Sei sulla pagina 1di 137

MULTI-STOREY

LIGHT TIMBER-FRAMED BUILDINGS


IN NEW ZEALAND – ENGINEERING DESIGN
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Author: David M Carradine


Contributing authors: Tobias Smith
Daniel Moroder
Emma O'Neill
Patricio Quintana Gallo
Roger Shelton
Design and layout: David Ronalds
ISBN: 978-1-98-852228-9 (pbk)
978-1-98-852229-6 (epub)
First published: October 2019
Copyright: BRANZ Ltd, 2019
Address: BRANZ Ltd
1222 Moonshine Road
RD1, Porirua 5381
Private Bag 50908
Porirua 5240
New Zealand
Phone: +64 4 237 1170
Fax: +64 4 237 1171
BRANZ shop: www.branz.co.nz

ALL RIGHTS RESERVED. The information contained herein is entitled to the full protection given by the Copyright Act
1994 to the holders of this copyright. Details may only be downloaded, stored or copied for personal use, by either an
individual or corporate entity, for the purposes of the carrying out of a construction-related or other business or for private
or educational use. Copying for the purposes of inclusion in trade or other literature for sale or transfer to a third party
is strictly forbidden. All applications for reproduction in any form should be made to the Channel Delivery Team Leader,
BRANZ Ltd, Private Bag 50908, Porirua City, New Zealand.
Disclaimer: The information contained within this publication is of a general nature only. BRANZ does not accept any
responsibility or liability for any direct, indirect, incidental, consequential, special, exemplary or punitive damage, or
for any loss of profit, income or any intangible losses, or any claims, costs, expenses, or damage, whether in contract,
tort (including negligence), equity or otherwise, arising directly or indirectly from or connected with your use of this
publication, or your reliance on information contained in this publication.
Any standard referred to within this publication can be purchased from Standards New Zealand by phoning 0800 782
632 or by visiting www.standards.govt.nz. Please note, the BRANZ books and bulletins mentioned in this publication
may be withdrawn at any time. For more information and an up-to-date list, visit BRANZ Shop online: www.branz.co.nz
or phone BRANZ 0800 80 80 85, press 2.

2
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

ACKNOWLEDGEMENTS
The author would like to thank all of the contributing authors for their
work on this document. Without their assistance, expertise and time, these
guidelines would not have been possible. A number of scientists, technicians
and administrators at BRANZ have also been extremely helpful in providing
suggestions and taking the time to help with a variety of tasks that also made
this document possible, so a big thank you to them.

3
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

PREFACE
There is currently a lack of published information in New Zealand around the
design of light timber-framed buildings that go beyond 2.5 storeys and are outside
the scope of NZS 3604:2011 Timber-framed buildings. This document provides
guidance on how designers can develop resilient Code-compliant designs for light
timber-framed structures up to 6 storeys within current New Zealand building
standards.

This design guidelines document has been prepared within BRANZ research
project QR00941 Specific design of light timber-framed buildings.

4
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

CONTENTS
1 INTRODUCTION 10
1.1 Background and motivation 10
1.2 Global LTF design methods 11
1.2.1 Europe 12
1.2.2 North America 12
1.3 Australian LTF design resources 14
1.4 New Zealand LTF design resources 14
1.5 Materials and capacity 17
1.6 Objectives and scope of these guidelines 17
1.7 Guidelines overview 18
2 LOADING DEMANDS FOR MULTI-STOREY LTF BUILDING 19
2.1 Introduction 19
2.2 Gravity and imposed vertical loads 20
2.3 Wind actions – lateral 20
2.4 Seismic actions – lateral 21
2.4.1 Seismic design in NZS 3604:2011 21
2.4.2 Capacity design 21
2.4.3 Building period 23
2.4.4 Higher modes 24
2.4.5 Ductility 24
2.4.6 P-delta effects 25
2.4.7 Use of the capacity factor φ in ductile and capacity protected elements 26
2.4.8 Calculation of non-linear horizontal deflections 26
2.4.9 Additional deflection of LTF LLRS due to pinched hysteresis loops 27
2.4.10 Calculation of local ductility demand 27
2.4.11 Calculation of inter-storey drift and drift limits 28
2.4.12 Podium structures and seismic loading 29
3 STRUCTURAL ANALYSIS AND DESIGN OF LTF BUILDINGS 30
3.1 Introduction 30
3.2 Gravity and wind design 30
3.2.1 Floors 31
3.2.2 Lintels and trimming studs 31
3.2.3 Walls 33
3.2.4 Roofs 35
3.3 Lateral load demand 35
3.3.1 Analytical modelling 36
3.3.2 Numerical modelling 49
3.4
Summary 51
4 DIAPHRAGM DESIGN FOR LTF BUILDINGS 52
4.1 Introduction 52
4.1.1 The role of diaphragms 52
4.1.2 Components of diaphragms 53
4.1.3 Timber diaphragm materials 54

5
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

4.1.4 Diaphragm design to NZS 3604:2011 57


4.1.5 Forces from higher modes on diaphragms during seismic loading 57
4.1.6 Capacity design and diaphragms 57
4.1.7 Transfer and deformation incompatibility forces 59
4.2 Design of diaphragms in LTF structures 59
4.2.1 Determination of wind diaphragm forces 60
4.2.2 Determination of seismic diaphragm forces 60
4.2.3 Dependency of the design capacity and the load direction 62
4.2.4 Distribution of forces within the diaphragm – flexible and rigid
diaphragms 62
4.3 Determination of diaphragm component demand 63
4.3.1 Girder analogy 64
4.3.2 Extended girder analogy 65
4.3.3 Finite element (shell) modelling 66
4.3.4 Special consideration for wind and large diaphragm demand 68
4.4 Component capacity and connection design in timber diaphragms 70
4.4.1 Panel connections 70
4.4.2 Sheathing panels 72
4.4.3 Chord, strut/drag and collector beams 73
4.4.4 Diaphragm deflection 75
5 CONTROL OF FLOOR VIBRATIONS IN MULTI-STOREY LTF BUILDINGS 76
5.1 Introduction 76
5.2 Floor vibration and critical design parameters 77
5.3
Design method 77
5.3.1 Calculation of vibration-controlled span 77
5.3.2 Vibration-controlled joist span tables 81
5.3.3 Influence of additional perpendicular floor elements 82
5.3.4 Non-rigid supports 82
5.4
Summary 83
6 CLADDING, WEATHERTIGHTNESS AND DURABILITY FOR LTF BUILDINGS 84
6.1 Introduction 84
6.2 Cladding and weathertightness 85
6.3
Durability 86
6.4
Summary 87
7 FIRE RESISTANCE IN MULTI-STOREY LTF BUILDINGS 88
7.1 Introduction 88
7.2 NZBC requirements 89
7.3 Methods for achieving fire resistance ratings in multi-storey LTF buildings 89
7.4 Fire during construction of LTF buildings 90
7.5 Summary 90
8 ACOUSTIC PERFORMANCE IN MULTI-STOREY LTF BUILDINGS 91
8.1 Introduction 91
8.2
NZBC requirements 92

6
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

8.3 Methods for achieving acceptable acoustic ratings in multi-storey LTF buildings 92
8.4
Summary 93
9 VERTICAL MOVEMENT IN MULTI-STOREY LTF BUILDINGS 94
9.1 Introduction 94
9.2 Individual contributions to vertical movement in LTF buildings 94
9.2.1 Settlement within the timber structure 94
9.2.2 Elastic deformation 95
9.2.3 Creep 95
9.2.4 Shrinkage 95
9.3 Estimation of vertical movement in LTF buildings 96
9.3.1 Estimating elastic deformation 96
9.3.2 Estimating creep deformation 97
9.3.3 Estimating shrinkage deformation 98
9.4
Summary 98
10
Design example 99
10.1
Gravity design 100
10.1.1 Floor design 100
10.1.2 Wall design 102
10.2
Lateral load design 103
10.2.1 Determination of seismic demand 103
10.2.2 Determination of wind demand 106
10.2.3 Lateral load distribution into walls 106
10.2.4 Design of wall components 109
10.2.5 Wall deformation and drift checks 113
10.2.6 P-delta effects 117
10.2.7 Building period 118
10.2.8 Numerical model 118
10.3
Diaphragm design 123
10.3.1 Diaphragm design: north-south direction (y-direction) 123
10.3.2 Diaphragm design: east-west direction (x-direction) 131
10.3.3 Finite element analysis 131
10.4
Summary 131
11 REFERENCES 132

7
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

FIGURES
Figure 1. Capacity design procedure. 22
Figure 2. Capacity design procedure and the chain analogy. 22
Figure 3. Typical hysteretic performance of plywood clad LTF wall subject to P21
bracing testing. 25
Figure 4. Design of lintels and trimming studs: a) aligned openings; b) non-aligned openings. 31
Figure 5. Structural model used for strength design of lintels (Shelton, 2013). 32
Figure 6. Structural models used for strength design of studs (Shelton, 2013). 33
Figure 7. Wall deformation contributions. 44
Figure 8. Stress spread in bottom plates for continuous and discrete supports. 48
Figure 9. Cumulative rotation of walls. 49
Figure 10. Spring and stick representation of LTF wall. 50
Figure 11. Definitions of diaphragm components. 53
Figure 12. Irregular floor geometry with typical diaphragm components for the given
lateral load direction shown with arrows. 54
Figure 13. Light timber frame and massive timber diaphragm examples, with schematic
cross-sections. 54
Figure 14. Blocked (left) and unblocked (right) diaphragms. 55
Figure 15. Blocked (top) and unblocked (bottom) diaphragms, fastener demand on
sheathing panels and framing elements, modified from Kessel & Schönhoff (2001). 56
Figure 16. Strength hierarchy in diaphragm design. 58
Figure 17. Static forces for ESA and pESA envelopes, modified from Gardiner, Bull and
Carr (2008). 60
Figure 18. Diaphragm behaviour. 63
Figure 19. Girder analogy for regular diaphragms. 64
Figure 20. Chord discontinuities and redistribution of forces in diaphragms. 67
Figure 21. Force introduction via the diaphragm edges: (a) via the bending of the chord beam;
(b) via fasteners into the sheathing panel, modified from Moroder et al. (2015). 69
Figure 22. Diaphragm connection details for sheathing panels. 71
Figure 23. Diaphragm connection details for mass timber panels. 72
Figure 24. Force transfer in collector or strut beams. 74
Figure 25. Force transfer in mass timber panels. 75
Figure 26. Architectural rendering of example LTF building (used with permission from
Interspace Architects). 99
Figure 27. Possible structural floor layout. 100
Figure 28. Wall layout with centre of mass and centre of rigidity. 107
Figure 29. Wall geometry in numerical model. 121
Figure 30. Numerical models with semi-rigid and rigid diaphragms and lateral loads. 121
Figure 31. Deformation of walls PY1 to PY9 under SLS loading (blue lines denote the
deformed wall shape). 123
Figure 32. Diaphragm portion considered for the analysis. 125
Figure 33. Shear force and moment distribution. 127
Figure 34. Diaphragm nailing – area in blue with nail spacing at 200 mm centres, area in
green nail spacing at 100 mm centres. 130

8
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

TABLES
Table 1. Short-term and long-term load factors. 32
Table 2. Summary of diaphragm analysis methods 64
Table 3. Floor joists SG8 in dry conditions considering vibration control. 81
Table 4. Floor joists LVL13 in dry conditions considering vibration control. 81
Table 5. Bracing wall lengths. 107
Table 6. Base shear force distribution (ULS in y-direction, including accidental eccentricities). 109
Table 7. Force demand and geometric information on wall PY1. 109
Table 8. Load demand in terms of shear flow and chord forces for capacity protected
elements (CPE) in wall Y1. 111
Table 9. Hold-down capacity values using = 1.6 and = 1.14. 112
Table 10. Required section sizes and fixings for wall PY1. 113
Table 11. Shear and moment demand for wall PY1 at ULS without accidental eccentricity. 113
Table 12. Wall PY1 rotations due to flexural deformation and anchorage slip and as relative
wall displacement. 115
Table 13. Elastic wall deformation contributions, elastic inter-storey deformation, elastic
inter-storey drift and total elastic displacement of wall PY1 under ULS loading. 116
Table 14. Ultimate inter-storey deformation, inter-storey drift and total displacement of
wall PY1 under ULS loading. 116
Table 15. Average ultimate inter-storey deformation, inter-storey drift and total
displacement of the structure in the y-direction under both SLS and ULS loading. 116
Table 16. Ultimate fastener deformations in the y-direction. 117
Table 17. Calculation of stability coefficients in the y-direction. 118
Table 18. Seismic mass, storey forces and average elastic deformation in the y-direction
of the building under SLS loading. 118
Table 19. Equivalent modulus of elasticity, shear modulus and rotational stiffness of wall PY1. 120
Table 20. Base shear distribution on the walls in the y-direction for the ULS load case
without accidental eccentricities. 122
Table 21. Base shear distribution on the walls in the y-direction. 123
Table 22. Wall overstrengths and input parameters for wall Y1. 124
Table 23. Wall overstrengths for all walls in the y-direction. 124
Table 24. Diaphragm storey forces in y-direction. 125
Table 25. Wall forces assuming a flexible diaphragm. 126
Table 26. Wall forces assuming a rigid diaphragm. 126

9
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

INTRODUCTION
1
In New Zealand and around the world, there is a growing interest in accelerating the
development of new types of multi-storey timber structures. The aim is to increase
the use of this renewable resource material using a broad range of connectors and
skill levels in simple structural systems. There are designers globally suggesting
that timber can be utilised for very tall buildings, including multi-storey structures
beyond the range of many international building codes for timber.

1.1 Background and motivation


There are various reasons for this move towards increased use of timber in multi-
storey buildings, but of most interest are the situations in New Zealand. Increased
steel prices have been cited as contributing to the surge in timber interest (Bowden,
2008). It was also mentioned that, at the time of writing, timber industries would
find it difficult to replace steel for non-residential applications without improving
manufacturing methods and finding support from engineers and quantity
surveyors. Maclaren (2010) used the Nelson-Marlborough Institute of Technology
Arts and Media Building in Nelson, New Zealand, as an example of extensive use of
timber in multi-storey buildings. He also provided numerous reasons for why this
type of construction is attractive to designers, builders and users.

In more recent years, the primary reasons for the worldwide interest in larger
timber structures includes a global focus on sustainability, the wide availability
of timber and timber-based products and efforts to develop building systems
that make more extensive use of renewable materials that store carbon, such
as timber and timber-based products. A more recent article in The Guardian
(Cathcart-Keays, 2014) discussed the possibilities surrounding high-rise timber
buildings and pointed out that New Zealand is one of the most promising places
in the world in terms of realising the value of larger timber buildings. It goes
on to note that, even in China, they are considering using wood for high-rise

10
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

construction. A survey of 10 large timber buildings conducted in 2014 (FII &


BSLC, 2014) indicated that timber is certainly a feasible option for providing
high-performance and cost-effective tall buildings up to 9 storeys.

In early 2014, the New Zealand Government announced the creation of a


$1 million award for innovative use of timber in larger buildings to increase
the use of timber. The intention was to provide an incentive for constructing
more sustainable buildings as well as for utilising material resources available
within the country (Norman, 2014). Another important aspect surrounding the
promotion of larger timber buildings in New Zealand is that both Auckland and
Christchurch are poised to have a significant increase in demand for mid-rise
buildings over the next 10–20 years. In Christchurch, the demand is largely
due to the rebuild of the city following the 2010/11 earthquakes. Auckland is
experiencing a significant increase in population, which is resulting in increased
demand for housing. Anderson (2014) noted that the construction of multi-unit
apartments, which has typically not been a strong market in New Zealand, is
increasing, but mostly in Auckland and to a lesser extent in Wellington.

Light timber framing (LTF) has stood the test of time as an efficient and practical
structural system. It maintains a strong foothold as one of the leading types
of construction for houses and low-rise buildings in New Zealand (Rosevear &
Curtis, 2017). These guidelines are aimed at providing methods for designing LTF
structures that fall beyond the scope of the prescriptive timber frame building
standard in New Zealand. The purpose is to provide guidance for designers who
want to make use of this simple and well understood building system for larger
buildings. There are significant initiatives around the world for the use of timber
in large buildings, including post-tensioned laminated veneer lumber (LVL),
cross-laminated timber (CLT) and glue laminated timber (glulam). There are
also many examples of LTF being used for multi-storey buildings, particularly
in the 3–6-storey range. These LTF buildings are being constructed in areas
of moderate to high seismicity, which further reinforces the notion that these
structures are resilient and efficient and have a role to play in the New Zealand
mid-rise landscape.

These guidelines have been developed so that designers and developers can
have robust options for constructing multi-storey buildings in timber using
local resources and construction techniques that are well understood in the
construction sector. The appetite for mid-rise buildings is rapidly growing in
New Zealand with population surges in major centres as well as government
initiatives that will result in housing densification. This will require medium-
density housing solutions, which are ideally suited to using LTF structures.
The guidelines have focused on structural solutions but include discussions
and suggestions for how to address a variety of other necessary design aspects
of multi-unit residential buildings such as fire and acoustic performance, floor
vibrations and weathertightness. In conjunction with other timber building
resources that are currently under development in New Zealand, these
guidelines will allow designers to provide a range of low-carbon timber building
options that meet the needs of the growing population and shifting community
profiles throughout the country.

1.2 Global LTF design methods


An important consideration for the implementation of design and construction
methods for multi-storey LTF buildings in New Zealand is how these buildings are
approached in other parts of the world. There are LTF buildings up to 6 storeys

11
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

being constructed in Europe and North America (Barber & Gerard, 2014), and
Australia is developing methods and guidelines for multi-storey LTF buildings
that can be used as examples and a roadmap for mid-rise timber buildings in
New Zealand. It is important to determine what methods are being used globally
and how these can be implemented within the New Zealand context. It is also
useful to know what has already been investigated so that choices can be made
about the most useful design options that are most likely to be practical for
adaptation.

1.2.1 Europe

There are examples of LTF construction throughout Europe, but in more


recent years, the focus for larger buildings has tended to be on massive
timber construction including the use of CLT and glulam. These massive
timber innovations have resulted in huge steps for timber buildings including
construction of an 18-storey building in Norway (Abrahamsen, 2018). Smith
and Frangi (2014) provided extensive information on taller timber buildings
and included a chapter on important structural design considerations for this
type of building. Different design philosophies were discussed with emphasis
on the importance of robust horizontal diaphragms and redundancy within
the structural system. In the United Kingdom, there has been more action
in the LTF space, and changes in the fire safety regulations in 1991 allowed
timber to be used for buildings up to 8 storeys (Grantham & Enjily, 2003).
Feasibility studies and example buildings in this part of the world provided
evidence of the effectiveness of LTF buildings for multi-storey applications.
While the UK is not seismically active, the design aspects for fire resistance,
differential movement and overall robustness of multi-storey LTF buildings
can be helpful in understanding what methods should be applied in New
Zealand.

1.2.2 North America

The United States and Canada have also been pushing forward with different
types of multi-storey timber structures. The Brock Commons building in
Vancouver, British Columbia is an 18-storey hybrid timber structure (Poirier
et al., 2016) that includes CLT and glulam supported laterally using concrete
cores. In the US, development also continues with the 6-storey steel and timber
hybrid T3 Atlanta Office Building in Georgia and the 3-storey post-tensioned CLT
Peavy Building in Corvallis, Oregon. North American LTF buildings have a lot of
similarities with New Zealand LTF buildings, and while mass timber such as CLT
is on the rise there, LTF still has a place in the multi-storey building space.

Resistance against earthquake actions is of paramount concern for larger timber


buildings around the world. An extensive literature review was conducted by
Mostafaei et al. (2013) on the seismic performance of mid-rise timber structures.
The focus of this investigation was primarily LTF as it is commonly used in North
America. It showed that, with a proper design, the likelihood of failure and even
damage of larger timber structures could be minimised during seismic events.
In terms of experimental testing, it was shown that, for shear walls, in general,
cyclic testing was more demanding than monotonic and that failure modes
changed when going from a single-storey wall up to a 6-storey wall. Sheathing
materials and wall-end hold-downs were noted as having significant impact
on the performance of the building. It was concluded that the characteristics
of the dynamic response of multi-storey timber buildings play a major role in
determining the seismic performance of these types of buildings.

12
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

In the United States, the limitations on the height of a building are related to
fire performance. For buildings with more than 4 storeys, a concrete first storey
is required, with an ultimate limit of 5 storeys for timber as it is considered a
combustible material. In the 2015 International Building Code (IBC), there is
information for classifying buildings by construction type (ICC, 2015). This allows
designers to determine maximum floor areas and heights. The definition of the
building type is based mostly on fire resistance and the combustibility of the
materials used to construct a building. The use of the building and the floor area
also have an influence on the allowable height. The IBC contains information
for designing and building with timber in prescriptive and specific design
applications. It also allows for buildings to be designed with the methods provided
in the Wood Frame Construction Manual (AWC, 2017) for one and two-family
dwellings up to 3 storeys in height. The IBC includes information on acceptable
materials and methods for construction but also provides span tables and tables
that describe necessary fastening, sheathing and connection requirements for
different loading conditions and components throughout a building. Most of
the specific engineering design components in the IBC refer engineers to other
documents including the National Design Specification for Wood Construction
(AWC, 2018) and the Special Design Provisions for Wind and Seismic (AWC,
2015). Prescriptive or conventional construction methods in the IBC are limited
to 3 storeys and are not allowed in high seismic regions, where additional height
restrictions are implemented. The International Residential Code (IRC) also
provides prescriptive and specific design details but it is also limited to 3 storeys
and is specified for one and two-family dwellings only (ICC, 2018).

In addition to the building codes mentioned above, there are many published
documents that serve as design guides for prescriptive and engineered
construction on topics such as wall bracing based on IRC provisions (ICC & APA,
2012), shear wall and horizontal diaphragm design (APA, 2007) and the use
of engineered wood products (APA, 2011a). While helpful for designers, these
guides are not a replacement for engineering design required when timber
buildings exceed the 3-storey limitation for prescriptive designs.

Some of the APA documents provide more rigorous detailing for high wind
regions. There are some methods that could be useful for New Zealand, such as
details for additional nailing and for overlapping sheathing across floor levels to
provide increased uplift and shear loading resistance (APA, 2012). This requires the
use of longer than typical sheets of plywood, which are available in some plywood
thicknesses around New Zealand. This is also suggested for the ground floor level,
where the sheathing should be wrapped across the rim joist and onto the sill plate.
This would be difficult for a slab on grade, but it is feasible for larger buildings with
at least one timber floor on a podium or suitable concrete foundation. Making sure
that the bottom plate is bolted or otherwise solidly secured to the foundation will
minimise problems due to the sliding or differential movement between the plates
and foundations. Another APA report (APA, 2011b) notes that, for high wind uplift
resistance, at least 7/16” (11.1 mm) thick sheathing is required. This report also
provides some details for tension splice plates that could be used to carry tensile
load across wall-to-floor connections.

Canada also has building code provisions that limit prescriptive designs of
LTF structures to 3 storeys with a maximum inter-storey height of 3.5 m and a
maximum building area of 600 m2 (CWC, 2009). Less restrictive requirements
for timber construction allow for the design of larger buildings using the Wood
Design Manual (CWC, 2010). In recent years, the province of British Columbia
has altered the provincial building code to allow for LTF buildings to extend

13
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

up to 6 storeys (Wood Works!, 2011). The increase in allowable heights for


LTF buildings has made it necessary to consider alternative design options
for vertical and horizontal load resistance. Vertical load carrying capacity can
be increased by using engineered wood products and ensuring precision of
construction, which can be greatly increased by using prefabricated components
and computer-aided manufacturing methods. Horizontal loads can be resisted
more effectively by using stronger shear walls than typically used in lower-rise
buildings as well as by including continuous self-adjusting steel tie rod systems
in strategic locations throughout the building. Horizontal diaphragms must also
be detailed and constructed to resist the increased loads imposed on larger LTF
buildings (Wood Works!, 2011). Due to the location and function of this case
study building in British Columbia (Wood Works!, 2011), it provides an excellent
example of the kind of structures that are possible for New Zealand using multi-
storey LTF construction.

A handbook published in Canada (Ni & Popovski, 2015) provides guidance on


the design of larger LTF buildings. It incorporates several research programmes
and structural aspects along with fire resistance, acoustic performance and
other non-structural considerations that aid designers of LTF buildings up to 6
storeys. The limitations are described from Canadian building code documents
on the inclusion of contributions from gypsum plasterboard towards lateral
load resistance. For LTF buildings over 4 storeys, there is to be no contribution
allowed from gypsum plasterboard when evaluating lateral load resistance.

1.3 Australian LTF design resources


Australian interest in multi-storey timber buildings has been growing in recent
years, also utilising CLT, LVL and post-tensioned timber solutions (Lavisci,
Buchanan & Burdon, 2018). Changes in the Australian National Construction
Code in 2016 allow for timber buildings up to 8–9 storeys (25 m) to be
constructed, with taller buildings also feasible based on verification of their
performance.

WoodSolutions (www.woodsolutions.com.au/publications) has developed a


series of guidance documents on the design of timber buildings, and these
include multi-storey LTF structures. This series of guides provides a range of
information on designing for acoustic performance, fire resistance and other
relevant topics for multi-storey timber buildings. It is important to note that the
different building code structure in Australia means that designers will always
have to keep in mind the differences between the countries and ensure that
New Zealand requirements are adhered to, particularly with respect to seismic
resistance.

1.4 New Zealand LTF design resources


Design and construction of LTF buildings in New Zealand are described primarily
by either prescriptive means using NZS 3604:2011 Timber-framed buildings or,
for timber structural solutions requiring specific engineering design (SED),
using NZS 3603:1993 Timber structures standard. NZS 3604:2011 is intended to
be used by builders, architects, engineers and designers and provides detailed
guidelines for the design and construction of LTF buildings within a limited
scope of building types. NZS 3603:1993 is more for use by qualified professional
design engineers with a knowledge of timber structures for the design of timber
building components that fall outside the scope of NZS 3604:2011. NZS 3603:1993
is approved as a Verification Method for proving compliance with the New

14
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Zealand Building Code (NZBC). At the time of writing, NZS 3603:1993 is under
revision and will be replaced with NZS AS 1720.1 Timber structures – Part 1: Design
methods, which is based on the Australian standard AS 1720.1 with modifications
for New Zealand. This standard will provide design methods and requirements
for all ranges of timber buildings. The design methods provided in this document
provide explanations that include what is proposed for the new standard, but
currently this standard is not officially adopted.

Due to the prescriptive nature of the standard, NZS 3604:2011 (section 1.1.2)
specifically defines the limitations on buildings that can be built or designed
using these methods. The most relevant limitations in NZS 3604:2011 are on the
size of the building. Total height measured from the ground level to the highest
point on the roof cannot exceed 10 m. Additionally, while 2-storey structures
are allowed, in the event of a 3-storey building, there are limits that only two of
the storeys can be supported on timber framing and one of the storeys must be
a part storey in a roof space. This effectively limits the number of storeys to 2.5,
with additional stipulations on the lower storey walls that includes the use of
concrete and/or masonry. Also included in NZS 3604:2011 are a flowchart (Figure
1.1) and schematic drawings (Figure 1.2) for determining the applicability of this
standard to a particular building.

In addition to NZS 3604:2011, testing methods to determine bracing units used


for buildings designed using NZS 3604:2011 limit their applicability to modest
sized timber buildings that incorporate structural elements that are well
distributed. However, it is mentioned that they could potentially be used for
buildings beyond the scope of NZS 3604:2011. This would require appropriate
engineering judgement (Shelton, 2010). While some of the construction
methods provided in NZS 3604:2011 would be acceptable for taller buildings,
additional techniques are needed for increased building heights, especially when
considering foundations, subfloors, walls and diaphragms.

Rather than prescriptive methods for building design, NZS 3603:1993 provides
specific characteristic properties and methods for calculating design strengths
for timber structural components to be used by design engineers. Methods are
included for sizing structural members in solid sawn timber and other timber-
based materials, as well as for the design of connections using nails, screws,
bolts, adhesives and other fastening systems. Section 5.2 of this standard
provides guidance for designing LTF shear walls and horizontal diaphragms
where timber-based panels are nail connected to timber framing. This standard
does not specifically limit the size of these components. Some limitations are
placed on the nails and thickness of the sheathing used to ensure fully ductile
design during cyclic earthquake loading. As with most nailed systems described
in NZS 3603:1993, an overstrength factor of 2.0 is recommended for all other
system components, including sheathing, chords, hold-down connections
and foundations (section 5.2.4 of NZS 3603:1993). Designers are encouraged
to investigate shear wall and foundation anchorages and chord and splice
connections to verify compatibility with the shear wall strength. Equations
are provided for determining deflections of LTF horizontal diaphragms and
shear walls. Section 6 of NZS 3603:1993 and its appendices provide additional
resources for designing composite plywood and timber members although
without specifically addressing plywood shear walls and diaphragms. Included
in NZS 3603:1993 are also methods for testing structural components and
assemblies for compliance with the standard. As noted previously, at the
writing of these guidelines, there is a revision of NZS 3603:1993 under way with
expected completion during 2019.

