance, performance requirements, test

methods, quality assurance and trademarking. The second phase, the design manual development

in Canada was initiated by the Canadian Wood Council (CWC) to include CLT in the Canadian

Standard Association (CSA) O86 Engineered Design of Wood Standard. In the U.S., the America

Wood Council (AWC) took the same initiative to include CLT in the National Design Specification

for Wood Construction (NDS). The last step, the adoptation of CLT in the building code of Canada

(NBCC) and US (IBC) has already started. In fact, the petition for changes was submitted in 2012

and was approved in 2015, and the current NDS contains a chapter for designing CLT in the US.

7
In addition, a CLT handbook (Gagnon et al. 2011) has been published by FPInnovations to

assist the regulatory authorities with the code proposal and users with their design. These efforts

to foster the application of CLT in North America has resulted in many research and development,

including the recently opened Candlewood Suites (March 2016), a commercial CLT hotel project

in Huntsville, Alabama. Currently in North America, a total of four companies are APA certified

members capable of producing structural use CLT panels (several others are producing matting CLT

for temporary access roads and working platforms in oil & gas industry). They are Structurlam and

Nordic in Canada, and SmartLam and D R Johnson in the U.S. In 2015, SmartLam announced their

plan to massive expand CLT capacity to 110,000 m3 /year, which quadruples the current output and

would become the largest CLT mill in the world upon completion. Meanwhile, European CLT

companies view North America as a promising market. For example, KLH started a sales joint

venture with Idaho Forest Group (March 2014).

From a market prospective, North America might experience a different start of CLT. Since

the traditional light frame construction still remain popular for the majority of the single and multi-

family housing, the market needs a few showcase projects, especially for commercial or industrial

end uses, to demonstrate the advantage of using CLT as building material. As mentioned above,

pre-fabrication of CLT enables fast speed of construction, thus the erection time and ultimate the

cost (from labor, crane and tool rental, etc) are reduced. Over the past several years, a number

of tall wood projects have been completed around the world, demonstrating successful applications

of new wood and mass timber technologies. Among them, the 8-story Wood Innovation & Design

Center (Figure 2.2) in British Columbia Canada was built to LEED Gold status. Similarly, a 7-story

mass timber office building named T3 (timber, technology, transportation) in Minnesota is the first

modern tall wood buildings in the US (Figure 2.3). These pilot projects are crucial to gain further

awareness of CLT in architecture and design community as they are important decision makers in

building projects.

In addition to Europe and North America, there is also CLT development in Asian and

the ‘Pacific Rim’. In Asian the major manufacturer and development is taking place in Japan,

where currently there re three CLT producers in the country (Length Cooperative, Meiken Lamwood

and Yamasa Mokuzai ). CLT development is also underway in Australia and New Zealand. One

of the early CLT projects which still remains the tallest timber apartments during the draft of

the dissertation is the Forte (Figure 2.4) apartment building built in 2012 at Victoria Harbor in

8
Figure 2.2: 8-Story Wood Innovation Design Center (Archdaily)

Figure 2.3: 7-story mass timber office building T3 (Archdaily)

Melbourne, Australia. It currently holds the title as the worlds tallest timber building, but it may

not hold it for much longer. At end of summer 2017, the University of British Columbia will finish

Brock Commons student residence (Figure 2.5). At 53 meters tall with 17 stories, with housing for

404 students, it will soon be the tallest mass wood hybrid building in the world.

2.2 Application of CLT

Cross-laminated timber can be used for a wide variety of constructions. Small-scale projects

like single or multi-family residences, as well as large-scale multi-story buildings for residential,

commercial or public service purposes can be realized. Bridge, towers and other special monumental

structures are also feasible. The load-carrying CLT element can generally be used as wall, floor

or ceiling. In addition, CLT is commonly seen in hybrid buildings, where CLT panels are used

in combination with other wood products, for instance, glulam, nail laminated tumber, but also

9
Figure 2.4: 10-story Forte apartment building (Lendlease)

Figure 2.5: 17-story Brock Commons student residence in UBC(Archdaily)

mineral materials such as concrete and steel (Gagnon et al. 2011).

Due to its simplicity in design and erection, platform construction is most frequently used

as CLT building technique, see Figure 2.6. Platform construction usually involves erection of the

lower story walls and then installation of diaphragms (floor elements), providing a ‘platform’ for the

construction of next story. With the panels properly fastened, CLT elements are able to transfer

both lateral and gravity loads. The fast speed or erection, simple connection system and well-defined

10
load paths are among the advantages of platform construction using CLT. Other than platform

construction, CLT building can also be erected using balloon construction, though it is rarely seen.

In balloon construction system, walls are lifted up continuing several stories without a platform.

Intermediate floors are then attached to the walls, providing rigidity to the finished floor. Given

the size of large wall panels, heavy-duty temporary shorings are often of paramount importance

for construction site safety. In fact, the size of the panels is often controlled by the length of the

semi-tractor trailer: 53 ft (longer trailers will need special permitting). Due to the limitation of

panels size that can be delivered to the job site as well as the extra cost associate with shoring,

balloon systems are usually applied for low-rise, commercial or industrial shell-like buildings where

the absence of internal load or non-load bearing walls would enable a faster erection of the building.

Given the dimensional stability and fast erection speed, among other advantages, mass

timber structural system can be very versatile. The following section discusses the popular framing

plans and the options for lateral system design.

Figure 2.6: Practice of platform construction at Redstone Project (Lendlease)

2.2.1 Gravity framing styles

A buildings framing works in conjunction with its foundation to provide strength and sta-

bility for the structure; it is also a critical component of the load path. Properly designed and

constructed building framing is important in all locations. Framing must transfer all gravity, uplift,

and lateral loads to the foundation. As the section name indicates, the gravity framing is responsible

for sustain and ensure vertical load path of all gravity forces. For mass timber construction, the

11
following three framing plans are commonly seen in the practice: post & beam, two-way panel deck,

and ‘honeycomb’.

Figure 2.7: Types of gravity framings

Post & beam construction is a method of building with heavy timbers (posts and beams)

rather than dimensional lumber such as 2x4s. Traditional timber framing is the method of creating

structures using heavy squared-off and carefully fitted and joined timbers with joints secured by

large wooden pegs. It is commonplace in wooden buildings from the 19th century and earlier. The

method comes from making things out of logs and tree trunks without modern high-tech saws to

cut lumber from the material stock, thus leaving a strong wood architectural finish to the building.

With the advanced cutting tools, post & beam framing has been modernized with the help of the

industrial tools, such as CNC machine. The horizontal deck has a versatile options that literally

could include all composite timber products, e.g. CLT, NLT, GLT, LVL, tongue and groove decking,

etc. The primary framing members (for posts and beams) can be selected from solid-sawn lumbers,

SCL, glulam or other non-wood materials.

Figure 2.8 shows the framing detail of the Carbon 12 –currently the tallest wood structure in

the US. Glue laminated timber columns and beams were used in this project for the gravity framing.

5-ply CLT was selected for their horizontal decking system.

Two-way panel deck, also known as the two-way slab, is designed to resist load in two

perpendicular directions. Contrary to one-way slab, the plate-like two-way slab has bending action

in both directions and the load is transferred to all supporting walls or beams. Due to the nature

of the crosswise layups in CLT, a two-way panel deck system is able to fully leverage this material

to its fullest potential. Two-way slab is common in concrete and steel decking system, however, it

12
Figure 2.8: Post & beam framing: Carbon 12 in Portland OR.

is rather unusual in mass timber construction. The open timber pavilion named Chicago Horizon

has a square, flat canopy roof, measuring 56 feet on each side–the maximum length of timber that

can be legally carried by a truck. This canopy is the first point-supported, two-way slab structure

built with timber, see Figure 2.9. The canopy consists of two piles of CLT panels laid crosswise

to each other. A total of 14 panels, each 8 feet wide and 4.125 inch thick, are topped by rubber

waterproofing membrane and gravel blast. The load of the roof slab is carried in perpendicular

directions and bears directly on 13 glulam columns distributed in a radial pattern, as opposed to

the typical post & beam grid. The canopy spans up to 30 feet between columns at isolated points.

It is also worth mentioning that the Brock Commons at UBC utilize a two-way deck design as well,

the decking system consists of 5-ply CLTs spanning on the 9 feet x 13 feet grid columns.

The concept of honeycomb framing has overlaps with platform constructions. A typical

honeycomb framing involves erecting the wall panels a setting floor decking panels on top, creating

individual, isolated spaces, just like honeycombs. The floor decking often spans in a two-way fashion

between the walls, however, it is only designed for the longer direction (as one-way slab). Therefore

honeycomb framing plan is suitable for accommodating shorter spans that are commonly seen in

hotel or apartment buildings. The Redstone Arsenal hotel in Huntsville Alabama (Figure 2.6) and

13
Figure 2.9: Two-way slab roof: Chicago Pavilion in Chicago IL (Hedrich Blessing)

Stadthaus Murray Grove apartment in London (Figure 2.10) are examples of honeycomb framing

where the entire structures are constructed entirely from pre-fabricated solid timber.

2.2.2 Lateral force system

CLT buildings are becoming increasingly popular with architecture and developers in areas

of high seismic hazard, for example Italy and Japan. This is mostly due to the advantages resulted

from the attractive weight-to-strength ratio of timber buildings. The reduced design inter-story

shear forces and the scale of the foundation will essentially drive down the cost and speed up the

erection process. Joint research activities in Italy and Japan showed that the CLT system can be

designed to provide seismic-resilience self-centering mechanism, thus providing a possibility for its

application in the seismic region (Ceccotti 2008). They tested shear wall components, a three-story

and a seven-story building (SOFIE project) in Japans E-Defense shake table facility and discovered

that connections and joints experienced large inelastic deformations and provided ductility to the

system. The CLT panels, on the contrary, perform rigidly and showed little deformation. In fact, the

tests concluded that the design and layout of the mechanical fasteners have a significant influence on

the performance of the CLT rocking wall system. A number of studies on CLT systems undertaken

in Europe and Canada: i.e. (Ceccotti and Follesa 2006; Dujic et al. 2008; Dujič and Žarnić 2006;

14
Figure 2.10: Honeycomb framing in Murray Grove apartment. Hackney, London (Waugh Thistleton
Architects)

Popovski et al. 2010) have demonstrated that a CLT system can be utilized effectively as a lateral

force resisting system. However, using CLT as lateral force resisting system in the US seismic design

can only be performed through ‘alternative methods’ since the material is not currently recognized

in seismic design codes. In 2009, the Applied Technology Council (ATC) proposed a methodology

published as Federal Emergency Management Agency report P695 (FEMA 2009) which provides a

rationale to evaluate seismic performance factors (SPFs) including the response modification factor

(R-factor), the system overstrength factor, and the deflection amplification factor for seismic design

in the U.S. The objective of the methodology is to provide an equivalent level of safety for all the

structures comprised of different seismic force resisting systems, i.e. approximately a 10% or lower

probability of collapse when subjected to an earthquake having the intensity of an earthquake with

a 2500 year return period (known in the U.S. at the Maximum Credible Earthquake). An overview

of this CLT seismic performance quantification using P695 can be found in Amini et al. (2014).

15
During the prepation of this manuscript, the CLT lateral system, including both vertical

(shear wall, moment resisting frame, etc) and horizontal (diaphragm) lateral systems, are addressed

in neither ASCE 7-10 nor NDS 2015. However, two approaches are available to design the vertical

lateral system as illustrated in Figure 2.11.

Figure 2.11: Lateral force resisting system of mass timber

If code recognized lateral system is selected, typically a hybrid construction can be seen.

Those lateral systems are well understood and have been in practice for long enough where architec-

tures and engineers are familiar with. The R values published in the ASCE 7-10 are readily available

and therefore the lateral system is clearly defined. As a matter of fact, many of the mass timber

structures erected in the US use a hybrid system where CLT or other mass timber product is design

for gravity load only. The examples are Carbon 12 with buckling restrained braces (Figure 2.12),

120 SE Clay with steel moment frame (Figure 2.13) and Roosevelt High School with RC columns

(Figure 2.14).

If the lateral system is not recognized in building code, or as a result of a innovative research,

ASCE 7-10 provides a code compliance pathway to incorporate such system given that analysis or

16
Figure 2.12: Buckling restrained braces in Carbon 12

Figure 2.13: Steel moment frame in 120 SE Clay Street

17
Figure 2.14: Reinforced concrete columns in Roosevelt High School dining hall

combination of analysis and test provides a reliability no less than the expected component designed

per ASCE section 1.3.1.1. It’s up to user’s responsibility to prove the equivalency and reliability of

the proposed lateral system and will be screened by jurisdiction peer review panels for final decision.

In fact, the state of Oregon is one step ahead of the rest where building code advisory board issued

a statewide alternate method to incorporate CLT shear wall as a recognized system (with R=2).

This provision can be found in the Section 602.4 of the 2015 Oregon Structural Specialty Code.

Second part of the lateral system design is to address the diaphragm. Diaphragm works as a

horizontal deep beam, where the loads are usually assumed to be uniform distributed, and resisted by

the diaphragm in bending. The resulting forces in the chords and collectors are the diaphragm panels

as well as connectors need to be designed for. The European approached proposed by Follesa et al.

(2013) assumed a rigid diaphragm bahavior for CLT wall and floor system. Typically, non-yielding

connections specified in the diaphragm panels where the connectors are designed to overstrength

factor 1.3 to 1.6 of yielding connection strength. The US CLT handbook typically assumes a semi-

rigid diaphragm, where the chords are designed with non-yielding connections while other internal

panel joint connections are designed with yielding ones. These yielding connectors shall be governed

by NDS yield modes III and IV and strength of other (non-yielding) limit states to be designed to

nominal yielding connection capacity. A white paper design example published by Spickler et al.

18
(2015) should provide a good start on diaphragm design and selection of connections.

2.3 Joints and Connections

“Connections in heavy timber construction, including those built with CLT, play an essential

role in providing strength, stiffness, and ductility to the structure; consequently, they require careful

attention by designers”(Gagnon et al. 2011).The light weight of CLT combined with high level of

prefabrication, in addition to the need to provide wood based alternative products and systems to

steel and concrete, have significantly contributed to the development of CLT products and systems.

The structural efficiency of the floor system acting as a diaphragm and that of walls in resisting

lateral loads depends on the efficiency of the fastening systems and connection details used to

interconnect individual panels and assembly together. According to the classification given in the

CLT Handbook, there are basically five types of connection assemblies for CLT buildings, they are:

(A) Floor panels connections

(B) Wall-to-wall connections

(C) Wall-to-floor connections

(D) Wall to roof connections

(E) Wall-to-foundation connections

Preferably, large-sized panels are to be used since this minimize the number of connections

and joints. However, as mentioned above, given the dimension in transportation and production,

this is not always possible.

Currently practices mainly select two types of fasteners: dowel type fasteners like nails, self-

tapping screws, rivets, dowels and bolts; and bearing-type fasteners such as medal brackets, split ring

and shear plates. Structural adhesive can also be used in combination with the mechanical fasteners

to achieve if a more rigid connection is desired. Given the introduction of CLT as construction

material in the market of North America, it is expected that new and innovative connection details

will be developed over time.

The database of CLT connector catelog is too large to cover thoroughly, some of those are

made for special situations and are rarely seen in typical details. Therefore the rest of this section

19
Figure 2.15: Illustration of CLT connections (Gagnon et al. 2011)

will be focused on a selection of generic CLT connectors.

Self-tapping screw Self-tapping screws has been extensively researched and used in Eu-

rope for the assembly of CLT elements. In many cases they are installed in an inclined angle to take

advantage of the high combined lateral and withdraw capacity. It receives popularity among designer

and builders as it is easy to design and install. Self-tapping screws are available in diameter up to

0.55 inch and length up to 59 inches, while common lengths are up to 39.4 inches (1000 mm). They

do not require pilot holes in most cases, unlike lag screws or through bolt used in traditional wood

construction, the size of which depends on the density of the wood and the diameter of the screw. It

is recommended that the design capacity of screws in CLT to account for potential gaps in unglued

cross layers. 45 degree is a typical angle of installing such long self-tapping screw. To facilitate

the installation, special washers are available to assist the measurement of angle at job site. These

washers (Figure 2.16) guides the installation of self-tapping screws and are usually countersunk into

20
the steel plate underneath.

Figure 2.16: Inclined self-tapping screw installed with washer guide (Rothoblaas)

Beam hanger Beam hangers are commonly seen in post & beam type of framing where

beams need to either connect to a wall or to a post (column). The connecting object can be either

wood, steel or concrete. Based on prescribed aesthetic requirements, architects or building owner

may select between a hidden joint or a visible joint. A hidden joint may cost more initially, however,

the joint assembly has a potential to achieve better fire rating since the connection is adequately

protected and isolated by the surrounding timber. Unprotected joint, on the other hand, is directly

exposed to the fire and fasteners quickly loss their penetration depth during the char development.

Figure 2.17: Exposed (left) and concealed (right) beam hanger (Rothoblaas)

21
With the help of CNC capability, ends of the connecting pieces can be pre-routed and

attached with countersunk hook connectors. Wood screw are typically used as fasteners for the hook

and many products in the market are referred as ‘identical twins’, meaning there’s no distinguished

male and female part of the connection assembly. So installation error can be minimized. The

installation of beam hooks can be easily done at fabricator, thus enables a fast beam drop-in which

requires only crane and some adjustments (with plumb, carpenter square and level).

Figure 2.18: Concealed hook connector connecting a glulam to a ledger beam (Rothoblaas)

Angle bracket Angle brackets are connectors that are widely used for mass timber con-

struction. After punching, shearing and cold forming, plane galvanized sheet medal is transformed

into three dimension angles. Provided its dimensional force resistance capacity, the application of

angle brackets can be very versatile that include connections of right-angled walls, wall to floor, wall

to roof and wall to foundation. Generally speaking, timber rivet, nail or screws are used to fasten

the angle bracket for wood to wood application, while for concrete and steel, bolts are more likely

to be used. Angle brackets are often engineered and reinforced based on its intended application.

For example, if the angle is to be installed at the end of the wall where uplifting and over-turning

demand is high, a portion of the angle will be curled up to act as stiffeners (see Figure 2.19). For

shear applications, special ribs will be cold formed to achieve better resistance.

For CLT shear wall as lateral system, angle brackets play a important role as they provide

ductility and shear resistance to the building. However, due to the lack of generic design of geometry

22
Figure 2.19: Hold down and shear bracket(Rothoblaas)

and material, angle brackets from different manufacturers often comes with different resistant char-

acteristics. Therefore the design (location, number of brackets, etc) and selection of angle brackets

should proceed with great caution since CLT shear wall is not a recognized lateral system yet and

there are no available R value that could guide the lateral design. On this regard, research has been

conducted to examine the performance of CLT shear walls (Popovski and Karacabeyli 2012) and

preliminarily quantified the seismic performance factor using generic angle brackets (Amini et al.

2016; Pei et al. 2012).

Perforated plates Perforated plates are plates with pre-pierced holes, they are are in-

tended to transfer tensile forces between timber elements, such as beams, structural panels and

claddings. Perforated plates can be also used as bracing element. At locations where CLT panels

are jointed in together, multiple perforated plates are installed to control the construction joint and

provide integrity of the diaphragm, see Figure 2.20.

2.4 Fire performance of CLT

Timber structures have experienced a renaissance during the recent few decades due to their

environmental credentials, and societal goals striving for sustainable development with lower energy

demands and less pollution in all sectors including the construction sector that stands for a major

part of the overall community economy. However, the combustibility of timber is still the concern

of general public as using wood as a building material. It is restricted in building regulations in

most countries, especially for higher and larger buildings. In the United States, compliance with

23
Figure 2.20: Perforated plate installed at diaphragm joint in Carbon 12 project

the building code is generally accomplished by construction in accordance with the International

Building Code (IBC) or Building Construction and Safety Code (NFPA 5000). The intention of

these documents is to establish minimum requirements to the public safety through a list of targets,

for example, life safety, structural strength, property protection, fire-flighter’s safety, etc. As such,

the fire design may deal with situations such as providing adequate structural integrity in a fire

scene, limit fire spread through a building and to adjacent properties as well as limit personal and

property damage due to a fire impact.

2.4.1 Fire resistant requirements

The minimum fire resistance requirements set forth in the IBC and NFPA 5000 depend on

the structural element, type of construction, occupancy, distance from property line and other factors

such as other detailing requirements articulated in IBC Chapter 4. The type of construction and

occupancy classification directly dictates the height and area limitations for code compliance. Mass

timber construction with CLT is able to comply with the provision in Types of Construction III, IV

and V, since Type I and II requires that major building element to be built with non-combustible

materials. Based on the type of construction, locations where CLT can be used are summarized

below:

24
Type III construction Type III construction requires that the exterior walls to be non-

combustible while allowing combustible material being used as interior building elements. Type III is

further divided into sub-classifications A and B based on fire resistance requirements (e.g. sprinkler

system).

Type IV construction Type IV is commonly known as ‘Heavy Timber’ construction,

it requires the exterior walls of being noncombustible materials and the interior building elements

of solid or laminated wood without concealed spaces. The 2015 IBC prescriptively allows the use

of CLT in Heavy Timber construction, including exterior walls (section 602.4.2), floors (602.4.6),

roofs (602.4.7) and interior walls (602.4.8). For exterior wall assembly, CLT is permitted with a

fire rating 2 hours or less and surface of the CLT being protected from one of the following: fire-

retardant-treated wood sheathing not less than 15/32 inch thick; gypsum board not less than 1/2

inch thick or a noncombustible material. If CLT is used as floor deck, it shall be not less than 4 inches

in thickness and shall be continuous from support to support while mechanically fastened to one

another. CLT used in roof assembly has the identical requirements except the minimum thickness

decreases to 3 inches. Load bearing interior walls shall have a one hour fire rating, regardless the type

of wood product being used in the assembly. While if interior partitions are formed by laminated

construction, they shall not be less than 4 inches.

Type V construction Type V is defined as a type of construction in which the structural

elements, including exterior and interior walls, can be any material permitted by the IBC code. For

CLT, the 2015 IBC and NFPA 5000 both refers to the ANSI/APA PRG 320 standard for code

permitted commodity. As such, CLT complying the Standard would be permitted for use in Type

V construction.

Type I and II construction Types I and II are those types of construction in which

the building elements are of noncombustible materials. These building elements listed in Table 601

of the 2015 IBC include primary structural frame, bearing walls, non-bearing walls and partitions,

floor and roof assemblies. However, combustible wood products are still permitted in Type I and

II constructions, given they are fire-retardant-treated and used in the following situations: 1) non-

bearing partitions where the required fire resistance rating is 2 hours or less; 2) non-bearing exterior

walls where fire-resistance-rated construction is not required; 3) roof construction.

25
Fire ratings Once the occupancy classification and type of construction are determined

and approved (so as the floor area and number of stories), fire rating requirements are assigned to

the building sub-assemblies. IBC 2015 fire-resistant rating (hour) requirements for various building

elements are screen shot here as Figure 2.21. Based on the required fire rating hours, architects and

engineers work together to select material and perform calculations so that casualty and property

damage are minimized in case of a fire.

Figure 2.21: Fire-resistant rating requirements for building elements. IBC 2015, Table 601.

A fire-resistance rating is defined as the period of time of a building element, component or

assembly maintain the ability to perform its separating functions (i.e. confining fire by preventing

or retarding the passage of excessive heat, hot gases or flames), continued to perform a given load-

bearing functions, or sometimes both. Structural fire performance of buildings is often assessed by

conducting fire-resisting in accordance with ASTM E119 (ASTM 2016b), where the test assembly

is subjected to a standard furnace temperature time history.

In case of exploring the deterioration of load-bearing functions under fire, specific test

method outlined in the E119 requires specimens to be loaded while being subjected to prescribed

temperature profile. It also requires the superimposed load to be the maximum load condition

26
allowed under natinoally recognized structural design criteria. As such, a test performed under

maximum load ensures that the fire-resistant rating obtained is appropriate for use in any equal

or less loading conditions. Primarily, a standard fire-resistance test entails three failure/acceptance

criteria:

1. Load bearing function: the assembly must support the applied load for the dura-

tion of the test;

2. Separating function: the assembly must prevent the passage of flame or gases hot

enough to ignite a cotton pad;

3. Insulation function: the assembly must prevent the temperature rise on the unex-

posed surface from being greater than 325 ◦ F at any location, or an average 250 ◦ F measured

at a number of locations, above the initial temperature.

2.4.2 Experimental studies in CLT fire performance

Recent fire tests (Frangi et al. 2008, 2009; Hasburgh et al. 2016) confirmed that the CLT

has the potential to provide good fire resistance, often comparable to typical mass assemblies of

non-combustible construction. This is found to stem from the nature of the thick timber members

to slowly char at a predicted rate, allowing the building system to maintain significant structural

capacity for extended durations. The charred layer further serves as a protection layer and reduce

the char rate of the core. Hasburgh and Bourne (2016) fire tested a series of Southern Pine CLT with

various layer configurations, adhesives and gypsum protections and they pointed out the significant

effect of adhesive bonds. Delamination could occur at the glue line where the char layer falls off

before being completed charred, leaving the remaining system unprotected and further vulnerate

the integrity of the structural system. They recommended that formaldehyde based wood adhesive

to be selected for CLT since they perform well under fire.

CLT panels with different ply configuration with and without gypsum protection were tested

by FPInnovations at the National Research Council’s fire lab (Osborne et al. 2012). Char rates and

fire-resistant ratings were determined for the panels, generating a good data set for subsequent

studies.

Another experimental fire study at CLT compartment level was conducted by McGregor

at Carleton University considering different level of fire load and protection levels(McGregor 2013).

27
The tests involve typical furniture fire load and a propane fire scenario. Fire in a protected room

with only furniture fuel was shown to self-extinguish, with CLT remain unaffected. The unprotected

CLT panel can delaminate and add to fire load. With propane fire, even protected room can fail

and initiate second flashover after the gypsum board failed (due to extensive charring).

Frangi and his collaborators (2008) conducted a full size fire test where a residential dwelling

in a three-story CLT building was exposed to a fire load of 790 MJ/m2 for an hour. The building

was constructed with 85mm wall panels and 142 floor panels. In addition, the walls were protected

with fire rated gypsums boards and the ceiling with mineral-wood insulation and gypsum board.

Post fire analysis of the building showed a char rate of 5-10mm on the CLT elements. No window

failure or fire penetration in the upper story was observed during the one-hour fire experiment. The

authors concluded that it is possible to limit the fire spread to one room by protecting the CLT with

sufficient non-combustible gypsum material.

Protected compartment fire tests were conducted by American Wood Council in 2015 for

a 16 by 12 ft CLT room with typical fume and fire generated from furniture and contents (book

shelf, TV, sofa, table, etc). The compartment was build with 5-ply CLT all around and had a

door opening of 6.5 ft by 3.5 ft. Test results confirmed observations from earlier compartment fire

tests that gypsum protected CLT (2 layers of 5/8 Type X gypsum wallboard) can achieve nearly

damage free performance during a fire burn out event. A room with only one wall unprotected has

acceptable performance under fire, while the room with two walls unprotected can experience CLT

delamination.

Figure 2.22: CLT compartment fire test (left) and post-fire compartment (right). American Wood
Council

Currently, Chapter 16 of the 2015 NDS incorporates a design methodology to calculate fire

28
resistance of CLT using a linear char rate approach. The effective char rate (inch/hr) βef f adjusted

for exposure time (t) using nominal linear char rate based on a 1-hour exposure can be expressed

as:

1.2βn
βef f = (2.1)
t0.187

A nominal char rate (βn ) of 1.5 inches/hr is commonly assumed for solid-sawn, glulam and CLT.

The effective char depth assumes linearity with the char rate:

achar = βef f t = 1.2βn t0.187 (2.2)

Subtract the char thickness from original member size, the remaining section can be used to

approximate the member strength under fire conditions. On the conservative side of the calculation,

when the char depth falls within a cross ply, it is recommended to increase achar to the nearest inner

ply of the major strength direction. NDS limits the linear char model methodology to ratings not

exceeding 2 hours. For extended fire ratings, additional data is needed to validate this model.

2.5 Wood Adhesive

2.5.1 CLT wood species

There two main types of wood in construction use– softwoods and hardwoods. Botanically

speaking, softwoods are those woods from gymnosperms (mostly conifers) and hardwoods are from

angiosperms (flowering plants). In the temperate portion of the Northern Hemisphere, softwoods

are generally needle-leaved evergreen trees such as pine and spruce, where as hardwood are typically

broad-leaf, deciduous (shedding annually) trees such as maple and birch. It should be noted that the

terms softwoods and hardwoods refer to the botanical characteristics of the wood and are not related

to the material hardness. The single most important distinction between the two general kinds of

wood is that hardwoods have a characteristic type of cell called a vessel element (or pore), whereas

softwoods lack of these, see Figure 2.23. Wood has a well defined cellular structure. In softwoods

and ring-porous hardwoods, the cells produced in the early growing season, the early wood, are

larger than those produced later in the season, the latewood. In diffuse-porous hardwoods the cells

are more uniform in size. The differences in the cell structure are responsible for the variation in

29
density observed among wood of difference species and even in wood of the same species, depending

on local growth condition.

Majority of the European CLTs are made of softwood tree species. Examples are Nor-

way spruce (Picea abies) timber, white fir (Abies alba), scots pine (Pinus sylvestris), and rarely

European larch (Larix decidua), douglas fir (Pseudotsuga menziesii) and Swiss stone pine (Pinus

cembra) (Brandner et al. 2013). Industries and academics have begun to look into the use of

hardwood/softwood and hardwood only species for CLT product development (Mohamadzadeh and

Hindman 2015; Kramer et al. 2013). In the Southern United States forests (also known as the

Southern Forests), softwood species of Southern Pine would be an ideal candidate to replicate the

European success of CLT. Its group species (loblolly, longleaf, shortleaf and slash pines) make up

the largest proportion (17% in 2012) of the softwood resource in the Southern Forests (Oswalt et al.

2014). The development of softwood CLT in the US would surely boost the saw timber industry as

well as generate more employment opportunities around this commodity.

Figure 2.23: Microscopic pictures of hardwood (balsa) and softwood (pine)(Andersson et al. 2006)

30
2.5.2 Background on wood adhesive

The recorded history of bonded wood dates back at least 3,000 years to the Egyptians (Skeist

and Miron 1990) and adhesive bonding goes back to early mankind (River 2003). Although wood

and paper bonding are the largest application for adhesives, better understand of the critical aspects

of wood adhesion should lead to improved wood composites. Understanding how an adhesive work

is difficult since adhesive performance is not one science of its own but the combination of many

sciences. Ones needs to consider both the chemical and mechanical aspects of bond strength and the

interrelation between the two factors. The bonding mechanism is beyond the scope of this chapter,

however, extensive research has been performed on this subject. The interested readers are referred

to the textbooks including Marra (1992) and Dillard (2002).

There are generally three steps in the process of adhesive bonding. The first is usually the

preparation of the surface to provide the best interaction of the adhesive with the wood. Preparation

of the surface can involve either mechanical or chemical treatment or a combination of the two. The

second step is that the adhesive needs to form a molecular-level contact with the wood; thus, it

should be a liquid form so that it can develop a close contact with the object. The third step is

the setting, which involves the solidification or curing of the adhesives. The solidification change

the physical state of the adhesive during bonding process and solidification depends on the type

of adhesive. For polymer adhesives, polymers are dissolved in solvent, which can be simply water

(white glues) or an organic (rubber cement). The lost of solvent converts these liquids to solids.

The other type of adhesive is made up of small molecules that polymerize to form the adhesive,

for example super glue and two-part epoxies. Most wood adhesive involves both solvent loss and

polymerization processes. Once the bond is prepared, the strength of the bond will have to resist

both internal and external force seen during the life of the assembly: forces developed during the

curing of the adhesive, the differential expansion/contraction between adhesive and wood during

environmental changes, and any applied external forces.

2.5.3 Wood surface preparation and setting

On the larger scale, wood is porous, cellular and anistropic material. It is porous in that

water and low molecular weight compounds will be rapidly absorbed and move along the wood fiber.

As discussed in the earlier section, cell types and sizes are remarkably different between hardwood

31
and softwood. The individual species in each of these classes varies considerably in their ability

for liquid to penetrate, let along the difference between early wood and late wood, sapwood and

heartwood. The surface preparation has been shown to have a large effect on the quality of the bond

(River et al. 1991). The best method for preparing a wood surface for bonding is to use sharp planer

blades. Unsharpened blades can crush cells and cause a very irregular surface (River and Miniutti

1975). For wood laminates, ASTM D2559 prescribes that the wood surface be planned with sharp

blades and then be bonded within 24 hours to provide the most bondable surface (ASTM 2012b).

Once an adhesive is applied to the wood, the adhesive needs to set to form a product

with strength. Set is to convert an adhesive into a fixed or hardened state by chemical or physical

action, such as condensation, polymerization, oxidation, vulcanization, hydration, evaporation or

mechanical pressure. More often than not, a combination of settings are required for curing the

adhesive. Penetration of the adhesive in the wood is an important part of the bonding process.

Green wood is difficult to bond with most adhesives because there is little volume into which the

adhesive can penetrate. On the other extreme, overly dry wood can also be difficult for the adhesive

penetration because the wood surface is harder to wet (Christiansen 1994). Therefore, wood with a

4% to 12% moisture content range is typically good for optimum adhesive penetration and setting.

2.5.4 Common wood adhesives

The heat-cured adhesives are very commonly used in the wood furniture industry as it

provide sufficient time for component assembly before heated press, and soften the wood, allowing the

jointing wood surface to be brought into close contact. However, a room temperature cure is better

for thick laminated panels because the heating a bulk element is more difficult. For manufacturing

of CLT, low cost and rapid setting of adhesive are more important factors. Typical wood structural

adhesives include formaldehyde reactive adhesives, isocyanates adhesives, epoxy adhesives and bio-

based glues. The rest of the section will focus on several of the commonly used, low cost and

commercially available structural wood adhesives.

Formaldehyde adhesives The most common wood adhesives are based on reactions of

formaldehyde with phenol, resorcinal, urea, melamine, or mixtures of thereof. The formaldehyde

adhesives are usually water-borne resins so that the curing process is not only polymerization, but

also the loss of water solvent. Since the setting process involves water, too much water or too little

32
water retard the reaction. Therefore control of both open and closed assembly times are important

for both adhesive penetration and water content of the bondline (open assembly time refers the

period between the application of adhesive and the time when the pieces have to be bonded; closed

assembly time refers to the period after the bonding surfaces have been joined during which the

pieces can still be repositioned).

Phenol formaldehyde (PF) adhesive is the oldest class of synthetic polymers, having been

developed at the beginning of the 20th century. The resins are widely used in both laminations and

composites because of their outstanding durability, which derives from their good adhesion to wood,

the high strength of polymer and excellent stability of the adhesive. The PF adhesives could serve

in almost all wood bonding applications, except that the adhesive cures slowly at room temperature

that often requires heating for fast setting. The most common additive to PF is resorcinal to

provide room temperature curing ability. Phenol-Resorcinal Formaldehyde (PRF) is curable at

room temperature with sufficient time to mix the components, spread them on the wood and press

the wood pieces together prior to adhesive setting. Like the PF resins, PFR forms very durable

bonds for failure and degradation. The only drawback of the PRF has been the cost of resorcinal.

Like PF and PRF, Melamine Formaldehyde (MF) adhesives has acceptable water resistance,

but it is much lighter in color compare to others. As such, MF is typically used for exterior or

architectural component as light-colored resins are more aesthetic pleasant. The MF adhesive needs

to be activated to give good polymerization to the final product. The catalyst (activation agent)

usually involves acids that lowers the pH of the resin to accelerate the polymerization process.

Therefore, respiratory protection during adhesive mixing is key for manufacturing safety. Similar to

PRF adhesive, the limitation of MF lies in its cost of melamine.

Isocyanates wood adhesive Several classes of adhesive used in wood bonding involves

the use of isocyanates. Isocynates are most often used to produce polyurethane by reacting with

liquid diols. The high reactive of isocyanates is both an advantage and a disadvantage. The advan-

tage is that polymerization processes rapidly and usually to high conversion. The disadvantage is

that they can react rapidly with many compounds in the human bodies. These reactions are rapid

under physiological conditions and are not readily reversible which means the safety of handling

isocyanates is a concern. The most common type is an self-curing isocyanate, which reacts with

water in the wood to initiate the curing process.

33
The water activated, self-curing Polyurethane (PUR) adhesives are widely used in coating

and wood bonding. To obtain good wetting, a solvent is typically spread prior to the application of

adhesive to reduce the viscosity for good wetting. Polyurethane can be either one or two component

system, while European CLT manufactures often incline to the one-component moisture active

Polyurethane due to its simplicity in the adhesive spraying system.

Another common adhesive is the emulsion-polymer isocyanate (EPI), which is a two-component

adhesive. It has been used for bonding of plastic to wood surfaces, OSB web to the flanges to make

I-joints. The ability to bond plastics and other non-wood materials is an advantage of the EPI resins

over many other wood adhesives. EPI also forms fairly durable bonds depending on the formula,

however, numerous CLT fire tests has reported poor heat performance of isocyanates wood adhesives

(Kippel et al. 2014; Hasburgh et al. 2016).

34
Chapter 3

Commercial CLT Manufacturing,

Clemson Pilot Production and

Specimen Preparations

The first objective of the this chapter is to overview the current practice of commercial

manufacturing of CLT panels in the Europe and United States. The entire production process is

divided into seven steps, from the sawn timber selection to final product packaging. Key machineries

necessary for the production are introduced. Followed by the discussion of industrial standard

practice, the exploratory small scale production of CLT panels in Clemson University is presented.

The pilot production was performed in the Forestry Department’s wood shop, with a 4 feet by 4

feet hydraulic press. Due to the limited size of the press, the production process was evaluated for

the possibility of accommodating larger test specimens.

After evaluation of the production process, ‘V3’ designated 3-ply Southern Pine CLT was

produced in a laboratory setting for the purpose of investigating its adhesive as well as structural

performance. The second objective is to understand the performance requirements outlined in the

PRG-320 document and to experimentally test the laboratory produced CLT panels. As part of

the structural performance quantification, rolling shear and bending properties are mandatory to be

examined for the plant pre-qualification. Note the target here is not to certify a APA recognized CLT

product, but to explore the feasibility of CLT made of domestic softwood lumber species: Southern

35
Yellow Pine. The test methods, specimen preparation are discussed as the rest of this chapter.

3.1 CLT production—European Experience

In general, the production of CLT is quite similar to that of glulam (glued laminated timber).

Base on the knowledge accumulated in Europe, the production of CLT can be categorized into the

following major steps:

1. Raw lumber selection

2. De-stacking, M.C. measurements, E-rating (optional) and trimming

3. Finger joint of lumbers

4. Planning and cutting of finger-jointed continuous lumbers

5. CLT layer formation, adhesive application and assembly pressing

6. Finishing and CNC machining

7. Marking, packaging and shipping handling

In 2011, APA-the Engineered Wood Association in collaboration with American National

Standard Institution (NIST) published a harmonized bi-country North American CLT production

standard used by both United States and Canada: Standard for Performance Rated Cross-Laminated

Timber, also known as ANSI/APA PRG 320 (ANSI 2012b). This standard has provided dimensions,

manufacturing tolerance, minimum performance requirements, test and quality assurance methods

for manufacturing of CLT of different grade, wood species and layup configuration. The above

mentioned CLT production steps will be discussed in detail with discussion of PRG-320.

3.1.1 Raw lumber selection

CLT is manufactured with laminations of dimension lumber or SCL (structural composite

lumber) such as laminated veneer lumber (LVL), laminated strand lumber (LSL) or orientated

strand lumber (OSL). Although PRG-320 permits the use of SCL when qualified with ASTM D5456

(ASTM 2017), however, in reality, it is still years away before SCL could be used in CLT production

because of the apparent challenge in the face bonding of SCL to SCL or SCL to lumber (Yeh

36
et al. 2012). Therefore, the use of SCL in CLT is beyond the scope of this study. PRG-320 draws

the European experience in the manufacturing processes of CLT, and takes into the consideration

of the North American lumber resources, manufacturing preference and end-use. For example, the

standard permits the use of any softwood lumber species and species combinations recognized by the

American Lumber Standards Committee (ALSC) with a minimum specific gravity 0.35, as published

in the National Design Specification for Wood Construction (NDS). One advantage of using standard

grade lumber is that such lumber will be typically marked as ‘HT’ (heat treated), meaning that the

resulting CLT product will also meet national or international phytosanitary requirements when

the traceability requirements of the lumber laminations can be properly demonstrated and certified.

Consistent lumber properties and qualities are of great importance in plant process of raw material.

Usually, lumber arrives at production facility pre-graded and kiln-dried to meet specified demands in

strength, stiffness and moisture content. Lumber grades in the parallel layers of CLT are required to

be at least 1200f 1.2E-MSR or visually graded No.2 and visually graded No.2 for perpendicular layers.

The moisture content of the lumber at time production should be 12%±3%. It is worth mentioning

that sometimes moisture content of the lumbers is purposely adjusted for adhesive applications, since

many of the Polyurethane based adhesives are very sensitive to the bond surface moisture content.

The net lamination thickness of all CLT at time of adhesive application is required to be at least

5/8 in., but not thicker than 2 in. to facilitate bonding. In addition, the lamination thickness is not

permitted to vary within the same layer except when it is within the lamination tolerance: ±0.008

in. across the width of a lamination and ±0.012in. along the length of a lamination at time of face

bonding. The tolerance requirements are typically achieved through planning process which will be

discussed shortly.

3.1.2 Trimming and finger jointing of lamellas

Lumbers enter the production line at a feed-in station, where they are first de-stacked.

At this moment, the additional measurements of moisture content and visual/machine grading are

optional. However, a trimming process is often administered either fully automatically or manually

with a fluorescent pen which can be scanned and detected by a trimming machine. Lumbers with

excessive defects therefore can be expelled from the production line.

After the trimming saw, the lumbers are of random length. To form a continuous lumber

piece suitable for desired parallel and longitudinal layers, piece-wise lumbers are often jointed by

37
means of finger joint.

The joints are milled in a clear wood area where no defects such as wane or knots are located

using a finger joint bit (Figure 3.1 shows a cutting bit for horizontal joint). European CLT producers

utilize two geometries of finger joints: vertical finger joints and horizontal joints (Figure 3.2). On

an aesthetic prospective, horizontal finger joints are harder to identify on the lumber wider surface

and thus they are barely noticeable on the out layers of the finished CLT product. This added

visual quality thus can be used for architectural CLT product, for instance, exposed beams, arches,

columns. In addition, horizontal joints also can be advantageous for thermal insulation of the CLT

elements such as air tightness and heat transfer (Brandner et al. 2013).

Figure 3.1: Horizontal finger joint bit (Infinity Cutting Tools)

Figure 3.2: Horizontal (left) and vertical (right) finger joint

Once the lumber ends are machined, wood adhesive is applied for mating process, see Figure

3.3. In Europe, finger joints in CLT productions are mainly facilitated with the Melamine-Urea

38
formaldehyde (MUF) or one component Polyurethane (PUR) adhesive. Those adhesives establish

an almost colorless bond line. Furthermore, they are characterized by resistance to exposure of

sunlight, humidity and hydrolysis. After adhesive application, the joints have to be aligned by

means of bonding or mating pressure, see Figure 3.4 where one piece of the lumber is coated with

adhesive, waiting for the mating process. This is typically done by a ‘crowder’ : the stop-and-go

system or a mini-joint system. Regardless of which system is used, enough pressure force is required

to ensure a solid alignment of the lumbers (Jokerst 1981). PRG-320 requires that the quality and

strength shall be qualified in accordance with the ANSI/AITC A190.1 (ANSI 2012a), which is a glue

laminated timber product specification. Brandner et al. (2013) also indicated that if manufactured

correctly, finger joint could exceed strength properties of the lumber segments.

Figure 3.3: Vertical finger joint machined with glue applied

3.1.3 Planning and cutting of jointed continuous lumbers

After finger joint process, the lumbers need to be planned for thickness tolerance require-

ments and adhesive application. Commonly, this is done for all four sides of the lumber to ensure

specification in terms of width and thickness. The planning process conditions the lumber for CLT

lamella face bonding since it reduces oxidation and smooths the surface for better adhesive spray.

As a matter of fact, both adhesive suppliers and ASTM D2559 (ASTM 2012b) recommend that

planning or any surface treatment to be done within 24 hours of the adhesive application. After the

39
Figure 3.4: Two pieces of lumber ready to be mated in a crowder

planning procedure, the lumbers are cut to desired length based on their intended layer.

3.1.4 CLT layer formation, adhesive application and assembly pressing

Based on European experience, the CLT layer formation can be done in two ways. Some

of the procedures manufacture single layers by adhesive edge bonding to limit gaps to an absolute

minimum. Then these layers are further processed to form final CLT product. Others form CLT

panels directly from face bonding lamellas without the intermediate steps of edge bonding (Brandner

et al. 2013). Both manufacturing options are discussed in the following paragraph. Production of

single layers has several advantages. Since single layers are edge bonded under pressure, its physical

properties are more stable since it reduces gaps between lamellas in a single layer. In addition, pre-

manufactured layers are already smooth and with its thickness in tolerance. Therefore, the pressing

pressure for final CLT element can be reduced. Another advantage that comes with single layer is its

relatively light in weight that can be easily maneuvered by vacuum lifting devices. However, single

layer pre-manufacture is an intermediate production step which may introduce additional initial

equipment cost that is not absolute necessary in the CLT production. Meanwhile, the disadvantages

of edge glued layers comes from the checks and splits of lamella that are resulting from swelling

and shrinkage due to climate change, see Figure 3.5. This degrades the visual appearance of the

panels and also negatively impact the physical properties over time. Unbonded lamellas, on the

40
other hand, can be used for single layer formation and CLT assembly. Following this approach, the

lamella edges are sporadically connected during face bonding (due to the adhesive being squeezed

into edge surfaces during press). In this case, swelling and shrinkage takes place between the boards

and thus check and splits can be avoided to a certain extend.

Figure 3.5: Swelling and shrinkage checks in CLT with edge bonded layers (left) and without edge
bonded layers (right)(Brandner et al. 2013)

The CLT layer forming is realized in several ways and depends on the production line design

and capacity, different lifting machines are used. In low capacity plants, CLT layups may also be done

by human operators. The assembly process is comparable to plywood production where adjacent

layers are also assembled perpendicular to each other (Gagnon et al. 2011). Between each layer,

adhesive is spread mechanically on single lamellas in continuous through-feed nozzles. A line-wise

applicator will run through the bonding surface within the adhesive open assembly time, see Figure

3.6 for fully automatic adhesive application machine.

Figure 3.6: Line-wise adhesive application(Ledinek Engineering)

After the layup of the CLT element, a pressure device is need to bond the lumber lamellas

and set the adhesive to form a composite wood product. According to Brandner et al. (2013),

pressure is a parameter that depends on the following manufacture specifications: (1) adhesive

system; (2) timber species; (3) the geometry of the adherents in regard to roughness and flatness of

the surface and allowed tolerance in the thickness; (4) the adhesive application system and (5) the

41
applied quantity of adhesive. Those specifications are discussed in the following.

The pressing pressure and time have to meet the requirement of the adhesive related pa-

rameters. Above all, the recommendations of the adhesive representatives need to be respected.

Enough pressure is important to equally distribute the line-wise applied adhesive over the whole

layer surface to guarantee a consistent and defined bond line thickness. Furthermore, the adhesive

type and spread rate have to be factored into consideration. As Brandner et al. (2013) stated, types

of adhesives can be differentiated first into close contact and gap-filling adhesives and secondly into

swelling (e.g. Polyurethane) and shrinkage adhesives. These adhesive characteristics have a signif-

icant impact on the required clamping pressure. Formaldehyde based adhesive systems typically

requires a lager pressure of 200-300 psi, while Polyurethane bonding can theoretically be done with

2-15 psi (Sellers Jr 2001).

In addition to pressure considerations associated adhesives, lumber characteristics and

species also need to be accounted for. Though the lamellas are finger jointed and planned, the

natural variation in the lumber thickness still exists. Therefore, enough pressure is necessary to

smooth those characteristics. However, excessive pressure can also lead to damage of wood cell

structures, and thus to a reduced shear resistance and adhesive penetration.

In general, there are three types of mechanical press system available in the engineered

wood industry, each with an optimal operating pressure: hydraulic press (<300 psi), pneumatic

press (<150 psi) and vacuum press (<15 psi). It should be noted that hydraulic and pneumatic

press are usually used for medium to large scale CLT plate due the size of the panels they able to

produce. In addition to surface pressure, some of the presses are able to provide side pressure to

reduce gap within the panel. Depend on the adhesive application requirement, a hot press is also

available for accelerated curing.

