University of Wollongong Research Online

University of Wollongong Thesis Collection 2017+ University of Wollongong Thesis Collections

2017

Development of a sustainability approach for the structural design of buildings

Mehdi Robati University of Wollongong

Follow this and additional works at: https://ro.uow.edu.au/theses1

University of Wollongong Copyright Warning You may print or download ONE copy of this document for the purpose of your own research or study. The University does not authorise you to copy, communicate or otherwise make available electronically to any other person any copyright material contained on this site. You are reminded of the following: This work is copyright. Apart from any use permitted under the Copyright Act 1968, no part of this work may be reproduced by any process, nor may any other exclusive right be exercised, without the permission of the author. Copyright owners are entitled to take legal action against persons who infringe their copyright. A reproduction of material that is protected by copyright may be a copyright infringement. A court may impose penalties and award damages in relation to offences and infringements relating to copyright material. Higher penalties may apply, and higher damages may be awarded, for offences and infringements involving the conversion of material into digital or electronic form. Unless otherwise indicated, the views expressed in this thesis are those of the author and do not necessarily represent the views of the University of Wollongong.

Recommended Citation Robati, Mehdi, Development of a sustainability approach for the structural design of buildings, Doctor of Philosophy thesis, Sustainable Buildings Research Centre, University of Wollongong, 2017. https://ro.uow.edu.au/theses1/253

Research Online is the open access institutional repository for the University of Wollongong. For further information contact the UOW Library: [email protected]

DEVELOPMENT OF A SUSTAINABILITY APPROACH FOR THE STRUCTURAL

DESIGN OF BUILDINGS

A thesis submitted in fulfilment of the

requirements for the award of the degree

DOCTOR OF PHILOSOPHY

from

UNIVERSITY OF WOLLONGONG

by

MEHDI ROBATI

SUSTAINABLE BUILDINGS RESEARCH CENTRE

FACULTY OF ENGINEERING AND INFORMATION SCIENCE

2017 i

In this path (of knowledge), you will not reach your goal without

effort; Follow thy master (murshid), if you seek the reward.

Hafez (Ghazal 250), 14th century Persian poet

ii

Declaration

I, Mehdi Robati, declare that this thesis submitted in fulfilment of the requirements for the conferral of the degree of Doctor of Philosophy from the University of Wollongong, is wholly my own work unless otherwise referenced or acknowledged. This document has not been submitted for qualifications at any other academic institution.

Mehdi Robati

Jun 7th, 2017

iii

Abstract

Sustainability in the building industry means ensuring that a building is ecologically friendly and economically feasible, as well as providing a healthy internal atmosphere for the occupants. Recent developments in low CO2-e emissions design have highlighted the need to comprehend the characteristics and constraints of design alternatives at a global scale before making an appropriate choice. Despite the improvements in low CO2-e emissions design, the guidance currently available to structural engineers on how to incorporate whole of life CO2-e emissions impact in building design is still limited. This research seeks to identify the structural systems needed to sustain the long-term performance of a commercial building. To accomplish this goal, a typical 15 story office building in Australia was analysed to evaluate the potential impact of various forms of construction and structural over the building’s lifetime. This particular building is one of four benchmark buildings proposed by the National Standard Development

Organization. The effect of different and structural flooring systems on its overall life cycle costs and carbon emissions (CO2-e emissions) are quantified. This research adopted different life cycle assessment tools and databases to measure the energy consumed by this building from its construction to the day it no longer exists.

This research also assessed existing literature in the field of minimising the CO2-e emissions impact of concrete as a main structural material, and also quantified the CO2-e emissions impact and thermal performance of different concrete mixes. The results confirmed that embodied CO2-e emissions can be reduced significantly using supplementary cementitious materials. The thermal conductivity of concrete is strongly

iv

influenced by thermal properties of the concrete mixes and the proportions of its constituents. The results reveal there are many variations in embodied CO2-e emissions values across different inventory databases.

An uncertainty analysis was used to quantify the variations associated with the CO2-e emissions embodied in building materials and structural systems. The results reveal the contribution and variation of each type of construction material and structural systems in their whole life cycle, from the extraction of raw materials to the construction site and end of life building. The sources of uncertainty are the variations in the method of analysis used for each assessment, the different system boundaries, the sources of data, and quality of input used to calculate the upstream process.

A detailed energy simulation analysis via DesignBuilder was used to quantify the possible impact that construction forms and type of structural concrete might have on the energy consumed by the reference commercial office buildings across five Australian cities. The energy analysis revealed that the thermal capacity can be utilised to shift loads to reduce peak demand and reduce operational energy consumption if it is used as thermal energy storage.

The final part of this study is a proposed method to combine life cycle cost analysis with relative CO2-e emissions costs of a concrete structure during its lifetime. The results provide a quantitative value for evaluating the global CO2-e emissions impact made by different building structures in five Australian climate zones. The findings of this study also show that selecting an optimal structural design based on a single phase of life cycle assessment might not be the best way to choose ideal design alternatives.

v

Acknowledgments

Undertaking this PhD has been a truly life changing experience that would have been impossible without the support and guidance I received from many people; I am grateful to everyone who has support my journey towards this finished thesis. There are far too many to name but there are some I would like to especially thank.

First and foremost, I want to express my sincere gratitude to my main supervisor,

Professor Tim MacCarthy and Co-Supervisor Dr Georgios Kokogiannakis. Their patience, encouragement, and immense knowledge have been key motivations throughout my PhD, and their advice on research and on my career have been priceless; I thank them for allowing me to grow as a research scientist.

Thanks also to my friends and colleagues at the University of Wollongong for listening, and for offering advice support throughout this entire process. Special thanks to Ms Tina

Anderson at student support who trusted me enough and granted opportunities to be part of a team to enhance the studies for students who are disadvantaged and show signs of disabilities. I would like to thank A/Professor Alex Remennikov, A/Professor Neaz

Sheikh for offering me to work as part of their teaching team. Thanks to the Australian

Government Research Training Program for supporting this research study.

It is also a pleasure to thank my friends at Wollongong and all the other friends scattered around. Thank you for your thoughts, well-wishes, phone calls, e-mails, texts, and being there whenever I needed a friend.

vi

Last but not least, my deep and sincere gratitude to my family for their continuous and unparalleled love, help, and support. I am grateful to my sister Maryam for always being there as a friend. I am forever indebted to my parents Hassan and Zahra for giving me the opportunities and experiences that have made me who I am. They selflessly encouraged me to explore new directions in life and seek my own destiny. This journey would not have been possible if not for them, and I dedicate this milestone to them.

Mehdi Robati

vii

Table of Contents

Declaration ...... iii

Abstract ...... iv

Acknowledgments ...... vi

List of Figures ...... xv

List of Tables ...... xx

List of publications associated with this thesis ...... xxiii

Chapter 1 Context statement ...... 24

1.1 Research background and motivation ...... 24

1.2 Sustainability in the structural design of buildings ...... 24

1.2.1 Environmental impact ...... 25

1.2.2 Design efficiency ...... 31

1.2.3 Economical aspect ...... 33

1.3 Aims and objectives ...... 34

1.4 Research question ...... 35

1.5 Research method in brief ...... 36

1.6 Structure of thesis ...... 37

Chapter 2 Literature review ...... 40

2.1 Introduction ...... 40

2.2 Sustainability in structural design ...... 41

2.3 Sustainable structural design strategy ...... 43

2.4 Sustainable structural design parameters ...... 45

viii

2.5 Quantifying the sustainability in buildings ...... 45

2.5.1 Standardisation work on sustainable construction ...... 46

2.5.2 Environmental footprint schemes ...... 48

2.5.3 Energy rating tool ...... 51

2.5.4 Benchmarking process ...... 54

2.5.5 Life cycle sustainability analysis ...... 56

2.6 Life cycle cost assessment ...... 57

2.7 Environmental life cycle assessment ...... 57

2.8 Environmental Life cycle assessment in building sector ...... 61

2.8.1 Life cycle assessment for residential buildings ...... 64

2.8.2 Life cycle assessment for non- residential buildings ...... 65

2.8.3 Quantify environmental life cycle assessment in building ...... 66

2.8.4 The energy embodied in a building ...... 67

2.8.5 Calculating the embodied energy ...... 69

2.8.6 Operational energy ...... 73

2.9 Environmental impact of material selection ...... 75

2.9.1 Embodied CO2-e emissions ...... 76

2.9.2 Thermal performance ...... 78

2.10 Environmental LCA tools and databases ...... 80

2.10.1 ICE (Hammond et al. 2011) ...... 83

2.10.2 Envest 2 (Whitehead et al. 2014) ...... 83

2.10.3 Ecoinvent (Takano et al. 2014) ...... 83

2.10.4 Athena (Tharumarajah & Grant 2006) ...... 84

ix

2.10.5 BEES (Suh et al. 2014) ...... 84

2.10.6 ELCD (Finnveden et al. 2009) ...... 85

2.10.7 GaBi (Takano et al. 2014) ...... 85

2.10.8 AusLCI (Islam, Hamidul et al. 2015) ...... 85

2.10.9 eTool (eTool 2014) ...... 86

2.10.10 Crawford (Crawford 2011) ...... 87

2.11 Life cycle uncertainty analysis ...... 89

2.12 Structural design procedure: ...... 91

2.12.1 Structural system ...... 91

2.12.2 Floor system ...... 93

2.13 Summary ...... 96

Chapter 3 Research design ...... 99

3.1 Introduction ...... 99

3.2 Overview of this methodology ...... 101

3.3 Benchmark building ...... 101

3.4 Structural design parameters ...... 103

3.5 Structural analysis and design method ...... 105

3.6 Environmental assessment method ...... 108

3.7 Life cycle environmental assessment approach ...... 109

3.8 Uncertainty analysis of Life cycle embodied CO2-e emissions uncertainty

analysis ...... 113

3.9 Analysis of Life cycle cost ...... 113

3.10 Environmental impact cost analysis ...... 115 x

3.11 Summary ...... 117

Chapter 4 Incorporating an environmental evaluation and the thermal mix designs ...... 119

4.1 Introduction ...... 120

4.2 Low carbon concrete material ...... 120

4.3 Thermal performance of concrete ...... 122

4.4 Environmental aspects of concrete...... 123

4.5 Methodology ...... 128

4.5.1 Materials and mix designs ...... 128

4.5.2 Embodied carbon dioxide equivalent emissions ...... 131

4.6 Results and discussion ...... 133

4.6.1 Embodied emissions ...... 133

4.6.2 Variations in embodied CO2-e emissions coefficient ...... 136

4.6.3 Thermal conductivity of concrete mix design ...... 141

4.7 Conclusion ...... 145

Chapter 5 A method of uncertainty analysis for embodied CO2-e emissions impact of building materials in Australia ...... 149

5.1 Introduction ...... 150

5.2 The need for an embodied CO2-e emissions analysis in construction ...... 150

5.3 Environmental life cycle analysis ...... 152

5.4 Methodology ...... 156

5.5 Results and discussion ...... 165

xi

5.6 Conclusion ...... 171

Chapter 6 Impact of structural design solutions on the energy and thermal performance of an Australian office building ...... 173

6.1 Introduction ...... 174

6.2 The role of structural design in the energy performance of buildings ...... 175

6.3 Methodology ...... 179

6.3.1 Description of base building ...... 179

6.3.2 Structural design parameters ...... 181

6.3.3 Structural materials ...... 184

6.3.4 Operational energy analysis ...... 184

6.4 Results and discussion ...... 187

6.4.1 Structural analysis and design ...... 187

6.4.2 Energy performance of the building (Energy consumption) ...... 188

6.4.3 Analysis of thermal performance ...... 192

6.4.4 Design week-free floating analysis ...... 196

6.5 Conclusion ...... 202

Chapter 7 Integrated life cycle cost method for low CO2-e emissions structural design by focusing on a benchmark office building in Australia ...... 204

7.1 Introduction ...... 205

7.2 CO2-e emissions impacts and life cycle cost analysis ...... 205

7.3 Methodology ...... 208

7.3.1 Initial cost assessment method ...... 212

xii

7.3.2 Operation and maintenance costs assessment method ...... 214

7.3.3 Environmental costs estimation method ...... 215

7.3.4 Building design and construction ...... 216

7.3.5 Energy modelling ...... 218

7.4 Results and Discussion ...... 221

7.4.1 Lifetime environmental impacts ...... 221

7.4.2 Present value environmental costs of buildings ...... 223

7.4.3 Life cycle cost analysis ...... 225

7.4.4 Combined life cycle environmental and cost net present value ...... 227

7.5 Conclusion ...... 229

Chapter 8 Conclusion ...... 231

8.1 Introduction ...... 231

8.2 Review of aims and objectives ...... 232

8.3 Research Findings Pertinent to the Research Objectives ...... 234

8.4 Contributions of the Research Findings ...... 242

8.5 Recommendations for further research ...... 245

8.6 Summary ...... 247

References ...... 248

Appendix A: Mix properties of different batches of concrete...... 270

Appendix B: Detailed structural design ...... 272

Section B-1: Summary of a detailed structural design ...... 272

Section B-2: Manual calculation and verification for flat of slab ...... 274

xiii

Appendix C: Hourly air room temperature distributions ...... 298

Appendix D: Free-floating analysis during the design weeks ...... 303

Appendix E: Discomfort degree hours during the design weeks ...... 306

Appendix F: Life cycle CO2-e emissions for the analysed buildings ...... 309

xiv

List of Figures

Figure 1-1 Three pillars of sustainability ...... 25

Figure 1-2 Locations where Chapters with specific research objectives are linked ...... 39

Figure 2-1 summary of the key knowledge areas of this research ...... 41

Figure 2-2 European Standardisation suites for assessing the sustainability of buildings;

adopted from (Rossi 2014; Tsimplokoukou et al. 2014) ...... 47

Figure 2-3 Relationship between LCA, GHG Standards and PEF method (EC 2013;

Tsimplokoukou et al. 2014) ...... 50

Figure 2-4 Development of the OEF method based on International methodologies (Ding

2004; EC 2013) ...... 50

Figure 2-5 Building energy benchmarking process (Pérez-Lombard et al. 2009) ...... 54

Figure 2-6 Energy performance assessment scales using self-reference benchmarks

(Wang et al. 2012) ...... 55

Figure 2-7 Life cycle assessment stages (Cradle to Grave) ...... 58

Figure 2-8 Life cycle assessment framework (ISO14040 2006) ...... 59

Figure 2-9 Overall scheme of life cycle impact assessment(Bengtsson & Howard 2010)

...... 60

Figure 2-10 Building life cycle assessment stages (Cradle to Grave) defined by EN

15978 (De Wolf et al. 2017) ...... 63

Figure 2-11 Boundary of a Life cycle energy system, adopted from (Dixit et al. 2010) . 66

Figure 2-12 Passive design framework (Wang et al. 2007) ...... 74

xv

Figure 2-13 Offices (All), Electricity End Use Shares, 1999 – 2012 (Pitt&Sherry 2012)

...... 75

Figure 2-14 Offices (All), Share of Natural Gas End Use 1999 – 2012 (Pitt&Sherry

2012) ...... 75

Figure 2-15 eTool boundary system(eTool 2014) ...... 86

Figure 2-16 Interior structures (Ali & Moon 2007) ...... 92

Figure 2-17 Exterior structures (Ali & Moon 2007) ...... 93

Figure 2-18 floor systems (CCAA 2003; Warner et al. 2010) ...... 95

Figure 2-19 Interaction between key areas of knowledge reviewed ...... 97

Figure 3-1 Overview of the methodology in this study ...... 100

Figure 3-2 Plan of the benchmark building (Source: NS11401.1)...... 103

Figure 3-3 Front view of the benchmark building (Source: NS11401.1) ...... 103

Figure 3-4 Structural design parameters ...... 105

Figure 3-5 structural analysis & design flow ...... 106

Figure 3-6 Building Life cycle ...... 108

Figure 3-7 Life cycle cost of building ...... 114

Figure 3-8 Environmental cost estimation method ...... 116

Figure 3-9 Global assessment framework ...... 117

Figure 4-1 Damping and lag effect of thermal mass (Lemay & Leed 2011) ...... 122

Figure 4-2 Concrete embodied CO2-e emissions system diagram ...... 132

Figure 4-3 Embodied CO2-e emissions for different grades of Concrete ...... 133

Figure 4-4 Variation in embodied CO2-e emissions for 32 MPa and 40 MPa concrete

mixes ...... 134

xvi

Figure 4-5 Embodied CO2-e emissions for 32-35 MPa ...... 135

Figure 4-6 Embodied CO2-e emissions for 38-42 MPa concrete ...... 135

Figure 4-7 Embodied CO2-e emissions across inventory databases for 32 MPa concrete

...... 136

Figure 4-8 Embodied CO2-e emissions across inventory databases for 40 MPa concrete

...... 137

Figure 4-9 Variations of embodied CO2-e emissions for different databases (32-35 MPa)

...... 140

Figure 4-10 Variations of embodied CO2-e emissions for different databases (38-42

MPa) ...... 140

Figure 4-11 Thermal conductivity of concrete mix designs ...... 142

Figure 4-12 Variations in thermal conductivity between concrete mix designs ...... 143

Figure 4-13 Embodied CO2-e emissions versus thermal conductivity across all the

concrete mix designs ...... 143

Figure 4-14 Embodied CO2-e emissions against the thermal conductivity for Grade 32-

35 MPa and 38-42 MPa ...... 144

Figure 5-1 Boundary study of Chapter 5...... 157

Figure 5-2 the workflow and methodology used in this Chapter ...... 160

Figure 5-3 Embodied CO2-e emissions variations for two grades of concrete (N32 and