15
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Engineers in New Zealand also have other options for seeking assistance with
the design of larger timber buildings, including LTF constructions. Buchanan
(2007) provides general performance and design criteria for timber shear walls
and diaphragms, including information on stress distributions in the various
components of these composite structures. Buchanan suggests that, for multi-
storey timber buildings using LTF, the maximum likely height is between 5 and
8 storeys if the lateral loads are resisted by plywood shear walls. In the case
of lateral loads being resisted by gypsum plasterboard linings, the more likely
maximum height would be 4 storeys.

The design of plywood shear walls is briefly summarised in Buchanan (2007),


whereas design examples and formulae can be found in Beattie (2001).
This manual provides more comprehensive information on the design of
larger timber buildings within New Zealand. Beattie provides evidence for
the feasibility of larger timber buildings using economic and construction
considerations specific to New Zealand and establishes 6 storeys as the possible
maximum number of storeys. The manual provides general guidance on possible
LTF construction typologies, ductile versus elastic designs and horizontal
diaphragm and shear wall design considerations. Beattie also provides
background and design recommendations for the control of changes in moisture
conditions, fire resistance and acoustics for multi-storey timber buildings
with ample information on typically constructed LTF systems with sound
transmission ratings. The methods for elastic designs are outlined with some
detail, but ductile design methods are provided in other documents (Deam,
1996). Beattie refers to standards that have been superseded by more recent
ones, which limits the usefulness of the elastic design examples.

Some work has also been presented by Banks and others regarding the feasibility
and design of multi-storey timber buildings in New Zealand. Banks (1999)
considered the feasibility of multi-storey timber construction in New Zealand.
He noted that, while in 1992 a change in the NZBC removed restrictions on the
height, the number of storeys for timber buildings being constructed remained
at 5 storeys. Numerous benefits of using timber for larger buildings were
discussed, and three potential 14-storey buildings were presented as feasible
options. These buildings utilised timber for mostly secondary components, but
one did include a timber moment-resisting frame as the primary lateral load
resisting system. None of the buildings used LTF for the structural system.

Milburn and Banks (2004) described the design and detailing of a 6-storey
apartment building in Wellington, the region of highest seismic activity in New
Zealand, which utilised plywood-lined walls for most of the required lateral
bracing. It described the unique details required to provide adequate resistance to
earthquake and wind actions and provided information on the analyses conducted
to derive them. This paper also provided several instances where LTF construction
was beneficial in terms of time and cost for a project of this size. More specific
details on designing plywood shear walls was provided by Banks (2007), including
recommendations on pitfalls to avoid during the design in an aim of simplicity.

More recent guidelines on the construction of larger timber buildings have


been developed by STIC, but these did not include information on LTF
structural systems.

At the time of writing, a project is under way by the Wood Processors and
Manufacturers Association of New Zealand (WPMA) on the development of
several guidance documents that are specifically for larger timber buildings.

16
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

There is not a great deal that is specific to LTF, but fire and acoustics in multi-
storey timber buildings are included topics along with important aspects about
costing and quantity surveying aspects of these buildings.

A significant amount of research has been conducted at BRANZ over the years
in relation to the design and performance of LTF buildings. Because most of the
LTF construction in New Zealand is based on NZS 3604:2011, some research has
focused on the underlying engineering principles behind the standard (Shelton,
2013). This work not only helps with an understanding of the assumptions and
technical criteria used to develop the standard, but also provides information to
designers who wish to go beyond the limitations of NZS 3604:2011 and develop
specific engineering designs using LTF. BRANZ resources and study reports on
a wide range of topics related to timber buildings can be found on the BRANZ
website (www.branz.nz).

1.5 Materials and capacity


When considering methods for designing taller LTF buildings, it is also critical
to have a comprehensive understanding of the failure mechanisms that could
potentially occur and how these can be accommodated, either within current
design practice or through updated methods. This requires information on the
strength and stiffness of the materials used, which would include solid sawn
radiata pine and Douglas fir timber, glulam, LVL and plywood. Strength and
stiffness values are available from manufacturers and standards such as NZS
3603:1993, but in some cases, research has suggested that these published
values may not be conservative enough for robust designs, particularly where
the designs are outside the boundaries of NZS 3604:2011 (Franke & Quenneville,
2011; van Beerschoten, Carradine & Palermo, 2013).

Most buildings constructed using the prescriptive requirements of NZS 3604:2011


are laterally braced using gypsum plasterboard screwed to LTF. Therefore,
the potential limits of this building method should be considered, as they may
be able to reach greater storey heights than are currently allowed for in NZS
3604:2011. Research conducted at BRANZ (Thurston, 2003) comprising full-
scale building tests indicated that, at least for a single-storey house, the bracing
performance was 50% greater than design predictions, assuming that the
walls were restrained against uplifting. The effectiveness of plasterboard shear
walls for resisting seismic actions was considered by Newcombe and Batchelar
(2012), who analysed existing P21 test data and performed several finite element
numerical analyses on plasterboard-lined bracing walls. They found that
current design procedures using NZS 3604:2011 may be unconservative due to
differences in the performance of walls of different lengths. Recommendations
were made for the design and evaluation of plasterboard braced houses to
provide more robust and conservative methods. The methods described helped
to identify some weak areas that need to be addressed in design, particularly for
2-storey houses in high seismicity regions.

1.6 Objectives and scope of these guidelines


While there is information available globally and some in New Zealand, there
remains a lack of guidance for the design of LTF buildings in the 3–6-storey
range. These guidelines are primarily for structural engineers and building
designers who need to know what options are available and viable for these
types of buildings in New Zealand. Designs for these types of buildings are
already under way, so the viability is not so much a question but rather it is the

17
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

methods that will result in Code-compliant designs that can be easily justified to
consenting authorities. These guidelines also provide information for consenting
officials to assist with their understanding of the critical aspects of these
buildings that require attention and appropriate detailing.

1.7 Guidelines overview


Section 2 presents descriptions of how load demands for multi-storey LTF
buildings should be derived and includes recommendations for calculating
seismic loads as these are highly dependent on building weight and the
structural system used.

Section 3 provides methods for conducting a comprehensive structural analysis


of a multi-storey LTF building and determining the structural actions on systems
and components throughout the building.

Due to the importance of understanding the behaviour of floor diaphragms,


section 4 is dedicated to the design of diaphragms that can be used in LTF
buildings and includes recommendations for massive floor systems that can be
used in conjunction with LTF buildings.

Section 5 includes recommendations for addressing issues around floor


vibrations in multi-storey LTF buildings.

Section 6 considers solutions for cladding, durability and weathertightness for


LTF buildings up to 6 storeys and provides some discussion on how these issues
can be addressed when increasing the height of timber buildings.

Fire resistance for multi-storey LTF buildings is covered in section 7 including


applicable NZBC requirements and how they can be addressed using
conventional methods that are already in common use in New Zealand.

Section 8 considers options for how to achieve Code-conforming acoustic


performance in multi-storey LTF buildings up to 6 storeys and includes
proprietary solutions as well as recommendations from research and practice.

Because timber has the potential to creep and induce movement in multi-storey
buildings, section 9 looks at vertical movement and how it can be calculated
and addressed with recommendations from overseas adapted for use with New
Zealand materials and methods.

Section 10 incorporates the information from the previous sections into a


worked-through design example using a 4-storey demonstration building so that
designers can implement the methods provided and see an application in action.

18
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

LOADING DEMANDS FOR


2
MULTI-STOREY LTF BUILDING

2.1 Introduction
In order to design any building, it is necessary to determine the loads that will
need to be resisted by the building as required by the NZBC and applicable
standards. This includes gravity and other vertical loads, lateral wind loads and
lateral seismic loads. It is also necessary for the design of the cladding systems
and other non-structural elements to understand the loading on the building so
that those elements can be effectively specified and Code compliant. It should be
noted that, for most types of loading, the demands on LTF buildings are similar
to those for any other structural system considered. The differences occur
primarily when considering seismic loads where the performance of the system
in terms of ductility and damping is affected by the structural system being used
and also by the weight of the building, which is typically less with timber than
other systems such as concrete.

Five applicable parts of the AS/NZS 1170 set of design action standards are
used for building design in New Zealand, and these are explained in the
following sections. Any building project starts with AS/NZS 1170.0:2002
Structural design actions – Part 0: General principles, which provides general
principles for determining load actions on structures. This contains
explanations of design procedures, load action combinations and analysis
methods for Code-conforming limit state designs of buildings as well as
information on exceedance requirements, how to utilise experimental data
and serviceability limitations. All load action determinations described in the
following sections must be made in conjunction with AS/NZS 1170.0:2002 to
ensure a comprehensive understanding of how the actions are calculated,
combined and applied for structural design purposes.

19
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

2.2 Gravity and imposed vertical loads


Determination of vertical actions due to gravity based on the self-weight of
the building and imposed loads due to the intended use or occupancy of the
building must be done according to AS/NZS 1170.1:2002 (Reconfirmed 2016)
Structural design actions – Part 1: Permanent, imposed and other actions. Building
self-weight includes the building structure but also claddings, partitions, interior
linings, permanent equipment and all other materials incorporated within the
building such as services. Imposed actions depend on the use of the building
and can be in the form of concentrated loads, partial loads or fully distributed
loads. In addition to self-weight and imposed actions, additional vertical
actions sometimes need to be considered due to snow and ice accumulation
as determined using AS/NZS 1170.3:2003 (Reconfirmed 2016) Structural design
actions – Part 3: Snow and ice actions. These actions are primarily based on
geographic location, height above sea level, topographic features and the shape
of the roof.

All applicable vertical actions must be applied as forces or pressures to floor and
roof systems and subsequently transferred through these systems to the columns
or walls that carry them to the building foundation. Section 3 of this document
provides methods for analysis of the necessary components and systems of LTF
buildings that transfer and resist applied vertical actions.

2.3 Wind actions – lateral


The determination of lateral wind actions for LTF buildings is no different than
for any other type of multi-storey building. The procedures in AS/NZS 1170.2:2011
(Reconfirmed 2016) Structural design actions – Part 2: Wind actions provide
methods for determining wind actions on buildings. Wind actions are based on
normal pressure created by blowing wind in conjunction with characteristics of
the shape and orientation of the building, which can create increases in pressure
and sometimes dynamic effects that require design attention.

Wind speed is calculated based on annual probability of exceedance, building


height, location, topographic effects and wind direction. Topographic effects
include terrain roughness, shielding and the relationship between the building
and the surrounding topography like hills and escarpments. Wind speeds need
to be calculated for all eight cardinal directions and the worst-case scenarios
used for design, considering the orientation of the building with respect to
prevailing winds.

The wind speeds are used to calculate pressures that are applied to normal
surfaces of the building such as walls and roofs and in some instances also need
to be applied to internal surfaces. Pressures are calculated for various parts and
areas on a building based on the wind speed in specified directions, an assumed
density of air, aerodynamic shape factors and dynamic response factors.

The shape factors provided in section 5 of AS/NZS 1170.2:2011 provide adjustments


for pressures applied to buildings or parts of buildings including interior and
exterior surfaces of walls and roofs and are described in terms of suction or
pressure, depending on the direction in which the action must be applied. These
factors account for a variety of building features including the shape of the roof
and overhangs, parapets and openings such as windows. Combination factors
also need to be considered, which allow for reductions in wind pressures caused
by non-simultaneous gusts. It is particularly important when designing cladding

20
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

systems that designers apply the appropriate local pressure factors Kl, described
in section 5.4.4 of the standard, as these factors can as much as triple the basic
pressures applied to certain areas of walls and roofs.

Dynamic response factors are dependent on the building frequency and are
only required for dynamically wind-sensitive buildings as described in section 6
of AS/NZS 1170.2:2011. While the aspect ratios of mid-rise LTF buildings are not
likely to result in them being dynamically wind sensitive, it is valuable to have an
understanding of how buildings can be proportioned to avoid this issue.

2.4 Seismic actions – lateral


As a high seismic risk country, designing for earthquake loading is of crucial
importance in New Zealand. A key to this design is the use of ductility, or
controlled damage in the sheathing fasteners in the case of a traditional LTF
building, to dissipate energy and reduce seismic load. This section outlines the
specific considerations required for the design of LTF buildings for earthquake
loading in New Zealand.

2.4.1 Seismic design in NZS 3604:2011

A detailed description of the assumptions used in the establishment of


earthquake bracing demand used in NZS 3604:2011 can be found in Shelton
(2013). While these assumptions are acceptable for the design of dwellings
within the scope of NZS 3604:2011, specific calculation of the seismic
demand in accordance with NZS 1170.5:2004 Structural design actions – Part
5: Earthquake actions – New Zealand is required for structures within the
intended scope of this document. The bracing unit concept and P21 test
results should therefore not be relied on when designing multi-storey LTF
buildings taller than 2.5 storeys.

2.4.1.1 Ductility in accordance with NZS 3604:2011

The ductility factor used in NZS 3604:2011 is μ = 3.5, and Shelton (2013)
recommends that this value also be used where specifically designed bracing
elements are used in buildings within the scope of NZS 3604:2011.

It is however not considered that the buildings covered within this document are
equivalent to those covered in NZS 3604:2011. This is mainly due to the increased
height of the structures with up to 6 storeys and therefore the increased loading
demand. Ductility values higher than 3 are also not specifically covered in NZS AS
1720.1 and would therefore require an Alternative Solution. Further discussion on
a suitable ductility assumption is described in section 2.4.5.

2.4.2 Capacity design

Capacity design is a fundamental principle of seismic design. As outlined in


Figure 1, capacity design requires the selection of a mechanism with designated
ductile regions or ductile connections. These ductile components must be
designed to withstand large cyclic deformations, and all other (less ductile
or brittle) elements must be designed to account for the overcapacity and
overstrength of the ductile elements.

The overcapacity is the difference between the actual strength of the ductile
elements and the seismic demand and should be kept as small as possible.

21
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The overstrength accounts for the probable higher strength of the ductile
elements, inaccuracies between analytical and tested values and other
hidden effects like friction.

Choose a kinematically Design the ductile connections Design all other brittle elements
ACTION

ACTION

ACTION
admissible mechanism – i.e. to fail in a ductile manner, for a demand greater than the
define the type and position with their capacity as close as actual capacity of the ductile
of the ductile element that possible to the demand – i.e. element including overstrength.
dissipates energy. ductile fasteners.
DEFINITION

DEFINITION

DEFINITION
Ductile mechanism as per Design of ductile elements Capacity protected elements
NZS 1170.5:2004. as per NZS AS 1720.1. as per NZS AS 1720.1.

Ensure a suitable global ductile Guarantee a fastener failure Guarantee that the brittle
behaviour in the structure can mechanism with at least one elements can withstand the
REASON

REASON

REASON
be achieved (i.e. nailing of the plastic hinge with the possibility probable higher forces from
sheathing and/or hold-down of resisting repetitive cycles with the ductile elements within a
connectors). minimal strength degradation. certain margin.

Figure 1. Capacity design procedure.

This principle can be best explained with a chain analogy shown in Figure 2,
where the maximum capacity is dictated by the weakest link. When the weakest
link is ductile, the overall chain is ductile. In order to ensure that the ductile
link in the system is indeed the weakest component along any given load path,
an overstrength factor is defined for ductile links.

brittle ductile brittle

Eseismic Eseismic

Rn,1 ≥ ɸ0 Rn,ductile Rn,ductile ≥ Eseismic Rn,1 ≥ ɸ0 Rn,ductile

Rt,min Ry,eff Rt,min

brittle ductile brittle

Figure 2. Capacity design procedure and the chain analogy.

22
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

As described below, these guidelines assume that the panel fasteners provide
ductility to the lateral load resisting system (LLRS) consisting of timber-based
panels fixed to timber framing to create shear walls. Other elements in the
system are capacity protected. The overstrength capacity of the panel fastener
can be calculated as:

(2.1)

where:
vCPE unit shear capacity at overstrength of the wall
φo, fastener overstrength factor of the ductile fastener
φ capacity factor of ductile fastener
Nd,j design fastener strength
k17 factor for multiple fastener effect
s fastener spacing

When calculating the design fastener strength Nd,j as per NZS AS 1720.1, an
amplification factor k17 can be used with connections with more than one fastener.
This increase in strength is justified by the lower probability of failure of the group of
fasteners and moves the design capacity from a lower characteristic value closer to
the mean value. Fastener overstrength factors consider the difference between the
lower and the upper characteristic strengths, and hence the group factor k17 needs to
be discounted when calculating the overstrength of the ductile fasteners.

The sheathing panel, chords and hold-downs need to be able to resist the
demand based on this overstrength capacity. This means that all capacity
protected elements need to be designed for a demand that is independent from
the actual seismic demand, as their demand is governed by the ductile fastener
capacity. The design demand on all capacity protected members is however
limited by the forces obtained by an elastic equivalent static analysis (ESA),
i.e. from an elastic response spectrum with a structural performance factor Sp
relating to the global ductility of the structure.

2.4.3 Building period

The fundamental period of the building T1 is a critical design perimeter that is


used to define the design base shear due to earthquake loading.

In most design cases, an approximation of fundamental building period is required


to begin sizing LLRS elements. LTF structures are wall buildings, and as such, the
empirical equation provided in section C4 of NZS 1170.5:2004 can be used with kt
= 0.05. LTF walls are generally more flexible than other forms of wall construction,
and although this increased flexibility is offset in part by their lightweight nature, an
additional factor of 2 is applied to the empirical formal as follows:

where:

(2.2)

T1 fundamental period of vibration


h n total building height

23
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Once the walls have been designed and detailed, the fundamental period can be
refined by using Rayleigh’s method:

(2.3)

where:
Fi seismic lateral force at level i
wi seismic weight at level i
Δi elastic static lateral displacement at level i due to the forces Fi
g acceleration due to gravity

When calculating the building period using Rayleigh’s method, the mean values
for the modulus of elasticity and shear shall be used. Unless a more precise
force-displacement curve is available, the tangent value for the fastener slip
modulus at serviceability limit state (SLS) loads should also be used.

By using Rayleigh’s method, a very high building period may be obtained, and
it is therefore recommended that a limit of two times the empirical period from
Equation (2.2) for ultimate limit state (ULS) is considered as an upper limit during
design. This is to prevent the use of very high calculated building periods due
to uncertainties in including the stiffening effect from non-structural elements,
friction between the individual building elements, the differences between the
design and the as-built conditions and possible accuracies in the modelling.

If a numerical model has been made of the structure including the real wall and
floor stiffness (i.e. modified wall and floor elements as described in section 3.3.2
and section 4.3.3, respectively), it will be relatively straightforward to apply
Equation (2.3) by reading the displacement at the centre of mass of the structure.
In addition, most modern numerical analysis programmes can provide a direct
calculation of the fundamental period that can be used.

In the analytical analysis procedure described in section 3.3.1.4, a rigid


diaphragm assumption is made and the wall forces are distributed in accordance
with this assumption. In order to have a realistic gauge of the building period,
the average displacements of all walls from the analytical model, neglecting
the effects of accidental eccentricity, should be used as the static lateral
displacement at level i.

2.4.4 Higher modes

Higher modes do not need to be specifically accounted for in the design of the
LLRS – an ESA or modal analysis can be adopted.

The effect of higher modes is normally of short duration only and is difficult to
assess. In order to consider any higher mode effects and higher than expected
seismic events, diaphragm yielding can be relied upon as discussed in section
4.1.5.

2.4.5 Ductility

Ductility in an LTF wall is generally provided by the nailed connection of


the sheathing panel to the stud framing. As the building racks over, the nails

24
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

(or ductile screws or staples) connecting the sheathing are yielded cyclically
providing hysteresis. This mechanism crushes the timber in both the panel
and the wall stud creating a ‘pinched’ hysteresis loop as shown in Figure 3. The
impact of this crushing on seismic performance is further discussed in section
2.4.9.

10

Load (kN)
6

-45 -35 -25 -15 -5 5 15 25 35 45


-2

-4
Displacement (mm)
-6

-8

-10

Figure 3. Typical hysteretic performance of plywood clad LTF wall subject to P21 bracing testing.

Typically assumed values of building ductility for LTF wall buildings are in the
range of 3–3.5. NZS AS 1720.1 limits the ductility of timber buildings to 3 as part
of the Verification Method, and this value is recommended for LTF buildings
within the scope of this document. Assumptions in this range do not typically
exceed the limits of local ductility that will be checked in section 2.4.10.

2.4.5.1 Fastener local ductility

Not all small dowel type metal fasteners are capable of providing the local
ductility required to ensure adequate levels of global ductility in the LTF system.
Hardened screws – for example, may fail in a brittle manner under cyclic
yielding, compromising load paths and structural stability.

Care in the specification of fasteners should be taken by the designer to ensure


that local ductile capacities are adequate for the target local ductility. Procedures
for checking the local ductility demand are presented in section 2.4.10.

2.4.6 P-delta effects

When a structure is displaced, the effect of gravity on the structure increases


lateral demand, an effect known as P-delta.

According to clause 6.5.1 of NZS 1170.5:2004, P-delta effects do not need to be


considered if (a) the largest translational period is less than 0.4 seconds or (b) the
height of the structure is less than 15 m from the base and the largest period is
less than 0.6 seconds. Although the lightweight nature of LTF buildings in general
leads to a decrease in the effects of P-delta, low initial stiffness and moderate
ductility targets can lead to stability coefficients θ that are greater than 0.1.

25
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The stability coefficient at each storey can be calculated as:

(2.4)
where:
θi stability coefficient at level i
Wi seismic storey weight at level i
Δstorey u,i ultimate inter-storey displacement at level i
Vi seismic shear demand at level i
hi inter-storey height at level i

If P-delta effects need to be considered, the more conservative Method A to


amplify loads and deformations from NZS 1170.5:2004 is adequate:

(2.5)

where:
kp (0.015 + 0.0075(μ – 1)) with limits of 0.015 < kp < 0.03
μ structural ductility factor
Vb seismic base shear
W t total seismic weight of the building

Where P-delta effects are seen as being prohibitive when calculated using
Method A, the more accurate Method B can be used.

2.4.7 Use of the capacity factor φ in ductile and capacity protected


elements

NZS AS 1720.1 establishes the capacity factors to be used in design. For the
design of sheathing fasteners or any other ductile element, the appropriate
capacity factor to that element should be used (ɸ = 0.8 for nails under lateral
load, for example).

For the capacity protected elements, the overstrength loading applied to the
element is derived from the strength of the ductile element being displaced
beyond its elastic limit (a large displacement). As such, a capacity factor of ɸ = 1
is used.

2.4.8 Calculation of non-linear horizontal deflections

LTF buildings are flexible, so careful consideration of deflections needs to be


made. Methods for calculating elastic deflections are described in section 3.3.1.4.

As described previously, when ductility is targeted in LTF buildings, it is


provided by the sheathing fasteners with all other elements capacity protected.
This essentially limits the structure to one possible sidesway mechanism,
meaning that the design horizontal deflections will be equal to the elastic
deflection envelope multiplied by the structural ductility factor μ:

26
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

(2.6)

where:
Δu,i ultimate deformation of level i
μ structural ductility factor
Δel,i elastic deformation of level i
kdt displacement amplification as per section 2.4.9

P-delta effects also impact on displacements. If P-delta effects need to be used,


as discussed in section 2.4.6, deflections are multiplied by the scale factor in
Equation (2.5) as described in NZS 1170.5:2004.

When calculating elastic deflections, designers sometimes use linearised force-


displacement curves to describe the slip of connections. Formulations often
consider distinct stiffness values for SLS and ULS analysis as per EN 1995-1-
1:2004 Eurocode 5: Design of timber structures – Part 1-1: General – Common rules
and rules for buildings for instance. Since all connections apart from the ductile
panel sheathing are designed for overstrength, their stiffness can be evaluated
using the assumptions based on SLS analysis and also when determining
ultimate deformations.

2.4.9 Additional deflection of LTF LLRS due to pinched hysteresis loops

The ULS deflection is calculated by multiplying the elastic deflections by the


structural ductility factor. This is based on the assumption that the system is able
to develop full hysteresis loops under reverse cyclic loading, akin to modern
steel and reinforced concrete structures. As shown in Figure 3, LTF walls are
characterised by pinched hysteresis behaviour due the irreversible crushing of
the timber around the fastener.

The pinching of the hysteretic loops leads to an LTF wall system having less
hysteretic damping capacity than assumed in the procedure presented in NZS
1170.5:2004 for a ULS displacement. As the structure is already non-linear and
without the ability to significantly increase its capacity, a reduction in hysteretic
damping will not impact on the design base shear of the structure. However,
it will increase ultimate displacements as the system continues to search for
energy dissipation.

A factor has been derived in NZS AS 1720.1 to allow for the impact of additional
ultimate displacements created by pinched hysteretic systems:

(2.7)

where:
kdt displacement amplification factor to be applied to the inelastic
displacement – see Equation (2.6).

2.4.10 Calculation of local ductility demand

The ductility of the LTF system is provided by the local yielding of fasteners, and
these must be checked to ensure that they have adequate local ductile capacity.

27
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Due to the use of capacity design, the inelastic displacement demand on the wall
due to nail slip is the only potential source of inelastic behaviour and therefore:

(2.8)

where:
Δ4,μ,i inelastic deformation component at the top of the wall due to nail slip
at level i
Δμ,i total inelastic deformation component at the top of the wall at level i
Δu,i total ultimate deformation at the top of the wall at level i
Δel,i total elastic deformation at the top of the wall at level i

The total displacement at the top of the wall due to nail slip then becomes:

(2.9)

where:
Δ4,u,i ultimate deformation at the top of the wall due to nail slip at level i
Δ4,i elastic deformation at the top of the wall due to nail slip at level i
(=Δ4,el,i)

It is possible to calculate the total fastener demand as follows (see also section
3.3.1.3.1):

(2.10)

where:
l wall length
hi inter-storey height
n number of horizontal joints in the sheathing panel
m number of vertical joints in the sheathing panel
en,u ultimate fastener slip

NZS AS 1720.1 provides maximum inelastic displacement limits for mild steel
nails equal to 6 mm. Alternatively, manufacturers’ testing data or research
literature could be used to established acceptable inelastic limits.

2.4.11 Calculation of inter-storey drift and drift limits

The calculation of the inter-storey drift is performed in accordance with


NZS 1170.5:2004 with the acceptable drift limits under ULS loading from this
standard being checked for compliance.

The following checks are recommended:


■■ The average storey drift in each direction.
■■ SLS drift without accidental eccentricity.
■■ ULS drift without accidental eccentricity including non-linear deflection
limits as described in section 2.4.8.

28
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

2.4.12 Podium structures and seismic loading

The use of podium structures, where one or more storeys of reinforced concrete
structure are positioned below the LTF framing, is very common in overseas
applications of LTF structures. This is often due to a requirement to create an
open space on the first floors of the building either for parking or commercial
activities. This may also include an increased inter-storey height for this first
floor.

Currently, guidance in the design of this type of structure is not explicitly


provided within NZS AS 1720.1. However, consideration of their unique seismic
performance, especially in the presence of non-linearity, is important.

Modal analysis can be used where the numerical model of the building
incorporates both the concrete and timber portions with their own stiffness
properties. The modal analysis of the structure should use the elastic code
spectrum with the timber portions of the building then having their shear force
and moment demands modified by the appropriate Sp and kμ factors to obtain
design forces. This approach will be iterative, as for the correct stiffness to be
input into the structural model, the LLRS will need to be already defined and
modelled as described in section 3.3.1.

Alternatively, ESA could be used by adopting a two-stage analysis procedure. In


this procedure, the interaction between the concrete and timber portions of the
building is virtually ignored.

In a two-stage analysis, the timber portion of the building is treated as a separate


structure supported on the ground. The base shears and overturning moments
from the timber portion are then applied to the concrete structure amplified by
the overstrength and overcapacity of the wall, with a limit of the elastic design
forces. For the two-stage approach to be used, the stiffness of the lower portion
must be 10 times the stiffness of the upper portion.

A form of the two-stage analysis procedure is presented in ASCE 7-16 Minimum


design loads and associated criteria for buildings and other structures. Additional
information on this two-stage approach is provided by Thompson (2015), but it
is worth noting that some of the philosophy in this document around seismic
design differs from that normally used in New Zealand. Therefore, should
Thompson (2015) and the accompanying example be used to determine the
seismic demand following the two-stage approach, designers should refer to the
BRANZ guidelines for the design of the timber structure.

29
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

STRUCTURAL ANALYSIS AND


3
DESIGN OF LTF BUILDINGS
3.1 Introduction
Buildings constructed using LTF must be analysed such that the systems and
components making up the whole of the building are able to resist the applied
loads with an adequate margin of safety. Limitations for ULS and SLS can be
found in AS/NZS 1170.0:2002. While these loads include wind, gravity, seismic
and live loads, it is also important to keep in mind that fire resistance, acoustic
performance and serviceability deflections including vibrations must also be
considered for a comprehensive design. Often in LTF buildings, these other
design parameters can impact the structural design. It is therefore important to
consider the sequence of the design process to optimise efficiency. Horizontal
diaphragms are a critical component for the design of LTF buildings and are
discussed in section 4 of these guidelines.