Hydraulic and pneumatic presses (Figure 3.7 and Figure 3.8) are among the more popular

choice for larger CLT producers in Europe since it is relatively more automatic and more efficient

than a vacuum press. Some producers may prefer pneumatic presses since no hydraulic fluid is needed

for accumulate pressures thus provides a cleaner and environmentally friendlier plant environment.

However, in practice a pressure of around 120 psi is applied and both systems are equally capable

of.

Another mechanical device for bonding CLT elements is the vacuum press, see Figure 3.9.

Vacuum press is able to offer a pressure of less than 15 psi. Due to the lower pressure forces, the

42
Figure 3.7: Hydraulic CLT press(D.R.Johnson)

Figure 3.8: Pneumatic press (Ledinek Engineering)

requirements of timber surface quality, thickness tolerance and wood defects like wane and warp

needed to be strictly enforced. The limitation of the vacuum press is its capacity to press thick

elements. However, utilizing a vacuum press for CLT production allows great flexibility in the

shape of the finished product. For example, panels with large cutouts, curvature or complicated

3-dimensional shapes can be easily enclosed in the vacuum cover. Meanwhile, the initial investment

and yearly production capacity is less than those of hydraulic and pneumatic presses. Therefore, it

makes a good option for market entry strategies or product portfolio extension for small enterprises.

43
Figure 3.9: Hydraulic CLT press (Ledinek Engineering)

3.1.5 Finishing and CNC machining

Due to the external pressure in the assembly process, excessive adhesive resins are squeezed

onto the edge of the panels. Meanwhile, layers in panel are not always aligned perfectly. Therefore,

standard CLT panels usually get an initial edge trimming to eliminate adhesive residuals and to align

all four edges. Subsequently, standard panels may receive regional planning or sanding if needed.

Layers of fire-retardant layers (e.g. gypsum board) or acoustic boards may be attached at this time

depending on the end use of the panel.

Over the last decade, the excitement for CLT not only lies in its sustainable, beautiful

and carbon positive building material, but also fast speed of erection due to its high level of pre-

engineering. Shop drawings generated by architect and engineer of record are usually further pro-

cessed by CLT manufacturer for custom cutting and drilling, for examples, windows, connections,

wire and hardware routes, etc (Figure 3.10). Such process on the job site will be difficult, time

consuming and inaccurate. Given the excellent workability of wood, CLT panels are usually ma-

chined and milled as construction ready components using Computerized Numerical Control (CNC)

machine before packing and shipment, see Figure 3.11 . The use of CNC machine in addition to the

factory cutting machine, 2 by lumbers can be pre-fabricated and transformed into multi-story resi-

dential or commercial structures even high-rise buildings as allowed by code officials. The panels’

ability to be CNC-routed into complex geometries, as both structural and interior finish also has

great potential for new complex, architectural experiences to create highly evocative experiences of

space and light. At the job site, these CNCed panels are tilted or craned into position with the hope

that all components will ‘fit like a glove’.

44
Figure 3.10: CNC milling of CLT panels (Hans Hundegger AG)

Figure 3.11: CNC Machining Center (IHB International)

3.1.6 Marking, packaging and shipping handling

Similar to glulam or other engineered wood, marks on the product ensures the quality

and provide the traceability to the material and manufacturer. According to PRG-320, certified

producers should provide the marks containing the following information:

1. CLT grade qualified in accordance with PRG-320

2. The CLT thickness and identification number

3. The mill name or identification number

4. The approved agency name or logo

45
5. The symbol of ‘ASNI/APA PRG 320’ signifying conformance to this standard

6. Any manufacturers designations which shall be separated from the grade-marks or

trademarks of the approved agency by not less than 6 inches (152mm),and

7. ‘Top’ stamp on the top face of custom CLT panels used for roof or floor in manufac-

tured with an unbalanced layup.

Upon final cleaning and dust blowing, CLT panels are packaged in water retardant foil to

protect from harsh weather conditions. Orders are usually transported by truck or by container if

exported overseas.

3.2 Clemson pilot CLT manufacturing

While CLT has gained popularity in Europe, Canada and elsewhere over the last two

decades, it has yet to be become established in the US. During the preparation of this manuscript,

there are only a limited number of CLT buildings constructed in the US and the CLT panels used

in these buildings were imported from Europe and Canada. While there are US manufacturers that

produce non-structural CLT panels (e.g. for crane mats), currently, there is only two manufactur-

ers in the United States that are certified to produce structural grade CLT panels for construction

(SmartLam and D R Johnson).

The structural performance and behaviors of CLT made from European and other non-US

wood species have been extensively studied by others (see Chapter 4 and 5). However, only scant

body of knowledge is available for CLT made from domestic US wood species. With a strong desire

for using CLT as potential structural elements among building owners, architects and engineers,

there is an urgent need to research mechanical properties of CLT made of domestic fast growing

wood species. In the Southern U.S. forests , Southern Yellow Pine would be an ideal candidate

for such application. Its group species (loblolly, longleaf, shortleaf and slash pines) make up the

largest proportion (17% in 2012) of the softwood resource in the Southern Forests (Oswalt et al.

2014). An authoritative test method, along with comparison of experimental test results to the

characteristic values in PRG-320, can be an appropriate method to ensure performance of CLT made

of domestic US species. For structural use CLT, PRG-320 regulates its performance by mandating

the minimum required test values in bending stiffness (EI ), bending moment (fb S ) and interlaminar

46
shear capacity (Vs ). In this particular study, 3-ply Southern Pine CLT bending and rolling shear

test specimens were produced in a laboratory setting. Both major and minor direction capacity were

experimentally investigated. Finally, results were adjusted for shear deformation and compared to

the ‘V3’ designated CLT characteristic values in PRG-320. The rest of this chapter will discuss the

pilot CLT manufacture in Clemson, and preparation of test specimens. Chapter 4 and 5 will report

the test results in rolling shear and bending, respectively.

3.2.1 Adhesive selection

Four different adhesives were used to produce the test specimens: Melamine Formalde-

hyde (MF), Phenol-Resorcinol Formaldehyde (PRF), Emulsion Polymer Isocyanate (EPI), and

Polyurethane (PUR) adhesive. The former two belong to the formaldehyde adhesive family while

the latter two belong to the isocyanate family. Based on the poor fire performance and adhesive

delaminations observed in rolling shear tests (see Chapter 4), EPI was excluded for the bending test

specimens. MF is typically used for manufacturing glulam and has good water resistance. Com-

pared to other formaldehyde-based adhesives, MF also has a much lighter color (i.e. it does not

stain the wood). However, the limitations of MF are the relatively high cost of melamine compared

to phenolic resin and the so called ‘off-gassing’ behavior of formaldehyde. PRF has the advantages

of curable at room temperature, durable, and economical (i.e. Phenol-Resorcinol compounds are

generally more economical than melamine). PRF is also an off-gassing adhesive, as it contains

formaldehyde. However, the off-gas rate of PRF over the life of the product (i.e., beyond the pro-

duction phase) is much lower than other formaldehyde resins such as MF (Rowell 2012). In addition,

the cured PRF may show unpleasing dark stains in the glue lines. EPI is a two-part adhesive that

requires mixing prior to glue application. Due to the emulsifability of isocyanate, adequate mixing

is important. Despite EPI is not an off-gassing adhesive, the main disadvantages are high cost of

EPI compared to formaldehyde-based adhesives and low workability during mixing of the adhesive.

PUR is a moisture-activated adhesive that is commonly used in CLT panels produced in European

and Canada, because it does not contain formaldehyde and is a simply one part adhesive system.

However, the fire resistance of PUR is generally poorer than other adhesives such as MF. As dis-

cussed thereof, all four adhesives have its advantages and disadvantages and it is of interest to see

the adhesive performance in structural use of CLT. The application of adhesives in the Clemson

manufactured CLT (hereafter refer to as Clemson pilot panels) followed the instructions provided

47
by their respective suppliers. The mix ratio, spread rate and minimum pressing time for each of the

four adhesives can be found in Table 3.1.

Table 3.1: Working properties of adhesives used in this study

Mix Ratio Spread Rate Platen Pressure CAT* Required Press Time
Adhesive
Rasin:Hardener lb/1000f t2 psi min HR

MF 100:60 63 150 30 3

PRF 100:40 60 150 30 3

EPI 100:15 72 150 30 3

PUR NA 60 150 40 4

*Close Assembly Time (CAT) refers to the time interval from substrate assembly to the application of full

pressure.

In addition to cost and working properties, each adhesive type possesses other attributes

that might be important. For example, among the adhesives mentioned above, PRF is dark brown,

whereas EPI and PUR are light/transparent color. PUR has the advantage in the glue spray system

since it does not require mixing and is moisture activated. EPI is free of formaldehyde. Adhesive also

plays key roles in CLT fire performance. For fire design of CLT, heat performance understanding of

each adhesive may be a good area for future research.

3.2.2 Lumber preparation

Several bundles of donated No.2 & No.3 2x6 Southern Yellow Pine were received and stored

at the Clemson University’s woodshop. They were conditioned at a constant temperature 70◦ F and a

relative humidity 68% for 1 month, see Figure 3.12. PRG 320 recommended that the lumber having

a moisture content (M.C.) of 12%±3% be targeted to ensure proper bond quality of the product.

Note this standard M.C. specification for lumber may not be suitable for all CLT manufacturing

process. Some adhesives (e.g. moisture activated ones) are more sensitive to M.C. than others. It

is recommended that the adhesive supplier would provide such information regarding the optimal

M.C. at time of adhesive application. Another reason for limiting the M.C. variation is to minimize

the development of internal stress within the same layer and also in the depth of the panel. Internal

stress may result from differential swell & shrinkage, which is dependent on the M.C. profile, growth

48
ring orientation and species. The CLT Handbook (Gagnon et al. 2011) further recommended that

the maximum difference in the M.C. between adjacent piece that are to be joint not exceed 5 %. A

electrical-resistance type moisture meter (Figure 3.13) was used to check lumber M.C. in the pilot

panel manufacturing.

Prior to CLT panel manufacturing, M.C. and density were spot-checked to ensure compliance

with the code’s minimum requirement:12%±3% for M.C. and 0.35 for specific density.

Figure 3.12: No.2 & No.3 2x6 Southern Yellow Pine being conditioned at woodshop

A 8 in. daylight, cold Newman hydraulic press was used to assemble all pilot CLT panels.

The press was designed at maximum pressure of 325 psi, with the platen size 4ft. by 4ft. Due to

the structure of the press, width of the CLT panels was limited to 4ft, whereas the opening of the

press allowed producing of longer panels with staged press. Therefore, No.3 lumbers were were cut

into 4ft. segments to serve as crosswise boards (Figure 3.14 showing a industrial miter saw cutting

a 2x6 lumber).

Once lumbers were cut into desired length, a surface planning was performed on the two

wider surfaces to remove dirt, defects and roughness. This helps activate or refresh the wood surface

to reduce oxidation for improved gluing effectiveness (Julien 2010). In most cases, planing only face

and back (two wider surfaces) may be sufficient if the width tolerance is acceptable and lumber

49
Figure 3.13: DELMHORST moisture meter with hammer electrode

Figure 3.14: Lumbers trimmed into 4ft. segments

50
edges are not glued. However, if those are the cases, planing all four sides are required to ensure

dimension uniformity.

As far as the lamination thickness for each layer, PRG 320 imposes a limitation of minimum

5/8 in. up to 2 in. maximum for both parallel and crosswise directions. For 2 by material, it is

economic to plane off 1/16 in. on both wider surfaces, leaving individual lamination thickness being

1 3/8 in. The Northfield lumber planner installed at woodshop is a singer surfacer, meaning each

run it only plans one surface. The lumbers were then flipped and fed into the machine for the second

run. See Figure 3.15, the machine was planing the top surface of a 12 ft. long 2x6.

Figure 3.15: Lumber planing in progress. Newman press standby on the right

3.2.3 Adhesive application and assembly press

Before the adhesive application, bonding surface of planed lumbers must be cleaned and

free from any adhesive-repellent substances such as oils, greases or release agents, which would have

a detrimental effect on bonding quality.

In a typical industrial glue application system, the extruder head moves and apply parallel

lines/threads of adhesive in an air tight system with direct supply from a adhesive container. The

extruder travels through every inch of the target area before full pressure is generated for assembly

51
process.

There are two main factors that could possibly affect the quality of the glue bond: mix

ratio and close assembly time (CAT). Mix ratio defines the portion of resin and hardener in a mix

(usually by weight). Typically, a higher mix ratio will result in a slower cure rate but allows a

longer assembly time for larger panels. It is a also a function of room temperature and humidity:

conversation with the adhesive agent indicated that for warmer and more humid plant environment,

one should increase the mix ratio and spread rate since the evaporation of water could result in a

accelerated curing. The actual mix ratio and spread rate is usually tabled by the adhesive supplier

and can be easily read with the input of room temperature, relative humidity and close assembly

time (CAT). CAT refers the time interval from substrate assembly (adhesive mix) to the application

of full pressure. It is easily understood that a higher mix ratio allows a prolonged working time that

are needed for panel layup, additional adhesive application. To summarize, the adhesive application

is influenced by many factors and should be properly planned before execution. The adhesive mix

ratio can be determined by the room temperature, humidity and desired close assembly time.

In manufacturing of the pilot panels, 70◦ F lab temperature, relative humidity 68% and a 30

minute CAT (for a 4ft by 4ft 3-ply CLT) was used to determine the mix ratio and spread rate for the

adhesive application. Those data shown in Table 3.1 guided the adhesive preparation throughout

this specimen manufacturing.

For the adhesive preparation in a laboratory setting, weight of each component was cal-

culated based on the mix ratio, spread rate and the desired cover area. The respective adhesive

components were then weighted in party-cups (Figure 3.16) and mixed using a popsicle stick. The

components were adequately mixed until the mixture shows consistent color. For the lab safety

measurement, latex gloves, masks and eye protection were always wore during the adhesive prepa-

ration due to the corrosive and skin-irritative nature of hardeners. Once laminations were laid up

and cleaned, the adhesive mixture was poured onto the surface as evenly as possible (Figure 3.17).

In fact, it was very difficult to achieve a close, parallel spaced and thin glue line during a manual

pour, an additional spread process was executed. A flat blade plastic spatula was used to spread

the adhesive within the target area, ensuring full adhesive coverage (Figure 3.18). Note this process

is not necessary in the mass produce of CLT in a plant environment since the pressure generated by

the press will squeeze and even out the glue lines.

Pressing is a critical step of the CLT manufacture account for proper bond development

52
Figure 3.16: Weighed resins and hardeners ready to be mixed

Figure 3.17: Adhesive application on a 16 f t2 area. Picture show malemin formaldehyde (MF)
adhesive

and CLT quality. The pressure supress the potential warping of layers and overcome their surface

irregularities. Structural cold set adhesives (those mentioned thereof) are commonly selected to

avoid having to heat the bulk panels during pressing.

In general, the pressure and pressing time required are determined based on the combination

of adhesive type and wood species. A low pressure would decrease the adhesive penetration and

result in a poor bond quality whereas high pressure is detrimental to the cell structure of the wood,

53
Figure 3.18: Spread the adhesive with a spatula

particularly for those low density softwood species. Adhesive supplier often times provide suggestions

based on their experience. For this study, a uniform pressure of 150 psi in the hydraulic press was

used for manufacturing of pilot panels. The pressing time required generally varies within several

hours and are highly dependent on the adhesive curing rate. For example, PRF takes the longest

pressing time, followed by MF, PUR, and EPI. If a shortened adhesive curing time is desired, radio

frequency (RF) technology used in the glulam and finger joint curing process can be applied for CLT

for accelerated curing. Meanwhile, adhesive spread rate can be reduced if curing is accelerated. All

pilot panels were pressed for their minimum hours required (Table 3.1 last column) and can be as

long as 6 hours. Figure 3.19 demonstrates a 3-ply Southern Pine CLT being assembled in the press

with the pressure shown. Note that the gauge shown in Figure 3.19 indicates the pressure developed

in five cylinders and to convert this value to platen pressure, a simple total force calculation will be

suffice. For example, 1350 psi in all five cylinders (each with 8 in. diameter) will generate 150 psi

pressure on a 4 ft. by 4 ft. platen:

Pcylinder ×Acylinder
Pplaten = (3.1)
Aplaten

54
Figure 3.19: A 3-ply CLT being pressed with pressure shown

3.2.4 Panel sizing and additional comments for longer panels

After assembly pressing, the panels were taken out of the press and stored in the indoor

environment for an additional 24 hours to allow the adhesive to cure sufficiently before the subsequent

sizing and testing tasks. A 7 1/4 in. diameter circular saw was used to trim the panel into desired

sizes. A blue chalk mark along with a straight edge from a OSB board was used to guide the circular

saw for the cut (Figure 3.20). For a 4 1/8 in. thick 3-ply CLT, it usually requires two runs of the

saw blade (top and bottom) to cut through the thickness. In the light of the manual cutting and

trimming experience obtained here, it is not recommended to use a larger saw blade for thicker

CLT. First off, wood is not a homogeneous material: its density varies from spot to spot, especially

when defects are presented (knots, decay etc.). The saw blade tends to have a strong kickback

when density changes. Second, the cutting process generate a lot of heat on the blade and this

gets severe when the blade is buried deep. Therefore running a shallow blade with multiple runs is

recommended, it is a safer way of cutting and prolongs the life of the blade.

The platen size of the press (4 ft.) did impact the dimension of the test specimens that

the lab is able to produce. For bending property investigations, PRG 320 documents requires the

55
Figure 3.20: Sizing a CLT panel with a circular saw

test samples to bear a minimum span to depth ratio 30, which would translate to be more than

10 ft in clear span for a 3-ply CLT. Adhesive suppliers were called in for potential solutions. They

recommended that different mix ratios in conjunction with multiple presses would achieve a span

longer than 4 ft. The purpose of using different mix ratios is to delay the adhesive curing in later

presses. Take MF as an example, in order to produce a 8 ft. panel, adhesives with two mix ratios

are applied to each of the 4 ft. sections. The first section will be cured in 3 hours under pressure

and the other in 6 hours. Therfore, by reducing the hardener percentage, the slower curing rate

would allow extra presses beyond the first 3 hours. Theoretically it may sound applicable, however,

this practice didn’t generate satisfactory glue bond quality as delaminations were observed in later

presses. The adhesive dried off quickly within a few hours and were not able to bond the layers

together (perhaps the cause of a short open assembly time). Therefore this practice was discarded

in the concern of glue bond quality. The 4 by 12 ft. panels made in the woodshop were pressed three

times with fresh adhesive applied on each 4 ft. section at time of assembly press (see Figure 3.21) to

achieve the desired span-to-depth ratio. Within each press, the adhesive mix ratio and spread rate

were maintained the same as indicated in Table 3.1.

3.3 Specimens preparation for structural evaluations

The pilot CLT panels produced in the woodshop were made according to designation ‘V3’

specified in the PRG-320 standard, meaning the parallel layers are No.2 Southern Pines and No.3

for all the cross-wise layers. For the rolling shear test, a total of ten 4ft. by 4ft. 3-ply CLT panels

56
(a) First press (b) Second press (c) Final press

Figure 3.21: Pilot panel staged press

were made. Among them, for MF and PRF adhesives, two panels were made for each of the strength

directions (major and minor), while for EPI, only two panels for the major strength direction were

made. These ten 4ft. by 4ft. panels were trimmed into forty 2ft. by 1ft. short beams. An additional

twenty PUR specimens with the same dimensions were produced by Structurlam in Canada and

shipped to Clemson for testing. Likewise, the twenty specimens include ten for major strength

direction and ten for minor. The number of short beams with dimension 1ft. by 2ft. prepared for

the rolling shear test is shown in Table 3.2.

Table 3.2: Number of specimens produced for major and minor axes rolling shear test
Adhesive Major Axis Samples Minor Axis Samples

MF 8 8

PRF 8 8

EPI 8 NA

PUR* 10 10

*All PUR specimens were made by Structurlam

The test method, results and discussion of 3-ply rolling shear will be presented in Chapter

4, while Chapter 5 will focus on the experimental study of bending stiffness and bending strength.

MF and PUR adhesives were used to produce the bending test specimens. Four 12ft. by 4ft panels

were made in Clemson using MF adhesive and they were trimmed into sixteen 1ft. wide strips.

The twenty PUR specimens were provided by Structurlam and they arrived test-ready. The total

57
number of bending specimens are shown in Table 3.2.

Table 3.3: Number of specimens produced for major and minor axes bending test
Adhesive Major Axis Samples Minor Axis Samples

MF 8 8

PUR* 10 10

*All PUR specimens were made by Structurlam

58
Chapter 4

Rolling Shear of Southern Pine

CLT

Cross-laminated timber (CLT) is a relatively new engineered wood product. While CLT has

gained acceptance and utilization in Europe, Australia and other countries, its applications in the

United States are limited. Limited information is available on CLT made from US wood species.

This chapter presents the rolling shear test results of CLT panels made using one of the US wood

species, Southern Pine (SP). Four types of adhesives, namely Melamine Formaldehyde (MF), Phenol-

Resorcinol Formaldehyde (PRF), Polyurethane (PUR) and Emulsion Polymer Isocyanate (EPI),

were used to make 3-ply Southern Pine CLT panels. The characteristic rolling shear strengths of the

test panels (5th percentile with 75% confidence tolerance limit) were determined for both the major

and minor axes and compared to the required characteristic values in PRG-320 for ‘V3’ layup. The

results showed that SP CLT panels made with PUR passed the PRG-320 rolling shear requirements

in both major and minor axes.

4.1 Introduction

Currently in North America, there has been a large research effort in exploring different

aspects of CLT, such as evaluation of rolling shear, manufacture parameters, two-way bending

performance and in-plane stiffness of CLT diaphragms. This section focuses on the experimental

59
investigation on the key structural performance: rolling shear capacity. It is required by PRG-320 as

plant pre-qualifications. Pre-qualified panels are certified with stamped grade and their performance

should exceed an established minimum in a statistical sense. The structural performance evaluations

conducted in the this study have a two-fold motivation. First, to evaluate the structural performance

properties of CLT made of domestic wood species, Southern Pine, in this case. Second, to deepen

the understanding of adhesives on both workability and their influences on the structural properties.

4.2 Background and literature review

Wood is an orthotropic material with different mechanical properties in the three mutually

orthogonal axes: Longitudinal (L), Tangential (T), and Radial (R), as illustrates in Figure 4.1.

Rolling shear is defined as the shear stresses due to shear strain in the RT plane, and can be

interpreted as the wood fibers parallel to the longitudinal direction rolling or sliding ‘over’ each

other, thus, wood producing a very low rolling shear strength value.

Figure 4.1: The principal axes of wood with respect to grain direction and growth rings

Currently, there are multiple test methods available to determine the rolling shear proper-

ties, and it has been found that the rolling shear properties are depend mainly on wood species,

growth ring orientation, lay-up, and span-to-depth ratio. A standard for performance-rated CLT

called ANSI/APA PRG-320 Standard for Performance-Rated Cross-Laminated Timber was recently

published (ANSI 2012b). This standard serves as an impetus towards acceptance of CLT made of US

60
wood species, as recently the International Code Council approved the inclusion of PRG-320 qualified

CLT panels as building materials in the 2015 International Building Code (ICC 2015). Kramer et al.

(2013) evaluated the rolling shear behavior of three-ply hybrid poplar CLT with Phenol-Resorcinol

Formaldehyde (PRF) adhesive in accordance to the PRG-320 standard. The adhesive was mixed

using a ratio of 2.5:1 (resin-hardener, by weight) and a spread rate of 60 lbs/1000 f t2 . The hybrid

poplar CLT was pressed at a pressure of 102 psi through tightening of a series of C channels. They

observed that shear through the section was the primary failure mechanism in the short-span shear

tests and reported the 5th percentile value of rolling shear strength of 76.9psi. Fellmoser and Blaß

(2004) evaluated the rolling shear modulus and the effect of span-to-depth ratio for spruce (Picea

abies.) CLT through a dynamic vibration test. Shear deformation depended significantly on the

thickness of the layers vulnerable to rolling shear failure (i.e. layers with lumber oriented perpendic-

ular to the span direction). The influence of shear on the overall bending behavior was found to be

significant for span-to-depth ratios smaller than 30:1 for bending parallel to the grain direction and

20:1 for perpendicular to the grain direction. Mestek et al. (2008) conducted a comparative analysis

using the shear analogy method and a finite element (FE) shell model. The increase of longitudinal

stress due to the effect of rolling shear deformation was a function of the span-to-depth ratio. Zhou

et al. (2014) evaluated the rolling shear properties of three-ply Black Spruce CLT by two-plate shear

test and three-point bending test described in ASTM D2718 (ASTM 2011) and ASTM D198 (ASTM

2015), respectively. Black Spruce CLT was edge glued with fast-curing epoxy and one-component

PRF, and then pressed under a pressure of 87 psi for 24 hours. The two-plate shear test was a

more appropriate test method for assessing the rolling shear modulus and the bending test is an

appropriate test method for determining the shear strength. The rolling shear strength was reported

as 310 psi, obtained via a three-point bending test using a span-to-depth ratio of 6:1. Zhou et al.

(2014) concluded that the shear analogy method can be used to accurately predict the deflection of

CLT specimens under bending when the span-to-depth ratio is relatively small. More recently, Li

et al. (2014) studied the rolling shear properties using torsional shear tests and bending tests for

Spruce-Pine-Fir (SPF) three-ply and five-ply CLTs. Panels were glued with polyurethane (PUR)

adhesive and pressed at two different pressures (15 psi and 58 psi). Higher pressure produced higher

average rolling shear strength, with 293 psi and 319 psi for the abovementioned pressures, respec-

tively. Specimens with thinner cross layers also had higher average rolling shear strength than the

specimens with thicker cross layers. In summary, there is limited information regarding the rolling

61
shear performance of Southern Pine (SP) CLT tested at desired span to depth ratio.

4.3 Adhesives and specimens prepared

Two batches of CLT panels were manufactured using visually graded Southern Pine (SP)

lumbers, namely small pilot-scale panels in a laboratory setting and full-size panels in a plant

environment. The CLT panels for rolling shear evaluation were all 3-ply ones and were manufactured

using visually graded SP lumbers in accordance to the ‘V3’ layup in PRG-320: No. 2 SP lumber in

outer layers parallel to the panel length and No. 3 SP lumber in center layer perpendicular to the

panel length. Both the Clemson and Structurlam CLT panels were made from 2x6 SP lumbers. The

first batch of small pilot-scale 4ft. by 4ft. CLT panels were manufactured at Clemson University.

The second batch of full-scale CLT panels (10’x40’) were produced by Structurlam in Penticton,

Canada and shipped to Clemson University for testing.

The 4 ft. x 4 ft. panels were made using MF, PRF and EPI while the full-size 10 ft.x 40 ft.

panels (from Structuralam) were made using PUR adhesive. Two 4 ft.x 4 ft. panels were made each

using MF and PRF. Each of the 4 ft.x 4 ft. panels were cut into eight short-span rolling shear test

specimens for major and minor axis rolling shear tests. Only one 4 ft.x 4 ft. panel was made using

EPI for major axis rolling shear strength evaluation. EPI was not evaluated for minor axis shear

strength. This is because the major axis shear tests were first conducted and EPI was found to have

the lowest major axis shear strength (See Results and Discussion). Two 10 ft.x 40 ft. PUR panels

were manufactured by Structurlam. Ten major axis and minor axis rolling shear test specimens each

were produced from one of the two 10 ft.x 40 ft. panels. Except for PUR, all other adhesives are

two component adhesives, which require the mixing of resin and hardener. PUR is a one-component

adhesive. The resin-to-hardener mix ratio, spread rate and minimum pressing time for each of the

three adhesives used in the production of the pilot-scale panels can be found in Table 3.1 in Chapter

3.

All panels were pressed with a platen pressure of 150 psi for approximately 6 hours, which

exceeded the minimum 3 hours pressing time by the manufactures. In trimming the panel into

desired size of short beams, each of the 4 ft. x 4 ft. panels was cut into 8 pieces of short-span shear

test specimens (12 in. wide, 23 5/8 in. long and 4 1/8 in. thick specimens). 60 specimens were

produced for rolling shear tests, in which 40 specimens were made at Clemson using MF, PRF and

62
EPI and 20 specimens were made using PUR in an industrial setting by Structurlam (see Table 3.2

in Chapter 3). EPI was not tested for minor axis. This is because major axis shear tests were

first conducted and it was found that EPI had the lowest major axis shear strength; therefore, was

eliminated from further minor axis shear strength evaluation. Prior to the test, moisture content

(M.C.) and density were measured for each specimen. The average M.C. were all greater than 8%

(Table 4.1), which satisfied the PRG-320 minimum MC requirement for the evaluation of mechanical

properties.

Table 4.1: Density and moisture content (MC) of major and minor axis specimens

Major Axis Minor Axis

Adhesive Type Quantity Density MC Quantity Density MC

lbs/f t3 %(COV) lbs/f t3 %(COV)

MF 8 34.5(2.1%) 9.53%(7.3%) 8 34.7(3.3%) 9.74%(3.6%)

PRF 8 35.9(2.2%) 9.52%(4.9%) 8 33.7(2.1%) 9.59%(5.3%)

EPI 8 34.7(1.9%) 10.5%(3.7%) – – –

PUR 10 30.7(2.6%) 8.66%(4.7%) 10 30.2(2.5%) 8.79%(7.2%)

4.4 Test method

The rolling shear capacity of SP CLT panels were evaluated using the three-point short-span

bending test approach described in ASTM D4761 (ASTM 2013). Schematic views of the three-point

bending test setups for major and minor axes rolling shear tests are shown in Figure 4.2. All

specimens were 12 in. wide and 4 1/8 in. deep, with a clear span of 20 5/8 in. and a span-to-

depth ratio of 5, as recommended in ASTM 3737 (ASTM 2012a), to ensure shear failure mode.

The bearing length was 1.5 in. at each of the supports. All specimens were loaded using a 150-kip

capacity actuator with a loading rate of 0.05 in/min to ensure a minimum test-to-failure time of 4

minutes. The deflection was measured using a 12-in stroke draw-wire displacement sensor attached

to the bottom face of the specimen at mid-span. In each test, the complete load displacement curve,

peak load and failure mode were recorded. Figure 4.3 shows the typical test setup for one of the

major axis specimens.

63
Figure 4.2: Three-point short-span bending test setup for rolling shear (left: major axis ; right:
minor axis).

Figure 4.3: Typical experimental setup for major axis rolling shear test

The rolling shear strength (τrolling ) for each test was computed using Eq.(4.1) and (4.2).

Pmax /2
τrolling = (4.1)
(Ib/Q)ef f

(EI)ef f
(Ib/Q)ef f = (4.2)
n/2
P
Ei hi zi
i=1

where Pm ax is the peak force observed during the initiation of rolling shear failure; EIef f is

the effective bending stiffness; Ei is the modulus of elasticity (MOE) of layer i; hi is the thickness of

layer i, except the middle layer, which is taken as half of the middle layer thickness (Figure 4.4); zi

is the distance from the centroid of the layer to the neutral axis, except for the middle layer, where

it is measured from neural axis to the centroid of the top half of the middle layer.

The MOE parallel to the major strength direction (E0 ) for calculating effective bending

stiffness EIef f was obtained from the National Design Specifications for Wood Construction (AWC

2012). The MOE values perpendicular to major strength direction(E9 0) was taken as 1/30 of E0 as

suggested by the PRG-320.

64
Figure 4.4: Schematic view of 3-layer ‘V3’ cross section for calculating rolling shear

4.5 Result and discussion

Table 4.2 summarizes the major and minor axes rolling shear strength computed using

Eq.(4.1) and (4.2). It should be noted that the initiating failure of five out of eight of the major axis

EPI specimens was glue bond failure. These five specimens were excluded from the rolling shear

strength calculations. The major axis rolling shear strengths exhibited less variability than that of

the minor axis specimens. This is because the major axis test specimens failed mainly in a ‘pure’

rolling shear mode while the minor axis specimens showed a combined rolling shear and bending

failures. More details on the failure mechanisms of the minor and major axis tests are discussed in

the next section.

Failure mechanism of major axis specimens Figure 4.5 shows the load-displacement

curves of the major axis rolling shear tests categorized by adhesive type. Except for five of the

EPI specimens that exhibited glue bond failures, the first major drop in load in each of the curve

shown in Figure 4.5 marks the initiation of rolling shear failure. The typical rolling shear mechanism

observed during major axis tests followed a sequence of three failure stages:

Table 4.2: Summary of rolling shear strengths for major and minor Axes

65
Major Axis Minor Axis

Adhesive Type Number of Mean RS Strength Number of Mean RS Strength

Test psi(COV) Test psi(COV)

MF 8 289(9.9%) 8 254(18.3%)

PRF 8 353(5.4%) 8 279(22.0%)

EPI 8 246(12.2%) – –

PUR 10 259(7.8%) 10 286(16.4%)

1. initiation of diagonal rolling shear cracks in the middle layer near the supports

(Figure 4.6),

2. extension of diagonal cracks to the top and bottom layers and followed by delam-

ination failure at one or two of the horizontal glue lines (Figure 4.7), and

3. bending failure occurred due to compression (top) and/or tension failures of the

longitudinal fibers in the outer layers (Figure 4.8).

From Figure 4.5, it can be seen that the CLT panel retained portion of the load carrying

capacity beyond the first major failure event or drop in load. This phenomenal was attributed to

redistribution of shear force occurred after the initiation of the first major diagonal rolling shear

crack. This transverse shear redistribution behavior was also observed in the shear tests by Zhou

et al. (2014). The limiting failure mode for all MF, PRF and PUR specimens was rolling shear.

Pre-mature delamination along the glue lines were observed for five of the eight EPI specimens with

no sign of rolling shear failure (Figure 4.8). These delamination failures are reflected in the first

peaks of the EPI load-displacement curves shown in Figure 4.5. As stated, these five specimens

with glue bond failures were excluded from the rolling shear statistics shown in Table 4.2. The glue

bond strength for each test specimen was estimated by using the peak forces shown in Figure 4.5

to compute the shear stress in the glue layers. Since glue bond failure was not the initiating failure

mode for MF, PRF and EPI specimens, one can expect the glue bond strengths for these adhesives

to be greater than the glue layer shear stresses computed using the peak loads. For those EPI

specimens with glue bond failures, the peak loads can be used to estimate the glue bond strengths.

The glue bond strength estimated in Table 4.3 shows the MF, PRF and PUR glue bonds were at

least 64%, 100% and 48% stronger than that of the EPI specimens, respectively.

66
(a) MF (b) PRF

(c) EPI (d) PUR

Figure 4.5: Load displacement curves of major axis rolling shear specimens

67
Figure 4.6: Typical rolling shear cracks

Table 4.3: Glue bond strength estimation for major axis specimens
Adhesive Bond Strength Mean (psi) STDEV

MF ≥2881 26.6

PRF ≥351 17.6

PUR ≥259 20.2

PUR 1752 19.1

1
Glue bond strength for MF, PRF and PUR are estimated as the shear stress at glueline when rolling shear

failure occured.
2
Glue bond strength for EPI is calculated as the shear stress at glueline when delamination occured.

Failure mechanism of minor axis specimens The load-displacement curves of minor

axis rolling shear tests are shown in Figure 4.9. The failure mechanism of minor axis rolling shear

specimens was different compared to the major axis. When load applied to a minor axis test

specimen was gradually increased, the minor axis shear test panel exhibited the following typical

failure sequence (Figure 4.10):

1. opening of vertical gap closest to the mid-span in the tension side (bottom layer)

of the panel. Note that the panels were not edge glued,

2. rolling shear failure and crack initiation in the top layer in compression, and

68
Figure 4.7: Extension of diagonal shear crack into glue lines

Figure 4.8: Glue bond delamination (observed in EPI specimens) and bending failure

3. bending failure in the middle layer longitudinal boards. This ultimate bending

failure typically accompanied with a very loud noise associated with rupture of wood fiber in

tension.

4.6 Characteristic rolling shear strength

In order to qualified for structural CLT application, the 5th percentile rolling shear strength

from tests with 75% confidence tolerance limit for SP CLT must meet the published minimum

values, known as characteristic test values, in Table 1 of PRG-320. To evaluate the characteristic

69
(a) MF (b) PRF

(c) PUR

Figure 4.9: Load displacement curves of minor axis rolling shear tests

70
Figure 4.10: Failure of minor axis specimen

test values, the major and minor rolling shear strengths for each of the adhesive type were fitted to

Normal, Lognormal, Gumbel, Frechet, Weibull and Unbounded Johnson distributions. For fitting

the major axis rolling shear strength, it was determined that the Unbound Johnson Distribution

(SU) is the most suitable distribution for capturing the extreme values at low percentile region. The

fitted Johnson SU distributions for major axis specimens are shown in Figure 4.11. Note that for

the EPI major axis specimens, only the three tests with rolling shear failures are fitted to Johnson

SU distribution (five other tests failed due to delamination of glue lines). For minor axis specimens,

normal distribution was determined to be the best-fit distribution (Figure 4.12).

The required characteristic test values listed in Table 1 of PRG-320 is the minimum required

one-sided tolerance limit (T L) associated with 5th percentile and 75% confidence level. To be more

specific, it is a one-sided lower TL as defined in ASTM D2915 (ASTM 2010).Eq.(4.3) was used to

determine the characteristic test values (YL ). For all test test data fitted, a tolerance factor k is

determined such that the T L intervals cover at least a proportion p of the population with confidence

γ:

YL = Ȳ − kS (4.3)

where Ȳ is the sample mean and S is the sample standard deviation. In this study, p and γ are set

to 5% and 75%, respectively, by the definition of tolerance limit. Dixon and Massey (1969) provided

71
Figure 4.11: CDF of major axis rolling shear strength

Figure 4.12: CDF of minor axis rolling shear strength

methods to estimate the tolerance factor k: in this study, k is estimated from the inverse cumulative

72
distribution function of the non-central t distribution:

tr,N −1,δ
k= √ (4.4)
N


δ = Zp N (4.5)

where δ is the non-centrality parameter used in the non-central t distribution, N is the sample

size, Zp is the Z value for a proportion of p (i.e. 5%) in the standard normal space. For major

axis specimens fitted in the Unbonded Johnson distribution, Z values are obtained by applying the

transformation function. In fact, the whole Johnson family distributions are based on standard

normal, each with a distinct transfer function to describe different data characteristics (Johnson

1949).

The final computed characteristic rolling shear strengths for are summarized in Table 4.4.

According to Table 1 of PRG-320, the minimum characteristic value for rolling shear of Southern

Pine CLT with ‘V3’ layup configuration is 180 psi for both major and minor axes. The PUR

specimens met and exceeded the 180 psi requirement for both the major and minor axes, while only

the major axis rolling shear strengths of the MF and PRF specimens passed the PRG-320 criteria.

Table 4.4: Characteristic rolling shear strengths

Major Axis Minor Axis

Adhesive Type δ Z Characteristic δ t k Characteristic

Strength (psi) Strength (psi)

MF 2.19 -2.19 201.1 4.65 6.19 2.19 146.4

PRF 2.19 -2.19 243.8 4.65 6.19 2.19 144.2

EPI 2.10 -2.10 190.4 – – – –

PUR 1.91 -1.91 198.7 5.20 6.65 2.10 182.7

4.7 Conclusion

Based on the above results and discussion, the following conclusions can be made:

1. All four adhesives (MF, PRF, EPI and PUR) can be used to manufacture Southern

Pine CLT panels that meet the minimum characteristic rolling shears trengths (lower 5th

73
percentile value with a 75% confidence interval) in the major axis direction for the ‘V3’ layout

in PRG-320;

2. Among the four adhesives, the EPI exhibited the lowest major axis rolling shear

strength. Several CLT specimens made with EPI adhesive showed premature delamination at

the glue lines between layers. This failure mechanism significantly decreased the rolling shear

strength. To use EPI for manufacturing SP CLT, further investigation such as varying the

EPI spread rate or pressing cycle should be carried;

3. In the minor axis direction, only the panels made with PUR adhesive passed

the PRG-320 requirement. While the MF and PRF specimens did not pass the PRG-320

minor axis rolling shear requirement, it is anticipated that modifying other parameters in the

manufacturing process of SP CLT such as varying the adhesive spread rate, lumber grade,

annual ring orientation, platen pressure, and/or layer thickness may resolve the minor axis

rolling shear requirement for these three adhesives.

74
Chapter 5

Bending Strength and Stiffness of

Southern Pine CLT

CLT panels are composite engineered wood products with cross-wised layups that provide

dimensional stability. It has been a promising building material to potentially replace concrete and

structural steel due to its economy, sustainability, carbon sequestration, speed and ease of erection,

and acceptable seismic performance. CLTs high strength-to-weight ratio makes it applicable to

timber building as slabs for horizontal structural members. Due to the limited knowledge of CLT

panels made of domestic US wood species, the bending stiffness and bending strength were tested

for the 3-ply Southern Pine CLT in both major and minor strength directions. This chapter presents

the test setup, results and discussion of the ‘V3’ Southern Pine CLT bending test. In addition to

reporting the test results, an interaction equation defining the failure criterion was used to estimate

the bending strength based on a Monte Carlo simulation. The strength parameters needed for

the Monte Carlo simulation was obtained from a mini in-house testing program on 2x6 lumbers.

The simulation shown a good match to the experimental data, indicating an appropriately defined

bending failure criterion.

75
5.1 Introduction

Cross-laminated timber (CLT) is an engineered wood product that is manufactured by face

gluing the cross-wise layers of dimension lumbers by means of either vacuum or mechanical press

to form large-sized solid composite panels. The finished panels are typically 2 to 10ft. wide, up to

60ft. long and up to 20 in. thick depending on the pressing capacity of the manufacturer. Such

systems were first developed in Switzerland in the 1970s in the furniture industry. In the mid-1990s,

an industry-academia joint research effort in Austria led to the commercialization of structural-

use CLT panels (Mohammad et al. 2012). These solid panels have versatile applications as to be

used as floor, wall, roof components in residential/commercial building constructions and have the

potential to replacement mineral building materials. In addition to the building sector, with research

development in timber/concrete composite material, CLT has become a promising alternative to be

used as hybrid bridge decks, as the high strength to mass ratio makes CLT an ideal candidate for long

span bending element. Economy, sustainability, carbon sequestration, speed and ease of erection,

acceptable seismic performance, and lower embodied energy are among advantages of CLT material

(Robertson 2011).

Although CLT has been a consistent presence in European markets for over 20 years, the

implementation of CLT product and systems have just began in United States and Canada (Brand-

ner et al. 2013). Only a handful of CLT structures have been erected over the past decade, with CLT

panels sourced from outside the USA and Canada. During the preparation of this manuscript, there

were only two APA/ANSI certified CLT manufacturer in the US, producing structural CLT panels

up to size of 10ft. by 24ft. with thickness up to 7 layers. The scant body of knowledge regarding

CLT performance manufactured from domestic US wood species has slowed the application of CLT

in buildings. It is therefore of paramount importance to gain access to certified material, modern-

ized building code and advanced technical capacity. This paper presents the experimental bending

stiffness (MOE) and bending strength (MOR) results of CLT panel manufactured of Southern Pine

(SP). Two batches of CLT were examined in the experiments: SP CLT manufactured in a laboratory

setting with Melamine Formaldehyde (MF) adhesive and commercially manufactured SP CLT with

polyurethane (PUR) adhesive. Both strong and weak axis bending properties were examined.

76
5.2 Background and literature review

In 2011, APA-the Engineered Wood Association in collaboration with American National

Standard Institute (ANSI) published a harmonized North American CLT product standard used by

both US and Canada, Standard for Performance Rated CLT, also known as ANSI/APA PRG320.

For various structural performance properties (bending, shear, bearing, etc.), required characteristic

test values along with allowable design values are tabulated for CLT manufactured of different wood

species, lumber grades and lay-up configurations. Corresponding test methods are also suggested in

this version of the standard as a part of the plant pre-qualification requirements. ANSI/APA then

released a newer version in 2012 (ANSI 2012b), addressing custom CLT grades for hardwood species.

This standard serves as an impetus towards the acceptance of CLT made of US wood species. As

recently the International Code Council approved the inclusion of PRG-320 qualified CLT panels as

building materials in the 2015 International Building Code (ICC 2015). The code allows CLT to be

used in Type IV and VB construction, with possibility of Type VA, IIIA and IIIB, depending on

the type of structural elements (wall, floor, roof etc.) and how the exterior and interior elements

are fire protected. Compared to the older versions of IBC code (2009 and before) where CLT are

normally permitted under ‘Alternate Materials and Methods’, the current 2015 IBC code creates

more potential opportunities for this engineered wood product.

Currently in North America, there has been a large research effort in exploring different

aspects of CLT, such as evaluation of rolling shear, manufacture parameters, two-way bending

performance and in-plane stiffness of CLT diaphragms. However, experimental research in full

size CLT bending is limited, mostly due to the cost and difficulties in manufacturing long beams

that satisfying PRG-320 bending test method in a laboratory setting. Mohamadzadeh and Hindman

(2015) evaluated the three layered yellow-poplar (Liriodendron Tulipifera) hardwood CLT in terms of

bending, shear by compression and delamination properties. The goal of their study was to measure

the abovementioned properties and to identify the viability of yellow-poplar in CLT manufacture

as an alternative to standard softwood species CLT. Yellow-poplar satisfies the specific gravity

requirement but, currently no tabulated values can be found in PRG-320. Twelve bending tests

include five-point and four-point bending were conducted on 72 in. long, 11.8 in wide, 4.2 in.

thick specimens, resulting a span-to-depth ratio 16. All No. 2 yellow poplar lumbers were edge

glued and cold pressed with Phenol Formaldehyde (PF) adhesive to form the CLT test specimens.

77
Non-destructive five-point bending tests were performed to evaluate bending stiffness (EI) and

interlaminar shear capacity (Vs ). Destructive four-point bending tests in accordance with ASTM

D198 (ASTM 2015) were performed to obtain the ultimate bending strength (Fb S). Reported

bending stiffness (EI: 5940 lb · ft/ft, 4470lb · ft/ft) and bending strength (Fb S:1.53x108 lb · in2 /ft,

1.37x108 lb · in2 /ft) values for low and high quality specimens satisfied the requirements for V1

designated (Douglas fir-Larch) CLT per PRG-320.

Kramer et al. (2013) tested the bending properties for three layered hybrid poplar (Populus

Deltoids) hardwood CLT specimens. Layers were surface glued using phenol-resorcinol formaldehyde

(PRF) adhesive and pressed using their in-house clamping fixture. Specimen dimensions were 15.7

in. wide, 3.9 in. thick and on-center span 100 in., resulting a span-to-depth ratio 26. All specimens

were loaded to failure using a third-point bending as suggested by PRG-320 and ASTM D198 (ASTM

2015). The obtained 5th percentile values for MOR and MOE were 2640psi and 1020ksi, respectively,

indicating a potential for hybrid poplar to meet the strength requirement for E3 designated CLT.

However, the tested MOE percentile value could not satisfy the requirement. The author attributed

the lower MOE value to the relatively low density of hybrid poplar and indicated that the design

using low density CLT would governed by deflection. Vibration characteristics of low density CLT

could also be a concern.

A study by Okabe et al. (2014) investigated experimental as well as numerical Monte Carlo

(M.C.) simulations of Sugi (Japanese cedar) CLT bending properties. MC method takes inputs in

the form of MOE, MOR, and tensile strength of sawn lumbers. Their specimens (1m in width, 3m in

length) with various thicknesses (90mm, 120mm and 150mm) and layup configurations had a signifi-

cant lower span-to-depth ratio than PRG320 recommended strip-shaped ones. Water-based polymer

isocyanate (EPI) adhesive was used for lumber lamination and finger jointing. The specimens were

tested using third-point bending according to Japan Industrial Standard JIS A 1414-2 (JIS 2010)

which is similar to the bending test method outlined in ASTM D198. Though bending test results

were not specifically tabulated in the table, comparison study showed that results obtained from

MC method was concordant with the experimental bending properties. Authors suggested that the

cross layers were able to transmit bending and shear stresses to increase moment carry capacity.

Their bending strength prediction equation takes a form of an interaction equation with sawn lum-

ber bending and tension strength. It provided a reliable estimate alternative to what is given in

the CLT Handbook (Gagnon et al. 2011) which is purely based on geometry and lumber bending

78
strength.

Hindman and Bouldin (2014) tested the bending and shear properties of five ply CLT

produced with No. 2 Southern Pine lumbers. Due to the limitation of test facilities, the bending

specimens were lap-jointed with vertical and inclined screws in several locations to reach span-to-

depth ratio of 30. The MOE and MOR test results indicated satisfaction for 5-ply V3 designated

CLT and low variability was observed for all specimens with lap joints and screws.