N40) ...... 165

Figure 5-4 Probability distributions of predicted embodied CO2-e emissions for the four

slabs ...... 167

Figure 5-5 Embodied CO2-e emissions associated with different structural forms ...... 168

xvii

Figure 5-6 Probability variation associated with the environmental impact (embodied

CO2-e emissions) of the materials used in the designed buildings...... 169

Figure 6-1 Relationship between thermal conductivity and density ...... 178

Figure 6-2 Plan of case study building (NS11401.1 2014) ...... 181

Figure 6-3 Section view of case study building (NS11401.1 2014) ...... 181

Figure 6-4 Structural analysis & design flow ...... 182

Figure 6-5 Summary of the annual heating and cooling degree-hours ...... 189

Figure 6-6 Predicted annual energy consumptions and national energy average usage

across five major climate zones ...... 190

Figure 6-7 Comparison between the annual energy requirements, structural forms and

construction materials of the office buildings ...... 191

Figure 6-8 Hourly air room temperature plotted against the hourly outdoor temperate for

the Waffle.low and 200.normal in the climate zone 2 (Brisbane)...... 193

Figure 6-9 Summer and Winter free floating temperature in climate zone 1 (Darwin) . 195

Figure 6-10 Summer and Winter free floating temperature in climate zone 2 (Brisbane)

...... 195

Figure 6-11 Summer and Winter free floating temperature in climate zone 5 (Sydney)

...... 195

Figure 6-12 Summer and Winter free floating temperature in climate zone 6

(Melbourne) ...... 196

Figure 6-13 Summer and Winter free floating temperature in climate zone 7 (Canberra)

...... 196

xviii

Figure 6-14 Probability of indoor air temperature during the summer and winter design

weeks for 200.normal and Waffle.low...... 197

Figure 6-15 Analysis of Winter design week free-floating for climate zones 1 (Darwin)

and 6 (Melbourne) ...... 198

Figure 6-16 Analysis of Summer design week free-floating for climate zones 2

(Brisbane) and 7 (Canberra) ...... 199

Figure 6-17 Discomfort degree hours during summer design week for climate zones 1

(Darwin) and 6 (Melbourne) ...... 201

Figure 7-1 Life cycle cost of building ...... 209

Figure 7-2 Environmental cost analysis method ...... 211

Figure 7-3 Global assessment framework ...... 212

Figure 7-4 Template of 15-storey commercial office building as visualised in

DesignBuilder ...... 217

2 Figure 7-5 Annual GHG (CO2-e emissions) normalised by net internal area (m ) ...... 222

Figure 7-6 Life cycle CO2-e emissions normalised by the gross floor area and separated

by the type of concrete and method of construction...... 223

Figure 7-7 Environmental costs of the buildings ...... 224

Figure 7-8 Initial capital costs (construction) of the building ...... 225

Figure 7-9 Present value of operational and replacement costs of the building ...... 227

xix

List of Tables

Table 2-1 Compares two proposed structural design strategies (adopted from

(Pongiglione & Calderini 2016)) ...... 44

Table 2-2 Analysis of sustainable reporting tools (CASBEE 2016; GBCA 2014; Kibert

2012; LEED 2012; Nguyen & Altan 2011; Y. J. Siew et al. 2013) ...... 51

Table 2-3 Overall score of four sustainable reporting tools (Nguyen & Altan 2011; Y. J.

Siew et al. 2013) ...... 52

Table 2-4 Advantage of a sustainable reporting tool (CASBEE 2016; GBCA 2014;

LEED 2012; LEED 2016; Michael 2013; Nguyen & Altan 2011) ...... 53

Table 2-5 Environmental impact assessment methods(Bengtsson & Howard 2010) ...... 61

Table 2-6 A review of life cycle analysis on concrete and concrete product (adopted

from (Hossain et al. 2018)) ...... 77

Table 2-7 Data sources for life cycle energy analysis ...... 81

Table 2-8 Summarises the basic information about each database and tools (adopted

from (De Wolf et al. 2017)) ...... 87

Table 3-1 Benchmark building features ...... 102

Table 3-2 Loading conditions for designing the building ...... 107

Table 3-3 Life cycle assessment procedure ...... 112

Table 4-1 Embodied CO2-e emissions for cementitious materials (Wimpenny 2009) .. 124

Table 4-2 Summary of concrete batches ...... 129

xx

Table 4-3 Final embodied CO2-e emissions coefficients ...... 132

Table 5-1 Uncertainty of physical parameters ...... 164

Table 6-1 Thermo-physical and structural properties of concrete classes as reported in

the literature ...... 177

Table 6-2 Overall specifications of the benchmark building ...... 180

Table 6-3 Loading conditions for design the building ...... 183

Table 6-4 Properties of selected concrete ...... 184

Table 6-5 Simulated assumptions for benchmark building ...... 185

Table 6-6 Physical properties of benchmark building ...... 186

Table 6-7 Summary of the structural design ...... 188

Table 6-8 Summer and Winter design weeks for the climate zones (DesignBuilder 2017)

...... 192

Table 6-9 Differences in the peak daily indoor air temperature between Waffle.low and

200.normal ...... 194

Table 6-10 Summary of discomfort degree hours during the design weeks ...... 200

Table 7-1 Summary of building materials and unit costs ...... 213

Table 7-2 Overall specification for the benchmark building ...... 217

Table 7-3 Summary of the structural design ...... 218

Table 7-4 Benchmark building physical properties ...... 219

Table 7-5 Benchmark building simulation assumptions ...... 220

Table 7-6 Whole life environmental and life cycle cost assessment of the building ..... 228

Table 8-8-1 Embodied CO2-e emissions associated with different structural forms ..... 237

xxi

Table 8-2 Variations in embodied CO2-e emissions of the materials used in the buildings

...... 238

xxii

List of publications associated with this thesis

The following papers were produced to disseminate the concept and outcomes of the study undertaken by the author during this PhD research. This thesis is formatted as a thesis by compilation.

Chapter 4 Robati, M, McCarthy, TJ & Kokogiannakis, G 2016, ‘Incorporating environmental evaluation and thermal properties of concrete mix designs’, Construction and Building Materials, vol. 128, pp. 422-35. Status: Published

Chapter 5 Robati, M, McCarthy, TJ & Kokogiannakis, G 2017, A method of uncertainty analysis for life cycle embodied CO2-e emissions of building materials in Australia, 28th Biennial National Conference of the Concrete Institute of Australia 2017, 22-25 October 2017. Status: Published

Chapter 6 Robati, M, Kokogiannakis, G & McCarthy, TJ 2017, ‘Impact of structural design solutions on the energy and thermal performance of an Australian office building’, Building and Environment, vol. 124, no. Supplement C, pp. 258-82. Status: Published

Chapter 7 Robati, M, McCarthy, TJ & Kokogiannakis, G 2017, ‘Integrated life cycle cost method for low CO2-e emissions structural design by focusing on a benchmark office building in Australia’, Energy and Buildings. Status: Accepted (Manuscript number: ENB_2017_2644)

xxiii

Chapter 1 Context statement

1.1 Research background and motivation

This study evaluates the potential effects of concrete structural systems on the lifetime performance of a building. This thesis comprises a series of papers that were published or accepted for publication to address the impact of CO2-e emissions associated with the structural design of buildings. Structural engineers aspiring to minimise the carbon embodied in a building must be able to understand how structural systems integrate with other building systems over the lifetime of the building. Structural engineers can influence the performance of a building from a sustainability perspective by considering how construction methods and structural materials will affect a life cycle analysis, while structural design can look into the future by understanding how structural materials are extracted, processed, transported, and installed. When considering the whole lifetime of a building a structural engineer must consider the impact the structural system will have on the building’s thermal performance and energy usage because at the end of life process, the structural design imposes constraints on how buildings are ultimately demolished, and components are removed, reused, or otherwise disposed.

1.2 Sustainability in the structural design of buildings

Sustainability involves the interaction of environmental, economic, and social factors (as shown in

Figure 1-1). With regards to the building industry, sustainability is ensuring that a building is ecologically friendly and economically feasible, as well as having a quality and healthy internal atmosphere for the occupants.

24

Figure 1-1 Three pillars of sustainability

In terms of achieving a sustainable structural design in a building, the structural engineer must recognise the characteristics and implementation constraints of the design alternatives at a global scale before choosing an appropriate one. This role gives structural engineers a particular responsibility to not only consider the structural effects and responses to loads, but also the sustainability aspects of structural design. Sustainable building design uses various methods to integrate the three sustainability aspects into design and practice (Ding 2008). The following sections provide a brief description on issues related to the environmental impact, design efficiency and economic aspect in the sustainable design process of the building.

1.2.1 Environmental impact

Building construction has a very important impact on the environment because all the processes of manufacturing and transporting materials as well as installing and constructing buildings consume

25

a lot of energy and emit large amounts of greenhouse gas (GHG) emissions. The carbon dioxide emissions from energy consumption in buildings accounts for most of the warming impact of current human greenhouse gas emissions (Pachauri et al. 2014).

The environmental impact of a building has temporary and long-term effects; temporary effects include initial construction activities such as noise and air pollution around the site, followed by the end of construction phase, whereas long-term environmental impact includes consumption of resources and production of waste through different phases of the building (construction, operation and end of life/demolition). Major environmental impact occurs during the lifetime of a building through the consumption of resource energy and the production of waste, all of which are explained below.

1. Resource consumption

Resource consumption is increasing as the demand for new buildings and upgrade of existing ones increases; globally the construction industry consumes 40% to 50% of the total mass of materials used (Storey 2008), and since concrete is the most widely used construction material, it consumes the second highest amount of natural resources (ISO15673 2005). The construction industry also generates large quantities of waste that is often incinerated or discarded in landfill sits. In Australia, the construction and demolition industry account for significant amount of waste generated and disposed of in landfill (Crawford 2011). The environmental impact due to resource consumption is attributed to different sources, because construction materials need water, energy for manufacturing, and transportation to deliver it to the site (Miller & Ip 2013).

26

2. Energy usage

Energy consumption attributed to buildings consists of the energy required for manufacture, supply, construct, and use during the useful life of a project, while the energy inputs needed to produce a building, from the extraction of raw materials, on site construction and final demolition and disposal represents the embodied energy of buildings (Hammond & Jones 2008). The operational energy of a building consists of the energy used for heating, cooling, hot water, ventilation, lighting and appliances, whereas the embodied energy represents a small fraction (13% to 19%) of its operational energy over a 50 year lifetime of the building (Dimoudi, A & Tompa, C

2008). However as operational energy decreases, the proportion of embodied energy in the total energy of building may gradually become a bigger issue; indeed studies in Sweden on buildings with the lowest energy consumption show that the embodied energy for a life time of 50 years may be 45% of the total energy (Thormark 2002).

The energy footprint consists of the fusil, nuclear, and renewable energy footprint, while the embodied carbon dioxide emissions footprint describes the emissions of carbon into the air as a result of a product or service (Radu et al. 2013). The carbon footprint describes the amount of greenhouse gas emissions (GHG) or CO2-e emissions produced during the life cycle of a product or system (Onat et al. 2014).

An appropriate choice of construction and building materials can possibly reduce the embodied energy and operational energy of a building. Materials with a low conductivity factors and high thermal resistances help to reduce energy consumption and associated greenhouse gas emissions

(Torgal & Jalali 2011). While dense materials such as concrete cannot meet these requirements, but if the concrete is thick enough it can improve thermal performances. Concrete and masonry

27

products can absorb and release heat at a rate which is almost the same as a building’s daily heating and cooling cycle stored heat (observing heat during the day and releases at night). A study where structural thermal performance was monitored in real time revealed that the thermal mass of a slab/wall caused a clear lag in the indoor and outdoor temperature change (Hajdukiewicz et al.

2015).

3. Waste production

The final stage of the built environment life cycle is the demolition and disassembling of a building.

Building construction and demolition produces waste through the life cycle of the building, from the construction phase, usage or maintenance phase, to the end of life or demolition phase. In fact in many countries, the waste produced from the construction and demolition area is one of the largest parts of the waste produced. In 2010, the Australian construction and demolition sector was

38% of the total waste generated, of which 33.9% was disposal (DEE 2013). Several regulations in Australia set strategic targets to minimise waste generation and maximise recovery; for example,

Queensland’s Waste Reduction & Recycling Strategy seeks to increase the rate of recycling to 80% by 2024 (DEHP 2014).

During the usage and maintenance phase, not a lot of waste is produced unless renovations take place (Tam & Tam 2006), but through the construction phase, the , concrete work, and masonry work typically accounts for 30%, 13%, and 13% of waste production (Poon et al. 2004).

The method of construction, project size, building type, technical problems, human errors and storage method are the main elements that influence waste generation in newly constructed buildings (Mokhtar et al. 2011).

28

The demolition phase produces large amounts of waste, and since most demolished structures end up as waste, the need to maximise the reuse and recycling of construction and demolition waste has been introduced as selective demolition (Kourmpanis et al. 2008). Selective demolition proposes to remove the building components in the inverse direction of its construction, and while this is more time consuming and expensive (Torgal & Jalali 2011), if the full potential of selective demolition is considered at the design stage it could result in many environmental, economic, and social benefits (Thormark 2007).

4. Environmental assessment tool

The environmental assessment tool provides a framework to assess and report environmental impacts in the building sector. This framework integrates three aspects of sustainability to evaluate the long-term performance of buildings. The sustainability reporting tools are categorised into the quality assessment, life cycle analysis, and energy consumption evaluation (Berardi 2012). The quality assessment system is a multi-criterion system which integrates the three aspects of the sustainability of buildings. This system is based on assessing criteria measured by several parameters and then comparing them across the real performance as a reference one (benchmark building).

Several other multi-criterion systems have been published to assess the sustainability of buildings and quantify the environmental impacts of human activities, as evidenced by the American

Leadership in Energy and Environment Design (LEED), UK’s Building Research Establishment

Environmental Assessment Methos (BREEAM), Green Star in Australia, and more recently the

Living Building Challenge (Ding 2004; Ding 2008; Kibert 2012). These systems are whole

29

building assessment systems where the frameworks are intended to iteratively assess the impact that a product, process, or service has on the environment (Cabeza et al. 2014). These assessments and rating tools are now being used to clarify the environmental impact of buildings even though they are criticised for lack of attention to the life cycle perspective and not understanding the total impact of decisions made during designing and operating the facilities (Whitehead et al. 2014).

In Australia, the National Standards Development Organisation (NSDO) is a non-profit company that developed a series of standards to specify the requirements needed to assess and declare the comparative environmental impact of building products and systems (NS11401.0 2014). NSDO proposed to use several benchmark buildings as a comparative tool to declare the environmental performance of building products and systems in buildings over their useful life (NS11401.0 2014).

This benchmarking tool can help designers compare design alternatives based on their potential impact (Rajagopalan et al. 2012).

To compare the impact of different design alternatives, a life cycle assessment method can be used.

A life cycle analysis assesses the materials used and the energy released by the system into the environment over the lifetime of a building; it is called “Cradle to Grave.” Several studies have pointed out the difficulty and uncertainties associated with applying whole life cycle assessment to the building sector due to different reasons— the long lifetime, size, and the intensive use of natural resources and inconsistencies inherent in estimating the environmental impact at every stage of a building (Cabeza et al. 2014; Ortiz et al. 2009; Sharma et al. 2011; Taborianski & Prado 2004).

30

1.2.2 Design efficiency

Another challenge for sustainable design is to improve the energy efficiency of buildings over their lifetime while reducing the environmental impact of the design. Research into sustainable structural design has attempted to address the impact that structural systems have on a building’s lifetime.

Several studies have shown the impact of structural systems on the environment by considering the amount of carbon emitted through pre-use, use, and the end of buildings life (DIIS 2013;

Pongiglione & Calderini 2016; Sarkisian et al. 2012; Webster 2004). Their studies showed that structural forms are responsible for 20% of the initial embodied energy and 10% of the whole building’s life cycle impact (DIIS 2013; Pongiglione & Calderini 2016; Sarkisian et al. 2012;

Webster 2004). Foraboschi et al. (2014) showed that flooring systems are responsible for most of the environmental impact when incorporated with other structural components (Foraboschi et al.

2014). Moreover their study of six construction systems indicated that type of floor and balance between thickness and the number of secondary beams directly affect the total energy embodied in buildings; this means that selecting an appropriate construction system can save significant amount of total energy embodied in a building (Foraboschi et al. 2014).

With regard to the type of structural material, choosing wisely helps to minimise the carbon emissions associate with building designs; however, selection of materials should address the specific geographical and economic context in which these technologies are applied. Cold-formed steel and thin walled aluminium profiles are commonly used in the construction industry due to their efficiency, lightweight, ease of erection, and low cost (Gilbert et al. 2014; Javed et al. 2017;

Mainey et al. 2015). Despite the advantage of cold-formed steel framing, there is little information available in Australian building design codes on fire ratings and sound transmission class ratings

31

of cold-formed steel in mid and high rise buildings (ABCB 2015; Hancock 2007). Currently, the cold-formed steel framing is limited to six stories height (Mujagic et al. 2012).

Our advanced understanding of timber buildings demonstrates the applicability of wood as a primary structural support system in mid-rise building construction (Robertson et al. 2012;

Skullestad et al. 2016), and the desire to design buildings with low carbon emissions has prompted several studies to consider timber as a part of the main structural materials (Abrahamsen & Malo

2014; Skullestad et al. 2016). Despite the advantages of a timber structure, our understanding of structural properties and safety of structural timber components is limited (Robertson et al. 2012;

Schmidt & Griffin 2013). Currently, the tallest timber building is a 14-storey residential building in Norway (Ramage et al. 2017).