3.2 Gravity and wind design


Gravity loads including snow and the weight of the completed building create
vertical static loads that must be resisted by various building components as
described in the following sections. Similarly, wind loads create static horizontal
loads that must be transferred from the exterior cladding through walls into
horizontal diaphragms and from there to the LLRS, which is typically sheathing
clad shear walls in the buildings being considered. Although not specifically
related to the structure of these buildings, it is important to note that wind
can also create areas of higher pressure on walls and roofs, and these must
be accounted for when designing the cladding and roof systems. Proprietary
systems can be used for both of these, but designers must be aware that systems
used for buildings designed using NZS 3604:2011 and NZBC E2/AS1 may be
inadequate if the site wind speeds exceed 55 m/s. Manufacturers’ technical

30
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

information needs to be carefully scrutinised to ensure these applications are


within the scope of usage.

3.2.1 Floors

Since the floor elements (sheathing, floor joists and bearers) are generally
independent of building height for residential multi-storey buildings, the design
information as per NZS 3604:2011 can be applied.

Elements that are part of the LLRS may need additional verifications to ensure
that both gravity and lateral loads can be transferred under diaphragm actions
(i.e. joists or beams and their splices, which also act as collectors, chords or strut
beams). These aspects are further discussed in section 3.3. Floors in multi-storey
buildings may also need to be designed as diaphragms to effectively transfer
lateral forces to the LLRS, which is discussed in section 4.

Care shall also be taken when designing floor joists, as the design tables in
NZS 3604:2011 do not necessarily provide the required vibration performance
expected in structures with separate living units. More information regarding
vibration performance can be found in section 5.

3.2.2 Lintels and trimming studs

Lintels need to transfer all gravity loads from the roof or floor on top of the
opening and need to transfer the respective forces to the trimming studs on
either side. In the case where wall openings are aligned along the building height
(see Figure 4a), the lintel sections as per NZS 3604:2011 can be used provided
design wind speeds are less than 55 m/s. The trimming studs in multi-storey
timber buildings required to support the lintels need specific engineering design
due to the increased cumulative load from the storeys above and the increased
wind load. The load from the tributary area above a lintel supporting several
floors can be mitigated if the wall above can be shown to be acting as a deep
beam carrying the loading from the floors above. Trimmer studs can be designed
using the same considerations for the design of standard wall studs as presented
in section 3.2.3. Additionally, concentrated lintel forces from wind face loads and
internal pressures need to be considered.

tributary area

Figure 4. Design of lintels and trimming studs: aligned openings (left-hand side); non-aligned openings (right-hand side).

31
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

When designing openings that need to resist the full wall loads from the storeys
above (Figure 4b), both the lintels and the trimming studs require specific
engineering design. The lintels can be designed as simply supported beams as
shown in Figure 5 with the following load combinations from AS/NZS 1170.0:2002:

1. 1.35 G
2. 1.2 G + 1.5 Q
3. 0.9G – Wu
4. 1.2G + ψc Q + 1.0 Su

lintel span

Figure 5. Structural model used for strength design of lintels (Shelton, 2013).

A combination factor ψc of 0.4 is used for floors and 0 for roofs, and appropriate
area factors for imposed actions on floors ψa from AS/NZS 1170.1:2002 should
also be included.

When determining the demand on the lintels, all the load contributions from
the storeys and roof above need to be taken into account. Since the lintels can
be assumed to be restrained at the top by the wall framing and the trimming
studs provide a torsional restraint, the factor for stability k12 can be taken as
1.0. No load sharing with other adjacent members should be considered. The
connection between the lintel and the trimming studs for uplift forces can be in
accordance with NZS 3604:2011 provided the design wind speed is below 55 m/s.

Lintels above doors and windows are especially sensitive to displacements and
need to be designed for SLS under the following load combinations:

1. G + ψl Q
2. G + ψs Q
3. G + ψs Q + W s
4. G + ψs Q + S u

The short-term and long-term load factors are shown in Table 1.

Table 1. Short-term and long-term load factors.

Load Short-term factor (ψs) Long-term factor (ψl)

Live (floor) 0.7 0.4

Live (roof) 0.7 0.0

Snow 0.5 0.0

For Load Combination 1, the effect of long-term creep must be accounted for by
multiplying the elastic deflections by a modification factor for creep j2 (same as
k2 from NZS 3603:1993), equal to 2. A capacity factor ɸ of 0.8 is appropriate for
the lintel design.

32
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

3.2.3 Walls

3.2.3.1 Studs

The wall studs are required to carry gravity loads from the supported roof and
floors as well as wind face loads. It is possible that studs could be additionally
loaded by heavy wall claddings at height – for instance, by brick ties – and this
would require some additional consideration. Trimming studs also resist wind
loading being transferred to them by the lintel they support. This will be in the
form of point loading at each lintel end.

Since uplift forces in the studs are normally not governing, the studs are
designed for bending and compression as per Figure 6. Gravity, live, snow and
wind load demands are to be taken from the respective loading codes.

The gravity demand needs to include the dead and live loads from the roof,
floors and walls above. Once the load demands are determined per metre of wall
length, the concentrated force on each stud can be calculated as a function of
the stud centre-to-centre spacing.

p p

stud height W

p p

Figure 6. Structural models used for strength design of studs (Shelton, 2013).

Wind load demands are determined using methods provided in AS/NZS


1170.2:2011 based on building shape and dimensions and site characteristics as
described in section 2.3.

The load combinations described in section 3.2.2 also require demand


combinations for the studs. A combination factor ψc of 0 can be used for
roofs and 0.4 for floor loads in combination with wind face loading. When
determining the slenderness coefficient, an effective length factor g13 of 0.9
can be assumed as per NZS AS 1720.1. No other out-of-plane restraint for the
studs should be considered. For in-plane loading, the studs can be assumed as
restrained by the dwangs or blocking required to support the sheathing panels.
When built-up members are used, each stud should be calculated independently
with its respective load share.

33
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Considerations regarding the use of NZS 3604:2011

The provisions in NZS 3604:2011 regarding mechanically fastened


built-up members are not appropriate to ensure effective connection
under the higher load demands of multi-storey timber buildings.
For a more detailed analysis of mechanically jointed columns, the
stiffness and strength of the fasteners should be included to determine
slenderness if mechanical jointing is being relied upon for reduced
slenderness. A possible approach is the gamma method as per
Eurocode 5.

A modification factor for load sharing k9 of 1.1 can be assumed, recognising the
load sharing provided by dwangs, lining and cladding as used in Shelton (2013).

For SLS, the load combinations as per section 3.2.2 are considered. The modulus
of elasticity E can be used for the studs, recognising that every wall element
consists of a number of studs and that the use of the lower bound modulus Elb is
too conservative. A maximum deflection under face loading of height over 200
with an upper bound limit of 15 mm should be used based on Shelton (2013).

3.2.3.2 Top and bottom plates

Top and bottom plates are loaded in compression perpendicular to grain due
to wall gravity loads and need to be verified. This is because of the higher
number of storeys and therefore higher compression loads acting on some floor
elements in the gravity load path, especially end studs of bracing elements. It is
important to note that this is a very critical aspect of the design of multi-storey
LTF buildings and must be considered in a realistic and preferably a conservative
manner. Some work has been done around this topic in Australia, and at the
time of writing, work is under way to develop more detailed recommendations
for how to design these components effectively.

One very critical aspect is to ensure that strength and particularly stiffness
values perpendicular to grain are on the conservative side, and research in
New Zealand has suggested that published values for radiata pine may be
unconservative. Franke and Quenneville (2011) found that perpendicular
to grain compression strength values published for radiata pine could be
unconservative, representing up to twice as much as test results indicated.
This depended on the type of test performed and was particularly evident
when the specimens were loaded as they would be in LTF construction. This
would include bottom plates in a wall system, for example. Van Beerschoten
et al. (2013) found that, for LVL loaded in compression perpendicular to
grain, it was necessary to include the effects of stress spreading around
loading points, which is similar to bottom plate loading in an LTF wall. To
cope with this, van Beerschoten et al. recommended reduced allowable stress
values depending on the application.

Explanations on the behaviour and modelling around loading of plates


perpendicular to the plane of the wall are provided in Shelton (2013) and can be
used as a guideline for plate designs in larger buildings with the understanding
that increased height and wind forces will result in higher forces that need to be
resisted by these members.

34
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

3.2.3.3 Stud connections

Connections between studs and plates are primarily based on shear demand
created due to wind loads acting on the walls that are bearing on the more
rigid floor or roof systems. Assumptions that contributions from wall linings
can be ignored is consistent with NZS 3604:2011 designs and can be used
for larger buildings. It is possible that stud-to-plate connections could be
additionally loaded by heavy wall claddings at height – for instance, by brick
ties – and this would require some additional consideration.

Roof uplift and wall overturning are resisted by more robust fixings and
considered in section 3.3.1.3. For wind speeds up to 55 m/s, it is reasonable
to use the typical connections used for plate-to-stud connections provided
in NZS 3604:2011. For higher wind speeds and increased stud heights, the
connections should be verified based on wind loading demands determined
for walls as described above as the typical use of two 90 x 3.15 mm end nails
will likely not be adequate.

3.2.4 Roofs

The design of roof trusses is usually specifically done by roof and truss
manufacturers, who will need to be made aware that these buildings may have
resulting wind speeds greater than 55 m/s. It is important to make sure that the
roof is adequately connected to the wall top plates and studs using commercially
available connections. Designers will be required to provide load demands that
are determined when determining wind loads on the building as described in
section 2.3. Local pressure coefficients must be considered for roof claddings,
whose capacities should be provided in technical literature on proprietary
products. In cases where wind speeds calculated from AS/NZS 1170.2:2011 are
determined to be less than or equal to 55 m/s, it is feasible to use rafter and roof
designs provided in NZS 3604:2011.

3.3 Lateral load demand


Two methods of determining lateral load demand on the lateral load-resisting
elements will be discussed in these guidelines.

For structures classed as regular in accordance with clause 4.5 of NZS


1170.5:2004, an analytical model is used that is based on the assumption of
a rigid diaphragm distributing forces to walls using their respective shear
stiffness. The rigid diaphragm assumption can be justified by the normally
short spans of diaphragm portions between shear walls. Additionally, when
seismic loading governs the design, diaphragms are overdesigned because of
capacity design principles.

For complex and irregular structures, a numerical modelling procedure


is presented. This allows the engineer to specifically allow for the various
stiffness contributions within the LLRS. Using this numerical modelling
procedure also allows the inclusion of specific diaphragm stiffness using
methods to be discussed in section 4. The numerical modelling procedure
requires prior understanding of the structural system in order to define
appropriate load distribution, therefore iteration is usually required. A first
step can be to use the assumptions of the analytical model, combined with
wind demand, to provide the first definition of wall composition including
member sizes and fastener schedule.

35
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

3.3.1 Analytical modelling

3.3.1.1 Force distribution into shear walls

Unless a more precise model with real wall and diaphragm stiffness is used
(see section 3.3.2), a rigid diaphragm is assumed when determining the force
distribution into the shear walls from seismic or wind demand. The following
sections are described for actions in the x-direction. Therefore, in practice, the
analysis would need to include actions in the y-direction as well.

3.3.1.1.1 Determination of centre of rigidity and mass

When assuming a rigid diaphragm and working with seismic loads, both
geometrical eccentricities as well as accidental eccentricities need to be
considered. The geometrical eccentricity is the distance between the centre
of mass and the centre of rigidity. The centre of rigidity at each level can be
calculated as:

(3.1)

where:
yR distance between the centre of rigidity and origin in the y-direction
yi distance between the centre of wall i and origin in the y-direction
kx,i stiffness of wall i in the x-direction

Since the wall deflection is mainly governed by shear deformation from


the sheathing panels and fastener slip, the wall stiffness can be assumed as
proportional to the wall length:

(3.2)

where:
li length of wall i

The centre of mass at each level is calculated as:

(3.3)

where:
yM distance between the centre of mass and origin in the y-direction
Ai area i on given floor
yi distance between the centre of area i and origin in the y-direction
Wi seismic weight on area i

36
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

3.3.1.1.2 Determination of eccentricities

The eccentricity of the equivalent force creating a moment at each level is


calculated as:

(3.4)

where:
ey eccentricity of the equivalent force in the y-direction
yM y coordinate of the centre of mass
yR y coordinate of the centre of rigidity

In the case where the eccentricity varies between the different floor levels,
the acting eccentricity on each floor needs to be calculated based on the
position of the equivalent forces above the floor level under consideration:

(3.5)

where:
ey,i eccentricity of the equivalent force at level i due to all equivalent
forces from the levels above in the y-direction
ey,j eccentricity of the equivalent force at level j in the y-direction
Fx,j equivalent force at level j in the x-direction

In accordance with NZS 1170.5:2004, the accidental eccentricity needs to be


taken as ±0.1b and needs to be added to the eccentricity at each floor:

(3.6)

where:
e’y eccentricity of the equivalent force including accidental eccentricity
by building dimension perpendicular to earthquake direction

When working with wind loads, the geometrical eccentricity is the distance
between the point of application of the wind resultant at each façade and the
centre of rigidity. Normally, the resultant wind force is applied at the midpoint
of the façade. Accidental eccentricity does not need to be considered. The
eccentricity is calculated as follows:

(3.7)

where:
by building dimension perpendicular to wind direction

37
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

3.3.1.1.3 Determination of the torsional moment

The torsional moment (due to eccentricities) at each floor due to equivalent


forces from all forces above the floor under consideration can be calculated as:

(3.8)

where:
MT,x,i torsional moment at level i from loads in the x-direction
Fx,j equivalent force at level j in the x-direction
Vx,i storey shear force at level i in the x-direction
e’y,i eccentricity at level i in the y-direction

The equivalent storey forces are distributed into the shear walls based on
their rigidity. This distribution needs to consider both the horizontal force
and the torsional moment due to eccentricities. A force in the x-direction will
activate all walls in the x-direction, and due to the torsion, an additional load
component needs to be considered both in the x and y-directions. Since some
load components due to torsion may act in the opposite direction from the main
driving force, the resultant shear force in the wall will be reduced.

3.3.1.1.4 Determination of the wall shear forces

The horizontal force due to external loads based on a rigid diaphragm


assumption and including eccentricities for each shear wall m at each level i in
the x-direction is calculated as follows:

(3.9)

where:
Vx,i[m] shear force of wall m at level i in the x-direction
Vx,i storey shear force at level i in the x-direction
MT,x,i torsional moment at level i from loads in the x-direction
kx stiffness of the wall in the x-direction
ky stiffness of the wall in the y-direction
x’ distance between the centre of the wall and the centre of stiffness in
the x-direction
y’ distance between the centre of the wall and the centre of stiffness in
the y-direction

The superscript [m] indicates that the variable is specific to wall m – the value in
the summations consider all walls in the structure.

The horizontal force in the y-direction can be calculated by swapping indices x


and y. When determining the force demand in the y-direction from the load in
the x-direction, the first term of Equation (3.9) will be 0, as Vy,i = 0.

38
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Under the assumption that the wall rigidity is governed by shear stiffness and
therefore proportional to the wall length, Equation (3.9) can be written as:

(3.10)
where:
l wall length

In the case where the wall properties change up the building height, in addition
to using Equation (3.5) for calculating the acting eccentricity, the stiffness and
location of each wall needs to be considered for each specific floor level:

(3.11)
where:
kx,i stiffness of the wall at level i in the x-direction
ky,i stiffness of the wall at level i in the y-direction
x’i distance between the centre of the wall and the centre of
stiffness at level i in the x-direction
y’i distance between the centre of the wall and the centre of
stiffness at level i in the y-direction

Since multi-storey LTF buildings are generally designed for a ductility of 3


and are hence not classified as nominally ductile nor brittle, no additional
earthquake actions in the orthogonal direction need to be considered.

3.3.1.2 Determination of wall demand

3.3.1.2.1 Horizontal force on each level

The horizontal force applied to each shear wall at each level is calculated directly
from the wall shear forces:

(3.12)
where:
Fi = ΔVi horizontal force on the wall at level i
Vi shear force of the wall at level i
Vi+1 shear force of the wall at level i+1

3.3.1.2.2 Overturning moment

The overturning moment at each level for each wall m is calculated as follows:

(3.13)
where:
Mi overturning moment at the base of wall at level i
Vi shear force of the wall at level i
hi height of shear the wall at level i

39
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

3.3.1.2.3 Shear flow

The shear flow v for each wall at level i is calculated based on the wall shear
forces:

(3.14)

where:
vi* shear flow (unit shear force) demand of the wall at level i
li length of the wall at level i
Vi shear force of the wall at level i

3.3.1.2.4 Normal forces (chord forces)

Based on the overturning moment, the tension and compression forces for each
wall at each level i are calculated as:

(3.15)

where:
Ni* axial forces in chord members of the wall at level i
Ti* tension force demand in chord of the wall at level i
Ci* compression force demand in chord of the wall at level i
Mi overturning moment at the base of the wall at level i
lc,i distance between the centre of the compression chords and tension
hold-down of the wall at level i

3.3.1.3 Verification of wall component capacity

3.3.1.3.1 Fastener capacity

The shear connection capacity (nails, screws, staples etc.) is based on the
individual fastener capacity and the fastener spacing:

(3.16)

where:
vi* shear flow (unit shear force) demand of the wall at level i
ɸ capacity factor
k modification factors for strength according to NZS AS 1720.1
k1 duration of loading
k13 = 1 for nails in side grain
k14 = 1 for nails in single shear
k16 = 1.1 for plywood or particleboard
k17 = 1.2 for connections with more than 50 nails
s fastener spacing
Qk characteristic fastener capacity

40
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

3.3.1.3.2 Sheathing panel capacity

The sheathing panels capacity verification depends on the specific material used.
In the case of plywood sheathing, the following verification shall be used:

(3.17)
where:
vi* shear flow (unit shear force) demand of the wall at level i
ɸ capacity factor
k modification factors for strength according to NZS AS 1720.1
g19 modification factor for plywood assembly
fs’ characteristic value in panel shear
t full thickness of plywood (sum of both panels if the walls are lined on
both sides)

3.3.1.3.3 Chord capacity

The tension and compression chords are to be verified as tension members and
columns, respectively.

The tension chords need to be verified as follows:

(3.18)
where:
Ti* tension force demand in chord of the wall at level i
ɸ capacity factor
k modification factors for strength according to NZS AS 1720.1
f t’ characteristic value in tension parallel to grain
A t net cross-sectional area of tension member

The compression chords need to be verified as follows:

(3.19)

where:
Ci* compression force demand in chord of the wall at level i
ɸ capacity factor
k modification factors for strength according to NZS AS 1720.1
f c’ characteristic value in compression parallel to grain
A c cross-sectional area of column

When determining the stability factor k12, buckling around both axes needs to
be considered. When compression members are made of multiple (built-up)
members, the following approaches with increasing accuracy can be used when
determining the buckling length about their minor axes:
■■ Calculate each individual member as an independent column member without
intermediate restraints with its respective share of the compression force.
■■ Calculate each individual member as a column member with intermediate
restraints provided by blocking or dwangs with its respective share of the
compression force.

41
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

■■ Calculate a mechanically connected built-up member based on stiffness


considerations (i.e. gamma method as per Eurocode 5) with intermediate
restraints if present with the total chord force.
■■ If the sheathing panels can be used to provide continuous restraint of the
compression member, the additional force component perpendicular to
the chord axes needs to be considered as well as the increased minimum
fastener spacing. Eurocode 5 provides a method to determine the additional
load in the continuous bracing elements required to prevent buckling.

For both tension and compression chords, the panel shear fasteners should be
evenly distributed to all built-up members.

3.3.1.3.4 Bearing capacity

The contact area between the compression chord and the bottom plate of the
wall undergoes compression stresses perpendicular to grain. Even if this bearing
capacity is verified, it can cause local compression deformations that can
increase the wall flexibility.

The bearing capacity is verified as follows:

(3.20)
where:
Ci* compression force demand in chord of the wall at level i
ɸ capacity factor
k modification factors for strength according to NZS AS 1720.1
fp’ characteristic value in compression perpendicular to grain
Ap bearing area for loading perpendicular to grain

Due to the potentially high chord compression forces in multi-storey buildings,


the bearing capacity may be a limiting factor in design and require larger contact
areas and hence chord cross-sections. This can be avoided by using horizontal
members with a higher compression perpendicular to grain capacity or local
reinforcement (fully threaded screws or dowels in contact with steel plates) or by
choosing a design where the load path does not involve any horizontal members
(i.e. continuous studs).

3.3.1.3.5 Hold-down capacity

The tension forces in the chord members need to be transferred between the
walls at adjacent levels. This connection can be realised by means of steel straps,
proprietary hold-down anchors or continuous tension rods. It is recommended
that the hold-down capacity be verified following the appropriate Code
verification or based on manufacturers’ data.

Walls need to be anchored into the foundation system. This connection can be
done with cast-in anchors or chemical/mechanical anchors. Capacity checks
normally require the verification of the anchoring system as well as concrete
strength and minimum distances to edges.

3.3.1.3.6 Shear connection capacity

At each level, the diaphragms transfer additional horizontal forces into the shear
walls. The diaphragm panels are normally connected directly to the top plate of

42
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

the wall, which is connected to the wall sheathing panels. The connection can be
verified as per section 3.3.1.3.1.

At the bottom storey, the bottom plate of the wall needs to transfer the base
shear into the foundation system. This connection can normally be carried
out with proprietary shear connectors or by bolting the bottom plate into the
foundation. For the latter, the bolts/anchors need to be verified for both the
foundation system and the timber.

3.3.1.4 Shear wall deformation

The deformation contributions in a cantilevered shear wall made of LTF can be


summarised as follows:
■■ Flexural deformation due to chord elongation/shortening.
■■ Shear deformation of the sheathing panels.
■■ Fastener slip in the panel connections.
■■ Rotation of the wall due to chord anchorage deformation/slip and bottom
plate compression perpendicular to grain.

These deformation contributions are well documented and codified for single-
storey structures. In the case of multi-storey structures, the walls are also loaded
by a moment in addition to the storey shear force. Further, the cumulative
rotation from all levels below the wall under consideration also needs to be
accounted for. This rotation is due to the flexural deformation from the storey
shears and overturning moments as well as the anchorage deformation.

The deformation contributions are described in the next sections. The equations
can be implemented in a spreadsheet and used to determine SLS and ULS
deflections of the individual walls. The deflection equations can also be used to
determine the building period and can be used to more accurately determine
the force distribution in the individual shear walls. An iterative procedure
will be required. The wall deformation taking into account all of the required
components (see Figure 7) is the following:

(3.21)

where:
Δ1,i flexural deformation due to horizontal storey shear force Vi at level i
Δ2,i flexural deformation due to overturning moment Mi at level i
Δ3,i shear deformation of the sheathing panel at level i
Δ4,i deformation due to fastener slip at level i
Δ5,i deformation due to wall rotation from anchorage system deformation
Δ6,i deformation due to cumulative wall rotation from levels below

When calculating the wall deflections for both the SLS und ULS cases, the mean
modulus of elasticity and mean shear modulus shall be used. All deformations
deriving from anchorage slip can be considered as elastic, as they are designed
as capacity protected elements. This means that an initial or tangent elastic
stiffness can be assumed also under ULS loading.

Under ULS loading, the only elements sustaining essential deformation are
the panel fasteners, whose deformation is normally determined from a non-
linear force displacement relationship (i.e. NZS AS 1720.1). If an approximated

43
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Δ1 and Δ2 Δ3

Δ1 + Δ2 flexural deformation Δ3 shear deformation due to panel


due to Vi and Mi shear deformation

Δ4 Δ5

Δ4 fastener slip Δ5 rotation due to


anchorage deformation

Figure 7. Wall deformation contributions.

approach with linear fastener slip Kser is used, the ultimate slip modulus
Ku = 2/3 Kser as per Eurocode 5 is used.

3.3.1.4.1 Flexural deformation

Flexural deformation is created by the extension and shortening of the wall


chord elements and is calculated as:

(3.22)

(3.23)
where:
Δ1,i flexural deformation due to horizontal storey shear force Vi at level i
Δ2,i flexural deformation due to overturning moment Mi at level i
hi inter-storey height
lc distance between chords
EI bending stiffness of the wall
EA axial stiffness of the tension and compression chords

In the case of a continuous rod system, the timber tension chord is


effectively loaded in compression as the chord force is taken by the tension
rod and transferred into the top of the timber chord. Hence, the same
equations as above can be used.

44
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

3.3.1.4.2 Panel shear deformation

Panel shear deformation is created by the shear distortion of panel sheathing as


during loading and is calculated as:

(3.24)

where:
Δ3,i shear deformation of the sheathing panel at level i
Vi horizontal storey shear force at level i
hi inter-storey height of level i
G shear modulus of sheathing panel
As shear area of sheathing panel (can be cumulative if both sides of the
wall are lined)
d sheathing panel thickness (can be cumulative if both sides of the wall
are lined)
l wall length

3.3.1.4.3 Fastener slip

The fasteners in the wall provide a major source of deformation. They will also
provide the only source of inelastic deformation under seismic loading. The wall
deformation due to fastener slip is calculated as:

(3.25)

where:
Δ4,i deformation due to fastener slip at level i
l wall length
h inter-storey height
n number of horizontal joints in the sheathing panel
m number of vertical joints in the sheathing panel
eni fastener slip resulting from the force on the individual fastener force
F1 due to the wall shear force Vi

The value for en can be determined by testing or can be calculated as per


NZS AS 1720.1. The force on a single fastener F1 can be calculated as:

(3.26)

where:
F1 force on a single fastener due to Vi
s fastener spacing

Instead of calculating the exact fastener slip en, following a non-linear force-
displacement curve, a linear approximation of the displacement can be used.
Eurocode 5 provides the slip modulus Kser for different fastener types.

45
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The approximated fastener displacement can be calculated with the


following equation:

(3.27)
Equation (3.25) then becomes:

(3.28)

When using a linear approximation to determine the fastener slip, the


ultimate fastener slip Ku = 2/3 Kser should be used when determining the
elastic deformation components under ULS loading.

Equation (3.25) is the same as the equation provided in NZS AS 1720.1:

(3.29)
where:
α h’/b’ sheathing panel aspect ratio, where h’ is the length of the
sheathing panel edge parallel to the chords
m’ number of panels along the chord beam

The difference between Equation (3.25) and Equation (3.29) is that the
former considers the number of sheathing panel splices as an integer and
the latter considers the number of panels as a ratio between wall and
sheathing panel dimensions. The equations lead to the same result if a large
number of splices are present. For typical shear walls in an LTF building,
the real number of splices by using Equation (3.25) should be used.

3.3.1.4.4 Rotation due to wall anchorage system deformation

In addition to fastener slip, the rotation due to the wall anchorage system also
provides a major source of deformation and is calculated as:

(3.30)
where:
Δ5,i deformation due to wall rotation from anchorage system deformation
δT elongation of hold-down/anchorage system due to chord tension forces
δC compression of bottom plate (or any horizontal timber member
between adjacent walls) due to chord compression forces
lc distance between centre of chords
hi inter-storey height

If the stiffness of the hold-down/anchorage system and the stiffness of timber


perpendicular to grain are known, Equation (3.30) can be written as:

(3.31)

46
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

(3.32)

where:
θi wall rotation at level i due to anchorage slip
Kt tension stiffness of hold-down/anchorage system
Kc compression stiffness of wall plates
lc distance between centre of chords
Mi overturning moment at level i
Kθi rotational stiffness at level i due to anchorage slip

Kt is normally provided in the manufacturers’ data or can be calculated based


on first principles. Since the hold-downs are normally designed as capacity
protected elements under seismic loading, the tangent slip modulus Kser (rather
than the ultimate value Ku = 2/3 Kser as per Eurocode 5) can be used.

When the compression chord sits on a horizontal timber member (i.e. bottom
plate, stringer, boundary beam or joist), the compression stiffness Kc can
normally be estimated as:

(3.33)

where:
Ap bearing area for loading perpendicular to grain
h c depth of wall plates and joists if appropriate with grain perpendicular
to grain direction
E90 modulus of elasticity perpendicular to the grain of the horizontal
timber members between two wall elements along the building height

If compression perpendicular to grain is avoided by local reinforcement or by


removing any perpendicular to grain contact as mentioned in section 3.3.1.3.4,
Kc can normally be assumed as infinitely rigid.

If more than one stiffness contribution acts in series, this needs to be


accounted for accordingly (i.e. hold-down above floor, hold-down under floor,
tension bolt etc.):

(3.34)

A more accurate value of Kc can be obtained by using the approach


provided in Blass and Görlacher (2004).

Due to stress spreading in the bottom plate (Figure 8), the compression
stiffness Kc can be increased as follows:

47
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

(3.35)

(3.36)

(3.37)

where:
c length of bearing area along plate
ξ c correction factor for plates on continuous support (foundation)
ξd correction factor for plates on discrete support (intermediate storey)

c c

hc

(a) continuous support (b) discrete support

Figure 8. Stress spread in bottom plates for continuous and discrete supports.

When a continuous tie rod system is used instead of discrete hold-downs,


all deformation components of this anchorage system needs to be taken
into account (i.e. rod elongation, anchor plate crushing, coupler elongation
etc.). When using proprietary systems, the displacement of the anchorage
system is normally provided.