With a strong desire for using CLT as potential structural elements among building owners,

architects and engineers, there is an urgent need to research mechanical properties of CLT made

of domestic fast growing wood species. In the Southern U.S. forests (‘Southern Forests’), Southern

Pine would be an ideal candidate for such application. Its group species (loblolly, longleaf, shortleaf

and slash pines) make up the largest proportion (17% in 2012) of the softwood resource in the

Southern Forests (Oswalt et al. 2014). An authoritative test method, along with comparison of

experimental test results to the characteristic values in PRG-320 (Table 5.1), can be an appropriate

method to ensure performance of CLT made of domestic US species. In this particular study, 3-ply

Southern Pine CLT bending properties (MOE and MOR) were investigated in both major and minor

strength directions, results were adjusted for shear deformation and compared to the V3 designated

CLT characteristic values in PRG-320. In addition, a bending failure criterion along with set of MC

simulations were performed to verify the bending strength of 3-ply CLT. With successful validation of

MC simulation, bending strength can be estimated for CLT with given thickness, layup configuration

and wood species material properties. Therefore, it relieves researchers from conducting expensive

yet labor intensive, repetitive experiments.

Table 5.1: Minimum required characteristic test value per PRG-320: CLT designation V31
Strength Direction fb S (psi) E (106 psi)

Major 2045 1.6

Minor 1205 1.4

1
V3: No.2 Southern pine lumber in all parallel layers and No.3 Southern Pine in all perpendicular layers

5.3 Test method

Details of the test method used to evaluate bending stiffness (EI) and bending strength

(Fb S) of 3-ply SP CLT as well as essential data analysis procedure are presented in this section.

79
Figure 5.1: Schematic third-point bending test setup (top:major axis; bottom:minor axis)

The bending properties of SP CLT panels were evaluated using the flatwise third-point loading test

approach described in ASTM D4761 (ASTM 2013) as recommended by PRG-320. Both standards

require a minimum span-to-depth ratio 30 to reduce shear deformation. The specimen on-center

span was designed to be 123.75 in. between supports with a span-to-depth ratio of exactly 30 (the

3-ply CLT has a thickness of 4.125 in.). Schematic views and photos of third-point bending test

setups for major and minor axes are shown in Figure 5.1 and 5.2.

With the setup shown, the geometry of all test strips complied with PRG-320 dimension

requirements. Prior to test, moisture content (M.C.) and density were measured for each specimen.

The average M.C. was greater than 8% (Table 5.2), which satisfied the PRG-320 minimum M.C.

requirement at the time of mechanical property evaluation.

The specimens were carefully transported to the test frame before symmetry of the system

was measured and verified with respect to support and loading points. An MTS hydraulic testing

machine with an integrated load cell of 150,000 lbs capacity was used to apply load on the specimen.

Testing speed was set constant with a rate of 0.1 in/min to ensure the time to reach specimen

80
(a) (b)

(c) (d)

Figure 5.2: Typical experimental setup for CLT bending test

81
failure to be within 10s to 10min (most of the specimens failed at around 7 minute). The deflection

was measured using a 12-in stroke draw-wire displacement sensor (sensitivity 0.5%) attached to the

bottom face of the specimen at mid-span. In each test, the complete load displacement history, peak

load, and failure mode were recorded. The relationship between load obtained from MTS load cell

and deflection recorded by displacement sensor was used for bending stiffness (EI) calculations.

Table 5.2: Density and MC of tested bending specimens


Manufacturer Number of Specimens Adhesive Density, lb/ft3 (COV) M.C.,% (COV)

Clemson 16 MF 36.44(5.6%) 8.66 (16.4%)

Structurlam 20 PUR 33.13(4.8%) 9.31(4.7%)

The equation for apparent modulus of elasticity (Ef ) of third point loading can be found in

many design manuals, including that of Steel Construction Manual (McCormac 2008):

23P L3
Ef = (5.1)
1296I∆

To adjust for shear deformation, a shear stiffness term was introduced to Eq. (5.1) : Eq.

(5.2) can be found in the appendix of ASTM D198 (ASTM 2015):

23P L3
E= PL
(5.2)
1296I∆(1 − 5GA∆ )

where P and ∆ are the load displacement history when the beam behaved in an elastic manner; L is

the clear span between supports; I is the moment of inertia of the section; GA is the shear stiffness

of the section.

A variety of methods have been adopted for the determination of basic mechanical prop-

erties of CLT in the Europe. ‘Mechanically Jointed Beams Theory’ was available to Eurocode 5

(Eurocode 2004). According to this theory, an efficient coefficient (γi ) was used to account for the

shear deformation of the perpendicular layer, with γ=1 representing rigidly glued member and γ=0

representing no connection at all. However, without experimental investigation, it is difficult to

physically quantify the rigidity of the adhesive between longitudinal and crosswise layers. Blass

and Fellmoser (2004) developed a ‘Composite Theory’ where a list of composition factors ki were

pre-calculated for various loading conditions. These composition factors provide the ratio between

strength and stiffness. However, ‘Composite Theory’ does not account for shear deformations. In

82
the mid 90s, a ‘Shear Analogy’ method was proposed by Kreuzinger (1995) and applied to solid

panels with cross layers. This method accounts for shear deformation of the cross layers and seems

to be the most accurate for CLT panels and therefore was adopted by CLT Handbook (Gagnon et al.

2011) and the CLT standard PRG-320. The shear stiffness based on ‘Shear Analogy’ method can

be obtained using Eq. (5.3):

a2
GA = Pn−1 (5.3)
[( 2Gh11b1 ) +( h1
i=1 2Gi bi ) + ( 2Ghnnbn )]

where a is the distance between the centroid of first layer to the centroid of the last layer; hi , bi and

Gi are the thickness, width and modulus of rigidity for layer i, respectively.

In the above equation, PRG-320 nomenclature provides the values for G as a ratio of E0

(MOE parallel to grain). It recommends G0 = E0 /16 to be used for modulus of rigidity for the lon-

gitudinal laminations, while for cross laminations, G90 = G0 /10. All PRG-320 suggested lamination

properties are tabulated in Table 5.3 as ratios of E0 .

Table 5.3: PRG-320 suggested ratios of E and G based on E0


Property E0 E90 G0 G90

Ratio based on E0 1 1/30 1/16 1/160

‘Shear Analogy’ method also recommended the way to calculate moment of inertia ICLT for

multi-layered cross sections. Parallel axis theorem is applied in this method with an adjustment for

effective width to account for the contribution of cross layers:

Ei
bi = b (5.4)
E0

Therefore based on Table 5, the width of cross layer is adjusted to 1/30 of that of longitudinal

layers. CLT moment of inertia ICLT with contribution of cross layers can be computed using Eq.(5.4):

X bi h 3
ICLT = ( i + Ai zi2 ) (5.5)
12

where Ai is the area of layer i and zi is the distance between the centroid of layer i to the neutral

axis. It is noted that for minor strength specimens, parallel axis theorem does not apply to the

tension layer since CLT is face glued only. Under bending load, the seams in bottom layer will open

up without generation of any composite action (see Test Result and Discussion section).

83
Effective section modulus SCLT is found by dividing the moment of inertia by distance from

centroid to top or bottom edge of the equivalent section:

ICLT
SCLT = (5.6)
y

Modulus of rupture (MOR) fb then can be determined from Eq. (5.7)

Mmax
fb = (5.7)
SCLT

where Mmax is the maximum moment produced based on peak force record in each third-point test:

1
Mmax = Pmax L (5.8)
6

MOE of Southern Pine No. 2 sawn lumber parallel to the major strength direction (E0 )

for composite section property calculation was obtained from the National Design Specifications

for Wood Construction (NDS) (AWC 2012) March 2013 addendum as a mean value. Using Eq.

(5.3) through (5.6), section properties of major and minor strength direction CLT can be found in

Table 5.4.

Table 5.4: Section property of 3-ply SP CLT


Strength Direction I(in4 /f t) GA(kip/f t) S(in3 /f t)

Major 67.78 4331.25 32.86

Minor 2.77 4331.25 4.03

5.4 Test result and discussion

As previously mentioned, in PRG-320 a span-to-depth ratio of approximately 30 is recom-

mended for evaluation for CLT bending specimens to reduce the effect of shear deformations. A

total 36 of bending specimens in both major and minor strength directions with a clear span to

depth ratio at exactly 30 were tested to ultimate failure. For each test, complete load-displacement

history, peak load, and failure mode were recorded for further analysis. Typical load displacement

curves of major and minor tests are shown in Figure 5.3.

It was clear that all bending specimens failed in a brittle manner without any ‘yielding’

exhibited before ultimate failure. Load-displacement segments up to 1000 lbf in major test specimens

84
(a) Pilot Major (b) Pilot Minor

(c) Structurlam Major (d) Structurlam Minor

Figure 5.3: Bending test load-displacement history for major (left) and minor (right) strength di-
rection specimens

and 500lbf in minor ones were selected to compute MOE. These thresholds were determined to be

far less than the failure load and therefore were chosen as beams remained elastic without any

permanent deformation. Linear regression was performed within the prescribed segment for each

test, obtaining the ratios of P/∆. Together with Eq. (5.2), MOE adjusted for shear deformation

was computed for all 36 specimens. For computing modulus of rupture (MOR), all peak loads were

identified and then the application of Eq. (5.6) through Eq. (5.8) led directly to MOR. The statistics

(mean and efficient of variation) for tested MOE and MOR were tabulated in Table 5.5. It is noted
PL
that on average, the shear deformation adjustment term (1 − 5δGA ) is around 84%.

Table 5.5: Mean values of MOE and MOR

85
Direction Provider MOE,106 psi (COV) MOE,ksi (COV)

Clemson 2.38(10.54) 7.40(12.59)


Major
Structurlam 1.20(8.14) 3.95(13.12)

Clemson 2.62(5.88) 9.55(20.13)


Minor
Structurlam 1.86(11.99) 7.00(17.19)

5.4.1 Failure mechanism of major strength direction specimens

The percentage differences of MOE and MOR for Clemson Pilot and Structurlam panels

were considerably high as seen from Table 5.5. MOE (the bending E) is mostly used for checking

serviceability requirements (e.g. deflection, vibration) and is mainly depend on the density of the

wood, quality of the wood (knot, split, checking etc.) as well as level of composite action of CLT.

During major strength tests, specimens behaved linear elastically before brittle failures took place,

and no signs of premature adhesive failure (delamination) were observed. Thus the above obser-

vations eliminated the possibility of glue failure that may cause decreased MOE (i.e. specimens

maintained high level of composite action before failure). Careful examination of lumber quality

revealed that Southern Pine lumbers used in Clemson Pilot panels were of better quality than that

in the Structurlam panels, though both of them were stamped as No. 2 SP lumbers. Pilot panel

lumbers contained overall fewer defects, and were 10% denser than Structurlam lumbers (see Ta-

ble 5.2 for density). Therefore, it is safe to conclude that a relatively low MOE mean value obtained

from Structurlam panels was a result of less density and lower quality lumbers being used.

Differences in MOR values can be explained in the perspective of failure mechanism. Pilot

panels displayed a clear bending failure where lower extreme fibers in constant moment region

experienced bending/tension rupture. The typical bending failure mechanism observed during Pilot

panel major axis tests followed a sequence of three stages (Figure 5.4):

1. Initiation of diagonal rolling shear cracks in the middle layer in the constant

moment region. Glue line remained intact;

2. bending failure occurred due to tension failures of the longitudinal fibers in the

bottom tension layer

3. followed by rupture in the tension layer, typical rolling shear failure was extended

86
(a)

(b)

(c)

Figure 5.4: Typical failure of major strength direction Pilot CLT

87
Figure 5.5: Wood fiber tension failure at lumber knot

in the middle crosswise layer. A portion of the lower glue line was compromised first and then

the crack continued to propagate upward in a zigzag pattern to upper glue line or sometimes

to the compression layer.

It is worth noting that rupture failure in the tension layer would often take place at lumber

defects if any, especially at knots due to stress concentration and material irregularity (Figure 5.5).

It is suggested that for critical bending elements, lumbers in outmost tension layer to be better

engineered and carefully inspected for defects before putting into manufacture and service. Using

higher grade material such as Machine Stress Rated (MSR) lumbers for tension layers can be a less

economic; however, the composite action of CLT can be fully exploited.

Specimens with finger joints displayed a similar fashion of failure propagation as the Pilot

panels; however, ultimate failure was due to the presence of finger joint (FJ). 9 out of 10 Struc-

turlam major strength specimens failed at FJs within constant moment region (Figure 5.6) with one

exception, which failed at lumber knot. Appendix A contains ten pictures taken after the major

strength tests. The finger joint itself constitutes a self-centering profile representing a folded scarf

joint. It enables a simple, fast and form-fit connection between lumber elements by maximizing

the bond surface and minimizing longitudinal losses of board material (Brandner et al. 2013). For

Structurlam panels, these edgewise finger joints (Figure 5.7) were machined, glued and infrared (IR)

accelerated cured (Figure 5.8) prior to surface preparation and adhesive application.

In the United States, PRG-320 requires that the strength, wood failure and durability of

88
finger joint to be qualified in accordance with the Standard for Wood Products - Structural Glue

Laminated Timber A190.1 (ANSI 2012a). During the preparation of this manuscript, it is unsure

whether Structurlam panels have plant pre-qualification of FJ according to ANSI/APA A190.1 or

CSA O177 (Association et al. 2006). Therefore due to the presence of FJ and its observed relatively

low strength, the average Structuralam specimen major strength MOR was 46.6% lower than Pilot

panels. The authors suggest that future studies to be performed at improving FJ quality in terms

of its typology, geometry and adhesive as well as smart engineering of the FJ to avoid its presence

in tension layer especially the area subject to significant amount of moment.

Figure 5.6: Finger joint failure in Structurlam major strength specimens

Figure 5.7: Edgewise finger joint (left); flatwise finger joint (right)

89
Figure 5.8: Structurlam lumber finger joint process

5.4.2 Failure mechanism for minor strength direction specimens

This section discusses the failure mechanism of minor axis bending specimens. Load-

displacement histories of minor axis bending tests are shown in Figure 5.3. The failure mechanism

of minor axis specimens was different compared to the major axis ones; however, all minor strength

specimens failed in a similar manner. When load on the specimen was gradually increased, the

minor axis test panel exhibited the following typical failure sequence (Figure 5.9):

1. Opening of vertical gap closest to the mid-span in the tension side (bottom layer)

of the panel. Note that the panels were not edge glued;

2. crack initiation at the gap in the longitudinal (middle) layer;

3. bending failure in the longitudinal layer. This bending failure would either prop-

agate to the upper glue line causing delamination in that interface or confined within the

longitudinal layer (Figure 5.9). Typically, the ultimate failure was accompanied by a very loud

noise associated with rupture of wood fiber in tension.

5.4.3 Characteristic Bending Stiffness and Strength

In order to quantify performance-rated CLT for structural applications, the tested charac-

teristic values in major and minor strength direction have to meet the minimum required criteria

provided in PRG-320. For ‘V3’ CLT bending properties, Table 5.1 lists the minimum values for

bending stiffness EI and bending strength Fb S. The ‘V3’ designated CLT was chosen because its

the closest configuration to the tested specimens, with only cross layers switched to No.2 Southern

90
(a) rupture and delamination in the glue line (b) rupture and propagation within the longitudinal layer

Figure 5.9: Failure mechanism of minor axis specimens

Pine lumber. PRG-320 defines the characteristic values as the structural property estimate, typically

a population mean for stiffness properties or a tolerance limit (5th percentile with 75% confidence)

for strength properties. Therefore, the reported mean values for bending stiffness in Table 5.5 be-

come characteristic values. Clearly, by comparison of Table 5.1 and Table 5.5, Structurlam 3-ply

CLT did not pass the stiffness requirement in major strength direction.

The minimum required characteristic strength values listed in Table 1 of PRG-320 is a one-

sided tolerance limit (TL) associated with 5th percentile and 75% confidence level. To be more

specific, it is a one-sided lower TL as defined by ASTM D2915 (ASTM 2010). Eq. (5.9) was used

to determine the characteristic test values:

YL = Ȳ − kS (5.9)

where Ȳ is the sample mean and S is the sample standard deviation. k is a parametric confidence

interval factor which determines that the interval covers at least proportion p (1-5%=95%) of the

population with confidence γ (75%). As a parametric factor, the value of k will depend on the type

of distribution which best describe the test results. Major and minor bending strength tested values

were fitted to Normal, Lognormal, Gumbel, Frechet, and Weibull distributions. Based on statistical

R2 and KS test, it is determined that Normal Distribution yields the best fit (Figure 5.10).

For Normal Distribution, ASTM D2915 (ASTM 2010) conveniently provides a k table that

can be looked up based on p, γ and sample size N (K8 = 2.189, k10 = 2.104). For T L based on

distribution other than Normal, a non-central t-distribution is needed to calculate the k factor. The

91
calculated lower T L strength values and population mean stiffness values are provided in Table 5.6.

(a) Major (b) Minor

Figure 5.10: Normal distribution for bending strength

Table 5.6: Characteristic bending test values

Direction Adhesive Finger Joint MOE,106 psi TL MOR,ksi MOE MOR

required,106 psi required,ksi

MF No 2.38 5.36
Major 1.6 2.045
PUR Yes 1.20 2.75

MF No 2.62 5.34
Minor 1.4 1.205
PUR Yes 1.86 4.47

According to PRG-320 performance CLT criteria, all bending strength characteristic values

have passed the standard. In terms of bending stiffness, Structurlam major strength direction

specimens failed the requirement by a margin of 25%.

5.5 Monte Carlo method for bending strength estimation

On the basis of above test, Monte Carlo simulation was proposed to estimate bending

strength of composite CLT based on random sampling individual lumber properties. The estima-

tion relied on shear analogy method and equivalent section area, and it took inputs in terms of

CLT configuration, layup, thickness of lumber and CLT panel, and different stiffness and strength

parameters of lumber. First, the bending stiffness and maximum moment carrying capacity of CLT

were derived.

92
5.5.1 Bending Stiffness ECLT of CLT Panel

In the shear analogy method, the characteristics of a multi-layer cross section are separated

into two virtual beams A and B. Beam A is given the sum of the inherent flexural and shear

stiffness of the individual layers along their own centers, while beam B is given the ‘Steiner’ points,

or increased moment of inertia because of the distance from the neutral axis (see Figure 15). These

two imaginary beams are coupled with infinitely rigid web members and therefore parallel axis

theorem is used to sum the bending stiffness of two beams (EIA and EIB ):

n
X
EICLT = EIA + EIB = Ei (Ii + Ai Ni2 ) (5.10)
n=1

where EICLT is the bending stiffness of CLT panel, usually calculated per foot of CLT width; Ii is

the moment of inertia in the i-th layer; Ni is the distance from neutral axis to centroid of each layer;

Ai is the equivalent section area calculated by width bi times thickness ti .

The location of the neutral axis is calculated in Eq. (5.10):

Pn
(bi ti Yi )
N = Pn=1
n (5.11)
n=1 (bi ti )

where N is the distance from neutral axis to bottom edge of the panel; Yi is the distance from the

centroid of i-th layer to bottom edge; n is total number of layers. Note that equivalent width of the

i-th layer can be obtained from Eq.(5.4) by using the mean value of parallel E0 .

Figure 5.11: Beam modeling using shear analogy method (adapted from Gagnon et al. (2011))

5.5.2 Maximum Moment Carrying Capacity

Maximum moment carrying capacity of CLT panel is estimated using the properties of lam-

ina which compose the CLT. Okabe et al. (2014) proposed an interaction equation which originated

from glulam bending strength calculation. Failure of CLT is determined by the interactive state of

93
bending and tensile stresses of lumbers in each layer:

σbi σti
+ =1 (5.12)
fbi fti

The derivation of the maximum moment carrying capacity is briefly repeated here, the

readers are referred to Okabe et al. (2014) if further details are needed. Assuming that the layers

are perfectly bonded together without any delamination or slippage, the bending curvature ρ of

panel should be universal in each layer:

1 M Mi
= = (5.13)
ρ EICLT Ei Ii

Ei Ii
Mi = M (5.14)
EICLT

where M and Mi are the moment of CLT panel and moment of i-th layer, respectively.

With the same assumption that that the CLT panel reserved its composite action, the

bending stress and tensile stress of each layer is expressed in Eq. (5.15) and (5.16):

Mi Ii Ei ti Ei
σbi = = · ·M = · ·M (5.15)
Si Si EICLT 2 EICLT

Ni Ni Ei
σti = i · Ei = · Ei = ·M (5.16)
ρ EICLT

where σbi is the bending stress in the i-th layer; σti is the tensile stress in the i-th layer; Si is the

section modulus of i-th layer; i is the axial strain of the i-th layer. Substituting Eq. (5.15) and Eq.

(5.16) in to Eq. (5.12) gives CLTs maximum moment carrying capacity in terms of its geometry

and lumber component characteristics:

EICLT 2fbi · fti


Mmax = · (5.17)
Ei (ti fti + 2Ni fbi )

where Ni is the distance from the centroid of i-th layer to the neutral axis; fbi and fti are the bending

and tensile strength of lumber at i-th layer, respectively. Due to the failure mechanism observed in

the bending test, the out-most longitudinal tension layer was used to compute the maximum moment

carrying capacity (i.e. i=3 for major strength direction and i=2 for minor strength direction).

94
5.5.3 Bending Property of Southern Pine No.2 Lumber

A statistical database for mechanical properties of various lumber species initiated in the

1978, known as the In-Grade Testing Program (IGTP) was selected as the baseline to compare the

No.2 Southern Pine lumber used in this study. The IGTP was motivated to establish a statistical

database for mechanical properties of various lumber species and to establish reference design values.

Bending, shear, tension and compression stiffness/strength were evaluated in this test program for

various common US wood species, including Douglas Fir-Larch, Hem-Fir, Southern Pine and other

softwood and hardwood species. The program took 12 years to test over 70,000 pieces of full size

dimension lumbers, randomly selected from sawn mills all over the nation. As a matter of fact,

several ASTM standards (e.g. D198, D245 and D2555) were published in advance to support the

testing program. In 1977, ASTM D1990 was published to outline the criteria for adjustment and

interpretation the data, and therefore established the design values of visually graded lumbers.

The IGTP test data was fit into five distributions (Normal, Lognormal, 2-P Weibull, 3-P Weibull

and non-parametric) with the parameters and statistical tests (KS, Anderson-Darling and Shapiro-

Wilk) given for each corresponding estimate. In this study, Southern Pine No.2 2x6 lumber at 12%

moisture content test data (Green and Evans 1987) fitted to a lognormal distribution was selected

as the baseline. The reason for selecting lognormal distribution is three-folded: first, lognormal

distribution is better at capturing extreme values and it has a long tradition being used in the

sawn lumber industry to model mechanical properties; second, it will generate positive-only random

numbers during Monte Carlo simulation; third, lognormal distribution passed all three statistical

tests in terms of their critical values. The lognormally fitted parameters for bending stiffness (MOE),

bending strength (MOR) and ultimate tensile strength (UTS) were tabulated in Table 5.7.

Table 5.7: Lognormal distribution parameters for IGTP No.2 2x6 SP lumbers at 12% moisture
content

Property Sample Size Logarithmic Mean Logarithmic STDEV

MOE 413 0.463 0.247

MOR 413 1.889 0.447

UTS 412 1.239 0.537

A ‘mini’ in-house lumber testing program was also carried out to verify whether the lumber

used in this study falls into the similar distributions described in Table 5.7 (with the exception of

95
UTS). It is worth noting that the in-house test program had no intention to replicate the effort

establishing design values, but simply to verify whether the lumber received for CLT Pilot panel

manufacture were on par with the ones tested in IGTP. ASTM D198 (ASTM 2015), D1990 (ASTM

2016a) and USDAs technical report on IGTP ‘Procedure for Developing Allowable Properties for

a Single species Under ASTM D1990 and Computer Programs Useful for the Calculations’ (Evans

et al. 2001) were strictly followed in terms of test procedure and data analysis.

5.5.4 In-house bending test

ASTM D1990 (ASTM 2016a) refers to ASTM D198 (ASTM 2015) and ASTM D4761 (ASTM

2013) for mechanical test methods. These standards offer a wealth of information on the mechan-

ical test methods that can be used. Missing from these standards is the identification of the test

conditions that are consistent with adjustment used in establishing allowable properties. The IGTP

was based on 2x8 bending tests that are third-point loaded with a span to depth ratio 17, the MOE

values have deflections measured at load heads. Such setup are referred to the standard conditions

hereafter. For test conditions that are deviated from this, ASTM D1990 as well as In-Grade Test

technical report (including detailed explanations and examples) provides all the adjustment equa-

tions necessary to bring the raw test data to standard conditions. A slight deviation from the IGTP

standard specimen is that the moisture content of raw MOE and MOR data was adjusted to 12%

rather than 15% percent, since PRG-320 target the average lumber moisture content on 12% at time

of manufacturing. It is worth noting that their applicability to the test conditions depends upon

how different those conditions (i.e. specimen size, span length. M.C., test temperature, etc.) are

from the assumed conditions. A flow chat published in the ASTM D1990 is shown in Figure 5.12,

which was helpful in understanding the steps of the adjustment process.

The sampling plan to pull out the in-house bending specimens from 2x6 bundle followed the

instruction provided in the ASTM D2915 (ASTM 2010). The lumbers were sequentially numbered

upon their arrival and twenty-one numbers were randomly generated to sample the lumbers. All

lumber pieces were tested on a third-point bending setup with clear span to depth ratio 76.5. Draw

wire deflection measurement was taken at the bottom of the mid-span. A MTS load head with

rate of 0.1 in./sec applied load to the specimens (see Figure 5.13) . Ultimate bending failures were

observed for all twenty-one specimens. Note that ultimate tensile strength (UTS) test was not

performed in this study. After the adjustment process based on ASTM D1990, the standardized

96
1
Figure 5.12: Flow chart of ASTM D1990 adjustment process. M.C. was adjusted to 12% in the
in-house tests

data was fit into lognormal distribution and compared to IGTPs published parameters (Figure

5.14). Two statistical goodness-of-fit tests, namely the KolmogorovSmirnov test and the Anderson-

Darling test were implemented to investigate the significance of the difference between observation

and distributions described in the IGTP. Note that all the critical values were calculated based

on degree of freedom equals to the number of specimen minus one (20) and 5% significance level.

The test statistics in Table 5.8 revealed that neither hypothesis was rejected, indicating the tested

values were robust as a representative of the In-Grade test program whereby large population were

examined. In addition to statistical tests, the correlation relationship between MOE and MOR were

investigated. It is found that MOE and MOR were positively correlated with a correlation coefficient

0.684 (Figure 5.15). Similarly, it is believed that UTS is also positively correlated with MOE and

MOR, however, their correlation was beyond the scope of this study and were not considered in the

Monte Carlo simulation. The lognormally distributed UTS obtained from IGTP is plotted in Figure

5.16.

97
Figure 5.13: In-house No. 2 2x6 Southern Pine bending test

(a) MOE (b) MOR

Figure 5.14: Lognormal distributions for IGTP and in-house bending test

5.5.5 Monte Carlo Simulation

It was concluded that the donated lumbers for manufacturing Clemson Pilot panels were

representative of the No.2 2x6 Southern Pine species group. On the basis of this conclusion, the

in-house lognormal estimates for MOE and MOR (solid line in Figure 5.14), along with lognormal

UTS estimates obtained from IGTP were used to generate random numbers for the simulation.

The MOE and MOR random numbers were correlated with a coefficient equals 0.684 based on the

explored correlation. For each simulation, six and three sets of random variables were generated

for each major and minor strength direction Monte Carlo specimen, respectively. In fact, six and

three pieces of longitudinal lumbers were most commonly seen in a major and minor strength 12 in.

98
Table 5.8: Test statistics for in-house bending and IGTP

Statistical Test MOE Test Statistics MOR Test Statistics Critical Value
KolmogorovSmirnov 0.137 0.098 0.288
Anderson-Darling 0.789 0.424 2.502

Figure 5.15: Correlation of MOE and MOR

wide CLT cross section (Figure 5.17), occasionally seven and four were observed. However, for the

sake of simplicity, six and three were selected as number of longitudinal boards. The variation of

lumber properties along its length due to the presence of finger joint is not specifically modeled in

the Monte Carlo simulation. The information regarding location of finger joints was not attainable

during the preparation of this manuscript. However, it can be fully implemented once the lengths of

lumbers before finger joint process are profiled. The crosswise lumbers were neglected in the Monte

Carlo simulation since bending failure was never observed in the cross layer.

Bending strength of CLT panels were simulated as the following procedure:

(a) Prepare cumulative distribution function of lognormal probability density for

MOE, MOR and UTS;

(b) generate random variables for each specimen and set total number of simulations

equals 1 million;

(c) adjust the equivalent width of crosswise layers by Eq. (5.4) and calculate section

properties using and Eq. (5.6) and Eq. (5.10). Note that the 2x6 tested mean bending stiffness

1.550· 106 psi was used as E0 in Eq. (5.4);

99
Figure 5.16: Lognormal distributions of IGTP UTS

Figure 5.17: Typical cross section of major (left) and minor (right) strength direction 3-ply CLT
specimen

(d) maximum moment carrying capacity of CLT panel was calculated using Eq. (5.17).

Although random numbers were generated for each longitudinal layer, only the lowest tension

layer was checked for failure criterion (Eq. (5.12)). For instance, the bottom layer and the

middle layer in a 3-ply CLT were checked for failure in major and minor strength direction,

respectively. With the section modulus S calculated in Step (3), the simulated CLT MOR was

obtained by applying Eq. (5.7).

5.5.6 Simulated Results

Figure 5.18 shows the Monte Carlo simulation results of 3-ply CLT MOR for both major

and minor strength directions. KS and Anderson − Darling statistical tests were performed with

the null hypothesis that observed test data was compatible with the proposed distribution from

100
Monte Carlo simulations. The test statistics shown in Table 5.9 indicates simulations matched well

with Pilot Panel test results but very poorly with the commercially made panels that contained

finger joints. This is due to fact that finger joint did not have sufficient bending capacity such

that premature failures were observed in the commercially made Structurlam specimens. In the

meanwhile, finger joints were not specifically modeled in the simulation. This is a very unique

situation where longitudinal boards in the Pilot specimens were continuous without any finger joint,

thus only one set of random variables was generated for each piece of lumber in the Monte Carlo

simulation. As a matter of fact, this is not the case in the mass production of CLT. Commercially

manufactured CLT or glulam will often contain finger joints. Their capacity varies due to different

mating orientation (Figure 5.7), adhesives, curing techniques employed in the manufacturing, etc.

In addition, the lumber pieces feeding the finger joint machine do not always come with same

length and each manufactures have their own quality control standards where large defects are often

automatically detected and cut off. In other words, the strength and location of the finger joint

in the CLT or other engineered wood product become two additional variables that have yet been

fully implemented in the Monte Carlo simulation. With bending tests on finger joint assembly and

lumber piece dimension measurement, distributions can be properly built around these two random

variables. This enables a more realistic and much desired Monte Carlo model to simulate CLT

bending capacity for commercially produced CLT.

(a) MOR Major Direction (b) MOR Minor Direction

Figure 5.18: Results of simulation on bending strength of 3-ply CLT panel

101
Table 5.9: Test statistics for CLT MOR and Monte Carlo simulations

Statistical Test MOR Major Direction MOR Minor Direction Critical Value
KolmogorovSmirnov 0.240 0.234 0.454
Anderson-Darling 0.755 0.251 2.571

5.6 Conclusion

The following results were obtained in this study:

1. 12 feet long 3-ply Southern Pine Clemson Pilot CLT panels were produced in a 4

ft. by 4 ft. hydraulic press using staged press. No premature glue line failure was observed in

the bending test;

2. for Structurlam major strength direction panels, it was found that bending stiff-

ness was below the minimum requirement by 25%. Low density of the wood was considered

as the cause for this.

3. a Monte Carlo method that utilizes interaction equation for failure criterion was

proposed to model the bending capacity of composite CLT. MOE, MOR and UTS of 2x6 sawn

lumber served as the input parameters for the model. An in-house bending test was performed

on 21 No.2 Southern Pine lumbers. The obtained results were adjusted to standardized data

based on ASTM D1990 and fitted into lognormal distributions. Results confirmed that those

lumbers were representative as a species group tested in the In-Grade Test Program. The

Monte Carlo simulation showed promising results compare to the Pilot panel bending strength;

4. the bending capacity and location of the finger joint are not considered in the

Monte Carlo model. The future research could be directed to include these two parameters

in the model and to look into how finger joint can be optimized or engineered in terms of

their configuration and location so that full composite capacity of the lumber can be better

exploited.

102
Chapter 6

Bending and Shear Strengths of

Composite CLT-Glulam Beam

6.1 Introduction

Wood construction offers many advantages that typically not found in other construction

materials. Wood is relatively inexpensive, renewable, easy to work with (light-weight), strong (high

strength to weight ratio) and adaptable. The economic, environmental and energy efficiency advan-

tages are the primary reasons that more residential buildings constructed of wood than any other

structural materials. The International Building Code (IBC) is a building code that establishes the

minimal requirements for building systems using prescriptive and performance related provisions.

For different building materials, the IBC refers to the material specific construction manuals. For

example, the latest version of IBC International Building Code, refers to AISC 360 (Specification for

Structural Steel Buildings), ACI 318-14 (Building Code Requirement for Structural Concrete) and

NDS 2015 (National Design specification for Wood Construction) for steel, concrete and wood con-

struction, respectively. To determine the load requirements, the IBC makes references to ASCE/SEI

7-10 (Minimum Design Loads for Buildings and Other Structures).

CLT has been formerly recognized as a construction material in the 2015 IBC. The appli-

cation of CLT in difference construction types is discussed in section 2.4 of this dissertation and it

is briefly summarized here for convenience:

103
1. Type III Construction: CLT is only permitted for use as interior building

elements;

2. Type IV Construction: Type IV is commonly known as ‘Heavy Timber’ con-

struction. It requires that the exterior walls made of non-combustible material and the interior

building elements of solid or laminated wood without concealed space. However, the 2015 IBC

has prescriptively allowed the use of CLT in Type IV construction, including exterior and

interior walls, floors and roofs, provided that the CLT elements meet the code specified fire

rating.

3. Type V Construction: This type of construction allows the use of any structural

materials permitted by the IBC code. For CLT, the 2015 IBC and NFPA 5000 (Association

et al. 2005) both refer to ANSI/APA PRG 320, Standard for Performance Rated CLT. As such,

any CLT complying with the PRC 320 standard is permitted for use in Type V construction.

4. Type I and II Construction: Type I and II are those types of construction

in which the building elements are of noncombustible materials. While untreated wood is

considered combustible, wood products are still permitted in Type I and II construction,

given that these wood products are treated with fire retardant and used in the code-specified

situations. However, during the preparation of this dissertation, there is no known commercial

production of fire retardant treated CLT; thus, Type I and II are not currently used for any

mass timber construction.

With the recent development and advancement of CLT, the wood construction is entering a

new renaissance period as increasing numbers of building owners and developers often consider mass

timber in lieu of steel and concrete in mid-rise construction because of its advantages (e.g. sustain-

ability). Since CLT panels are pre-fabricated, the accuracy and stability of the panel dimensions

allow for more rapid construction with fewer crews. The light weight nature of CLT, when compared

to precast concrete panels, results in smaller size of cranes and operational spaces needed at the job

site. Since heavy equipment is not needed, the job site is typically safer and the operational cost is

lower. Another benefit of mass timber system is sustainability as wood is considered a renewable

building material that can be replenished with a proper forest management. Wood is considered

a carbon sequester. It also takes less energy to harvest and process than steel or concrete. The

carbon sequestration nature of the wood itself along with other smart energy management systems

104
implemented in a mass timber building can actually result in a net negative carbon footprint over the

life cycle of the building. The structural efficiency and resiliency are also advantages of mass timber

construction. Wood has a very high strength-to-weight ratio, which often leads to a lightweight,

strong and stiff building. The shake table test performed for a full-scale seven stories CLT building

has proven that the mass timber building is able to withstand some of the most serve earthquakes

(Ceccotti et al. 2013).

Today’s mass timber has greater structural capabilities compared to the traditional light

frame construction. The development of CLT has opened up new opportunities to use mass timber

in larger and taller wood buildings. To date, at least two studies have systematically looked into the

possibility of constructing timber skyscrapers and the type of mass timber structural systems that

is needed to make this viable. In the first study (Green and Karsh 2012), mass timber is used as the

predominant load-resisting system in the design of a twenty-story tall building. For lateral system,

a mass timber core is used with steel W sections to create a strong column weak beam design. The

steel beams with Reduced Beam Section (RBS) ensures the plastic hinge to form in the beams, rather

than the strong laminated solid wood panels and glulam columns. With the strong column weak

beam design, the risk of weak story collapse is reduced. The second study examined the conceptual

design of a 42-story CLT-concrete hybrid building in Chicago. This study is called ‘Timber Tower

Research Project’ which is carried out by the architecture and engineering firm Skidmore, Owings

and Merrill (SOM). The design utilizes mass-timber as the main structural material to minimize the

embodied carbon footprint of the building. The study also benchmarked the carbon footprint of the

wood solution to that of a concrete system. It was found that the mass timber building is able to

reduce 65% to 70% of the total carbon footprint generated by reinforce concrete or steel solutions

(Skidmore and Merrill 2013).

With the recent trend of using wood in taller and larger buildings, there is a need to

develop a pre-fabricated composite wood floor system that can span a long distance (e.g. 40 ft.

or longer). Conventional light-frame or stick-frame wood construction may not be able to sustain

the structural demand for longer spans. The presented research proposed a build-up mass timber

floor system consisting of CLT as the top and bottom flanges and glulam beams as the webs. The

design, assembly and experimental investigations of both small and full-scale composite specimens

are discussed in this chapter. The serviceability and strength of the proposed floor system were

evaluated via non-destructive and destructive tests, respectively. Note that several research projects

105
on CLT and mass timber floor systems were also initiated about the same period of this study. In

the following section, a literature review is presented on the topic of mass timber floor systems and

the motivations of this study are discussed.

6.2 Background and motivation

The current commonly used or proposed mass timber build-up floor systems can be broadly

grouped into one of the two categories, namely timber-concrete composite system or timber-only

system. When concrete is used, it provides a higher stiffness and strength than that of an all-wood

floor system. The timber-concrete composite floor also provides improved acoustical performance in

the floor system.

The first composite floor system (i.e. timber-concrete composite) usually involves the use

of shear connectors or dowel type fasteners to transfer the shear between timber and concrete to

achieve composite action. One such example is the timber-concrete composite system used in the

University of British Columbia (UBC) Earth Sciences Building. The composite floor system used in
1
the UBC building is known as the Holz Beton Verbund (HBV) system where metal mesh plates

were adhered and embedded into the CLT panels and topped with concrete (Figure 6.1). Clouston

et al. (2005) performed shear and bending tests on the HBV composite system. They found that the

timber-concrete composite floor behaved nearly as a full composite action system and ductile failures

were consistently observed in the metal mesh plate. Other than the metal mesh plate, traditional

dowel type anchorages have also been used in timber-concrete composite system. Steinberg et al.

(2003) tested the shear capacity of five types of timber anchorages topped with light weight concrete

(Figure 6.2). They concluded that the capacity and stiffness of the anchorages were able to provide

adequate shear transfer at the timber-concrete interface. Steinberg et al. (2003) also found that the

overall behavior of the composite system was mainly governed by the stiffness and strength of the

connectors.

A different example of timber-concrete floor system can be found in the Life Cycle Tower

in Durbin, Austria. For each floor slab, a layer of reinforced concrete is bonded to glulam beams,

creating a decking that performs well both thermally and acoustically. The composite floor is

supported by glulam double posts to achieve a complete vertical load path. The schematic drawing
1 Holz Beton Verbund are German words for Wood Concrete Composite.

106
(a) (b)

Figure 6.1: Timber concrete composite with mesh steel plate (Equilibrium Consulting Inc.)

Figure 6.2: Dowel type anchorage used in timber-concrete composite (Steinberg et al. 2003)

of this system can be found in Figure 6.3.

(a) (b)

Figure 6.3: Timber concrete composite system in Life Cycle Tower (Buildipedia)

107
Jacquier and Girhammar (2015) performed bending tests on all-timber or timber-only com-

posite glulam-CLT beams connected with double sided punched metal plates and inclined screw

(individually and combined). The specimen tested were 20 ft. long inverted T shaped beams (see

Figure 6.4). They observed withdraw failure of the plates at peak bending strength when only the

plates were used to connect the glulam and CLT panel. When the combination of both the punched

metal plates and inclined screw were used to connect the glulam and CLT panel, a marginal im-

proved in the peak bending strength was observed. The failures in the specimen with both punched

metal plates and inclined screws were concentrated in the glulam and CLT, indicating a satisfactory

bond at the interface.

Figure 6.4: Cross section of composite glulam-CLT beam with double sided nail nail plate and screw
connection (Jacquier and Girhammar 2015)

Similarly, Masoudnia et al. (2016) experimentally tested a 20 ft. composite T-beam com-

posed of CLT slab and LVL web. The members were connected with 48 self-tapping screws and the

test setup is shown in Figure 6.5. Their results indicated that the consideration of composite action

in the design of CLT composite beams greatly decrease the required material and cost. By forming

a composite section, the flexural stiffness of the T-beam was increased by almost 2.5 times when

compared to the same T-beam configuration but without composite action.

Theoretically, a conventional flat CLT slab with adequate thickness or plies may be used for

long span application (e.g. greater than 30 ft.). The technical design guide published by Structurlam,

a CLT producer located in the Canada, indicated that for a typical office or classroom loading, a

108
Figure 6.5: CLT-LVL composite beam test setup (Masoudnia et al. 2016)

9-ply, 12.4 in. thick CLT is able to span 26 ft, with vibration controlled the design. While a thick

CLT panel can achieve the required span distance, it may not be economical to use thick solid

CLT panel. Furthermore, a thick CLT slab will increase the self-weight of the building and in turn,

increase the seismic demand of the building.

T-beams and boxed-beams have a long history in long-span applications in the buildings

and other infrastructures. For example, prestressed concrete T- beams are often used in parking

garages. The low profile of the T-beam allows sufficient vertical clearance for vehicular traffic. T-

shaped and boxed girders are also commonly seen in bridge applications, although the strength

and span demands often require a much deeper section. T-beams and boxed girders are efficient

geometric shapes. According to parallel axis theorem, the material further away from the neutral

axis has greater contribution in resisting bending whereas the material near the neutral axis has

minimum contribution in resisting bending. Thus, in pursuing of a more efficient cross section, a

more efficient T-shape or boxed girder shape can be utilized to improve the bending stiffness and

strength. In summary, a CLT slab with supporting glulam or other engineered lumber may be viable

and efficient solution in fulfilling long span requirements.

109
Unlike precast concrete T-beam, where tensile forces are mainly resisted by the prestressed

tendons, the tensile capacity of timber beams or webs is relatively low compared to steel tendons

or rebars. One popular alternative is to strengthen the structural composite lumber with fiber

reinforced polymer (FRP) and the FRP timber beam has already been successfully implemented

as bridge girders. Figure 6.6 shows several deep FRP reinforced glulam girders erected in place

during the bridge construction. However, FRP is relatively expensive and difficult to implement.

The presented research has favored the ‘all timber’ solution in the hope that the aforementioned

advantages of mass timber construction can be fully retained. Therefore, a CLT slab attached to

the bottom of the CLT-glulam T-beam (i.e. forming a boxed section) was proposed to provide

additional flexural and tensile capacity needed in long span situations. This has essentially led to

the concept of hollow-core mass timber composite beam, as shown in Figure 6.7. The proposed

hollow-core timber composite beam is analogous to the precast concrete boxed girders. The overall

thickness of the build-up members (i.e. the sum of the CLT slab thicknesses and glulam web depth)

is governed by the loading and span requirements; however, due to the limitation of lane width

during transportation, the width of the pre-assembled composite beam is limited to the dimension

of a standard semi-tractor trailer (around 93” or 7.75 wide).

Figure 6.6: Farfield bridge with FRP reinforced glulam girders (Advanced Structures & Composite
Center)

One challenge faced by this system is the need for shear transfer between the CLT and glulam

to ensure that satisfactory composite behavior is achieved. Typically, this is achieved with structural

110
Figure 6.7: Proposed concept of hollow mass timber composite beam

wood adhesive or mechanical fasteners. Both options either used individually or combined are able

to offer sufficient stiffness and composite action. However, using adhesive to attach CTL to glulam

requires one to maintain a large clamping force until the adhesive is fully cured, which may take

several hours. In addition, there is anecdotal evidence to show that, during a fire outbreak or under

an elevated temperature condition, the bonding strength of certain adhesives may deteriorate. Due

to these limitations, the proposed hollow-core composite beam system utilized mechanical connectors

(i.e. screws) to attach the CLT and glulam.

While there is an increasing demand for mass timber long span system, there is only a scant

body of knowledge existed on the behavior of composite CLT or mass timber beams, in particular,

full-scale experimental data are very limited. To address the research gaps, a series of modeling and

experimental studies were proposed and carried for the proposed CLT-glulam composite system.

The objectives of the presented study are summarized as the following:

1. To investigate the feasibility of using inclined self-tapping screws as means of

connecting build-up members via shear test of the CLT-glulam connection and to confirm the

theoretical estimation of the connection strength.

2. To design a composite floor system that is able to sustain the load of a typical

office occupancy (100 psf live load) using the best interface connector configuration determined

in via the connection test.

3. To evaluate the assembly process and provide suggestions for future research

4. To destructively test the build-up composite beam in both bending and shear and

use the test results to validate the design procedure for the composite beam.

111
6.3 Self Tapping Inclined Screw and Small Scale Tests

6.3.1 Selection of web member and connection for the composite beam

The work presented in this section was built on the foundation of the past CLT research

performed at Clemson University. The concept of hollow-core mass timber beam was initially pro-

posed by Montgomery (2014). A series of experimental studies of the connection between CLT and

glulam were carried out and the results were documented in the thesis by Montgomery (2014).

The main reasons that standard flat CLT panels were selected as the top and bottom flanges

of the composite system are as follows. First, the cross lamination or layup of standard CLT panels

provides bi-directional load resistance and dimensional stability in a two-way bending behavior.

Second, the layup of the CLT flanges may be modified with custom layups that are beneficial for the

composite system, provided that the custom CLT panels meet the performance rating of the PRG-

320 standard. For example, a thicker CLT and/or better grade lumber may be used in lieu of the

recommended layups in PRG-320 to increase the load resistance. Third, the standard dimensions of

CLT panels are well within the target dimensions of the proposed pre-assembled mass timber long

span system. In a typical CLT manufacturing facility, the dimensions of the panel can be as large

as 10 ft. by 50 ft. and 12 in. thickness (Stora Enso, Structurlam, etc). Therefore, CLT panels were

selected for use as the flanges.

A wide range of engineered wood products is suitable for use as the webs of the hollow-core

composite system. The list of possibilities considered for the web includes: dimensional lumber,

I-joist, glue laminated timber (glulam), structural composite lumber (SCL), parallel strand lumber

(PSL), laminated veneer lumber (LVL), truss (include wood truss and open web truss). The selection

of an appropriate web material is a matter of selecting a product that better suits the availability,

structural, non-structural and fire requirements of the intended application of the composite beam.

Based on the target design load (100 psf), dimensional lumber and I joist were excluded

from the consideration. Preliminary calculations revealed that for a 30 ft. span with standard

loading, dimensional lumber and existing I-joists available on the marker are impractical to use due

to unreasonably tight spacing.

SCL and LSL were initially considered in this study. Compared to SCL, LSL is typically

cheaper and it provides a higher shear strength. LSL beam offers many advantages such as cost

competitiveness, the ability to be fabricated straight and long. However, LSL was not selected for the

112
final design because the relatively poor screw withdraw capacities observed in the tests performed

by Closen (2013). Closen concluded that self-tapping screws had a tendency to split the LSL beam

section because, compared to solid wood, the LSL lumber is generally drier and more porous, which

are not good qualities for achieving strong mechanical connection.

Wood and open-web truss were ruled out because of fire concern and its construction in-

compatibility with self-tapping screws. The web of the trusses is very thin and the hollow portion

of the truss would allow fire to quickly spread transversely. Second reason is that the shallow top

flange (usually 2x material placed horizontally) does not have enough depth to embed the inclined

self-tapping screw.

Glue-laminated timber (glulam) was considered a good candidate as the web members used

in the composite floor system. The glulam beams have been mainly used as bending members;

however, they also provide sufficient shear resistance due to their deep sections. The dimension

requirements as a result of long span system are satisfied since the mature glulam industry is capable

of manufacturing long, deep sections with various widths. As far as the attachment method, glulam

is feasible since the flat surface and nature of deep section allows the application for both wood

structural adhesive and self-tapping screw.