In Australia, concrete is the most widely used construction material and concrete structures predominantly drive the Australian building industry (Kelly 2017). Concrete is a popular material due to its excellent mechanical and durability properties, and it is adaptable, relatively fire resistant, and generally available and affordable. Recent advances in concrete technology offer lower GHG emissions by using supplementary cementitious materials and recycled aggregates. There are improvements to thermal properties by using admixtures. Concrete as a building material can have high thermal properties because it absorbs and retains energy for a long period of time. In cold seasons, high thermal mass building elements that contain concrete such as walls and floor slabs, absorb radiant heat from the sun during the day and gradually release it back into the system (space) during the night when the outside temperature drops (Lemay & Leed 2011). The ongoing developments of novel construction materials such as Ultra-lightweight concrete (Huiskes et al.

2016; Roberz et al. 2017; Yu, QL et al. 2015) which offers lower thermal conductivity and density,

32

raise a question about their possible impact on the lifetime thermal performances. Another question is raised regarding the relationship between the life cycle environmental impact and thermal performance during the operational phase of a building. From a low CO2-e emissions design perspective, it is important to understand how to design a building structure with respect to its strength, the impact of CO2-e emissions, the energy performance and the thermal comfort over its lifetime. This question is often overlooked by structural engineers.

1.2.3 Economical aspect

This combination of a lifetime environmental impact and energy performance in economic terms provides a framework to compare and evaluate the sustainability of construction solutions. To evaluate environmental sustainability, a life cycle analysis must include the performance of a building from its manufacture to its disposal. This explains why the whole building system considers the materials, site activities, consumption of energy and resources, the thermal comfort and natural resources, as well as their interfaces with each other. This approach is encouraged as a cost-saving method or life cycle cost technique which allows more money to be invested into a new building system, even though it may be expensive on the first or overall cost basis (Kibert

2012). In fact many sustainable reporting systems do not mention integrating the economic aspects into their assessments (Y. J. Siew et al. 2013). The life cycle cost is a useful method in the design process to determine which system provides the maximum net saving when alternative choices provide the same performance requirements but varying with regards to the initial costs and operating costs (Fuller 2016; Harris et al. 2017). From a life cycle perspective, it appears that the economic driver is not only the initial cost, but also the operational, maintenance, and repair costs

33

are also important (Harris et al. 2017). The life cycle cost method attempts to establish a trade-off between the initial cost and long-term cost of alternative design solutions (Robinson et al. 2015).

The life cycle cost analysis is a useful tool, but there is a need to examine the integration framework to consider the environmental impact, design efficiency, and economical aspect in the decision making process.

In response to the concerns raised, this research program seeks to explore the potential impact of low CO2-e emissions structural design over the lifetime of a building. The scope of this PhD is to evaluate the effect of structural concrete and construction methods on the consumption of energy and resources over the lifespan of a building. It focuses on how low CO2-e emissions in structural design is achievable, and how it can directly relate to sustainability in the built environment.

1.3 Aims and objectives

The primary aim of this research is to investigate the impact of concrete structural systems on sustainability of buildings. It mainly focuses on the life cycle of a building by considering lifetime energy performance, carbon dioxide emissions and cost, i.e., cradle to grave.

This study investigates the potential CO2-e emissions impact associated with various types of concrete and construction systems. To achieve this end, a benchmarking method is used to compare and measure how structural design alternatives affect the lifetime impacts of a building; specifically, the life cycle CO2-e emissions impact associated with a typical office building in

Australia.

This research seeks to develop a framework that will include a life cycle assessment in the structural design of buildings. This framework will demonstrate how different design alternatives affect the

34

whole of life carbon emissions, energy consumption, materials used, and life cycle cost in a building. To achieve this goal, the research is classified into the following primary objectives:

1- Review existing literature in sustainability and then identify the main key strategies,

parameters, and tools associated with structural design;

2- Investigate the relative impact that the properties of selected concrete materials will have

on CO2-e emissions at the initial stage of structural design;

3- Investigate the effect of uncertainty on the impact of carbon emissions that each

construction material has on a typical high-rise office building;

4- Investigate the relative contribution of structural concrete and construction forms to the

lifetime energy demand and thermal performance of office buildings;

5- Develop a life cycle assessment model of several office buildings as benchmark buildings.

6- Propose a global assessment parameter as a result of whole life CO2-e emissions, and

structural costs during the lifetime of buildings.

1.4 Research question

The key questions for this research are:

➢ How is sustainability defined in structural design?

➢ How do we measure the effects of equivalent concrete structural system on the life

cycle of buildings?

➢ Can structure and structural design play a significant role in the life cycle of

buildings?

35

➢ What is the impact of structural materials and construction systems over the lifetime

environmental impact, the energy performance and cost of a building?

➢ Can life cycle assessment be applied to the structural design and construction

method?

1.5 Research method in brief

This research builds on the existing mix-assessment method to investigate the research questions from various perspectives. In the first stage, current literature is used to identify significant variables for the low carbon design of buildings; in the second stage, carbon emissions impacts associated with type of concrete as a structural material are analysed by collecting a number of concrete mix designs from published literature; in the third stage, a comparative method is used to compare the impact of carbon emissions associated with various construction forms and building materials by considering a typical concrete benchmark building in Australia. The results from this stage were then reprocessed through the uncertainty assessment method to quantify the variations associated with the carbon emissions from the study parameters. In the fourth stage, an energy analysis method is used to identify the potential impact of structural alternatives on the energy usage and thermal performance of the benchmark office building. This stage identifies the impact made by construction forms and the properties of structural concrete (Ultra-lightweight, normal weight). In the final stage, a framework is developed to describe the lifetime impact of CO2-e emissions associated with the structural design of a building in form of a single dimensional parameter which describes lifetime of CO2-e emissions and construction costs associated with

36

various design alternatives at the initial stage of the decision-making process. The contextual value of these conclusions are based on the finding from these multi-analysis approaches.

1.6 Structure of thesis

The outline of each Chapter is as below, and the location of the linking Chapters with specific research objectives is shown in Figure1-2.

Chapter 1 provides the background information pertaining to this study. It describes the main intention of this research by highlighting the research aims and objectives; it also provides research questions and an overview of the method used to conduct this research.

Chapter 2 reviews the literature and previous studies on sustainability and sustainable structural design of buildings. This Chapter highlights the sustainability aspects and the sustainability principles in building design. It also covers the common techniques and solutions for quantifying energy performance and minimising the carbon dioxide emissions from buildings. This Chapter points out the research gap in incorporating life cycle assessment tools with the structural design of buildings.

This is followed by a review of the sustainability reporting tools, life cycle sustainability assessment for energy consumption, life cycle assessment tools, and databases.

Chapter 3 presents the methodology used to integrate the life cycle assessment approach into the structural design of buildings. To illustrate the potential impact of CO2-e emissions on concrete structural systems, the benchmark building was used to compare the results of energy analysis and

CO2-e emissions based on different design alternatives.

37

Chapter 4 incorporates a carbon dioxide emissions evaluation (CO2-e emissions) and the thermal properties for various concrete mix designs. This Chapter also compares the impact of selecting a concrete mix design in terms of embodied carbon dioxide emissions (CO2-e emissions) with resulting thermal conductivity and density at the design stage of buildings.

Chapter 5 develops an uncertainty analysis model to quantify the variations associated with the calculation of CO2-e emissions. This Chapter uses a benchmark building to compare the uncertainty associated with the impact of CO2-e emissions in four different construction forms.

Chapter 6 contains the energy analysis and thermal performance of an Australian benchmark office building and estimates the impacts of selecting various structural concrete and construction forms.

Chapter 7 provides a framework to integrate the results of a life cycle assessment as a single parameter to enhance the decision-making process at the early stage of building design. This

Chapter compares the total life cycle CO2-e emissions and building costs associated with two different construction forms and types of concrete across five major cities in Australia.

Chapter 8 summarises the work presented in this thesis and outlines the contribution to the knowledge of this field; it also presents several recommendations for future research.

38

Figure 1-2 Locations where Chapters with specific research objectives are linked

39

Chapter 2

Literature review

2.1 Introduction

The previous Chapter introduced the main issues to consider with sustainable building design because sustainable buildings have now become a multi-disciplinary research investigation. From a structural design perspective, a concern is how to develop a decision-making process that will reduce the negative impacts on the natural environment. This means understanding the main strategies, parameters, and tools associated with sustainable building. This Chapter is a review of the available literature and a presentation of the fundamental concepts related to this research, including an identification of the factors which influence low carbon dioxide emissions structural elements, as well as the recognised methods and technologies used to reduce the CO2-e emissions of building structures such as sustainability reporting tools, protocols, and standards. The key knowledge areas of this research are shown in Figure 2-1.

40

Figure 2-1 summary of the key knowledge areas of this research

2.2 Sustainability in structural design

In terms of moving towards sustainability in structural design, a structural engineer must clearly understand the characteristics and constraints of the design alternatives at a global scale before choosing an appropriate one. This role gives structural engineers a particular responsibility to not only consider the structural effects and responses to loads, but also considers the environmental, economic, and social aspects of structural design while working to reduce the side impacts of the proposed design.

41

Research into sustainable structural design strives to address the impact that structural systems have on various aspects of sustainability, such as the environment, economy and society, which is why some researchers propose various principles of sustainable construction. In 1994, Kibert developed six principles for sustainable practices in civil engineering practices: minimise resource consumption, maximise resource use, use renewable/recycle resources, protect the natural environment, create a healthy, non-toxic environment, and pursue quality in creating the built environment (Kibert 1994). In 2004, Maydl pointed out the significance of a holistic approach for civil engineers to reduce the impact of building structures (Maydl 2004). Early studies attempted to quantify the effects of structural forms to the energy usage of buildings (DIIS 2013; Pongiglione

& Calderini 2016). Their studies showed that the structural forms are responsible for 20% of initial embodied energy (DIIS 2013). Webster (2004) studies on three buildings (located in Canada and the USA) revealed that the structural forms could account for 10% of the whole building life cycle impact. Several other studies argued that by moving towards minimising the operational energy, the environmental impacts associated with manufacturing and construction will become even more important (Pongiglione & Calderini 2016; Sarkisian et al. 2012; Webster 2004).

A number of studies highlighted the social and economic aspects of sustainability in structural design (Fernando et al. 2012; Kang & Kren 2007; Pongiglione & Calderini 2016). For instance,

Fernando et al. (2012) categorised the environmental, economic, and social impact of bridge structures. The environmental impact consists of embodied CO2-e emissions; the economic impact considers the costs associated with materials, equipment and property damage; the social impacts considers the risk of injury or death (Pongiglione & Calderini 2016).

42

Meanwhile, Alimoradi (2014) points out conflicts between some sustainable practices and some fundamental questions (Alimoradi 2014). Alimoradi’s questions introduced more general problems with sustainable structural design with concerns about prioritising the impacts to be covered, the strategies to be developed, and methods used to evaluate (Pongiglione & Calderini 2016).

Pongiglione and Calderini (2016) formulated these points into a single question: “What are the essential features and steps needed to achieve a sustainable structural design” (Pongiglione &

Calderini 2016). The following sections address this concern by focusing on sustainable design strategies and parameters as well as exploring the role structural engineers play in sustainable rating tools.

2.3 Sustainable structural design strategy

Sustainable structural design strategy is a method of selecting structural materials and forms to reduce their impact (Pongiglione & Calderini 2016). Since recognising the impact that building structures have on the environment, the economy, and society, several studies began to develop tangible design strategies to reduce this impact. Danatzko and Sezen (Danatzko & Sezen 2011) proposed several strategies for sustainable structural design as follows:

- Minimising material production energy by only selecting materials where the

manufacturing process allows for resource and energy saving. This method can be used to

conserve natural resources and reduce embodied emissions.

- Minimising embodied energy by providing for lower embodied energy and CO2-e

emissions.

- Minimising material use by reducing the quantity of materials used in a building structure.

43

- Life cycle assessment as a decision making tool to compare the impact associated with

different design alternatives.

- Maximising structural system reuse by using recovered structural materials and members.

Anderson and Silman (2009) also proposed a different classification to address the reduction of embodied CO2-e emissions associated with a building structure. Their classification includes (1)

Design for efficiency, (2) Design for materials, (3) Design for energy, (4) Design for adaptability and (5) Design for recycling (Anderson & Silman, 2009; Pongiglione & Calderini, 2016). A comparison of these two proposed classifications shows some overlap between them; Table 2-1 compares the two proposed structural design strategies.

Table 2-1 Compares two proposed structural design strategies (adopted from (Pongiglione & Calderini 2016))

Approach Strategy Danatzko and Sezen Anderson and Silman (2009) (Danatzko & Sezen 2011) Adaptability ---- Design for adaptability Maximising structural system Reuse ---- reuse Design for reuse and recycling ---- Design for recycling Low embodied energy and Minimising materials production Design for materials CO2-e emissions energy Minimising materials Minimising materials use Design for efficiency Minimising energy use Minimising embodied energy Design for energy Life cycle analysis/ Life cycle assessment ---- inventory/assessment

44

2.4 Sustainable structural design parameters

The sustainable structural design must have quantitative and accountable parameters; for instance, the parameters used to measure strategies to mitigate the impact of resources reduction are expected to be the usage of none renewable energy or raw materials, but not the impact of embodied CO2-e emissions (Pongiglione & Calderini 2016). Parameters which assess sustainable structural designs are still in progress (Hong et al. 2012). Traditionally, the structural design of buildings was limited to optimising cost and weight parameters (Moon 2008), but new approaches now provide frameworks to integrate the long-term impact of materials and systems into the design process. For instance, several studies pointed towards assessing the sustainability of structures based on their environmental impact (Hong et al. 2012; Moon 2008; Roaf et al. 2009), their embodied energy

(Crawford 2011; Foraboschi et al. 2014; Jiao et al. 2012; Yokoo et al. 2013), their embodied CO2- e emissions (Anderson & Silman 2009; Islam, H. et al. 2015; Oh et al. 2016; Yepes et al. 2015) and the use of a momentary element to address the environmental and structural characteristics during the lifetime of buildings (Tsimplokoukou et al. 2014).

2.5 Quantifying the sustainability in buildings

Measuring the environmental impact in the building could be a complicated task at the design and construction stage. However, building sustainability reporting tools could be used to build up a score of building performances against the sustainability criteria. Sustainable reporting tools are present industry standard guidelines which allow for comparisons across different projects (Y. J.

Siew et al. 2013). The building sustainability reporting tools are categorised as an assessment of

45

total quality, life cycle analysis and energy consumption (Berardi 2012). The following section reviews those tools presently available for evaluating and reporting sustainability in the building.

2.5.1 Standardisation work on sustainable construction

The International Organisation for Standardisation (ISO) and European Committee for

Standardisation (CEN) has attempted to define an objective for sustainable construction and provide a common framework to compare the results. The ISO technical committee provides several specifications and standards for the environmental and sustainable assessment of buildings.

These standards and specifications include a sustainability indicator (ISO21929-1 2016), an environmental declaration of products (ISO21930 2007), general principles (ISO15392 2008) and frameworks to assess the environmental performance of construction work (ISO21931-1 2010).

In parallel with ISO, CEN has published standardised approaches for measuring sustainability in construction. The CEN technical committee provides standardised methods for evaluating the sustainable aspects of new and existing construction activities as well as several standards for declaring the environmental products that form part of construction products. Figure 2-2 summarises the European standard publications for assessing the sustainability of buildings (Rossi

2014).

46

Figure 2-2 European Standardisation suites for assessing the sustainability of buildings; adopted from (Rossi 2014; Tsimplokoukou et al. 2014)

Similarly, European United provides regulations and policies to improve the environmental performance of buildings by including eco-friendly and energy efficient materials. Their publications are classified into the following parts:

- Energy Labelling Directive (energy related products)

- Energy Performance Building Directive (EPBD)

47

- Energy Performance Building Directive (energy related products)

- Energy efficiency action plan

- Ecolabelling Regulation

- Ecolabelling for Buildings

- Green Public Procurement

- Construction and Demolition Waste

- Resource efficiency road map

The resource efficiency report suggests that existing policies in the building industry, which primarily focus on energy efficiency, also need to integrate resource efficiency (EC 2014) because they consider the ecological impacts and resource usage during the lifetime of buildings (EC 2014).

2.5.2 Environmental footprint schemes

In line with ecological sustainability, various tools have emerged to address a unified footprint method for the environmental impact of production and consumption (Galli et al. 2012). The

European commission has presented two methods for assessing the environmental footprint associated with products and organisations (Galli et al. 2012; Tsimplokoukou et al. 2014). The environmental footprint of a product (PEF) applies to individual products or services, whereas the organisational environmental footprint (OEF) studies organisational activities as whole and considers all the activities associated with the product and services of organisations from the supply-chain (from extraction of raw materials through to the manufacturing process, usage, and end of life)(EC 2013).

48

The aim of these two publications (PEF and OEF) is to establish a common method to assess, display, and benchmark the environmental performance of materials, services, and companies in accordance with a comprehensive ecological assessment over the life cycle of a building (EC 2013)

The plans for an environmental footprint consist of energy and water consumption, embodied carbon emissions, and waste production (Tsimplokoukou et al. 2014). The environmental footprint represents a quantitative potential life cycle environmental impact of a specific product in the form of a single indicator (Čuček et al. 2012).