Note that, when a continuous tie rod system is used, potential shrinkage
in the anchorage system needs to be considered. This can be avoided if
shrinkage compensators are used.

3.3.1.4.5 Cumulative rotation from all wall levels below

The rotation of the wall is cumulative up the wall height with each rotation
adding to the rotation of the wall above as shown in Figure 9. This creates the
cumulative rotation defined as:

(3.38)

48
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

where:
Δ6,i deformation due to cumulative wall rotation from levels below
Θ cumulative rotation of all walls under level i
θ rotation due to wall anchorage system deformation as per section
3.3.1.4.4
hi inter-storey height at level i
αV rotation due to flexural deformation from horizontal shear force
αM rotation due to flexural deformation from overturning moment

(3.39)

(3.40)

lc distance between chords


EI bending stiffness of the wall
EA axial stiffness of the tension and compression chords

(Δ3 + Δ4)
Δ5
(Δ1 + Δ2)
Δ6
3
Θ3

2 Θ2 θ1

(αF + αM)i
1

Θ3
Θ2

Θ1

Figure 9. Cumulative rotation of walls.

3.3.2 Numerical modelling

The numerical modelling procedure includes creation of a three-dimensional


numerical representation of the structure comprising stick elements and
rotational springs as shown in Figure 10. The stick elements have an effective
flexural stiffness to account for chord shortening and elongation (due to bending
of the wall) and an effective shear stiffness to account for panel stiffness and
fastener slip. Rotational springs at the base of each wall represent the stiffness of
the anchorage system.

49
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

sheathing panel

sheathing panel splice Eleqi


GAeqi

tension/compression chord KΘi

anchorage system

Figure 10. Spring and stick representation of LTF wall.

The application of the numerical representation requires that each wall be


predefined, which can be done either using previous design experience of
through the use of the analytical model in section 3.3.1.

Once defined, a practical way to implement the equivalent stick model is to


input a solid element of the correct length l and width b and providing it with the
effective stiffness using the equations provided below. For the length l, the actual
wall length should be used, and the difference between the wall length l and the
distance between chords lc can usually be neglected. For the wall width b = beq,
either the sheathing panel thickness or the depth of the studs can be used.

3.3.2.1 Equivalent flexural stiffness

The equivalent flexural stiffness of the wall is defined as:

(3.41)
where:
Eeq equivalent elastic modulus of numerical stick element
E elastic modulus of chord elements
Achord area of chord element
lc distance between centre of chords
beq equivalent wall width

The stick member in the model shall be modelled with its equivalent dimensions
lc and beq.

3.3.2.2 Effective shear stiffness

The effective shear stiffness of the wall is defined as:

(3.42)

50
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

where:
Geq equivalent shear modulus of wall
beq equivalent shear wall thickness
h inter-storey height
d sheathing panel thickness (can be cumulative if both sides of the wall
are lined)
l wall length
s fastener spacing
Kser fastener slip modulus
n number of horizontal joints in the sheathing panel
m number of vertical joints in the sheathing panel

The factor 1.2 is required, as classic beam theory implemented in commercial


computer software considers average shear stresses when calculating shear
deformation. Testing has shown that the shear stresses in sheathed shear walls
and diaphragms are constant, and hence equivalent shear stiffness needs to be
artificially increased (ATC, 1981; Smith, Dowrick & Dean, 1986).

3.3.2.3 Rotational wall stiffness

The rotational wall stiffness at the bottom of the wall due to the anchorage slip is
defined as:

(3.43)

where:
Kθ rotational stiffness due to anchorage slip
Kt tension stiffness of hold-down/anchorage system
Kc compression stiffness of wall plates
lc distance between centre of chords

3.4 Summary
This section has provided methods for designing the majority of systems
and elements required for LTF buildings based on the load demands and
requirements for ULS and SLS. Design of diaphragms is covered in section 4 as
there are more complex loading and connection requirements to be considered
as well as methods of analysis that should be used. Load demands including
gravity, wind, live and seismic have been covered in this section, while following
sections will cover other performance states such as fire, vibrations and acoustic
separation that must be taken into account for an effective and holistic design of
a multi-storey LTF building. As noted previously, the sequence of design should
be considered so that the building is designed in a logical and efficient manner
that incorporates all of these aspects together rather than in isolation.

51
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

DIAPHRAGM DESIGN
4
FOR LTF BUILDINGS

4.1 Introduction
Research and observations following earthquakes have made it clear that
diaphragm design in multi-storey buildings must be considered in order to
provide robust and safe buildings. Diaphragms in LTF buildings of 3–6 storeys
should be designed using specific engineering design methods because the
assumptions inherent in the floor and ceiling designs provided in NZS 3604:2011
are not likely to provide satisfactory designs when larger loads are required to
be transferred and separation is required between residential units. Connection
details between diaphragms and shear walls also require attention to ensure that
loads are safely transferred between these systems when subjected to design
level lateral loads. This section provides definitions and explanations on how to
effectively design diaphragms for multi-storey LTF buildings.

4.1.1 The role of diaphragms

Suspended floors and roofs not only carry gravity forces and transfer them to
the next vertical loadbearing element, but by providing diaphragm action (in-
plane load transfer), they also resist and transfer horizontal forces from seismic
or wind loading into the bracing walls or other LLRS. However, diaphragms not
only provide the load path for these lateral forces but also tie structural and non-
structural elements together providing critical building integrity.

Additionally, due to the differential deformation behaviour of different lateral


load resisting systems (LLRSs, combination of walls and frames/portals, bracing
walls made of different materials, bracing walls of different lengths etc.),
diaphragms also generally need to accommodate deformation compatibility

52
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

forces. However, these forces are typically small in LTF buildings and can be
neglected in their design. In the case of mixed LLRSs (steel braces or portals,
LTF walls with different sheathing type or hold-down connections, presence
of mass timber bracing walls, stiff cores around lift shafts or staircases etc.),
deformation compatibility forces should be taken into consideration.

The roles of diaphragms are summarised in clause C5.7.1 of NZS 1170.5 Supp 1:2004
Structural design actions – Part 5: Earthquake actions – New Zealand – Commentary.

4.1.2 Components of diaphragms

The main components of a timber diaphragm can be grouped as (see Figure 11


and Figure 12):
■■ plate elements (diaphragm sheathing including connections)
■■ chord beams
■■ collector beams
■■ drag/strut beams
■■ connections to the LLRS.

During seismic events, forces are generated by floor accelerations acting on


masses that primarily consist of the structural elements such as the floor itself
and walls, as well as people and furniture etc., which are generally located on
the diaphragm plate elements. The plate element resists and transfers the unit
shear forces (forces per unit length) from these forces, and the chord beams
resist diaphragm bending via compression and tension forces. Around openings
and re-entrant corners, strut or drag beams collect the shear forces from the
disturbed area and transmit them into other parts of the diaphragm. Collector
beams transfer the horizontal diaphragm forces from the panel element into the
lateral load resisting elements. This force transfer can be at a point (i.e. a cross-
brace or cantilevered column) or continuous over a certain length (i.e. in bracing
walls), with the latter being the most common case in LTF buildings.

connection to LLRS connection to collector

opening connection to LLRS


wall panel
chord

collector column

chord

frame beam collector


plate element
column
lateral loading

Figure 11. Definitions of diaphragm components.

53
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

shear walls

chord beam

shear walls re-entrant corner


strut beam

chord beam

opening collector beam

chord/strut beam
strut
strut beam
collector beam

chord beam chord beam discontinuous diaphragm chord

Figure 12. Irregular floor geometry with typical diaphragm components for the given lateral load direction shown with arrows.

4.1.3 Timber diaphragm materials

Typical floor diaphragms in LTF buildings are made of plywood, particleboard


or oriented strand board sheathing panels nailed onto framing elements.
Framing elements can be made of solid wood, glulam or LVL as well as built-up
members such as I-beams or trusses. Individual sheathing panels are connected
by metallic fasteners such as nails, screws or staples. To avoid squeaking of the
floors, panels are often glued to the joists also. However, for the purpose of these
guidelines, gluing of sheathing will be neglected due to its brittle behaviour
and challenges with site monitoring. Glue should not be relied on to transfer
diaphragm forces.

The use of mass timber panels (glulam or CLT, for example) is also a viable
option for flooring within an LTF structure (Figure 13). For this solution,
the thicker engineered timber panels carry the gravity loads and also resist
diaphragm forces without the need for framing elements. In terms of diaphragm
design, only minor differences exist, and these will be discussed within shaded
boxes in this section.

Figure 13. Light timber frame (left) and mass timber diaphragm (right) examples, with schematic cross-sections.

54
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

4.1.3.1 Blocked and unblocked diaphragms

The primary mechanism behind diaphragm action is the transfer of shear forces
in the plane of the diaphragm panel. The principle behind this is explained
in the following sections and requires that shear transfer can occur along all
individual sheathing panels. This requires that all sheathing panel edges are
blocked (Figure 14). No flying (unblocked) joints are allowed.

blocked diaphragm unblocked diaphragm

Figure 14. Blocked (left) and unblocked (right) diaphragms.

Considerations for the use of unblocked diaphragms

Unblocked diaphragms are defined as diaphragms where not all sheathing


panel edges are connected to each other by blocking elements and generally
only provide a fraction of the capacity of an equivalent blocked diaphragm.
While often undesirable for most designs, unblocked diaphragms are often
more economic to build but result in displacements four times larger than
blocked diaphragms (Kessel & Schönhoff, 2001). They should therefore be
used only in situations with limited spans and reduced loading unless the
details have been shown to adequately transfer the diaphragm forces for the
specific situation.

Unblocked diaphragms withstand loads with a completely different


mechanism like the moment couple series normally used to design
diaphragms made of transverse boards based on a Vierendeel truss analogy.
This causes an additional force demand in the remaining fasteners and
causes force components perpendicular to the framing elements as shown
in Figure 15 (blue arrows). Because of the high loads, the connectors close
to the unsupported edge can undergo extensive yielding and therefore lead
to large deformations. This mechanism does not lead to immediate brittle
failure but leads to very large displacements. Guidance on the design of
unblocked diaphragms can be found in the German national appendix to
Eurocode 5 (DIN EN 1995-1-1/NA:2013 National annex – Nationally determined
parameters – Eurocode 5: Design of timber structures – Part 1-1: General –
Common rules and rules for buildings).

55
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

diaphragm
blocked
diaphragm
unblocked

diaphragm layout
diaphragm
blocked
diaphragm
unblocked

fastener demand on sheathing panels


diaphragm
blocked
diaphragm
unblocked

fastener demand on framing elements

Figure 15. Blocked (top) and unblocked (bottom) diaphragms, fastener demand on
sheathing panels and framing elements, modified from Kessel & Schönhoff (2001).

56
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

4.1.4 Diaphragm design to NZS 3604:2011

The design of diaphragms in NZS 3604:2011 can be divided into normal floors,
floor diaphragms and ceiling diaphragms. For all cases, standard detailing in
terms of nailing pattern, framing details, blocking requirements and chord
splices exist. For normal floors, the maximum distance between bracing lines
is limited to 6 m. For floor diaphragms, the maximum length can be 12 m
with a maximum aspect ratio of 2. The design principles behind the design of
floor diaphragms are not known (Shelton, 2013) and are likely based on good
practice and engineering judgement. Testing has shown that NZS 3604:2011 floor
diaphragms built as per NZS 3604:2011 can resist a unit shear force of 1.6 kN/m
(Shelton, 2004). By using this value in a girder analogy, larger chord strengths
than are currently provided by codified top plate splices are required.

The unit shear forces in multi-storey buildings are generally larger than those
expected for structures within the intended scope of NZS 3604:2011, therefore,
design principles as per NZS 3604:2011 should not be applied to buildings
outside of the scope of NZS 3604:2011. Specific engineered design is expected to
be required for most structures within the scope of these guidelines. This section
outlines procedures for the specific engineering design of diaphragms for use in
multi-storey LTF buildings.

4.1.5 Forces from higher modes on diaphragms during seismic


loading

Higher modes in flexible mid-rise buildings can possibly lead to higher inertia
floor forces than predicted with equivalent static analysis (ESA) as discussed
in section 2.4. These increased floor forces are random in direction, and their
maxima do not occur at the same time at each floor. For these reasons, even a
more sophisticated modal analysis might not capture acceleration peaks in the
diaphragm as it provides only an envelope of maxima that do not occur at the
same time and are not in equilibrium.

LTF buildings are generally built with wood-based diaphragm sheathing material
nailed to timber framing. This construction system provides a high level of
ductility and has historically performed well under earthquake actions in
laboratory and historical seismic events. Past performance and research have
also shown that nominal amounts of localised yielding in the timber diaphragms
do not affect performance and that the floor forces are reduced when compared
to purely elastic-behaving diaphragms (Fleischman, 2014).

Due to this reason and as discussed in the next section, the effects of higher
modes on diaphragm inertial forces for structures up to 6 storeys can be
considered to be within the envelope provided by the pseudo equivalent static
analysis (pESA) as described in section 4.2.2.

4.1.6 Capacity design and diaphragms

The capacity design procedure described in section 2.4.2 only needs to be


applied for seismic loading and requires the selection of a mechanism with
designated ductile regions or ductile connections.

To comply with the New Zealand seismic loading standard NZS 1170.5:2004,
diaphragms should be designed elastically. This infers that diaphragms should
have the capacity to resist probable design loads and redistribute these loads

57
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

within themselves in a manner compatible with the assumptions used for the
LLRS design. Some localised plasticity in the diaphragm can be allowed provided
it does not alter the basic assumption of the diaphragm mechanism and that the
gravity resisting functions of the diaphragm elements are not compromised (NZS
1170.5 Supp 1:2004).

To guarantee that local plasticity in diaphragms can occur and does not cause any
detrimental effects, all non-ductile diaphragm components (wooden sheathing,
chord and collector beams, connections with stocky bolts or cold drawn/hardened
steel such as type 17 screws etc.) theoretically need to be designed based on
capacity design principles (i.e. for the overstrength of the ductile fasteners of the
diaphragm panel splices). However, the practical application of these principles
is complex. Therefore, it is recommended that some margin be placed between
the ductile sheathing fasteners and the non-ductile capacity protected elements
even if strict capacity design principles are not adhered to. To do this, a 20%
margin is placed between the demand used to design the nailed sheathing and the
demand used to design all non-ductile diaphragm components. The maximum
limit of ESA floor forces for overstrength actions as per NZS AS 1720.1 (based on
the elastic spectrum with µ = 1 and a structural performance factor Sp as per the
ductile structure under consideration) does, however, remain applicable for the
determination of the diaphragm demand for the nailed sheathing connections.

This effectively leads to two levels of hierarchy in the diaphragm – LLRS ensuring
resilient load carrying behaviour is achieved as shown in Figure 16. Using this
second level of capacity design, high demands may be imposed in the capacity
protected diaphragm elements. However, the demand effectively cannot be
larger than the demand derived from an elastic ESA.

Design of capacity protected diaphragm elements based on


the minimum between 1.2 times the pESA forces or maximum
floor forces for overstrength actions as per NZS AS 1720.1

Design of ductile diaphragm elements based on pESA with average


overstrength of ductile LLRS

Design of ductile LLRS based on ESA

Figure 16. Strength hierarchy in diaphragm design.

For structures of more than 4 storeys where the diaphragm panel connections are
not failing in a ductile manner (i.e. panels connected with non-ductile screws or
glued connections without the presence of yielding metallic fasteners) or where
capacity design principles are not adhered to within the diaphragm, the effect of
higher modes should be explicitly considered. This can be done using time history
analysis or the inclusion of dynamic amplification factors from literature.

Any potential ductile response of the diaphragm is not considered to increase


the ductile capacity of the LLRS, and the use of capacity design within the
diaphragm cannot be considered to provide the LLRS with ductile capacity.

58
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

4.1.7 Transfer and deformation incompatibility forces

Inertial forces are generally not the only action to be considered in diaphragms.
Structures with non-uniform LLRSs up the building height can generate
large additional transfer forces in diaphragms. These are caused by the force
redistribution (transfer forces) and displacement incompatibility (deformation
compatibility forces) between LLRSs with different stiffness or deformation
patterns. Transfer forces can generally be observed in podium structures and
structures with vertically offset walls and consist essentially of concentrated
force introductions into the diaphragms. Deformation compatibility forces can
be observed with dual LLRSs with wall and frames and to a lesser extent in
structures with bracing walls of different lengths or materials.

Deformation compatibility forces can normally be neglected in LTF structures,


due to the flexibility of timber members and connections (Cobeen et al., 2014).
This statement is valid as long as all of the structural elements comprising the
LLRS are of similar stiffness. This is typically achieved by using the same type of
bracing walls and by designing bespoke bracing elements such as steel portals to
undergo similar displacements under horizontal loading. If very stiff concrete or
mass timber cores are used to brace the structure in combination with traditional
lined bracing walls, a more sophisticated analysis outside of the scope of these
guidelines will be required to account for displacement incompatibilities.

Transfer forces within the diaphragms need to be accounted for in the diaphragm
analysis by concentrated forces or additional diaphragm supports, as discussed later.

4.2 Design of diaphragms in LTF structures


The design of any system can be split into two stages: determination of demand
and establishment of capacity. This is also true of diaphragms in LTF buildings,
and this section is divided into these two stages as follows.

Determination of diaphragm demand:


■■ Calculate wind demand using AS/NZS 1170.2:2011 as described in section 2.3.
■■ Calculate seismic demand using NZS 1170.5:2004 as described in section 2.4.
Use the pESA as per section 4.2.2 to determine the floor forces based on the
building overstrength.
■■ Find maximum floor force between wind demand and pESA seismic
demand.

Establishment of component demand:


■■ At each level, distribute the floor design force to each component of the
LLRS by:
• using envelope of maximum forces considering diaphragm as fully
flexible then fully rigid
• OR modelling the diaphragm and LLRS considering their component
stiffness (requires iteration).
■■ Find the load demand on diaphragm components by:
• applying a girder analogy
• OR shell element modelling (finite element analysis)
• OR equivalent truss method (when finite element analysis is unavailable).

Check component capacity:


■■ Check component capacity according to the appropriate material standard,
experimental data or manufacturers’ data.

59
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

4.2.1 Determination of wind diaphragm forces

The determination of wind loading is described in section 2.3. Once determined,


these are applied as line loads to the edges of the diaphragm.

Wind loading on façades

Often it is possible to move straight to the establishment of component


demand once the line load is determined. However, the nature of wind
forces should not be overlooked.

Wind loading is a combination of an external push on the windward side


and a suction force on the leeward side. Care needs to be taken that this
suction force has a clear load path back into the diaphragm. Although not
a frequent problem in sheathed diaphragms, mass timber diaphragms
spanning in a single direction can present a line of weakness. In this case,
ties can be required to adequately engage the diaphragm.

4.2.2 Determination of seismic diaphragm forces

Overstrength and dynamic effects in multi-storey LTF buildings can be


determined using the pseudo equivalent static analysis (pESA) method as
per clause C5.7.2 of NZS 1170.5 Supp 1:2004. pESA forces are determined by
multiplying the storey forces from the ESA by the overstrength of the ductile
LLRS (i.e. the global building overstrength based on the ductile metal connectors
providing the ductility and damping of the systems). A minimum acceleration
value, corresponding to the peak ground acceleration, is used at the lower
storeys as shown in Figure 17.

overstrength

pESA envelope
Height

ESA envelope

PGA Lateral force

Figure 17. Static forces for ESA and pESA envelopes, modified from Gardiner, Bull and Carr (2008).

60
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Overstrength values of the ductile elements of the LLRS (typically the metal
sheathing panel fasteners) are normally provided in relevant material standards,
manufacturers’ information or can be determined by testing. Due to the flexibility
of LTF walls, the spacing of the wall sheathing panel fasteners is often governed by
serviceability limit state (SLS) deflection, requiring more fasteners than required
from the ULS seismic demand in the walls to maintain sufficient stiffness. This leads
to an overdesign of the ductile link, which needs to be considered together with the
overstrength of the fasteners, defined as the overall building overstrength.

Since the ductile elements in the LLRS are the panel fasteners of each wall
up the building height, the average overstrength for each storey for each wall
needs to be considered. If it cannot be assumed that the walls at each storey
yield at the same time (the difference between actual demand and capacity
varies significantly for the different storeys), a weighted average based on the
shear demand should be considered. The overstrength factor for wall m can be
calculated as follows:

(4.1)

where:
ɸ om overstrength factor of wall m based on the average overstrength up
the wall height
ɸ o, fastener overstrength factor of the ductile fastener
ɸ strength reduction factor
vCPE, i unit shear capacity at overstrength of the wall at level i
vi* unit shear demand at level i from an unamplified ESA based on
NZS 1170.5:2004 or from a modal analysis
Qd,j,i design fastener strength at level i
k17 factor for multiple fastener effect from NZS AS 1720.1
si fastener spacing at level i

These guidelines consider that the majority of LLRSs will consist of LTF walls,
with nails as sheathing fasteners. NZS AS 1720.1 defines an overstrength factor of
1.6 for small diameter metal fasteners. The overall building overstrength factor
can then be calculated as the average from all walls as follows:

(4.2)

where:
ɸ ob overall building overstrength factor

As the connection of the diaphragm sheathing to the joists and chords has by its
nature the ability to provide stable non-linear response, the overstrength design
demand on the diaphragm elements is limited by the maximum overstrength
demand acting on the building as per NZS AS 1720.1. These values can be
determined by an elastic ESA, i.e. from an elastic response spectrum with a
structural performance factor Sp as per NZS 1170.5:2004 for the ductile building
under consideration. This limit does not apply however if the pESA forces are
derived from the PGA W forces.

61
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The demand on the diaphragm due to seismic loading FD* to be used to design
the ductile diaphragm components is defined as:

(4.3)

where:
FD* diaphragm demand for ductile components
FESA floor force from ESA (with µ and Sp from the ductile building)
FESA max overstrength
floor force from ESA at maximum overstrength (µ = 1, Sp as per
NZS 1170.5:2004 for the ductile building under consideration)
PGA peak ground acceleration
W total seismic weight of building

Accidental eccentricity does not need to be taken into account when using the
pESA approach.

Any non-ductile elements along the diaphragm load path (sheathing panels,
chords, non-ductile connections) should be designed with an additional 20%
demand. While this is not a strict application of capacity design principales, it is
acceptable as the sheathing connection is already being designed as a capacity
protected element.

(4.4)

where:
FD,CPE* diaphragm demand for capacity protected diaphragm components

4.2.3 Dependency of the design capacity and the load direction

It is common practice to consider seismic loads along the two principal


directions of the building. Also taking into account the directional effects of wind
and the reversed action of seismic loads, a number of load combinations need to
be considered when designing the diaphragm system.

Panels and panel fasteners must be designed for the maximum unit shear force
from all combinations. Depending on the loading direction considered, the
functions of chord and collector beams can be interchanged. Boundary beams
will need to be designed to resist both chord and collector actions because of
the arbitrary nature of the actions. Chord or collector action do not need to be
considered simultaneously, but separate analysis for both loadings in the two
principal building directions will be required. Any diaphragm loading needs
to be combined with the relevant gravity loading under the appropriate load
combination.

4.2.4 Distribution of forces within the diaphragm – flexible and rigid


diaphragms

The rigidity of the diaphragm will influence how forces are distributed between
the LLRSs. Examples of different diaphragm behaviours are shown in Figure 18. In
the case of a flexible diaphragm, the force distribution into the lateral load resisting
elements and within the diaphragm itself can be determined according to a tributary
area approach. For rigid diaphragms, on the other hand, the horizontal forces are
distributed as a function of the stiffness of the lateral load resisting elements.

62
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Diaphragm loading Flexible diaphragm


(Δdiap ≫ΔLLRS)

Rigid diaphragm Semi-rigid diaphragm


(Δdiap ≪ ΔLLRS) (Δdiap ≅ ΔLLRS)

Figure 18. Diaphragm behaviour.

However, in reality, an intermediate behaviour of the two extreme cases often


occurs. Timber diaphragms often work in this intermediate range, commonly
defined as semi-rigid, where the force distribution is a function of both the
diaphragm and the LLRS stiffness.

Although the diaphragm can be classified as either rigid or flexible – for example,
within Appendix A of NZS 1170.5:2004 – the classification criteria is not always
straightforward to calculate, providing little information regarding diaphragm
demand. These guidelines do not consider it necessary that such a classification
be made. This is because, in diaphragm design, a conservative assumption
can be made where the diaphragm is assumed as first fully rigid and then fully
flexible with component forces being taken as the maximum demand resulting
from the two assumptions.

Alternatively, a numerical model of the diaphragm representing its actual


stiffness and that of the LLRSs can be built and incorporated into the structural
model described in section 3.3.1.

4.3 Determination of diaphragm component demand


Numerous procedures are available that can be used to analyse timber
diaphragms and provide the designer with essentially the same result. The
appropriate selection of a procedure will depend on the geometry and
complexity of the structure and the analytical tools available.

Two procedures, summarised in Table 2, are described in these guidelines.


These procedures vary in their input requirements. In general terms the girder
analogy can be used for simple diaphragm geometries, with the numerically
based finite element diaphragm analysis being required for more complex
geometries.

63
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Table 2. Summary of diaphragm analysis methods.

Analysis method Floor geometry Note

Design values readily available.


Girder analogy Regular geometry Floor geometries can be
included using first principles.
Determination of the design
values require a comprehensive
Finite element analysis Any geometry
understanding of the design
software.

Interested readers can refer to other methods available in literature that allow
for the design of openings (Elliott, 1979; Jephcott & Dewdney, 1979; Diekmann,
1982; Dean, Moss & Stewart, 1984; Kessel and Schönhoff, 2001; Prion, 2003;
Tissell & Elliott, 2004; Ni & Popovski, 2015) or design for irregular diaphragms
in general (Malone & Rice, 2012). Most procedures can easily be implemented in
the presence of only one opening or irregularity but will lead to complex designs
when multiple irregularities need to be considered.

A third analysis method is the equivalent truss method (Moroder et al., 2015)
which models the timber diaphragm components by using diagonals on a
rectangular grid and by including fastener flexibility, similar to the strut and tie
approach for concrete structures. This method is not included within the scope
of these guidelines.

4.3.1 Girder analogy

For regular and rectangular diaphragms or regular rectangle areas within the
diaphragm, the load path and the internal actions in the diaphragm components
can be determined using the girder analogy as shown in Figure 19. Shear forces
are resisted by the diaphragm panels and can be assumed as constant along the
diaphragm depth. Bending moment is resisted by the chord beams acting in
tension and compression. The edges of the rectangular diaphragm area parallel
to the external loads are considered as linear supports and effectively need to be
able to transfer the horizontal forces into the vertical LLRS.

chord action T (tension) M=wL2/8

L
H
V
V=wL/2

unit shear

horizontal load w
supports (lateral load resisting system)

chord action C (compression)

Figure 19. Girder analogy for regular diaphragms.

64
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The unit shear force demand v*, defined as the shear force per unit length, is
used to determine fastener spacing and is calculated as:

(4.5)
where:
V reaction force at the diaphragm supports based on the diaphragm
demand F*D from either wind or seismic loading
H diaphragm depth

A simplifying assumption that diaphragm demand is based on the maximum


shear force demand can be made as follows:

(4.6)
where:
w uniformly distributed load from storey diaphragm demand. For wind
actions, w can be determined directly from the storey wind demand.
For seismic actions, w can be determined from FD* for ductile
elements (i.e. the ductile sheathing panel fasteners) or FD,CPE* for
brittle elements (i.e. sheathing panels) as per section 4.2.2.
L diaphragm span

For larger diaphragms, fastener spacing can be reduced towards the centre of
the diaphragm by calculating the maximum shear demand at the specific point
that the individual sheathing panel is connected. It should be noted, however,
that this can add complexity to construction and will require a higher level of
monitoring to ensure that errors in nail spacing are not made when matching the
specified nail spacing within the floor.

The maximum chord tension and compression forces can be determined as


follows:

(4.7)

where:
T* tension force in chord beam
C* compression force in chord beam
M moment from the uniformly distributed load
w uniformly distributed load from storey diaphragm demand. For wind
actions, w can be determined directly from the storey wind demand.
For seismic actions, w should be determined from FD,CPE* for brittle
elements as per section 4.2.2.

4.3.2 Extended girder analogy

For diaphragms with a limited number of irregularities, hand methods based on


first principles are still feasible to determine unit shear forces and axial forces
in framing elements. Due to the presence of inter-tenancy walls, irregular floor
diaphragms can normally be broken down into smaller regular diaphragms
allowing the use of the girder analogy.

65
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

In the presence of interrupted chords or re-entrant corners, chord extensions


are typically included as part of the floor framing. The forces in these extensions
need to be transferred into the adjacent parts of the diaphragm as shown in
Figure 20a. This is done by adding their respective shear force (axial force in the
chord at the start of the extension divided by the length of the extension) to the
panel unit shear force in that part of the diaphragm as shown in Figure 20b.

Alternatively, the discontinuous chord can also be ignored and the next
continuous internal beam can be considered as the chord beam as long as its
continuity is guaranteed (Figure 20c). The internal chord beam needs to carry
higher forces because of the smaller lever arm. Higher displacements and
therefore potential damage at the discontinuities need to be taken into account
with this approach.