In conclusion, CLT flange with glulam web, attached by self-tapping inclined screw is pro-

posed as a cost effective solution for mass timber long span system. The design example of a 40

ft. long single span composite CLT-glulam system for office occupancy is presented in section 6.4.

The determination of the member dimensions for CLT and glulam and the spacing of the screws are

discussed in the next few sections.

6.3.2 Joint with inclined screw

The preliminary analysis indicated a feasible solution of mass timber long span system

that consists of CLT flanges and glulam beam webs, connected using self-tapping inclined screw

(Montgomery 2014). However, the actual screw type and required spacing of the connection, which

ensures the proper engagement of composite action and the design procedure for the composite

system were not determined by Montgomery (2014). The stiffness and the strength of the connection

are the most important parameters that govern the overall behavior of the composite beam.

The important geometric parameters of screws are specified in Figure 6.8. In general, fully

threaded inclined screw is distinguished between lag screw and self-tapping screw. Lag screws used

113
in the traditional timber design have part of the shank without threads (denoted ls in (a)), and

the lag screws often require pilot holes. In contrast, self-tapping screw do not require pilot holes.

However, it is beneficial to insert self-tapping screws into pre-drilled holes to provide a proper guide

for achieving the target angle, especially for those long and large diameters screws. Self-tapping

screws can have a portion of non-threaded shank, either near the screw head or in the mid-section of

the screw, which enables pre-tension of the screws (Figure 6.8(b)). In a timber-to-timber connection,

self-tapping screws with a non-threaded shank in the mid-section with two different thread pitches

are often used to induce clamping force into the connecting timber pieces. In order to develop the

clamping force, the contact surface will have to be lied within the segment of the non-threaded

shank. For better reinforcement purpose, fully threaded screws are used to allow for a continuous

load development and force transfer along the whole length of the screw.

Figure 6.8: Geometric parameters for different wood screws (Jockwer et al. 2014)

In the past two decades, new high strength fully threaded screws have been widely used in

many engineered wood products. There are two main advantages for using fully threaded inclined

screws. First, the continuous thread gives more grip and more efficient force transfer between the

main member and side member. Second, inclined embedment of dowel type fasteners utilize com-

bined tension/compression and shear capacity that greatly increase the stiffness and load carrying

capacity of the joint compared to fasteners installed without an angle (Pirinen et al. 2014). Inclined

screws with diameter up to 1/2 in. and length up to 2 ft. are commonly available in the market.

The availability of self-tapping screws with continuous threads, characterized by high withdraw ca-

pacity, has led to the opportunity to conceive new engineered configuration and geometries for wood

structures using connection endowed with higher values of stiffness and load-carrying capacity. In

summary, in an inclined screw, the load transfer mechanism involves not only the bending capac-

114
ity of the screw and the embedment strength of the wood, but also the withdraw capacity of the

fasteners as well as the friction between the timber members.

The force interaction of the inclined screw in the shear connection can be explained using a

simplified strut-and-tie model (Figure 6.9). The connection force in the inclined screw is transferred

as shear and tension. The tension component works as tie (blue), while wood works as compression

strut (red), which creates friction force (yellow) between the members, increasing the stiffness and

load carrying capacity of the connection. Over the years, different models have been proposed by

many authors and code councils to predict the capacity of inclined screws. European standards (EN

1995) suggested the use of a quadratic combination formula of to combine the screw shear and axial

strengths for the definition of to determine the load carrying capacity or reference design load. Note

that this analytical equation is only in the loading condition given the for screws simultaneously

subjected to shear and axial stress load, as which in is the case of the inclined screw.

Figure 6.9: Force interaction in a inclined screw connection

According to CEN (European Committee for Standardization), the inclined screw strength

can be estimated using the quadratic shear and axial strengths interaction equation in EC5 (EN

1995), commonly known as the European code:

Fax.Ed 2 Fv.Ed 2
( ) +( ) ≤1 (6.1)
Fax.Rd Fv.Rd

where Fax.Ed is the design axial force, Fax.Rd is the axial strength, Fv.Ed is the design lateral force

(in the shear plane), Fv.Rd is the shear strength. The design axial strength values are calculated

115
on the basis of the material coefficient γm and the modification factor Kmod . The shear strength is

defined using the Johansen Model for timber-to-timber joints. However, over the years, this formula

has been shown to grossly underestimate the ultimate strength of inclined screws (Tomasi et al.

2006), making it inappropriate for designing the inclined screw connections.

To this end, the work carried out by Blass and Bejtka is of fundamental importance. Those

two authors have set forth a theoretical analytical model for inclined screws subject to combined

shear and tension forces (Blass and Bejtka 2001; Bejtka and Blass 2002). Based on the Blass and

Bejtka’s model and Eurocode 5 quadratic equation formula, a general model has been developed

for the calculation of strength and stiffness of the geometry configurations depicted in Figure 6.10.

The model can be used to estimate the capacities of the inclined screws in shear-compression joint

(Figure 6.10 left), shear-tension joint (Figure 6.10 middle), and crossed joint (Figure 6.10 right).

Based on the work of Kevarinmäki (2002), on the deformability and joint ductility, a specific model

for the calculation of stiffness for inclined screws has been developed. It allows the evaluation of the

instantaneous slip modules Kser for various inclined angles.

Figure 6.10: Common timber connection joints. Left to right: shear tension, shear compression and
crossed joint

Bejtka and Blass (2002) extended the Johansen’s yield theory (also known as the European

yield models) (Johnsen 1949) and incorporated several additional parameters in the model such as

withdraw capacity of the screw, force direction, friction resistance between members, etc. They

investigated three failure modes for single- and double-shear connections considering the bending

capacity of the fasteners. For instance, in failure mode 3, where two plastic hinges are formed,

solving the equations of equilibrium for the main and side members yields the load carrying capacity

of the joint (Figure 6.11). However, this method have very limited application since it relies on the

withdraw test values, which are not readily available for different wood species, types of screws and

116
inclined angles.

Figure 6.11: Forces and stresses for European Yield Mode 3 for a inclined screw connection (Bejtka
and Blass 2002)

Based on the work of Bejtka and Blass, Jockwer et al. (2014) developed a simple model

to estimate the load carrying capacity of the inclined screw. The resistant of the connection when

loaded in an angle α to the grain direction can be calculated from the resistances in axial (Rax ) and

in lateral (Rv ) direction of the screw:

R0 = Rax (µsin(α) + cos(α)) + Rv (sin(α) − µcos(α)) (6.2)

where µ is the friction coefficient between two contact surfaces. The axial lateral strength of the

screw can be obtained using Eq. (6.3) and (6.4):

W
Rax = (6.3)
1.2cos2 α + sin2 α

p
Rv = 2My fh def (6.4)

where My is the screw yield moment, fh is the dowel bearing strength, def is the effective diameter,

117
W is the withdraw capacity of the screw. During the joint behavior shear and pulling test, it was

found that (1) density of the timber and inclined angle α are the parameters with the most influences

on the stiffness and strength of the joint; (2) fully threaded screw shows highest strength and stiffness

properties.

In Finland, Kevarinmäki (2002) developed an equation to estimate the capacities for both

crossed joint and shear-tension joint. The type of screw is selected based on the anticipated force

demand of the joint. For example, for the sheathing nail in a shear wall, crossed joint is desired

since it is more resistant to reverse cyclic load. On the other hand, shear-tension joint is often

used for composite beams under sustained gravity load and shear connection where the direction

of the force remains unchanged during the lifetime of the structure. Kevarinmäki (2002) conducted

145 inclined screw joint tests, including both shear-tension and crossed joints. Based on the tested

strength values and failure modes, Kevarinmäki (2002) proposed an equation that can predict the

mean withdrawal strength of the shear capacity of the joint. Since the loading on a floor system

under gravity load is uni-direction and it does not involve reversed cyclic loading, the shear-tension

screw joint is used in the proposed CLT-glulam composite beam. The capacity of the crossed joint

is slightly more complex to calculate and it is not discussed here. Interested readers are encouraged

to refer to Kevarinmäki (2002). For the shear-tension screw joint, shear capacity can be calculated

as:

In Finland, Kevarinmäki (2002) developed the capacity estimations for both crossed joint

and shear tension joint. The type of screw is selected based in the nature of the anticipated force

introduced to the system. For example, when assemble wall panels, cross joint is desired since it

is more resistant to reverse cyclic load. On the other hand, shear tension joint is often used when

assemble bending and shear element where direction of the force remains unchanged during lifetime

of the structure. Based on this argument, he conducted 145 inclined screw joint tests, including

both shear tension and cross joints. Based on the tested strength values and failure modes, he

proposed the mean withdraw strength equation that can be used to predict the shear capacity of

the joint. Since the loading on a floor system is uni-directional and does not involve reverse cyclic

load, the shear tension screw joint is used in the presented study. The capacity of the cross joint

is slightly more complex to calculate and is not covered here, interested readers are encouraged to

refer to Kevarinmäki (2002) if relevant information is needed. For a shear tension screw joint, shear

118
capacity can be calculated as:

Rd = nRT,d (cos(α) + µsin(α)) (6.5)

where n is the number of screws in a joint, RT,d is the screw withdrawal capacity, α is the inclined

angle, and µ is the kinetic friction coefficient between two contacting members. The screw withdrawal

capacity depends on the withdrawal strength in the main and side members, tension capacity of the

screw and pull through strength of the screw head. The minimum strength of these three failure

modes (i.e. withdrawal, screw tension failure and pull through) is used for RT,d :


fa,1,d πds1 + fhead d2h ,







RT,d = fa,2,d πd(s2 − d), (6.6)





Fu.d

where d is the outer diameter of the thread (nominal size), s1 is the length in member 1 towards the

screw head (side member), s2 is the length in member 2 towards the screw tip (main member), Fu.d

is the screw tension capacity, fa,1,d is the withdrawal capacity in member i, dh is the head diameter,

fhead is the pull through strength of the screw head. These parameters are usually available from

the manufacturer’s product information sheet.

The validity of the method proposed by Kevarinmäki (2002) has been tested by many

researchers, including Prat-Vincent et al. (2010); Angeli et al. (2010); Tomasi et al. (2010). The

inclined screw test results led to the following conclusions:

1. The increase of screw inclination provides an increasing of the resistance and

stiffness of the joints

2. The mechanical behavior of screws subjected to shear-compression is relatively

complex. The test results of shear-compression joints showed that the shear-compression joint

is less efficient.

3. The experimental studies confirmed that the connection resistance can be pre-

dicted using both Blass & Bejtka and Kevarinmäki’s formulations.

119
6.3.3 Selection of self-tapping inclined screws and theoretical strength
predictions

With the previous section discusses the behaviors of different inclined screw joints and

the determination of a suitable configuration for composite beams under gravity load, namely, the

shear-tension joint, this section discusses the determination of the inclined angle used in this study.

The experimental studies conducted by Prat-Vincent et al. (2010), Angeli et al. (2010) and

Tomasi et al. (2010) have discovered that increase in the screw inclination angle provides an increase

in the resistance and stiffness of the joint. This finding was also observed in the inclined screw joint

tests performed at Clemson University. Montgomery (2014) tested the shear-tension CLT-to-glulam

joint with different combinations of screw diameter, length and inclined angle. The four screws used

in his study are shown in Figure 6.12:

Figure 6.12: ASSY and SFS Intec screws studied in Montgomery (2014)

• The top screw shown in Figure 6.12 is the ASSY SK2 screw. It was used in Montgomery (2014)

to clamp the glue bond during adhesive curing. It has a wide head with threads in the main

member and shank in the side member to improve clamping pressure.

• The second screw from the top of Figure 6.12 is the ASSY VG Cylindrical screw that is 5/16

in. in diameter and 11 3/4 in. in length.

• The third screw from the top of Figure 6.12 is the same ASSY VG Cylindrical screw but

significantly larger, with 3/8 in. in diameter and 15 3/4 in. in length.

• The fourth screw is an SFS Intec WT-T3 that is 5/16 in. in diameter and 11-13/16 in. in

length.
2 ASSY is the name for a family of screws with asymmetrical thread produced by SWG (Schraubenwerk Gaisbach
GmbH). http://www.swg-produktion.de/en/produktion/products/assy.html
3 WT-T is a double-threaded lag bolt system developed and manufactured by SFS Intec Switzerland.

https://sfsintec.biz

120
For a pure mechanically fastened joint with inclined screws, Montgomery (2014) investigated

six configurations with inclined angles at 30◦ , 45◦ and 90◦ . For each configuration, three tests were

performed. The failure loads averaged for three tests are summarized in Table 6.1. All six tests

were performed using a 2 ft. long CLT-glulam joint specimen that consisted of 3-ply Southern Pine

CLT flanges and a single glulam web. Two screws spaced at 6 in. on center were installed at each

flange-web interface for all test configurations, except for Test 2, where the number of screws was

doubled to investigate the group effect. The typical specimen geometry and inclined screw test setup

are shown in Figure 6.13.

Table 6.1: Shear Test Results Montgomery (2014)


Test Number Test Description Averaged Peak Load (kip)

1 Small ASSY screw 45◦ 20.01

2 Small ASSY screw 45◦ (group study) 40.36

3 Large ASSY screw 30◦ 35.91

4 SFS Intec screw 45◦ 23.03

5 SFS Intec screw 30◦ 22.90

6 SFS Intec screw 90◦ 15.74

Figure 6.13: Specimen for inclined screw joint Test 3 (Montgomery 2014)

Comparing the strength and stiffness values published in Montgomery (2014) indicates that

a larger diameter ASSY VG screw installed at 30◦ provides a desirable performance in terms of

balancing between ultimate capacity and stiffness. The results of a cost-benefit analysis also favored

121
the larger ASSY VG screw since it costs less per kip of shear capacity provided. Therefore, the 3/8

in. diameter ASSY VG cylindrical screw, installed at 30◦ angle was selected to connect the flanges

(CLT panels) to the web (glulam). The packaging/label of the ASSY VG screw, a sample screw and

the bit are shown in Figure Figure 6.14.

Figure 6.14: ASSY VG Cylindrical 3/8 in. screw

A preliminary estimation of the capacity of the selected inclined screw joint configuration

(ASSY VG screw at 30◦ angle) was carried out based on the information obtained from the literature

review. Once the theoretical values were obtained, two shear tests of the screw joint were conducted

to verify the theoretical predictions. Consider a shear-tension joint consisted of 3-ply Southern Pine

V3 CLT and in-house manufactured No.2 glulam connected by 3/8 in. ASSY VG inclined screw

installed at 30◦ . The theoretical joint capacity can be calculated using Eq. (6.2) to Eq. (6.6).

The method proposed by Jockwer et al. (2014) and Kevarinmäki (2002) were used to compute the

ultimate strength of the screw joint. The calculations are presented in Appendix B. Both methods

yielded close estimations for strength level capacity per screw, one at 10.6 kip and the other at 11.4

kip.

6.3.4 Small scale inclined screw joint test

The dimensions for the inclined screw joint test specimen are shown in Figure 6.15. The

3-ply Southern Pine CLT flanges and glulam webs were manufactured at Clemson using pilot-scale

press (Newman HP-188) with a 4 ft. x 4 ft. platen. In order to test the inclined screw in shear, a

force generated by actuator was applied to the top of the glulam offset while letting the bottom of

the two CLT flanges reacting against a strong floor (see Figure 6.18). This causes the applied load

122
to be evenly resisted by each side of the CLT-glulam connection. Two nominally identical static

shear tests were performed for the ASSY VG screws with each specimen contained four screws, two

on each side at a 6 in. on center spacing. This symmetrical two-sided test setup with two screws at

each interface eliminated eccentric load and moment from the specimen.

Figure 6.15: Specimen dimensions for the inclined screw joint shear test

The specimen assembly started with drilling pilot holes for the CLT flanges. For pilot

drillings, 7/32 in. is the diameter recommended by My-ti-con4 for softwood species (Closen 2015).

However, due to market availability, a slightly larger 1/4 in. diameter 12 in. long drill bit was

purchase for the job. A ZL WT-U mode angle guide was set at 30◦ of inclination and the drill

bit passed through all CLT layers (Figure 6.16). Note that the two darker metal inserts in Figure

6.16(a) were milled and hardened in the in-house machine shop at Clemson for guiding the 1/4 in.

diameter drill bit. An angle guide with a larger diameter hole could cause wobbling of the bit tip

and even breaking of the screw. Once the pilot holes were drilled, the unattached CLT and glulam

beams were fastened with temporary pipe clamps to facilitate the screw installation. Figure 6.17

shows a Milwaukee hammer drill driving in the long inclined screw. However, this low speed high

torque hammer drill was not able to drive in the screws completely; for a few last turns, a ratchet
4 My-ti-con is the distributor of ASSY screws in the North America

123
with long extension had to be used to fully embed the screw head.

(a) Angle guide (b) Pilot drilling

Figure 6.16: Drilling pilot holes using an angle guide

Figure 6.17: Inclined screw installation with hammer drill

A self reacting frame with strong floor was used to apply the load and provide reactions. At

bottom of the specimen, a thin sheet metal with two square tubing welded at two ends restrained

124
the CLT flanges. In fact, CLT flanges had a tendency to spread outward and split the glulam web

when under load, the restraints helped concentrate all applied forces on the screw joints. An MTS

actuator with 150 kip capacity was used to apply the load on the glulam web, with a displacement

controlled procedure at a rate of 0.1 in./min. This loading rate was used for mechanical connections

based upon the requirement of ASTM D1761 (Mechanical Fasteners in Wood ). Two LVDTs were

mounted on one side of the web member and reported displacement was the average of these two

readings. Figure 6.18 shows the setup for the small-scale shear test.

Figure 6.18: Test setup for small scale shear test

The load-displacement results for two small-scale static shear tests is shown in Figure 6.19.

It is observed from the figure that two shear test curves showed almost identical stiffness and the

first yield a slightly higher ultimate capacity at 34.75 kip whereas the second one 32.55 kip. The

averaged strength 33.65 kip was close to the value (35.91 kip) reported in Montgomery (2014). The

strength and stiffness values are important design parameters to ensure the composite behavior of

the build-up section, therefore they are discussed here. In comparison of the strength values, it

was found that both theoretical predictions overestimate the screws ultimate capacity (8.4 kip from

test, 10.6 kip from Jockwer et al. (2014) and 11.5 kip from Kevarinmäki (2002)). Apparently, using

theoretical predictions will result a inadequate design, therefore the tested ultimate capacity was

used in the next section when determine the spacing for the screws. In addition to the ultimate

125
capacity, the stiffness values were also extracted from the shear test. Due to the nonlinear response

of the inclined screw joint, the stiffness were categorized based on their displacement regions: initial

stiffness with displacement less than 0.06 in.; linear stiffness with displacement between 0.06 in. and

0.25 in.; and secant stiffness covers the displacement up to failure. These values are tabulated in

Table 6.2.

Figure 6.19: Results for two static shear tests

Table 6.2: Inclined screw stiffness values

Items Stiffness per screw


00
Static initial (< 0.06 ) 13.33 kip/in
Static linear (0.0600 ∼ 0.2500 ) 31.47 kip/in
Static secant 24.33 kip/in

The failure of the two shear tests were consistent: one side of the two screw sheared off at

the connection interface, see Figure 6.20. However, this may not be the true failure mode since the

experimental setup did not allow any separation of the connecting members, only limited sliding. If

the long composite beam were to be subject to uniformly distributed load, the curvature resulted

from deflection and the failure of the screws will cause separation of the members. It is highly likely

that the inclined screw joint will show a more ductile behavior than what is tested in the smalls

scale shear test.

126
Figure 6.20: Screws sheared off in a inclined screw joint

6.4 Design and Assembly of CLT-Glulam Composite Beam

6.4.1 Determination of beam size and design parameters

The objective of this section is to estimate the size of the long span composite beam and its

desired loading. The actual member size and screw spacing are determined in the design presented

in the next section.

The geometry of the design composite beam was largely constrained by the size of the

existing testing facility and size of CLT that can be delivered by semi-tractor trailer. The strong

floor at the Clemsons Wind and Structural Engineering Research Facility was utilized to test the

inclined screw joint. The height of the testing bay is adjustable by relocating the two cross beams

that hold the MTS actuator (Figure 6.18). The 5 ft. by 40 ft. CLT panels were produced by

Structurlam, a CLT manufacturer located in Penticton Canada. The glulam beams (24F-V4-1.7E)

were produced by Anthony Forest Products (Washington, GA). The decision was made at this point

to set the overall geometry of the composite beam at 5 by 40 ft., the beam depth was then determined

based on the loading demand. At this size, both CLT and glulam can be easily transported by a

standard flat-bed trailer.

One of the most common commercial occupancies, Office, was selected as the target design.

For dead load, ASCE 7-10 requires an itemized list to estimate the total dead load. However, a

moderate 20 psf dead load was assumed. For live load, ASCE Chapter 4 provides a minimum

uniformly distributed live load table for all intended occupancies. For office buildings, lobbies and

first floor corridor are designed with the largest live load demand, at 100 psf, followed by corridor

127
above first floor (80 psf) and regular office space (50 psf). The most severe demand of 100 psf live

load was used to design the CLT-glulam composite beam.

6.4.2 Design of the long span composite beam for office occupancy

As the same for design of all structural elements, the design of long-span mass timber

composite beam includes strength limit state and serviceability limit states. The strength limit state

for a floor beam usually includes bending and shear capacity checks. In general, the serviceability

limit states usually govern the design for long span systems, and this is certainly the case for wood.

Serviceability limit states normally include the maximum deflection limits for total load (live +

dead), live load only and long-term load. In fact, it has been shown that the deflection limits alone

may not be enough to design a satisfactory floor system. This is not the only case for wood, but

are generally true for all construction materials such as concrete and steel. Long-span system, in

particular, are prone to vibration from normal machinery and human footstep loadings.

Of course, there are many other important serviceability requirements that worth careful

considerations. Acoustical performance is one of them, which deals with the isolation of airborne

and structure-borne sound. A good acoustical design can greatly improve the comfort level of

the occupants. For example, acoustic performance may affect the employee productivity in office

settings, performance quality in auditoriums, and the market value of apartments, condominiums

and single-family homes. Each occupancy type may have a different comfort level threshold. Unlike

the strength level design, the acceptance criteria of acoustic is subjective; thus, acoustic design has

its own challenges. Another key aspect, which is unique to CLT or other mass timber construction, is

the building enclosure design. Building enclosure design has important implications for the energy

performance and durability of the structure as well as indoor air quality and occupant comfort.

Design considerations for building envelope include water intrusion, control of heat flow, air flow

and moisture flow. The satisfaction of these serviceability requirements is equally important, if not

less important than the safety or strength level designs. However, only the basic structural design

along with deflection checks was performed in this study. Vibration serviceability was addressed

separately in Chapter 7. It is recommended that future research to look into other aspects of the

mass timber serviceability designs.

After a few iterations of trial and error, the cross section of the composite beam was de-

termined based on material availability, structural capacity as well as serviceability deflections, see

128
Figure 6.21. Conventional 3-ply layup for Southern Pine CLT was selected for flange members, with

the lumber of the outer two longitudinal layers of the CLT parallel to the span direction; whereas the

webs are 24F-V4-1.7E glulam beams with stock width 3.125 in. and height 11 in. The designation

of the glulam beam indicates the following stress class and design values: 2400 psi allowable bending

stress, 1.7 · 106 psi modulus of elasticity, ‘V’ stands for visually graded lumbers and 4 is the layup

combination number.

Figure 6.21: Cross section for the CLT-Glulam composite beam

In the following sections, in addition to the determination of inclined screw spacing, the

designs for strength and serviceability limits states are presented. While the calculations are per-

formed assuming full composite action, the actual performance of the long-span composite beam

was examined later during the full-scale destructive test.

Check for bending and vertical shear assuming full composite action Section

properties of the composite beam were calculated based on the equivalent width method, analogous

to the transformed section method of reinforced concrete T-beam design, proposed by Okabe et al.

(2014) where the width of the cross layers are taken as 1/30 of the longitudinal layers. This ratio

is based on the proposed E0 /E90 value in the PRG-320. The MOE value for the longitudinal layers

of CLT with No.2 SYP lumber was taken as 1400 ksi (NDS), whereas the 1700 ksi was used for the

glulam beams. The calculated section properties are summarized in Table 6.3. More details for the

section property calculations and limit state checks may be found in Appendix C.

Note the first moment of area (Q) is calculate at the CLT-glulam interface with the neutral axis

coincides the section centroid. To calculate the LRFD moment and shear carrying capacities, the

129
Table 6.3: Section properties of the composite beam (assume full composite)

Section Property Value


I (moment of inertia) 2.056 · 104 in4
S (section modulus) 2.136 · 103 in3
EI (flexural stiffness) 2.899 · 107 kip · in2
A (area) 3.915 f t2
Q (first moment of area) 1.269 · 103 in3

following two equations were used:

MCapacity = λ φb KF b fb S (6.7)

2
VCapacity = λ φv KF v Fv d twebs (6.8)
3

where K, λ and φ are adjustment factors that can be found in NDS, fb is the allowable bending

stress for SP No.2 lumber, d is the overall depth of the cross section.

On the demand side, the factored shear and moment demand were computed by assuming

uniformly distributed load on a simply supported beam:

Wu L2
MDemand = (6.9)
8

L
VDemand = Wu (6.10)
2

The calculated demands and capacities are tabulated in Table 6.4. It is seen that the beam

has satisfactory capacity in bending and vertical shear.

Table 6.4: Bending and shear capacities of the composite beam

Bending (kip · f t) Bending with FJ (kip · f t) Shear (kip)


LRFD Demand 206.2 206.2 21.1
LRFD Capacity 322.8 224.7 24.3

Note that the contribution of CLT flanges to shear strength was ignored as other studies

have shown that the flanges of I- or T-beams have little influence on the overall shear strength.

130
Observation of Table 6.4 indicates that the member sizes selected for the beam are adequate under

the full composite assumption. However, to fully engage the build-up members for composite action,

appropriate spacing for the self-tapping inclined screws must be determined for horizontal shear

transfer.

Check for deflection assuming full composite action Since the long-span composite

beam is used as a bending member, the expected deflections under the design loads (i.e. without load

factors) should be computed and checked against the serviceability limit states to ensure satisfactory

service conditions. The live load only deflection along with a long-term (creep) deflection were

checked, their limits were taken as ∆LL < L/360 and ∆Long < L/240, respectively. The formula for

calculating long-term deflection is provided in the NDS:

∆T = Kcr ∆LT + ∆ST (6.11)

where ∆T is the total deflection, Kcr is the time dependent deformation factor (taken as 1.5 for

CLT), ∆LT is the immediate deflection due to the long-term component of the design load (typically

dead load or a part of the dead load) and ∆ST is the short-term or normal component of the design

load (typically the live load only). The deflections were calculated at mid-span for a uniformly

loaded beam:

5wL4
∆= (6.12)
384 EI

Table 6.5 summarized the calculated deflections and their respective limits. Under the full

composite assumption, the deflections are all under the limits.

Table 6.5: Live load and long term deflections

Deflection (in) Deflection Limit (in)


Live load 0.90 1.30
Long Term 1.54 1.95

Spacing of the inclined screws and horizontal shear Spacing of the inclined screws

and horizontal shear The horizontal (interface) shear demand is was determined using the equation

131
of shear flow and assumesassuming full composite action of the section:

V (x)Q
τ= (6.13)
It

where V (x) is the shear demand at location x along the beam, Q is the second moment of area at

the connection interface, I is the moment of inertia and t is the beam width at the interface (taken

as 6.25 in. the width of two glulam beams). For a uniformly loaded and simply supported beam, the

shear demand is maximum at the supports and decreases linearly as the distance from the support

increases. Apparently, the spacing of the inclined screws depends on the magnitude of the shear

force. Under the demand of uniformly distributed load, the screws have minimum spacing at the

support and maximum spacing at mid-span.

On the resistance side, due to the lack of established allowable design values and appropriate

adjustment factors for self-tapping inclined screws, the tested ultimate strength of the inclined screw

was taken as the factored LRFD design value. For proper reliability calibration of the connection,

large number of screw shear tests have to be performed. In this study, φ factor was assumed to be

1.0. Inverting Eq. (6.13) and cancel out the interface width term t, the maximum spacing allowed

to transfer the vertical shear V (x) can be written as:

V (x)Q φR
≤ (6.14)
It t spacing(x)

φRI
spacing(x) ≤ (6.15)
V (x)Q

where R is the LRFD resistant for a single screw.

Figure 6.22 shows the maximum allowable spacing of the screws as a function of distance

from the support. Also shown in the figure are the spacings selected for the composite beam. The

adequacy of the horizontal shear design is achieved if the provided spacing falls under the maximum

spacing envelope. For design purpose, three spacings were used for three sections of the beam:


400 o.c., 00 < x ≤ 50






Spacing(x) = 600 o.c., 50 < x ≤ 130 (6.16)




1000 o.c., 130 < x ≤ 19.50

132
Note Eq. (6.16) only shows the spacing for half of the span. Since both the beam geometry and

loading are symmetrical, the placement of the inclined screws is also symmetrical.

Figure 6.22: Maximum screw spacing envelop

6.5 Assembling of the Composite Beam

The previous sections determined the dimensions for CLT and glulam as well as the screw

spacings needed to transfer the horizontal shear. This section covers the assembling process from

beam delivery all the way to the finished specimen. As an old saying goes, ‘a picture is worth a

thousand words’, thus many of the information related to assembling process are conveyed graphically

in this section.

First, a shop drawing that specifies the spacing and orientation of the inclined screws was

prepared to facilitate the assembling process as well as the drilling of pilot holes for the inclined

screws. In section 6.3.2, it has been discussed that the shear tension joint is the most efficient

and easy to design type of timber-to-timber connection. Its improved strength and stiffness come

from the combined utilization of shear and tension strengths of the screws. The easiest way to

configure the inclined screw profile is to consider the beam deflection shape assuming no composite

action. If the CLT and glulam were stacked together without any adhesive or mechanical fasteners,

each layer would deflect and bend with respect to its own neutral axis under bending load. In this

circumstance, the stress state at the joint will be tension above the interface and compression below

the interface. Thus creating a potential for relative movement. The purpose for inclined screw is to

133
prevent relative movement at the interface of members. Figure 6.23 shows the deflection shape of

a non-composite section with the tension and compression side labeled at the connection interface.

Take the top left corner of the beam as an example. The compression side of the glulam web has

a tendency to move towards the mid-span while the tension side of the CLT to move towards the

beam end. Therefore, orienting the screw with its tip pointing towards the mid-span subjects the

fastener in combined shear and tension. Note that for those screws connecting the bottom CLT to

the glulam, the screw tip should be oriented towards the beam end. A shop drawing containing the

spacing and orientation of the inclined screws was produced and it is shown in Figure 6.24 for half

of the span (the other half is mirrored).

Figure 6.23: Exaggerated deflection shape for a non-composite beam

The 3-ply Southern Pine CLT and 24F-V4 glulam beams were delivered separately by flatbed

trailers. Figure 6.25 shows the unloading of a package of four CLT panels using an overhead crane

and spreader beam. These materials were stored under a shed and covered with vinyl tarp prior to

the assembly of the composite beam specimen.

To begin the assembling process of the composite section, two glulam beams were placed

134
Figure 6.24: Shop drawing for inclined screw

on top of the two CLT panels. The top CLT panel shown in Figure 6.26(a) was used as the bottom

flange. The pilot holes were first drilled for the bottom CLT flange. Next, the two glulam beams were

oriented vertically and temporarily secured to the bottom CLT flange with angle brackets (Figure

6.27(a)). The top assembly (glulams and bottom CLT flange in an inverted position) was then lifted

away leaving the unattached top CLT flange in place (Figure 6.27(b)). Finally, two fork lifts were

135
Figure 6.25: CLT unloading from a flatbed trailer

used to lift the top CLT flange and the temporary glulams and bottom CLT flange assembly was

rolled underneath the top CLT flange (see Figure 6.28). At this point, the CLT flanges and glulam

beams were all in position, pending the installation of inclined screws.

(a) Placement of the build-up members (b) Bottom flange pilot drilling

Figure 6.26: Placement of the material for bottom flange pilot drilling

The drilling of pilot holes was also performed on the top CLT flange. With the beam

temporarily secured on saw horses, a total of 168 screws were installed from top and bottom of

136
(a) Temporary erection of glulam beams (b) Top assembly lifted away

Figure 6.27: Temporary erection of glulam beams and lifeted away top assembly

Figure 6.28: CLT panel topping the temporary assembly

the assembly, see Figure 6.29. Recall in the small-scale shear test of the inclined screw, an impact

hammer drill was used and it lacked torque to drive the screws completely in. For the full-scale

specimen, a Milwaukee 18V cordless impact drill rated at 700 lb · f t of torque was sourced and used

during the assembly process (Figure 6.30(a)). The impact motion of the drill allows the screws to

be easily driven into the CLT and glulam. One disadvantage of the cordless impact drill is its harsh

noise during operation, therefore ear protection was worn at all times. During the screw installation,

the inclined angle was constantly checked with a protractor angle finder(Figure 6.30(b)).

Once the assembly of the composite section was completed, the end of each side of the

beam was sandwiched in between a steel W18 wide flange beam and 2x6 lumber to simulate the

bearing contact of CLT wall above the composite beam (Figure 6.31.). The W8 above the lumber

was connected to the W18 through threaded rods and the W18 wide flanges were bolted down to

137
the strong floor. This support condition resulted in a center-to-center span of 39 ft. with 6 inches

of overhang on each side of the beam (see Figure 6.31.).

(a) Install from above (b) Install from below

Figure 6.29: Installation of inclined screw on the composite beam

(a) Cordless impact drill (b) Angle finder

Figure 6.30: Tools used in inclined screw installation

6.6 Full Scale Destructive Tests of CLT-Glulam Composite

Beam

In the section 6.4.2, full composite action was assumed for the design of the composite beam.

To evaluate the composite action and the actual bending performance of the composite beam, a series

of non-destructive and destructive tests were conducted to examine the following parameters:

138
Figure 6.31: Support of the long span composite beam

• Linear apparent stiffness;

• Service level deflections;

• Interface slips;

• Ultimate capacity and failure mode.

6.6.1 Test setup and sensor installation

Test Setup The setup and sensors selected for the test are mainly driven by the intended

purposes, namely, for determining the stiffness, deflections or ultimate strengths for flexure and

shear. A third-point test and a three-point test were used for flexure and shear tests, respectively.

The third-point bending test is typically used for the investigation of bending properties of wood

members, such as ASTM (2013) and ?, since a well-defined constant moment region is desired when

determining the ultimate capacity and failure modes for flexure.

String Potentiometer To track the differential movement between the flanges and webs,

eight string potentiometers were attached to the composite beam, four on each side (Figure 6.32).

These eight string pots occupied only half span. It was assumed that the beam would deform

symmetrically under load; hence, only half of the span was monitored. On each side of the beam,

(1) two spring pots were mounted on the top wide surface of the glulam to measure the slip between

top flange and web, and (2) two spring pots were mounted to the bottom of the glulam to measure

the slip between bottom flange and web. Note that two string pots (top and bottom of the glulam)

were installed in Zone I of the beam and two spring pots in Zone II (see Figure 6.32 left). Zones I

139
and II correspond to regions where the screws were spaced at 4 in. o.c. and 6 in. o.c., respectively.

For tracking the vertical deflection of the beam, a long stroke string pot was utilized to monitor

vertical displacement at the bottom center of the specimen.

Figure 6.32: Attachment of string potentiometers on the composite beam

Data Acquisition System The complete data acquisition (DAQ) contains two stand-

alone systems: MTS Flextest software suite that controls and records the movement of the hydraulic

actuator and three USB-1208FS units from Micros Measurements that records the string pots. The

string pots were operated with a constant 15V power supply and the string pots output signals

(0-10V) were recorded by two USB-1208FS devices. A third USB-1208FS device was used to send

custom voltage signal (0-5V) to the MTS systems for the purpose of data synchronization.

Figure 6.33: USB-1208FS Data acquisition aystem

140
6.6.2 Bending test–1

With the experimental setup and data acquisition system ready, the first bending test ap-

plied a third-point loading. The loading beam was fabricated with a 20 ft. long W18 with two

8x8x0.5 HSS oriented perpendicular to the W18 and welded to the bottom of the W18 section as

the loading points (Figure 6.34). The two HSS loading bars were spaced at 1/3 of the span: 13.3 ft.

Figure 6.34: Third-point bending setup

A general loading procedure for bending as well the shear test performed later was separated

into two steps. First, the load controlled procedure (500 lbf/min) was utilized to evaluate perfor-

mances of the composite beam under dead load only, service load and factored ultimate load. No

failure shall be observed during the load controlled procedure if the beam was designed adequately.

Following the load controlled procedure, the displacement controlled procedure (0.25 in./min) was

then utilized to determine the ultimate strengths (flexure and shear). For destructive test, the dis-

placement controlled procedure is a safer procedure to operate because a sudden loss of strength in

the test specimen would not cause abrupt actuator cylinder movement. In contrast, the force con-

trolled procedure could result in a sudden and large movement of the actuator if a brittle failure was

to occur. For the safety of the experimental study, two levels of displacement-triggered interlocks

and one force-triggered interlock (at 120 kip) were configured in the MTS Flextest software to avoid

unexpected and excessive movement of the actuator.

Load controlled procedure Figure 6.35 shows the moment and shear diagrams of a

fixed-fixed beam subject to uniformly distributed loading (a) and third point loading (b). The load

increments (in terms of 2P) in the load controlled procedure were determined such that the moment
PL
within the constant moment region in the third point loaded beam 9 equals to that of uniformly

141
W L2
loaded beam in the mid-span 12 . Table 6.6 shows the five target applied loads (actuator forces)

determined based on the design dead load (20 psf) and live load (100 psf) for bending or flexture

test. The moments shown in Table 6.6 contains dead load only, service load, factored load and two

additional increments in between. For each load level, the actuator load was ramp from zero to the

target load. Upon reaching the target load, the load was held for 120s to examine and document any

failure mechanisms that occurred. The load was then release to zero and the process was repeated

for the next load level.

Figure 6.35: Moment and shear diagrams for bending test 1

Table 6.6: Mid-span deflection and stiffness of the 1st bending test

Load Deflection Secant Stiffness Moment Note


(kip) (in) (kip/in) (kip · f t.∗ )
6 0.29 15.00 13.0 Dead Load Only
12 0.72 14.76 26.0 --
24 1.67 13.57 52.0 --
36 2.60 13.76 78.0 Service Load
54 4.02 13.68 117.0 Factored Load
76 5.90 13.33 164.7 Displacement Controlled

*The moment is calculated based on fixed-fixed B.C.

Displacement controlled procedure Followed by the load controlled procedure, a dis-

placement controlled procedure was applied to the specimen to bring it to the strength limit. During

the displacement controlled procedure, the actuator traveled at a constant rate of 0.25 in./min and

this rate was used in the subsequent displacement controlled procedure. The test setup remained

the same, and the mid-span deflection and interface slips were monitored through the same string

142
pots.

Figure 6.36 shows the load versus mid-span displacement curve obtained from the bending

test. Note that the mid-span displacement was obtained from the string pot installed underneath

the specimen rather than the MTS actuator internal LVDT reading. In fact, under a third-point

loading setup, the mid-span deflection is expected to be slightly larger than that of the MTS LVDT,

due in part to bearing deformation of the CLT at the supports. Since the use of actuator LVDT

may not be suitable since it does not measure the true deflection of the beam, the deflection and

secant stiffness tabulated in Table 6.6 were all computed from mid-span string pot readings.

Figure 6.36: Load displacement curve for the 1st bending test

Observations in bending test–1 During the displacement controlled test, the W18 wide

flange loading beam experienced top flange local buckling near the connection between the steel beam

and the actuator, although steel was stiffened at all loading points. The test was terminated at 76

kip (Figure 6.36 dotted line).

A residual deflection of 0.35 in. was carried into the displacement controlled loading. Figure

6.37 shows a picture of the test. As seen from Figure 6.36, the beam deflected linearly with some

loss of stiffness at 76 kip when the steel loading beam failed. The composite CLT-glulam beam

showed no observable strength degradation and survived almost 140% of the factored design load.

Occasional popping sounds were heard during test, when the load was approximately above 60 kip).

143
It was later discovered sounds came from partial withdrawal of some screws from the glulam beams.

Figure 6.37: Deformed beam during a displacement controlled loading procedure

Though the beam performed well capacity wise under service and factored load, it is of great

interest to examine the efficiency of the inclined screw connection at the CLT-glulam interfaces, as

the differential movement between the members may indicate loss of composite action. Figure 6.38

plots the interface slips on top of the beam response under displacement controlled loading. The

displacements from eight string pots (SP) were presented separately according to their locations:

the slips for the four string pots located in Zone II (see Figure 6.32)are shown in the left of Figure

6.38 while the slips for the other four string pots located in Zone I are shown in the right of Figure

6.38. The numbering of the string pots can be found in Figure 6.32. The slips were very minimum

and barely noticeable to human eyes. The slip at 76 kip load was 0.07 in., which is roughly the

thickness of a gauge 14 sheet metal. The differential movements can be attributed to many reasons,

including re-sitting and extension of the screw, screw withdraw from wood members, extension and

contraction of the wood member at interface, etc.

One of the beam’s under-performance was the deflection at service load (2.6”, 36 kip). The

measured displacement at service load was equal to the L/180 limit (2.6), which was below the

serviceability requirement of L/240 (1.95) set forth in IBC for floor beam. However, the assump-

144
Figure 6.38: Interface slippages during the displacement controlled loading–1st bending

tion of boundary condition heavily influences the interpretation of the applied load and associated

deflections (see Table 6.7). Under the pinned boundary condition assumption, the beam’s deflection

was acceptable at 1.61 in. However, the actual boundary condition of the test was neither fixed or

pinned. The clamping fixture used in the test was thought to result in a partially fixed boundary

condition. It is concluded from the results of the first bending test that stiffness of the beam was

under-designed, albeit it maintained high level of composite action during the test. The CLT and

glulam member sizes shall be adjusted to satisfy the deflection requirements.

Table 6.7: Deflection with fixed and pin boundary condition assumptions

Fixed B.C. Deflection (in.) Pinned B.C. Deflection (in.)


Load (kip) Assume Fixed Load (kip) Assume Pinned
Dead only 6 0.29 6 0.29
Service 36 2.6 18 1.61
Factored 54 4.02 27 1.90

L/360=1.3”, L/240=1.95”, L/180=2.6”

Stiffness reductions The deflections shown in Table 6.6 indicate that the beam’s com-

posite action was compromised at higher loads and resulted in a reduced flexural stiffness. The true

stiffness and its reductions are valuable parameters that may be used to improve the current design

and perhaps provide insight for future ribbed floor system design. However, the interpretation of

145
the beam flexural stiffness heavily depends on the assumption of support condition. Consider two

sets of support conditions shown in Figure 6.39, the left with pinned-pinned boundary condition

and the right with fixed-fixed. For a third-point loading, a is one third of the span (a = 1/3L). The

flexural stiffness can be obtained using the following equations:

23P L3
EIpinned = (6.17)
648∆
5P L3
EIf ixed = (6.18)
648∆

Figure 6.39: Boundary conditions for a third-point bending setup

where ∆ is the mid-span deflection. It it also worth noting that shear deformation is ignored due to

the long span length and the relative small contribution of shear deformation observed in Chapter

5.

Using deflections measured during the bending test, the flexural stiffness can be calculated

with the assumed boundary conditions. Figure 6.40 visualizes this stiffness reduction at different

load levels and the theoretical flexural stiffness calculated assuming full composite action using

equivalent width method is also shown in the figure. Regardless of the assumed boundary condition,

the stiffness reduction was consistent since it is a function of only load and associated deflection.

Compare to the dead load case, the stiffness reduction was 33% at service loading (36 kip) and 35%

at factored loading (54 kip). The reduction at service load was substantial: almost 1/3 of the initial

stiffness. Based on the test observations, the reduction in stiffness has the most contribution from

withdraw and flexibility of the inclined screws, despite the fact that the deformation in the interface

was very limited.

146
Figure 6.40: Stiffness reduction in bending test

6.6.3 Bending test–2

The purpose of the second bending test was to examine the bending strength of the CLT-

glulam composite beam. As mentioned in a previous section, the steel loading beam experienced

local buckling during Bending Test 1. In Bending Test 2, a concrete infilled 8x8x1/2 HSS was used

along with a three point bending test (i.e. only one loading bar at the mid-span). The 5 ft. long

HSS was fully filled with 3000 psi Quikrete ready mix concrete. The three point flexural test was

performed two weeks after casting of the concrete into the HSS section (see Figure 6.41 for test

setup), allowing the concrete to reach at least 75% of its compressive strength.

The three-point bending test followed a similar combination of load and displacement con-

trolled procedures, with the load applied in the load controlled procedure modified based on the

new loading diagram. Figure 6.42 shows a set of new moment diagrams used for computing the load

increments in the load controlled procedure. Note that the load which caused the failure of the W18

loading beam (76 kip) was also converted in the center point loading procedure. Table 6.8 shows

the four load increments used.

Figure 6.43 shows the beam’s load displacement response under three-point loading, the

specimen eventually failed in the displacement controlled procedure at 61 kip. The failure load at

147
Figure 6.41: Three point flexural test setup

Figure 6.42: Moment and shear diagrams for bending test 2

148
Table 6.8: Load increments used in 2nd bending test

Load Level (kip) P (kip)


Dead Only (20 psf) 3
Service (120 psf) 16
Factored (1.2D+1.6L) 24
76 kip 32

61 kip resulted a over-strength factor of 2.54 and it equivalent to around 300 psf life load capacity at

failure. A noticeable difference in the 2rd load controlled procedure is that the beam clearly shows

creep effect at holding loads (i.e. 3 kip, 16 kip, 24 kip and 32 kip). This could indicate damage in

the system, including both the wood member and the screw connection. Wood typically does not

show creep effect under a very short duration of sustained load (2 mins). Therefore, the observed

creep most probably came from the re-withdrawal of the screws. Rotation of the beam end can be

seen in Figure 6.44, where it shows the beam is lifted from the support.

Figure 6.43: Load displacement curve for the 2nd bending test

The inclined screws still held well in the 2rd bending test. Figure 6.45 shows the limited

interface slips from eight string pots up to the beam’s three-point bending capacity at 61 kip. This

near perfect composite action (deformation wise) has ‘forced’ the beam to fail in the wood members

149
Figure 6.44: Rotation in the beam support

and the failure mode is discussed in the next few paragraphs.

Figure 6.45: Interface slippage during the displacement controlled loading–2rd bending

150
Observations in bending test–2 In Chapter 5, small scale bending test was performed

on the strip-like 1x12 ft. specimens. It was learned that the majority of the bending failures can

be attributed to the finger joints and lumber defects. Vertical finger joints were present in the CLT

panels. The vertical finger geometry and the lumber defects often cause stress concentrations and

those eventually lead to the beam pre-mature failure. The failure mechanism of the composite beam

was similar to that observed in small-scale bending tests, where tension rupture of the wood fibers

in the tension face of the CLT flange. For the composite beam, additional failures were observed in

the glulam beams.

At 41 kip, the beam experienced a loud ‘hiccup’ accompanied with a small load decrease

(around 1 kip). It was decided to pause the loading procedure for inspection. However, nothing was

found consider the crack may hide inside the hollow core. The test was resumed. At around 52 kip,

the specimen started to give warning signals of its failure: continuous wood fiber snapping sound

from beam’s tension side and small wood chunks falling off from underneath. The beam eventually

failed at 61 kip with a loud ‘explosion’ sound. The failure mechanisms can be categorized as the

following:

1. The most majority of the failure took place at the tension side mid span. After the destructive

bending test, the middle 10 ft. section was cut off for thorough inspection, it was found that

the tension side of the CLT panel contained many finger joints and lumber defects, whereas

the compression flange was intact. Figure 6.46 shows the tension side failures;

2. One of the finger joint failures passed through the thickness of the tension CLT and caused

delamination and rolling shear failure, see Figure 6.47;

3. The glulam beams also experienced substantial rolling shear and bending failures (Figure 6.48).

From small-scale timber specimen tests, it was observed that the CLT panels often failed

in a brittle manner with little to no sign of partial failure prior to the ultimate failure. However,

this is not the case for the tested full-scale composite beam. Although not ductile, the composite

timber beam gave loud noise prior to reaching its strength limit. The failure can be identified with

wood fiber snapping sound and falling objects. This can be advantageous to the building occupants

as they gain additional time to escape prior to any catastrophic failure.

151
Figure 6.46: CLT tension side failures. Red circle: finger joint failure; Rectangle: lumber knot
failure; Green rhombus: survived finger joint

Figure 6.47: Tension CLT delamination and rolling shear failure

152
Figure 6.48: Glulam rolling shear and bending failure

6.6.4 Shear test

The same composite beam specimen tested for bending was also utilize for shear test. In

the bending test, the peak moment demand occurred at the mid-span. The segments near the ends

of the beams or supports were largely undamaged. Thus, the beam segments near the supports were

utilized for shear test.