The energy footprint consists of the fusil, nuclear, and renewable energy footprint, while the embodied emissions footprint describes the emissions of carbon into the air as a result of a product or service (Radu et al. 2013). The carbon footprint describes the amount of GHG emissions or CO2- e emissions produced during the life cycle of a product or system (Onat et al. 2014). There are three main international standards to address the carbon footprint associated with products and services, and they are known as the Publicly Available Specifications (PAS) 2050 (PAS2050 2011), the

GHG Protocol Products Standard (Protocol 2011) and ISO 14067 (ISO14067 2013). Along with those three standards, ISO 14067 also aims to provide a standardisation line on the carbon label to enhance comparable and reliable carbon plans in the future (Wu et al. 2014).

In summary, Figure 2-3 and Figure 2-4 summarise the relationship between life cycle assessment, and the GHG standards PEF and OEF.

49

Figure 2-3 Relationship between LCA, GHG Standards and PEF method (EC 2013; Tsimplokoukou et al. 2014)

Figure 2-4 Development of the OEF method based on International methodologies (Ding 2004; EC 2013)

50

2.5.3 Energy rating tool

Several multi-criterion systems have been published to assess the sustainability of buildings and quantify the environmental impact of human activities, as evidenced by the American Leadership in Energy and Environment Design (LEED), UK’s Building Research Establishment

Environmental Assessment Methos (BREEAM), Green Star in Australia, and more recently, the

Living Building Challenge (Ding 2004; Ding 2008; Kibert 2012). These reporting tools review different phases of a project and provide guidelines for the design (as shown in Table 2-2). These tools use viewpoint and performance based criteria to evaluate existing and new construction buildings. Third parties are used to verify the final results of all the tools, as noted by Y. J. Siew et al. (2013).

Table 2-2 Analysis of sustainable reporting tools (CASBEE 2016; GBCA 2014; Kibert 2012; LEED 2012; Nguyen & Altan 2011; Y. J. Siew et al. 2013)

Item BREEAM LEED CASBE Green Star Design, Preliminary Design , Design, review and Stage of Project construction design, design, review and construction and operation and post design as built Offer training and × × × × certification Use viewpoint and × × × × performance based criteria Existence of different systems (existing building, × × × × new construction, ...) Provision of case study × × × × and manuals Third party verification × × × × Percentage of credits End product presentation Total Score BEE graph Total score achieved (%) Note: × shows the presents of an attribute

51

The final presentation of a project can be in the form of a graph (such as a BEE graph for CASBEE), the percentage of archived credit (such as BREEAM), and the total score (like LEED and Green

Star).

These four sustainable reporting tools for buildings are scored against criteria such as: Availability,

Popularity, Methodology, Applicability, Accuracy, Data collection process, User friendly,

Development, and Presentation. Table 2-3 compares the overall score for four sustainable reporting tools.

Table 2-3 Overall score of four sustainable reporting tools (Nguyen & Altan 2011; Y. J. Siew et al. 2013)

Attributes LEED BREEAM CASBEE Green Star Popularity (out of 10) 10 10 6 5 Availability (out of 10) 7 7 7 8 Methodology (out of 15) 10 11 13 9 Applicability (out of 20) 13 13 11.5 10 Data collection process (out of 10) 7 7 6 9 Accuracy (out of 10) 7 8 9 5 User friendly (out of 10) 10 8 6 8 Development (out of 10) 8 8 7 8 Presentation (out of 5) 3 3 4 3 Final Score (out of 100) 75 75 69.5 65

The advantage of different sustainable reporting tools is summarised in Table 2-4.

52

Table 2-4 Advantage of a sustainable reporting tool (CASBEE 2016; GBCA 2014; LEED 2012; LEED 2016; Michael 2013; Nguyen & Altan 2011)

Sustainable Advantage reporting tool • Criteria are publicity reviewed and arguably more transparent than BREEAM LEED • Credit is allocated for the heat island effect ( trees as shadows, high solar reflectance materials) • Credit is given for verifying the thermal comfort ( post occupancy) • Encourages energy reduction which leads to zero net CO -e emissions BREEAM 2 • sets a minimum requirement for sub metering energy use • High accuracy and adaptable methodology CASBEE • Graphical representation of final assessment which can be interpreted easily Green Star • Designed specifically for unique Australian conditions

Despite their advantages, these building sustainable reporting tools have been criticised for not considering the lifetime impacts of buildings (Y. J. Siew et al. 2013), and for not incorporating the financial aspects by assessing projects, as well as imposing non-scientific benchmarks (being too idealised) (Ding 2008; Mitchell 2010; Sharifi & Murayama 2013; Y. J. Siew et al. 2013).

The literature suggests several points to enhance building sustainable reporting tools

• Y. J. Siew et al. (2013) suggests expanding the list of criteria to include more measurable

economic and social issues.

• Y. J. Siew et al. (2013) suggests considering the life cycle effect of buildings based on

sustainable reporting tools.

• Mitchell (2010) suggests integrating building performance tools with the potential

environmental impact report in order to obtain a reliable result about the decisions made.

53

• Kestner et al. (2010) suggests applying some changes to allow more engineering disciplines

to contribute to the sustainability of projects.

2.5.4 Benchmarking process

Energy benchmarking uses data to measure the energy performance of a building over time and then compare the results across similar buildings. Building benchmarking can reveal the energy efficiency of a building by indicating its energy performance (Shahrestani et al. 2014).

Benchmarking aims to identify the right action where great savings can be made (Bosteels et al.

2010), which means evaluating the building energy efficiency by comparisons with standards or established energy benchmarks.

Pérez-Lombard et al. (2009) suggests four steps in the benchmarking process, as shown in Figure

2-5. Firstly, provide a database with information about energy performance, the type of building, and the size of a significant number of buildings which can be extracted from available literature, and from commercial and non-commercial databases. Secondly, evaluate the energy performance of the case study by collecting relevant information, and thirdly, quantify the quality of the case study by comparing the results of an energy performance analysis against the samples held in the databases. Lastly, the compression result shows the feasibility of energy strategy from both the technical and economic points of view (Pérez-Lombard et al. 2009).

Figure 2-5 Building energy benchmarking process (Pérez-Lombard et al. 2009)

54

There are two models where a database can be used to establish statistical benchmarks, the simple normalised model and the regression model. The normalised model uses an energy performance indicator such as the floor area where normalised energy uses intensity to account for one feature that affects the energy performance of a building (Wang et al. 2012), while the regression model considers the effects of other energy usage factors for the energy performance of the benchmark building (Wang et al. 2012).

In terms of benchmarking a certain building by simulation, implementing a self-reference building as the comparison criteria is preferable because a self-referencing building is presented as the total identical buildings in terms of geometrical dimensions, shape, functional layout (Wang et al. 2012).

This method has been adapted to evaluate the energy performance of both new and existing buildings across different environmental schemes and energy certifications. Generally, the performance of an asset building is determined by the percentage of annual energy reduction of a building in relation to a reference building. Figure 2-6 shows a different assessment scale for classifying the energy performance of a building.

Figure 2-6 Energy performance assessment scales using self-reference benchmarks (Wang et al. 2012)

55

2.5.5 Life cycle sustainability analysis

The demand in developing methods to improve understanding and address the impact and benefit of the product during its lifetime has led to in life cycle sustainability assessment (LCSA). The

LCSA has recently developed methods to evaluate the environmental, economic and social impact of product life cycles by integrating the three LCA tools, environmental life cycle assessments (E-

LCA), life cycle cost assessment (LCCA) and social life cycle assessment (S-LCA) (Finkbeiner et al. 2010). The general concept formulates (Equation 2-1) of LCSA are given as a summation of

ELCA, ELCC and SLCA (Kloepffer 2008; O’Brien et al. 1996).

LCSA = (E-LCA) + (LCCA) + (S-LCA) (2-1) LCSA = Life Cycle Sustainability Assessment E-LCA = Environmental Life Cycle Assessment E-LCC = Life Cycle Cost Assessment S-LCA = Social Life Cycle Assessment The Life Cycle Cost Assessment (LCCA) and Social Life Cycle Assessment (SLCA) follow

Environmental Life Cycle Assessment (ELCA) which is based on Environmental Life cycle

Assessment procedure, as given in the ISO 14040 series (Klöpffer & Grahl 2014). Development under the Environmental Life Cycle Assessment differs from the substantiality aspects because the environmental aspects can be covered by Environmental Life Cycle Assessment (Cabeza et al.

2014; Takano et al. 2014), while economic and social aspects still demand essential research progress (Finkbeiner et al. 2010; Kloepffer 2008). In the following section, a literature review has been carried out on the Life cycle cost and environmental assessment by focusing on the building and construction industry.

56

2.6 Life cycle cost assessment

Life cycle cost assessment defines all the costs associated with the lifetime of a building, including owning and operating a facility over a period of time (Hunkeler et al. 2008; Mearig et al. 1999).

For the building industry, a number of studies have tried to optimise the structural design of buildings by considering the economic and environmental aspects. These studies have proposed a conceptual framework (called “Eco- efficiency method”) to measure the sustainability of buildings by simultaneously selecting the optimal product design and considering its environmental and cost characteristics (Cha et al. 2008; Hahn et al. 2010; Ji et al. 2014; Saling et al. 2002). Others have proposed a method to evaluate the environmental impact and economics by converting embodied

CO2-e emissions into a momentary term (Gu et al. 2008; Hong et al. 2013; Itsubo & Inaba 2003; Ji et al. 2014; Kim, J et al. 2012). For example, several studies on structures and construction materials as products, have evaluated the environmental impact and cost effect over various stages of the life cycle (Chou & Yeh 2015; Huang et al. 2009; Silvestre et al. 2014). According to Cabeza et al. (2014), a combination of the life cycle cost analysis with the life cycle environmental impact provides a better understanding of the total impact of a proposed project or policy (Cabeza et al.

2014).

2.7 Environmental life cycle assessment

An Environmental Life Cycle Analysis (LCA) is an assessment to identify and evaluate the environmental aspects of a product throughout all the activities during its life (ISO14040 2006).

This method assesses the materials used and energy released by the system to the environment; it

57

is called “Cradle to Grave” and includes the extraction of raw materials, production processes, transportation, and the use and disposal depicted in Figure 2-7 (ISO14040 2006).

Figure 2-7 Life cycle assessment stages (Cradle to Grave)

The last version of ISO14040 (2006) defines four phases for any LCA, these phases are definitions of the goal and scope, analysis of the inventory, and impact assessments and interpretations (as shown in Figure 2-8). The definitions of the goal and scope state the purpose of the study and the boundaries of the system. The inventory analysis identifies and quantifies the amount of materials and energy inputs and outputs of a system, while the impact assessment characterises the inventory flows in the form of an environmental impact. This stage consists of three fundamentals that include the impact categories, classification of life cycle inventory, and modelling the category characterisation. Classifying the results of the life cycle inventory involves assigning the emissions, waste, and resources used to the impact categories chosen; the final result of the life cycle inventory assessment becomes the characterisation (indicator) result. This combination of environmental impact and scope of LCA study is summarised at the interpretation stage, while the last stage summarises the results for the conclusions, recommendation, and decision making (ISO14040

2006; Karimpour et al. 2014).

58

Figure 2-8 Life cycle assessment framework (ISO14040 2006)

Two methods are used to assess the life cycle impact, these include the damaged oriented methods

(end points) and the problem oriented methods (midpoints) (Bengtsson & Howard 2010). The problem oriented approach (Midpoint impact category), translates impacts into 15 environmental themes such as climate change, acidification, and human toxicity, etc., whilst the damaged oriented method (endpoints) classifies flow into different environmental themes which demonstrate the damage caused by each theme to the recourse, the natural environment, and human beings ( as shown in Figure 2-9) (Bengtsson & Howard 2010).

59

Figure 2-9 Overall scheme of life cycle impact assessment(Bengtsson & Howard 2010)

The problem oriented impact and the damaged oriented impact can be evaluated by different assessment methods, as summarised in Table 2-5.

60

Table 2-5 Environmental impact assessment methods(Bengtsson & Howard 2010)

Method Included Boundary Place Time frame • Global impacts: global warming, Midpoint AusLCI and resource depletion, and ozone depletion Related to which (some to ALCAS best • Regional impact: human and eco Australia indicator is endpoint) practice guide toxicity, land use, water use, soil considered

salinization, photochemical smog • Human health (as lack of illness) • Ecosystem health (potentially affected fraction of species) Western IMPACT2002+ • Resources (impact on human and eco Midpoint Infinite Europe system health) • Climate change (greenhouse gases emitted per person per year) • Human health (as lack of illness) • Ecosystem health (potentially affected Eco-indicator 99 fraction species) Endpoint Europe Around 200 years • Resources (impact on human and eco system health) • Human health (as broad sense, eg WHO) • Ecosystem production capacity EPS v2000 • Biodiversity Endpoint Global Infinite • Recreational values • Abiotic resources

2.8 Environmental Life cycle assessment in building sector

Applying a life cycle assessment to the building sector is unique in life cycle assessment practice

(Ortiz et al. 2009; Taborianski & Prado 2004), not only because of the complexity, size, and intensive use of natural resources across all stages of building, but also because the following factors make this sector exclusive in comparison to other complex products (Sharma et al. 2011).

61

Firstly, buildings often last for more than half a century so it is difficult to predict the whole lifetime impacts of a project from cradle to gate (Cabeza et al. 2014). Secondly, since most of the environmental impact of a project occurs during the operation phase, the initial stage of design and the selection of materials is critical in order to reduce those impacts (Kibert 2012). Thirdly, during the lifetime of a project the building may undergo many changes in terms of form and function, which can be more significant than the original product (Stephan & Crawford 2014), so they should be considered at the early stage of design to minimise the environmental effects of changes (Crawford 2011), and finally, there are many stakeholders and shareholders involved in the building industry.

The European standards Technical committee (CEN) TC350 has published a series of standards to define the cradle to grave environmental impact of buildings (Moncaster & Symons 2013) and construction works (De Wolf et al. 2017) (as shown in Figure 2-10).

62

Figure 2-10 Building life cycle assessment stages (Cradle to Grave) defined by EN 15978 (De Wolf et al. 2017)

Figure 2-10 shows how the different stages of the building life cycle assessment are organised. The product stage covers the supply of raw material (A1), the transportation of materials from extraction to manufacturing plant (A2), and the manufacturing process (A3). The construction process phase includes transport from the manufacturer’s gate to the construction site (A4), as well as the construction process (A5). The use phase covers the time from completion of the construction work to the end of life deconstructed /demolition stage. The use stage consists of the impact arising from the use of components (B1), maintenance (B2), repair (B3), replacement (B4), and refurbishment (B5). The operational stage energy (B6) and water (B7) includes the energy and water used by the building during its operational lifetime. The end of life phase consists of deconstruction and demolition (C1), transport from site to landfills or recycling facilities (C2), and waste processing (C3) and disposal (C4). Beyond the system boundary (cradle to grave), stage D

63

represents the benefits and loads of components for reuse, materials for recycling, and energy recovery for future use.

Based on EN 15978 (EN15978 2011), the data associated with life cycle assessment must be recent and be geographically coherent with the location for production (which is rarely the case due to lack of databases(Dixit et al. 2012)). Data are also needed to correspond to the system boundaries set for future assessment (De Wolf et al. 2017).

According to usage, buildings can be categorised as residential and non- residential; residential buildings are either single or multiple family houses and non- residential building are those used for multi-purposes such as public buildings, office buildings, and industrial and commercial buildings (Sharma et al. 2011).

2.8.1 Life cycle assessment for residential buildings

The result of a study the life cycle assessment of buildings shows that the occupancy phase of buildings is responsible for 70% to 90% of the total environmental impact (Asdrubali et al. 2013), while another study of high and low populated building indicated that the structures of buildings, windows, bricks and drywalls are in the main contributors of embodied energy and greenhouse emissions (Norman et al. 2006). A study on the demolition phase of buildings in Italy showed the advantage of waste recycling in sustainability of building in terms of the environment and energy

(Guggemos & Horvath 2005). Indeed the available literature reveals that the operational phase of buildings is responsible for 80% to 85 % of life cycle energy consumption of buildings (Richman et al. 2009; Sharma et al. 2011).

64

2.8.2 Life cycle assessment for non- residential buildings

Junnila and Horvath (2003) studied on environmental impacts of office buildings over a 50 year project lifetime (Junnila & Horvath 2003). They considered a building in Finland with one kWh/m2/year as a functional unit, and then considered three phases of life cycle assessment

(inventory analysis, impact assessment and interpretation of results) to define the most important aspects of building and building materials. The result of this study shows that electricity usage has the biggest impact because it is mainly used for heating, lighting, HVAC through manufacturing and maintenance of steel, concrete and paint, structure, and office waste (Junnila & Horvath 2003).

Scheuer et al. (2003) studied the life cycle performance of a six storey office building (7,300 m2) with a 75 year lifetime. Their results showed the initial energy intensity over the life of the building as 316 GJ/m2 and electricity with a HVAC system accounted for 94.4% of the life cycle energy consumption of the building.

Kofoworola and Gheewala (2009) evaluated the life cycle performance of a 38 story office building in Thailand. The function unit of this study was carried out as the gross floor area (60,000 m2) of the building over a 50 year lifetime. This study considered its whole of lifetime impacts (materials production, consumption, construction, operation, maintenance, demolition and disposal). The results showed that steel and concrete are the most significant materials in terms of energy use and the environmental impact at the manufacturing phase. The whole life cycle environmental impact was dominated by the operational stage, which was 71% of total photo oxidants, 66% of total acidification, and 52% of global warming.