4.3.3 Finite element (shell) modelling

Finite element analysis is arguably the most accurate way to analyse diaphragms,
since it allows for the continuous nature of the system. The load paths including
tension, compression and shear stresses can be accurately predicted as well as
the diaphragm deflections. However, the accuracy of the results is in proportion
to the model complexity, which again is in proportion to the knowledge and
time required in setting up the model and post-processing the results. Finite
element modelling does, however, require the pre-definition of element sizes
and fastener spacing, meaning some iteration is necessary.

Most engineering software allows for the definition of shell elements with
membrane action. Orthotropic shell properties are normally not required
for standard timber diaphragm analysis, since the shear stiffness is generally
equal in the two plan directions, and the axial stiffness in the longitudinal and
transverse direction does not have a relevant impact in the analysis.

The shear stiffness needs to be reduced to account for the sheathing panel
splices, which under most circumstances account for the majority of the
diaphragm flexibility.

A reduced shear stiffness can be calculated by (Moroder et al., 2015):

(4.8)
where:
Gef equivalent shear modulus of the sheathing
G sheathing shear stiffness
t sheathing panel thickness
Kser || slip modulus of the fastener parallel to the panel edge
s fastener spacing
b panel width
h panel length
ci number of fasteners along sheathing panel edge
c1 number of rows of fasteners between adjacent panels along the
sheathing panel length h
c2 number of rows of fasteners between adjacent panels along the
sheathing panel width b

66
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

area 1 area 2 area 3 chord

chord

chord interrupted chord

moment

area 3 area 3

H1 H2

chord

chord shear flow


continuous chord
V*
L1
Vmax
shear

V*

moment

M* M*max

V* = V*/H1 + T*/L1 shear flow T* = M*max/H2

T* = M*/ H1 chord flow

Figure 20. Chord discontinuities and redistribution of forces in diaphragms.

67
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Typically, c = 2 for plywood sheathing where the force is transferred with one
line of fasteners to the framing member and with a second line of fasteners
into the next panel. For mass timber panels, c = 1 for half-lap joints and c = 2 for
connections with splines.

In accordance with NZS AS 1720.1, from 0 to 0.5 mm nail deformation, the


average deformation δ can be determined using:

(4.9)
where:
k37 fastener deformation factor, generally 1 for diaphragm loading as it is
short term (under 5 minutes)
N specified nail load demand in N
nα,y short-term design yielding nail strength in N

Since the nail load demand changes along the diaphragm length, it is necessary
to derive a linear stiffness Kser || rather than using non-linear force-displacement
curves describing the nail slip. A simplifying assumption can be made by taking
Kser || to be the secant stiffness at 0.5 mm nail displacement as follows:

(4.10)

Depending on the software used, it may be simpler to calculate an in-plane


stiffness modifier defined as the ratio between the equivalent shear stiffness Gef
and the material shear stiffness G.

With this approach, individual sheathing panel splices do not need to be modelled.
However, it is important that axial stress and forces in both principal directions
be checked to quantify the force transfer from diaphragm irregularities. While
axial forces do not occur in regular diaphragms, they are often encountered in the
presence of irregularities. While not usually governing the design of the timber
elements themselves, these forces become a critical weakness in fastener design.

Drag, collector and chord beams can be modelled directly, and their forces can
be obtained directly from the beam elements.

Most software packages provide the option to calculate resultant forces over a
defined shell section (by numerical integration of all stresses along the section),
which can be used to design reinforcing elements at openings, discontinuous
chords and re-entrant corners.

4.3.4 Special consideration for wind and large diaphragm demand

In the case of large wind loads and larger diaphragm spans, horizontal loads
on façade elements can become substantial and care needs to be taken that the
horizontal forces are effectively transferred through the diaphragm into the LLRS.

In seismic design, the majority of the forces are generated by the inertia of the floor
and contents, which can be considered as smeared over the whole diaphragm. For
this reason, the considerations below often apply only to wind governed designs.

68
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Loads applied to a diaphragm edge, a combination of wind suction and internal


pressure, for example, are not necessarily activating the whole diaphragm, and
load paths need to be established within the floor framing to transfer the force
components perpendicular to the panel edges, which have the potential to pull
the diaphragm elements apart.

Establishing load paths from the diaphragm edges into the diaphragm body can
be done in one of the following ways:
■■ Force transfer via chord beam (LTF diaphragms only) (see Figure 21a):
The chord beam resists the forces in bending (in the diaphragm in-plane
direction) and transfers the forces into the framing elements parallel to the
load direction. Since this force can act in both tension and compression, a
tension connection between the chord beam and the framing members is
also required. The framing elements continuously transfer the axial load into
the sheathing panels. Splices in the joists or blocking elements are designed
considering the demand from both the tension and compression forces.
■■ Force transfer via sheathing panels (LTF and mass timber diaphragms)
(see Figure 21b):
The force is introduced from the chord beam directly into the sheathing panels,
requiring the fasteners to resist forces perpendicular to the panel edges. This
mechanism also creates normal panel stresses. Since the fasteners already
need to resist the unit shear force (blue arrows) from the diaphragm action, it
is necessary to check if they can also resist the additional load component (red
arrows) from the direct load introduction. The additional force component
normally also requires an increased edge distance of the fasteners.

F F

chord beam
bending of beam
force introduction into the
tension/compression
panels over the fasteners
connection
continuous force
tension/compression introduction into the
in framing elements framing elements
(continuous force
unit shear force
introduction into the
sheathing panels force transfer
framing element
fastener sheathing panel
a) Force transfer via chord beam b) Force transfer via panels

Figure 21. Force introduction via the diaphragm edges: a) via the bending of the chord beam; b) via fasteners into the sheathing panel,
modified from Moroder et al. (2015).

69
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

A simple approach to account for the resultant force, which accounts for
both the unit shear force and the additional transfer force component, is
to artificially reduce the diaphragm depth and therefore increase the unit
shear force in the panel fasteners. This approach is used in the national
appendix to Eurocode 5 (DIN EN 1995-1-1/NA:2013), which requires that
the effective diaphragm depth be taken as a quarter of the diaphragm span
for force introduction along one edge or half the diaphragm span for force
introduction along both edges (which would generally be the case with wind
loading).

4.4 Component capacity and connection design in timber


diaphragms

The strengths of all diaphragm elements including sheathing panels, fasteners,


frame elements and chord/collector/drag beams need to be established, together
with the connection of the collector beam to the LLRS.

4.4.1 Panel connections

In order to define the strength of panel connections, the design capacity of the
single fastener is divided by its spacing and compared against the unit shear
force (shear flow) demand:

(4.11)

where:
v* unit shear force demand
s fastener spacing
Nd,j strength of fastener in shear

k modification factors for strength according to NZS AS 1720.1


Qk characteristic fastener capacity

For load components orthogonal to the panel edge, as described in section 4.3.4,
force orientation must be considered, including allowance for the minimum
distance to the loaded edge. Minimum fastener edge distances can be found in
chapter 4 of NZS AS 1720.1.

The sheathing panels are generally supported by joists or beams whose primary
function is to support gravity loads. The edges of adjacent sheathing panels are
connected through these framing members, transferring the unit shear force to
the adjacent panel.

Floor layout and construction sequencing will often govern the choice of
connections, and a selection of examples is shown in Figure 22. Gluing of
diaphragm panels to transfer shear is not recommended as it tends to result in
brittle failure modes and requires an increased level of construction monitoring
and quality control to ensure adequate bonding.

70
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

a) Nailing/screwing of the panel to the next joist

b) Inclined fully threaded screws or regular screws at 90° between joists

c) Blocking

Figure 22. Diaphragm connection details for sheathing panels.

Mass timber diaphragm design

Mass timber panels made of CLT, LVL or glulam are often profiled or
machine cut and allow for specific connection detailing. These panels
normally carry both gravity and diaphragm loads and do not require
joists or beams other than at their end supports. Butt joints are generally
avoided to prevent relative movement between the panels and to prevent
gaps opening between the panels. The most common connection details
for mass timber panels are shown in Figure 23.

71
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

a) Half-lapped or shiplap joint with b) Single surface spline (wooden strip


screws or nails in recess between panels) with screws
or nails

c) Internal spline with inclined fully d) Connection with steel plates


threaded screws (single or double
inclination)

Figure 23. Diaphragm connection details for mass timber panels.

4.4.2 Sheathing panels

Sheathing panels need to resist the unit shear force in the diaphragm including
any overstrength demand where required. The capacity of the sheathing is a
function of the panel thickness and the in-plane shear strength. In cases with
very thin sheathing panels or where large framing spacing is used, panel shear
buckling should also be considered.

Based on NZS AS 1720.1, once a sheathing material and thickness has been
selected, the following verification needs to be carried out:

(4.12)
where:
v* unit shear force demand
ɸ capacity reduction factor
k modification factors from NZS AS 1720.1
g19 modification factor for plywood assembly
fs’ characteristic in-plane shear stress of the panel
t full thickness of plywood

The strengths of joist and blocking elements normally do not need to be


checked, since their axial load is generally relatively small. In cases where large
concentrated forces are introduced into the diaphragm by framing elements (for
example, large wind loads from the façade), compression or tension strengths
checks might be required. Buckling can normally be ignored since lateral
displacements are prevented by the sheathing panels.

72
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Mass timber diaphragm design

The same design checks as LTF diaphragms need to be carried out for mass
timber diaphragms with some additional considerations.

Due to the missing framing elements, longitudinal forces are carried


directly by the mass timber panels and their connections. Axial buckling
should be checked for very large compression forces.

In a mass timber diaphragm system, connectors are normally loaded both


parallel and perpendicular to the panel edges, and increased minimum
distances to the loaded edge are often required compared to a shear
parallel to the edge only connection.

Brittle failures in tension perpendicular to the grain shall be avoided in


design, with checks available in Appendix ZZ of NZS AS 1720.1.

4.4.3 Chord, strut/drag and collector beams

Tension and compression forces in chord, strut/drag and collector beams need
to be checked in accordance with NZS AS 1720.1. Under seismic loading, the
chords, struts and collector beams need to be designed based on FD,CPE* as per
section 4.2.2

Similar to the frame elements, compression chords are normally not


subjected to buckling, since the diaphragm is restraining any lateral
displacements. For chord splices, a conservative design should be adopted,
since a loss in chord capacity could compromise global building behaviour.
This is automatically achieved by considering chord splices as capacity
protected elements.

Collector and strut beams collect all shear forces and transfer them either
to the LLRS or redistribute them into other parts of the diaphragm. As
such, they carry tension and compression forces and need to be designed
accordingly.

4.4.3.1 Tension members

Members required to resist tension forces must satisfy the following:

(4.13)

where:
T* tension force demand in chord of the diaphragm
ɸ capacity reduction factor
k modification factors for strength (from NZS AS 1720.1)
f t’ characteristic stress in tension parallel to grain
A t net cross-sectional area of tension member

73
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

4.4.3.2 Compression members

Members required to resist compression forces must satisfy the following:

(4.14)
where:
C* compression force demand in chord
ɸ capacity reduction factor
k modification factors for strength (from NZS AS 1720.1)
f c’ characteristic stress in compression parallel to grain
A c cross-sectional area of chord

4.4.3.3 Splicing a chord member

The biggest challenge in the design of chord, drag/strut or collector beams is the
potential intersection with other members. The transfer of axial forces correctly
in tension and compression represents a critical load path within the structure,
and connections need to be designed accordingly.

Figure 24a shows a possible connection for axial forces between two
intersecting collector/strut beams. Tension forces are transferred by the bolts
and the steel angles, and compression forces are transferred by the bearing
of the collector/strut beam on the continuous beam. For smaller axial forces,
a nailed steel strap or plate may be enough to transfer tension forces as
shown in Figure 24b.

continuous beam

collector/strut beam

bracket with bolt


(a)

steel plate continuous beam

collector/strut beam

(b)

Figure 24. Force transfer in collector or strut beams.

74
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Mass timber diaphragm design

In the case of mass timber panels with small collector forces, dedicated
strut beams can sometimes be avoided by transferring tension forces
through nailed steel strips to the adjacent panel(s) as in Figure 25. For such
cases, possible tension forces perpendicular to grain at the panel edge
need to be considered as discussed in section 4.4.3.

mass timber panel

steel reinforcement

floor opening

Figure 25. Force transfer in mass timber panels.

Most collector and drag/strut beams also carry gravity loads. This means that their
section size and connection design needs to be carried out based on appropriate
combinations of tension and compression and bending/shear.

4.4.4 Diaphragm deflection

In the case of regular diaphragms spanning between two supports, the girder
analogy can be used, and the mid-span deflection Δ can be determined as per
NZS AS 1720.1

The deflection of regular and irregular timber diaphragms can also be


determined directly with a shell model as outlined in section 4.3.3.

75
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

CONTROL OF FLOOR VIBRATIONS


5
IN MULTI-STOREY LTF
BUILDINGS

5.1 Introduction

Floor vibrations are a very subjective phenomenon and are difficult to describe
and quantify in engineering terms. Different users have different levels of
perception of floor bounciness and also have different levels of tolerance.
Vibrations are an SLS governed by small displacements, and numerous factors
including the presence of non-structural elements, support conditions and
damping can influence the performance significantly.

The current vibration control criteria of 1–2 mm deflection under a point load of
1 kN as per AS/NZS 1170.0:2002 Appendix C (Table C1) is a coarse approximation
and does not necessarily provide adequate performance to users as a number of
complaints have recently shown (CEAS, 2017). This tends to be more prevalent
where beam supporting beam scenarios exist and where cantilever beams are
supported by other beams. While current floor span tables in NZS 3604:2011
have been developed to provide adequate performance for structures within
the scope of that standard, in some cases, including multi-storey construction,
there may be a need for higher-performing floor systems. For this reason, a more
stringent criterion is presented in this section.

As discussed in the next section, there are a number of parameters that


can describe vibration performance. A combination of some of the criteria,
calibrated against numerical analysis and compared to as-built floors, has led
to a design procedure (Ni & Popovski, 2015) and is presented in an adapted

76
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

form in this section. For situations where the design approach presented does
not lead to practical design or more stringent criteria need to be applied, more
sophisticated approaches based on finite element and modal analyses are
available in literature (Smith, Hicks & Devine, 2009).

5.2 Floor vibration and critical design parameters


A person walking on a floor creates impact forces that lead to dynamic effects.
A part of this impact is perceived as noise, but mechanical vibrations also
occur. Research has shown that a person walking generates harmonic waves in
multiples of 2 Hz. Normally, only the first three to four harmonics are significant,
meaning that floors with frequencies lower than 8 Hz can be brought into
resonance by a person walking on it. The magnitude of resonance depends of
the damping of the floor system and the actual frequency (Ni & Popovski, 2015).

For floors with a frequency above 8 Hz, the induced vibrations are normally
dominated by a transient response caused by an individual step or heel drop. This
transient vibration decays quickly and is a function of the floor stiffness and mass.

When designing floors for buildings with multiple tenancies, a more detailed
analysis that considers floor stiffness, mass and damping should be adopted.
Normally, these procedures involve frequency, deflection and acceleration checks.
Adding mass normally reduces transient vibration and impact noise, but the
added mass reduces the frequency, possibly creating resonance problems. The
use of concrete also increases the seismic mass. More information can be found
on shared tenancy floor systems requiring higher vibration limitations in Hamm
et al. (2010), but it can be observed that this work includes some rather stringent
criteria around floor layouts that are not often required for New Zealand buildings.

5.3 Design method


5.3.1 Calculation of vibration-controlled span

The following design method was developed in Canada (CCMC, 1997; Ni &
Popovski, 2015) and adapted to New Zealand design practice. The method is
applicable to timber-framed floors that have frequencies of more than 9 Hz and a
maximum span of 14 m.

The procedure is presented for timber joists (sawn or engineered timber) and a
flooring made of plywood sheathing panels or similar timber-based panels. Joists
can also be made of wood and steel web trusses, I-beams or steel elements, but
in these cases, the equations provided below need to be adjusted accordingly.
This would be done using appropriate modification factors (i.e. g41) and EIeff and
EI’ as described below.

Although concrete and screed toppings can improve the vibration performance
due to their two-way stiffness and hence reduce floor deformations, the
increased mass tends to reduce the maximum allowable length of the joists.
Furthermore, the additional mass will impact negatively on the seismic
performance of multi-storey buildings by adding seismic weight. If a concrete
topping needs to be accounted for in the proposed procedure, the concrete
bending and axial stiffness can be summed with the individual contributions
from the plywood. In addition, when a concrete topping is present, the full span
length can be used for the flooring when determining the effective bending
stiffness of the joist-floor system as discussed later.

77
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

When assessing the vibration performance of floors, the continuity of the joists
should not be accounted for, meaning that each span needs to be considered as a
simply supported span.

Based on the fundamental design parameters to assess vibration performance


(i.e. natural frequency and static deflection) and in conjunction with empirical
studies, the following equation can be used to determine the maximum joist
span where vibration performance can normally be considered as acceptable:

(5.1)
where:
lv vibration-controlled joist span in m
EIeff effective composite bending stiffness in Nm2 of the joist-flooring
system as per Equation (5.3)
mL mass per linear metre of joist length in kg/m
g41 modification factor for concentrated loads on a grid system as per
NZS AS 1720.1

The effective (composite) bending stiffness takes the interaction between the
joists and the flooring above under consideration. The presence of plywood
flooring normally increases the stiffness of joists, and this effect can be taken
into account with the gamma method as explained below.

The mass per linear metre mL normally consists of the self-weight of the joist
and the flooring based on a strip with a width s equal to the joist spacing. In
addition, an allowance for services, finished floor and suspended ceiling should
be considered. By assuming an additional 20 kg/m2 of superimposed dead load, a
surface load of about 40 kg/m2 is obtained, which aligns with the load assumptions
for residential houses as per Shelton (2013). Additional loading may need to be
added for inter-tenancy floors having additional acoustic requirements.

The static deflection check considers a concentrated load applied to the mid-span
of a single joist. The presence of the flooring spanning perpendicular to the joists
allows for some load redistribution to the adjacent joists. This effect is considered
in the design standard NZS AS 1720.1 by reducing the applied load by the factor g41:

(5.2)

where:
EIjoist bending stiffness of the joist
EIf,90 bending stiffness of the flooring perpendicular to the joist span
calculated over the whole joist span
l joist span
i spacing of joists

The presence of the flooring fastened to the joist increases the bending stiffness
of the joists. Unless the joist and the flooring are rigidly connected (i.e. stressed
skin floors), the flexibility between the elements needs to be taken into account.
The gamma method for mechanically jointed beams as described in Eurocode

78
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

5 can be used to calculate the effective composite bending of the joist flooring
system as follows:

(5.3)

where:
EI’joist apparent bending stiffness of the joist including shear stiffness
EIfloor bending stiffness of the flooring parallel to the joist span
EAfloor axial stiffness of the flooring parallel to the joist span
EAjoist axial stiffness of the joist
γ coefficient to account for the composite action of mechanically jointed
beams
afloor distance between the centroid of the flooring and the centroid of the
mechanically jointed floor system
ajoist distance between the centroid of the joist and the centroid of the
mechanically jointed floor system

In order to account for both the bending and shear deformation of the joists,
the Timoshenko beam equation as below is recommended (Timoshenko & Gere,
2006). If the shear deformation of the joist is small, the joist bending stiffness
EIjoist can be used instead of the apparent bending stiffness EI’joist. The shear
component is normally more significant for deeper cross-sections:

(5.4)
where:
G shear stiffness of the joist (can normally be taken as E/15 for sawn and
glue-laminated timber and E/20 for LVL)
A cross-sectional area of the joist
fs shear form factor (1.2 for rectangular sections, ratio of total section
area to web area for I-sections)

Most codes provide an apparent E, such as NZS 3603:1993 and AS 1720.1 - 2010. If an
apparent E is used, then the shear stiffness term can be ignored from Equation (5.4).

When using I-joists, the apparent bending stiffness can be calculated based on
the shear factor Sf. Equation (5.4) is rewritten as:

(5.5)

where:
S f shear factor as provided by the manufacturer or from literature

The γ-coefficient accounting for the flexibility between the joist and the flooring is
calculated as:

(5.6)

79
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

where:
s spacing of flooring fasteners
K slip modulus of the flooring fasteners
EAfloor axial stiffness of the flooring parallel to the joist span
lf distance parallel to the joist direction, between plywood splices

Floors including wood panels like plywood or other timber-based panels are
limited in size, and the presence of panel joints limits the continuity of the
mechanically jointed beam. This is accounted for by using the distance lf
between the panel joints. In the presence of a concrete topping, the gap between
panels can be neglected and the entire joist span can be used to determine the
effective bending stiffness.

The distances between the centroid of the flooring and the joist from the
centroid of the mechanically jointed floor system are:

(5.7)

(5.8)

where:
afloor distance between the centroid of the flooring and the centroid of the
mechanically jointed floor system
ajoist distance between the centroid of the joist and the centroid of the
mechanically jointed floor system
h height of the joist
t thickness of the flooring

In case of mass timber floors where thicker wooden panels are spanning
between supports, the equation to determine the allowable vibration-
controlled span becomes:

(5.9)
where:
lv vibration-controlled joist span for mass timber floors in m
EIeff effective bending stiffness for a 1 m wide panel in Nm2
In the case of a glulam beams on flat (parallel laminated timber),
the effective bending stiffness equals the bending stiffness of the
floor for a 1 m wide panel. For cross laminated timber, the
effective bending stiffness needs to take the presence of the cross
layers into account. The effective bending stiffness for CLT is
normally provided by the supplier or can be calculated (i.e.
gamma method or shear analogy).
m mass per linear metre of a 1 m wide strip of floor in kg/m (self-
weight of the panel and any superimposed dead load)

80
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

5.3.2 Vibration-controlled joist span tables

The procedure presented above has been applied to the joist span Table 7.1 from
NZS 3604:2011 with results presented in Table 3. The following assumptions have
been made when determining the vibration-controlled span of the joists:
■■ Average modulus of elasticity of SG8 timber as E = 8,000 MPa considering
the low probability of a person walking on the same joist all the time and the
redistribution of loads to adjacent joists through the flooring.
■■ Fastener spacing at 300 mm as per the connection to intermediate supports
in standard plywood flooring.
■■ Fastener diameter of 2.8 mm.
■■ Fastener stiffness of 1,100 N/mm as determined from the force-displacement
curve provided in NZS 3603:1993 as the tangent stiffness at a deformation of
0.5 mm.
■■ Plywood thickness of 15 mm for joist spacing of 400 mm and 450 mm and of
19 mm for joist spacing of 600 mm with grade F11 plywood.
■■ Additional superimposed dead load of 20 kg/m2 for a total dead load
including the self-weight of the structural elements of about 40 kg/m2 (inter-
tenancy floors may have higher superimposed dead loads).

The increase of 10% used for continuous joists does not apply when determining
the vibration-controlled spans.

Table 4 shows the vibration-controlled spans for floor joists made of LVL13.

Table 3. Floor joists SG8 in dry conditions considering vibration control.

Maximum vibration-controlled span of joists at maximum spacing of


Floor joist size
400 mm 450 mm 600 mm

90 x 45 mm 1.45 m 1.40 m 1.25 m

140 x 45 mm 2.20 m 2.15 m 2.00 m

190 x 45 mm 2.80 m 2.75 m 2.60 m

240 x 45 mm 3.40 m 3.30 m 3.10 m

290 x 45 mm 3.95 m 3.85 m 3.65 m

Table 4. Floor joists LVL13 in dry conditions considering vibration control.

Maximum vibration-controlled span of joists at maximum spacing of


Floor joist size
400 mm 450 mm 600 mm

150 x 45 mm 2.65 m 2.55 m 2.40 m

170 x 45 mm 2.90 m 2.85 m 2.65 m

200 x 45 mm 3.35 m 3.25 m 3.05 m

240 x 45 mm 3.85 m 3.75 m 3.55 m

300 x 45 mm 4.60 m 4.45 m 4.25 m

360 x 45 mm 5.30 m 5.15 m 4.90 m

400 x 45 mm 5.75 m 5.60 m 5.30 m

81
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The tables above are not intended as replacements for tables found in
NZS3604:2011, but have been developed as examples of spans where more
stringent vibration performance is desired.

5.3.3 Influence of additional perpendicular floor elements

The presence of additional structural elements perpendicular to the joist span


direction like blocking or strutting can reduce the perceived floor deflections
because of increased load redistribution to adjacent joists. This can be accounted
for by taking these elements into account when calculating g41 in Equation (5.2).

Alternatively, the following values can be used to increase the vibration-


controlled floor span if perpendicular elements are placed not more than 2 m
centre to centre or from the joist supports (Ni & Popovski, 2015):
■■ 8% increase with inclusion of timber blocking, 35 mm wide and the same
depth of the joist and fixed with three 3.15 mm nails per end.
■■ 8% increase with inclusion of herringbone strutting, 35 x 35 mm and fixed
with at least two 2.8 mm nails per end.
■■ 5% increase in the presence of continuous strapping/timber elements
of at least 20 x 90 mm as commonly used for ceiling battens fixed to the
underside of the joists with two 2.8 mm nails.

5.3.4 Non-rigid supports

The procedure as presented above assumes a simply supported floor system on


rigid supports. In certain conditions, the floor joists are supported by beams or
lintels, which cause additional flexibility and hence lead to higher deflections.
If the floor supports are not rigid enough, this needs to be taken into account
when determining the vibration performance of the floor.

It is recommended that the static deflection check under a 1 kN point load of


the beam supporting the floor is verified for a more stringent deflection of 0.25
mm or a maximum limit of 0.5 mm if a lower level of vibration performance is
acceptable. In addition, the frequency check of 8 Hz needs to be verified on the
global beam-floor system. This frequency can be determined with computer
software and a two-dimensional model or alternatively can be approximated
using the modified Dunkerley equation (Hamm, 2008):

(5.10)

where:
fsystem natural frequency of the floor-beam system
ffloor natural frequency of the floor based on the effective composite
bending stiffness EIeff
fbeam natural frequency of the supporting beam/lintel

The frequency of the beam and joist/floor can be calculated with the following
equation:

82
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

(5.11)

where:
f estimated natural frequency of a linear element in Hz
EI bending stiffness of the element under consideration in Nm2 (effective
bending stiffness of the joist-flooring system or bending stiffness of
beam element)
m mass per linear metre applied to the member under consideration
in kg/m
l span in m

Alternatively, the units for the equation above can be span l in m, bending
stiffness EI in Nmm2 and mass m in kN/m.

5.4 Summary
While there are existing criteria within New Zealand building standards, it
has been proposed that these criteria may not provide adequate mitigation of
annoying floor vibrations in some instances. Based on this, recommendations
from Europe and Canada have been incorporated in this section to provide
additional options for designers to consider when evaluating the performance
of floors used for multi-storey LTF buildings. These criteria are intended for
higher-performing floor systems, which may be dictated either by the needs
of a client or more stringent requirements for the specific building being
designed. It is important to note that other considerations such as the addition
of layers for acoustic separation may have additional influence on floor vibration
performance and should be accounted for as effectively as possible during
design. Continued research is needed around floor vibrations for New Zealand
buildings that can be part of a holistic approach to potential combinations
of materials and structural timber systems for vibration, acoustic, fire and
structural performance optimisation.

83
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

CLADDING, WEATHERTIGHTNESS
6
AND DURABILITY FOR LTF
BUILDINGS

6.1 Introduction

Cladding for multi-storey LTF buildings can include a broad variety of materials
and systems that can effectively enclose the building and create an exterior
envelope. Most cladding systems will minimise moisture infiltration into the
building cavity and keep the timber structure from getting excessively wet.
Should failure of the cladding system occur, the timber structural components
can potentially be exposed to moisture during the life of a building, and as such,
one method of ensuring the minimisation of potential deterioration in strength
and stiffness is to have the timber protected from decay through treatment. The
NZBC provides a path to compliance by providing an indication of what levels
of treatment are recommended for different parts of a timber building in order
to protect the timber from decay and deterioration throughout the life of the
building, based on the perceived risk of high moisture content occurring. A multi-
storey timber building will have timber elements used in a variety of applications
including the structure, cladding and other non-structural elements, and in order
to comply with the Acceptable Solution, each will require different protection
depending on the function within the building and accessibility.

Treatment types and hazard levels required for compliance with the Acceptable
Solution for timber used in New Zealand buildings are described in NZS
3640:2003 Chemical preservation of round and sawn timber, and locations for
different timber applications are provided in NZS 3602:2003 Timber and wood-

84
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

based products for use in building. Together, these standards will inform building
designers about what they should be specifying for the different elements
throughout a building to fulfil Acceptable Solution requirements.

Compliance can also be shown through the use of durable timber products
within building design or clear establishment of hazard level through
building physics analysis to establish that they can perform for the necessary
timeframes specified by the NZBC, with clause B2 Durability providing
guidance on whether a life of 5, 15 or 50 years is necessary for different
building components. Structural systems and all elements providing support
for the structure of the building must have a minimum 50-year life, and this
would include all timber structural components comprising LTF buildings
such as the framing, bracing systems (framing and sheathing products) and
floor diaphragms. Cladding elements typically are required to have a minimum
15-year life due to their visibility and the ability to evaluate or observe
deterioration within these systems.