After the destructive bending test, the composite timber beam was cut into three sections:

13 ft. and 17 ft. from the ends and 10 ft. from the middle section. A 16.75 in. deep circular saw

and a reciprocating saw, both with carbide blade, were used to cut the beam. Prior to cutting the

composite beam, screws near the cutting locations were removed. However, some of the damaged

screws were not able to be backed out. Nevertheless, the circular saw cut was able to cut through

these steel screws (see Figure 6.49).

The 13 ft. section was used for the shear test and the test setup can be found in Figure 6.50.

The 13 ft. section was clamped at the ends with a point load applied at 5 ft. from the left support

(Figure 6.50), with 6 in. of overhang at each end of the beam. The point load was applied right at

the boundary of two inclined screw spacing zones: Zone-I with 4 in. o.c. and Zone-II with 6 in. o.c.

The same string pots were used in the shear test specimen. The string pots were instrumented at 1

ft. from the center of the supports.

One of the key steps before performing the test was to estimate the short beam’s capacity,

in both shear and bending. Short beam has a much greater bending capacity due to the shortened

moment arm and the vertical and horizontal shear capacity have to be carefully considered since

153
Figure 6.49: Cutting the CLT-glulam beam for shear test

Figure 6.50: Shear test setup

154
the testing capability is heavily bonded by the capacity of actuator and load frame. Table 6.9

summarizes the estimated support reactions at different failure scenarios and the corresponding

actuator forces P . The forces associated with different failure scenarios were estimated based on the

following assumptions:

1. Bending Assume that the 13 ft. span specimen has the same moment carrying

capacity as the 40 ft one, as obtained in bending test 2.

2. Vertical Shear Assume two glulam beams fail in shear.

3. Horizontal Shear Based on the capacity of a single inclined screw, spacing and

the shear flow equation (Eq. (6.13)).

Table 6.9 indicates that the most likely failure mode would be horizontal shear failure (i.e.

failure of the inclined screws). Since Zone II (with screws spaced at 6” o.c.) had a lower horizontal

shear capacity compared to Zone I, it was anticipated that failure would most likely initiated in Zone

II. The estimated actuator loads for horizontal shear failures are within the capacity of the actuator

(150 kip).

A load controlled procedure was applied first to examine the performance of the beam under

service loads through in a series of load increments. These increments were calculated based on the

reactions computed using the design loads (20 psf dead and 100 psf live) for a 39 ft. span and then

converted into the force needed in the actuator to achieve the same reaction force at the left support.

The moment and shear diagrams for the asymmetrical shear test is shown in Figure 6.51. Table 6.10

shows target actuator loads for dead load only, service load and factored design load. In addition,

a 150% factored load was added to the loading procedure to observe the serviceability potential for

the combination of 20 psf dead and 150 psf live load. Note the actuator forces P shown in Table 6.10

were rounded up to the nearest kip. Followed by the load increments and unloading, a displacement

controlled procedure was applied to determine the ultimate shear capacity. The short beam was

loaded to ultimate failure in a displacement controlled procedure at 0.25 in./min. All the interface

string pots were removed prior to the ultimate destructive shear test in concern that sudden zipping

of the screw lines may damage the sensors.

Figure 6.51 shows force versus displacement response of the shear test. During the load con-

trolled procedure, the short shear specimen deformed linear elastically with very minimum residual

155
Figure 6.51: Moment and shear diagrams of the shear test

Table 6.9: Shear Test Failure Estimations

Failure Mode Shear (5’ Support) (kip) Shear (7’ Support) (kip) P (kip)
Bending 119.0 85.0 203.7
Vertical Shear 60.6 60.6 104.0
Horizontal Shear (Zone-I) 55.6 39.7 95.4
Horizontal Shear (Zone-II) 51.9 37.1 89.0

deflection after unloading. Meanwhile, there was no wood damage or screw failure observed up to

150% of the factored design load. This observation is collaborated by the very low interface slips

shown in Figure 6.53 (peak slip at 0.05 in.). It can be seen that the 5 ft. section experienced slightly

larger interface slips (Figure 6.53 left) than that of the 7ft. section (Figure 6.53 right). This is

expected since the shorter section received more shear demand in an asymmetrically loading. Based

on the results of the two connection tests of the inclined screws, the slips at ultimate failure for the

two tests occurred at approximately 0.29 and 0.38 inches, which were significantly higher than the

peak slip of 0.05 in. observed in the shear test with 150% of the factored design load (see Figure

6.19). In summary, the test results show that the composite beam has adequate shear capacity at

service load with reserved capacity to safely carry the shear demand for more than 150% of the

factored design load or more than double of the service load. This is expected, as shear typically is

not the controlling design factor for long span beams.

156
Table 6.10: Load increments used in the shear test

Load Level Reaction V (kip) P (kip)


Dead Only 2.0 4
Service 11.7 20
Factored Load 17.4 30
150% Factored Load 26.1 45

Figure 6.52: Load displacement curve for the shear test

Observations in shear test During the displacement controlled test, the specimen safely

passed the 150% factored design load (45 kip) and the first failure occurred at 79.6 kip (actuator

force). The failure took place in one of the bottom CLT flange corners of the 7 ft. section. This failure

started with delamination in the CLT glue line, which extended towards the support, accompanied

with minor rolling shear failures (Figure 6.54(a)). At support, the delamination failure passed

through the bottom CLT layer and contained in the width of a 2x lumber (Figure 6.54(b)) . This

delamination in the CLT created 45◦ angle rolling shear cracks, very similar to the shear failure

cracks typically observed in a reinforced concrete beam. However, in the composite beam, the

damage did not propagate upward into the glulam webs. The shear-induced cracks were contained

in the bottom tension flange. Right after the delamination failure, the load decreased from 79.6 kip

to 25 kip due to the partial loss of a tension layer.

157
Figure 6.53: Interface slippage during the load controlled shear test

The destructive shear test was extended beyond the first failure to investigate the reserved

capacity of the beam. Figure 6.52 shows that the beam had reserved capacity to carry load beyond

the first failure load (79.6 kip). When the load was increased to above 80 kip, the screws started

to withdraw from the glulam beams, making continuous popping sound. The zig-zag load pattern

at the tail of the load-displacement curve (Figure ShearLoadDisplacementResponse) indicates the

gradual withdrawal of a series of screws from the glulam beams. At 108.6 kip, one of the glulam

beams in the 7 ft. section failed in the interface connections, likely due to loss of composite action

from the screw failures. The applied load was quickly shared by the other glulam beam, resulted

in unbalance load distribution and bowing of the top flange. Due to safety concern, the test was

terminated at this point. Consider the shear force at 5 ft. section, the failure load resulted a over-

strength factor of 3.62. The ultimate shear capacity is equivalent to 400 psf live load on a 40 ft. 5

ft. wide floor beam.

During the after-test inspection, the residual slips (after removing the load) of the glulam

beams relative to the CLT were measured. The 4” o.c. screws in the 5 ft. section show no trace of

movement, while majority of the damage was concentrated on the screw lines in the 7 ft. section,

where the slips were measured at 1/4 in. at both sides, see Figure 6.55. Due to the withdrawal of

screws, separations between webs and flanges were observable at the interface and the gap varied

from location to location with the maximum gap measured around 3/8 in (Figure 6.56). No other

wood failure was observed for the shear test.

158
(a) Delamination in the tension CLT flange (b) Delamination extended through the 2x at support

Figure 6.54: Delamination failure at 79.6 kip

6.7 Conclusion and Future Research

With the growing need for long span solutions in the building sector, this chapter explored

the possibility of developing a structurally efficient mass timber long span floor element. Mechanical

connectors, namely, inclined screws were considered in this study. Theoretical predictions as well

as small-scale screw joint shear tests were performed to deepen the understanding of such timber-

to-timber connection. With the information in hand, a full-scale 40 ft. long span composite beam

was designed and fabricated using CLT and glulam beams connected by inclined screws fastened

at 30 ◦ . The beam was further subjected to destructive flexural and shear tests to confirm its load

serviceability as well as ultimate strength. The beam performed well with flying colors in terms with

structural capacity, however, its flexural rigidity may be under-designed.

Lesson learned is that serviceability is equally important as structural design as it impacts

the building occupants’ comfort on a daily basis. An excessive deflection in the floor system may

cause crack in the gypsum fire protection board, misalignment of mechanical and electrical pipes,

etc. In fact, deflection is only one aspect of the serviceability design, an occupant-friendly floor

system consider other serviceability requirements such as vibration and acoustics.

The proposed long span composite beam solution can be packaged as a preassembled unit

with integrated MEP systems and thus enables fast plug-n-play construction in the job site. This

further accelerates the mass timber construction from element level to assembly level. Of course,

the logistics of the material and planning will be the key behind the scene. Two items were not

159
(a) Interface slip at 5 ft. section near support (b) Interface slip at 7 ft. section near support

Figure 6.55: Interface slippages after the shear test

considered in the presented study and this could be viewed as the proposed future research:

1. The connection between long span composite elements. In this regard, traditional

lap/butt joint or perforated plate can be use with short wood screws to fasten the elements

together. The potential of two way slab can be further explored.

2. Cost benefit analysis. The presented study has chased more of a feasible long

span solution than the cost behind it. The integrated benefit must be carefully balanced when

assessing the construction feasibility of such element because the structural cost could seem

prohibitive.

160
Figure 6.56: Opened gap at connection interface

161
Chapter 7

Vibration Properties of Composite

CLT-Glulam Beam

7.1 Introduction

Wood light-frame construction was the first building system unique to early America. De-

veloped in the first half of the 19th century, light-frame construction became the standard for U.S.

suburban housing shortly after World War II. Freed from heavy timbers of the post-and-beam system,

wood light-frame offers ease of construction. Despite its increasing use of timber for single-family

and multi-family house, the user experience has however not always been fully satisfying (Gliso-

vic and Stevanovic 2010). Particularly timber floors in multi-family houses are very susceptible to

annoying vibration cause by human activities. Woeste and Dolan (2007) has pointed out that the

annoying vibration is most probable when the floors are design to the building code minimum but

various steps can be taken to prevent the likelihood of such vibration, probably with minimal added

cost. The vibration response is also dependent on the damping properties of the structures. It has

been shown that damping is a key factor for the human perception of the vibration response, and

thus annoyance (Lenzen 1966).

The out-of-plane low amplitude motions resulted from surface impacts (eg. foot falls) are

not uncommon in timber buildings and are commonly designed against. In North America, dy-

namic response for timber floors are mostly associated with residential and mercantile occupancies

162
like offices, and joisted floor constructions having flexible subfloor layers and supported on timber

stud walls. Typically, vibration serviceability design of such floors presumes that dynamic motions

building occupants experience are dominated by flexibility and masses associated with particular

building occupancies. Associated technical measures are to stringently limit static deflections of

joist under concentrated load (assume direct relationship between static and dynamic response); or

empirically relating parameters like estimated fundamental frequencies and static deflection with

particular floor systems. Both approaches are simple, but fail to directly address dynamic response

observed by building occupants. For light frame wood construction, composite deck framed with

I-joist beams and girders, sheathed with OSB or plywood have been designed and constructed for

decades. This type of floor system is easy to build and is cost effective. However, the span of timber

I-joist is limited to around 20 to 30 feet, often controlled by deflection and vibration issues. In fact,

the recent popularity of open space floor plans have led to specifications of larger bay sizes and

longer spans in floor system. In such cases, the traditional wood joist type of floor is barely able to

accommodate the load that come with the increasing span.

Among the long span solutions, Cross-Laminated Timber (CLT) panel has been favored as

primary load bearing members due to its dimensional strength. The range of applications of CLT

as a construction material in North America is far from fully developed but it is clear that some

applications will seek to employ this material in increasing structural-demanding applications. To use

the material more efficiently, composite floor decks can be proposed in large variety of constructions

and occupancies, such as administration and commercial buildings, hotels and bridges. Due to the

increased floor mass and longer span, floor vibrations have become an area of concern. Floor decks

with low frequencies may be resonance with the vibrations due to human activities and the resulting

acceleration may exceed human comfort levels. Therefore, the design of composite floor structures

are often limited by serviceability criteria such as deflection limits and vibration behavior, rather

than strength criteria.

Building occupants’ discomfort resulted from a poor vibrational serviceability may related

to annoyance, fatigue, discomfort and pain (Griffin 2012). If the discomfort of the affected personnel

becomes overwhelming, the occupants may opt for extreme measures such as litigation or relocation,

and consequently the building would suffer diminished commercial value or even decommissioned

in rare cases. While most floor vibration problems are not this extreme, the scenarios highlight

the potential significance of this serviceability problem. The efforts required to adjust an annoying

163
floor vibration problem after the building is commissioned can be very difficult and expensive to

implement. In most cases, the costs of fixing the problem in-situ are much greater than tackling

the problem in the design phase prior to the floors construction, which doesnt even include the

potential liability cost, such as legal expenses, loss of rental, or consulting fees (Hanagan 2005). The

easiest ‘fix’ for a floor vibration problem is avoidance: simply design a floor such that its unlikely

to experience a vibration serviceability problem when put into service. The key to avoidance is the

ability to predict the dynamic behavior of the floor in the design stage and to determine if it will be

a problem under service conditions, a task much easier said than done, and the subject of much of

the floor vibration research to date.

To investigate the vibration characteristics and possible remedies for annoyance/discomfort,

this chapter is organized as follows: first a literature review is performed to understand the human

perception of vibration discomfort and available design criteria. Topics related to CLT vibration

will be covered as well. Second, given the potential vibration discomfort in long span timber system,

the need for vibration modal test and expected research outcome are justified. Following the mo-

tivations, basic structural dynamics and modal testing techniques will be reviewed for the purpose

of conducting experimental modal analysis on the composite beam assembled in Chapter 6. Third,

a heel-drop modal analysis will be performed and modal information will be extracted for analysis.

Finally, a finite element model calibrated based on the experimental data will be presented.

7.2 Motivation and Scope of Research

To note is that limited investigation of behaviors of CLT panel have been undertaken and

have yielded preliminary understanding of vibration characteristics of such material (Hu and Gagnon

2011; Fitz 2008; Weckendorf and Smith 2012; Weckendorf et al. 2016). However, all these investiga-

tions were focused solely on non-composite CLT panels. With the demand for longer span length,

direct application of thicker CLT panel will result in less economic use of material, introducing

redundant seismic weight that burdens the lateral system and foundation size.

The proposed CLT-Glulam composite hollow core concept provides a viable solution for the

demand of long span load bearing structural element. There is no doubt that the composite beam

assembled by engineered wood product will satisfy the stiffness and strength requirement. However,

information regarding vibration serviceability of full scale composite timber beam is scarce and much

164
needed for final concept delivery. Associated technical measures are to stringently limit static de-

flections under concentrated load (assume direct relationship between static and dynamic response);

or empirically relating parameters like estimated fundamental frequencies and static deflection with

particular floor systems. Both approaches are simple, but fail to directly address dynamic response

observed by building occupants. Vibration evaluation can be used to assess the serviceability of the

floor system (typical activities: identification of natural frequencies and damping ratios). Mean-

while, full size modal testing of the floors is able to provide additional information to the dynamic

behavior and possible remedies.

The purpose of this chapter is to examine the dynamic behavior of grounded CLT-glulam

composite beam using experimental modal analysis techniques. The testing results will be extracted,

analyzed and finally, human comfort and testing floor’s vibration serviceability will be quantified.

The desired goals of the research are as follows:

1. Conduct literature survey on human perceptibility to vibrational motions and

vibration serviceability design of timber floors. This will include a initial assessment of floor

vibration serviceability based on hand calculations;

2. Identify and summarize best practice in dynamic testing of floor systems, including

equipment selection, modal testing theories, experimental techniques, and signal processing.

Also included in this goal is the design and fabrication of in-house force plate, a device that

measures the impact of foot heel drop on the specimen;

3. Perform experimental modal testing on the grounded composite CLT-Glulam

beam using the heel-drop excitation. The global modal properties (natural frequencies, damp-

ing ratios, mode shapes) of the floor system will be identified;

4. Construct finite element model and validate using the modal properties obtained

in the experimental study. The FE model is then used to perform additional parametric

studies;

5. Evaluate the vibration design criteria and provide additional comments on indus-

trial construction practice.

165
7.3 Literature Review and Background

Annoying floor vibrations are common in many types of building structures. Problems of this

nature have been reported in the office building, shopping malls, airport concourse, and restaurants,

among others (Hanagan and Murray 1997). Although floor vibrations can result from many sources

(e.g. reciprocating machinery, boiler, heavy truck traffic), the most common and problematic are

caused by the occupants themselves. Occupants generate foot-fall impacts from activities such as

walking, dancing, jumping, etc. Such floors are particularly problematic because they cannot be

easily isolated from the structure and they occur so often. The assessment ‘annoying’ is in general,

determined by the occupancy requirements and human perception of floor vibrations.

7.3.1 Human perception of vibration in floors

Vibration perception is subjective–some people feel it annoying while other people might

find it barely perceptible. Murray et al. (2003a) indicated that human response to floor motion is a

very complex phenomenon, involving the magnitude of the motion, the environment surrounding the

sensor, and the human sensor. For example, a large amplitude continuous motion is more annoying

than motion caused by infrequent low magnitude impact. The threshold of discomfort perception of

floor motion in a busy commercial work place can be higher than in a private home. The reaction of

a senior citizen living on the fifteenth floor can be considerably different from that of a young adult

living on the second floor of an apartment complex, if both are subjected to the same motion. The

way a person experience a floor vibration can be divided into two main categories distinguishing

between vibrations that are felt through the balance organs and through sight or hearing:

1. The floor vibration makes people on the floor vibrate and they experience discom-

fort (Ohlsson 1988);

2. The floor vibration causes loose objects on the floor or attached to ceiling or wall

to rattle, shake or swing, or making the structures squeak, which causes discomfort (Smith

2003).

Nonetheless, many attempts have been made to statistically determine what properties

should be fulfilled for an acceptable floor (Sjökvist and Brunskog 2013; Weckendorf 2009). Their

research is not unambiguous but certain conclusions can still be drawn.

166
Perhaps the most frequently cited reference for human perception of vibration is by Reiher

and Meister (1946). The Reiher-Meister scale is based on a displacement range of 0.01-10 mm

and frequency range of 1-100 Hz. The modified Reiher-Meister scale shown in Figure 7.1 was

proposed by Lenzen (1966) for vibrations due to walking impact. For floors with less than 5% critical

damping, Lenzen suggested the original scale be applied if the displacement is increased by a factor

of ten. Wiss and Parmelee (1974) suggested that a constant product of frequency and displacement

existed for a given combination of human response and damping. Allen and Rainer (1976) developed

vibration criteria in terms of acceleration and damping intended for quiet human occupancies such

as residential buildings and offices. As damping increases, the steady-state response due to walking

becomes a series of transient responses; resulting in a less significant response. Murray (1979)

suggested a human perception scale for required damping as a function of the product of initial

displacement and frequency, which are the same parameters used in the Wiss-Parmelee scale. Allen

et al. (1985) suggested a design procedure for assembly floors subjected to rhythmic activities such

as dancing and exercises. The International Standards Organization (ISO) (ISO 1989) recommends

vibration limits in terms of acceleration root-mean-squared (rms) and frequency. As shown in Figure

7.2, a baseline curve is used by ISO and different multipliers are used for different occupancies.

Various design guides and codes around the world address vibration serviceability in the

same manner: outline simplified hand calculation methods for estimating the dynamic properties of

a floor system, estimate an applied dynamic load that simulates walking forces, and finally compute

the dynamic response of the system of the applied load for comparison with an established level

of human comfort acceptability. The following section describe the present design criteria for floor

vibration.

7.3.2 Present and Historical Development of Vibration Design Criteria

United States The majority of the floor vibration assessment and design criteria devel-

oped in U.S. is based on the ISO criteria, which essentially imposing limits on both acceleration and

frequency. Among all, the guidelines published by the American Institute of Steel Construction:

Floor vibrations due to human activity, Design Guide 11 (Murray et al. 2003b) is one of the widely

accepted and referenced manuals. It is applicable to a variety to building materials since it provides

simple hand calculation on the key components of the vibration design: the estimation of fundamen-

tal frequency, handling of the composite action, approximation of the harmonic force and evaluation

167
Figure 7.1: Lenzen’s modified Reiher-Meister scale. (Lenzen 1966)

of the peak acceleration. The criteria are provided both for walking and rhythmic excitation.

Each vibration serviceability evaluation involves two sets of calculations: one to compute

the fundamental frequency of the floor element and another to compute the maximum amplitude

of acceleration. The floor is deemed satisfactory if the calculated performance point falls below the

dashed line in Figure 7.2. ISO Standard suggests limits in terms of RMS acceleration as a multiple of

the base line curve (the bottom line in Figure 7.2). The multipliers are 10 for offices, 30 for shopping

malls and indoor footbridge, and 100 for outdoor footbridge. Evaluation results are then compared

with Figure 7.2 to determine if the serviceability design requirement is met. The following equation

represents the comparison:

ap P0 e(−0.35fn ) a0
= ≤ (7.1)
g Wβ g

168
Figure 7.2: Recommended peak acceleration for human comfort for vibrations due to human activ-
ities (Murray et al. 2003b; ISO 1989). The root mean square acceleration is the square root of the
area under the acceleration spectral density (ASD) curve in the frequency domain

where a0 /g is the human tolerance level of peak acceleration (as fraction of gravity); ap /g is the

estimated peak acceleration due to walking excitation (as fraction of gravity); P0 is a constant force

(65lb. for floors and 92lb. for foot bridges); fn is the fundamental natural frequency of the system;

β is the modal damping ratio and W is the effective weight of the system, including self-weight, dead

load and live load, without any load combinations. The numerator P0 e−0.35fn represents an effective

harmonic force due to walking which results in resonance response at the natural floor frequency fn .

The right side of the inequality (7.1) represents the human tolerance level of peak accelera-

tion, which can be easily obtained from Figure 7.2 upon selection of occupancy type. The left side

of the inequality computes the peak acceleration response assuming a spring-mass-damper system

driven at its natural frequency by an effective harmonic force due to walking, resulting in a resonance

169
response. Here, only the lowest resonance response (1st harmonic) is considered under the assump-

tion the response due to other harmonics of the step frequency is small in comparison. As Murray

et al. (2003b) observed, motions due to quasi-static deflection and footstep impulse vibration can

become more critical than resonance if the fundamental frequency of a floor is greater than about

8Hz, a more stringent acceleration limit and static deflection stiffness can be imposed–horizontally

extend the acceleration tolerance limit beyond 9.5Hz and a minimum stiffness of 1kN/mm under

concentrated load.

DG11 suggest the peak acceleration used as the threshold for human comfort in office and

residences subjected to vibration frequencies between 4Hz and 9.5Hz is 0.005g or 0.5% of gravity

(see Figure 7.2). The lower threshold with in the frequency of 4Hz to 9.5Hz can be reasoned by the

fact that human organs are very sensitive to low frequencies of vibration. Broner (1978) indicated

exposure to low frequency vibration (or infrasound) would cause nausea, disorientation and general

unpleasantness, in some extreme cases, suicidal attempt. Another reason for this reversed plateau

at 4Hz to 9.5Hz is the possible excitation of 2nd or 3rd harmonic of walking, where a harmonic is

defined as the an integer multiply of the occupancy step frequency. For example, human footfalls

average about 2 steps per second (Ji et al. 2005) under normal walking condition, If each footfall

is seen as an excitation, the following 2nd, 3rd as well the 4th harmonic will be at 4Hz, 6Hz and

8Hz, respectively. Floor that have a natural frequency at or near one of those harmonic frequency

may exhibit an excessive response since the harmonic frequency would coincide with the resonant

frequency of the floor.

DG11s procedure for estimating the natural frequency assumes a uniformly loaded, simply

supported beam:

r
π gEI
fn = (7.2)
2 wL4

which is essentially identical to the equation offered in the CLT Handbook. Note w is the uniformly

distributed weight per unit length, without any load combination.

The last two parameters required for the inequality (7.1) are the effective weight W and the

modal damping β. DG11 outlines a series of calculations for effective weight as a function of relative

stiffness of components, continuity factor and floor geometry. As far as damping ratio β, an accurate

way of obtaining such value would be through experimental measurement. If direct measurement is

170
not practical, DG11 demands that β to be estimated based on experience and recommended values

for the expected occupancy type of the evaluated floor. For office building (the type of occupancy

which the CLT-Glulam beam is designed for), DG11 recommended the following values for damping

ratios: 0.02 for floor with few non-structural components (ceiling, ducts, partitions, etc.), 0.03 for

floor with non-structural components and furnishings, but with only small demountable partitions,

and 0.05 for full height partitions between floors. It is not sure whether these recommended values

are valid for timber composite beams, however, the dynamic testing will provide preliminary modal

damping ratios for bare, unfinished floor.

European work and Eurocode 5 In Europe, the work of Ohlsson has been very impor-

tant, since it lay down the basis for Eurocode5. Ohlsson (1983) tested floors made of steel joist and

timber to develop design criteria. He conducted both in-field and laboratory studies where people

were invited to provide their feedback. His then design criteria was focused on limiting initial impulse

velocity, frequency transformed compliance as well as the spacing of adjacent natural frequencies.

Based on his work, the present Eurocode5 consider the following limits on pertinent parameters:

• The first eigenfrequency should be higher than 8Hz. If not, special investigation should be

made.

f1 > 8Hz (7.3)

• The maximum instantons deflection under 1kN (224.8lb.) point load applied at anywhere on

the floor taking the account for load distribution is limited to

w1kN < 1.5mm (7.4)

• The unit impulse velocity response, i.e. the maximum initial value of the vertical floor vibration

velocity (m/s) caused by an ideal unit impulse (1 N·s) applied at any point of the floor giving

maximum response (disregarding components above 40 Hz) is limited to:

v < 100(f1 ζ−1) m/s (7.5)

171
where f1 is the floor fundamental frequency and ζ is the modal damping ratio, taken as 0.01 unless

otherwise proven more appropriate.

Canadian work Onysko (1986) made a timber floor study in Canada through a survey

among residents and testing floors in existing houses. He found that the traditional North American

criterion of deflection under a uniformly distributed load was not well correlated to acceptability

among residents. The deflection under 1kN point load was better correlated. He also discovered that

the span length should be factored into consideration. The correlation to the duration of a transient

vibration also had a good correlation to acceptability, but since damping is difficult to estimate, this

could not be included in a design criterion. He stated that if damping can be properly identified,

the transient vibration response would be a reliable criterion for floor vibration serviceability design.

Three North American researchers, Hu, Chui and Onysko summed their work in a publication (Hu

et al. 2001) where they agreed that the frequency components, magnitude of response and damping

are the most important factors. Hu (2002) continued with field testing of timber floors and developed

five different criteria through the statistical logistic regression:

1
f1 0.39 ≥ 15.3
w1kN
L0.38
f1 ≥ 10.3
m 0.27 a0.19
RM S
f1 0.22 ≥ 15.3
w1kN (7.6)
L0.49
f1 0.14 ≥ 10.7
L0.35 apeak
f1 0.21 ≥ 17.5
v1kN

7.3.3 Past vibration studies on CLT

Information on vibration properties of CLT plate system is very limited, and almost non-

existent for a new generation of CLT products entering the North American marketplace. Typical

vibration study on CLT involves using modal analysis tools to determine the dynamic characteristics.

Modal data extraction focused on determination of fundamental natural frequencies, mode shapes,

separations between modal frequencies, and modal damping ratios; which are parameters used to

assess the vibration serviceability if floor system in buildings. Occasionally, finite element models

were proposed and calibrated based on the experimentally obtained data to further assist future

172
research.

The following is a literature review with particular regard to experimentally tested CLT

floor system (composite or non-composite), measurement comparison to Finite Element Method and

general composite beam modeling techniques. The realm of modal testing will be briefly introduced

here (though a more detailed experimental modal testing theory and application will be provided in

section 7.6), since it helps better understanding of the past CLT vibration studies.

Modal testing Modal testing refers to a referenced type of dynamic measurement which

both input force and output response will be used to form frequency response functions (FRF)

between the excitation and measured locations. In contrast, in some situations where response-

only measurements (unreferenced modal testing) may be the only measurements available because

capturing the input force is not practical or possible, such is the case for walking or heel-drop

excitations. Unreferenced excitation can still yield important information about a test floor, however,

it is not nearly as fully descriptive as a force-referenced test. The advantage of knowing FRF relations

in a test grid is that in addition to natural frequency and damping estimates, the mode shapes can

be estimated from the magnitude and phase information contained in the complex domain of FRFs.

Hanagan et al. (2003) notes the ability to express the behavior of a floor system in terms of its modal

properties provides a consist baseline for comparing the results of finite element analysis and actual

measured behavior.

In a modal test, the floor is excited by some mechanical means and its response is measured

using a set of accelerometers. A variety of floor excitation methods, as documented in the literature,

have been successfully implemented. The most common approach can be classified as transient

excitation, including instrumented hammer (Williams and Falati 1999), damped impact hammer

(Eriksson 1991), and heel drops (Hanagan et al. 2000). Continuous excitation can be created using

electric and hydraulic shakers. It is important to assess the different excitation source to identify

those most suitable for the test object, to evaluate walking induced vibration. In an effort to charac-

terize the best suitable modal testing technique and providing a set of guidelines for future research,

Hanagan et al. (2003) applied different modal testing techniques and quality control procedures to

a steel composite laboratory floor. They recommended that a transient excitation in the form of

a heel drop and a continuous excitation resulting from a chirp signal to a shaker yielded the best

results. When cost and portability are at issue, the heel drop is the best practice.

173
Heel drop is a very simple method that has been used for years (Lenzen and Murray 1969)

in which a person stands in the middle of the floor, assuming a natural stance, maintaining straight

needs, rising onto his toes approximately 2.5 in. and suddenly relaxing to allow the full weight to

freefall and strike the floor with his heels. The heel drop impact is normally idealized as a triangular

pulse with initial magnitude 600 lbs., decrease to zeros over 0.05 seconds, totaling an impulse of 67

N·s. Some design guides that use heel drop (such as Canadian Steelwork Code 1989) often treat

it as an impulse of 70 N·s. The obvious advantages of the heel drop test are that it is easy to

perform and relatively low burden on the budget. By using a simple force plate placed on the floor,

the well-known heel drop test can be adapted to be a part of an effective modal testing technique:

instrumented heel drop test. Blakeborough and Williams (2003) evaluated the use of instrumented

heel drop (where force impulse history is obtained through a force plate) test as an alternative

source of excitation for conducting modal testing on a floor. They discovered that the application

of instrumented heel drop test provide good coherence across the bandwidth. In addition, heel drop

test gives a much smother FRF, making it easier and more importantly to obtain accurate estimates

of natural frequency modal damping. They concluded that instrumented heel drop test showed

excellent resolution of frequency response functions in the range of 2 to 15 Hz, making it ideally

suited to determine the modal properties of potential floors.

The literature covering the applications of different modal testing techniques and quality

assurance measurements is quite vast as it mostly spans the automobile, aerospace and mechanical

industries. However, vibration testing related to engineered wood products are relatively rare,

especially those involves CLT. On top of this, the readers are directed to Ewins (2000) for a thorough

guidance through the theory, practice and application of experimental modal analysis.

Laboratory/In-situ testing of timber floors The following is a chronological survey of

the literature involving experimental testing and modeling of floor systems. The literature reported

here is purposely selected among those which test specimens involve engineered wood floor system.

In addition, several well documented studies that contained valuable modal testing experience are

cherry-picked since these experience helped shaping the experimental testing procedures adopted in

this study.

Hu and Gagnon (2012) performed a complete set of dynamic testing to assess the vibration

characteristics of bare CLT floors, these tests include: subjective evaluation, static concentrated

174
load of 1kN, instrumented hammer modal test and forced vibration test. The tested floor system

consisted of CLT panels made with different thickness (230mm, 182mm and 140mm), various length

(8m and 4.5m) and two types of joint details (step joint and spline joint). Their discovery confirmed

the combination of natural frequency with 1kN deflection, or with acceleration or velocity was well

correlated to human perception. Although CLT specimens had different thickness, joint means or

span length, the modal testing revealed a consistent value of around 1% for modal damping ratio.

Based on human perception study and experimental testing, the author proposed a simple from of

the design method that controls the span based on the geometry, stiffness of the floor system. It is

worth mentioning here that this serviceability equation was advised in the CLT Handbook, guiding

the preliminary vibration design of CLT floor.

Weckendorf and Smith (2012) performed an operational modal analysis on bare 5-ply CLT

(same thickness in each layer) specimens with different test parameters: support material (timer

and steel beam), boundary condition (clamped or not), and whether there’s person on the floor.

The floor specimens were excited by a person pushing and pulling a broad-band, low weight two-

wheel trolley while walking around the floor perimeter. Note that operational modal analysis is

an unreferenced testing technique, where only the response signal from accelerometers are analyzed

in the frequency domain, therefore ambient noise in the environment is picked up as well. In such

scenario, maintaining a low noise testing environment is a key. By comparing the frequency response

of tested specimen, they discovered that doubling the floor width results almost unchanged natural

frequencies but introduced additional modes. On the boundary condition comparison, they observed

a notable difference in both natural frequencies and modal damping ratios. The first order mode

damping ratio without people standing on the floor was reported to be below 1%, however, it

almost tripled (3%) if a person is located at the center of the floor. They suggested that a simple

extrapolation of vibration serviceability design practice proposed for other types of wood floor system

is not appropriate to design CLT plate floor system and should be discouraged.

Labonnote and Malo (2010) approached to the CLT vibration serviceability in a different

way. They conducted finite element analyses for the CLT floor system with various material prop-

erties. They believed that thicker plate with higher quality wood in the outer layer will increase

the vibration performance. Choosing an inner layer with optimal annual ring angle (close to 45◦ )

showed higher shear stiffness which led to improve vibration properties. However, the support condi-

tion of the CLT floor system in the construction usually constrained the plate in a two-way bending

175
situation, where controlling the ring angle in the inner layer may seem unpractical.

The dissertation of Barrett (2006) provides a well-documented study of the evaluation of

in-situ steel composite floors. Three areas were examined in his research: the best practice in high

quality dynamic testing of in-situ floor system, modal testing of multi-bay steel composite floor and

examination of dynamic properties, and development of a set of fundamental finite element modeling

techniques to adequately represent the dynamic response of the tested floor system. Based on his

experience using multiple excitation signals, burst chirp excitation generated using an electrodynamic

shaker is recommended as the most accurate and consistent source of excitation for acquire high

quality measurement suitable for use in the parameter estimation, operating deflection shape and

calibration/validation of FE models. He further addressed that the measurement, analysis and

computation of floors FRF is the core premise linking all areas of the modal study. Based on United

Kingdoms Dynamic Testing Agency (DTA) previously classified five different Levels of Test in its

Handbook of Modal Testing, the author also provided a meaningful classification of floor vibration

test further determined in terms:

1. the methods and equipment used for testing

2. the extend of the useful information that can be extracted from the tests

Hanagan et al. (2003) summarized their experience conducting modal analysis on the steel

joist floor system using both transient and continuous excitations. Several measurement issues

pertaining to the experimental modal analysis were explained and resolved: selection of measurement

bandwidth, excitation source, excitation level, excitation placement within the floor plan, recording

length and need for averaging. The definition of these terminologies and their settings within the

current study will be outlined later in sectionX 7.2. Meanwhile, Hanagan and her collaborators also

offered practical suggestions on the selection of testing equipment, e.g. FFT analyzer, vibration and

force transducers and their sensitivity requirements. Experience of modal analysis from the above

authors has been very beneficial in improving the quality of signal collection and analysis procedure

for assessment of vibration serviceability performed in the presented research.

176
7.4 Initial Vibration Assessment Using DG11 Guidelines

As discussed thereof, DG11 provides a quick assessment tool for the floor vibration due to

walking excitation. Fundamental natural frequency (Eq.(7.1)) and peak acceleration (Eq. (7.2))

are the key parameters that relates the floor dynamics to human vibration perception limits. How-

ever, in Eq. (7.1), the damping ratio β is difficult to estimate without direct measurement. The

damping associated with floor system primary depends on non-structural components, furnishings

and occupants. DG11 recommends modal damping ratios in a range of 0.01 to 0.05. The lower

value is suitable for floor with no non-structural or furnishings and few occupants, while the value

0.05 is suitable for offices and residents with full-height room partitions between floors and other

non-structural components. In this preliminary assessment, damping ratios are varied in the range

0.01–0.05 with 0.01 interval to investigate its sensitivity on the calculated peak acceleration.

The flexural rigidity EI of the CLT-Glulam beam, considering the composite action can be

expressed as:

n
X
EI = (Ei Ii + γEi Ai Zi2 ) (7.7)
i=1

where 0< γ ≤1 (γ=1 for rigid full composite action, γ=0 for non-composite). With the design

loading assumed in Chapter 6 (self weight, 20psf dead, 100 psf live), flexural rigidity EI calculated

using equivalent width method (shown in Appendix B), the fundamental frequencies estimated using

Eq. (7.2) are tabulated in Table 7.1.

Table 7.1: Estimation of fundamental natural frequencies based on DG11 guideline


Properties Full Composite Non-Composite

Span Length (ft) 39 39

EI (kip· in2 ) 2.878 · 107 1.918 · 106

fn (Hz) 7.08 1.02

It is worth noting that in Eq.(7.2), only the self weight of the beam is used for w; however,

in estimating the peak accelerations, the W in Eq.(7.1) includes the self weight, dead and live loads,

without any load combinations. Applying the estimated natural frequencies in Table 7.1, the peak

accelerations consider the variation of damping ratio and composite action is shown in Figure 7.3.

Observation of the figure reveals that ISO baseline and the limit for office and residential occupancies

177
both fall below the estimated peak accelerations, regardless of the selection of damping ratio and

composite factor γ. This indicates a poor vibration serviceability in terms of excessive acceleration

that the occupants might experience.

DG11’s quick hand calculation vaguely quantifies the dynamic characteristics of the com-

posite floor in the sense that it lacks estimation of true damping ratio and beam bending stiffness,

let along the idealized boundary condition and loading scenarios, the calculated natural frequencies

and peak accelerations can not be highly credited. In an effort to accurately estimate the dynamic

properties, the rest of this chapter will focus on the introduction of the modal analysis theory, test

execution and analysis of the of the instrumented heel-drop test. The measurement of floor acceler-

ations subjected to footstep impact serves as a direct indicator of the floor’s vibration serviceability.

Figure 7.3: Estimated peak accelerations (% gravity) for the CLT-Glulam composite beam

7.5 Basic Structural Dynamics of Floors

A very basic introduction of the subject of structural dynamics is presented in the following

section. For a more thorough guidance on the theory of structural dynamics testing as well as the

essential signal processing techniques, the readers are directed to (Craig jr and Kurdila 2006; Ewins

178
2000).

Before embarking on the detailed concept of structural dynamics, it is appropriate to put

different stages of concept into the context of a typical theoretical vibration analysis progress. Gen-

erally, the analysis starts with a description of the structure’s physical characteristics, usually in

terms of mass, stiffness and damping properties, this is referred as the Spatial Model. Then it is

customary to perform a theoretical modal analysis of the spatial model which leads to a description

of the structure’s behavior as a set of vibration modes: the Modal Model. This model is defined as

a set of natural frequencies with corresponding modal damping and vibration mode shapes. It is

important to remember that this solution always describes the various ways in which the structure

is capable of vibrating naturally, i.e. without any external forcing or excitation. Therefore, these are

called the ‘natural’ or ‘normal’ modes of the structure. The third stage is generally that in which

we have the greatest interest, namely the analysis of exactly how the structure will respond under

given excitation and especially, with what amplitudes. Clearly, this will depend not only upon the

structure’s inherent properties but also on the nature and amplitudes of the imposed excitation and

so there will be innumerable solutions of this type. However, it is convenient to present an analysis

of the structure’s response to a ‘standard’ excitation (from which the solution for any particular

case can be constructed) and to describe this as the Response Model. An example of ‘standard’ ex-

citation will be that of a unit-amplitude sinusoidal force applied to each Degree-of-Freedom (DOF)

individually, and at every frequency within a specified range. Thus the response model will consist

of a set of Frequency Response Functions (FRF) which must be defined over the applicable range

of frequency. Figure 7.4 illustrates the theoretical route to the vibration analysis. As indicated,

it is possible to proceed from Spatial Model to Response Model through a response analysis. It is

also possible to proceed to undertake an analysis in the reversed direction: i.e. from a description

of the response properties (such as measured frequency response functions) we can deduce modal

properties and less commonly, the spatial properties. This is an ‘experimental route’ to vibration

analysis which is shown as the reversed direction in Figure 7.4.

Although very few practical structures could be realistically modeled by a single-degree-

of-freedom (SDOF) system, the properties of such system is very important because those for a

more complex multi-degree-of-freedom (MDOF) system can always be represented as the linear

superposition of a number of SDOF characteristics. Considering a SDOF spatial model with mass

m connected to a spring with stiffness k. The equation of motion without external forcing can be

179
Figure 7.4: Theoretical and experimental routes of vibration analyses

obtained by applying Newtons second law:

mü(t) + k u̇(t) = 0 (7.8)

The trial solution x = Xeiωt leads to the requirement that:

k − ω2 m = 0 (7.9)

Hence the modal model consists of a single solution with a natural angular frequency given by
p
(rad/s). The natural frequency is easily obtained as: ωn = k/m (rad/s). The natural frequency

is easily obtained as:

r
ωn 1 k
fn = = (7.10)
2π 2π m

For a damped system, damping describes the rate of decay of the vibration response. There

are some different models for describing damping, e.g. viscous damping and structural damping,

Ewins (2000) gives a good explanation of different dampings. Essentially, all structures exhibit

a degree of damping due to the hysteretic properties of the material from which they are made.

Damping cause mechanical vibration energy to dissipate by transforming into other sources of energy.

When considering a multiple degree of freedom (MDOF) system, there exist more than

one natural frequencies or eigenfrequency (one for each DOF). Each eigen-frequency is associated

with a specific mode shape that describes the form of the vibration for that eigen-frequency. The

spatial parameters such as mass and stiffness along with the boundary conditions define the modal

properties.

180
7.6 Modal Testing Theory and Equipment Preparation

The key for accurately characterizing the dynamic behavior of a floor system and validating

the FE models to represent that behavior, relies solely on the acquisition of high quality dynamic

measurements. Thus it is of paramount importance to introduce the fundamental theory of modal

analysis that allows experimentalists to estimate the dynamic properties of a floor through testing. It

is also important to discuss the available testing methods and their associated equipment/techniques

since the acquisition of high quality modal data is a product of the available equipment and its

effective employment.

The following section provides an overlook of the theory of modal testing and the processing

of the dynamic measurements, the equipment used to acquire dynamic measurements, a description

of applied and recommended experimental modal testing technique for floor systems. It should be

noted that many of the techniques used in this study were based on recommendations outlined by

Hanagan et al. (2003) and others (Reynolds et al. 1999; Pavic and Reynolds 1999; Barrett 2006),

from their experience testing in-place steel composite floors and concrete floors; however, in many

cases, the recommended techniques are improved or expanded upon the experience of testing the

floor specimen of the presented research.

7.6.1 Modal testing method

Unreferenced Response-Only Modal Testing Unreferenced response-only modal anal-

ysis, more often known as operational modal analysis (OMA), is a technique for extracting modal

parameters from vibration response signals of mechanical structures. As inferred from its name, a

main difference compared to the referenced modal analysis technique is that the measurement of the

input forces is not required. This enables a testing of structures under operating conditions or in

other situations where the input forces are impossible to measure. It is therefore also called Ambient

Modal or Output-only Modal. This means, for instance, that if a bridge is going to be tested, the

bridge traffic and normal operation need not to be interrupted during the test. Similarly, if an engine

is going to subject to output-only modal testing, it is more desirable to perform such test with the

engine running at normal operating conditions. Implementation economy is one primary advantage

of ambient vibration tests as only the output vibration of the structure needs to be measured. This

is particularly attractive for civil engineering structures (e.g., buildings, bridges) where it can be

181
expensive or disruptive to carry out free vibration or forced vibration tests (with known input).

Perform modal test under operating (ambient or natural) conditions means that the struc-

ture is subjected to realistic vibration behavior, which might be difficult to obtain by use of artificial

excitation. It also means that the test can be performed simultaneously with other response test and

it provides the possibility for extraction of modal information under conditions where a traditional

accelerance based modal test is very difficult or impossible to perform. The ambient loadings that

are suitable for testing the floor system could include but not limited to: heel-drop, walking parallel

or perpendicular to the floor framing, or bouncing at harmonics of the bays dominant frequency.

The measured responses are governed by the dynamic characteristics of the system and

the forces, which excite the system. The derived model thus contains information of both the

system characteristics as well as the excitation signals. Therefore, identifying modal properties

using ambient data does have disadvantages:

1. The identification methods are more sophisticated. As the loading is not measured,

in the development of identification methods it needs to modeled by some stochastic process or

its dynamic effects on the measured response have to be removed. Otherwise it is not possible

to explain the characteristics in the data based solely on modal properties;

2. Without loading information, the identified modal properties can have significant

identification uncertainties. In particular, the results are as good as the broadband assumption

applied;

3. The identified modal properties only reflect the properties at the ambient vibration

level, which is usually lower than the serviceability level or other design cases of interest. This

is especially relevant for the damping ratio which is commonly perceived to be amplitude-

dependent.

The easiest dynamic measurement to take when testing a floor system is the unreferenced

response-only measurement because it requires the least amount of equipment and technique. A

single-channel hand–held spectrum analyzer, Figure 7.5, will have sufficient ability to conduct un-

referenced measurement, while being much less expensive and more portable. Examine the accel-

eration levels from various types of unreferenced excitation can indicate a lot about the behavior

of the floor under service conditions. Despite whether the input force is measured or not, it can

still provide adequate excitation to a floor system within frequency range of interest for floor vi-

182
bration serviceability. Thus, analysis of the frequency content of the response form an unreferenced

excitation can give insight into the active frequencies of the floor. The only information obtained

from a single channel measurement is the accurate measurement of time-history acceleration and

frequency domain of that history. Given certain types of excitation dynamic properties of the floor

can be estimated to a limited degree from the unreferenced measurement; however, it should be

understood that only fully referenced modal testing provides the accuracy typically needed for any

detailed characterization of the behavior of a floor system. As long as this limitation is understood,

then unreferenced measurement is an effective and low cost tool that has its place in dynamic testing

of floor systems.

Figure 7.5: Handheld spectrum analyzer (TestWorld Inc.)

Referenced Modal Testing As discusses thereof, results of response-only modal testing

contain information on both the structure itself and the excitation source. Therefore the Operating

Deflection Shape (ODS) shows the response of a structure to both resonant and forced vibration.

On the contrary, referenced modal analysis is similar to ODS in that deformation of the structure

can be viewed, but it is primarily concerned with resonance frequencies (or natural frequencies) of

a structure. Modal analysis is typically performed by exciting the structure with a known force

and measuring the response of the structure at many locations using accelerometers (or any other

motion sensors).

Base on the size of the specimen and desired signal-to-noise level, the excitation source can

be categorized into two classifications:

183
1. Impact testing is sometimes called bump testing, tap testing, or resonance testing.

It involves the use of an instrumented device to excite a structure in order to measure the

response. The commonly used excitation source includes instrumented hammer and force

plate. Due to the limited excitation energy induced by the impact, the size of the specimen

and the desired frequency bandwidth needs to be carefully planned out before execution of the

modal testing.

2. For larger or more complex structures it may be necessary to use an electrodynamic

shaker to excite the structure with sufficient energy to achieve the required signal to noise levels.

It is also useful to user shakers if extremely consistent and accurate results are required because

shaker input eliminates the potential for human error that might be found with impact testing.

With shaker excitation there are different types of signals to be used to excite the structure

under test: random, burst, sine, and chirp. Readers interested in referenced modal testing

using electrodynamic shaker should direct to Barrett (2006).

It is noted here that the referenced modal testing is selected for the completion of the

study since it’s a classic and mature dynamic analysis technology, where manuals, bulletins, journal

publications are abundantly available. Its ability in generating high quality dataset, obtaining

natural frequencies and mode shapes, along with its relatively easy and straightforward parameter

estimation process also contribute to the preference of referenced modal testing. The following

section explains the fundamental theory of the referenced modal testing.