65

2.8.3 Quantify environmental life cycle assessment in building

The environmental life cycle assessment method can be quantified by the relative energy sequestered and expended through different life cycle stages of a building. A life cycle energy analysis is an approach that assesses lifetime building energy inputs to the building systems. The life cycle energy assessment concept can be used to illustrate the lifetime benefits of a strategic design to minimise the operational energy and embodied energy of buildings (Cabeza et al. 2014).

Building life cycle energy system can classify the energy as embodied energy, operational energy and end of life embodied as shown in Figure 2-11.

Figure 2-11 Boundary of a Life cycle energy system, adopted from (Dixit et al. 2010)

66

2.8.4 The energy embodied in a building

There are different approaches to this definition due to the extraction of raw materials to the factory gate (A1-A3:cradle to gate), the extraction of raw materials to the site work (A1-A5:cradle to site), and the extraction of raw materials to the demolition and disposal phase (A1-C4:cradle to grave)

(Torgal & Jalali 2011).

In the first scenario (A1-A3: cradle to gate), the energy needed for transport and the energy executed in the work are both considered as the construction phase energy consumption of the building. In this case, the embodied energy represents 85% to 95% of the total energy of materials and the remaining 5% to 15% belong to the construction, maintenance, and demolition of the building. The last scenario (A1-C4:cradle to grave) shows the necessity of using local materials because of variations in embodied energy across different modes of transport (plane, Highway,

Railway, boat) (Berge 2009).

Although extensive studies have been carried out on embodied energy and carbon assessment, further studies are needed to address the significant of emissions that influence the accuracy of the life cycle assessment in buildings. The following section summarises the research undertaken to quantify embodied energy and CO2-e emissions associated with building and building materials.

5. Building scale

This review of literature considered variations in the embodied energy of materials across different types of buildings, regions, and sources of data; as an example, Langston (2008) evaluated the initial embodied energy of thirty completed buildings in Melbourne, Australia. These case studies are a random mix across a broad range of functional purposes, including office workplaces, health

67

facilities, residential and educational buildings, as well as commercial and hotel accommodation.

The results indicate that the initial embodied energy varies across different types of buildings, for instance the initial embodied energy of the case studied office buildings changed from 60,326 GJ to 121,541GJ as the total gross floor area increased from 2,543 to 4,704 square metres. In the other words, the average initial embodied energy of the office building is 27.77 GJ per m2 gross floor area (Langston & Langston 2008).

Ding’s (2004) research on twenty educational (high school) projects in New South Wales, Australia shows that the initial embodied energy across all the projects is 8.05 GJ/m2 (Ding 2004), while on the other side of the world, Chang (2012) shows the embodied energy for a 21 storey educational building is 6.3 GJ/m2 (Chang et al. 2012). Oka and Suzuki (1986) studied the embodied energy of office buildings in Japan by collecting the data from 10 buildings. The results of their study show that the energy required to construct 1 m2 of floor area varies from 6.5 to 13 GJ/m2 with an average value of 8.9g GJ/m2 (Suzuki & Oka 1998).

Treloar (1993) and Tucker (1993) showed that the embodied energy for an office building in

Australia vary from 8 GJ/m2 to 9GJ/m2 (cited by Treloar 1996b), while Cole (1996), Oka (1993), and Treloar (1993) showed that the embodied energy of commercial buildings vary 4.3 GJ/m2 to

19 GJ/m2 (cited by (Ding 2004)).

6. Building materials

In terms of construction materials, making the appropriate choices can reduce the embodied energy of a building by 17% (Thormark 2006), and reduce carbon dioxide emissions by almost 30% by avoiding 38 tonnes of carbon dioxide (González & García Navarro 2006). Dimoudi and Tompa

68

(2008) state that the embodied energy of an office building can represent from 13% to 19% of operational energy over a 50 year life of the project (Dimoudi, A. & Tompa, C. 2008). They also report the energy consumption and carbon dioxide emissions for two different office buildings, where the first building consumed 1.93 GJ/m2 and provided 198 kg CO2-e emissions /m2 and the second building consumed 3.97 GJ/m2 and purchased 289.4 kg CO2-e /m2. These authors also note that the embodied energy in structural materials (concrete and steel reinforcement) are the largest percentage of energy embodied in these buildings because the first building had 66.73% and the second building reached 59.57% of the total embodied energy (Dimoudi, A. & Tompa, C. 2008).

The parameters responsible for variations in embodied energy data can be summarised into discrepancies in the system boundaries (Dixit et al. 2012; Hammond & Jones 2010; Lenzen 2000;

Reap et al. 2008), methods of computing embodied energy (Joseph & Tretsiakova-McNally 2010;

Lenzen 2000; Optis & Wild 2010), the geographic location of the study (Dixit et al. 2012; Optis &

Wild 2010), the type of energy used (primary and delivered)(Dixit et al. 2010; Dixit et al. 2012), and the quality and source of data (Lenzen 2000; Menzies et al. 2007; Peereboom et al. 1998).

2.8.5 Calculating the embodied energy

Embodied energy is defined as the total energy required to construct a building, including the direct energy used in construction and assembly, as well as the indirect energy needed to manufacture the materials and building components. The energy content of materials refers to the energy used to extract raw materials, as well as manufacture and transport to the building site. The primary methods used to determine embodied energy are Input-Output economic based analysis, Process-

69

based analysis, and Hybrid analysis. These methods differ in how data about input energy is collected in the main materials production and administration process.

An Input-Output method of analysis could account for most of the direct energy and indirect energy by using economic data whereby money flows between different sectors of industry in the form of

Input and Output tables, as applied by a national government. In other word, this method is transcribing economic flow into energy flows by applying average energy tariffs. For instance,

Junnila and Horvath (2003) applied the Input-Output method to evaluate the life cycle of CO2-e emissions embodied in a 5-storey concrete building, while Cho et al. (2012) used it to determine the life cycle of CO2-e emissions associated with three steel buildings constructed from rigid-frame, braced-frame and outrigger structures, respectively. However, this method still suffers from problems such as differences in the economic data from country to country such as energy tariffs, product costs, and even how the sectors are grouped (Goggins et al. 2010). The Input-Output method can provide an industry wide environmental analysis but it cannot quantify a detailed and process specific analysis of CO2-e emissions from material production (Gan et al. 2017;

Kofoworola & Gheewala 2008).

The Process-based method considers building materials as the final products and then all the possible direct input energy works upstream of the main process and to deliver more accurate and reliable results. Process-based analysis is a widely used method because it leads to more accurate and reliable results (Crawford & Treloar 2003; Dixit et al. 2010), however, it is also deemed to be incomplete because some of the upstream processes are truncated or excluded due to the attributes needed to identify and quantify each product input, and small energy of the complex upstream process (Crawford & Treloar 2005). To calculate the energy embodied in a building, the quantities

70

of all the materials must be calculated and multiplied with the respective Process-based embodied energy intensities (Dixit 2017; Dixit et al. 2012; Rauf & Crawford 2015; Stephan & Stephan 2016).

Several studies used Process-based assessment methods in their analysis to determine the embodied

CO2-e emissions associated with materials such as wood, brick, masonry and low-medium strength concrete (Gan et al. 2017; Gustavsson et al. 2010; Scheuer et al. 2003). Nadoushani and

Akbarnezhad (2015) applied a process-based life cycle assessment to determine the embodied CO2- e emissions of various materials in several mid-rise buildings with rigid frame and shear walls.

Yan et al. (2010) also used a Process-based method to evaluate a 30-story concrete building by using the embodied CO2-e emissions coefficient for production, transportation, construction and waste disposal.

Because of the inherent problems with Input-Output analysis and Process-based, Hybrid methods of embodied energy analysis have been established. A Hybrid analysis unifies the benefits of two first methods to mitigate errors and limitations of Process-based and Input-Output analysis. This method of analysis commences with a Process-based analysis of accessible energy input data at the final production stage, and then shifts to the Input-Output base analysis when achieving reliable and consistent data are difficult due to the complexity of the upstream process. Hence, the Hybrid analysis method is categorised into Process-based Hybrid analysis and Input-Output Hybrid analysis (Suh et al. 2004; Treloar 1997).

Input-Output Hybrid analysis is based on the identification and extraction of a direct energy path from the Input-Output based method, as well as integrating Process-based data to prevent indirect effects. As Treloar (2003) states, Input-Output Hybrid analysis is considered as an appropriate method in the life cycle assessment of buildings (Langston & Langston 2008).

71

Process-based Hybrid analysis integrates Input-Output data to complex parts of upstream processes of material production to reduce the inherent problem with the Process-based method. The process based Hybrid model is an appropriate method for analysing large atypical products, if all the inputs are drawn far enough back, however, complex materials and overestimated prices of products could raise problems for this method (Dixit et al. 2010). A review of these methods indicates that they all have strengths and limitations, but the Hybrid analysis is the preferred approach for embodied energy due to its systemic completeness and use of reliable data, if available (Crawford 2011;

Crawford 2008; Lenzen & Crawford 2009; Majeau-Bettez et al. 2011; Rauf & Crawford 2015).

The system boundary of a Process-based Hybrid analysis or Hybrid material energy coefficients has the same limitations as a process analysis because many of the direct inputs to a process can be excluded (Lenzen & Dey 2000; Rauf & Crawford 2015). An Input-Out Hybrid analysis addresses the truncation issues associated with a process-based Hybrid analysis by using Input-Output data to fill in any gaps remained in the data (Rauf & Crawford 2015; Yu et al. 2017). An Input–Output based Hybrid analysis is more complete and less data and labour intensive (Rauf & Crawford 2015;

Yu et al. 2017). The study of a residential building and other household products shows that e process analysis values can be up to 87% lower than equivalent Input-Output based hybrid analysis values (Crawford 2008); so, analytical methods have been developed to estimate the embodied energy and CO2-e emissions in building products. Section 2.10 summarises the databases and tools available to determine the embodied energy and CO2-e emissions in building products.

72

2.8.6 Operational energy

The operational energy consists of the energy used to maintain comfortable conditions and daily maintenance; basically, it’s the energy used for heating, cooling, hot water, lighting, and for running appliances and operating equipment. Operational energy varies according to the range of comfort required, the climatic conditions, and the function of the building. Operational energy over the lifetime of a project is expressed as the annual operating energy multiplied by the lifetime of the project.

The operational energy is generally greater than the embodied energy over the lifetime of a building

(Torgal & Jalali 2011). Many studies suggest different passive and active technology to reduce the amount of operational energy (Kestner et al. 2010). With passive design, the amount and type of insulation in the external walls and roofs, the building’s orientation, the use of shading and glazing for windows, passive solar heating and thermal mass, contribute to improving the overall energy usage of a building (Kestner et al. 2010).

Wang et al. (2007) developed a framework which begins with climate data to state the design objectives, then defines the passive design alternatives based on building orientation, form, and the energy efficiency of building services (as shown in Figure 2-12).

73

Figure 2-12 Passive design framework (Wang et al. 2007)

Heating, cooling, ventilation, and the provision of hot water are generally responsible for most of the energy consumed in buildings (Karimpour et al. 2014; Pérez-Lombard et al. 2011). Space heating is often the highest share of the total operational energy; this amount of this energy can be calculated by subtracting the total heat loss from the free heat gain. The annual amount of energy consumed in office buildings varies between 100 to 1000 kWh per square metres depending on the type and use of office equipment, the climatic condition of the project, and the use of heating, ventilation, and air conditioning (HVAC) (Pitt&Sherry 2012). In Australia, a typical office building

74

from 1999 to 2012 shows that electricity is used for HVAC (43%), Lighting(26%) and equipment

(20%), as shown in Figure 2-13 (Pitt&Sherry 2012).

Figure 2-13 Offices (All), Electricity End Use Shares, 1999 – 2012 (Pitt&Sherry 2012)

Data on the amount of natural gas used from 1999 to 2012 shows that space heating accounts for the majority of gas consumed (as shown in Figure 2-14).

Figure 2-14 Offices (All), Share of Natural Gas End Use 1999 – 2012 (Pitt&Sherry 2012)

2.9 Environmental impact of material selection

The previous section shows the environmental impact of buildings by considering embodied energy and operation over the lifetime of a building. The embodied energy shows the amount of energy consumed to extract, refine, process, transport and fabricate a material or product (including

75

buildings), while the operational energy defines the energy consumed during the usage of a building. This section summarises the challenging section of the concrete as one of the most used construction materials, by focusing on the embodied carbon footprint (CO2-e emissions) and thermal properties.

2.9.1 Embodied CO2-e emissions

Concrete is the most widely used material in every type of construction due to its strength and durability. It is estimated that around 25 billion tons of concrete are produced globally each year.

Unlike other construction materials, concrete has a lower environmental impact because its production and its ingredients do not require a great deal of energy (Marinković et al. 2014). The amount of CO2-e emissions associated with the production of concrete and its relative impact is quite small, but due to the large consumption of natural resources and the waste generated, the overall negative environmental impact is quite significant. The production of as a key component of concrete has attracted much attention in life cycle assessments due to its energy- intensive production process which produces large amounts of CO2-e emissions (Shen et al. 2015;

Zhang et al. 2013). Due to the increasing sustainability concerns in the building and construction industry, concrete technology uses supplementary cementitious materials to replace the cement, that are generated as industrial waste or by-products; the most widely used are Granulated Blast

Furnace Slag (GBFS), , and (Abdalqader et al. 2016; Li et al. 2016). Hossain et al. (2018) summarised the impact of supplementary cementitious materials in concrete mix design and found that by selecting appropriate materials, the environmental impact can be reduced up to 38% compared to convention concrete (Hossain et al. 2018). A summary of the

76

methodological aspects associated with a life cycle analysis of concrete and concrete products is shown in Table 2-6.

Table 2-6 A review of life cycle analysis on concrete and concrete product (adopted from (Hossain et al. 2018))

Reviewed literature Reviewed Life cycle

Functional unit Functional Type of SCMs Type System boundary Application inventory data

Cou

Literature

Cradle Cradle Cradle

Database

ntry

Primary Primary

grave

gate data

site

- - -

to to to to

(Flower & Sanjayan FA, Concrete Australia 1 m3 * * * 2007) GBFS (O’Brien et al. 2009) FA Concrete Australia 1 m3 * * * * (Habert & Roussel FA, Concrete France 1 m3 * 2009) GBFS (Chowdhury et al. FA Road construction USA 1 kg * * * 2010) FA, Comparison of (Chen et al. 2010) France 1 kg * * * GBFS SCMs with cement FA;GBF (Habert et al. 2011) Concrete France 1 m3 * * * S: SF Czech (Hájek et al. 2011) SF Concrete 1 m3 * * * * Republic Comparison: SCMs (Van den Heede & De FA, & cement; different Belgium 1 kg * * * Belie 2012) GBFS LCA methods FA, South (Park et al. 2012) Concrete 1 m3 * * * * GBFS Korea FA;GBF (Crossin 2012) Concrete blend Australia 1 m3 * * * * S;SF FA, (Proske et al. 2013) Concrete Germany 1 m3 * * * GBFS (Knoeri et al. 2013) FA Concrete Switzerland 1 m3 * * * * (De Schepper et al. FA Concrete Belgium 1 m3 * * * 2014) FA, (Randl et al. 2014) Concrete Austria 1 m3 * * GBFS (Van den Heede & De FA Concrete Belgium 1 m3 * * * Belie 2014) (Zhang et al. 2014) FA Concrete Hong Kong 1 m3 * * * * (Blankendaal et al. FA, the Concrete 1 m3 * * * * 2014) GBFS Netherland (Seto 2015) FA Concrete Canada 1 m3 * * * (Liu et al. 2015) FA Concrete China 1 m3 * * (Turk et al. 2015) FA Concrete Slovenia 1 m3 * * * * (Saade et al. 2015) GBFS Cement Brazil 1 t * * * (Anastasiou et al. FA; Concrete Greece 1 km * * * 2015) GBFS (Celik et al. 2015) FA, LP Concrete USA 1 m3 * *

(Crossin 2015) GBFS Concrete Australia 1 m3 * * * *

77

FA, (Berndt 2015) Concrete Australia 1 m3 * * * * GBFS FA, (Lawania et al. 2015) Concrete wall Australia 1 m3 * * * * * GBFS (Li et al. 2016) GBFS Cement China 1 kg * * * * Concrete paving (Hossain et al. 2016) FA Hong Kong 1 t * * * * eco-blocks (Pushkar & Verbitsky FA; Blended cement Israel 1 m3 * * * 2016) GBFS (Teixeira et al. 2016) FA Concrete Portugal 1 m3 * * * * FA;GBF (Robati et al. 2016) Concrete Australia 1 m3 * * S: SF (Dossche et al. 2016) FA Concrete Belgium 1 m3 * * * * (Yang et al. 2017) FA Cement China 1 t * * * (Hossain et al. 2016) FA Cement Hong Kong 1 t * * * * FA; (Miller et al. 2016) Concrete USA 1 m3 * * GBFS FA;GBF South (Kim & Chae 2016) Concrete 1 m3 * * * S Korea FA;GBF (Pineda et al. 2017) Mortars Spain 1 kg * * * * S: SF (Mohammadi & South FA;GBF Concrete Australia 1 m3 * * * 2017) S: SF GBFS: (Panesar et al. 2017) Concrete Canada 1 m3 * * * * SF FA;GBF South (Kim & Tae 2016) Concrete 1 m3 * * * S Korea FA;GBF (Tait & Cheung 2016) Concrete UK 1 m3 * S FA= Fly ash; GBFS=Granulated blast furnace slag; SF= Silica fume; LP=Limestone powder

2.9.2 Thermal performance

Another advantage of concrete is its high thermal mass which affects the thermal performance of a building. Constructions with a high thermal mass usually contain heavyweight materials such as concrete (normal weight) and typically result in indoor environments with relatively small variations in temperature (Zhu et al. 2009). The thermal mass is directly related to the mass of a building which is stored in floors, walls, ceilings and partitions, and the thermal storage capacity is affected by the surface area, thickness of the building’s component, and is also a function of the material used in the building (Hensen & Lamberts 2012). The primary factors affecting the thermal properties of the building materials are thermal conductivity, specific heat capacity, and the density of the material (Hensen & Lamberts 2012).