6.2 Cladding and weathertightness


Wall and roof claddings are the primary protection for buildings against external
moisture penetration and therefore must be designed to create a weathertight
envelope that will protect the building interior and structural systems
supporting the building. The NZBC provides the requirements that must be met
by all structures in clause E2 External moisture. Buildings must be constructed
to provide resistance to penetration and accumulation of external moisture
to protect individuals from illness or injury. This is particularly important for
LTF buildings where the primary structural systems are timber, which can be
susceptible to degradation in high moisture environments. These structural
systems are often enclosed by wall, ceiling and exterior linings and therefore are
not typically visible in the event that moisture was accumulating and potentially
creating a hazard. Windows, doors and other penetrations to the exterior
building envelope must also be detailed and installed correctly to avoid external
moisture penetration.

Choosing an appropriate cladding system should include an understanding


of how it will interact with the structural system while also protecting that
structural system from the direct effects of the environment. It is also critical to
know that the selected cladding will have adequate attachment to the building
to be able to withstand required loads due to wind. Heavier cladding systems
may also need to be checked to ensure they will able to withstand an earthquake
without failing and becoming detached from the building. Often a cladding
system is chosen based on architectural considerations and must be checked
to verify that it will be suitable for the geometry, height and location of the
building. These systems can include timber panels or weatherboards, aluminium
panels or monolithic autoclaved aerated concrete panel systems.

For buildings that have wind pressures that fall within or below the extra high
category according to NZS 3604:2011, it is feasible to utilise proprietary cladding
systems that have been verified for buildings designed using the standard.
Some proprietary systems have also been approved for use in higher-pressure
scenarios, and information on the allowable specific engineering design (SED)
wind pressures is available in manufacturers’ literature. As noted previously, it
is important to consider all local pressure coefficients when determining wind
pressures on buildings as these relate directly to cladding loads and must be
checked for any building outside the scope of NZS 3604:2011.

85
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The majority of cladding systems for use with larger multi-storey buildings
will require the inclusion of a cavity between the cladding and the air barrier
that is created using battens or other proprietary elements. The reason for
this inclusion is so that water making its way through the cladding system will
be able to run down the cavity due to gravity instead of becoming held within
the interior building space and possibly initiating decay and deterioration of
the building. Most mid-rise LTF buildings designed using these guidelines will
include plywood shear walls for the exterior of the building and will therefore
have a rigid air barrier that can be sealed, and battens can be installed on top of
the plywood for the cavity system.

There are currently some gaps in the NZBC when it comes to acceptable
methods for providing weathertight solutions for buildings that exceed 10 m
and are beyond the scope of NZS 3604:2011. Acceptable Solution E2/AS1 and
Verification Method E2/VM1 do provide design, detail and testing options
for cladding and penetration systems but are limited as noted above. It is
recommended that, when specifying cladding systems that are unique or
unfamiliar to the building designer, it is best to consult with a façade engineer to
ensure that the cladding will provide adequate protection against the elements
and fulfil the necessary NZBC requirements. Proprietary cladding systems will
include technical literature such as details on how the system can be integrated
within a multi-storey timber structure to provide a suitable means of protecting
the building from environmental damage while providing a durable and
attractive exterior.

Research at BRANZ has been conducted that aims to provide test methods for
the verification of cladding systems greater than 10 m and up to 25 m (Overton,
2018). The resulting evaluation method (BRANZ, 2018) provides a means of
assessing cladding systems to ensure that adequate weathertightness is achieved
for tested systems against wind-driven rain and other environmental loads that
could result in moisture infiltration for taller buildings. The expectation is that
these systems include a drained and ventilated cavity and the testing provides
a set of consistent parameters to be used for the design of mid-rise cladding
systems.

In other parts of the world, the issues around cladding and weathertightness
of multi-storey LTF buildings have also been considered and can provide some
direction around what can be done for New Zealand buildings. Sources from
Canada (Ni & Popovski, 2015; HPO, 2011) include recommendations on how LTF
buildings can be designed to provide protection from exterior moisture but also
discuss issues around effective insulation and how to achieve better-performing
and more energy-efficient structures.

6.3 Durability
Timber that is kept dry can last for centuries as shown in examples of timber
buildings from around the world. Regardless of this, the Acceptable Solution
contained within the NZBC operates on the assumption that water will find
its way into buildings either through condensation, infiltration or plumbing
failures, and therefore as part of this Acceptable Solution, all timber is assumed
to be vulnerable to moisture levels that could potentially result in degradation
of untreated timber. NZS 3602:2003 provides specifications for the treatment
of timber within buildings including the range of applicable hazard classes for
different areas of the building. While other parts of the world design buildings
using philosophies that allow for the use of untreated timber, this is not in line

86
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

with the base assumption of the Acceptable Solution where all timber structural
elements require some level of protection should moisture content become
elevated. This can be provided either through naturally durable timber species
or more often through chemical treatments as described in NZS 3640:2003.

Should moisture contents within the timber remain low, it is unlikely that
a loss of stiffness or strength will occur. Obtaining durability of otherwise
vulnerable timber material through careful design and an understanding of
moisture movement within a structure is possible. Research in the UK ( Jones &
Brischke, 2018) describes some of the innovations and applications for timber
products in recent years and how the durability of timber materials can be
evaluated for use in European and UK building applications. Global innovations
in timber products and design have significant merit and should be considered
for multi-storey timber buildings in New Zealand, but it should be noted that
environmental conditions can differ from other parts of the world and product
durability can vary from what has been verified elsewhere. It is therefore
necessary to fully validate the performance of any timber products that will be
included in buildings in New Zealand across the necessary lifespan as required
by the NZBC depending on the use and accessibility within the building.

6.4 Summary
A wide range of proprietary roof and wall cladding systems are available that
can provide adequate protection to multi-storey LTF buildings in New Zealand
using a variety of materials and architectural features. These systems should
be verified as being fit for purpose for New Zealand conditions, and this can be
achieved through BRANZ Appraisals or CodeMark programmes. While timber
can easily last for the life of a building, it is necessary to understand how it is
being used within a building and what levels of treatment are required to comply
with NZBC requirements. NZS 3602:2003 and NZS 3640:2003 provide guidance
and specifications and make up an Acceptable Solution on how to evaluate
durability hazards in timber buildings and how to appropriately select treated
timber components that will stand the test of time and be Code compliant.
Compliance can also be shown through the use of naturally durable timber
products within building design or clear establishment of hazard level through
building physics analysis to establish that they can perform for the necessary
timeframes specified by the NZBC.

87
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

FIRE RESISTANCE IN
7
MULTI-STOREY LTF BUILDINGS

7.1 Introduction

A major concern for multi-storey LTF buildings is fire resistance, similar to any
other type of non-domestic multi-storey building. While buildings comprised
of large timber members such as glulam are known to perform adequately
in fire (Buchanan & Abu, 2017), light timber structures are often perceived as
having inadequate fire resistance. Because timber is a combustible material, it is
assumed that it will ignite during building fires, add to the fire load present in a
building and possibly result in catastrophic failure of the structure. Research has
shown that structural framing timber in LTF buildings does not add significantly
to the fire load and that the risk of damage to framing can be managed using
typical cladding solutions already in place. Barber and Gerard (2014) provided
an overview of the issues that require consideration regarding the fire resistance
of mid-rise and high-rise timber buildings. These included items such as the
perceptions and realities of fire risk, exposed timber elements, structural
connections, service penetrations and means of achieving approval through
regulatory agencies when designing and building larger timber buildings.

Once constructed, LTF buildings are clad with wall, floor and ceiling linings
on the inside surfaces and exterior cladding systems on the outside, including
the roof. This results in the timber framing members making up the structural
system being covered and therefore not exposed to any direct flame or heat
sources during the early stages of a fire event. As described by Buchanan (2007),
if designed appropriately using approved linings and fastenings, LTF can provide

88
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

excellent resistance to fire. These fire-resistant LTF systems need to meet all
applicable building codes and are most often verified by full-scale fire resistance
tests (Beattie, 2001).

Single-household detached dwellings designed according to NZS 3604:2011 and


Acceptable Solution C/AS1 will usually have no fire resistance requirements
except if part of the building lies within 1 m of a property boundary. In this
case, there is a requirement that the wall facing the boundary has a specified
(30-minute) fire resistance rating. When multiple household units exist, it will
be necessary to provide separating walls and floors in multi-storey buildings
that have specified fire resistance ratings. Similar systems to those used for
NZS 3604:2011 buildings can be employed, but they must be checked for NZBC
compliance. It is recommended that fire engineering practitioners be included as
part of the design team for a multi-storey LTF building as early in the design as
possible.

7.2 NZBC requirements


The NZBC includes fire resistance requirements for buildings such that adequate
protection from fire is provided for occupants of the building from injury and for
adjacent buildings from damage and to allow for suitable firefighting and rescue
operations to occur. These requirements are provided in clauses C1–C6 of the
NZBC, which include two Verification Methods and seven Acceptable Solutions.
The requirements of clauses C1–C6 also cross over other NZBC clauses including
structural and access specifications. The options of Alternative Solutions are
also available for designs falling outside the existing Verification Methods
and Acceptable Solutions, and these need to include consultation with an
experienced fire engineering practitioner.

7.3 Methods for achieving fire resistance ratings in multi-


storey LTF buildings
Several information sources are available on how LTF buildings in New Zealand
can provide adequate fire resistance and be compliant with the NZBC, including
Buchanan (2007) and Beattie (2001). NZBC clauses C1–C6 provide guidance
as noted in the previous section. NZS 3603:1993 also provides options for
determining fire resistance ratings for loadbearing timber systems such as walls
and floors in LTF buildings. The most applicable of these methods for LTF
includes testing according to AS 1530.4-2005 Methods for fire tests on building
materials, components and structures – Part 4: Fire-resistance test of elements of
construction and extrapolations from these tests using well established criteria.
These tests establish fire resistance ratings based on the structural adequacy,
integrity and insulating ability of the tested system, expressed in minutes for
each in the form of 60/60/30 (Bano-Chapman, 2008). At the time of writing, the
fire resistance portions of NZS 3603:1993 are in the process of being replaced by
NZS AS 1720.4, which will include these test requirements in the future.

Internal LTF wall and floor systems are typically clad with proprietary lining
materials, such as gypsum plasterboard, which provide adequate resistance
to fire if detailed and installed correctly according to technical information
provided by the product manufacturers. These linings provide fire protection
to the timber elements, depending on the lining thickness, stopping methods
and fixing to the timber framing, all of which need to be specified based on
testing and subsequent extrapolations of test results. Linings to internal walls
and ceilings other than plasterboard can also be considered for LTF buildings,

89
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

but designers have to ensure they have acceptable flame spread and smoke
development indices, which are also determined through testing. Sprinklers
are an option for reducing the required level of fire resistance in LTF buildings
and can be specified using accepted design methods, usually by fire engineering
practitioners.

Exterior cladding must have adequate resistance to fire to ensure that fire spread
involving the building façade is not excessive, to avoid the behaviour observed
in recent fire disasters in Australia and the UK (White & Delichatsios, 2014). Full-
scale façade fire tests as well as small-scale cone calorimeter tests can be used
to determine fire spread characteristics of façade systems and exterior cladding
materials, respectively, for multi-storey buildings. These data are available in
technical literature from the manufacturers of cladding systems and should be
considered as part of the façade design.

7.4 Fire during construction of LTF buildings


Another serious aspect of fire safety is during the construction of LTF buildings
where the potential fire loading can be significantly higher due to the lack
of linings and large areas of exposed timber that can easily ignite. The fire
engineering design of every LTF building should include the possibility of a fire
during construction (Spearpoint, 2008). A New Zealand guide by STIC (2012) for
fire safety in timber buildings sets out that the overall strategy is to:
■■ prevent access to the building site after hours
■■ undertake a rigorous fire risk assessment as part of an overall safety plan
■■ plan for all possible emergencies
■■ prevent ignition and mitigate fire risk as much as possible
■■ identify and manage potential fuel
■■ implement early installation of fire-safe construction.

Similarly, the Structural Timber Association provides recommendations around


separation distances for timber buildings during construction (STA, 2014), and
the UK Timber Frame Association has guidelines around fire safety for timber
frame construction sites (UKTFA, 2008).

7.5 Summary
Despite having a structural system comprised primarily of small sections of
timber, with appropriate protection and detailing, multi-storey LTF buildings can
provide the same Code-compliant fire resistance that other structural materials
achieve. This includes reduction of occupant injury, the ability to fight a fire
when it occurs and protection to surrounding properties from damage caused by
a fire. Because LTF buildings most typically include plasterboard interior linings
for walls and ceilings, the timber framing is protected from fire by commonly
used construction methods and materials. Tested wall systems are specified in
manufacturers’ literature and can be included in designs to achieve fire safety
compliance. Interior linings and exterior cladding also need to be considered to
ensure adequate protection against fire spread and smoke generation.

90
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

ACOUSTIC PERFORMANCE IN
8
MULTI-STOREY LTF BUILDINGS

8.1 Introduction
Acoustic performance is something that is not typically considered for buildings
designed according to NZS 3604:2011, but for multi-storey LTF buildings that
will likely have multiple residents or occupants, it must be included as part of
the design of the building. The overall objectives are to reduce the transmission
of sound between occupied units and from public spaces such that noise levels
within inhabited spaces are within acceptable limits. This includes both airborne
noise such as people talking and impact noise transmitted through direct contact
with the building components such as footsteps (Buchanan, 2007). The NZBC
provides requirements for acceptable sound transmission levels for wall and
floor systems for buildings in New Zealand.

As multi-storey buildings become more common throughout New Zealand, there


have been research projects conducted that consider the acoustic performance
of these buildings and methods for providing adequate sound transmission.
A recent study was conducted around the acoustic performance of medium-
density housing, which includes multi-storey buildings up to 6 storeys (Dunn et
al., 2018). The study highlighted the need for more easily accessible technical
information on design solutions for acoustic performance. It was recommended
that an online information repository be developed to assist designers, which
could be updated easily and kept current. It was also noted that more work
is required to allow for overseas solutions to be incorporated within the New
Zealand built environment for these applications.

Much like fire resistance design, it is good practice to consider how to achieve
acceptable acoustic performance early in the design of a building. It can lead

91
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

to more economical and efficient designs when fire, structure and acoustics
are considered together rather than in isolation. This is due to the conflicting
requirements of fire and structural design that seek well connected building
components, whereas acoustic design methods include separation and isolation
of building components (Walther & Beattie, 2013). It is possible to satisfy all these
requirements including acoustic performance with LTF construction utilising
Acceptable Solution G6/AS1 of NZBC clause G6 Airborne and impact sound or
using proprietary systems verified using Verification Method G6/VM1.

8.2 NZBC requirements


The NZBC provides performance requirements for acoustic transmission in
clause G6, which state that the sound transmission class (STC) must be no less
than 55 and the impact insulation class (IIC) must be no less than 55. The limit
on STC applies to walls, floors and ceilings, while the limit on IIC is for floors
only. These ratings are derived for different building assemblages through testing
according to accepted test methods described in the NZBC. Acceptable Solution
G6/AS1 provides some construction techniques that can be used to minimise
sound transmission through building components, which include physical
separation of building elements, use of noise control building elements, avoiding
rigid connections of services passing through building elements and creating
airtight seals around penetrations and service fittings through building elements.
This Acceptable Solution also provides details for floor and ceiling systems
that are deemed to comply with the NZBC, which include timber stud and joist
assemblies for use with LTF buildings. Building assemblies not falling within the
Acceptable Solutions can be verified using Verification Method G6/VM1, which
provides the required test methods for determining STC and IIC.

It is worth noting that the Acceptable Solutions and Verifications Methods for
sound transmission provide minimum solutions that address the STC and IIC
requirements, which are single figures and may not fully capture the dynamic
nature of sound transmission experienced by occupants. Additional aspects that
can affect acoustic performance include wall and connection detailing, floor
plan layouts, details around furnishings and external noise (Beattie, 2001).

8.3 Methods for achieving acceptable acoustic ratings in


multi-storey LTF buildings
As described in the previous section, there are Code-compliant systems that can
be included within multi-storey LTF buildings that will provide the necessary
levels of noise transmission reduction included in the NZBC. Additional
options for detailing of walls and ceiling constructions can be found in Beattie
(2001) and Buchanan (2007), which address some of the additional aspects of
noise control mentioned above. It is noted by Buchanan that LTF can provide
sufficient acoustic performance if detailed appropriately and can compete with
other structural materials such as masonry and concrete. There are numerous
proprietary systems available to the New Zealand market that have been
verified utilising G6/VM1 and are detailed in technical literature provided by
manufacturers for LTF applications.

Research has been conducted to develop building systems and components


that provide options for improved acoustic performance of multi-storey LTF
buildings. Walther and Beattie (2013) developed and verified through acoustic
and fire tests a series of wall, floor and wall-to-floor construction details for LTF
buildings that are compliant with the NZBC requirements. They also provide

92
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

sound transmission ratings for a range of floor and wall constructions including
different floor coverings, which can be incorporated within designs of multi-
storey LTF buildings. Emms et al. (2015) developed a structural connection
system to be used in multi-unit residential LTF buildings that was able to transfer
seismic and wind loads without compromising on acoustic performance. The
system was tested and verified, and estimates were developed around the cost of
implementing the system within LTF buildings.

8.4 Summary
While there exist methods for achieving the NZBC clause G6 requirements for
sound transmission that are applicable to multi-storey LTF buildings, other
solutions are available through proprietary systems and research programmes
specific to multi-storey LTF buildings. Because LTF buildings have less mass
than other structural systems, there can be challenges to achieving acceptable
acoustic performance for multi-storey applications. It is important to consider
how to incorporate sound transmission performance with structural and fire
resistance requirements to create robust LTF designs, especially for multi-unit
residential buildings.

93
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

VERTICAL MOVEMENT IN
9
MULTI-STOREY LTF BUILDINGS

9.1 Introduction

All structures must cope with the influence of short-term and long-term
movement during the lifetime of the building, and this must be considered
during design. Typically, this is of more importance in a tall LTF building, which
can display increased levels of vertical movement over time than other structural
materials or systems. Vertical movement includes contributions from settlement,
shrinkage and the effects of loading and can be described as follows:

settlement + elastic deformation +


total vertical moment =
creep + shrinkage/swelling

The following sections will describe sources of vertical movement and how these
movements can be estimated. A range of solutions for reducing and accommodating
these movements can be found in Ni and Popovski (2015) and TRADA (2011).

9.2 Individual contributions to vertical movement in LTF


buildings
9.2.1 Settlement within the timber structure

No interface between structural members is perfectly flat or square, and these


imperfections in surfaces will tend to close under loading. This is due to gaps
closing as the structure resists the applied dead and live load forces. Typically,
however, settlement deflections will occur during construction and are not
critical for consideration during the life of the building because gaps will be

94
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

already closed before other components are installed. However, exceptions to


this will occur if a component that will not settle (for example, a concrete lift
shaft) is present during construction of timber-framed portions of the building.

The prefabrication of framing, typical to the New Zealand construction industry,


can help to reduce the impact of settlement.

9.2.2 Elastic deformation

Timber is an anisotropic material with a relatively low modulus of elasticity


when compared to other common construction materials. This is especially
true in the perpendicular to grain direction, and timber members loaded in
this direction may undergo sizeable amounts of elastic deformation. In LTF
buildings using platform construction, bottom plates, top plates, floor joists
and occasionally floor sheathing are all loaded in the perpendicular to grain
direction. While for buildings of less than 3 storeys the elastic deformation of
studs loaded parallel to the grain can be ignored, in multi-storey LTF buildings,
the long total length of studs being loaded means that the elastic deflection of
the studs should also be considered.

Similar to settlement, elastic deformation of the framing will occur progressively


during construction. Depending on construction sequencing (i.e. platform or
balloon framing, when cladding is applied), this may be accounted for in the
estimation of total vertical movement.

9.2.3 Creep

While the elastic deflections discussed in section 9.2.2 are completely


recoverable and proportional to the level of load, additional deflection due to
stress will occur over time and is known as creep. Creep is non-recoverable upon
the removal of load and is especially pronounced under elevated levels of load
and/or large fluctuations of moisture content. Buchanan (2007) provides more
details on the effects of creep in timber. Creep can occur both parallel to grain
and perpendicular to grain, but the effects parallel to grain tend to be minimal.
If flexural elements are included within the vertical load resisting system of a
building, it is important that their creep contribution be considered.

Fluctuations in moisture content, such as the use of green timber that is then
allowed to dry in service, have the potential to significantly increase the impact
of creep. Mechano-sorptive creep, created by the molecular shifting of wood
fibres during moisture cycling, further increases creep effects. The use of kiln-
dried timber and controlling of humidity inside wall cavities are the best ways of
reducing the impact of creep deflections.

9.2.4 Shrinkage

The term ‘dry’ wood is a misnomer as all wood contains some moisture in
service. Changes in this moisture content create dimensional changes of
the timber. Once the moisture content (MC) of wood drops below the fibre
saturation point (FSP), shrinkage occurs. Timber in constant conditions
of humidity and temperature will gain (or lose) moisture until it reaches
equilibrium with the atmosphere around it. This level of MC is known as
the equilibrium moisture content (EMC). As such, timber with an MC below
FSP placed into a wall cavity will reduce or increase dimension until it has
reached its EMC.

95
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The dimensional stability of timber as the moisture content changes depends


on the grain direction. Considerably more shrinkage and swelling occurs in the
transverse (radial and tangential) direction than the longitudinal direction. In
the longitudinal direction, shrinkage of most timbers from the FSP to 12% MC is
approximately 0.1% –less than one-tenth that of the transverse direction – and
is normally not significant enough to be considered in 1–2-storey structures.
In multi-storey LTF structures, however, it should be accounted for due to the
cumulative lengths of stud used in the multi-storey walls.

As mentioned above, shrinkage and swelling in the transverse direction is


significantly more pronounced than the longitudinal direction. Radial dimensional
change is typically less than tangential change. However, designers, specifiers and
builders have little to no control over the orientation of the radial and tangential
components of structural timber. As such, an average value can be used as discussed
in section 9.3.3. Shrinkage transverse to the grain in members transferring vertical
load perpendicular to the grain should be accounted for in the design.

The EMC for the timber within the framing of an LTF building will depend on the
temperature and relative humidity within the wall cavity. NZS 3602:2003 provides
a range of allowable moisture contents for framing at the time of enclosure with
an upper limit of 20% in most cases. There is currently no available data on the
temperature and relative humidity conditions within wall cavities of multi-storey
LTF buildings or for LTF stand-alone houses. It has been suggested that the EMC for
framing members is 8–12% for buildings having central heating and air conditioning
and 10–14% for buildings that are intermittently heated (Buchanan, 2007). It is
assumed that most multi-storey LTF buildings will have some degree of climate control
that will be used on a regular basis. Therefore, for determining potential shrinkage
in framing members, it is suggested that an EMC of 11% would provide a reasonable
estimate of timber conditions for most larger LTF buildings in New Zealand.

9.3 Estimation of vertical movement in LTF buildings


The total vertical movement of the building will be a summation of the
movements due to elastic deformation, creep and shrinkage, with settlement
generally considered negligible in design.

While the equations for estimating vertical movement presented below are
simple, a certain amount of engineering judgement will be required to assess
where vertical movements are critical. In most multi-storey LTF buildings,
numerous and varying load paths will exist with members being loaded to a
wide range of load/capacity ratios. For example, although there is likely to be a
concentration of studs at each end of a lateral loadbearing wall, gravity loading is
evenly spread across the wall. This means that, while under an LLRS check, the
end studs and plates will be critical as discussed in section 3.2.3, and the more
widely spaced central studs will be critical for the check of vertical movement.

While the sheathed LTF walls themselves will act as a sort of deep beam resisting
settlement, it should not be assumed that the system is capable of averaging
vertical movement across a floor.

9.3.1 Estimating elastic deformation

The elastic deformation of each member can be calculated based on engineering


principles. Due to the high number of timber members along the load path, the
average modulus of elasticity can be used.

96
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

For the studs, the elastic deformation is calculated as:

(9.1)

where:
Δel,st,i elastic deformation of the wall stud at level i
Ni load on the wall stud at level i
lst,i length of the stud at level i
E parallel to grain modulus of elasticity
Ast area of stud under consideration

For the top and bottom plate and joists, members loaded in the perpendicular to
grain direction, the elastic deformation is calculated as:

(9.2)

where:
Δel,0,I elastic deformation of the top and bottom plate and joist at the level i
nplate number of top and bottom plates at level i
hplate vertical dimension of plates at level i
hjoist vertical dimension of joists at level i

Values for perpendicular to grain stiffness E0 are not available in NZS AS 1720.1.
However, this value can be taken as follows for dry timber:

(9.3)
9.3.2 Estimating creep deformation

Creep deformation is estimated in NZS AS 1720.1 by multiplying the short-


term elastic deformation by an appropriate modification factor for creep. For
members in compression, this factor j2 in most cases will be 2 as long-term
loading will be for longer than 1 year and initial MC levels should be less than
15%. Therefore:

(9.4)

where:

Δtotal total deformation due to elastic deformation and creep.

Figures in NZS AS 1720.1 provide j2 values for higher levels of initial moisture
content.

97
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

9.3.3 Estimating shrinkage deformation

Similar to the estimation of elastic deformation, the estimation of the impact of


shrinkage will be a combination of shrinkage due to the studs and shrinkage due
to the top and bottom plates and joists.

In order to calculate the shrinkage as the building reaches its EMC, the
percentage moisture change must be calculated. This is done by subtracting
the average moisture content within the wall cavity from the initial moisture
content. Conservatively, this initial value could be taken as 20%, giving a
potential moisture content change of 9%.

The following discussion considers the estimation of shrinkage effect on radiata pine
as this species is most commonly used in construction in New Zealand. Shrinkage
properties of other New Zealand timbers are available in Buchanan (2007).

As discussed in section 9.2.4, shrinkage in the longitudinal direction is


considerably less than in the transverse direction. However, due to the
long length of studs used in a multi-storey LTF building, it still needs to be
considered. Most timbers shrink along the grain by approximately 0.1%
when drying from green (FSP) to 12%. For radiata pine, this gives a shrinkage
coefficient longitudinal to the grain of 0.006% per % MC change. The
longitudinal shrinkage of the stud is calculated as:

(9.5)

where:
Δshr,st stud deformation due to shrinkage

As described in section 9.2.4, it is not practical to control if the plates and joist
will be oriented in the tangential or radial direction and therefore the average
shrinkage/swelling properties will be used. For radiata pine, the tangential
and radial shrinkage from green (29% MC) to dry (12% MC) are 3.9% and 2.1%,
respectively. The average of these values gives a 0.18% dimensional change per %
MC change.

The shrinkage due to timber placed in the transverse direction is calculated as:

(9.6)

where:
Δshr,0 plate and joist shrinkage

9.4 Summary
This section has provided sources for vertical movement in multi-storey LTF
buildings and methods for calculating these deformations. In cases where
deformations become significant, it may be necessary to detail façades or other
parts of the building to accommodate these movements. More details on how to
address these issues are available in TRADA (2011).

98
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

10
DESIGN EXAMPLE

In order to assist designers with using these guidelines for multi-storey LTF
buildings, a building example was chosen that would be simple enough to be
an example yet have enough complexity to be realistic. This section provides
descriptions of a 4-storey block of flats in Wellington and goes through the
design of the building systems and components as previously discussed in this
guidance document. The original building design was provided by Allan Eng
of Interspace Architects as an example of a typical multi-storey LTF residential
building. Some minor changes to the original design had to be implemented to
get a Code-conforming solution. Elevations of the building are shown in Figure
26 and a typical floor plan is provided in Figure 27.

Figure 26. Architectural rendering of example LTF building (used with permission from Interspace Architects).

99
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

10.1 Gravity design


The gravity design of the building is separate from the lateral design and will be
briefly explained here. The roof, walls, lintels, floors and subfloor need to be
designed for the appropriate load combinations. The roof and subfloor design
are not shown in this example as their design methods are not unique for LTF
buildings, and their designs vary depending on the system being used.

The dead and live loading on the building has been calculated in detail in section
10.2.1.

10.1.1 Floor design

10.1.1.1 Floor joists

The floor has been designed assuming the layout shown in Figure 27, with
a maximum joist span of 4.18 m. Table 3 in section 5.3.2 has been used to
determine the required joists assuming that the span is vibration controlled.
The longest span using SG8 sawn timber would be 3.95 m using 290 x 45 mm
members, so LVL13 joists are required. Therefore, 300 x 45 mm LVL13 joists at
600 mm centres having a maximum span of 4.25 m according to manufacturer
literature have been chosen for this design.

4008
LIFT
BED 2 BED 3
LOBBY

STORE

BATH LAUNDRY
STUDY
SCULLERY
KITCHEN

ENSUITE
5200

2530 3922 3680


DRESSING
700

DINING

LOUNGE
BED 1

DECK

4180

POSSIBLE LOCATIONS OF BEAMS

Figure 27. Possible structural floor layout.