7.6.2 Modal testing theory

Before diving into the modal testing theory, it is necessary to recognize and consider the

difference between referenced measurement data and unreferenced measurement data. If both ex-

citation force and resulting response are recorded, in either time or frequency domain, the test is

considered as a referenced modal testing. In some testing situations, response-only measurements

(unreferenced modal testing) may be the only measurements available because capturing the input

force is not practical or possible, such is the case for walking or heel-drop excitations. Unreferenced

excitation can still yield important information about a test floor, however, it is not nearly as fully

descriptive as a force-referenced test. The unreferenced modal testing would generate results with

relative magnitude and phase angle of a floor. It should be stressed that this would only yield

184
information about the frequencies and relative shape (not mode shape), and no information about

the acceleration response per input force, which is key in validating FE models.

In one of the comprehensive texts available on this subject, Ewins (2000) used three phases

to describe the theoretical multi-degree-of-freedom (MDOF) vibration analysis, which is based on the

premise that the response of a structure is just a combination of the response of its individual modes

of vibration. The first phase is to set up the governing equations of motion (i.e. determine mass,

stiffness and damping for the MDOF system of equations). The second phase is to perform a free

vibration analysis using equation of motion, the solution to free vibration yields the eigenvalue and

eigenvectors of the system of equations, containing modal properties in terms of natural frequencies,

damping ratios and mode shapes. The last stage is a forced vibration analysis where the applied

forces are often harmonic or sinusoidal in nature. By solving system of equations under applied force,

the complete solution is presentable using a single matrix, known as frequency response function

H(ω), which often short for FRF. Thus, Hij (ω) represents the harmonic response Xj in one the

degree-of-freedom j, caused by a single harmonic force, Fi , applied at a different degree-of-freedom, i.

The FRF is a complex frequency domain function that contains real part (amplitude) and imaginary

part (phase angle). With the definition explained, the FRF can be obtained as the response divided

by the input force:

X(ω)
H(ω) = (7.11)
F (ω)

Note that ω is the circular frequency in units in radians/second and is used simply for

convenience; it is interchangeable with circular frequency f , in units of cycles/second. The term

H(ω) in Eq. (7.11) is a general form often used to annotate a frequency response function. Using

the same notation seen in Ewins (2000), a more specific designation for the quantity describe in

Eq. (7.11) is a α(ω), the displacement response per input force, which is known as the ‘receptance’.

Equally, other forms of FRF exist based on selected response quantity. If the velocity response per

input force is used, the corresponding FRF is called the ‘mobility’ and the notation Y (ω) is used.

Likewise, of the acceleration response per input force is used, the FRF term is often referred as

‘accelerance’ and notated by A(ω). Displacement, velocity, and acceleration can be calculated by

185
derivatives or integrals of one another, and they can be related as follows:

X(ω)
Receptance : α(ω) =
F (ω)

Ẋ(ω)
M obility : Y (ω) = = (jω)α(ω) (7.12)
F (ω)

Ẍ(ω)
Acceleranec : A(ω) = = (jω)3 α(ω) = −ω 2 α(ω)
F (ω)


In mobility and accelerance expression, j = −1 and the jω terms are derived from the com-

plex representation of the harmonic forcing function f (t) = F ejωt , and the harmonic displacement

response x(t) = Xejωt . Therefore harmonic velocity and acceleration responses are ẋ(t) = (jω)Xejωt

and ẍ(t) = (jω)2 Xejωt = −ω 2 Xejωt , respectively. Note that these referenced FRF functions is in-

dependent of the excitation, meaning the vibrational characteristics (natural frequencies, damping

ratios) are irrelevant to how the structure is excited.

In a MDOF system like floors, not just one frequency response function is computed, but the

entire matrix called the frequency response function matrix [H(ω)]. Each entry of the FRF matrix

is a complex function in the frequency domain representing the input/output relationship between

two degrees of freedom on the structure for a given frequency ω. Thus, the accelerance FRF matrix

is an assemblage of terms Aik (ω), which relates the acceleration response of a particular location on

the floor i, to an input force at another location k, as shown:

 
A (ω) A12 (ω) · · · A1k (ω)
 11 
 .. 
Ẍi (ω) αi (ω) A21 (ω) A22 (ω) · · · . 
[A(ω)] = = = . (7.13)
 
Fk (ω) Fk (ω)  . .. .. .. 
 . . . . 

 
Ai1 (ω) Ai2 (ω) · · · Aik (ω)

While the derivation of the harmonic acceleration response of a MDOF system from the uncoupled

equations of motion is straight forward, it is presented in Appendix D for any interested reader. The

fundamental solution is shown in Eq. (7.14), which constitute a single term in the accelerance FRF

186
matrix:

R
X(ω) X −ω 2 φir φir /mr
Aik (ω) = = (7.14)
F (ω) r=1
(ωr − ω 2 ) + j(2βr ωr ω)
2

where R=the number of modes; ωr =the circular frequency of the rth mode; mr =the modal mass

of the rth mode; βr =modal damping ratio of the rth mode; [Φ] = [φ1 φ2 · · · φR ]=matrix of R mode

shape.

Note the expression is in a form that assumes viscous damping, and although not shown

here, there are more general forms that accommodate other damping models such as proportional

or hysteretic damping (Ewins 2000). The denominator of Eq. (7.12) contains the global properties

of the structure, the resonance frequencies and damping ratios, and does not hold any information

pertaining to the location of the excitation or response. This information is contained in the numer-

ator, which holds terms from the mode shape that scales the response. This fundamental analytical

expression of the accelerance FRF is the basis from which all experimentally determined dynamic

properties of the system are derived, as it is a mathematical expression of these quantity of interest

(frequencies, damping and mode shapes) in a formulation directly relate to the quantities measured

during modal testing (acceleration response and input force).

7.6.3 Signal processing

As stated previously, the frequency response function goes by many names, but it is simply

a transfer function which describe the relationship between two simultaneously measured signals:

input signal and output signal. For modal testing of a structure, it is often intuitively set up with the

input signal as the excitation of the structures, such as applied surface force or constraint excitation,

and the output signal as the physical response of the structure, such as displacement, velocity or

acceleration. It is the accelerance FRF matrix between an applied force and the acceleration response

of floor structures that is of interest in this study. The basis of the whole process relies on the accurate

measurement of signals in the time domain and Fourier analysis of those signals to transfer them

into the frequency domain:

Z ∞
1
x(t) = X(ω)ejωt dω (7.15)
2π ω=−∞

187
Eq. (7.15) is the Fourier Transformation and it simply states that a function in time

can be written as an infinitely series of sines or cosines with varying frequencies and amplitudes.

Obviously, anything involves infinity will be a headache to deal with, so several assumptions and

simplifications are made along the way to make this process manageable. Rather than infinite

time, only a window of time is analyzed. The signal within the time window is sampled a finite

number of times at discrete increments, which can be analyzed using a Simplified Discrete Fourier

Transformation (DFT) analysis. DFT process can be accomplished in several ways depending on the

type of modal testing setup and instrument employed. The most convenient way is to use a spectrum

analyzer (also known as signal analyzer or digital signal processor) which receives analog signals from

the force and acceleration transducers and immediately digitize the signal using an analog-to-digital

converter. It is this digitization that creates a discretely sampled time record of the signals, which

the spectrum analyzer uses to perform a Fast Fourier Transformation (FFT) on the time domain

data to transform it into the frequency domain data. Due to the instrument limitation in the

current research, both signals (excitation input and acceleration response) are collected using NImax

(National Instrument) and a Matlab build-in function fft (Fast Fourier Transformation) computes

the DFT of x(t) using a FFT algorithm. The FFT is a computationally efficient implementation of

the DFT when the digital time window is set up in a specific way, namely the time record length

must be a power of 2, such as 128, 512, 2014, 2048 or 8192. Each spectral values of the FFT is a

complex number, as it represents the corresponding Fourier coefficient that belongs to the sine and

cosine terms of the Fourier series at the given frequency, in addition to the phase lag angle. Thus,

for the two measured signals in time domain, the input force f (t) and acceleration response a(t), the

DFT computes F F Tf (ω) and F F Tx (ω), respectively. It is at this point, with the simultaneously

measured force input signal and acceleration response signal digitized in the time domain and their

respective FFTs computed, frequency domain functions used in the modal testing to describe the

dynamic properties of the floor are captured. From the FFTs of the digitized time history analysis,

the first frequency domain function computed is the cross-spectrum of the acceleration response

signal and the input force signal,Gxf (ω), which is a measure of the correlation between the two

signals:

Gxf (ω) = F F Tf∗ (ω) · F F Tx (ω) (7.16)

188
where F F Tx (ω) is the FFT of the acceleration response; F F Tf (ω) is the FFT of the input force;

F F Tf∗ (ω) is the complex conjugate of F F Tf (ω).

Because the values of the FFTs are complex numbers, the cross-spectrum are complex as

well, and phase information represents the relative shift between the two signals. The next frequency

domain functions computed is the auto-spectrum of each signal, Gf f (ω) for the input force signal

andGxx (ω) for the acceleration response:

Gf f (ω) = F F Tf∗ (ω) · F F Tf (ω)


(7.17)
Gxx (ω) = F F Tx∗ (ω) · F F Tx (ω)

where * again, represents the complex conjugate of the signal.

The auto-spectrum, which is the cross spectrum of a signal with itself, can be seen as a

measure of the relative strength of the signal at each of the spectral frequencies at which it is

computed. Note that this is a real valued function, as there is no phase lag between a function and

itself (also because the product of a complex number and its complex conjugate is a real number).

The auto-spectrum can be displayed in several formats, such as unit2 /Hz (commonly called the

Power spectral Density, or PSD), but the one used in the presented research is in root-mean-squared

(RMS) units, which is actually the square root of the values that are computed by Gf f (ω) and

Gxx (ω). This simply put the units in the auto-spectrum back into the intuitive units of the original

measured signal (lbf RMS and g RMS).

In fact, the computation of frequency response function (FRF) in its general form is:

Gxf (ω)
H(ω) = (7.18)
Gf f (ω)

Eq. (7.18) states that the FRF is computed by dividing the cross-spectrum of the excitation

and response signal (a complex valued function holding the phase information between the signals)

by the auto-spectrum if the excitation signal (a real valued function that simply scales the correlation

in cross-spectrum). This makes FRF a complex valued function, complete with magnitude and phase

information (or real and imaginary parts) at each spectral frequency in the units if response per

unit of input excitation. For typical floor vibrations, the FRF measured is the accelerance, the

acceleration response per unit input of force, and is displayed in units of in/s2 /lbf for convenience

189
of presenting the measured values, which generally ranging from 0–15 in the presented research.

Other units, such as g/lbf or m/s2 /N are acceptable, but they lead to much smaller values that can

be hard to work with (1in/s2 /lbf = 2.59 · 10−3 g/lbf = 5.71 · 10−3 m/s2 /N ).

It should be noted that while the cross-spectrum and auto-spectrum that formulate the

FRFs are computed between two signals each time a measurement is taken, it is the average of

repeated measurement that builds the level of confident that the computed spectral values are well

correlated (poorly correlated signal may indicates either high level of noise or very little response or

excitation). Thus, the coherence function is computed with each average, providing a quality check

on the correlation of all measured data. This computation is:

|Gxf (ω)|2
γ 2 (ω) = (7.19)
Gxx (ω)Gf f (ω)

At first glance, this function resembles the correlation function between two random vari-

ables, as it should, the coherence function measures the causality between the input and output

signals. Eq.(7.19) is often referred as the magnitude-squared coherence since it generates a non-

negative real number with range of 0 to 1. Similar to the correlation coefficient, the coherence being

0 indicates absolutely no correlation and 1 indicates two signals are perfectly correlated. Departures

from unity can indicate input noise, output noise, a non-linear process or any combination of these

things. When measuring the input-output behavior of a system, there is always noise present that

obscures the input and output measurements. An important measure is how much of the measured

output is actually caused by the measured input and a coherence spectrum serves this purpose well.

However, since the auto-spectrum and cross-spectrum are analyzed and computed in the frequency

domain, it is possible that the same frequency content appeared at different time. Eq. (7.19) may

not be subtle enough – an algorith is needed to locate frequencies present in both signals at the

same time. It is known as the Welch method and it was proposed by Welch in 1967. The underlying

signal processing techniques are beyond the scope of this study, but it can be easily found in many

textbooks, including that of Hayes (1996). A Matlab (Version R2016a) built-in function welch was

used to generate all the magnitude-squared coherence functions for this study. During the signal col-

lection of each foot-impact, coherence functions were continuously monitored to ensure the highest

quality measurements obtained whenever possible.

190
7.6.4 Parameter estimation

In general, modal testing can either be classified as gathering a set of fixed-input, roving-

response measurement (one column of the accelerance FRF in Eq. (7.13), and restated here for

convenience), or conversely a fixed-response, roving-input set of measurement (one row of FRF

matrix).

 
A (ω) A12 (ω) · · · A1k (ω)
 11 
 .. 
Ẍi (ω) αi (ω) A21 (ω) A22 (ω) · · · . 
[A(ω)] = = = .
 
Fk (ω) Fk (ω)  . .. .. .. 
 . . . . 

 
Ai1 (ω) Ai2 (ω) · · · Aik (ω)

If done properly, each yields the same information for the system. The former is typically

accomplished using a shaker excitation in a fixed location while accelerometers (or any other re-

sponse detectors) sweeping across the area of interest at measurement points. The latter is usually

accomplished with accelerometers placed in a fixed location while provide the excitation across the

area of interest. In this scenario, an instrumented impact hammer is traditionally used for the im-

pulse excitation in the fixed-response, roving-input test protocol. It is to note that the two testing

technique is not limited to the described setups, however, it is much less common to use an impulse

excitation in the fixed-input, roving-response setup. An alternative to shaker excitation in the fixed-

input, roving-response setup is using an instrumented heel-drop on a force plate at a fixed location,

which is an attractive low cost technique which yet provides the same high quality data. By mea-

suring the input excitation using a force plate, the modal test becomes a referenced test, therefore

high quality data can be assured. Analysis of set of measurement, the experimentally measured

accelerance FRFs is the final step in the processing of experimental modal analysis. Typically, the

measured FRFs contains one column or one row of entries, representing fixed-input, roving-response

and fixed-response, roving-input setup, respectively. These measurements are used to extract the

dynamic properties of frequencies, damping and mode shapes. This process is called the Parameter

Excitation. The different methods for conducting parameter estimation, and the theories behind the

mathematics for each of the methods, is also vast and is better left to texts devoted to the subject.

A good set of reference goes to the textbook by Ewins (2000) and a paper by Avitabile (2006) which

191
gives a brief overview and history of the development and implementation of the various methods

for parameter estimation. The method adopted in this research will be addressed in the section ded-

icated to presenting experimental results and testing techniques. Each term in Eq. (7.13), Aik (ω),

represents the accelerance FRF of acceleration response on one location on the floor with respect to

an input force at another location. In a special case, where the excitation and response are measured

at the same place, such FRF entry is called the driving point measurement and is located at the

main-diagonal of the FRF matrix.

The process of parameter estimation is based on relating these FRF measurements to the

mathematical formulation (from analytical derivation) of the accelerance in Eq. (7.12). In the most

basic term, the peak in the FRF indicate the presence of at least one mode, and sharpness of the peaks

indicates the level of damping in that mode. The relative magnitude and phase between different

Aik (ω) characterizes the mode shapes. While various methods are used for parameter estimation,

they all essentially formulate a mathematical formulation simulating the measured accelerance in

terms of estimated parameters (frequencies, damping and mode shape vectors) that would closely

approximate the analytical expression of the accelerance in Eq. (7.12). In such sense, the term

‘parameter estimation’ can be equally replaced with the term ‘curving fitting’. Specific methods used

for curve fitting and estimation of dynamic properties are presented in the data analysis section of the

chapter. With a cursory review of the basics of the modal analysis, the fundamental of the digital

signal processing, the parameter estimation, it is important to describe the commonly available

testing techniques and testing equipments.

7.6.5 Equipment preparation

A referenced modal test was performed on the CLT-Glulam composite beam to analysis its

vibration properties. The response signals from a array of accelerometers (4) were referenced against

the force plate which records the foot step impact. For such a setup, a fixed input, roving response

test was proposed. Figure 7.6 shows a schematic view of the experimental setup.

To briefly introduce the test setup: the laptop interface with the NI cDAQ with NIMAX

(National Instrument Measurement & Automation Explorer) software. Two NI modules were in-

stalled on the compact USB chassis: NI9234 for the force plate (three channels used) and NI9234

for the piezoelectric accelerometers (all four channels used). The NIMAX configured the physi-

cal channels in terms of their terminal locations, data type and sampling rate, whereas the signal

192
Figure 7.6: Schematic view of the vibration test setup

recording was performed by the University owned software DAQScribe. Synchronized excitation

and response signals were recorded in a time domain for post-processing. Detailed description of the

testing equipment used are presented in the following paragraphs:

Measurement of Acceleration Response–Accelerometers Acceleration response mea-

surements were taken using PCB model 333B50 shock accelerometers produced by PCB Pieztronics.

The 333B50 accelerometer has a frequency range of 0.5–3K Hz and operating measurement range of

±5g, they are well suited for the purposes of the vibration test of the timber floors. Each accelerom-

eter weights only 7.5 gram (roughly equals to a pencil) and with measured mean peak acceleration

around 0.7g, the accelerometer tends to flip over or jump around without fastening to the test

structure. The PCB engineers were consulted upon the feasible approaches to attach the Titanium

accelerometer housing to wood surface. They recommended super glue or beewax, however, the use

of super glue requires solvent for detachment and it really takes time considering that the accelerom-

eters are constantly roving their positions; the beeswax was also deemed unpractical since the porous

nature of the wood fibers causes the roughness of the was paste and decreases the adherent capacity.

Finally, they recommended a simple solution of hot glue that works well with wood surface and

Titanium housing (see Figure 7.7). It is advised that the thin soft layer of glue acts as a mechanical

filter and it is not feasible for high frequency modal testing (>1K Hz), however, for the problem at

hand with interest in a low frequency range, the attaching method should perform well.

193
Figure 7.7: Accelerometer attachment with hot glue

One of the most overlooked item in the modal testing might be the cables as Brillhart

and Hunt (2005) claimed accelerometer cables are often the weakest link in the setup. The 333B50

accelerometers requires a cable of 10-32 to BNC connection. The fiber glass coaxial cable is the most

expensive cable type and are extremely vulnerable to bent or kink. Nearly half of the time required

to test a floor system is relocating the accelerometers and managing the coaxial cables. Extreme

care was taken to ensure the working condition of the cables and it is recommended to purchase

the TFE jacketed cable as the reasonable extra cost would well worth the future frustration if cable

fails.

Measurement of Excitation Forces–Instrumented Force Plate The force plate is an

essential component for the referenced modal test. It is designed to be used on large civil engineering

plate structures such as building floors or pedestrian bridges. Its construction can easily be scaled

up or down depending on the range of the excitation forces. The weight of the device depends on

the plate material and the size of the load cells, and the frequency of interest are normally less

than 30 Hz (depends on the excitation source). The force plate is designed to sit on top of the

structure being excited. Howard et al. (1999) recommended that the weight of the force plate plus

the excitation source to be less than 1% of the weight of the test structure so that the addition of

equipment would have minimum effect on its dynamic response.

The in-house force plate was manufactured with two one quarter inch thick steel plates of

194
Figure 7.8: In-house instrumented force plate

Figure 7.9: Calibration of force plate

dimension 2ft by 2ft. Sandwiched in are three loadcells connected in parallel to the top and bottom

steel plates by threaded rods. Loadcells were manufactured by OMEGA Engineering with model

number LC401-1K, each with an operating capacity of ± 1000lbf. These load cells were individually

calibrated and leveled with respect to steel plates. Upon complete assembly of the force plate, the

device was once again calibrated to ensure voltage linearity. Steel blocks were sequentially placed

on the force plate and the output was recorded, the linearity of the force plate was remarkable with

a linear regression R2 infinity close to 1, see Figure 7.9.

One of the problems discovered during the modal testing is the incompatibility of the signal

to noise ratio (SNR) between the load cell signal and acceleration signal. Piezoelectric accelerometer

has proven its superior noise immunity to white noise (Levinzon 2004), however, the LC402 loadcells

195
are relatively less immune to the noise.

Loadcell, according to the type of output signal, there are many types of constructions,

for instance, strain gauge cells, mechanic cells, and other types (fiber optic, piezo-resistive, etc.).

The ones used in this study were a Omega Engineering low profile LC402, they are strain gauge

type with nominal resistance of 350Ω, and suitably wired in a four-arm bridge configuration. In

such configuration, two of the gauges are in tension and the rest two in compression, each wired

with compensation adjustments. Due to the nature of the resistor type of strain gauges, the load

cell is buried in a bard-band noise background where the unwanted information and the relevant

signal sometimes share a very similar frequency spectrum (Hernandez 2006). The more the noise

and the relevant signal share the same frequency spectrum, the less the designer can remove the

unwanted information by using classical filtering techniques (Kailath 1974; Anderson and Moore

2005). Among those classic filtering techniques, averaged based filters (moving averaging, Savitzky-

Golay filter, etc.) and resampling are beyond the consideration in smoothing the load cell output,

provided the following reason:

1. by smoothing the data, extreme values are inevitably clipped, while severity of

this clip increase with the length of the moving window;

2. it cause a temporal auto-correction at a lag. For modal testing which signal peak

and location carry important information regarding natural frequency and damping charac-

teristic, the moving average filter and resampling technique were deemed inadequate for the

smooth purpose.

Based on aforementioned disadvantages and the nature of high frequency load cell noise,

it was necessary to apply a filter that is able to eliminate high frequency unwanted signals. In

such scenario, it is ideal to use a FFT filter along with a low pass filter. Easily understood from

its name, an FFT filter performs a fast Fourier transformation to analyze the frequency content of

the input signal, then, followed by an inverse FFT to transform the filtered data into time domain

values. There are five types of filters available in the FFT filter: Low Pass, High Pass, Band Pass,

Band Block and Threshold. For the load cells use in the in-house force plate, low-pass filter was

selected since it is able to block all frequency components above the cutoff frequency, allowing only

low frequency signal (real heel-drop impact signal) to pass. A low-pass filter was designed using a

196
cutoff ratio:

Sampling Frequency
Cutoff Ratio = (7.20)
Cutoff Frequency

To visualize the functionality of FFT filter as well as to determine a reasonable frequency

cutoff ratio, the signal to noise ratio (SNR) of the original signal was investigated through a para-

metric study. One of the heel drop time histories was selected for this study, which was sampled at

1651 Hz. Figure 7.10 shows that the vast majority of the noise was filtered out at frequency cutoff

ratio of 2.1, meaning, for the force plate assembly, 1651/2.1=786 Hz represents the dominant noise

frequency. As a matter of fact, a higher cutoff ratio of 8 was selected since a higher ratio doesn’t

increase SNR anymore. Comparison of the original and filtered heel-drop time domain signal is

presented in Figure 7.11. Close inspection of the figure reveals that the peak magnitudes were well

preserved and no lag was observed. For the subsequent modal analysis, loadcell signals were first

filtered before forming the accelerance matrix; while the acceleration signals were untouched.

Figure 7.10: Loadcell noise in decibels (dB)

197
Figure 7.11: Original and filtered loadcell signal

7.6.6 Channel configuration

The signal acquiring equipment consisted of a NI cDAQ 9172 compact USB chassis, a NI 9234

module and a NI 9237 module. NI 9234 (Figure 7.12 slot 1) is a four-channel sound and vibration

input module that is able to measure signals from integrated electronic piezoelectric (IEPE) sensors,

such as accelerometers, tachometers etc. NI 9237 (Figure 7.12 slot 3) on the other hand, is a four-

channel analog input module. It has build-in Wheatstone Bridges (full, half and quarter bridges)

that can be activated based on the wiring diagram. For the cable connections, 9234 utilize a BNC

connector whereas 9237 uses a regular RJ-50.

A standard laptop running NIMAX 16.0 was used to configure the chassis and individual

modules. The configuration task involved assigning the physical terminal locations of each sensor,

set up the bridge type (full bridge for loadcell), signal input range, resistant (loadcell only), and

sensitivity (accelerometer only).

The NI chassis has a internal master time base that can be used to synchronize all the

modules it carries. The data sampling rate of the chassis, without any filter or resampling procedure,

198
Figure 7.12: Modal test signal acquiring system.

can be calculated as follows:

fM
fs = (7.21)
256 · n

where fM is the internal master time base 13.1072 MHz; n is an integer from 0 to 31. If n is taken

as 31, the minimum sampling frequency of the data acquisition system can be calculated as 1651.6

Hz and this was used throughout the modal test.

7.7 Experimental Testing Method

The experimental modal testing is a dedicated research activity that is often subject to

compromises and influences. Understandably, the limitation of the available equipment is the most

notable compromise that constrains the information that can be obtained. Time, however, is an

other critical factor that has the greatest influence on the quality, quantity and perhaps the level of

detail during the dynamic measurements. A fine frequency resolution is often critical and particular

when the damping ratio is low. In fact, fine frequency resolution comes at a cost of longer record

length. In addition, the existence of white noise, regardless if the testing is performed in field or in a

laboratory, requires higher number of averages to improve the quality of FRF measurement. Large

test area, fine testing grid, need for signal averaging all take time. The best way to address the

inevitable compromise is to clearly define the overall objectives of the dynamic testing of the floor

system and match those needs with the equipment and time needed for testing.

199
In the presented study, the test piece is a 5ft by 40ft composite timber beam that has

a potential vibration serviceability issue related to humman perceptions. The objectives are to

capture enough measurements to adequately estimate the modal properties and validate an FEM

model. This translate to the typical activities of the experimental modal analysis in identifying the

natural frequencies, estimating level of damping in the floor system and determine the mode shapes

(for corresponding natural frequencies). And the frequency band of interest is in the low 100, which

building occupants are sensitive at. While all measurements satisfy the overall objective of the

proposed modal testing to some degree, different types of measurement and methods of testing have

different goals. In traditional modal testing setup using a shaker, the sweeping sine or other types of

broadband excitation (eg. instrumented heel-drop) is used to measure the accelerance FRF over a

certain frequency range of interest. Steady state sinusoidal excitation is used primarily to verify the

accelerance FRF values at specific frequency lines and also serve as the initial excitation for decay

measure. In the presented study, the instrumented heel-drop was selected as the excitation source,

since it is easy to perform and requires no expensive equipment but a force plate. As Raebel et al.

(2001) indicated, the force plate provides better portability compare to the heavy electrodynamic

shaker and their study has proven the heel-drop is a good way of exciting the floor, there is no need

to use the shaker unless very high quality data is required or the frequency range of interest exceeds

30 Hz.

7.7.1 Instrumented heel-drop test

Heel-drop is a form of a broadband excitation where the foot heel impulse is performed on

the force plate and excites the test structure. It can be effectively used in the referenced modal

test in forming the accelerance FRF matrix. Heel drop is a very simple method that has been used

for years (Lenzen 1966) in which a person stands in the middle of the floor, assuming a natural

stance, maintaining straight needs, rising onto his toes approximately 2.5 in. and suddenly relaxing

to allow the full weight to freefall and strike the floor with his heels. A demonstration of heel-drop

excitation on a force plate in shown in Figure 7.13. The heel drop impact is normally idealized as

a triangular pulse with initial magnitude 600 lbs, decrease to zeros over 0.05 seconds, totaling an

impulse of 67 N·s. Some design guides that use heel drop (such as Canadian Steelwork Code 1989)

often treat it as an impulse of 70 N·s. The obvious advantages of the heel drop test are that it is

easy to perform and relatively low burden on the budget. By using a simple force plate placed on

200
Figure 7.13: Heel-drop impact on a force plate

the floor, the well-known heel-drop test can be adapted to be a part of an effective modal testing

technique: instrumented heel-drop test. Blakeborough and Williams (2003) evaluated the use of

instrumented heel drop (where force impulse history is obtained through a force plate) test as an

alternative source of excitation for conducting modal testing on a floor. They discovered that the

application of instrumented heel drop test provide good coherence across the bandwidth. In addition,

heel drop test gives a much smother FRF, making it easier and more importantly to obtain accurate

estimates of natural frequency modal damping. They concluded that instrumented heel drop test

showed excellent resolution of frequency response functions in the range of 2 to 15 Hz, making it

ideally suited to determine the modal properties of floors.

The force plate was placed on the slab-on-grade to investigate the measuring consistency over

a few heel-drop impacts. Figure 7.14 shows the time history of the main impact and subsequent

body vibrations of seven heel-drops performed by the same person. the signal of the seven time

histories was synchronized based on the time of the peak force and it can be seen that a single

person is able to achieve high degree of repeatability. Figure 7.15 shows a single heel-drop history

for an experimentalist (myself) weighted 170lb. The weight of the top steel plate was included in

the total weight. In the first part of the signal the experimentalist rises from a normal standing

position onto his toes. This causes small fluctuations in force as he accelerates upwards and then

comes to temporary rest. The experimentalist then balances on his toes for about half a second

before bringing his heels down, causing an initial reduction in vertical force as his center of gravity

accelerates downwards, followed by a very sharp increase in force at the moment of heel impact.

201
Following the main impact there is a short period of heavily damped oscillation at a frequency of

around 5 Hz: this is due to vertical vibration of the human body. As discussed thereof, heel-drop

is fixed input, roving response modal test. Upon explanation of how the impact signal is generated,

the following sections further discuss the location of the excitation, the placement of accelerometers

on testing grid and details of the signal acquiring.

Figure 7.14: Time history of 7 heel-drop impacts

Test piece discretization and accelerometer placement The general rule of thumb

of determining the test grid is to estimate the modal shapes and place enough grid points where

the mode shapes have credible chance to be captured. It is anticipated the first and second bend-

ing modes and the torsional mode, all in the long direction (due to the aspect ratio 8, the short

direction has much larger bending stiffness and torsional rigidity) will be captured. Therefore, for

the transverse direction, the beam surface was evenly discretized into four grid lines, while the the

long direction, the grid was spaced at 1ft on center. The dense grid of 156 nodes in Figure 7.16 (it

only shows half of the span) was used for analysis of the topic above. To clearly mark the nodes

for its convenient handling in the parameter extraction phase, the surface of the test specimen was

pencil-marked with 39 column and 4 row lines. The following equation was used to assign the node

202
Figure 7.15: A single time history of heel-drop impact

number based on its location:

N odeN umber = (Row − 1) × 39 + Column (7.22)

For example, the node on column 10, row 3 was marked with number 88.

Figure 7.16: Grid lines used in the heel-drop modal testing (only half of the span shown)

Four 333B50 accelerometers were ordered from PCB and they occupied all channels in

NI9237 during the test. These accelerometers were grouped in an array and were used to scan

through the specimen, one column at a time. Figure 7.17 shows one of the measurements where the

accelerometers were placed at Column 25 while force plate placed at Column 22, Row 2.

203
Figure 7.17: An array of accelerometers scanning the test piece surface at Column 25

Placement of excitation Placement of the excitation will not greatly affect the quality

of the data provided the excitation is not located on a nodal line or directly above a support

column. The nodal lines are referred to the line locations where the modal shape has no magnitude

(Schwarz and Richardson 1999). For a rectangular floor, the excitation should not be placed along

the centerline of the floor in either direction, it also should not be placed along the lines occurring

at the third points of the floors as there are nodal lines for some of the low frequency modes. If

the excitation or force input is located on a nodal line, the modes will not be seen in the frequency

response function and thus can’t be obtained. Therefore the excitation was placed at Row 2, Column

22 with node number 61. The photo and the schematic view of the excitation placement can be

found in Figure 7.18 and Figure 7.19.

Record length and bandwidth One of the key parameters in modal testing is the

determination of the bandwidth of interest. As discussed thereof, the frequency above 50 Hz do not

contribute significantly to the response of a floor subject to foot step impact. In fact, the heel-drop

excitation only provides sufficient energy in the 0–30 Hz frequency range, therefore a bandwidth of

30 Hz was chosen for the data collection. Once the bandwidth is determined, the minimum required

204
Figure 7.18: Picture of the excitation placement on node 61

Figure 7.19: Schematic view of the excitation placement

sampling rate can be calculated using the Nyquist Criterion (Golnaraghi and Kuo 2010):

fs > 2B (7.23)

205
The Nyquist Criterion clearly indicates the minimum sampling rate to sufficiently capture a contin-

uous signal is twice the bandwidth, 60 S/s in case. However, as discussed in section named Channel

Configuration, the sampling rate of the NI system was set at its lowest minimum: 1651.6 Hz. There-

fore the current setup of sampling rate was more than enough to completely capture the signal in

the desired bandwidth. In an effort to sufficiently record the decay of the motion, a 40 second time

window was used after each impact, making sure the motion is completely waned out. So for each

impact recording, each accelerometer will generate 66064 samples. Using the closest power of 2, the

number of samples used for data post processing is determined to be 216 = 65536. The following

table summarizes the digital signal process settings:

Table 7.2: Digital signal process settings


Bandwidth Sampling Rate Record Length Frequency Resolution

(Hz) (S/s) (Samples) (Hz)

30 1651.6 65536 (216 ) 0.025

Signal averaging It should be addressed that the experimental measurement of the ac-

celerance involves signal averaging to reduce the effect of ambient noise. In other word, repeated

excitations are needed for each of the measuring nodes. The cross-spectrum and auto-spectrum in

Eq. (7.18) are updated with each heel-drop impact, using the recorded force and acceleration re-

sponse. Note Fourier transformations is needed to convert time domain signals to the cross-spectrum

and auto-spectrum in frequency domain. In the following sections where frequency domain functions

are presented, they reflect the averaged functions. The number of averages needed for a measure-

ment is a compromise between data confidence and lengthy experimental time. The confidence in the

data is measured by the high level of coherence (Eq. (7.19)), while the minimum number should be

whatever ensures quality measurements as indicated by good coherence function. For experimental

measurement of the accelerance FRF, it was determined five averages of each heel-drop impact were

needed to ensure the quality of the measurement within the frequency range of interest, as a higher

number of average did not improve the coherence, only lengthened the time required to take the

measurement.

206
7.7.2 Flow chart of the test procedure

Prior to the destructive tests described in Chapter 6, two sets of non-destructive heel-drop

modal analyses were conducted for the composite beam. Each of the modal analyses varied in their

boundary condition of the end supports. The two boundary condition considered in the tests were

fixed and pinned, which can be seen in the Figure 7.20. The composite beam was fastened on a pair

of W18s on its two far end, while these two W18s were bolted down to the strong floor at all times

during the test. Such configuration is referred to as the grounded specimen test, compared to the

other two types: free support and in-situ. For the fixed boundary condition, the composite beam was

sandwiched in between two W sections (a W18 and a W8). In addition, a piece of 2x6 was clamped

underneath the top W8 to mimic the true loading scenario where a partition or load-bearing wall

sits on top of the floor assembly. It is to be noted that ‘fixed’ boundary condition achieved here is

an idealization of its true definition, the joint will allow small amount of rotations in three rotation

degree of freedoms. The pinned boundary condition involved clamping of the CLT bottom flange

to the grounded W18s through threaded rods. Therefore the translational DOFs were constrained,

allowing rotations in its rotational DOFs.

(a) Fixed B.C. (b) Pinned B.C.

Figure 7.20: Boundary conditions considered in the experimental modal analyses

The experimental modal test started with preparation activities such as setting up the lap-

top and force plate, gluing the accelerometers on the beam surface, running the wires, and checking

connectivity. A timer was used to remind the experimentalist of the 40 second signal collection

window. Once ready, the experimentalist heel-drop impacts the force plate (as well as the beam)

207
for 5 times before roving the accelerometers to the next column. It was repeated 39 times with 4

accelerometers in a column to cover whole specimen, as Figure 7.21 shows the measurement being

recorded at Column23. During the measurement of each column, coherence functions were contin-

uously monitored to ensure capturing the highest quality data whenever possible. To summarize, a

flow chart of the test procedure is Figure 7.22.

Figure 7.21: Schematic view of the excitation placement

7.8 Presentation of Measured FRFs

The previous sections described the theory, equipment and methods involved in experimental

modal testing using heel-drop excitation. To note, it virtually can be applied to any type of floor

system, not just composite timber floor. Laboratory floor system typically differs from in-situ floor

system, mainly due to the idealized boundary conditions, and the finite space available to achieve a

full scale multi-span, multi-bay setup. Contrast to grounded laboratory test specimens, on-situ floor

system typically covers larger area and the presence of partition walls and building contents often

contribute to the modal parameters. Unfortunately, opportunities to test large in-situ CLT floor

208
Figure 7.22: Modal test flow chart

system are rare as its difficult to locate a CLT building on the east United States. Ever rare are

the opportunities to test the floor system made of mass timber and fastened using the state-of-art

self-tapping inclined screws.

The presented research includes the extensive modal testing and complete parameter ex-

traction, all in the hope to characterize the dynamic property of the grounded composite timber

floor system. The floor assembly served as valuable test specimen and provide significant insight

into the improvement of the currently design, and possibly the application of multi-bay multi-span

floor systems. The rest of this section presents the measurement results and behavior of the test

209
floor in each boundary condition are summarized, including natural frequencies, damping, and mode

shapes. Finally, the vibration serviceability of the test floor is assessed again and compared to the

DG11’s estimation.

For each of the tested floor, the driving point accelerance FRF is presented first. This FRF is

the key measurement for the modal parameter extraction and provides the strongest measurements

for estimating frequencies and damping. The natural frequencies will be identified for both sets

of the experimental modal analysis. In each entry of the FRF accelerance matrix, the peak on

the magnitude plot (or a dive in the phase plot) is evidence of at least one mode, however, some

are clearly more significant than others. The dominant frequency can be visually identified as the

frequency with the largest magnitude. Therefore, the term ‘dominant frequency’ is used rather than

the ‘fundamental frequency’ since it is commonly reserved to describe the lowest frequency of the

structure and the dominant frequency may or may not be the fundamental frequency. Estimations

of damping using the half power bandwidth method were performed on the driving point accelerance

and composite FRF. For comparison of manually identified natural frequencies and damping ratios

for the FRFs, results from modal software ABRAVIBE were also presented. This modal analysis

software was further used to extract the mode shapes using SDOF curve fitting and fitting algorithm

will be presented.

Modal Analysis Software The modal analysis software used in this research was the

ABRAVIBE Version 1.2 (Brandt 2011). It is free, open Matlab toolbox for analysis of noise and

vibration signals. It was developed not only for vibration analysis, but also a tool for learning as

it provides numerous examples for common acoustic and vibration tasks. The software contains a

suite of Matlab executables that are able to handle the basic signal processing required in modal

analysis, it should be appeal to any one in the signal processing area who wishes to perform spectral

analysis or generate FRF matrix using FFT based methods. For a typical referenced modal analysis,

ABRAVIBE provides a template executable that allows a semi-automatic modal parameter extrac-

tions (among them the most important ones are frequencies, dampings and mode shapes) with the

user-provided FRF matrix. The software also allows the user to construct the 3-D model for the test

floor by importing the node location and connectivity matrices. The software also has the ability

to perform the parameter estimation by curve fitting a set of imported FRFs. Although multiple

fitting algorithms are available in the software, the chose technique was the SDOF rational fraction

210
polynomial curve fitting algorithm. The MDOF has the advantages of fitting several FRF peaks

simultaneously and capturing closely spaced modes, however, it was not needed in the presented

research since the natural frequencies were well separated and the application of SDOF algorithm

greatly reduced the processing time.

Driving point and full accelerance With the shaker at node 61, a total of 5 driving

point measurements were taken from the heel-drop impacts (same for other nodes). Figure 7.23

and Figure 7.24 contain the accelerance FRF magnitude, phase and coherence for five driving point

measurements. Notice the FRF magnitude and phase plots are very consistent between each mea-

surement and there are excellent coherence at the accelerance magnitude peaks, indicating a well

correlated, good quality data set were obtained. This full set of accelerance matrix contains 156

frequency response functions, with one at each measuring node. As Eq. (7.14) indicates, each entry

of the accelerance matrix contains the same information regarding the global modal frequencies and

damping ratios, however, driving point accelerance are typically selected thanks to their stronger

signal and better coherence. As a matter of fact, it is of interest to verify and compare the modal

parameters extracted from full set of data in a average sense and from driving point measurements.

Similar to Figure 7.23 and Figure 7.24, accelerance and coherence functions of 156 measurements

are presented in Figure 7.25 and Figure 7.26.

In addition to the 2D plots, commonly displayed FRF formats include waterfall plot and

spectrogram (although waterfall is usually the inverted spectrogram), these plot provide better

visualizations of the FRF magnitude or even sometimes the mode shape if the test piece possess a

very simple geometry (Avitabile 2006). Spectrograms presented in this study are three dimensional

plots made by stacking up the consecutive 2D plots shown in Figure 7.25. Figure 7.27 shows a

spectrogram of the FRF magnitudes measured at all nodes, i.e. the Y-axis (DOF Number) notates

the FRF measurement location. The ridge-like peaks in Figure 7.27 indicates the existence of a

mode and the spreadness of the peak at the mode frequency informs the damping ratio of this very

mode. The following section will discuss the extraction of natural frequencies and mode shapes from

driving point as well as complete set of FRF measurements.

211
Figure 7.23: Driving point accelerance and coherence for the fixed boundary condition. Force plate
and accelerometer in node 61

7.9 System identification from measured FRF data set

System identification from measured FRFs plays a crucial role in structural dynamics as

it bridge the gap from experimental measurement to system dynamic properties. The most pop-

ular identification approach is based on the modal analysis method, leading to a interpretation of

visualized eigenmodes.

In practice, near all vibration problems are related to structural resonance behavior (that

is natural frequency excited by input force). Earlier in this chapter, it has been shown that the

complete dynamic behavior of a structure (in a frequency range) can be viewed as the summation

of a set of individual modes of vibration, each having a characteristic modal parameters: natural

frequency, damping ratio and mode shape. The modal parameters can be identified from a set

212
Figure 7.24: Driving point accelerance and coherence for the pinned boundary condition. Force
plate and accelerometer in node 61

of frequency response measurement between a reference point (driving point) and a number of

measurement points. Such measurement point, is usually referred as a degree-of-freedom. The

modal frequencies and damping ratios can be found from all FRF measurements on the structure.

These two parameters are therefore called the ‘global parameters’. A FRF measurement on a typical

DOF is often repeated several times for improved accuracy and reliability. However, to accurately

extract the mode shapes, FRF measurement must be made over a number of DOFs, to ensure

sufficient detailed covering over the test piece.

213
Figure 7.25: Accelerance and coherence plot for all measured 156 nodes for fixed boundary condition

7.9.1 Determination of modal frequencies

The purpose of the modal tested performed in this study involves the identification of the

modal parameters, with particular interest in the natural frequencies. Natural vibration frequencies

are directly related to human perception of annoyance as low range frequencies often cause resonance

of human organs, and this could lead to nausea, headache, elevated mental stress or other discomforts.

Observation of spectrogram (Figure. 7.27) averaged at all DOF indicated the test specimen to be

a lightly-coupled (or simple) structure. The modes are not closely spaced and are not heavily

damped. It is therefore assume that the CLT-Glulam composite beam to be a simple structure,

as it behaves predominately as a combination of a suite of SDOF systems. So far, a number of

mathematical modes on the modal parameter extraction have been developed and can be roughly

classified by either parametric or non-parametric method. Generally, parametric methods such as

214
Figure 7.26: Accelerance and coherence plot for all measured 156 nodes for pinned boundary condi-
tion

Ibrahim Time Domain (ITD), Eigenststem Realization Algorithm (ERA), Least-square Complex

Frequency Analysis (LSCF), Algorithm of Mode Isolation (AMI), etc. are preferable in accurate

estimation of modal parameters. However, they tend to be more computational expensive. Non-

parametric ones such as Peak-picking (PP), Frequency Domain Decomposition (FDD) have the

advantages being fast in process and easy to visualize the results.

Among the non-parametric methods, peak picking has been widely used to quickly assess the

natural frequencies due to its simplicity, straightforwardness and reliability (De Roeck et al. 2000).

215
(a) Fixed B.C. (b) Pinned B.C.

Figure 7.27: Spectrogram of all FRF measurements

On the other hand, the accuracy of the peak picking method often depends on the freuquency

resolution of the analysis. In this study, 40 seconds of signal was recorded for each heel-drop impact,

resulting a frequency domain resolution of 0.025 Hz, which is acceptable for using the peak pick

method. The method itself is named after its key step: the identification of the natural frequencies

as the peaks of an FRF plot. The peak picking method is based on the fact that the FRF goes

through an extremum around the natural frequencies. The frequency at which this extremum occurs

is a good estimate for the natural frequencies.

Theoretically, natural frequency and damping are global system parameters that do not

vary with respect to measurement points, however, equipment noise, leakage and white noise may

contribute to variations in the detected parameters. Therefore, in addition to the parameter ex-

traction from driving point accelerance, it is also a common practice to identify natural frequencies

and damping ratios on each measured FRF and take their mean as the system modal parameters.

However, this is sometimes easier said than done since the DOF on the nodes might not capture

any resonance frequency at certain vibration mode (a node is the point that remains still when the

structure is vibrating at a certain natural frequency. Therefore a peak picking at composite FRF

defined as the average of the magnitude at all response points gives a good overall estimate of the

natural frequency and damping ratios. Mathematically, the composite FRF can be calculated as:

Ni X
No
1 X
Hc (ω) = composite(realH(ω)) = real(Hjk (ω)) (7.24)
No Ni j=1
k=1

216
Using Eq. (7.24), the composite FRF for fixed boundary condition with the peaks identified

can be shown in Figure. 7.28. The composite FRF gives a good global picture of the information

contained in the individual FRFs. For example, all of the modal peaks seen in Figure 7.27 are

evident in Figure 7.28. Note that it is difficult to find a single FRF put of the set in Figure 7.27

that gives as good a representation as the composite does of all pf the modes of the structure. One

reason for this is the noise in the individual magnitude FRFs is canceled when averaging them from

to form the composite FRF. Use composite FRF has the following advantages:

(A) Decreased noise level due to the average process. The curve are smoothed for

easier frequency identification;

(B) Increased visibility of peaks in the frequency domain FRFs since certain measure-

ment DOFs at node might not carry all the modal parameters.

(a) Fixed B.C. (b) Pinned B.C.

Figure 7.28: Composite FRF magnitude created from the heel drop test. Natural frequencies are
identified as red dots

For driving point accelerance (Figure 7.23) and composite FRF maginitude (Figure 7.28),

peak picking method identified the same natural frequencies for the fixed and free boundary condition

test pieces, with 7.2 Hz and 6.7 Hz being as the dominant frequencies, respectively. The other non-

dominant natural frequencies are reported along with the damping ratios in the following section.

217
7.9.2 Determination of modal damping ratios

The first method to obtain damping of a mode is the half power bandwidth. The half power

use the transfer function FRF directly to obtain the damping. The half power band is defined as the

frequency range enclosed within the -3 dB drop, which is roughly at 2 = 0.707 of the maximum

amplitude (exact:20log10 ( √12 ) ≈ −3.0103dB).

Once a mode is isolated, the peak magnitude and natural frequency can be identified as A1

and fr . Then magnitude at half band frequency fa and fb is:

A1
A2 = √ (7.25)
2

The frequency fa and fb associated with the half power points on either side of the peak

are obtained, as shown in Figure 7.29. Then the damping ratio ζ is obtained using the following

equation:

fb − fa
ζ= (7.26)
2fr

It is worth noting that half power bandwidth method limit its application to lightly damper

structures (less than 5%), as the damped natural frequency ωd is typically close to natural frequency

ωn (see Chopra et al. (1995) for derivation).

Alternatively, the damping can be obtained from time domain analysis which requires no

new measurement. Consider free vibration of a damped system, the impulse response function can

be found in many articles, including that of Harris and Beyer (1988):

p
h(t) = A0 e−ζωn t sin( 1 − ζ 2 ωn t + ζ) (7.27)

where ωn is the natural frequency, A0 is the residual magnitude, and ζ is the damping ratio associated

with this natural frequency. Here the oscillating term is given by a sine wave at damped natural

frequency and the damping is represented by the exponentially decaying curve. It is well known

that the dissipating mechanism can be detected by the analysis of the decaying envelop A(t) of

the impulse response function. For a SDOF system, the constitutive function A(t) is known in the

218
Figure 7.29: Illustration of half power bandwidth method

explicit form:

A(t) = A0 e−ζωn t (7.28)


where ζ = 2c km and m and k are mass and stiffness of the system respectively. The envelop

function can be be obtained using a Hilbert transform (Bracewell and Bracewell 1986) of the impulse

response function:

ln A(t) = −ζωn t + ln A0 (7.29)

Thus the damping ratio ζ of the system can be estimated form the slope of the straight

envelop line A(t) plotted in a log in a semi-logarithmic scale. This procedure is well known in the

literature and can be found elsewhere (Staszewski 1997; Brancaleoni et al. 1993).