78

With concrete and concrete products, the thermal properties are influenced by the amount of cement paste, the aggregate and the presence of moisture (ACI122R 2014). Damdelen et al. (2014) experiments have shown that the GGBS content in concrete mixes provides the highest thermal admittance while concrete mixes with silica fumes have a similar thermal admittance value as

Portland cement (Damdelen et al. 2014). Robati et al. (2016) also show that the type of aggregate affects the density and thermal conductivity, while replacing normal aggregate with lightweight aggregate reduces the density and thermal conductivity of concrete. Specific heat/heat capacity is another indicator of thermal inertia, but as highlighted in the ACI122R, the specific heat does not change across concrete with different densities (from 1280 to 2240 kg/m3) (ACI122R 2014).

Recent developments in concrete admixtures have led to the production of a novel Ultra- lightweight concrete (1,350 to 1,900 kg/m3) with a much lower thermal conductivity than normal concrete (2,200 to 2,600 kg/m3) (Marinkovic et al. 2010; Wu et al. 2015; Yun et al. 2013; Zhang

& Poon 2015). Roberz et al. (2017) state that buildings with a high thermal mass require more time to heat up or cool down, which might lead to thermal discomfort and increase the heating and cooling load. However, there is some concern about higher sensitivity as the temperature of concrete fluctuates with a lower thermal conductivity (Ponechal & Staffenova 2017; Robati et al.

2017; Roberz et al. 2017). In response, Al-Sanea and Zedan (2011) point out that an adequate amount of thermal mass and types of materials are associated with the case study building, and this depends on many interrelated factors such as weather conditions, the occupancy pattern, internal heat gains, HVAC type, building orientation, fenestration and the shading system (Al-Sanea &

Zedan 2011). The following section summarises the literature reviewed on the environmental life cycle assessment tools and databases.

79

2.10 Environmental LCA tools and databases

The sources of data for life cycle analysis can be classified in accordance with the different phases of manufacture, operation, and demolition. The possible source of data for the manufacturing phase of building materials can be found by considering the economic input and output table, process based or hybrid based, or even from literature by considering the quantities of materials used, the bills, and the average distance material is transported (Whitehead et al. 2014).

Evaluating the energy performance of buildings depends on factors related to local climate conditions. So, a possible source of data for the operational phase can be extracted from annual electricity usage and fuel consumption for heating purposes, household survey and energy simulation software (Cabeza et al. 2014). The total amount of energy used in a building over its lifespan can be estimated by a summation of all the energy consumed during all phases, some of which have been carried out to compare the features and capability of an energy simulation program

(Crawley et al. 2008; Foucquier et al. 2013; Nguyen et al. 2014). Harish and Kumar (2016) reviewed the most used modelling methodologies that were developed and adopted to model the energy systems of buildings (Harish & Kumar 2016). EnergyPlus is a proven building performance simulation program tool in common practice in the Australian context (Asadi et al. 2012; Daly

2015; Ryan & Sanquist 2012; Yalcintas 2008), while DesignBuilder is a third-party graphical user interface for EnergyPlus that has been used in many studies in Australia (Chowdhury et al. 2008; Daly 2015;

Rahman et al. 2010; Rahman et al. 2011).The source data for the demolition phase can be stated as operating the equipment and the average distance material is transported over the demolition time

(Langston & Langston 2008).

80

Finally, a life cycle analysis estimates the impact that a building makes on the environment

(Adalberth 1997). Table 2-7 summarises the data sources for life cycle energy analysis of a building at different phases.

Table 2-7 Data sources for life cycle energy analysis

Life cycle phase Activity Source of Data

• Manufacturing energy data of the building materials from literature, economic process Building material analysis, input/output tables, hybrid analysis production • Quantities estimated from building drawings, bill

of materials and from interviews with building • Designer, contractor/owner (a)Manufacturing • Average distances for material transport Transport • Energy data for transport operations

Building construction • Energy use from site visit including refurbishment • Simulation software (EnergyPlus, VisualDOE, e- Use of electricity and fuels QUEST, DesignBuilder, ENORM, TRNSYS, (b) Use for heating, sanitary, water Ecotect, SunCode, etc., annual electricity bills, and lighting house hold survey on energy use, Inventory data for fuel production, Electricity mix data. • Demolition operations and quantities from specific Building demolition measured data • Use of equipment and explosives from database • Average distances for material transport (C) Demolition Transport • Energy data for transport operations

Recycling • Specific measurement data

Total energy use and Life cycle energy embodied CO2-e emissions and embodied CO2- • Phase (a+b+C) of the building in its life e emissions cycle Life cycle material and energy flow estimation • Phase (a+b+C)

Life cycle • Greenhouse effect or global warming, ozone assessment Impact assessment that depletion, acidification, eutrophication, building makes on the photochemical smog, etc. estimated using environment software: SimaPro, ECOBAT, LEGEP, BEES, ATHENA, etc.

81

For the purpose of life cycle analysis, there are various inventory databases and software tools which are classified as product assessment tools, whole of building decision support tools and whole of building assessment system framework. Within the product assessment tool, the building products are considered to be the smallest part of any analysis. These tools compare and evaluate building products across each other, and as such, provide a platform such as Building for

Environmental and Economic Sustainability (BEES)(Cabeza et al. 2014).

A building assembly tool is a group of interdependent components that create a specific system in a building. This means that all of the materials and elements needed to build a wall system can be evaluated by a building assembly tool that will analyse the effects that each alternative will have on the environment (Cabeza et al. 2014). There is a wider scope of tools in the product assembly, such as ATHENA eco calculator tools (Cabeza et al. 2014).

Whole of building asset tools evaluate the environmental effects of a building as a unique system by considering all the systems and assemblies. These tools can compare the design alternatives which are very useful in early stage of a project design. For example, the ATHENA Impact

Estimator considers the whole building as an input by considering the building geometry and assemblies. The final results of these tools are amassed for the whole building and presented as the environmental impact due to variety of life cycle phases and the specific impact made by the building (Cabeza et al. 2014). The following sections provide a brief description of the inventory databases available.

82

2.10.1 ICE (Hammond et al. 2011)

In England, Hammond and Jones developed a database to determine the embodied energy and embodied carbon of a large number of building materials (Hammond et al. 2011). This database includes data that fulfils the ISO standard requirements and it has been used to release the Inventory of Carbon and Energy (ICE) which covers the boundary for each material, from the extraction of raw materials, the process and manufacture of product to dispatch at the company “cradle to gate”

(Hammond et al. 2011; Hammond & Jones 2008; Lamnatou et al. 2014).

2.10.2 Envest 2 (Whitehead et al. 2014)

Envest 2 was established by the Building Research Establishment (BRE) in the UK to allow users to access the environmental impact as well as a comparison across each other by considering whole life costing of construction materials and the energy consumed during the operational phase of buildings. Although this tool considers the lifetime of project by including the energy consumed in heating, cooling, and daily operating, it does not have enough data to cover the embodied energy of building components (Watson & Jones 2005; Whitehead et al. 2014).

2.10.3 Ecoinvent (Takano et al. 2014)

In Switzerland, the Ecoinvent database was developed in late 2000 to supply a life cycle inventory data that includes all the economic activities at a unit process level (Takano et al. 2014). The last version of this Ecoinvent database (version 3.3) considers datasets at the unit process level where they can be linked to different system models. The philosophy behind this method is that the primary production of materials is only assigned to primary materials, while products are classified into one of three categories (ordinary by-product, recyclable material, and waste) which are based

83

on the level of products, not on individual activities (Ecoinvent 2013; Frischknecht et al. 2005;

Frischknecht et al. 2007).

2.10.4 Athena (Tharumarajah & Grant 2006)

In the USA, the Athena life cycle inventory database was created by the National Renewable

Energy Laboratory and the Athena institute. This database provides individual gate-to-gate, cradle to gate, and cradle to grave boundaries to account for the amount of energy and materials flowing into and out of the environment. The resulting boundaries are associated with activities for producing materials as well as manufacturing and assembling products within the United States

(Rebitzer et al. 2004; Tharumarajah & Grant 2006).

2.10.5 BEES (Suh et al. 2014)

The national Institution of Standard and Technology developed the BEES software to measure the environmental performance of building products using a life cycle assessment. This software provides a product to product comparison in terms of environmental and economic performance

(Suh et al. 2014). BEES only includes materials which have significant cost, energy, or weight.

This tool covers 280 building products and aims to assist in the selection of potentially green and cost effective products (Cabeza et al. 2014; Suh et al. 2014). BEES is primary useful for analysing products during the manufacturing process because it does have some limitations when used to analyse building processes and building materials such as the use and maintenance of insulation

(Forsberg & von Malmborg 2004; Haapio & Viitaniemi 2008).

84

2.10.6 ELCD (Finnveden et al. 2009)

The European reference Life Cycle Database (ELCD) consists of Life Cycle Inventory (LCI) data from front-running EU-level business associations and additional sources for key materials, energy carriers, transport, and waste management (ELCD 2007; Finnveden et al. 2009).

2.10.7 GaBi (Takano et al. 2014)

The GaBi database is a product system assessment tool that first appeared in 1992 from PE

International as a German company. This database provides information on the life cycle inventory for commercial purposes and also compares principal life cycle inventory datasets collected from industries, associations and public sections worldwide (Takano et al. 2014). The GaBi tool uses an integrated product database developed through technical literature and industry review. This database has been used as a reference database in forms of life cycle assessment calculation in the context of the German building certification system (Takano et al. 2014). However, the use of phase impacts have not yet been addressed adequately (Cabeza et al. 2013).

2.10.8 AusLCI (Islam, Hamidul et al. 2015)

The Australian National Life Cycle Inventory Database (AusLCI) is a major initiative currently being delivered by the Australian Life Cycle Assessment Society (Islam, Hamidul et al. 2015). The development of AusLCI was requested by stakeholders from industry, government, and academia in order to deliver a methodology for applying ISO 14040 measurements in Australia. AusLCI assists with providing LCA for a whole life of product as well as benchmarks for eco labelling.

The results of AusLCI data cover all the main impact areas by considering the use of land, the reduction of water, eutrophication, human toxicity, eco toxicity, ozone depletion, photo-chemical

85

ozone formation and global warming (Rebitzer et al. 2004; Renouf et al. 2015; Tharumarajah &

Grant 2006).

2.10.9 eTool (eTool 2014)

This online tool was developed in Australia as life cycle assessment tool called eTool; it uses a life cycle analytical method to include the energy embodied over the total design life of a building.

This eTool complies with ISO 14044, ISO14040 and EN 15978 (Iyer-Raniga et al. 2014), and the default materials and energy are based on an Australian life cycle inventory database. It also covers embodied energy from cradle to gate by considering all the energy used to extract the raw materials and process them into useable building products. The eTool life cycle boundary is shown in Figure

2-15 (eTool 2014).

Figure 2-15 eTool boundary system(eTool 2014)

86

2.10.10Crawford (Crawford 2011)

In terms of the inventory energy database in Australia, Treloar and Crawford (2010) proposed an embodied energy coefficient for selected building materials by considering all the production processes from the extraction of raw materials, transporting to the company, and from manufacturing and finishing the material (cradle to gate). Treloar and Crawford (2010) used an energy based Input-Output model developed by Professor Manfred Lenzen at the University of

Sydney, as well as Australian process data compiled by Grant (2002). The energy used in construction, operation, maintenance, refurbishment, demolition, and reuse and recycling is excluded from this study (Crawford 2011; Crawford 2008; Stephan 2013).

Table 2-8 is a summary of the national and international inventory databases, industry data report, and available software.

Table 2-8 Summarises the basic information about each database and tools (adopted from (De Wolf et al. 2017))

EEC ECC LCA Method Boundaries Region Free Industry data reports Inventory of Carbon and Energy ✓ ✓ Literature Cradle to gate UK ✓ (ICE) Structure and Carbon (Carbon ✓ Engineering Cradle to gate US ✓ working group) Economic I/O Hutchins UK Building Blackbook ✓ Cradle to gate UK LCA WBCSD on cement ✓ Manufacturing Cradle to gate World ✓ NRMCA on concrete ✓ ✓ Manufacturing Cradle to gate US World Steel ✓ ✓ Manufacturing Cradle to gate World CORRIM on timber ✓ ✓ Manufacturing Cradle to gate US Software and tools Carbon Calculator Environmental Economic I/O ✓ Cradle to gate UK ✓ Agency LCA Economic I/O BEES ✓ ✓ ✓ Cradle to gate US ✓ LCA Athena Sustainable Materials Cradle to ✓ ✓ Process LCA N. America ✓ (North America) gate/grave

87

Cradle to CCaLC Tool ✓ ✓ Process LCA UK ✓ gate/grave Cradle to GaBi ✓ ✓ Process LCA Germany grave GEMIS ✓ ✓ Process LCA Cradle to gate Germany LEGEP Software GmbH ✓ ✓ Process LCA Cradle to gate Germany LTE OGIP ✓ ✓ Process LCA Cradle to gate Germany Economic Cradle to Sankey Editor ✓ ✓ Germany I/O* LCA grave Cradle to Umberto ✓ ✓ Process LCA Germany grave Cradle to SimaPro & OpenLCA ✓ ✓ Process LCA Netherlands grave Cradle to OpenLCA ✓ ✓ Process LCA Netherlands ✓ grave Cradle to EQUER and novaEQUER ✓ ✓ Process LCA France grave Qantis suite ✓ ✓ Process LCA Cradle to gate France Cradle to Eco-Bat 2.1 ✓ ✓ Process LCA Switzerland grave Cradle to Bousted Model ✓ ✓ Process LCA UK grave Cradle to eTool ✓ ✓ Process LCA Australia ✓ grave Databases European Life Cycle Database ✓ ✓ EPD Cradle to gate Europe ✓ (ELCD) US LCI ✓ ✓ EPD Cradle to gate US ✓ Quartz ✓ ✓ Literature Cradle to gate US ✓ IVL Swedish Environmental ✓ ✓ EPD Cradle to gate Sweden ✓ Research Institute EcoInvent ✓ ✓ LCIA Cradle to gate Switzerland Oekobaudat.de (German National ✓ ✓ EPD Cradle to gate Germany ✓ Database) Milieudatabase.nl (Dutch National ✓ ✓ EPD Cradle to gate Netherlands ✓ Database) INIES (French National Database) ✓ ✓ EPD Cradle to gate France ✓ IVAM ✓ EPD Cradle to gate Netherlands EPD database BBRI (Belgian ✓ ✓ EPD Cradle to gate Belgium National Database) Australian Life Cycle Inventory ✓ ✓ EPD Cradle to gate Australia ✓ (AusLCI) Building Product Life Cycle ✓ ✓ EPD Cradle to gate Australia ✓ Inventory (BPLCI) Crawford ✓ Hybrid I/O Cradle to gate Australia ✓ New Zealand building materials ✓ EPD Cradle to gate NZ ✓ embodied energy EEC: Embodied Energy Coefficients; ECC: Embodied Carbon Coefficients; LCA: Life Cycle Assessments; Environmental Product Declaration (EPD); Life Cycle Inventory Analysis (LCIA); Input/Output (I/O)

88

2.11 Life cycle uncertainty analysis

The use of sustainable building materials to minimise the environmental impact and to improve the energy performance has been increasingly at the centre of attention. A comprehensive life cycle approach must consider how material section impacts the entire life of building, including the extraction of raw materials, manufacturing processes, transportation to the construction site, construction processes, the operational phase, and the end of life recycling and potential for reuse

(Ding 2014). However, applying life cycle procedures to buildings may not be straightforward because they can present difficulties and uncertainties at different stages of a building’s life.

In the operational stage building, energy models have been used to predict the average energy consumption of a building (Faggianelli et al. 2017). Quantifying the uncertainties for operational energy has been the topic of a large number of studies (Coakley et al. 2014; Daly et al. 2014;

Eisenhower et al. 2012; Jain et al. 2015; Tian 2013); so, this will not be discussed any further in this thesis. However, little is known about uncertainties associated with the selection of building materials over the life cycle of embodied CO2-e emissions of buildings. Several studies have revealed the growing significance of embodied emissions and have shown its relationship to the lifecycle carbon emissions of buildings (Ayaz & Yang 2009; Dixit et al. 2010; Ibn-Mohammed et al. 2013; Lee & White 2008).

Uncertainty analysis is a useful tool to quantify the risk associated with variations in input parameters and their influence on the overall life cycle environmental impacts (Hong et al. 2016).