100
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The loads on the floor joists, including self-weight, are:

(10.1)

(10.2)

(10.3)

The maximum shear and moment in the joists, with a critical load combination
of 1.2G + 1.5Q, are:

(10.4)

(10.5)

Another option for the floor joists could be to use I-joists rather than solid timber
or LVL, which would allow for larger penetrations in the floor – an issue that is
often of concern for multi-storey buildings. Care must be taken when designing
a floor with I-joists however, to ensure that they have adequate vibration
performance.

10.1.1.2 Floor beams

The design of the floor beam between the lounge and the kitchen/dining area has
been shown, and all others would be designed similarly. The tributary width of the
beam is 3.8 m, and the length of the beam is 5.2 m. For the calculation of the beam
self-weight, it is assumed that a 250UB37 with a self-weight of 0.37 kN/m will be
used, which will then be verified.

(10.6)

(10.7)

(10.8)

The maximum shear and moment in the beam, with a critical load combination
of 1.2G + 1.5Q, are:

(10.9)

(10.10)

Therefore, the member is verified. An alternative approach could be to use a


timber beam instead of a steel beam in the floor. However, space is often an issue
in the floor cavity of a multi-storey building, so the choice of member would
largely depend on architectural detailing.

101
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Typically, the floor beams in an LTF design can be supported on studs that are
a part of the wall and act as columns. Depending on the load on the column,
built-up members can be used, or larger glulam, LVL or steel columns could be
integrated into the wall. As the loads on the columns are cumulative down the
building, studs may be able to be used in the upper levels while larger members
may be required in the lower levels. The columns would be designed in the same
way as the studs, which is shown in detail in section 10.2.4.

10.1.2 Wall design

10.1.2.1 Studs

The design of the LTF walls is based on the assumption that all walls and
openings are aligned along the height of the building. The walls in the
X-direction are non-loadbearing because the floor spans parallel to the walls, so
it is assumed that only the walls in the y-direction are loadbearing for gravity.
The maximum wall tributary width is therefore 4.1 m, for wall PY3 (see Figure 28).

The loads on the walls are:

(10.11)

(10.12)

(10.13)

(10.14)

(10.15)

For the calculation of the studs, the loads on the walls are cumulative down the
building. The critical stud is therefore on level 1.

With a critical load combination of 1.2G + 1.5Q, the distributed load on the
maximally loaded internal walls are:
Level wi wi (kN/m)

Roof 1.2(wroof + wwalls,4) (1.2)∙ (1.8 + 2.7) = 5.4

3 wroof + 1.2(wfloor,3 + wwalls,3) + 1.5(Qfloor) 5.4 + 1.2∙ (3.8 + 1.8) + 1.5∙ (6.2) = 21.4

2 w3 + 1.2(wfloor,2 + wwalls,2) + 1.5(Qfloor) 21.4 + 1.2∙ (3.8 + 1.8) + 1.5∙ (6.2) = 37.4

1 w2 + 1.2(wfloor,2 + wwalls,2) + 1.5(Qfloor) 37.2 + 1.2∙ (3.8 + 1.8) + 1.5∙ (6.2) = 53.4

Assuming a stud spacing of 600 mm centres, the load per stud at level 1
becomes:

(10.16)

102
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Wall studs are 140 x 45 mm SG8 members, with dwangs at 600 mm centres from
the bottom of the wall:

(10.17)

Refer to section 10.2.4 for a detailed explanation of the calculation of the


compression capacity of the studs. For an external wall, the stud would also
need to be checked for wind loading on the 45 mm wide face of the member.

Bottom plates and connections of walls can be designed according to NZS


3604:2011.

10.1.2.2 Lintels

The lintels in this building are aligned up the height of the building, so if the
lintels are designed with trimmer studs to carry the cumulative gravity load
down the structure, the lintels only need to carry the weight of the façade or
roof directly above them. The lintels can therefore be designed according to NZS
3604:2011.

The lintels off the bath, ensuite and dressing rooms have a span of 700 mm and a
loaded dimension of 2.1 m.

The building has a light roof and floor joists that span perpendicular to the
lintels, so Tables 8.9 and 8.11 of NZS 3604:2011 can be used to determine
required lintel sizes:

Level Lintel size (SG8) (mm)

4 90 x 70

3 140 x 70

2 140 x 70

1 140 x 70

10.2 Lateral load design

10.2.1 Determination of seismic demand

The two-stage approach has been used to determine the seismic demand of the
building, as outlined in section 2.4.12.

In order to calculate the design base shear, the seismic mass of the structure must
first be determined. The weight of the split roof level has been added to level 4.

10.2.1.1 Dead loading

The following material densities have been assumed for the calculation of the
dead loads of the building:
■■ Radiata pine – 500 kg/m3
■■ Ply sheathing – 550 kg/m3
■■ LVL – 600 kg/m3

103
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The dead loading will be approximated by considering the floor, wall and roof
loads.

The load from the 300 x 45 mm LVL13 joists at 600 mm centres, as determined
in the gravity design, is:

(10.18)

The weight of the 15 mm ply sheathing as well as an additional 0.5 kPa for the
floor finish, architectural fittings, acoustic separation and so on are added to this
to find the total weight of the flooring:

(10.19)

The typical weight per square metre of external walls will be taken as 0.6 kPa,
as the wall is assumed to be between a light and medium wall in Shelton (2013).
With a total perimeter wall length of 59.1 m, the typical weight at each floor
becomes:

(10.20)

For level 4, which has higher walls, the weight of the walls becomes:

(10.21)

An additional partition load of 0.2 kPa will also be applied over the floor area to
account for internal walls.

Considering a light roof of 0.2 kPa with a ceiling of 0.24 kPa, as assumed in
Shelton (2013), the roof load will be taken as:

(10.22)

10.2.1.2 Live loading

As a self-contained occupancy unit, the live loading on each floor is:

(10.23)

10.2.1.3 Seismic weight

The area of each floor and the roof is:

(10.24)

In order to calculate the area reduction factor for each floor, the area of each
apartment will be calculated, neglecting the stairwells, storage area, lift and
lobby (26 m2) and the decks (16 m2):

(10.25)

104
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The area reduction factor for each floor is calculated in accordance with
NZS 1170.5:2004 as:

(10.26)

The seismic weight of each floor then becomes:

(10.27)

Wi Wi (kN)

Groof + Wwalls,4/2 (0.44 180) + (159.6/2) = 159

(0.72 180) + (106.4 + 159.6)/2 + (0.2 180) +


Gfloor + (Wwalls,4 + Wwalls,typ)/2 + Gwalls,int + Qfloor 352
(1.5 0.3 (138 0.56 + 42)) =
(0.72 180) + 106.4 + (0.2 180) + (1.5 0.3
Gfloor + Wwalls,typ + Gwalls,int + Qfloor 326
(138 0.56 + 42)) =
(0.72 180) + 106.4 + (0.2 180) + (1.5 0.3
Gfloor + Wwalls,typ + Gwalls,int + Qfloor 326
(138 0.56 + 42)) =

The total building weight Wtot hence becomes 1163 kN.

10.2.1.4 Building period

An initial approximation of the building period is calculated based on the height


of the structure and the empirical formula presented in section 2.4.3:

(10.28)

Once the walls have been designed and detailed, the fundamental period will be
refined.

10.2.1.5 Elastic site spectra

The building site has been classified as soil type C. Knowing the building period,
the spectral shape factor becomes:

(10.29)

The remaining factors required for the calculation of the elastic site spectra are:

Hazard factor (Wellington) Return period factor (ULS) Near fault factor

Z = 0.4 RULS = 1 N=1

10.2.1.6 Design site spectra and design base shear

For a ductility μ = 2 and a structural performance factor of Sp = 0.7, the following


design spectral shape factor can be determined:

(10.30)

105
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The base shear of the structure hence becomes:

(10.31)

10.2.1.7 Overstrength limit

The overstrength limit is determined by taking the ratio of the elastic site spectra
from the equivalent static analyses with ductility equal to 1 and Sp = 0.7 and
ductility equal to 2 and Sp = 0.7:

(10.32)

This value provides a maximum limit of the required overstrength reasonably


expected for the capacity protected elements.

10.2.2 Determination of wind demand

As noted in section 2.3, the determination of wind loads is no different for a


multi-storey LTF building than for any other multi-storey building. For this
example, assumptions were made around the topography and the surrounding
obstructions to derive wind speed and pressures. For the purposes of the lateral
load response and design of the building, the seismic loading has been used to
derive applied loads to be resisted and subsequent design of the LTF systems
and components.

10.2.3 Lateral load distribution into walls

The storey forces or storey shears from the lateral loads need to be distributed
into the individual bracing walls. Since the floor diaphragms have been designed
for both a flexible and rigid diaphragm approach, the forces can be distributed
assuming a rigid diaphragm and thus also take into account geometric
and accidental eccentricities. In this design example, only the step-by-step
calculations for the seismic forces are shown. The wind forces can be calculated
in an analogous way.

Assuming the wall lengths and positions as per Figure 28 and Table 5, the
centre of rigidity was determined assuming the wall stiffness is proportional to
its length. This assumption can be justified, as the deformation components Δ3
and Δ4 from the panel shear and fastener slip, respectively, are predominant
(about 80% of the total wall deformation as per section 10.2.5) and are hence
proportional to the wall length:

(10.33)

The centre of mass has been determined as the centre of the floor area,
assuming that the seismic mass is evenly distributed across the whole floor.

(10.34)

106
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Figure 28. Wall layout with centre of mass and centre of rigidity.

The same positions for centres of rigidity and mass have been assumed for all levels.

The geometric eccentricity results in the following:

(10.35)

Table 5. Bracing wall lengths.

Wall Length (m) Wall Length (m)

PY1 4.9 PX1 3.2

PY2 4.2 PX2 6.6

PY3 2.6 PX3 2.3

PY4 4.6 PX4 1.6

PY5 4.8 PX5 1.6

PY6 3.4 PX6 3.0

PY7 2.1

PY8 12.2

107
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The following torsions due to geometric eccentricity need to be taken into


account. These values will mainly be used for deformation checks.

(10.36)

According to NZS 1170.5:2004, an accidental eccentricity of ±10% of the building


dimension perpendicular to the load direction under consideration needs to be
taken into account.

(10.37)

The following torsions including both geometric and accidental eccentricities


need to be considered:

(10.38)

The base shear forces Vb = Vb,x = Vb.y and the torsions can now be distributed
across each shear wall m at each level i with the following equations:

(10.39)

(10.40)

Table 6 gives wall forces and relative input data for seismic loading in the
y-direction.

108
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Table 6. Base shear force distribution (ULS in y-direction, including accidental eccentricities).

Wall ID l (m) x (m) y (m) a (°) x’ (m) y’ (m) V[m] (kN)

PY1 4.9 0.0 -2.5 90 -7.6 1.5 39.36

PY2 4.2 0.0 -11.1 90 -7.6 -7.1 33.37

PY3 2.6 4.0 -1.3 90 -3.6 2.7 17.40

PY4 4.6 4.0 -10.9 90 -3.6 -6.9 30.73

PY5 4.8 8.0 -1.3 90 0.4 2.6 26.35

PY6 3.4 10.5 -0.6 90 2.8 3.4 19.20

PY7 2.1 10.5 -11.2 90 2.8 -7.2 11.89

PY8 12.2 14.1 -6.1 90 6.4 -2.1 71.72

PX1 3.2 4.0 0.0 0 -3.6 4.0 0.76

PX2 6.6 7.4 -3.8 0 -0.2 0.2 0.07

PX3 2.3 12.8 -3.8 0 5.1 0.2 0.02

PX4 1.6 13.3 -1.9 0 5.6 2.1 0.20

PX5 1.6 13.3 0.0 0 5.6 4.0 0.39

PX6 3.0 5.5 -12 0 -2.1 -8.0 1.45

The individual wall base shear forces can be distributed up the height of the
walls using the approach as per clause 6.2.1.1 of NZS 1170.5:2004.

In this example, only wall PY1 will be analysed and detailed further. All other
walls can be detailed and verified in the same way.

The storey shear and overturning moments for wall PY1 are summarised in Table 7.

Table 7. Force demand and geometric information on wall PY1.

M* v*
Level V* (kN) F* (kN) l (m) h (m) hcum (m)
(kNm) (kN/m)

4 12.6 0.0 12.6 4.9 4.5 13.5 2.6

3 26.5 56.6 13.9 4.9 3.0 9.0 5.4

2 35.1 136.0 8.6 4.9 3.0 6.0 7.1

1 39.4 241.2 4.3 4.9 3.0 3.0 8.0

0 39.4 359.3 0.0

10.2.4 Design of wall components

The shear flow (unit shear force) v can be obtained with the following equations,
and the values for all storeys can be found in Table 7.

(10.41)

109
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

First, the fastener capacity and spacing need to be verified, being the ductile link
of the lateral load resisting system. Standard 2.80 mm diameter flathead nails at
100 mm spacing at the ground level have been selected. The fastener spacing up
the building height was increased based on decreasing shear as shown in Table 8.

(10.42)

All other structural members including the sheathing panels, chords and hold-
downs are designed as capacity protected elements and need to be verified
against the overstrength capacity of the fasteners. Since their demand is coming
from an element undergoing large displacements, the capacity reduction factor
ɸ can be taken as 1.0. The nail overstrength factor has been taken as 1.6 as per
NZS AS 1720.1.

A 15 mm thick grade F8 plywood has been used for all walls.

(10.43)

The chords need to be checked for both tension and compression forces. Their
demand needs to be calculated taking into consideration the overstrength
capacity of the fasteners at each storey.

The chord force component from the ductile fastener capacity including
overstrength at each storey can be calculated as follows:

(10.44)

The total chord demand at overstrength for the multi-storey structure is the sum
of all force components from the levels above the level under consideration.

(10.45)

A summary of the shear capacity at overstrength from the panel fasteners


and the respective chord forces can be found in Table 8. Where the ratio of
overstrength capacity to demand is greater than the overstrength limit of 2
calculated in section 10.2.1, the limit has been applied. For wall PY1, the limit has
been applied for levels 2–4.

110
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Table 8. Load demand in terms of shear flow and chord forces for capacity protected elements (CPE) in wall PY1.

Level V (kN/m) vCPE (kN/m) Overstrength (ΔT=ΔC)CPE (T = C)CPE


(kN) (kN)

4 2.6 5.1 2.0 23.2 23.2

3 5.4 10.7 2.0 33.1 56.3

2 7.1 13.9 2.0 43.7 100.1

1 8.0 13.9 1.7 45.0 145.1

Wall PY1 is carrying gravity loads that need to be considered when verifying wall
elements. When calculating the tension capacity of the studs and the anchorage
rotation capacity of the hold-downs, a reduction due to G + ψEQ is considered,
and when calculating the compression capacity of the studs and the bearing
capacity of the bottom plates, an additional G + ψEQ is considered. The gravity
loads are the same as those used to calculate the seismic weight of the building
in section 10.2.1.

The chord members are verified for both tension and compression forces. The
chords at the ground floor are selected to be made of eight 45 x 140 mm studs:

(10.46)

For the compression check, the buckling loads in both the strong and weak axis
need to be considered. In the strong axis (wall out-of-plane), the buckling length
can be considered as 0.9 the real height, since the lining provides some restraint
to the stud ends. In the in-plane direction, the studs are restrained by dwangs,
which are assumed to be at 600 mm centres.

(10.47)

where:

(10.48)

(10.49)

where S3 and S4 are the slenderness ratios in the strong and weak axis,
respectively.

111
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The stability factor k12 was calculated based on the greater of S3 and S4 and:

(10.50)

Therefore,

(10.51)
and
(10.52)

Next, the bearing capacity of the bottom plate under the chords needs to be
verified:

(10.53)

Using eight 45 mm studs may not provide the most economic design, but
the cross-section is required to fulfil the bearing area check. To improve the
design, an LVL bottom plate or compression reinforcement could be used, thus
limiting the cross-section of the chords. Additionally, larger sawn LVL or glulam
members could be used instead of built-up chords made of individual studs.

Finally, the hold-down brackets to resist the uplift forces need to be verified. For
this design, assumed hold-down values as shown in Table 9 have been used. It is
important that appropriate material strength reduction factors for the different
failure modes are used.

Table 9. Hold-down capacity values using ɸ0= 1.6 and k1 = 1.14.

Characteristic strength
Hold- (kN)
ɸ min min
Kt
down (T,S) (T,S,C)
Timber Steel Concrete 1 0.9 0.65

XD1 35.0 40.0 60.0 12000 39.9 36.0 39.0 36.0 36.0

XD2 50.0 80.0 100.0 17500 57.0 72.0 65.0 57.0 57.0

XD3 75.0 80.0 100.0 25000 85.5 72.0 65.0 72.0 65.0

XD4 100.0 100.0 150.0 35000 114.0 90.0 97.5 90.0 90.0

XD5 150.0 160.0 180.0 45000 171.0 144.0 117.0 144.0 117.0

For the bottom storey, two XD4 brackets are used. The following capacity can be
achieved:

(10.54)

(10.55)

112
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

These are generic, non-proprietary brackets with assumed strength and stiffness
values, which in practice will need to be verified by the manufacturer. For this
bracket size, the steel failure is governing. However, for other bracket sizes (such
as the XD3 and XD5 brackets), the concrete failure (ɸ ˙ Qk concrete) may govern.
The concrete failure only occurs at the bottom storey where the wall is fixed to
the foundation, but for the upper storeys, only the steel and timber capacities
need to be verified, so the next lowest capacity value can be used. Table 10
summarises all wall elements and their capacity for wall PY1 including fastener
spacing, plywood thickness, chord dimensions and hold-down types.

Table 10. Required section sizes and fixings for wall PY1.

Sheathing Fasteners Chords


Bearing Hold-down
(plywood F8) (nails 2.80 mm) (140 x 45 mm SG8 studs)
Level
vd,I vd dchord Nd,t Nd,c Nd,p Type Qd,t
T (mm) S (mm) nstuds (-)
(kN/m) (kN/m) (mm) (kN) (kN) (kN) (-) (kN)

4 15 42 250 5.1 1 45 38 30 40 1/XD1 36

3 15 42 125 10.7 3 135 113 200 94 1/XD2 57

2 15 42 100 13.9 5 225 189 333 142 2/XD2 114

1 15 42 100 13.9 8 360 302 533 227 2/XD4 180

10.2.5 Wall deformation and drift checks

In this section, the deformation of wall PY1 is calculated under ULS conditions,
with all walls in both directions requiring verification in the same way.
Accidental eccentricity has not been considered for deflections checks.

The wall deflections up the building includes six different contributions:


(10.56)

Only the deflection of the wall PY1 at level 4 is calculated below. All other storey
deflections can be calculated similarly. First, the elastic deflections under ULS
loads are calculated, and these then need to be amplified to account for the
ductile behaviour of the structure. SLS deflections are calculated in a similar
manner with respective SLS loads.

Table 11. Shear and moment demand for wall PY1 at ULS without accidental eccentricity.

Level V (kN) M (kNm)

4 10.3 0.0

3 21.6 46.1

2 28.6 110.9

1 32.1 196.7

0 32.1 292.9

Flexural deformation due to storey shear:

(10.57)

113
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Flexural deformation due to overturning moment:

(10.58)

Shear deformation of the sheathing panel:

(10.59)

Deformation due to fastener slip:

(10.60)

The fastener slip modulus Kser has been derived based on the actual load
demand on the nail. The number of panel splices n and m are derived based on a
standard plywood panel dimension of 1.2 x 2.4 m. Individual fastener slip can be
calculated as:

(10.61)

The anchorage slip, which is effectively causing a rotation at the base of the wall
at each storey, can be calculated as:

(10.62)

where:

(10.63)

The hold-down stiffness Kt is normally provided by the manufacturer or needs to


be derived from analytical equations or from literature. For the ULS, the elastic
stiffness needs to be reduced by a factor 2/3 as per Eurocode 5. When equating
the rotational stiffness, the number of hold-downs need to be taken into account
considering that multiple hold-downs increase the stiffness, but the presence of
hold-downs at the bottom of the wall at the level under consideration and the
hold-downs at the top of the wall at the level below reduce the stiffness by half.

114
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

When calculating the compression stiffness of the bottom plate, stress spreading
can be taken into account as follows.

(10.64)

where:

(10.65)

The deflection component Δ6 for level 4 depends on the rotations from the levels
below. These rotations depend on the flexural and anchorage slips from the
walls from level 1 to level 3.

(10.66)

The flexural rotations can be calculated as:

(10.67)

(10.68)

The anchorage rotation can be calculated as per Equation (10.63). All rotations of
wall PY1 are summarised in Table 12.

Table 12. Wall PY1 rotations due to flexural deformation and anchorage slip and as relative wall displacement.

Level �V,1(mrad) �M,1 (mrad) θ (mrad) �α mrad) �θ (mrad) ∆6 (mm)

4 0.17 0.00 0.54 0.48 2.08 11.5

3 0.06 0.08 0.90 0.34 1.18 4.6

2 0.05 0.12 0.86 0.17 0.32 1.5

1 0.03 0.14 0.32 0.00 0.00 0.0

The inter-storey deformation can be calculated as the sum of all deformation


components Δ1 to Δ6:

(10.69)

When calculating the total building deformation, all deformation components


need to be considered:

(10.70)

115
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Once the elastic deflection ULS contribution at all levels is calculated, the
ultimate ULS deflections can be obtained by multiplying by the ductility μ = 2,
the displacement modification factor kdt = √2 = 1.41 as per NZS AS 1720.1, the
inter-storey deflection factor km = 1.2 as per NZS 1170.5:2004 and the deflection
scale factor kd = 0.91 as per NZS 1170.5:2004.

The elastic and ultimate ULS deflections without accidental eccentricity are
summarised in Table 13 and Table 14, respectively.

Table 13. Elastic wall deformation contributions, elastic inter-storey deformation, elastic inter-storey drift and total
elastic displacement of wall PY1 under ULS loading.

∆1 ∆2 ∆3 ∆4 ∆5 ∆6 ∆storey drift ∆tot


Level
(mm) (mm) (mm) (mm) (mm) (mm) (mm) (%) (mm)

4 0.5 0.0 1.4 2.2 2.41 11.5 16.4 0.36% 41.1

3 0.1 0.1 1.9 1.9 2.71 4.6 10.3 0.34% 24.7

2 0.1 0.2 2.5 2.1 2.58 1.5 8.2 0.27% 14.4

1 0.1 0.2 2.9 2.7 0.96 0.0 6.2 0.21% 6.2

Table 14. Ultimate inter-storey deformation, inter-storey drift and total displacement of wall PY1 under ULS loading.

Level ∆storey μ (mm) drift μ (%) ∆tot μ (mm)

4 55.7 1.24% 139.4

3 35.0 1.17% 83.7

2 27.8 0.93% 48.8

1 20.9 0.70% 20.9

Since all walls are connected through the diaphragms and restrained to similar
deflections at each storey, instead of verifying the individual wall deflections,
the average deflections of all walls in the same direction are considered.
This is a simplified approach and may not always accurately capture specific
deformations of individual walls. The average values at SLS and ULS without
accidental eccentricity are shown in Table 15.

Table 15. Average ultimate inter-storey deformation, inter-storey drift and total displacement of the structure in the
y-direction under both SLS and ULS loading.

SLS ULS

Level Total Inter-storey Inter-storey Total Inter-storey Inter-storey


deformation deformation drift deformation deformation drift
(mm) (mm) (%) (mm) (mm) (%)

4 19.0 8.0 0.18% 142.5 61.3 1.36%

3 11.0 4.8 0.16% 81.2 36.4 1.21%

2 6.2 3.7 0.12% 44.9 26.9 0.90%

1 2.5 2.5 0.08% 18.0 18.0 0.60%

As can be seen from Table 15, the inter-storey drift limits of 0.33% and 2.5%
under SLS and ULS loading, respectively, have been satisfied.

116
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

In section 10.2.1, a global ductility of 2 has been selected. Since all the non-
linear behaviour of the structure is limited to the fastener deformation, the local
ductility of the fasteners needs to be verified. The ultimate wall deformation
component due to fastener slip ∆4 under ULS loading with accidental
eccentricity in wall PY1 at level 1 can be calculated as follows:

(10.71)

This value can be used to calculate the actual ultimate fastener displacement:

(10.72)
This value is smaller than the maximum allowable ultimate displacement as per
NZS AS 1720.1 of 6 mm. The local ductility of the fasteners can be calculated as:

(10.73)

As per the inter-storey drift, instead of verifying the individual wall


deformations, the average of all wall deformations in the considered direction
should be taken into account. The average ultimate fastener deformation of the
structure in the y-direction is summarised in Table 16.

Table 16. Ultimate fastener deformations in the y-direction.

PY1 PY2 PY3 PY4 PY5 PY6 PY7 PY8 Average


Level
(mm) (mm) (mm) (mm) (mm) (mm) (mm) (mm) (mm)

4 1.45 1.66 1.50 1.28 0.98 1.18 1.52 0.76 1.29

3 1.88 2.10 1.92 1.64 1.27 1.52 1.93 1.00 1.66

2 2.10 2.28 2.03 1.79 1.39 1.62 2.00 1.14 1.79

1 2.04 2.11 1.67 1.63 1.24 1.35 1.50 1.20 1.59

As can be seen, the average fastener slip is less than the maximum allowed value
of 6 mm.

10.2.6 P-delta effects

According to clause 6.5.1 of NZS 1170.5:2004, P-delta effects do not need to be


considered if (a) the largest translational period is less than 0.4 seconds or (b)
the structure is less than 15 m from the base and the largest period is less than
0.6 seconds. Since both these criteria are not satisfied, criteria (c), the maximum
value of the stability coefficient is less than 0.1, needs to be checked.

The stability coefficient at each storey can be calculated as:

(10.74)

The stability coefficients with the respective input values are shown in Table 17.

117
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Table 17. Calculation of stability coefficients in the y-direction.

∆ storey u,i
Level V (kN) W (kg) hi (m) �y
(mm)

4 68 15,900 4.5 70.6 0.04

3 143 35,200 3 42.2 0.03

2 190 32,600 3 31.6 0.02

1 213 32,600 3 21.7 0.01

As can be seen, the stability coefficients are always smaller than 0.1, and
therefore P-delta effects do not need to be considered in the analysis.

10.2.7 Building period

The building period can be estimated by using Rayleigh’s method and the elastic
SLS deflection of the structure. As discussed, the average deflection of all walls
in the given direction is considered. The accidental eccentricity is not to be
considered when determining the building period.

(10.75)

Table 18 summarises the parameters required in Equation (10.75), which


provides a building period of 0.60 seconds in the y-direction. The kd factor has
been set to 1.0 as per NZS 1170.5:2004 for the determination of the fundamental
period of the building.

Table 18. Seismic mass, storey forces and average elastic deformation in the y-direction of the building under
SLS loading.

Total elastic SLS


Level W (kN) F (kN) deformation in Y
(mm)

4 159 40.2 20.8

3 352 44.5 12.1

2 326 27.5 6.8

1 326 13.7 2.8

10.2.8 Numerical model

In order to develop a simple numerical model without the use of shell elements,
all bracing walls are modelled as cantilevered beam elements with equivalent
properties. The equivalent properties take all stiffness sources of the LTF walls
into account.

In some software packages, it is easier to define standard beam elements with


well defined material properties. To adjust for the wall assembly and all sources
of stiffness, the properties of these elements can then be reduced by stiffness
multipliers as defined in Equations (10.77) and (10.80).

118
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The beam elements should be placed at the centre of each wall with the section
length equal to the length of the wall. The beam width is arbitrary but should
ideally be taken as the thickness of the wall/studs.

The equivalent modulus of elasticity can be calculated based on the chord cross-
section and properties as well as the chosen equivalent thickness. For wall PY1 at
level 1, the following modulus of elasticity can be obtained:

(10.76)

This value can either be directly inserted in the material properties, or a bending
stiffness modifier can be used for the beam at the given level:

(10.77)

This modifier is used to reduce the bending stiffness of the element representing
the wall to reflect the fact that the flexural stiffness of the wall is given by the
deformation of the end chords, not the solid block representation.

Most common software programs enable the user to use this type of modifier
(i.e. a reduction of concrete beam stiffness to allow for cracking).

The equivalent shear stiffness, based on the sheathing panel properties and the
fastener type and spacing can be calculated as follows:

(10.78)

As explained in section 10.2.5, fastener slip modulus Kser has been derived based
on the actual load demand on the nail.

(10.79)

Similar to the equivalent modulus of elasticity, this value can either be directly
inserted in the material properties, or a shear stiffness modifier can be used for
the beam at the given level:

(10.80)

119
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Finally, the rotational stiffness at the wall base can be calculated. Since the
wall is fixed to the concrete foundation, the full stiffness of the two hold-down
brackets are taken into consideration.

(10.81)

In equation (10.81), the SLS stiffness has been used. For ULS analysis, 2/3 of this
stiffness would need to be considered. The compression stiffness was calculated
taking into account the stress spreading in the bottom plate:

(10.82)

where:

(10.83)

For clarity, the rotational stiffness at level 2 is also calculated:

(10.84)

In Equation (10.84), the hold-down stiffness has been halved due to the presence
of hold-downs at both the bottom of the wall under consideration and at the
top of the wall at the level below (two equal springs in series, halving the total
stiffness). Two hold-downs per chord are present. The compression stiffness has
been calculated taking into account the stress spreading in the horizontal timber
members as already shown in Equations (10.82) and (10.83).