Table 7.3 and 7.4 shows the damping ratios estimated by the half power bandwidth method

and the time domain impulse response functions. Both of them yielded close estimations. Further

observation of Tables reveals that all damping ratios are less than 5% of the critcal damping. This

has indicated that the assumptions made earlier which treat the beam as a simple structure as well

as light coupling between modes are valid.

219
Table 7.3: Damping estimates for composite beam with fixed boundary condition

Mode number Natural Frequency (Hz) Half power bandwith Impulse Response Function
Mode1 7.2 1.00% 0.99%
Mode2 15.8 1.14% 1.14%
Mode3 20.5 1.39% 1.31%
Mode4 24.0 2.97% 3.00%

Table 7.4: Damping estimates for composite beam with pinned boundary condition

Mode number Natural Frequency (Hz) Half power bandwith Impulse Response Function
Mode1 6.7 1.03% 1.01%
Mode2 15.5 1.13% 1.14%
Mode3 19.4 1.41% 1.36%

7.9.3 Determination of mode shapes

The previous sections discussed the extractions of the key modal parameters: natural fre-

quency and damping ratios. If the structure is excited with any of its natural frequencies, the

resonance occurs. The structure naturally vibrates and displace in a pattern and this configuration

is often described as the mode shapes. The animate display of mode shapes describes the pattern of

the vibration and it’s very useful for N V H (noise, vibration, and harshness) engineers. The open-

source ABRAVIBE Version 1.2 (Brandt 2011) software was used to analyze the vibration signals

measured in the experimental modal test. Before diving into the mathematical formulas and curve

fitting, it is of paramount importance to introduce the concept of system poles, which will be used

throughout the ABRAV IBE program.

The motion of a single degree of freedom system can be described by Eq. (7.30):

mẍ(t) + cẋ(t) + kx(t) = f (t) (7.30)

with m the mass, c the damping coefficient, and k the stiffness. The equation states that the sum of

all forces action on the mas m should be equal to zero, with externally applied force f (t), −mẍ(t)

the inertia force, −cẋ(t) the viscous damping force and −kx(t) the restoring force. The variable

stands for the position of the mass m with respect to its equilibrium point. Transforming Eq. (7.30)

to the Laplace domain (assuming zero initial conditions) yields:

Z(s)X(s) = F (s) (7.31)

220
with Z(x) the dynamic stiffness:

Z(s) = ms2 + cs + k (7.32)

The transfer function H(s) between displacement and force (Laplace domain Receptance),

X(s) = H(s)F (s), equals the inverse of the dynamic stiffness:

1
H(s) = (7.33)
ms2 + cs + k

The roots if the denominator of the transfer function, i.e. d(s) = ms2 + cs + k, are the poles

of the system. In other word, at system poles, the vibrating structure reaches its global ultimatum

(resonance). The poles, denoted λ usually resulting in a complex conjugate pair:

λ = −σ ± iωd (7.34)

p
with fd = ωd /2π the damped natural frequency; fn = ωn /2π natural frequency where ωn = k/m =

|λ| and ζ = c/2mωn = σ/|λ| the damping ratio. If, for instance, a mass δm is added to the original
p
system, its natural frequency decreases to ωn = k/(m + δm). If c = 0, the system is not damped

and the poles becomes purely imaginary, λ = ±iωn .

The Frequency Response Function (FRF), denoted by H(ω) is obtained by replacing the

Laplace variable s in Eq. (7.33) by iω resulting in:

1 1
H(ω) = = (7.35)
−mω 2 + icω + k (k − mω 2 ) + icω
p
Clearly, if c = 0, then H(ω) goes to infinity for ω approximating to ωn = k/m.

Although very few practical structure could realistically be modeled by SDOF system, the

properties of such a system are important because those of a more complex multiple-degree-of-

freedom system can always be represented as the linear superposition of a number of SDOF charac-

teristics.

For a structure with N degree-of-freedom, the transfer function H(s) similar to Eq. (7.33)

can be written as:

1 N (s)
H(s) = = (7.36)
M s2 + Cs + K d(s)

221
with the numerator polynomial matrix N (s) given by:

N (s) = adj(M s2 + Cs + K) (7.37)

where adj is the adjugate and the common denominator polynomial d(s), also known as the charac-

teristic polynomial:

d(s) = det(M s2 + Cs + K) (7.38)

When the damping are small, the roots of the characteristic polynomial d(s) are complex

conjugate pole pairs, λm and λ∗m , m = 1, ..., Nm , with N.m the number of modes of the system.

The transfer function can be rewritten in a pole residual term (Balmès et al. 1996), i.e. the so-called

‘modal’ model:

Nm ∗
X Rm Rm
H(s) = + (7.39)
m=1
s − λm s − λ∗m

where s = jω if converted into modal mode. The residual matrices Rm , m = 1, ..., Nm , are defined

by

Rm = lim H(s)(s − λm ) (7.40)


s→λm

It can be shown that the matrix Rm is of rank one, meaning that Rm can be decomposed

as:

 


 Ψm (1) 



 

 
 Ψm (2) 
  
T
Rm = Ψm Ψm .. Ψm (1) Ψm (2) ··· Ψm (Nm ) (7.41)
.

 


 


 


Ψ (N ) 
m m

with Ψm a vector representing the ‘mode shape’ of mode m. From Eq. (7.39), it is concluded that

the transfer function of a MDOF system with Nm DOFs is the sum of Nm transfer functions and

that the full transfer function matrix is complete characterized by the modal parameters, i.e. the

poles λ = −σ ± iωd and the mode shape vectors Ψm . As such, the following section will focus on

the the mathematical extraction of system poles and mode shape vectors.

222
Estimation of system poles If a measurement can be taken in a reliable and true

fashion, then it can be assumed that a parametric model can be defined that describe the data.

This process is often referred to as the ’curve fitting’ as the modal testing community strive to

parametrically express the measured signals, regardless whether the data are presented in frequency

domain (FRFs) or time domain (impulse response function). For impulse response function (IRF),

the relation between measurement and modal parameters is expressed:

Nm
X ∗

hij = (rijm eλm t + rijm eλm t ) (7.42)
m=1

where λm is the pole value for mode m. Convert Eq. (7.39) into modal space, the similar relationship

in frequency domain can be expressed:

Nm ∗
X rijm rijm
Hij (jω) = + (7.43)
m=1
jω − λm jω − λ∗m

With the established relationship between measurement and mathematical model, curve

fitting techniques are required to interpret first the system poles λ and the corresponding mode

shapes Ψm . A variety of fitting method are available, selection of a method will primary base on

the application, DOF, domain and type of inputs. A modal analysis software manual published

by Simens LMS Lab (Simens 2011) provides a brief overview of the popular extraction methods.

However, only the method used in the ABRAVIBE program is discussed: least square complex

exponential method (LSCE). LSCE has its unique advantages of handling multiple input multiple

output (MIMO) data set and global parameter estimation ability in the time domain. Global

parameter estimation refers to fitting of multiple systems poles simultaneously, however, LSCE is

often challenged with highly coupled structure with close spaced modes. For better estimation of the

system poles and mode shapes, LSCE was applied at each manually select frequency which contains

only one peak.

The first step involved in using the LSCE method is to select the frequency range of interest

(see Figure 7.30 for example selection for the dominant mode) and convert the accelerance to recep-

tance since the transfer function shown in Eq. (7.36) is defined upon displacement. This conversion

can be found in Eq. (7.12). The frequency domain FRF were then converted into time domain

equivalent impulse response function (IRF). Since the measured data was sampled (not continuous),

223
so rather than working with Eq. (7.42), it necessary to work with:

Nm
X
n ∗ ∗n
hij = (rijm zm + rijm zm ) (7.44)
m=1

where zk = eλm δt . Instead of damped complex exponentials, the characteristics are now power series

with base number zm . It can be shown that the sampled data is the solution of a linear finite

difference equation with constant real coefficient of order 2Nm :

hij,n + a1 hij,n−1 + ... + a2Nm hij,n−2Nm = 0 (7.45)

The characteristics zk as well the system conjugate poles can be found by solving:

zk2Nm + a1 zk2Nm −1 + ... + a2Nm = 0 (7.46)

Figure 7.30: Section of frequency range for LSCE method. Figure shows the range selected for the
dominant frequency

Such procedure was repeated for each peak in the accelerance plot and thus a total of seven

poles were identified for the measured signals. Table 7.5 identifies the seven system poles and shows

agreement with previously identified natural frequencies and damping ratios (Tables 7.3 and 7.4).

Curve fitting for mode shapes Once the system poles were identified, the curve fit-

ting process for mode shapes were fairly easy considering the use of Eq.(7.43). The residual term

224
Table 7.5: System poles identified for the composite beam during modal testing

Mode number System pole


Fixed–Mode1 -0.4524 ± 45.2367i
Fixed–Mode2 -1.1317 ± 99.2679i
Fixed–Mode3 -1.7904 ± 128.7929i
Fixed–Mode4 -4.4787 ± 150.7299i
Pinned–Mode1 -0.4336 ± 42.0951i
Pinned–Mode2 -1.1005 ± 97.3882i
Pinned–Mode3 -1.7187 ± 121.8817i

containing the mode shape vector Ψ can be obtained by flipping Eq. (7.43):

Rm = Hij (jω)(jω − λm ) (7.47)

Figure 7.31 shows the curve fitting result at DOF-1 for the composite beam with fixed

boundary condition. the fitted curve captured the peak location and the openness of the curve

indicated the correct amount of damping was parameterized. During the curve fitting process, each

synthesized curve was visually inspected to ensure accuracy.

Note λm is the complex conjugate system pole and Hij (jω) is the measured Receptance FRF. Curve

fitting loops through all measurement point i and generates the mode shape vector Ψ. It is well

understood that there is no unique solution for mode shapes and the mode shapes presented in this

study were scaled individually for better visualization. The 2D horizontal line and 2D orthographic

view visualization functions developed in ABRAVIBE were used to plot the mode shapes. Recall

that the test pieces was grided into 156 DOFs with 39 columns and 4 rows (Figure 7.16), the display

models were constructed in the exact same fashion with the black dot indicates the measuring point.

So one should expect to see four dots in one measuring column in the 2D plot. Figures 7.32 and

7.33 show the all mode shapes at identified mode shapes. Upon visually inspection of the plotted

mode shapes, the recognized vibration modes are summarized in Table 7.6.

Table 7.6: Vibration modes for composite CLT-Glulam beam

Mode number Natural frequency (Hz) Vibration mode


Fixed–Mode1 7.2 First bending mode
Fixed–Mode2 15.8 First torsional mode
Fixed–Mode3 20.5 Second bending mode
Fixed–Mode4 24.0 Second torsional mode
Pinned–Mode1 6.7 First bending mode
Pinned–Mode2 15.5 Torsional mode
Pinned–Mode3 19.4 Second bending mode

225
Figure 7.31: Example curve fitting at DOF-1

7.9.4 Assessment of signal quality

It is clear that the mode shapes presented in Figure 7.33 contain irregularities and these ir-

regularities are not uncommon to the experimental modal analysis. Multiple causes could contribute

to the slightly degraded mode shapes: quality of the signal, the compatibility of the excitation and

response signals, the content of white noise and equally important is the curve fitting process. One of

the direct measurement of signal quality is the coherence function discussed in the Signal Processing

section. The coherence functions indicates the correlation between excitation and response and were

therefore continuously monitored during the test to ensure linear dependence. Figure 7.34 presents

the coherence functions which are plotted on top of the composite FRF. Note since each of the

measurement DOF should generate a coherence function, a set of 10 and 90 percentile functions are

provided to indicate the variance within the 156 measurements. The coherence functions indicate

excellent linear dependence between excitation and response above 5 Hz. In fact, at FRF peaks

226
(a) Fixed–Mode1–7.2Hz–First bending mode

(b) Fixed–Mode2–15.8Hz–First torsional mode

(c) Fixed–Mode3–20.5Hz–Second bending mode

(d) Fixed–Mode4–24.0Hz–Second torsional mode

(e) Pinned–Mode1–6.7Hz–First bending mode

(f) Pinned–Mode2–15.5Hz–Torsional mode

(g) Pinned–Mode3–19.4Hz–Second bending mode

Figure 7.32: 2D mode shapes at identified natural frequencies

227
(a) Fixed–Mode1–7.2Hz–First bending mode (b) Pinned–Mode1–6.7Hz–First bending mode

(c) Fixed–Mode2–15.8Hz–First torsional mode (d) Pinned–Mode2–15.5Hz–Torsional mode

(e) Fixed–Mode3–20.5Hz–Second bending mode (f) Pinned–Mode3–19.4Hz–Second bending mode

(g) Fixed–Mode4–24.0Hz–Second torsional mode

Figure 7.33: 3D mode shapes at identified natural frequencies

228
(a) Fixed B.C. (b) Pinned B.C.

Figure 7.34: Coherence function with 10th and 90th percentile values

where natural frequencies were identified, the 90th percentile values consistently reach beyond 0.92.

It is therefore concluded that the signal obtained was of good quality. The curve fitting, on the

other hand, was only subjected to visual inspections. Some of the DOFs may not fit well due to

the complex shape of FRF or a mode close by. However, contrary to the identification of natural

frequency and damping ratios, the extraction of mode shapes is not of greater importance in the

presented modal analysis. Should mode shapes be in need of higher quality, advanced multi-input-

multi-output (MIMO) algorithms such as Algorithm of Modal Isolation (Allen and Ginsberg 2006)

can be utilized.

One of the statistical indicators, similar to coherence function, is the Modal Assurance

Criterion (MAC) for quantitative comparison of modal vectors. The MAC was originally introduced

in the modal testing in connection with the Modal Scale Factor, as an additional confidence in the

equivalent of modal vector fro different excitation locations (Pastor et al. 2012). However, it is also

often used to compare the measured mode shape against vectors determined by analytical model,

or the difference of vector from different curve fitting process or one mode shape before and after a

change in the physical structure caused by a wanted modification. In the presented study, it is used

to verify whether two modes are coupled and their correlation. The correlation between two mode

shapes is computed using:

|ΨTj Ψi |2
MAC(i, j) = (7.48)
(ΨTj Ψj )(ΨTi Ψi )

229
(a) Fixed B.C. (b) Pinned B.C.

Figure 7.35: Modal assurance criteria values for obtain mode shapes

Supply Eq. (7.48) with the mode shape vectors obtained through curve fitting, the MAC

plot for the modal test with two different boundary conditions is shown in Figure 7.35.

The MAC computation gives a real-values scalar quantity from 0 to 1.0, with 1.0 representing

perfect correlation between the two vectors and 0 indicating that the vectors are orthogonal. MAC

values in Figure 7.35 show mode shapes are lightly coupled except modes 3 and 4 in the fixed B.C.

specimen, which correlation reaches almost 0.5. Close inspection of the FRF magnitude in Figure

7.25 indicates that mode 4 is relatively a weak mode and is slightly coupled with mode 3. Therefore

it is concluded that the fitted mode shapes are reasonable and acceptable.

7.10 Finite Element Modeling of Composite Floor System

The objective of this section is to discuss the finite element (FE) modeling approach and

result for composite floor system. The capability of creating valid computational model for dynamic

frequency domain analysis provides a valuable tool for engineers, consultants and researchers to

evaluate the vibration serviceability of in-situ or proposed composite floor. The actual interpretations

of the FE modeled results are still subject of debate, let along various simplified method for evaluating

serviceability. However, without doubt, the ability to create computational models which adequately

represents the dynamic behavior provides a easiest way to avoid vibration issues in the design phase

and is much easier and cost-effective then testing even a small portion of a problematic in-situ floor.

This section discusses the fundamental principles used in developing partial composite finite

230
element models of the tested floor system. The two vibration tests include the identical member

sizes, number of inclined screw and the inclined angle, with the only exception of the boundary

conditions, one is fixed and the other is pinned. The intention of the study is not to independently

creating ideal model for each boundary condition, but rather, to model both floors with fundamental,

logic techniques that allows for future parametric studies of different design parameters. To name a

few, these parameters could include lumber species, composite layup, span length, spacing of girders,

etc.

SAP2000 Version 18 (CSI 2015) finite element software was used in the presented research.

SAP2000 was selected due to its availability, its dynamic finite element capability as well as its

popularity among practicing design engineers, especially when newer versions allows structural model

import from Revit and ETABS. Design engineers are more likely to be more familiar with commercial

FE package like SAP2000 then other packages such as ANASIS, ABAQUS or OpenSees, which gears

more towards research. Their experience maybe limited in the traditional structural static analysis,

however, dynamic model analysis in SAP2000 are very straight forward and user friendly.

Many of the basic techniques used for modeling composite floor in SAP2000 were obtained

from CSI knowledge base, a cluster of user manuals for software developed by Computers and

Structures. Within the documentation, a example of RC slab on girder with shear studs provides

the fundamental knowledge needed for composite section modeling. In addition, technical research

in the field of building component vibration (Davis et al. 2013; Davis 2008; Sladki 1999; Nguyen

et al. 2012; Alvis 2001) further improved the FE model through comparison of the modeling results

with experimental modal measurements.

Creating FE dynamic models require adequately representing mass, stiffness and boundary

conditions to analytically compute frequencies and mode shapes. In SAP200, damping of the re-

spective vibration frequency relates to the material modal damping ratio γ, which is a user input.

Due to the lack of supporting document/studies on composite lumber, damping is not studied in

this section.

The essence of modeling process includes creating the FE model and manually updating the

models to bring the computed modal properties into accordance with the experimentally measured

ones. The floor models are created in the X − Y plane with the negative Z axis represents the

gravity direction. The basic steps of creating FE model of a composite floor system for evaluation

vibration serviceability can be summarized in the following steps:

231
1. Based on the geometry of the composite floor, create grid system for the location

of shell, frame and link elements;

2. Define materials and sections for shell and frame elements; define the correct link

stiffness to represent connections between flange and web;

3. Adjust model to adequately represents mass, stiffness and boundary conditions;

4. Perform SAP2000 modal analysis to compute frequencies and mode shapes;

5. Perform additional parametric study for evaluation of vibration serviceability of

composite floor in a boarder sense.

7.10.1 Assignment of Mass, shell Laminates, Stiffness, and Links

Mass & Composite Laminates Mass is used for the inertia in dynamic analyses. Within

SAP2000, the mass of the individual elements is computed from its volume and density of the mate-

rial, and lumped at its joints and assigned to the translational degrees of freedom. The distribution

of the inertia mass is therefore represented by the mass of the assembled elements.

SAP2000 does not carry a stock glulam library, therefore, the two 24F–V4 Southern Pine

glulam were modeled with user defined rectangular cross section with dimensions (11 in.x 3.125 in.),

mass density (0.02 lb/in2 ), Young’s modules (1.55·106 psi), Poisson Ratio (0.3, typical for pines).

The CLT flanges which attached to the glulams were represented by user defined rectangular

shell elements. It should be noted that the shell elements were selected in lieu of plate elements

since the ability to handle in-plane deformations is a advantage for vibration analysis of composite

floor made of laminated material (CLT in this case). The build-in layer shell elements was used to

defined the stiffness and fiber orientation of 3-ply CLT: two longitudinal layers are oriented parallel

to the X-axis with E0 = 1.55 · 106 psi, the middle transverse layer is rotated 90 degress to be parallel

to Y-axis with E90 = E0 /30 = 5.16 · 104 psi. The location and stacking sequence of the laminate

composite is specified using ply thickness and the distance from its centriod to the neutral axis, see

Figure 7.36 where the top longitudian layer is highlighted in darker blue.

Due to the nature of FE program where the mass and stiffness are distributed among the

elements, it was important to use enough elements to mesh the object, so accurate results can be

obtained. A convergence study was performed to benchmark the composite model, with the criterion

that the mid-span deflection of the beam under 0.4 klf to be within 0.1 in. (see M odelBenchmark).

232
Figure 7.36: User defined layer shell element in SAP2000

It was found that the mesh size of 24 in. along each side of the shell elements gave convergent

results. Similarly, frame elements were meshed along its length at each mesh joint. The layout of

the composite floor model is show in Figure 7.37.

Figure 7.37: Composite floor model. Red: shell elements; blue: frame elements; green: links and
restraints

Stiffness The fundamental concept of hollow core long span system is to create the com-

posite action between wood members to increase its structural mechanics properties. In the presented

study, the composite action between CLT and glulam beams was achieved by mechanical fasten-

233
ers, to be more specific, self-tapping screws installed with a angle. The continuous contact and

the limited slippage between the members therefore provide the composite action of the system.

By modeling both the shell element and the framing element which represent the CLT flanges and

Glulam beams, a series of links must be added to the stiffness property to properly account for the

composite behavior.

In SAP2000, three modeling approaches are available to simulate the composite action of a

build-up member: link, body constraint and joint offset. Among them, link is selected in this study

since the option to adjust the link stiffness allows for the modeling of partial composite actions where

the interface slippage can be made to be known. However, body constraint and joint offset methods

are only capable of modeling full composite action.

To start building up the FE model, frames and shells were drawn at the elevations of their

respective centroids. Rigid links were assigned to connect the shell and frame elements at the grid

lines defined by the actual location of the screws. At this moment, the beam is considered full

composite since all members are connected rigidly and no interface slip is allowed. To introduce the

partial composite action, zero-length links with user defined stiffness were assigned to the location

where shell element and rigid link meet, on both top and bottom CLT. These links were fixed in the

vertical direction (global Z direction), see Figure 7.38. The fixed stiffness in the vertical direction

(local x axis) creates a ‘Hard Contact’ at the interface, preventing elements from intruding into each

other. However, without specifying the link stiffness in other DOFs, the assembly would still act as

non-composite. As mentioned before, link element is a powerful restrictive element where stiffness

of all DOFs can be user specified. For example. to model fully composite action, the links will have

all 6 DOFs fixed, in another word, interface slippage or rotation is not allowed in this case. On the

other side of the spectrum, a non-composite assembly will have links with fixed x direction, while

the rest being all free. Somewhere in between where the interface movement is allowed but limited,

the links will have a stiffness specified with a numerical value in girder longitudinal direction. The

property of links used in the model benchmark is summerized here:

1. Links for fully composite action: fixed all DOFs

2. Links for partial composite action: fixed x direction, specify stiffness in y direction,

and set the rest DOFs free

3. Links for non composite action: fixed only x direction

234
Figure 7.38: Centroids of respective elements

It is recommended to take advantage of the ‘Group’ command in the SAP2000, where

elements can be selectively picked and assign different section properties when needed. As long as

the geometry stays the same, the model can easily accommodate adjustment for parametric study

or any other analysis created by a design engineer.

Model benchmark Prior to the model benchmark, a mesh size convergence study was

performed and concluded that 24 in. by 24 in. shell element size satisfy the convergence criteria.

Therefore, the reported benchmark results were based on this mesh dimension. The benchmark

contained a CLT-Glulam beam with cross section shown in Figure 7.38, span length 40ft, and pin

connection at one end and roller at the other. A uniform load of 0.4 kip/ft was applied at the frame

elements (0.2 kip/ft per glulam) and deflection was measured at the mid-span. Two cases were

considered in the benchmark problem: fully composite and non-composite. Analytical predictions

and SAP2000 outputs are summerized below:

Table 7.7: Benchmark results of the FE model

Benchmark Analytical Prediction SAP2000 Output Percentage Difference


Case (in.) (in.) (%)*
Full Composite 0.723 0.732 1.274
Non Composite 10.850 10.841 -0.079
*The percentage difference=(SAP-Analytical)/Analytical x 100%

The calculated difference shown in Table 7.7 confirmed the SAP2000 user manual (CSI 2015)

where it over-predicted the the fully composite case. It is because the shear deformation is considered

in the deflection while the analytical solution does not. Meanwhile, SAP under-predicted the non-

composite case by fraction of a percentage, possible reasons could be the mesh size or the deflection

235
Figure 7.39: Deflection of a composite beam

measuring point. However, based on the small difference, it was concluded that the composite model

developed was accurate enough for further modal analysis in the presented research.

7.10.2 FEM modal analysis and parametric study

The previous section presented the general steps for creating an FE model of a floor sys-

tem for evaluation of vibration serviceability. Here, FEM modal analysis and parametric study was

performed and the results are evaluated. There are a variety of dynamic case types available in

SAP2000. including multi-step static, modal, response spectrum, time history, moving load, buck-

ling, steady state, and power spectrum density. Modal analysis was the only case type studied

and discussed in the presented study. It should be noted that since the evaluation of vibration

serviceability mainly concerned with the dominant or fundamental frequency, only the first mode

was interested in the presented study. The following paragraph discusses the general procedure for

modal analysis with SAP2000, however, the detailed background on the computational methods

used in the FEM program is best left to the SAP2000 user manual.

SAP2000’s modal analysis of the floor’s FEM model computes the frequencies and mode

shapes based on the assembled mass and stiffness matrices, without consideration of the presence

or level of damping in the structure. The most important aspects of the modal analysis feature for

evaluation of floor serviceability are the computed dominant frequency and the visual shapes of the

vibration for each mode. For the experimental measurements, the lowest modes of a floor system

generally have single curvature, and double curvature is not encountered until higher frequency

modes. For a computed mode shape, the response may be localized to its members, such as isolated

CLT or glulam vibration modes. These localized mode shapes usually come with low amplitude

and are therefore difficult to detect in the experimental measurement with accelerometers attached

only to the top surface of the test piece. Another distinguish difference between the experimental

modes and computed modes lies in the direction of vibration measured. PCB 333B50 is a single

236
CLT-GlulamwithInclinedScrew_V16.sdb 6/13/2017
CLT-GlulamwithInclinedScrew_V16.sdb 6/13/2017

YZ X
YZ X

(a) transverse mode (b) axial mode (compression side on the right)

Figure 7.40: Computed transverse mode and axial mode

axis accelerometer, and it was placed with the arrow up measuring the vertical vibration of the

structure. Therefore all experimental mode shapes are vertical mode shapes. On the contrary,
SAP2000 18.1.0 Deformed Shape (MODAL) - Mode 11; T = 0.00779; f = 128.33923 lb, in, F
SAP2000 18.1.0 Deformed Shape (MODAL) - Mode 6; T = 0.01574; f = 63.54239 lb, in, F
computed modes include mode shapes in all three axes. Due to the higher stiffness in beam’s

transverse and axial direction, their natural frequencies are typically higher than the corresponding

ones on the vertical direction, see Figure 7.40 which shows a second bending mode in transverse

direction a compression–tension mode in the axial direction. The natural frequencies associated

with these mode shapes are beyond the interested range of floor vibration serviceability study and

thus excluded from the presented study. If the need of measuring vibration in all three axes does

present itself, triaxial accelerometer would be a ideal candidate for such purpose. For the FE aided

modal analysis, frequencies were visually inspected along with the computed mode shapes to ensure

consistency in the results interpretation.

SAP2000 offers two methods of modal analysis, eigenvector analysis and Ritz-vector analysis.

Eigenvector analysis determines the undamped free vibration natural frequencies and mode shapes

using only the mass and stiffness matrices to solve generalized eigenvalue problem, whereas Ritz-

vector analysis looks and computes the modes that are excited by a particular applied loading pattern

(CSI 2015). Theoretically, Ritz-vector method yields excellent results in forced response analysis

where spatial distribution of the loading is accounted for. For example, an acceleration load or any

defined load cases can be a form of spatial distribution of loading. As Barrett (2006) indicated, the

Ritz-vector method may be advantageous in filtering unnecessary modes by a starting load vector

in the direction of interest. However, for the plane rigid floor modes, the direction of vibration was

never in question. Eigenvector analysis consistently produced logical mode shapes and this it was

237
used exclusively for the floor model in the presented research.

Following the general steps previously discussed, the floor skeleton model was laid out for

the composite beam. A user defined frame section and area plate element were created to represent

the glulam beams and CLT panel, respectively. The input values of the material used in the model

development are presented in the Table 7.8. Also included in this table are the stiffness properties

obtained from small scale inclined screw shear tests. These stiffness values were assigned to the

zero-length link property in the longitudinal direction along the span length of the structure (global

X in the FE model).

Table 7.8: Composite floor modeling parameters

Timber Mass Properties Link Stiffness Properties


E0 1.55 · 106 psi Static initial (< 0.0600 ) 13.33kip/in
E90 5.17 · 104 psi Static linear (0.0600 ∼ 0.2500 ) 31.47kip/in
Mass density 0.02lb/in3 Static secant 24.33kip/in
Poisson ratio 0.3 Fatigue initial 15.90kip/in

It was found that the use of static linear stiffness (31.47 kip/in) to represent the connection

between flange and web gave the most acceptable natural frequencies when compare to the ex-

perimentally obtained ones. With this configuration, the boundary conditions were modeled using

restraint commands in the SAP2000. For pinned connection, pins were assigned to each side bottom

CLT flange, restraining the CLT only; whereas for the fixed boundary condition, all six DOFs of the

respective glulam beam and CLT ends were restrained. The boundary condition resulted in a clear

span of 39 ft with 6 in overhang on each side, same as the experimental setup during the heel drop

testing. Eigenvector modal analyses were performed on the floor model to compute the frequencies

and mode shapes. The frequencies of the first four modes are presented in Table 7.9 with the values

from experimental modal analysis also included, the corresponding FE mode shapes are presented

in Figure 7.41.

There are some interested items to note in the computed natural frequencies and mode

shapes shown in Table 7.9 and Figure 7.41. First and most obvious one is that the 2nd torsional

mode in the pinned B.C. was picked up in the analytical modal analysis. As a matter of fact, this

mode do and should exist in the experimental FRFs, it most probably falls out of the 0–30 Hz range

of interest and not analyzed in the curve fitting process. Second observation is that the fixed B.C.

238
CLT-GlulamwithInclinedScrew_V16.sdb 6/13/2017
CLT-GlulamwithInclinedScrew_Fixed.sdb 6/13/2017

CLT-GlulamwithInclinedScrew_Fixed.sdb 6/13/2017 CLT-GlulamwithInclinedScrew_V16.sdb 6/13/2017

YZ X
Z
Y X

(a) Fixed–Mode1–7.9Hz (b) Pinned–Mode1–6.1Hz

SAP2000 18.1.0 Deformed Shape (MODAL) - Mode 8; T = 0.16366; f = 6.11041 lb, in, F

CLT-GlulamwithInclinedScrew_Fixed.sdb 6/13/2017 CLT-GlulamwithInclinedScrew_V16.sdb 6/13/2017

SAP2000 18.1.0 Deformed Shape (MODAL) - Mode 5; T = 0.12636; f = 7.91402 lb, in, F

YZ X
Z
Y X

(c) Fixed–Mode2–14.7Hz (d) Pinned–Mode2–13.8Hz

SAP2000 18.1.0 Deformed Shape (MODAL) - Mode 15; T = 0.07262; f = 13.77078 lb, in, F

CLT-GlulamwithInclinedScrew_V16.sdb 6/13/2017
CLT-GlulamwithInclinedScrew_Fixed.sdb 6/13/2017

SAP2000 18.1.0 Deformed Shape (MODAL) - Mode 8; T = 0.06795; f = 14.71778 lb, in, F

Z
YZ X
Y X

(e) Fixed–Mode3–19.8Hz (f) Pinned–Mode3–19.3Hz

SAP2000 18.1.0 Deformed Shape (MODAL) - Mode 14; T = 0.07501; f = 13.3307 lb, in, F

SAP2000 18.1.0 Deformed Shape (MODAL) - Mode 9; T = 0.05959; f = 16.78106 lb, in, F

Z
YZ X
Y X

(g) Fixed–Mode4–29.8Hz (h) Pinned–Mode4–28.0Hz

Figure 7.41: Floor model computed mode shapes

SAP2000 18.1.0 Deformed Shape (MODAL) - Mode 19; T = 0.03561; f = 28.08577 lb, in, F

SAP2000 18.1.0 Deformed Shape (MODAL) - Mode 13; T = 0.0336; f = 29.76494 lb, in, F 239
Table 7.9: Comparison of experimental and computed natural frequencies

Fixed B.C. Pinned B.C.


Experimental (Hz) Analytical (Hz) Experimental (Hz) Analytical (Hz)
Mode 1 7.2 7.9 6.7 6.1
Mode 2 15.8 14.7 15.5 13.8
Mode 3 20.5 19.8 19.4 19.3
Mode 4 24.0 29.8 N.A. 28.0

created a slightly more stiff specimen compared to the pin connection one, causing a increase in all

natural frequencies with corresponding mode shapes. This increase on average is 5.0% and 11.1% for

experimental analysis and FEM analysis, respectively. The third observation is that the computed

natural frequencies are close to the measured ones, with average 14.2% difference in all detected

modes.

Comparison of frequencies and mode shapes was a necessary step in the model validation

process. If the FEM model is further used for force response analysis or other types of transient

analysis, the matching frequencies to experimental measurement will be a pre-requisite for such

application. However in the presented study, the intention of the FE model is to create a quick

assessment tool for floor vibration serviceability analysis that has reasonable level of complexity

and accuracy and can be used by practicing engineers. It is therefore concluded that the the

computed frequencies for these modes and their respective mode shapes are in good agreement with

the measured frequencies of the long span floor system.

For a typical eigenvalue problem, it is known that the characteristic values, i.e. natural

frequencies in the presented research, are dependent on the mass, stiffness and damping matrices.

A general rule of thumb is that the addition of mass decreases the vibration frequency whereas

the addition of stiffness increases the frequency. What sets the task at hand unique to the typical

eigenvalue problems is the introduction of self-tapping inclined screw. These screws not only serve

the connection between flanges and webs, but also provide composite behavior to the floor system.

The stiffness gain resulted from the screw connections directly translates into increased frequencies

since the additional mass was minimal compared to the stiffness gain. Since both experimental

and computational modal analyses indicated a low dominant frequency in terms of human vibration

annoyance perception, it is of great interest to explore the sensitivity of screw stiffness contribution

to the vibration frequencies. In the presented parametric study, only dominant frequency (first

240
bending mode) was investigated since they usually possess the largest magnitude and it happened

to fall inside the annoyance frequency range (4-9.5 Hz).

A front-end program MATLAB was used to control the SAP2000 program through Applica-

tion Programming Interface (API), where the frequency output was recorded during each parameter

iteration. The two FE models (Fixed and Pinned B.C.) were the same as those used in the FE

validation, only with link element stiffness being variable during the parametric iteration. Figure

7.42 shows the analysis results with stiffness values plotted in log scale. Also shown in the figure is

the linear stiffness of inclined self-tapping screw (31.47 kip/in in green dash-dot line) obtained from

small scale shear test, and the DG11/ISO 4-9.5 Hz frequency basin. It is easily seen that the the

floor system provides a poor vibration serviceability since its dominant frequencies fall inside the

undesired annoying frequency range. Based on the provided screw spacing, the dominant frequency

is sensitive to the connection stiffness in the range of 10 to 100 kip/in. In another word, as soon

as the inclined screw engages the connecting members, the stiffness gain resulted from the com-

posite behavior increases the natural frequencies. However, the frequency gain was not substantial:

the model with fixed boundary condition varied from 4.2 Hz at non-composite to 15.0 Hz at fully

composite; whereas the pin connection model varied from 1.86 Hz to 9.44 Hz. While nothing being

fully composite, the use of interface connection didn’t quite bring the floor system into satisfactory

in terms of vibration serviceability, its contribution is limited since the stiffness of the floor system

(CLT and glulam) was under-designed in the first place.

7.11 Vibration Serviceability Assessment and Recommenda-

tions

7.11.1 Assessment

In section 7.3, the initial assessment of floor vibration serviceability discussed the use of

DG11’s simplified hand-calculation where the computation was performed in two steps: frequency

and peak acceleration. Through experimental and analytical modal analysis, the true frequency pro-

file has been identified, and dominant frequency of the grounded specimen was found to be falling

inside the undesired frequency range. The low frequency is sensitive to human organs and may cause

annoyance, frustration or medical conditions that designers should stay away from. To completely

241
Figure 7.42: Frequency gain in link element stiffness

quantify the floor vibration serviceability using DG11/ISO guide, the peak acceleration needs to be

estimated in addition to the discovered frequency profile. Note that the acceleration measurement

specified in the DG11/ISO is expressed in terms of Root Mean Square (RMS) acceleration (Figure

7.2), rather than the peak value during each impact. The RMS acceleration (GRM S ) is the square

root of area under the Acceleration Spectral Density (ASD) vs. frequency curve, it expresses the

acceleration history in a ‘average’ sense and therefore should be less than the value of peak accelera-

tion. The GRM S was computed for each heel-drop response measurement in the frequency range of

0-30 Hz, resulting a total of 780 data points for each test. As one may realize, the magnitude of the

GRM S varies as the DOF closer to the driving point often experience large accelerations. However,

the distribution of GRM S accelerations depicts the true variation of the acceleration response where

the impact source may or may not be direct adjacent to the building occupants at all times. The

GRM S values computed for fixed and pinned B.C. specimens were fitted into various distributions

and it was found that the 2-parameter Weibull distribution well fits the data, see Figure 7.43. A

notable difference observed when comparing the distribution of GRM S reveals that the acceleration

measured in the pinned B.C. specimen is slightly higher than that of the fixed B.C. one. This behav-

242
ior is anticipated since under the same impact, vibration in a higher frequency often accompanies

with a lower magnitude.

Figure 7.43: 2-parameter Weibull distribution fitting for GRM S

Upon obtaining the acceleration profiles, the vibration serviceability of CLT composite floor

system can be further assessed using the DG11 guide. In Figure 7.44, the Weibull probability density

functions (PDF) for GRM S profiles are plotted on the DG11 occupancy comfort chart. The perfor-

mance of the floor system under footstep impact can be therefore evaluated by marking its dominant

frequency and RMS acceleration on the occupancy comfort chart. Since GRM S is best described by

Weibull distribution and not deterministic in nature, their mean values are used as a acceleration

response indicator. In Figure 7.44, the circular dot and the filled diamond shape represent the

vibration performance points of the grounded composite floor for the fixed and pinned boundary

conditions, respectively. It is clearly seen that based on the predefined occupancy comfort level, the

as-build CLT composite beam is considered Acceptable for floor designed for rhythmic activities

and outdoor footbridges. For office or residential occupancy and indoor footbridge, shopping mall,

dancing or dining venues, the performance points may lead to the conclusion of unsatisfactory per-

formance. However, the grounded composite beam tested was at its bare minimum, without dead

load and non-structural components. The additional mass and associate damping may dampen out

the response and therefore decrease the RMS acceleration. In addition, the boundary condition used

in the experimental modal analysis consider a single, separated floor beam; whereas in forming a

complete floor diaphragm, a series of composite beams are to be parallelly connected and share the

load together. The real boundary condition may result in a more stiff floor system and therefore

243
drives up the frequency and RMS acceleration. Field measurement of in-situ mass timber floor

dynamic properties is recommended for future research. Since the boundary condition, along with

true loading on the floor, can be considered in the performance interpretation.

Figure 7.44: Vibration serviceability assessment for CLT-Glulam composite beam floor system

As discussed earlier in the chapter, it is the easiest to tackle the problematic floor in the

design phase, rather than performing a costly and time consuming rehabilitation project for a in-situ

floor system. There are a handful of solutions that can be used to improve its performance. For

stiffness requirements, selecting deeper and/or wider glulam beams may drive up the frequency as

well as reducing the response magnitude. Another common practice for the finish of engineered

wood subfloor is to pour a thin underlayment of concrete or gypcrete. It is extensively used in the

condominium, apartments and hotel construction for sound reduction, fire ratings, radiant heating

and floor leveling. The light weight cement based underlayment with dowel connections provides

244
additional stiffness gains as it serves the same purpose of increasing the beam’s stiffness without

introducing tremendous of mass. As far as the acceleration response, the as-tested floor performed

poorly since it was a bare floor and the damping was minimum. The finished floor system with various

underlayment, bearing and partition walls as well as occupational contents would definitely cushion

the acceleration response. Future research should look into the viability of using floor underlayment,

such as dense board, rubber underlayment, acoustic board etc, to improve the impact and acoustic

performance of composite mass timber floor system.

7.11.2 Recommendations

This chapter covers the preliminary vibration serviceability assessment using DG11 guide,

the experimental modal testing of a ground composite floor system and its data analysis and results

interpretation. Similarly, the recommendations will be addressed in those areas.

Criteria used in assessment of floor vibration serviceability If the true stiffness of

the floor system can be accurately estimated, DG11’s simplified hand calculation provides a reliable

estimates on fundamental frequency and acceleration response. However, it is only recommended

for preliminary vibration serviceability analysis since it was one of the least stringent criteria existed

in the literature. DG11 essentially limits the acceleration response within each occupancy category

to a concave curve, with its flat lowest basin at 4 to 9.5 Hz. There is no absolute frequency limit set

for the floor system.

The DG11 guideline may not be directly applicable to a mass timber floor system since

timber is much lighter than that of a steel joint concrete floor and the lack of mass could result in

excessive acceleration response. In addition, for a lighter system, it is much easier to excite harmon-

ics resonance of walking excitations, which are the integer multiply of occupancy step frequency. If

the floor system’s fundamental frequency can be designed far above the occupants’ activity forcing

frequency, regardless whether it is walking or rhythmic activities, the acceleration response will be

drastically decreased as the harmonic diminishes exponentially with the increasing integer multiply.

Therefore it is recommended to apply a lower fundamental frequency boundary to limit the harmonic

response, and this limit can vary with the indented occupancy. For example, the 2005 Canadian

National Building Code specifies that the light weight joisted floors, the minimum natural frequency

shall not be less than 15 Hz. This can be a good start for vibration design of mass timber floor

245
system, however future research should investigate occupants comfort sensitivity to higher funda-

mental frequency. The second recommendation is that the additional mass and stiffness provided by

cement based topping, such as light weight concrete or gypcrete, may improve occupants’ comfort

level as they feel ‘less’ vibration.

Experimental modal analysis The capacity of the experimental modal analysis largely

depends on the availability of the testing equipment. However, simply having the appropriate

equipment and ample time does not guarantee high quality measurement. It is the combination of

proper selection of equipment, continuous monitoring of the data throughout the testing, along with

the reliable parameter extraction techniques that yields the acceptable vibration characteristics of

the floor system. The following is a list of recommendations that includes the planning, conducting

and data analysis of referenced modal testing, to ensure satisfactory measurement quality and results

interpretation:

1. Identify the overall objectives of the investigation. The objectives vary

from project to project as it may simply involve a rough estimation of floor natural frequency

or a complete set of FRF measurements for forced vibration analysis. These objectives need

to be clearly identified prior to any other decisions such as selection of excitation source or

equipment purchase.

2. Select proper testing method and allot time and equipment. Based on

the intended modal analysis purposes, select testing (referenced or unreferenced) and obtain

test equipment. The selection of equipment relies extensively on the decision of the excitation

source and data analysis software/hardware. The associated cost varies significantly from

a in-house made force plate plus software to a electrodynamic shaker with fully functioned

spectrum analyzer. In the presented study, the heel-drop and force plate were selected as

the excitation source, as it is considered to be an economic setup to perform basic reference

modal analysis. However, the heel-drop impact only provides adequate energy in the 0-30 Hz

frequency range and is only suitable for exciting light test piece with small to moderate area

coverage. For advanced testing purposes and/or larger area coverage, electrodynamic shaker

will be ideal albeit its higher initial cost.

3. Estimate test parameters and allot ample time for the testing. Upon

246
the section of testing equipment, the recorded length and sampling frequency should be be

determined since these are drive by the desired frequency domain resolution. It is strongly

suggested to perform a few measurements and examine the signal quality before fill effort of

measurement. Auto, cross spectrum and coherence functions are among the key plots that

need to be continuously monitored during the signal measurement process.

4. Conduct floor vibration testing with the recommended practice and

expect the unexpected. Depending on the testing environment, the number of averages

need to be determined on the spot. A perfect balance between signal quality and time needed

for the test is what experimentalists strive for during the test. Extreme caution should be

taken when handling the cables as they tends to tangle together. Never force a cable since a

kink would irreversibly destroy it. Keep a spare cable whenever possible to avoid any delay or

possible frustration.

5. Post processing. The last process involved in the experimental modal testing

which explained in the least detail in the presented study is the post processing and curve

fitting process. A full suite of parameter extraction often include frequency, damping and

visually display mode shapes. Depending on the complexity of the measured FRFs, different

approached can be applied to estimate the parameters.

Experimental modal analysis is a entertaining and rewarding study activity as it often

requires decision making for compromises. For first timers, it may be difficult to initiate the test,

however, it is more of a learn by doing type of activity. The experimental testing method and

practice presented in this chapter provide an accessible experience for the dynamic modal testing

for floor systems.

247
Chapter 8

Conclusion

8.1 Summary of Research

With the surging demand for mass timber construction, the presented dissertation discussed

various aspects of the structural performances and application of cross-laminated timber (CLT)

panels. There were two phases to this research. In phase one, a pilot-scale study was carried out

to evaluate the manufacturing process of CLT panels using Southern Yellow Pine, a group of tree

species native to the Southeast US. A series of small-scale tests were conducted to verify the adhesive

and structural performances of the CLT panels for flexure and rolling shear.

The small-scale test specimens consisted of a combination of laboratory and commercially

manufactured CLT panels. The test results generated from the pilot scale study were used in the

second phase of this study where the concept of a composite CLT-Glulam beam, suitable for long

span applications, was proposed. The proposed composite section takes the geometry of a hollow

core composite beam, in which the CLT panels and glulams were fasterned using inclined screws.

To investigate the feasibility the proposed composite beam for floor application, a non-destructive

vibration test and a destructive structural test were performed on a full-scale 40 ft. long composite

beam. The selection of a suitable timber-to-timber joint connection, the design and assembling of

the composite beam were discussed in this dissertation. The destructive test in full-scale verified

the composite beam can safely carry over 100 psf of live load and 20 psf of dead load. In addition

to destructive test, a non-destructive modal test was performed utilizing an instrumented heel-drop

excitation and acceleration response to quantify the vibration performance of the composite floor.

248
8.2 Summary and Findings of Research Tasks

Following are the major findings of the presented research:

Chapter 4: Rolling Shear The rolling shear strengths of Southern Pine CLT made using

four different adhesives (MF, PRF, EPI and PUR) were evaluated experimentally. Rolling shear

(horizontal shear) failure in the CLT refers to the peeling of the crosswise layers when subjected to

out-of-plane load. It could lead to premature CLT panel failures as the rolling shear cracks often

propagate to the glue lines and compromise the glue bond. The rolling shear failure is typical to

many laminated materials with orthotropic plies. It was found that with proper application of the

adhesive, Southern Pine lumber can be used to produce CLT with rolling shear resistance that meets

and exceeds the PRG-320 minimum requirement for V3 layup.

Chapter 5: Bending Strength and Stiffness The bending stiffness (MOE) and strength

(MOR) were evaluated experimentally for Southern Pine CLT panels made in a laboratory and

commercial manufacturing plant. The test results show that Southern Pine can be used to produce

CLT panels that meet the reference flexural properties specified in the PRG-320 standard. However,

it was observed that the presence of finger joints and lumber defects (e.g. knots) in high flexural

demand region, namely mid-span of the bottom tension layer, can severely affect the bending strength

of the CLT panel. Lumbers with minimum defects (e.g. No.1 or MSR) are recommended to be used

in the tension layer. In addition, avoiding having finger joints located in high tension region is also

recommended.

A Monte-Carlo simulation model was developed to estimate the bending strength of CLT

numerically. The model uses a bending-tension interactive failure criterion to predict the bending

strength based on lumber species, grades, thickness, and lay-up. The developed simulation model

may be used by CLT manufacturers to design layups and optimize the CLT panel for bending

strength.

Chapter 6: Bending and Shear Strength of Composite CLT-Glulam Beam An

investigation of timber joint fastened with inclined screws was conducted to select a suitable connec-

tion method for attaching the CLT panels to the glulam beams. Inclined screws installed at a 30-deg

angle, measured from the surface of the wood, were used in the CLT-glulam joints. The strength of

the inclined screw was predicted using two analytical models and the predictions were found to be

slightly higher than that of the actual strengths measured from two small-scale shear tests of the

249
30-deg inclined screw joints.

A 5 ft. wide and 40 ft. long CLT-glulam composite floor system was constructed and tested

in full scale to validate the as-designed bending and shear strengths. It was found that the composite

floor system designed in accordance to the NDS provisions has adequate capacities in bending and

shear to safely carry the design loads (100 psf live + 20 psf dead) with overstrength factors of 2.5 and

3.6 for bending and shear, respectively. In addition to the ultimate strength, the deflection of the

composite beam was also evaluated under service level loads. It was found that, when the composite

beam was loaded as a standalone element, the beams deflection under the service live load (2.60 in.)

exceeded the L/240 (1.95 in.) limit. Note that this the bending test was conducted under a very

pessimistic boundary condition. In a real floor system, multiple composite beams are connected side

by side to form the floor. The plate action is expected to increase the overall stiffness of the floor.

All in all, the design of the composite beam was found to be limited by the serviceability, namely

deflection, and not strength.