Several studies state that the lack of comprehensive production records and differences in manufacturing process result in many variations in the CO2-e emissions coefficient and the embodied energy coefficient factor (Ozoemena et al. 2017; Robati et al. 2016). For instance, Robati

89

et al. (2016) identifies some inconsistencies in the calculation of embodied CO2-e emissions across the different databases. This is attributed to variations in embodied CO2-e emissions coefficients and a lack of in-depth consideration of the detailed properties of each individual concrete mix design (Robati et al. 2016).

Several studies proposed stochastic modelling and interval calculation as a reliable and accepted method for uncertainty analysis in life cycle assessment (Ozoemena et al. 2017; Wang & Shen

2013). For instance, Blengini and Di Carlo (2010) used Monte Carlo analysis to understand the reliability of life cycle assessment for a case study building. They considered the uncertainties associated with the quantity of materials (production and maintenance), transport distance, and energy consumption for heating, cooling, and domestic hot water and cooking. A Monte Carlo analysis using behavioural modelling of data uncertainty through probability distributions can be used to define the optimal solutions for a project. Another study acknowledged the impacts of uncertainties in materials and the life span of buildings over the life cycle performance of buildings

(Aktas & Bilec 2012). Aktas and Bilec (2012) showed that selecting a random service life for building materials and systems leads to a significant variation in the results of a life cycle assessment (Aktas & Bilec 2012). Also, Leung et al. (2015) showed an ongoing trend to improve methods for predicting uncertainties and modelling (Leung et al. 2015). Their study is based on

134 journal articles of uncertainty analysis methods associated with the environmental impact of buildings (Leung et al. 2015). Despite all the previous studies, there is still a lack of work aimed at quantifying uncertainties associated with structural design and construction forms.

90

2.12 Structural design procedure:

Concrete flooring systems provide a designer with a variety of alternative floor systems for a specific building. AS3600 (2009) provides the minimum criteria for the design of concrete structures in Australia. The structural systems of mid to high-rise building are based on considerations that regularly address the efficiency of structural systems in terms of unit structural materials (Ali & Moon 2007; Cho et al. 2004); however, the unit structural quantity cannot reflect the complexities of mid to high rise buildings because as different structural materials are integrated, their costs and variations with the speed of construction must be considered too.

2.12.1 Structural system

Ali and Moon (2007) classified the structural systems of tall buildings into interior and exterior structures and this is based on how the components for the primary lateral load resisting system will be distributed in the building. The interior is assigned to the building when the main lateral loading resisting system is located inside, and an exterior structure is given to the building when the main part of lateral loading resisting system is located at the primate of the building. Interior structures are suggested for buildings up to 20 storeys and they have two main structural components: moment resisting frames and shear trusses/ shear walls, while exterior structures are for buildings of up to 160 stories and they typically consist of a tube in the core of the structures.

Figures 2-16 and 2-17 show the concept of each system diagrammatically.

91

Figure 2-16 Interior structures (Ali & Moon 2007)

92

Figure 2-17 Exterior structures (Ali & Moon 2007)

2.12.2 Floor system

A multi-disciplinary approach to selecting building components has been investigated. For instance, Takano et al. (2014) demonstrated the impact of selected building materials on embodied

CO2-e emissions and on the building cost in a Finish context, as the authors studied three building component categories— the structural frame, the inner components (insulation and sheathing), and the surface components (exterior cladding and flooring). They found that the materials selected influenced the embodied CO2-e emissions and cost of the building (Takano et al. 2014). The selection of a concrete structural system is based on several factors specified by Australian Building

Codes Board (ABCB 2015).

93

The ABCB (2015) provides a simplified and uniform set of regulations designed to establish essential construction standards for structural adequacy, fire resistance, public health and general amenity. The technical requirements of this code refer to AS3600 (2009) and other standards such as AS/NZ1170.0 (2002). There are several concrete floor systems that a designer can select and are economical and technically-satisfactory solutions. Concrete floor systems are generally reinforced using steel reinforcement, fabric, or high-strength strands that are pre-stressed. Pre-stressing concrete with a straight or a draped cable stabilises the applied loads by the uplift force and this limits how far the floor component can deflect. This is very beneficial for long-span floors because it removes the need to camber the formwork or to provide deeper sections.

With all systems, careful attention should be given to shrinkage and shortening due to prestressing because both processes generate large forces in a floor system constrained by stiff columns or walls (Warner et al. 2010). Figure 2-18 summarise floor systems with regard to the length of their spans.

94

Figure 2-18 floor systems (CCAA 2003; Warner et al. 2010)

95

2.13 Summary

The literature shows the role played by structural engineers in the sustainable design of buildings.

The key strategies and parameters associated with sustainable structural design strategies have been reviewed together with a detailed consideration of current international codes and specifications; the results highlight the main strategies and parameters which could affect the life cycle impacts of buildings. Various aspects of sustainable design and construction of buildings have been considered by researchers to minimise the environmental impact of structural design, but most focused on minimising the energy and resources consumed by buildings. Figure 2-19 provides the interactions between the key areas of knowledge reviewed.

96

Figure 2-19 Interaction between key areas of knowledge reviewed 97

There are noteworthy gaps in the field of sustainable structural design; one of the key challenges in the sustainable structural design of buildings is to improve their lifetime energy efficiencies and reduce thier CO2-e emissions impacts. Moreover, inconsistencies in estimating the CO2-e emissions impact raises the need to address uncertainties associated with CO2-e emissions impact made by different construction forms and concrete materials. The reviewed literature points out that structural engineers often overlook the influence that construction form and materials has on the thermal comfort and energy performance of a building, and not enough attention has been made to integrate the building and environmental costs into the decision making process. Chapter 3 presents a research methodology designed to meet the current study objectives and also cover those aforementioned knowledge gaps in the sustainable structural design of buildings.

98

Chapter 3

Research design

3.1 Introduction

Chapter 2 was a review of the relevant literature, and the contextual factors necessary for designing building with sustainable structures. The low carbon structure is linked to a reduction in their environmental impact resulting from the production of building materials, construction activities, maintenance and operational phases, and end of life processes. This Chapter proposes a methodology to integrate the lifetime of CO2-e emissions, and the structural and economic aspects of a building structure over its intended life cycle.

To evaluate the environmental aspects of alternative structural designs, a Life Cycle

Assessment (LCA) methodology is adopted because it has been used extensively to address environmental impact in the manufacturing and construction sectors (Ding

2014). The LCA in this study is used to quantify the carbon emissions associated with different design alternatives. The structural performance of buildings is also considered using the limit state method in the design process (AS3600 2009). The limit state design is a method used to design a structure by considering all actions likely to occur during its design life. The last step of this research methodology will combine the results obtained from a lifetime CO2-e emissions study and structural design and express them as a single economic value. The structural and environmental costs are presented in the form of monetary units because they also represent the final outcomes from the sustainable parameters. The research methodology of this thesis is shown in

Figure3-1. 99

Figure 3-1 Overview of the methodology in this study

100

3.2 Overview of this methodology

This study uses a benchmarking design against which structural design alternatives can be compared in terms of their performance and environmental impact. The benchmarking system in this study is used to examine different climate zones in

Australia, as well as different structural materials and methods of construction. This is carried out using a simulation model from a reference building as a base model and a sensitivity analysis. This comparative analysis is then presented in terms of lifetime energy and thermal performance as well as the life cycle cost and life cycle carbon emissions across design alternatives. This study then recommends ways to enhance the sustainable structural designs of buildings.

3.3 Benchmark building

Benchmarking aims to identify the right action where great savings can be made

(Bosteels et al. 2010). This refers to evaluating building energy efficiency by comparisons with standards or established energy benchmarks. This study examines the potential CO2-e emissions associated with the concrete structure by using a benchmarking method to compare and quantify how construction forms and concrete alternatives affect the lifetime impacts of a building in Australia.

A typical 15-storey office building is used because it is one of the four benchmarking buildings proposed by the National standard Organization (NSDO) in Australia

(NS11401.0 2014). This benchmark building will be used to simulate the carbon emissions, energy performance, and the cost of a building in Australia.

As mentioned in section 1.4.1, National Standards Development Organisation (NSDO) is a non-profit company which developed a principal standard NS11401.0 (NS11401.0

101

2014) in early 2014 to specify the requirements for assessing and declaring the comparative environmental impact of building products and systems in Australia

(NS11401.0 2014).

The system for constructing the proposed office building (reference building) is a typical high rise concrete structure in Australia (NS11401.1 2014); it has a square plan shape and a total floor area of 1000 m2 metres, with an average 3.30 m height per storey, as shown in Table 3-1.

Table 3-1 Benchmark building features

Building features Basement dimensions m 31.62 × 31.62 Number of Stories - 15 Average elevation per floor m 3.3 Total floor Area (including parking, Stairs & Verandas) m2 15,000 Total habitable area (external dimensions) m2 8,807.1 No of floors above ground level --- 11 No of rooms - 176

The building has two parts; the first three storeys are underground parking and storage areas, and the remaining twelve storeys are office areas. This building has a concrete frame structure with concrete columns and slabs, as well as walls. The exterior sides of this building are shielded by non-opening windows. The square plan shape is designed as 31.62×31.62 m and there are doors on all four sides of the ground floor, as shown in Figure 3-2 and Figure 3-3.

102

Figure 3-2 Plan of the benchmark Figure 3-3 Front view of the building (Source: NS11401.1) benchmark building (Source: NS11401.1)

3.4 Structural design parameters

As highlighted in sections 1.4 and 2.9, developments in concrete technology now offer a concrete mix with lower embodied CO2-e emissions and lower thermal conductivity.

It was shown that the appropriate choice of construction and building materials can potentially reduce the life cycle energy of buildings (section 2.9). The flooring of structural systems is a critical part of a building in terms of the CO2-e emissions

(section 2.12) and thermal performance (section 2.9).

As such in this study two aspects of structural design are examined over the whole life cycle of a building: floor systems (slab thickness) in different systems of construction, and different types of concrete (low density and high density).

A structural analysis is used to verify that the proposed building is a building design which could be used in practice for the purpose of this thesis. Based on the categories

103

in section 2.12.1, the structure of this building is designed with moment resisting frames and shear walls. The structural analysis first considered Flat plate as a floor system with a conventional column arrangement (5.27m) (NS11401.1 2014); based on the literature reviewed (section 2.12.2), the proposed building was then analysed and designed for different floor systems to cover a wider range of designs. For the alternative designs, the floor system is designed for slabs from 185 mm to 260 mm thick, depending on the type of floor systems. These flooring systems consist of the following:

- A post tensioned slab

- A Waffle slab

- A flat slab with a drop panel

Two types of concrete are used as the main structural materials; normal weight and

Ultra-lightweight. Normal weight concrete has a density between 2,400 to 2,500 kg/m3 and Ultra-lightweight concrete has a density lower than 1400 kg/m3. Detailed information about those types of concrete is provided in Chapters 4, 6 and 7, and the structural design alternatives are summarised and shown in Figure 3-4.

104

Figure 3-4 Structural design parameters

3.5 Structural analysis and design method

In terms of structural analysis and design, a detailed concrete structure design is considered by following the Australian Standards Concrete structures (AS3600 2009).

Two main aspects of this code, i.e., the strength and serviceability, were taken into the account during the structural design of this building. An overview of the method used for structural analysis is shown in Figure 3-5.

105

Figure 3-5 structural analysis & design flow

In this study the amount of live load came from the Australian and New Zealand standard for imposed actions (AS/NZ1170.0 2002). The live loads for office storage, parking, and work rooms are 5kPa and 3 kPa, respectively. The dead load for the concrete members was obtained by multiplying the volume of the member with the unit weight of concrete. Wind loads on the building were determined in accordance with the Australian and New Zealand standard wind actions (AS/NZ1170.2 2011). The magnitude of wind pressure on the structure was calculated based on height above ground, size, importance, and location. The importance level of the building is level 3, due to the consequences of failure based on the expected high rate of occupancy and use AS 1170.2002. For ultimate limit states and serviceability, the annual probability 106

exceedance came from AS 1170.2.2002, Table 3.1, for designed working life of 50 years in a cyclic zone in Australia. To calculate the wind load, Zone D is used to ensure there is enough strength in the structure as well as validating the potential of constructing the building in other zones. For the loading conditions, a combination of actions is used to check the serviceability and strength of the building in accordance with clause 4.2.1 and 4.2.2 of the AS1170.2002, as shown in Table 3-2. The Computer

Aid Design (CAD) package Etabs, Safe and a Microsoft Excel spreadsheet are used to verify the minimum requirements of the concrete design code. Details of the structural analysis and design are shown in Appendix B.

Table 3-2 Loading conditions for designing the building

Loading conditions Type of load Load (kPa) Live load-Office storage and parking area 5 Live load-Work rooms 3 Dead Load 4.3 Wind Load- Windward Ultimate limit states 6.6 Serviceability limit states 5.4 Wind Load- Lee ward Ultimate limit states 4.1 Serviceability limit states 3.4 Wind Load- Side wall Ultimate limit states 1.3 Serviceability limit states 1.1 Load combinations for Load combinations for serviceability states design Ultimate states design 1.35G G+ Ψl Q 1.25G+1.5Q G+ Ψs Q 1.25G+1.5ΨlQ G+ ΨsQ + Ws 1.2G+Wu+ΨcQ 0.9G+Wu G: permanent action (dead load); Q: Imposed action (Live load); Wu: ultimate load action; Ws: serviceability wind action; Ψl: Factor for determining quasi-permanent values (long term) of actions; Ψs: Factor for determining quasi-permanent values (long term) of actions; Ψc: Combination factor for imposed action;

107

3.6 Environmental assessment method

An environmental assessment method is used to estimate and quantify the potential environmental impact of different structural systems over the lifetime of the benchmark building. A life cycle environmental analysis of the building is divided into the following five phases: production, construction, use, end of life, and beyond life

(as shown in Figure 3-6) (Moncaster & Symons 2013).

Figure 3-6 Building Life cycle

The production and construction phases considered the extraction of raw materials, manufacturing, processing, and transportation to site; this is often called the “cradle to site” portion. The use phase considers the energy required to operate the building over its lifetime, while the end of life phase assumes the total estimated energy consumed at the end of life; beyond life represents the benefits and loads of components for reuse,

108

materials for recycling, and energy recovery for future use. The last stage of life cycle assessment (beyond life) is excluded from the scope of this study.

3.7 Life cycle environmental assessment approach

A life cycle environmental analysis assesses the lifetime building GHG emissions associated with the energy inputs; this concept can be used to illustrate the lifetime benefits of strategies design to minimise the GHG emissions (CO2-e emissions) of buildings (Cabeza et al. 2014). The boundary of this system categorises the embodied

CO2-e emissions at the initial stage, operational stage, and end of life stage of buildings. The initial embodied CO2-e emissions considers the impact of CO2-e emissions as a result of the extraction of raw materials to the manufacturing processes, and from transportation to the construction site and construction activities. The operational phase accounts for the energy consumption and material replacement during the lifetime of the building, while the end of life provides information about embodied CO2-e emissions associated with the demolition, transportation, and landfilled at the end of building life.

Information about embodied CO2-e emissions associated with different concrete was extracted from several published outputs to address the lack of information on CO2-e emissions of different concrete components in Australia (as shown in Table 4-2). The boundary of this system for calculating the total embodied CO2-e emissions is shown in Figure 4-2. This study considered the embodied CO2-e emissions associated with concrete and concrete materials from cradle to gate. This study includes every step from the extraction of raw materials, transport to the , mixing, and the production of concrete by considering the energy consumed (Diesel fuel, LPG fuel and

109

electricity). The process of transportation and placement of concrete was excluded from this study. Table 4-3 summarises the final embodied CO2-e emissions coefficients that are related to individual concrete components based on Australian data.

For the purpose of comparison, the results of embodied CO2-e emissions associated with various types of concrete were compared against six publicly available inventory databases and guidelines that are commonly used in Australia:

- ICE (Hammond et al. 2011),

- Crawford (Crawford 2011),

- Alcorn (Alcorn 2003),

- eTool (eTool 2014),

- BPIC (BPIC 2014)

- AusLCI (AusLCI 2016)

The carbon dioxide (CO2-e emissions) associated with the operational phase is estimated based on the energy used by the benchmark building. The building energy simulation software DesignBuilder is used to determine how different structural systems perform based on the ongoing energy consumption of a building (section

2.10). As section 2.10 shows, DesignBuilder is a user interface for the EnergyPlus dynamic thermal simulation engine; this software requires hourly weather files (based on epw standard format) while the required inputs for equipment and occupancy heat gaining and schedules are extracted from (NS11401.1 2014) and ABCB (2015).

Detailed information about the assumption and building properties are shown in section 6.3.1 and section 6.3.4. The results of operational energy are presented in terms 110

of the total energy used as well as the energy loads which are dominated by cooling energy across different design alternatives (Robati et al. 2017). The total energy demand of the building across different Australian climate zones was multiplied by the

Australian national emissions factor in 2015 that was proposed by the Australian

National Greenhouse Accounts (DEE 2016; DEE 2016), and then converted to CO2-e emissions values (more information is provided in section 7.33). The embodied CO2- e emissions associated with construction work, transportation, and final demolition at the end of life are estimated at a level of about 1% (Ruuska & Häkkinen 2015; Sartori

& Hestnes 2007).

The mean distance from the manufacturing companies to the site (central business district for each city) is measured using the Google map tools (Poinssot et al. 2014).

The five major cities are Sydney, Melbourne, Brisbane, Canberra and Darwin. The following table (Table 3-3) summarises all the steps involved in a life cycle environmental assessment and the available tools used in this thesis.