The data in Table 19 has been inserted into the numerical model created with
the Finite Element Software RFem from Dlubal (Dlubal Software, 2019):

Table 19. Equivalent modulus of elasticity, shear modulus and rotational stiffness of wall PY1.

Level Eeq (MPa) Geq (MPa) Kθ (kNm/rad)

4 438 30 123,165

3 1314 37 177,132

2 2191 39 323,819

1 3505 38 1,323,506

120
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The diaphragms have been modelled as shell elements with a 15 mm plywood


membrane. To account for the fastener stiffness, the shear modulus of the
membrane was reduced using a stiffness modifier of 0.31 as shown in section 10.3.3.

The horizontal seismic loads have been inserted as surface loads applied to the
floor diaphragms at each storey. The seismic weights of each level were the only
vertical loads applied in the model because the gravity structural system was
designed by hand and did not require modelling.

The basic geometry is shown in Figure 29, and a three-dimensional view of the
analytical model with applied lateral loads is shown in Figure 30.

Figure 29. Wall geometry in numerical model.

Figure 30. Numerical models with semi-rigid and rigid diaphragms and lateral loads.

121
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The base shear distributions according to the numerical model for both rigid
and semi-rigid diaphragm stiffness are summarised in Table 20. These results
are compared to the values obtained using the analytical method. As can be
seen, the analytical model predicts forces relatively well when compared to the
numerical model with a rigid diaphragm. When using a more realistic diaphragm
stiffness, the force distribution changes slightly as expected.

Table 20. Base shear distribution on the walls in the y-direction for the ULS load case without accidental eccentricities.

Numerical
Numerical
Analytical Difference (semi-rigid
(rigid diaphragm) Difference
Wall ID diaphragm)

(%) (kN) (%) (kN) (%) (%) (kN) (%)

PY1 15% 32.1 15% 31.4 -2% 13% 27.3 -15%

PY2 13% 27.2 12% 25.8 -5% 11% 23.5 -14%

PY3 7% 15.6 6% 13.6 -13% 9% 18.6 19%

PY4 13% 27.5 13% 28.3 3% 17% 36.8 34%

PY5 12% 26.0 13% 28.1 8% 14% 29.2 12%

PY6 8% 17.3 8% 16.8 -3% 7% 16.0 -8%

PY7 5% 10.7 4% 9.4 -12% 7% 14.8 38%

PY8 27% 56.6 28% 58.5 3% 22% 47.9 -15%

Sum 100% 213.0 100% 213 100% 213

The differences between the analytical and numerical approaches can be


explained by the simplifying assumption in the analytical approach where the
wall stiffness is assumed to be proportional to length. The difference between
the methods is acceptable, considering all the assumptions and simplifications
used in the calculation of the different design parameters.

The obtained storey shear forces and overturning moments from each
wall can then be used to calculate the wall component demand and design
the individual elements in the same manner as when using the analytical
approach.

The deflections calculated in the numerical model are shown in Figure 31. Due
to the in-plane diaphragm flexibility, the walls are not all deforming by the
same amount.

The analytical and numerical wall deformations at level 4 are further


summarised in Table 21. Although differences in the individual wall
deformations can be observed, the average building deformation is captured
relatively well by the analytical model. Reasons for the difference in wall
deformation are that the numerical model also captures the deformation of
the walls due to the torsional nature of the building (even if small when loaded
in the y-direction) and that the diaphragms in the model are not modelled as
perfectly rigid and thus not having the same shear force distributions in the
walls. In the numerical model, the walls are tied together with a non-rigid
diaphragm, allowing a certain degree of differential movement. By using the
analytical model, no deformation restraint is applied as all walls are designed
independently.

122
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Figure 31. Deformation of walls PY1 to PY9 under SLS loading (blue lines denote the deformed wall shape).

Table 21. Base shear distribution on the walls in the y-direction.

Wall ID Analytical (mm) Numerical (mm) Difference (%)

PY1 32 37 16%

PY2 32 36 13%

PY3 41 47 15%

PY4 27 42 56%

PY5 24 32 33%

PY6 27 29 7%

PY7 38 44 16%

PY8 15 14 -7%

A natural frequency analysis with the numerical model provides a building


period of T1 = 0.87 s in the x-direction of loading, which, compared to the
analytical model with T1 = 0.80 s, is accurate. As mentioned above, the
differences are most likely due to the torsional behaviour of the structure, which
the analytical model is not able to capture. In the y-direction, the numerical
model provides a building period T1 of 0.6 s, which is identical with the
analytical period as calculated in section 10.2.7.

10.3 Diaphragm design


10.3.1 Diaphragm design: north-south direction (y-direction)

10.3.1.1 Determination of diaphragm floor demand

As discussed in section 4.2.2, the floor seismic demand based on the global
building overstrength needs to be determined. For this, the overstrength for
each wall at each level needs to be determined.

123
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The overstrength for wall Y1 can be determined as:

(10.85)

These values are summarised in Table 22 and are based on ɸ = 0.8, a nail
overstrength factor ɸ o, fastener = 1.6, a nominal nail capacity Qd,j of 0.9 kN for a
2.8 mm diameter nail and a k37 factor of 1.3.

Table 22. Wall overstrengths and input parameters for wall PY1.

Fastener shear
Fastener
Level
Shear demand
spacing
capacity at φO
v* (kN/m) overstrength (-)
S (mm)
vCPE (kN/m)

4 2.6 250 5.6 2.18

3 5.4 125 11.1 2.07

2 7.1 100 13.9 1.96

1 8.0 100 13.9 1.74

Average 1.99

The overstrength for all walls is summarised in Table 23.

Table 23. Wall overstrengths for all walls in the y-direction.

Level PY1 PY2 PY3 PY4 PY5 PY6 PY7 PY8

4 2.18 2.18 2.61 2.61 3.16 3.08 3.08 2.96

3 2.07 2.07 2.48 2.48 3.00 2.92 2.92 2.81

2 1.96 1.96 2.34 2.34 2.83 2.76 2.76 2.66

1 1.74 1.74 2.08 2.08 2.52 2.46 2.46 2.37

Average 1.99 1.99 2.38 2.38 2.88 2.80 2.80 2.70

This equates to an average global building overstrength in the y-direction of 2.49.

Table 24 shows the diaphragm forces as determined with the pESA method
(global building overstrength times ESA floor forces and PGA times floor weight)
and as determined with the ESA method based on the maximum overstrength
forced (elastic spectrum with Sp = 0.7). For the pESA method, a global building
overstrength of 2.49 and a PGA of 0.532 g has been used.

The force limitation from the maximum overstrength forces (elastic spectrum
Sp = 0.7) is used for this diaphragm design, as the sheathing fasteners are ductile
and a second level of strength hierarchy between ductile and brittle diaphragm
components will be used.

124
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

As can be seen from Table 24, the PGA component of the pESA method is
governing on levels 1–3, and only on the roof level can diaphragm forces be
reduced based on the maximum overstrength forces.

Table 24. Diaphragm storey forces in y-direction.

FpESA* FpESA* FESA, max overstrength* F D*


Level
ɸob FESA, i (kN) PGA Wi (kN) (kN) (kN)

4 169 85 136 136

3 187 187 151 187

2 116 173 93 173

1 58 173 46 173

All diaphragms will be designed for the maximum storey force on level 3 with
FD* = 187 kN. For simplicity, the floor is divided in two portions as shown in
Figure 32. This assumes that half the floor force is resisted by walls PY1, PY3,
PY5, PY6 and half of PY8. This is a simplification that allows the application of
the extended girder analogy. Alternatively, a finite element model can be used.

The diaphragm portion under consideration comprises approximately half of the


building area. Therefore, we will divide our floor demand by two:

(10.86)

PY5 PY6
PY1 PY3 interrupted chord PY8/2

4008
LIFT
BED 2 BED 3
LOBBY
6000

3000 3000
collector

chord strut STORE

BATH LAUNDRY
STUDY
SCULLERY
KITCHEN

chord
ENSUITE
14300
5200

2530 3922 3680


DRESSING
700

DINING

LOUNGE
BED 1

DECK

4180

POSSIBLE LOCATIONS OF BEAMS

Figure 32. Diaphragm portion considered for the analysis.

125
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

We do not know if our diaphragm is classified as rigid or flexible. Therefore, we


will use the upper and lower bound approach to form an envelope of forces.

10.3.1.2 Assuming a flexible diaphragm

Assuming a flexible diaphragm means that each wall will be loaded in accordance
with its tributary width, and we can apply the diaphragm force as a UDL:

(10.87)

The forces resisted by each wall are summarised in Table 25.

Table 25. Wall forces assuming a flexible diaphragm.

Wall ID Tributary width (m) Required force (kN)

PY1 2.10 13.0

PY3 4.00 26.0

PY5 3.20 21.3

PY6 3.00 19.9

PY8/2 1.90 11.8

10.3.1.3 Assuming a rigid diaphragm

Assuming a rigid diaphragm means we will need to apportion the wall loads in
accordance with their stiffness. As discussed in section 3.3.1, the stiffness of a
wall is generally governed by its shear deformation. Therefore, force distribution
will be done in accordance with the wall length, L. Although accidental
eccentricity does not need to be taken into consideration when using pESA
forces, the different location of the centre of mass and the centre of rigidity
influences the force distribution. The same approach as used to determine
the wall distribution earlier in the example is used to determine the forces
transferred from the diaphragm into the walls. The resulting forces are shown in
Table 26.

Table 26. Wall forces assuming a rigid diaphragm.

Wall ID Wall length (m) Required force (kN)

PY1 4.9 24.3

PY3 2.9 13.1

PY5 5.0 20.5

PY6 3.4 13.0

PY8/2 6.1 20.9

To determine the shear and moment demand in the diaphragm, a simple beam
model assuming both flexible and rigid diaphragms is used. Results for both the
flexible and the rigid floor assumption are shown in Figure 33.

126
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Flexible diaphragm

Rigid diaphragm

Figure 33. Shear force and moment distribution.

Based on the envelope (maximum) of the values from the two diaphragm
stiffness assumptions as shown in Figure 33, the unit shear forces in the
sheathing panels and the connection between the sheathing panels to the
collector beams/walls can be determined. For the latter, we have to know the
length of the collector beam or wall section that the shear is being transferred to.
In this example, we have collector beams running in the north-south direction,
each 6 m in length.

The maximum shear flow in the panels is calculated as the maximum shear force
divided by the depth of the diaphragm (also assumed to be 6 m):

(10.88)

127
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The maximum shear flow between the sheathing panels and the collector beams
is calculated as the maximum reaction force divided by collector length:

(10.89)

As the values are very similar, the maximum value will be used for all plywood
nailing.

The capacity of a single 2.8 mm nail in shear is:

(10.90)

A material reduction factor of 0.8 has been used, as the diaphragm demand is
based on the PGA component of the pESA method rather than an overstrength
component.

The nail spacing required is:

(10.91)

Therefore, space nails at 200 mm centre to centre. This nail spacing needs to be
used for all plywood sheathing edges and also to connect the plywood sheathing
to the collector and chord beams.

10.3.1.4 Design of capacity protected elements

All brittle elements in the diaphragm load path are designed by applying a 20%
force increase to the demand used to design the sheathing fasteners as discussed
in section 4.2.2.

Since the diaphragm design is governed by the PGA portion of the pESA, the
demand of any capacity protected element in the diaphragm should also be
increased by the inverse of the strength reduction factor ɸ of the sheathing
fastener (which is 0.8 for demands coming from PGA forces).

10.3.1.5 Design of plywood sheathing

The maximum shear flow in the sheathing panels is determined as follows

(10.92)

From our gravity load design, we know a 15 mm thick F8 ply sheet is required.
The shear strength per metre of this sheet is:

(10.93)

128
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

10.3.1.6 Design of tension and compression chords

It is normal to have chord beams that are the same cross-section as the floor joist.
From the gravity load design, we have 300 x 45 mm LVL13 joists at 600 mm centres.

The maximum moment demand in the diaphragm at overstrength is M* = 1.2/0.8


∙ 55 kNm = 83 kNm. With a distance between the two chords of 6 metres, this
results in an axial chord demand of 14 kN. This is the maximum chord demand
in the diaphragm system in tension and compression.

In compression:

(10.94)

In tension:

(10.95)

At the bottom edge of our nominated area, we do not have continuous support
for our chord beam, meaning that it will be subjected to combined moment and
compression demand.

Copying our moment verification from section 10.1.1 but with a reduced demand
due to the consideration of combined gravity and seismic loading (G + 0.3Q +E),
our moment capacity is:

(10.96)

Combined compression moment capacity ratio:

(10.97)

Combined tension moment capacity ratio:

(10.98)

Any splices in the chord members need to be designed to transfer the axial tension
demand, and this can typically be done with steel strapping or nail plates.

The chord member at the upper edge of the diaphragm portion is however
interrupted, and a strut beam needs to be introduced to transfer the moment
of M* = 1.2/0.8 ∙ 41.4 kNm = 62 kNm occurring in this area. Due to reduced
diaphragm depth of 6 m – 2.2 m = 3.8 m in the location of the floor opening,
the axial force in the chord and strut becomes 16.3 kN. This force can easily be
resisted by the chord beams and joist as shown in the design checks above.

129
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

The force in the strut needs to be transferred via the diaphragm into the
interrupted chord. The strut is extended by 3 m to either side of the opening. The
additional unit shear force in the diaphragm panels adjacent to the strut beams is:

(10.99)

The total unit shear force in the diaphragm around the opening hence becomes:

(10.100)

Additionally, the shear flow from the diaphragm reaction forces at the location of
the opening needs to be adjusted to reflect the reduced diaphragm depth:

(10.101)

Both unit shear forces can easily be resisted by the plywood sheathing. The nail
spacing in this area needs to be decreased to 100 mm:

(10.102)

In practice, the whole area highlighted in green in Figure 34 is nailed at 100 mm


centres, with the remaining diaphragm nailed at 200 mm.

The lower part of the diaphragm portion, which is resisted by walls PY2, PY4, PY7
and the remaining half of PY8, can be calculated in a similar manner. The
re-entrant corner due to the deck can be solved similarly to the staircase opening.

PY5 PY6
PY1 PY3 interrupted chord PY8/2

4008
LIFT
BED 2 BED 3
LOBBY
6000

3000 3000
collector

chord strut STORE

BATH LAUNDRY
STUDY
SCULLERY
KITCHEN

chord
ENSUITE
14300
5200

2530 3922 3680


DRESSING
700

DINING

LOUNGE
BED 1

DECK

4180

Figure 34. Diaphragm nailing – area in blue with nail spacing POSSIBLE
at 200 LOCATIONS
mm centres, area in green nail spacing at
OF BEAMS

100 mm centres.

130
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

10.3.2 Diaphragm design: east-west direction (x-direction)

The diaphragm in the east-west direction can be calculated similarly to the


diaphragm in the north-south direction.

10.3.3 Finite element analysis

Alternative to the girder analysis, the diaphragm demand can also be calculated
by using a finite element model. In order to do this, the membrane representing
the diaphragm needs to be modelled taking into account shear stiffness of the
sheathing panel and sheathing fasteners.

In this section, only the determination of the effective plywood stiffness is shown
without developing the full diaphragm analysis.

This requires us to define the fastener stiffness. According to NZS AS 1720.1, the
stiffness of a nail before yield is not linear. However, as discussed previously,
for our purposes, it is reasonable to assume that we can interpolate linearly.
Substituting the Code-defined yield deflection of δ = 0.5 mm, we can define the
following stiffness:

(10.103)

For the diaphragm area with 200 mm nail spacing, the effective shear stiffness of
the sheathing then becomes:

(10.104)

Depending on the software used for the finite element analysis, a stiffness
modification factor to be applied to a standard plywood membrane can be
used. This factor takes the additional flexibility from the plywood splices
into consideration if the diaphragm is modelled as a 15 mm thick plywood
membrane. The factor is calculated as:

(10.105)

10.4 Summary

This section has provided example methods for the design of a 4-storey
block of flats using LTF as the structural system. These methods are further
described in the previous sections of this guideline, and it is recommended that
the document be read in its entirety so that designers have a comprehensive
understanding of the recommendations made and the rationale for the design of
the timber systems and individual components comprising the building.

131
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

11
REFERENCES

Abrahamsen, R. (2018). Mjøstårnet – Construction of an 81 m tall timber building.


New Zealand Timber Design Journal, 26(1), 4–11.

Anderson, R. (2014). Sector review. Progressive Building, 105, 18–19.

APA. (2007). Diaphragms and shearwalls design/construction guide. Tacoma,


Washington, USA: APA – The Engineered Wood Association.

APA. (2011a). Engineered wood construction guide. Tacoma, Washington, USA:


APA – The Engineered Wood Association.

APA. (2011b). Design for combined shear and uplift from wind. APA System Report
SR-101B. Tacoma, Washington, USA: APA – The Engineered Wood Association.

APA. (2012). Building for high wind resistance in light-frame wood construction.
Tacoma, Washington, USA: APA – The Engineered Wood Association.

ATC. (1981). Guidelines for the design of horizontal wood diaphragms. Berkeley, CA,
USA: Applied Technology Council.

AWC. (2015). Special design provisions for wind and seismic. Leesburg, VA, USA:
American Wood Council.

AWC. (2017). Wood frame construction manual (WFCM) for one- and two-family
dwellings. Leesburg, VA, USA: American Wood Council.

AWC. (2018). National design specification for wood construction. Leesburg, VA,
USA: American Wood Council.

132
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Banks, W. (1999). Multi-storey timber construction – a feasibility study. New


Zealand Timber Design Journal, 8(2), 14–23.

Banks, W. (2007). Plywood shear walls – worked examples. New Zealand Timber
Design Journal, 15(2), 9–15.

Bano-Chapman, P. (2008). Fire resistance – attention to detail. Build, 104, 62–63.

Barber, D. & Gerard, R. (2014). High-rise timber buildings. Fire Protection


Engineering, 63, 11–20.

Beattie, G. (2001). Multistorey timber buildings manual. Wellington, New Zealand:


Carter Holt Harvey, Fletcher Challenge Forests Ltd, James Hardie Building
Products Ltd & Winstone Wallboards Ltd.

Blass, H. J. & Gorlacher, R. (2004). Compression perpendicular to the grain. World


Conference on Timber Engineering, Lahti, Finland, 14–17 June.

Bowden, J. (2008). Steel buckles, but is the timber industry ready to pounce?
Inwood, 81, 8–11.

BRANZ. (2018). Performance of mid-rise cladding systems. BRANZ Evaluation


Method EM7. Judgeford, New Zealand: BRANZ Ltd.

Buchanan, A. H. & Abu, A. (2017). Structural design for fire safety (2nd ed.). New
York, USA: Wiley.

Buchanan, A. H. (2007). Timber design guide. Wellington, New Zealand: New


Zealand Timber Industry Federation Inc.

Cathcart-Keays, A. Wooden skyscrapers could be the future of flat-pack cities


around the world. Retrieved from www.theguardian.com/cities/2014/oct/03/-sp-
wooden-skyscrapers-future-world-plyscrapers

CCMC. (1997). Development of design procedures for vibration controlled spans


using engineered wood members. Ottawa, Canada: Canadian Construction
Materials Centre.

CEAS. (2017). Indemnity matters No. 62. Wellington, New Zealand: AON New
Zealand.

Cobeen, K. E., Dolan, J. D., Thompson, D. & van de Lindt, J. (2014). Seismic design
of wood light-frame structural diaphragm systems: A guide for practicing engineers.
NEHRP Seismic Design Technical Brief No. 10. Gaithersburg, MD, USA: National
Institute of Standards and Technology.

CWC. (2009). Engineering guide for woodframe construction. CWC, Ottawa,


Ontario, Canada: Canadian Wood Council.

CWC. (2010). Wood design manual. Ottawa, Ontario, Canada: Canadian Wood
Council.

Deam, B. L. (1996). The seismic design and behaviour of multi-storey plywood


sheathed timber framed shearwalls (PhD thesis). University of Canterbury,
Christchurch, New Zealand.

133
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Dean, J. A., Moss, P. J. & Stewart, W. (1984). A design procedure for rectangular
openings in shearwalls and diaphragms. Pacific Timber Engineering Conference,
Auckland, New Zealand, May.

Diekmann, E. F. (1982). Design of wood diaphragms. Journal of Materials


Education, 8(1-2).

Dunn, M., Phillips, T., Emms, G., Stocchero, A., Mace, B., Kingan, M.,
Newcombe, M., Fea, P. & Fullbrook, D. (2018). Acoustical design of medium
density housing. BRANZ External Research Report ER30-A. Judgeford, New
Zealand: BRANZ Ltd.

Elliott, J. R. (1979). Wood diaphragm testing – past, present, planned. In


Proceedings of a workshop on design of horizontal wood diaphragms. Berkeley,
California: Applied Technology Council.

Emms, G., Gaunt, D., Stocchero, A., Banks, W., Dodd, G., Ballagh, K.,
Scheibmair, D., Chung, H., Mace, B. & Ng, I. (2015). Better acoustically performing
structural connections. BRANZ External Research Report ER1. Judgeford, New
Zealand: BRANZ Ltd.

FII & BSLC. (2014). Summary report: Survey of international tall wood buildings.
Vancouver, BC, Canada & USA: Forestry Innovation Investment & Binational
Softwood Lumber Council.

Fleischman, R. (2014). Seismic design methodology document for precast concrete


diaphragms. McLean, VA, USA: Charles Pankow Foundation.

Franke, S. & Quenneville, P. (2011). Compression strength perpendicular to


the grain of New Zealand radiata pine lumber. Australian Journal of Structural
Engineering, 12(1), 1–12.

Gardiner, D. R., Bull, D. & Carr, A. (2008). Internal forces of concrete floor
diaphragms in multi-storey buildings. New Zealand Society for Earthquake
Engineering Conference, Wairakei, New Zealand, 11–13 April.

Grantham, R. & Enjily, V. (2003). Multi-storey timber frame buildings: A design


guide. Watford, UK: Building Research Establishment (BRE).

Hamm, P. (2008). Schwingungsverhalten von decken bei auflagerung auf


unterzügen. Holzbau Quadriga, 1, 41–46.

Hamm, P., Richter, A. & Winter, S. (2010). Floor vibrations – new results. World
Conference on Timber Engineering, Riva del Garda, Italy, 20–24 June.

HPO. (2011). Building enclosure design guide – wood-frame multi-unit residential


buildings. Burnaby, BC, Canada: Homeowner Protection Office.

ICC & APA. (2012). A guide to the 2012 IRC wood wall bracing provisions. Country
Club Hills, IL, USA & Tacoma, WA, USA: International Code Council & APA – The
Engineered Wood Association.

ICC. (2015). 2015 International building code (IBC): Code and commentary: Volumes
1 and 2. Country Club Hills, IL, USA: International Code Council.

134
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

ICC. (2018). 2018 International residential code (IRC) for one- and two-family
dwellings. Country Club Hills, IL, USA: International Code Council.

Jephcott, D. K. & Dewdney, H. S. (1979). Analysis methods for horizontal


wood diaphragms. In Proceedings of a workshop on design of horizontal wood
diaphragms. Berkeley, California: Applied Technology Council.

Jones, D. & Brischke, C. (2018). Design life for wood and wood-based products.
High Wycombe, Buckinghamshire, UK: Exova BM TRADA.

Kessel, M. H. & Schönhoff, T. (2001). Entwicklung eines nachweisverfahrens für


scheiben auf der grundlage von Eurocode 5 und DIN 1052 neu (Development of
a design methodology for diaphragms based on Eurocode 5 and DIN1052:2008).
Braunschweig, Germany: Fraunhofer IRB Verlag.

Lavisci, P., Buchanan, A. & Burdon, N. (2018). New applications and challenges
for engineered timber structures. New Zealand Timber Design Journal, 26(4),
19–25.

Maclaren, P. (2010). The Nelson-Marlborough Institute of Technology’s glimpse


into the future of modern commercial buildings. New Zealand Journal of Forestry,
55(2), 3–4.

Malone, R. T. & Rice, R. W. (2012). The analysis of irregular shaped structures –


diaphragms and shear walls. Columbus, OH, USA: McGraw-Hill Education.

Milburn, J. & Banks, W. (2004). Six-level timber apartment building in a high


seismic zone. New Zealand Timber Design Journal, 12(3), 9–13.

Moroder, D., Smith, T., Pampanin, S. & Buchanan, A. H. (2015). An equivalent


truss method for the analysis of timber diaphragms. Pacific Conference on
Earthquake Engineering, Sydney, Australia, 6–8 November.

Mostafaei, H., Al-Chatti, Q., Popovski, M., Tesfamariam, S. & Bénichou, N. (2013).
Seismic performance of wood mid-rise structures. Research Report No. RR-345.
Ottawa, Ontario, Canada: National Research Council Canada.

Newcombe, M. & Batchelar, M. (2012). Seismic design of plasterboard wall


bracing systems. SESOC Journal, 25(2), 42–51.

Ni, C. & Popovski, M. (2015). Mid-rise wood-frame construction handbook. Special


Publication SP-57E. Pointe-Claire, QC, Canada: FPInnovations.

Norman, R. (2014). The value of wood. Progressive Building, 104.

Overton, G. (2018). Moving on up. Build, 165, 59–60.

Poirier, E., Moudgil, M., Fallahi, A., Staub-French, S. & Tannert, T. (2016). Design
and construction of a 53-meter-tall timber building at the University of British
Columbia. WCTE 2016 World Conference on Timber Engineering, Vienna,
Austria, 22–25 August.

Prion, H. G. L. (2003). Shear walls and diaphragms. In S. Thelandersson & H. J.


Larsen (Eds.), Timber engineering (pp. 383–408). Chichester, NY, USA: Wiley.

135
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

Rosevear, M. & Curtis, M. (2017). Physical characteristics of new houses 2015.


BRANZ Study Report SR367. Judgeford, New Zealand: BRANZ Ltd.

Shelton, R. (2004). Diaphragms for timber framed buildings. SESOC Journal,


17(1), 16–23.

Shelton, R. (2010). A wall bracing test and evaluation procedure. BRANZ Technical
Paper P21. Judgeford, New Zealand: BRANZ Ltd.

Shelton, R. (2013). Engineering basis of NZS 3604. Judgeford, New Zealand: BRANZ Ltd.

Smith, A. L., Hicks, S. J. & Devine, P. J. (2009). Design of floors for vibration: A new
approach. SCI Publication P354. Ascot, Berkshire, UK: Steel Construction Institute.

Smith, I. & Frangi, A. (2014). Use of timber in tall multi-storey buildings. Zurich,
Switzerland: International Association for Bridge and Structural Engineering
(IABSE).

Smith, P. C., Dowrick, D. J. & Dean, J. A. (1986). Horizontal timber diaphragms


for wind and earthquake resistance. Bulletin of the New Zealand Society for
Earthquake Engineering, 19(2), 135–142.

Spearpoint, M. (2008). Fire engineering design guide. Christchurch, New Zealand:


Centre for Advanced Engineering, University of Canterbury.

STA. (2014). Timber frame structures – separating distances during construction


for fire safety. Engineering Bulletin 6. Alloa, Scotland, UK: Structural Timber
Association.

STIC. (2012). Guidelines for fire safety on EXPAN construction sites. Christchurch,
New Zealand: Structural Timber Innovation Company.

Thompson, D. (2015). WoodWorks design example WW-005 – Five-storey wood-


frame structure over podium slab. Washington DC, USA: WoodWorks.

Thurston, S. (2003). Full-sized house cyclic racking test. BRANZ Study Report
SR119. Judgeford, New Zealand: BRANZ Ltd.

Timoshenko, S. P. & Gere, J. M. (2006). Mechanics of materials. New Delhi, India:


CBS Publishers & Distributors.

Tissell, J. R. & Elliott, J. R. (2004). Plywood diaphragms. Report 138. Tacoma, WA,
USA: APA – The Engineered Wood Association.

TRADA. (2011). Timber frame construction (5th ed.). Buckinghamshire, UK:


Timber Research and Development Association.

UKTFA. (2008). 16 steps to fire safety on timber frame construction sites. Alloa,
Scotland, UK: UK Timber Frame Association.

van Beerschoten, W., Carradine, D. & Palermo, A. (2013). Compressive strength


and stiffness of radiata pine laminated veneer lumber. European Journal of Wood
and Wood Products, 71, 795–804.

Walther, T. & Beattie, G. (2013). Optimised wall-to-floor junctions in multi-storey,

136
Multi-Storey Light Timber-Framed Buildings in New Zealand – Engineering Design

multi-residential light timber-framed buildings. BRANZ Study Report SR208.


Judgeford, New Zealand: BRANZ Ltd.

White, N. & Delichatsios, M. (2014). Fire hazards of exterior wall assemblies


containing combustible components. Quincy, MA, USA: Fire Protection Research
Foundation.

Wood Works! (2011). Mid-rise construction in British Columbia: A case study based
on the Remy Project in Richmond, BC. Ottawa, Ontario, Canada: Canadian Wood
Council.

137

Potrebbero piacerti anche