Chapter 7: Modal Testing and Analysis When designing a long span element, the

vibration typically controls the member design. A series of non-destructive modal and vibration

tests were carried out to experimentally quantify the vibration characteristics of the CLT-Glulam

composite beam, which included the natural frequencies, mode shapes and associated damping ratios.

It was found that the acceleration response exceeded the recommended serviceability requirement for

typical office/classroom occupancy. However, consider the pessimistic boundary condition used in

the experimental study, and the absence of non-structural components and finishes, the acceleration

in a realistic fully finished floor system is expected to be lower than the recommended lower bound

for office/classroom occupancy.

8.3 Limitations and Recommendations for Future Research

During the flexural experimental studies of both small-scale and full-size specimens, it was

found that the bending strengths were affected due to the presence of finger joints in tension critical

region. The CLT flexural capacity can potentially be improved with the enhanced performance

of finger joints. In this regard, it is recommended that future research be focused on the finger

geometry, adhesive selection, curing process, and strategic placement of the finger joint locations, in

the aim to improve the flexural and tensile strength of the CLT panels.

250
The vibration characteristics (frequency and acceleration response) of the floor system are

heavily influenced by its mass, stiffness, and boundary condition. As stated previously, the com-

posite beam was tested under a very pessimistic condition. Addition of non-structural components

and continues bracing along the span direction could reduce significantly the vibration. It is recom-

mended that in-situ vibration test be conducted on CLT floors in full-scale buildings.

The composite beam was tested as an isolated element in this study. Future research should

examine the connection and load sharing behavior between adjacent composite beams. Lap/butt

joint or perforated plate reinforced with wood screws may be a feasible connection solution. In

addition, the bottom tension flange may be omitted to form double-T composite wood beam, as

long as the webs are structurally detailed and braced along the length, and have adequate capacity

to carry the design loads. Future research may explore the use of double-T composite wood beam.

Finally, a cost-benefit study is recommended for future work to evaluate the potential mar-

kets for using the proposed long-span composite CLT-glulam beam in lieu of steel and concrete.

251
Appendices

252
Appendix A Pictures of Major Strength Direction Finger

Joint and Lumber Knot Failures

Figure 1: Specimen1-2 FJ and 1 knot failures

Figure 2: Specimen2-1 FJ and 1 knot failures

253
Figure 3: Specimen3-1 FJ failure

Figure 4: Specimen4-2 FJ failures

Figure 5: Specimen5-1 FJ and 2 knot failures

254
Figure 6: Specimen6-2 knot failures

Figure 7: Specimen7-1 FJ and 2 knot failures

Figure 8: Specimen8-1 FJ and 2 knot failures

255
Figure 9: Specimen9-1 FJ and 2 knot failures

Figure 10: Specimen10-2 FJ failures

256
Appendix B Screw Joint Capacity Estimation

This page is intentionally left blank.

257
1/3
Appendix A. Strength Estimation of Shear-Tension Inclined Screw
Joint
Problem Statement: 3/8'' diameter 15 3/4'' length ASSY VG Cylindrical screw is used in a shear tension joint for a
CLT-glulam composite beam. Screw is installed 30 degree to the plane of side member. CLT is a 3-ply South Pine
V3, glulam all No.2 lumbers.

Estimate the joint capacity.

Method 1. Using method from ref. [1] and [2]

Many of the ASSY screw properties can be found in the European Technical Approval ETA-13/0029. Note the prov
table contains characteristic values.
(http://x.etadanmark.dk/danish/eta/pdf/ETAer/ETA130029%20Wuerth%20ASSY%20Plus%20VG%20r1.pdf)

d  0.375in  9.525  mm use 10mm in the table

M yk  36N  m

Dowel bearing strength from NDS Table 11.3.3

Fpara  6150psi Fperp  4200psi

Bearing strength at an angle


Fpara  Fperp 3
Fθ   5.51  10  psi
2 2
Fpara   sin
π    π 
   Fperp   cos  
  6    6 

258
2/3
d 1  0.248in core diameter
factor 1.1 is recommended if the plastic hinge developed in the
d eff  1.1  d 1  0.273  in threaded part from reference [1]

Rv  2  M yk  Fθ  d eff  0.979  kip Lateral load carrying capacity

Lag screw withdraw design value (NDS table 11.2A)


lbf
W  307  15.75 in  4.835  kip withdraw strength
in
W
Rax   2.93  kip inclined screw withdraw strength
2 2
1.2   cos    1   cos  
π π
  6    6 

μ  0.26 kinetic friction

R0  Rax   μ  sin
π  π    π  π 
  cos    Rv   sin   μ  cos    3.188  kip
 6  6   6  6 

Bring R to strength level


KF  3.32

This is the strength value of shear capacity, per screw, calculated


RR  3.32  R0  10.583  kip
using method from ref. [1] and [2].

Method 2. Using method from ref. [3]

d  0.375  in

4.125in
S1   4.763  in Embedment length in side member
cos
π

6

S2  15.75in  S1  10.987  in Embedment length in main member

ASSY withdraw allowable strength

from the influence of end- and edge distance on the loading carrrying capacity in joints with
fax30k  4.5MPa
inclined screws. By Pernilla Malborg. The data is from Hasson and Thelandersson (2003)
listedin table 2.4

259
0.2
fa1d  fax30k  
8  d
  595.03  psi 3/3
 S1 
0.2
fa2d  fax30k  
8  d
  503.435  psi
 S2 

ASSY design tension capacity


π   d 1
2
Fud  136.6ksi   6.598  kip from ASSY screw design guide
4

ASSY head pull through Fhead  130lbf

Rtd  minFud fa1d  π  d  S1  Fhead fa2d  π  d   S2  d   3.469  kip

Kinetic friction coefficient μ  0.26

Rd  1  Rtd   cos
π  π 
  μ  sin    3.455  kip
 6  6 

Bring R to strength level


KF  3.32

This is the strength value of shear capacity, per screw, calculated


RR  3.32  Rd  11.471  kip
using method from ref. [3]

References:
[1] Design model for inclined screws under varying load to grain angles. By R Jockwer, Rene Steiger and Andrea Fran
[2] Joints with inclined screw. By I. Bejtka and H.J. Blass. CIB-W18/35-7-4
[3] Joint with inclined screw. By Ari Kevarinmaki. CIB-W18/35-7-3

260
Appendix C Composite Beam Capacity Estimation

This page is intentionally left blank.

261
1/8

Appendix B: Capacity of CLT Long Span Composite Beam

Modified June 29.2017


The design capacity of hollow CLT is calculated by the literature: Estimation of bending stiffness, moment carrying
capacity and internal shear force of SUGI CLT panel. By Minoru Okabe.
The following numbers can be found in C:\Users\mengzhg\Desktop\My project\Meeting Journal\Meeting 04.29.16\CLT
composite section property.xlsx

Section Properties
Calculate (EI)eff for Composite Floor Element 5ft wide_based on Minoru Okabe, Motoi Yasumura, Kenji Kobayashi
Layer E (ksi) Z (in) b (in) h (in) bh3/12 (in4) AZ2 (in4) Ii (in4) EI (kip in2)
1 1400.0 8.938 60.000 1.375 13.0 6590.0 6603.0 9244210.938
2 46.7 7.563 2.000 1.375 0.4 157.3 157.7 220793.4896
3 1400.0 6.188 60.000 1.375 13.0 3158.5 3171.5 4440132.813
4 1700.0 0.000 6.250 11.000 693.2 0.0 693.2 1178489.583
5 1400.0 6.188 60.000 1.375 13.0 3158.5 3171.5 4440132.813
6 46.7 7.563 2.000 1.375 0.4 157.3 157.7 220793.4896
7 1400.0 8.938 60.000 1.375 13.0 6590.0 6603.0 9244210.938
Sum 746.1 19811.6 20557.7 28988764.1

4 4
IHMT := 2.056  10 in This is the moment of inertia for composite CLT floor with conventional LCL layup.
IHMT 3 3
SHMT := = 2.136  10 in
9.625in

Eref := 1400ksi

7 2
EIHMT := 2.89887641  10 kip  in

262
2 2/8
AHMT := 4.125in  5 ft  2 + 3.125in  11in  2 = 3.915  ft

Calculate first mement of area (Q), neural axis at mid height


in 3 3
Q := 1.375in  60in  ( 1.375in  2.5 + 5.5in) + 1.375in  60  ( 1.375in  1.5 + 5.5in) ... = 1.269  10 in
30
+ 1.375in  60in   1.375 + 5.5in
in
 2 

L := 39ft

Loadings

Office building loading:ground floor or corridor: 100 psf live and 20 psf dead

lbf
35  L  AHMT
3
ft
SW := = 137.023  plf
selfweight L

dead load DL := 20psf  5 ft = 100  plf

live load LL := 100psf  5 ft = 500  plf

( SW + DL)
1.4  = 553.053  plf
0.6

[ 1.2  ( SW + DL) + 1.6  LL] 3


= 1.356  10  plf controls λ := 0.8
0.8
3
Wu := 1.2  ( SW + DL) + 1.6  ( LL) = 1.084  10  plf Factored demand

Simply supported at both ends, calculate shear and moment


L
VLRFD_demand := Wu  = 21.146  kip
2

2
Wu  L
M LRFD_demand := = 206.177 kip  ft
8

Check for bending-full composite


LRFD bending capacity, ignore all adjustment factors

NDS reference design value fb := 1050psi ft := 650psi fv := 175psi

KFb := 2.54 KFt := 2.70 KFv := 2.88

263
ϕb := 0.85 ϕt := 0.80 ϕv := 0.75
3/8

2
Wu  L
M LRFD_capacity := fb  KFb  λ  SHMT  ϕb = 322.829  kip  ft > M LRFD_demand := = 206.177 kip  ft
8

Check for shear-full composite


t web := 3.125in  2 = 6.25  in d overall := 19.25in

2
VLRFD := fv  KFv  λ  t web  d overall   ϕv = 24.255  kip > VLRFD_demand = 21.146 kip
3

This is an approximation for the I beam, reference see NS10-1beamstress.pdf

Plot shear and moment diagrams


x := 0in , 0.1in .. L

Vu( x) := ( VLRFD_demand - Wu  x) Max shear is at both supports

30
20
10
Vu( x)
0
kip
- 10
- 20
- 30
0 10 20 30 40
x
ft

Vmax := Vu( 0) = 21.146 kip Shear check

Moment Diagram
x

M u( x) :=  Vu( x) dx

0

264
300
4/8
200
M u( x)
100
kipft
0

- 100
0 10 20 30 40
x
ft

VLRFD_demand
Max momnet location is when shear=0 Xvo := = 19.5  ft
Wu

M max := M u( Xvo) = 206.177  kip  ft Moment check

Check shear for glulam beams:


For 24F V4 Southern Pine glulam, reference design values:

6 6
E := 1.7  10 psi Emin := 0.9  10 psi

Fb := 2400psi FbNeg := 1650psi

FVx := 175psi FcP := 740psi

Common Adjustment Factors NDS Supplement Table 5A

CM := 1.0 Moisture
Ct := 1.0 normal temperature
Cvr := 1.0 shear reduction factor

KFv := 2.88
ϕv := 0.75
Cvr := 1.0

Adjusted shear strength

λ = 0.8
Fvx_prime := λ  ϕv  KFv  FVx  CM  Ct  C vr = 302.4  psi

Vnx := Fvx_prime  t web  d overall = 36.383  kip Vmax = 21.146  kip Glulam beams are able to
carry shear alone

265
5/8

Spacing For Inclined Screws


Calculate corresponding shear flow in the joint, the spacing changes as the shear is different along the span.

3 3
Q = 1.269  10  in
4 4
IHMT = 2.056  10  in

Vu( 0ft)  Q
τ0 := = 208.766  psi for two glulams surfaces at support
IHMT  3.125in  2

LRFD Capacity for 1 ASSY VG 3/8'' installed at 30 degree RASSY := 8.135kip

RASSY
Spacing := = 6.235  in at support
τ0  3.125in  2

RASSY
Spacing( x) :=
Vu( x)  Q
 3.125in  2
IHMT  3.125in  2

Provide the following spacing along the beam: (symmetric)

S( x) := 4in if 0ft  x < 5ft


6in if 5ft  x < 13ft
L
10in if 13ft  x <
2

The profile of the inclined screw needs discussion and second opinion. But now check this: Assuming the third-point
bending test will achieve the same LRFD moment, design the shear for the specimen.

M LRFD_capacity = 322.829  kip  ft

L
Clear Span L = 39  ft Constant moment span = 13  ft
3

266
M LRFD_capacity
Reaction := = 24.833  kip 6/8
L
3
L
Vu3( x) := Reaction if 0ft  x <
3
L 2L
0kip if x< Max shear is at both supports
3 3
2L
-Reaction if x<L
3

30
20
10
Vu3( x)
0
kip
- 10
- 20
- 30
0 10 20 30 40
x
ft

Moment Diagram

L
M u3( x) := Reaction  x if 0ft  x <
3
L L 2L
Reaction  if x<
3 3 3

- Reaction   x -
L 2L  2L
Reaction   if x<L
3  3  3

400

300

M u3( x)
200
kipft

100

0
0 10 20 30 40
x
ft

267
RASSY L
Spacing3( x) := if 0ft  x < 7/8
Reaction  Q 3
 3.125in  2
IHMT  3.125in  2
L 2L
100in if x<
3 3
RASSY 2L
if x<L
Reaction  Q 3
 3.125in  2
IHMT  3.125in  2

40
Maximum spacing allowed: uniform load
35
Provided spacing
Spacing of inclined screw

30

25

20

15

10

0
0 4 8 12 16 20

Span (ft)

Check for deflections

SW = 137.023 plf

DL = 100 plf

LL = 500 plf

4
5 ( LL)  L L
Live load only deflection ΔLL := = 0.898 in = 1.3 in
384  EIHMT 360

Long term deflection Kcr := 1.5 structural composite lumber


4
5 ( DL + SW )  L
ΔLT := = 0.426 in
384  EIHMT

4
5 ( LL)  L
ΔST := = 0.898 in
384  EIHMT
L
= 1.95 in
ΔT := Kcr  ΔLT + ΔST = 1.536 in 240

268
Appendix D Frequency Domain Response of a Multi-Degree-

of-Freedom System under Harmonic Excitation

The finite element equation of a N degree of freedom viscously damped linear structure in

a matrix notation can be found in many textbooks, including that of Clough and Penzien (1993):

[M ]{ẍ} + [C]{ẋ} + [K]{x} = {f } (1)

where the square symmetric N xN matrices [M ], [C] and [K] contains the structure’s physical mass,

damping and stiffness characteristics, respectively. The N x1 vectors {f } = {f (t)} and {x} =

{x(t)} contain the input force applied to the structure and the resulting motion due to these forces,

respectively. The dots denote differentiation with respect to time.

Consider a case where the external force vector {f } is composed by a single harmonic force

that is applied at a single point (q th coordinate) on the structure. Mathematically,

{f } = {f0 }eiωt (2)

where

{f0 } = {0...fq ...0}T (3)

The time domain solution of Eq. (1) for the excitation of Eq. (2) is obtained by first un-

coupling the structures’ equation of motion through the following coordinate transformation (Ewins

2000):

{x} = [Φ]{y} (4)

Eq. (4) represents a linear transformation that relates the structure’s displacement in the

spatial model {x} to that in the modal model. This coordinate transformation is mapped by the

undamped mode shape matrix [Φ].

Substituting Eq. (2) and Eq. (4) into Eq. (1) and pre-multiply both sides of the equation

by [Φ]T results in:

[Φ]T [M ][Φ]{ÿ} + [Φ]T [C][Φ]T {ẏ} + [Φ]T [K][Φ]T {y} = [Φ]T {f0 }eiωt (5)

269
Eq. (5) reduces to a set of N uncoupled single degree of freedom equations of motion

on the modal coordinates if and only if the damping matrix [C] obeys the following orthogonality

requirement:

[Φ]T [c][Φ] = [Cr ] (6)

where the N xN [Cr ] diagonal matrix contains the structure’s modal damping coefficients. This

orthogonality requirement can be fulfilled if a proportional damping is assumed. In this case, the

damping matrix is expressed as the combination of the mass and stiffness according to Clough and

Penzien (1993):

X
[C] = [M ] ab [[M ]−1 [K]]b (7)
b

where ab is an arbitrary constant and b is an integer. All damping matrices that are defined based

on Eq. (7) are diagonalized by the same transformation that diagonalized [M ] and [K]. A particular

and widely employed case of proportional damping is the so called Rayleigh damping that is obtained

by setting b = 1 into Eq. (7). This gives:

[C] = a0 [M ] + a1 [K] (8)

Pre and post multiplication of both sides of Eq. (8) by [Φ]T and [Φ] gives the following

results for [Cr ]:

[Cr ] = [Φ]T [C][Φ] = a0 [mr ] + a1 [kr ] (9)

where the orthogonality relationship for modal mass [mr ] and modal stiffness [mk ] in the modal

model Eq. (7.35b) is used in the right hand side of Eq. (9):

[Φ]T [M ][Φ] = [mr ], [Φ]T [K][Φ] = [kr ] (10)

If proportional damping is assumed, Eq. (5) reduces to:

mr y¨r + cr y˙r + kr yr = [Φ]Tr {f0 }eiωt (11)

270
Division of both sides Eq. (11) by the modal mass results in an alternative form for the rth

equation on motion in the modal space:

y¨r + 2βr ωr y˙r + ωr2 yr = µr eiωt (12)

where βr is the modal damping ratio of the rth mode shape and is defined in terms of the corre-

sponding rth mode mass, stiffness and damping coefficient:

µr
βr = √ (13)
2 kr mr

The constant µr appears on the right hand side of Eq. (13) is the input force modal

participation factor and is defined as:

1
µr = {φ}Tr {f0 } (14)
mr

This constant shows how the structure’s rth mode shape interacts with the spatial distribu-

tion of input force that is applied on the structure {f0 }. When µr = 0, Eq. (12) presents a travial

solution for all values of the excitation frequency ω for zero initial conditions. Physically, this means

that the rth mode of vibration has no participation in the structure’s response to this particular

excitation.

If there is at least one non-zero µr , the time time domain solution of Eq. (11) is given as:

yr = Yr eiωt (15)

where Yr is the unknown rth mode amplitude. Substitute Eq. (15) into Eq. (11) yields the following

solution for Yr :

µr
Yr = (16)
ωr2 − ω2 + 2jβr ωr ω

The time domain solution is obtained simply by back substituting Eq. (16) into Eq. (15):

µr
yr = eiωt (17)
ωr2 − ω 2 + 2jβr ωr ω

The spatial domain displacement of a MDOF system now can be related to yr using the

271
modal transformation matrix again:

N
X
{x} = {φ}r yr (18)
r=1

Eq. (18) is frequently referred as modal superimposing (Craig 1981) since it indicates that

the finial solution x is obtained by superimposing (summing) the contribution of each mode shape

individually to the total displacement vector. Substituting Eq. (17) in Eq. (18) and use the definition

of µr leads to:

n
X {φ}r {φ}Tr {f0 }/mr iωt
{x} = e (19)
ω 2 − ω 2 + 2jβr ωr ω
r=1 r

From the time domain solution in the spatial mode, the frequency domain displacement in

the response model can be obtained as:

n
X {φ}r {φ}Tr {f0 }/mr
{X(ω)} = (20)
ω 2 − ω 2 + 2jβr ωr ω
r=1 r

At this point, Frequency Response Function (FRF) can be calculated by computing the

division of spectral displacement and spectral force, respectively. Mathematically in terms of Re-

ceptance:

X(ω)
{R(ω)} = (21)
F (ω)

Express Eq. (21) in a vector notion gives the fundamentally important Receptance FRF

solution of MDOF system incorporated with Raleigh damping:

n
X φir φkr /mr
Rik = (22)
r=1
ωr2 − ω 2 + 2jβr ωr ω

When the excitation is coincide with the response location, i = k, Rij is called the driving

point receptance FRF. In this case, Eq. (22) can be written as:

n
X φ2ii /mr
Rii = (23)
ω2
r=1 r
− ω 2 + 2jβr ωr ω

and when i 6= k, Rik is called transfer Receptance FRF.

Two additional FRF presentations which commonly employed in the modal testing are the

272
Mobility and Accelerance FRF. If the velocity response per input force is used, this is called the

Mobility and the notation Y (ω) is used. Similarly, if the acceleration response per input force is

used, this is called the accelerance and the notation used is A(ω). The relationship between the

Mobility and Accelerance FRF to the Receptance FRF are given by:

Y (ω) = ωjR(ω), A(ω) = −ω 2 R(ω) (24)

Therefore, in the most common case where accelerometers are used in the modal testing,

the accelerance FRF which contains the structure’s modal information is given by substitute Eq.

(24) into Eq. (22):

n
X −ω 2 φir φkr /mr
Aik = (25)
r=1
ωr2 − ω 2 + 2jβr ωr ω

where N is the number of modes; ωr is the circular natural frequency of the rth mode; mr is the

modal mass of the rth mode; βr is the modal damping ratio of the rth mode; φir and φik are the

corresponding i or k DOF term of the rth mode.

Eq. (25) is considered as the classical physical based FRF in the generalized coordinate

space. Rather than directly approximate this equation with measured FRFs, different mathematical

models were proposed for much faster fitting algorithm.

273
Bibliography

David E Allen and JH Rainer. Vibration criteria for long-span floors. Canadian Journal of Civil
Engineering, 3(2):165–173, 1976.
David E Allen, JH Rainer, and G Pernica. Vibration criteria for assembly occupancies. Canadian
Journal of Civil Engineering, 12(3):617–623, 1985.
Matthew S Allen and Jerry H Ginsberg. A global, single-input–multi-output (simo) implementation
of the algorithm of mode isolation and application to analytical and experimental data. Mechanical
Systems and Signal Processing, 20(5):1090–1111, 2006.
Steven Robert Alvis. An experimental and analytical investigation of floor vibrations. PhD thesis,
Virginia Tech, 2001.

M Omar Amini, John W van de Lindt, Shiling Pei, Douglas Rammer, Phil Line, and Marjan
Popovski. Overview of a project to quantify seismic performance factors for cross laminated
timber structures in the united states. In Materials and joints in timber structures, pages 531–
541. Springer, 2014.
M Omar Amini, John W Van de Lindt, Douglas Rammer, Shiling Pei, Philip Line, and Marjan
Popovski. Determination of seismic performance factors for clt shear wall systems. In Proceedings
of the World Conference on Timber Engineering. Vienna, Austria, 2016.
BDO Anderson and JB Moore. Optimal filtering dover publications. 2005.
Mats Andersson, Linda Persson, Mikael Sjöholm, and Sune Svanberg. Spectroscopic studies of
wood-drying processes. Optics express, 14(8):3641–3653, 2006.

Albino Angeli, Maurizio Piazza, Mariapaola Riggio, and Roberto Tomasi. Refurbishment of tradi-
tional timber floors by means of wood-wood composite structures assembled with inclined screw
connectors. In 11th World Conference on Timber Engineering–WCTE, pages 20–24, 2010.
APA ANSI. Ansi/aitc a190.1–2012 standard for structural glued laminated timber. ANSI/APA,
Tacoma, US, 2012a.

APA ANSI. Ansi/apa prg 320–2012 standards for performance-rated cross-laminated timber.
ANSI/APA, Tacoma, US, 2012b.
Canadian Standards Association et al. Qualification Code for Manufacturers of Structural Glued-
laminated Timber [electronic Resource]. Mississauga, Ont.: Canadian Standards Association,
2006.
National Fire Protection Association et al. NFPA 5000: Building construction and safety code. The
Association, 2005.

274
ASTM. D2915-10. Practice for Sampling and DataAnalysis for Structural Wood and WoodBased
Products. ASTM International, West Conshohocken, PA, 2010.
ASTM. ASTM D2718-00.Standard Test Methods for Structural Panels in Planar Shear (Rolling
Shear). ASTM International, West Conshohocken, PA, 2011.

ASTM. ASTM D3737-12.Standard Practice for Establishing Allowable Properties for Structural
Glued Laminated Timber (Glulam). ASTM International, West Conshohocken, PA, 2012a.
ASTM. D255912ae1. Standard Specification for Adhesives for Bonded Structural Wood Products for
Use Under Exterior Exposure Conditions. ASTM International, West Conshohocken, PA, 2012b.
ASTM. ASTM D4761 - 13.Standard Test Methods for Mechanical Properties of Lumber and Wood-
Base Structural Material. ASTM International, West Conshohocken, PA, 2013.
ASTM. ASTM D198-15.Test Methods of Static Tests of Lumber in Structural Sizes. ASTM Inter-
national, West Conshohocken, PA, 2015.
ASTM. D1990-16. Standard Practice for Establishing Allowable Properties for VisuallyGraded Di-
mension Lumber from InGrade Tests of FullSize Specimens. ASTM International, West Con-
shohocken, PA, 2016a.
ASTM. E119-16A. Standard Test Methods for Fire Tests of Building Construction and Materials.
ASTM International, West Conshohocken, PA, 2016b.
ASTM. D5456-17. Standard Specification for Evaluation of Structural Composite Lumber Products.
ASTM International, West Conshohocken, PA, 2017.
Peter Avitabile. Part 5: 101 ways to extract modal parameterswhich one is for me? Experimental
techniques, 30(5):48–56, 2006.
AWC. National Design Specification for Wood Constructions (NDS). Forest Product Association,
2012.
Etienne Balmès et al. Frequency domain identification of structural dynamics using the pole/residue
parametrization. OFFICE NATIONAL D ETUDES ET DE RECHERCHES AEROSPATIALES
ONERA-PUBLICATIONS-TP, 1996.
Anthony R Barrett. Dynamic testing of in-situ composite floors and evaluation of vibration service-
ability using the finite element method. Technical report, DTIC Document, 2006.
I Bejtka and HJ Blass. Joints with inclined screws. In Proceedings of the 35th Meeting of W018
of International Council for Research and Innovation in Building and Construction, pages 1–12,
2002.

A Blakeborough and MS Williams. Measurement of floor vibrations using a heel drop test. Proceed-
ings of the Institution of Civil Engineers-Structures and Buildings, 156(4):367–371, 2003.
Hans Joachim Blass and Ireneusz Bejtka. Screws with continuous threads in timber connections.
Joints in Timber Structures: Stuttgart, Germany, 12-14 September 2001, page 193, 2001.
Hans Joachim Blass and Peter Fellmoser. Design of solid wood panels with cross layers. In 8th world
conference on timber engineering, volume 14, page 2004, 2004.
Ronald Newbold Bracewell and Ronald N Bracewell. The Fourier transform and its applications,
volume 31999. McGraw-Hill New York, 1986.

275
F Brancaleoni, D Spina, and C Valente. Damage assessment from the dynamic response of deteriorat-
ing structures. In Safety Evaluation Based on Identification Approaches Related to Time-Variant
and Nonlinear Structures, pages 276–291. Springer, 1993.
Reinhard Brandner et al. Production and technology of cross laminated timber (clt): A state-of-
the-art report. In Focus Solid Timber Solution-European Conference on Cross Laminated Timber
(CLT), pages 21–22, 2013.
Anders Brandt. Noise and vibration analysis: signal analysis and experimental procedures. John
Wiley & Sons, 2011.
R Brillhart and D Hunt. Lessons learned in modal testing. Experimental Techniques, 29(6):58–61,
2005.
NORMAN Broner. The effects of low frequency noise on peoplea review. Journal of Sound and
Vibration, 58(4):483–500, 1978.
S. Cain. Standing tall, cross-laminated timber use increasing in mid-rise structures. , volume 17, no.
2, 12-19. Engineered Wood Journal, 17(2):12–19, 2014.

Ario Ceccotti. New technologies for construction of medium-rise buildings in seismic regions: the
xlam case. Structural Engineering International, 18(2):156–165, 2008.
Ario Ceccotti and Maurizio Follesa. Seismic behaviour of multi-storey x-lam buildings. In Proc.
International Workshop on” Earthquake Engineering on Timber Structures” Coimbra, Portugal,
2006.
Ario Ceccotti, Carmen Sandhaas, Minoru Okabe, Motoi Yasumura, Chikahiro Minowa, and Naohito
Kawai. Sofie project–3d shaking table test on a seven-storey full-scale cross-laminated timber
building. Earthquake Engineering & Structural Dynamics, 42(13):2003–2021, 2013.
Anil K Chopra et al. Dynamics of structures, volume 3. Prentice Hall New Jersey, 1995.

AW Christiansen. Effect of overdrying of yellow-poplar veneer on physical properties and bonding.


European Journal of Wood and Wood Products, 52(3):139–149, 1994.
Max Closen. Assy screw test. , My-ti-con, 2013.
Max Closen. Pre-drilling instruction for swg assy screws. , My-ti-con, 2015.

RW Clough and J Penzien. of structures. Machw Hill, pages 234–300, 1993.


Peggi Clouston, Leander A Bathon, and Alexander Schreyer. Shear and bending performance of
a novel wood–concrete composite system. Journal of Structural Engineering, 131(9):1404–1412,
2005.

Roy R Craig. Structural dynamics: an introduction to computer methods. John Wiley & Sons Inc,
1981.
RR Craig jr and AJ Kurdila. Fundamentals of structural mechanics. Address: John Wiley & Sons,
2, 2006.

Pablo Crespell and Sylvain Gagnon. Cross laminated timber: a primer. FPInnovations, Special
Publication, 52, 2010.
CSI. Linear and nonlinear static and dynamic analysis and design of three-dimensional structures.
Computers and Structures. Inc.: Berkeley, California, USA, 2015.

276
Brad Davis, Di Liu, and Thomas M Murray. Simplified experimental evaluation of floors subject
to walking-induced vibration. Journal of Performance of Constructed Facilities, 28(5):04014023,
2013.
Douglas Bradley Davis. Finite element modeling for prediction of low frequency floor vibrations due
to walking. PhD thesis, Virginia Tech, 2008.

Guido De Roeck, Bart Peeters, and Wei-Xin Ren. Benchmark study on system identification through
ambient vibration measurements. In Proceedings, 18th International Modal Analysis Conference,
pages 1106–1112, 2000.
DA Dillard. Fundamentals of stress transfer in bonded systems. adhesion science and engineering.
Elsevier, Amsterdam, The Netherlands, 2002.
Bruno Dujič and Roko Žarnić. Study of lateral resistance of massive X-Lam wooden wall system sub-
jected to horizontal loads. International workshop on Earthquake engineering on timber structures,
2006.
Bruno Dujic, Simona Klobcar, and Roko Zarnic. Shear capacity of cross-laminated wooden walls.
In Proceedings of the 10th World Conference on Timber Engineering, Miyazaki, Japan, 2008.
BS EN. 1-1: 2004 eurocode 5: Design of timber structuresgeneralcommon rules and rules for
buildings, 1995.
P Eriksson. modal analysis of a pre-cast concrete floor element. In Proc., 9th Int. Modal Analysis
Conf, pages 430–434, 1991.
CEN Eurocode. 5design of timber structurespart 1–1: generalcommon rules and rules for buildings.
CEN: Brussels, 2004.
James W Evans, David E Kretschmann, Victoria L Herian, and David W Green. Procedures for
developing allowable properties for a single species under astm d1990 and computer programs
useful for the calculations. Technical report, US Department of Agriculture, Forest Service, Forest
Products Laboratory, 2001.
DJ Ewins. Modal testing, practice and application, 2000.
Peter Fellmoser and Hans Joachim Blaß. Influence of rolling shear modulus on strength and stiffness
of structural bonded timber elements. In CIB-W18 Meeting, volume 37, 2004.
FEMA. Quantification of Building Seismic Performance Factors. FEMA P695, Washington, DC,
2009.
M Fitz. Untersuchung des Schwingverhaltens von Deckensystemen aus Brettsperrholz (BSP). PhD
thesis, Diplomarbeit, TU Graz, 2008.

Maurizio Follesa, Ioannis P Christovasilis, Davide Vassallo, Massimo Fragiacomo, and Ario Ceccotti.
Seismic design of multi-storey cross laminated timber buildings according to eurocode 8. Ingegneria
Sismica, 4, 2013.
Andrea Frangi, Giovanna Bochicchio, Ario Ceccotti, and Marco Pio Lauriola. Natural full-scale fire
test on a 3 storey xlam timber building. In Proceedings of 10th World Conference on Timber
Engineering (WCTE), 2008.
Andrea Frangi, Mario Fontana, Erich Hugi, and Robert Jübstl. Experimental analysis of cross-
laminated timber panels in fire. Fire Safety Journal, 44(8):1078–1087, 2009.

277
S. Gagnon, C. Pirvu, and FPInnovations (Institute). CLT handbook : Cross-Laminated Timber.
Quebec: FPInnovations, 2011.
Ivan Glisovic and Bosko Stevanovic. Vibrational behaviour of timber floors. In Proc., World Conf.
on Timber Engineering, Trees and Timber Institute, National Research Council, Italy, 2010.

Farid Golnaraghi and BC Kuo. Automatic control systems. Complex Variables, 2:1–1, 2010.
Karl Konstantin Grasser. Development of Cross Laminated Timber in the United States of America.
Master’s thesis, University of Tennessee, Knoxville, 2015.
David W Green and James W Evans. Mechanical properties of visually graded lumber. US Depart-
ment of Agriculture, Forest Service, Forest Products Laboratory, 1987.

Michael Green and E Karsh. The case for tall wood buildings. Canadian Wood Council on behalf of
the Wood Enterprise Coalition by Forestry Innovation Investment, North Vancouver, BC, Canada,
2012.
Michael J Griffin. Handbook of human vibration. Academic press, 2012.

Linda M Hanagan. Walking-induced floor vibration case studies. Journal of architectural engineering,
11(1):14–18, 2005.
Linda M Hanagan and Thomas M Murray. Active control approach for reducing floor vibrations.
Journal of structural engineering, 123(11):1497–1505, 1997.

Linda M Hanagan, Christopher H Raebel, and Eric Marsh. Modeling for controller design on a steel
floor system. In SPIE proceedings series, pages 1341–1347. Society of Photo-Optical Instrumen-
tation Engineers, 2000.
Linda M Hanagan, Christopher H Raebel, and Martin W Trethewey. Dynamic measurements of
in-place steel floors to assess vibration performance. Journal of Performance of Constructed
Facilities, 17(3):126–135, 2003.
Cyril M Harris and Robert T Beyer. Shock and vibration handbook, edited by cyril m. harris. The
Journal of the Acoustical Society of America, 84(3):1126–1126, 1988.
Laura Hasburgh, Keith Bourne, Perry Peralta, Phil Mitchell, Scott Schiff, and Weichiang Pang.
Effect of adhesives and ply configuration on the fire performance of southern pine cross-laminated
timber. In Proceedings of 14th World Conference on Timber Engineering (WCTE), 2016.
Monson H Hayes. Statistical digital signal processing and modeling john wiley & sons. New York,
1996.
Wilmar Hernandez. Improving the response of a load cell by using optimal filtering. Sensors, 6(7):
697–711, 2006.
Daniel P Hindman and John C Bouldin. Mechanical properties of southern pine cross-laminated
timber. Journal of Materials in Civil Engineering, 27(9):04014251, 2014.
Joseph N Howard, MA KAISER, and Thomas M Murray. Designing and building a force plate for
measuring input forces in modal tests. In SPIE proceedings series, pages 1083–1088. Society of
Photo-Optical Instrumentation Engineers, 1999.
L Hu and S Gagnon. Controlling cross-laminated timber (clt) floor vibrations fundamentals and
method. In Proceeding of the 12th World Conference of Timber Engineering, 2012.

278
Lin J Hu, Ying H Chui, and Donald M Onysko. Vibration serviceability of timber floors in residential
construction. Progress in Structural Engineering and Materials, 3(3):228–237, 2001.
LJ Hu. Development of a performance criterion for controlling vibrations in wood-based floors. In
Proceedings of 7th World Conference in Timber Engineering, pages 12–15, 2002.
LJ Hu and S Gagnon. Vibration performance of cross-laminated timber floors. CLT Handbook.
FPInnovations, Canada, 2011.
ICC. International building code. International Code Council, USA, 2015.
ISO. Standard 2631-2, Evaluation of human exposure to whole-body vibration. Internatinoal Standard
Organization, 1989.
Nicolas Jacquier and Ulf Arne Girhammar. Evaluation of bending tests on composite glulam–clt
beams connected with double-sided punched metal plates and inclined screws. Construction and
Building Materials, 95:762–773, 2015.
Tianjian Ji et al. Frequency and velocity of people walking. Structural Engineer, 84(3):36–40, 2005.
JIS. JIS A 1414-2:Performance test methods of panel components for building construction-Part 2:
tests for mechanical properties 5.3 bending test. Japanese Standards Association, 2010.
Robert Jockwer, René Steiger, and Andrea Frangi. Fully threaded self-tapping screws subjected to
combined axial and lateral loading with different load to grain angles. In Materials and joints in
timber structures, pages 265–272. Springer, 2014.
KW Johnsen. Theory of timber connections. Int Assoc Bridge Struct Eng., 9:249–262, 1949.
Norman L Johnson. Systems of frequency curves generated by methods of translation. Biometrika,
36(1/2):149–176, 1949.
Ronald W Jokerst. Finger-jointed wood products. Technical report, DTIC Document, 1981.
F. Julien. Manufacturing cross-laminated timber (clt):technological and ecomomical analysis. Tech-
nical Report 201001259-3257AAM, FPInnovations, Quebec,Canada, 2010.
Thomas Kailath. A view of three decades of linear filtering theory. IEEE Transactions on Informa-
tion Theory, 20(2):146–181, 1974.
A Kevarinmäki. Joints with inclined screws. cib-w18. Technical report, paper 35-7-3, Kyoto, Japan,
2002.
M Kippel, CLAUDE Leyder, ANDREA Frangi, and MARIO Fontana. Fire tests on loaded cross-
laminated timber wall and floor elements. Fire Safety Science, 11:626–639, 2014.
Josef Kolb. Systems in timber engineering: loadbearing structures and component layers. Walter de
Gruyter, 2008.
Anthonie Kramer, Andre R Barbosa, and Arijit Sinha. Viability of hybrid poplar in ansi approved
cross-laminated timber applications. Journal of Materials in Civil Engineering, 26(7):06014009,
2013.
H Kreuzinger. Mechanically jointed beams and columns. Timber Engineering-STEP, 1:1–8, 1995.
N Labonnote and KA Malo. Vibration properties of cross laminated timber floors. In Structures &
Architecture: ICSA 2010-1st International Conference on Structures & Architecture, July 21-23
July, 2010 in Guimaraes, Portugal, page 121. CRC Press, 2010.

279
Kenneth H Lenzen. Vibration of steel joist-concrete slab floors. Engineering Journal-American
Institute of Steel Construction Inc, 3(3):133, 1966.
KH Lenzen and TM Murray. Vibration of steel beam-concrete floor systems. Studies in Engineering
Mechanics, (29), 1969.

Felix A Levinzon. Fundamental noise limit of piezoelectric accelerometer. IEEE Sensors Journal, 4
(1):108–111, 2004.
Minghao Li, Frank Lam, and Yuan Li. Evaluating rolling shear strength properties of cross laminated
timber by torsional shear tests and bending tests. Stud, 10(18.5):280, 2014.
Alan A Marra. Technology of wood bonding. Van Nostrand Reinhold,New York, 1992.

Reza Masoudnia, Ashkan Hashemi, and Pierre Quenneville. Evaluation of the effective flange width
in the clt composite t-beams. In Proceedings of 14th World Conference on Timber Engineering
(WCTE), 2016.
Jack C McCormac. Structural steel design. Prentice Hall, 2008.

Cameron James McGregor. Contribution of cross laminated timber panels to room fires. PhD thesis,
Carleton University Ottawa, 2013.
Steve McKeand, Tim Mullin, Tom Byram, and Tim White. Deployment of genetically improved
loblolly and slash pines in the south. Journal of Forestry, 101(3):32–37, 2003.

Peter Mestek, Heinrich Kreuzinger, and Stefan Winter. Design of cross laminated timber (clt). In
Proceedings of the 10th World Conference on Timber Engineering, 2008.
Milad Mohamadzadeh and Daniel Hindman. Mechanical performance of yellow-poplar cross lami-
nated timber. Technical report, Virginia Tech, 2015.
M Mohammad, S Gagnon, BK Douglas, and L Podesto. Introduction to cross laminated timber.
Wood Design Focus, 22(2):3–12, 2012.
William Graham Montgomery. HOLLOW MASSIVE TIMBER PANELS: A HIGH-
PERFORMANCE, LONG-SPAN ALTERNATIVE TO CROSS LAMINATED TIMBER. Mas-
ter’s thesis, Clemson University, 101 Calhoun Dr, Clemson, SC 29634, 2014.

Thomas M Murray, David E Allen, and Eric E Ungar. Floor vibrations due to human activity.
American Institute of Steel Construction, 2003a.
TM Murray. Acceptability criterion for occupant-induced floor vibrations. Sound and Vibration, 13
(11):24–30, 1979.
TM Murray, DE Allen, and EE Ungar. Design guide 11, floor vibrations due to human activities.
American Institute of Steel Construction AISC, Canadian Institute of Steel Construction CISC,
2003b.
TH Nguyen, EF Gad, John L Wilson, and N Haritos. Improving a current method for predicting
walking-induced floor vibration. Steel and Composite Structures, 13(2):139–155, 2012.

Sven V Ohlsson. Floor vibrations and human discomfort. Chalmers University of Technology, Divi-
sion of Steel and Timber Structures, 1983.
Sven V Ohlsson. Springiness and human-induced floor vibrations: a design guide. Number 1988 in
12. Swedish council for building research, 1988.

280
Minoru Okabe, Motoi Yasumura, and Kenji Kobayashi. Estimation of bending stiffness, moment
carrying capacity and internal shear force of sugi clt panels. In Proceedings of World Conference
on Timber Engineering, Quebec City, Canada, August, pages 10–14, 2014.
Donald Onysko. Serviceability criteria for residential floors based on a field study of consumer
response. Canadas Wood Products Research Institute, 1986.

Lindsay Osborne, C Dagenais, and N Benichou. Preliminary clt fire resistance testing report. Point-
Claire, Canada, 2012.
SN Oswalt, WB Smith, PD Miles, and SA Pugh. Forest resources of the united states, 2012; a
technical document supporting the forest service 2015 update of the rpa assessment: Us depart-
ment of agriculture, forest service, washington office. Technical report, General Technical Report
GTR–WO–91, 218 p.[Also available at http://www. srs. fs. usda. gov/pubs/47322.], 2014.
Miroslav Pastor, Michal Binda, and Tomáš Harčarik. Modal assurance criterion. Procedia Engineer-
ing, 48:543–548, 2012.
A Pavic and P Reynolds. Experimental assessment of vibration serviceability of existing office floors
under human-induced excitation. Experimental Techniques, 23(5):41–45, 1999.
Shiling Pei, John W van de Lindt, and Marjan Popovski. Approximate r-factor for cross-laminated
timber walls in multistory buildings. Journal of Architectural Engineering, 19(4):245–255, 2012.
Matti Pirinen et al. Ductility of wood and wood members connected with mechanical fasteners. ,
2014.
Marjan Popovski and Erol Karacabeyli. Seismic behaviour of cross-laminated timber structures. In
Proceedings of the World Conference on Timber Engineering, 2012.
Marjan Popovski, Johannes Schneider, and Matthias Schweinsteiger. Lateral load resistance of
cross-laminated wood panels. In World Conference on Timber Engineering, pages 20–24, 2010.

Florian Prat-Vincent, Colin Rogers, and Alexander Salenikovich. Evaluation of the performance of
joist-to-header self tapping screw connections. In 2010 world conference on timber engineering,
Riva del Garda, Trento, Italy, ID256, volume 9, 2010.
CH Raebel, Linda M Hanagan, and Martin W Trethewey. Development of an experimental protocol
for floor vibration assessment. In IMAC-XIX: A Conference on Structural Dynamics,, volume 2,
pages 1126–1132, 2001.
H Reiher and FJ Meister. The effect of vibration on people. forsch gebeite ingenieurwes 1931; 2:
381–6 [in german] english translation: Report no. Technical report, F-TS-616-RE, Headquarters
Air Material Command, Wright Field, Ohio, 1946.

P Reynolds, A Pavic, and P Waldron. Modal testing, fe analysis and fe model correlation of a
600 tonne post-tensioned concrete floor. In PROCEEDINGS OF THE INTERNATIONAL SEM-
INAR ON MODAL ANALYSIS, volume 3, pages 1129–1136. KATHOLIEKE UNIVERSITEIT
LEUVEN, 1999.
Bryan H River. Fracture of adhesive-bonded wood joints. Handbook of Adhesive Technology, Revised
and Expanded, page 325, 2003.
Bryan H River and Victor P Miniutti. Surface damage before gluing-weak joints. Wood and wood
products, page 2, 1975.

281
Bryan H River, Charles B Vick, and Robert H Gillespie. Wood as an adherend. Treatise on adhesion
and adhesives, 7:1–23, 1991.
Adam Blake Robertson. A comparative life cycle assessment of mid-rise office building construction
alternatives: laminated timber or reinforced concrete. PhD thesis, University of British Columbia,
2011.

Roger M Rowell. Handbook of wood chemistry and wood composites. CRC press, 2012.
Brian J Schwarz and Mark H Richardson. Experimental modal analysis. CSI Reliability week, 35
(1):1–12, 1999.
Terry Sellers Jr. Wood adhesive innovations and applications in north america. Forest Products
Journal, 51(6):12, 2001.
Lars-Göran Sjökvist and Jonas Brunskog. An experimental and statistical study of the behavior
of the vibration field in two coupled lightweight wooden joist floors. Applied Acoustics, 74(4):
517–520, 2013.

Irving Skeist and Jerry Miron. Introduction to adhesives. In Handbook of adhesives, pages 3–20.
Springer, 1990.
Owings Skidmore and LLP Merrill. Timber tower research project. Report, Retrieved on, 4:16, 2013.
Michael Joseph Sladki. Prediction of floor vibration response using the finite element method. PhD
thesis, Virginia Tech, 1999.

Ian Smith. Vibration of timber floorsserviceability aspects. Timber engineering, pages 241–266,
2003.
W Brad Smith, Patrick D Miles, Charles H Perry, and Scott A Pugh. Forest resources of the united
states, 2007. Technical report, US Forest Service, 2009.

Kris Spickler, Max Closen, Philip Line, and Martin Pohll. Cross laminated timber horizontal di-
aphragm design example. My-ti-con, 2015.
WJ Staszewski. Identification of damping in mdof systems using time-scale decomposition. Journal
of sound and vibration, 203(2):283–305, 1997.

Eric Steinberg, Ricky Selle, and Thorsten Faust. Connectors for timber–lightweight concrete com-
posite structures. Journal of structural engineering, 129(11):1538–1545, 2003.
Roberto Tomasi, Maurizio Piazza, Albino Angeli, and M Mores. A new ductile ap-
proach design of joints assembled with screw connectors. In Proceedings of 9th
World Conference on Timber Engineering, Portland, OR. CD-ROM (see http://oregonstate.
edu/conferences/event/WCTE2006/proceedings. html), 2006.
Roberto Tomasi, Alessandro Crosatti, and Maurizio Piazza. Theoretical and experimental analysis
of timber-to-timber joints connected with inclined screws. Construction and building materials,
24(9):1560–1571, 2010.
Jan Weckendorf. Dynamic response of structural timber flooring systems. PhD thesis, Edinburgh
Napier University, 2009.
Jan Weckendorf and Ian Smith. Dynamic characteristics of shallow floors with cross-laminated-
timber spines. In Proceedings of 12th World Conference on Timber Engineering, July, pages
16–19, 2012.

282
Jan Weckendorf, Tomi Toratti, Ian Smith, and Thomas Tannert. Vibration serviceability perfor-
mance of timber floors. European Journal of Wood and Wood Products, 74(3):353–367, 2016.
Martin S Williams and Shahram Falati. Modal testing of a post-tensioned concrete model floor slab.
In SPIE proceedings series, pages 14–20. Society of Photo-Optical Instrumentation Engineers,
1999.

John F Wiss and Richard A Parmelee. Human perception of transient vibrations. Journal of the
Structural Division, 100(Proc Paper 10495), 1974.
Frank Woeste and Daniel Dolan. Is a spring in your step causing problems? Structural Engi-
neer(Atlanta, Ga), 8(5):24–27, 2007.

Borjen Yeh, Sylvain Gagnon, Tom Williamson, and Ciprian Pirvu. The cross-laminated timber
standard in north america. In World Conference on Timber Engineering, 2012.
Qinyi Zhou, Meng Gong, Ying Hei Chui, and Mohammad Mohammad. Measurement of rolling shear
modulus and strength of cross laminated timber fabricated with black spruce. Construction and
Building Materials, 64:379–386, 2014.

283

You might also like