111

Table 3-3 Life cycle assessment procedure life cycle assessment procedure Life Source of Data and method Tools for this Verification cycle Activity used study method phase Extracted initial

Manufacturing embodied CO2- coefficient from e emissions data for the several Building building materials from published material Compare

literature and publicly available Australian production against 5 databases, as shown in outputs (for publicly Chapters 4 (4.5) and Chapter 5 concrete); Excel available (5.4). spreadsheet for databases in analysis. Australia and an Average distances for material overseas one,

use (Manufacturing) use transport by considering central - including etool; business district; Available Pre Transport ICE; literature data about CO -e Statistical 2 Alcorn; (a) emissions for transport (Shown analysis tools Crawford; in Chapter 5.4). (JMP) and Excel BPIC; AusLCI Building spreadsheet Energy uses from published construction, literature (shown in Chapter including 5.4). refurbishment DesignBuilder as a simulation

Use of software based on the Energy- Australian electricity and Plus solving engine. The DesignBuilder national average fuels for results were validated by and Excel for commercial cooling, (b) Use (b) comparing them with spreadsheet building energy lighting and

(Operational)* published results from the consumption ventilation

literature (shown in Chapter 6).

Average CO2-e emissions associated with demolition extracted from Building Excel extracted from published published demolition spreadsheet

literatures (shown in Chapters literatures (Demolition)

(C) End of life of (C) End 5 and 7).

Total embodied CO -e Combining the 2 Phase (a+b+c); (shown in emissions of results of a, b --- Chapter 7)

the building and c Life cycle cycle Life during its life

environmental impact environmental cycle

*This type of highly glazed office building risked overheating during the summer and winter periods. As such, the cooling load is the dominant load across all five climates studied (Robati et al. 2017).

112

3.8 Uncertainty analysis of Life cycle embodied CO2-e emissions uncertainty

analysis

As shown in section 2.11, little is known about the uncertainties associated with the embodied CO2-e emissions of building materials over the life cycle of buildings.

Chapter 5 in this study aims to quantify uncertainties associated with the calculation of life cycle of embodied CO2-e emissions buildings (Cradle to Grave). That Chapter assesses the life cycle of CO2-e emissions from four different construction forms used (section 3.5) in a typical 15 storey office building in Australia. Monte Carlo techniques were used in the Chapter to quantify the variability associated with environmental impact (CO2-e emissions) for each construction material and structural system. Monte Carlo is a statistical method that uses random values from input parameters and presents a distribution for the output parameter as examined by numerous researchers (Bisinella et al. 2016; Bojacá & Schrevens 2010; Ciroth et al. 2004; Fitch & Cooper 2005; Grant et al. 2016). Section 5.4 provides detailed information pertaining to the main methodological stages.

3.9 Analysis of Life cycle cost

Assessing the life cycle costs includes the initial cost, the operating and maintenance costs, and the disposal cost (as shown in Figure 3-7). For the purpose of this study, the initial cost includes the expenses associated with the manufacturing materials and construction of the building. The operating and maintenance costs include the running energy and material replacement costs during the lifetime of the building. The disposal costs include the demolition costs.

113

Figure 3-7 Life cycle cost of building

The initial costs associated with building materials and construction activities are taken from the Australian construction handbook based on 2016 data (Rawlinsons 2016).

The operating energy cost is estimated based on the annual energy consumption (based on energy simulation results) and estimated future Australian energy prices

(Economics 2015; Jacobs 2016) for the cost of energy up to 2040. A future cost analysis is then used to extend the energy costs from 2040 to 2066 (50-year lifetime of the building- base year 2016). The Present Value (PV) of operational costs is estimated based on Equation (3-1) with a 7% discount as a nominal rate per year

(Lawania & Biswas 2016).

푏푢푖푙푑푖푛푔 푙푖푓푒푡푖푚푒=50 푓푢푡푢푟푒 푐표푠푡 푃푉 표푓 표푝푒푟푎푡𝑖푛푔 푐표푠푡 = ∑ 표푐푐푢푟푟𝑖푛푔 푦푒푎푟 (3-1) 푦=1 (1+푑푖푠푐표푢푛푡푒푑 푟푎푡푒)

The PV replacement costs for the building materials are estimated as having a shorter lifetime than the buildings (50 years) (Rauf & Crawford 2012). The replacement cost 114

is calculated based on the current price of the materials and an escalation (inflation) rate of 3% (RBA 2016). Equation (3-2) is used to represent the present value for maintenance cost (Fuller 2010).

50 푐푢푟푟푒푛푡 푚푎푡푒푟푖푎푙 푝푟푖푐푒∗(1+푒푠푐푎푙푎푡푖표푛 푟푎푡푒)표푐푐푢푟푟𝑖푛푔 푦푒푎푟 푃푉 표푓 푚푎𝑖푛푡푒푛푎푛푐푒 푐표푠푡 = ∑ 표푐푐푢푟푟𝑖푛푔 푦푒푎푟 푦=1 (1+푑푖푠푐표푢푛푡푒푑 푟푎푡푒)

(3-2)

The costs associated with demolition at the end of life are estimated based on the future cost analysis (shown in Equation 3-3). Future costs are was estimated based on the national average cost of demolition per square metre ($112/m2) for a fifteen storey reinforced concrete frame and slab buildings (Rawlinsons 2016) over a 50 year lifetime of buildings.

푐푢푟푟푒푛푡 푑푒푚표푙푖푡푖표푛 푐표푠푡∗(1+푒푠푐푎푙푎푡푖푛 푟푎푡푒)50 푃푉 표푓 푑푒푚표푙𝑖푡𝑖표푛 푐표푠푡 = (3-3) (1+푑푖푠푐표푢푛푡푒푑 푟푎푡푒)50

The expenses associated with refurbishment and development of the external site are not included in this study.

3.10 Environmental impact cost analysis

The CO2-e emissions (in section 3.6) calculated for each stage of the designed buildings (production, construction, use, and end of life) are used to quantify the relative environmental impact of several structural design choices. These structural choices include two flooring systems (Waffle slabs and Flat slabs) with two different types of concrete (Normal weight and Ultra-lightweight concrete). The total embodied

CO2-e emissions value of each phase of the building is then converted into costs. The monetary value of CO2-e emissions is based on Adams et al. (2014) method and the

Australia Emissions Trading Scheme (Combet 2012) with an inflation rate of 3% per year (RBA 2016). Future CO2-e emissions costs are discounted at a rate of 7% per year 115

(Lawania & Biswas 2016). Equation 3-4 is used to determine the present value of CO2- e emissions over the lifetime of the buildings. The base year (2016) is used for all calculations.

50 Current 퐶푂 −e price ∗ (1+escalation )푛표푐푐푢푟𝑖푛푔 푦푒푎푟 푃푉 표푓 푒푛푣𝑖푟표푛푚푒푛푡푎푙 푐표푠푡 = ∑ 2 (3-4) (1+푑푖푠푐표푢푛푡푒푑 푟푎푡푒)푛표푐푐푢푟𝑖푛푔 푦푒푎푟 푦=1

Figure 3-8 represents the methodology used for the environmental cost analysis in this study.

Figure 3-8 Environmental cost estimation method

The economic value obtained from the life cycle costing assessment and the environmental cost assessment is used to select the most suitable structural design; the resulting cost is useful because it compares total costs across the reference benchmark and the alternative designs. This then enables the comparison data to be described with a single global assessment parameter. A summary of this framework is shown in Figure

3-9 (Tsimplokoukou et al. 2014).

116

Figure 3-9 Global assessment framework

3.11 Summary

This Chapter describes the research methodology in order to include the life cycle carbon emissions and energy consumption in the process of structural design during decision making. This research methodology is used to evaluate the CO2-e emissions associated with several construction forms and structural concrete. The structural analysis part included in this study also examines the performance-based design of the buildings by considering the limit state method. The last step of this research methodology combines the results obtained from a lifetime CO2-e emissions study and the structural design and expresses them in terms of their economic value. All the CO2- e emissions impacts are converted into CO2-e emissions, which is then multiplied by the Australian National carbon price.

The integrated structural and lifetime CO2-e emissions costs are presented in terms of monetary units to provide a single global sustainable parameter that could be used to evaluate the alternative structural design of buildings. The results of this framework are used as an evaluation and comparison method across alternative structural design solutions. The following Chapters describe the environmental assessment and energy

117

performance analysis during operation, as well as the economics of the designed buildings.

118

Chapter 4 Incorporating an environmental evaluation and the thermal properties of concrete mix designs

Contribution of the candidate to the Published Work

The contribution of the candidate in the published paper was 90%. He co-authored them with his main and co-supervisors. The candidate collected and analysis data and wrote the paper. Other authors reviewed the papers and provided useful feedback for improvement.

Citation: Published

Robati, M, McCarthy, TJ & Kokogiannakis, G 2016, ‘Incorporating environmental evaluation and thermal properties of concrete mix designs’, Construction and Building Materials, vol. 128, pp. 422-35.

119

4.1 Introduction

Chapter 3 proposed a methodology to integrate the environmental, structural, and economic aspects of a building structure over its life cycle. The literature review in

Chapter 2 highlighted the sustainable structural design strategy defined as a way of selecting structural materials and forms to reduce their impact through lifetime of a project (Pongiglione & Calderini 2016). One of the main challenges in the sustainable design of buildings is to improve their energy efficiency over its lifetime whilst simultaneously reducing the environmental impact of the initial design.

As section 2.9 shows, recent advances in concrete technology result in a lower environmental impact due to the application of supplementary cementitious materials and recycled aggregates; there are also improvements in thermal properties due to the application of admixtures. However, the correlation between the environmental impact

(Cradle to Gate) and thermal performance of concrete mix designs has not been researched adequately because the GHG emissions associated with producing the individual components of concrete must be considered with greater refinement. This

Chapter correlates the impact of selecting a concrete mix design in terms of CO2-e emissions with the resulting thermal conductivity and density at the design stage of buildings.

4.2 Low carbon concrete material

Concrete is the most widely used construction material in the building industry and as such consumes the second highest amount of natural resources (ISO15673 2005). The main constituents of general purpose concrete are cement, water, and aggregates, and the most carbon intensive components in the manufacture of concrete are cement and

120

aggregates. A report released by the United States Geological Survey shows that the global production of cement increased by 100 million tonnes in one year to 4.18 billion tonnes in 2014 (Survey 2015). The American Association (PCA) estimates that the consumption of cement will continue to increase into the future

(PCA 2015).

Concrete is a popular material because it has excellent mechanical and durable properties, it is also adaptable, relatively fire resistant, and generally available and affordable. Concrete can absorb and retain energy for a long period of time, an action which reduces energy consumption by transferring heat in a natural daily cycle through the structural components (thermal mass) of the building. The mass components reduce the fluctuations in temperature in building spaces which in turn decreases the associated peak heating and cooling loads (Torgal & Jalali 2011).

Through its high thermal mass, a concrete slab will absorb heat during the day and release it back into the room at night. The relatively high specific heat of solid concrete makes it attractive as a passive thermal store, so an appropriately designed concrete mix can offer the benefits of thermal performance by reducing the consumption of heating and cooling energy in buildings (Anderson & Silman 2009; Appleby 2012).

This situation raises the question about how best to design a concrete mix with respect to its strength, thermal properties, environmental impact, and CO2-e emissions intensity. This Chapter seeks to identify the environmental impact and thermal performance of different concrete mix designs by examining the impact of CO2-e emissions on the thermal properties of concrete.

121

4.3 Thermal performance of concrete

Concrete is one of several building materials that possess high thermal properties. In cold seasons, high thermal mass building elements that contain concrete, such as walls and floor slabs, absorb radiant heat from the sun during the day and gradually release it back into the system (space) during the night when the outside temperatures drop

(Lemay & Leed 2011). The distinct benefit of high thermal mass is to moderate changes in the peak load of energy requirements due to fluctuations between inside and outside temperatures. High thermal mass causes a time lag between internal and external temperatures (Figure 4-1), but it also stores heat which dampens the fluctuations between peaks; this often improves the thermal comfort and decreases the demand for energy for heating and cooling (Lemay & Leed 2011). Besides its thermal mass, the thermal properties of a concrete mix design such as conductivity, also influence passive heating design strategy. An optimum design of concrete mix could either reduce the escape of passive heating before being absorbed or re-release stored heat before the colder night (DIIS 2013).

Figure 4-1 Damping and lag effect of thermal mass (Lemay & Leed 2011)

122

The thermal conductivity of concrete mix designs is influenced by the thermal properties of the cement, aggregates, and the existing moisture (ACI122R 2014). The thermal conductivity of concrete depends on the type of aggregates used in the mixture.

Some published construction properties databases associate thermal conductivity with the density of concrete, for example ACI122R (2014) and CIBSE (2006); this means that some thermal properties of concrete mixes at the initial stage of the structural design of buildings can be considered. This Chapter quantifies the thermal conductivity of different concrete mix designs.

4.4 Environmental aspects of concrete

The basic constituents of concrete are binder (cementitious materials), coarse and fine aggregates (or inactive mineral filler), and water. The properties and combinations of these materials affects the various admixtures, and how it is handled during construction determines the properties of the in-situ concrete.

A major source of greenhouse emissions during the production of concrete is Portland cement; indeed the cement sector is responsible for producing 2,823 million metric tons (Mt) of embodied CO2-e emissions in 2010 (Kajaste & Hurme 2016); this is almost 9% of global CO2-e emissions from burning fossil fuels in 2010 (Kajaste &

Hurme 2016). Traditional methods of responding to this issue are to develop energy efficient cement production plants through improved technology, change the energy sources used and use substitutes for clinker such as fly ash and ground granulated blast furnace slag (Ishak & Hashim 2015; McLellan et al. 2012; Rahman et al. 2015; Worrell

2008).

123

The concrete industry is addressing some of the worries about environmental issues by supplementing or replacing the use of cement and other components associated with high embodied CO2-e emissions. Several researchers have studied the possibility of replacing cement in concrete with recycled materials (de Castro & de Brito 2013;

Ingrao et al. 2014; Jacoby & Pelisser 2015). The use of alternative cementitious materials remains the main path to reducing embodied CO2-e emissions in the concrete industry (Mehta 2002). Wimpenny (2009) conducted a study in low CO2-e emissions alternatives to concrete by exploring the strategies being adopted and developed in 12 countries around the world; his results are classified into the seven groups shown in

Table 4-1.

Table 4-1 Embodied CO2-e emissions for cementitious materials (Wimpenny 2009)

Group Example Suggested embodied References

quantities CO2-e emissions Alternative Fly ash 40% Medium (Newlands et cementitious Slag 80% Low al. 2012; materials Silica fume 10% Low Wimpenny 10% Very high 2009) Municipal solid waste ----- Medium incinerator ash (MSWIA) Non-Portland Geopolymer ----- Low (McLeod cement binders Calcium sulphate based Low 2005; Taylor Calcium sulfoaluminate High & Collins Magnesite based High 2006) Low cement Lean Concrete ----- Medium (Wimpenny concrete 2009)

124

Ultra-high Fibre reinforced ----- Medium (Wimpenny strength superplasticiser silica fume 2009) concrete concrete (FRSSFC) Changes in Oxygen enrichment of kiln ----- Medium (Taylor & Portland atmosphere to enhance Collins cement burning 2006) manufacture Belite ----- Very high Alinite and Fluoralinite ------cement ------Portland limestone cement Alternative Bituminous based materials ----- Very low (Wimpenny binder types (Agent C) 2009) Carbon Sequestering carbon from the ----- Very low (McLeod capture kiln capturing carbon in the 2005) concrete, e.g. Hemp (Lime based binder and hemp) The most commonly used alternative cementitious materials are Ground Granulated

Blast Furnace Slag (GGBFS) and coal combustion fly ash. GGBFS is a by-product of iron and steel making and fly ash is a by-product of burning coal, mainly for generating electricity. These cementitious materials are used to replace a portion of the cement in concrete mix designs. The process used to produce fly ash and GGBFS involves less greenhouse gas emissions than ordinary Portland cement (Van den Heede & De Belie

2012).

Fly ash is widely available which, if not used in concrete, is an industrial waste with serious disposal problems. Worldwide, most of the annual production of fly ash is disposed of as waste material in ash dams or in a landfill (Dhir 2006). In Australia, about 20% of fly ash produced in coal-fired power stations is used in construction industry (Wang & Wu 2006). The Australian Standard, AS3582.1, sets specific requirements for fly ash and has classified it into three grades (fine, medium and 125

coarse)(AS3582.1 2016). If the physical properties of fly ash do not comply with the

AS3582.1 Standard requirements it cannot be used as a supplementary material in the cement and concrete industry (Dhir 2006). The proportion of fly ash in blended cement typically changed from 15% to 30% and for some particular applications, it can be increased to 50% and 60% (Malhotra & Mehta 2002; Marsh 2003). The positive contribution of fly ash for reducing concrete embodied CO2-e emissions can be up to

44% when it substitutes 40% of Portland cement in a typical concrete mix design(ADAA 2012), however, this decrease in the use of coal might also have a negative impact on the supply of fly ash (Gursel et al. 2016).

Other supplementary materials such as GGBFS can also be used to replace Portland cement, in fact, substituting a portion of Portland cement with GGBFS can substantially reduce the negative environmental impact of concrete (Obuzor et al.

2011). Fly ash and GGBFS can be added separately to a concrete mix, however, unlike fly ash, the amount of available GGBFS is limited. The worldwide production of

GGBFS is only 25 million tonnes per year (Malhotra 2006). The proportion of GGFS in concrete typically varies from 40% to 60% of the overall amount of blended cement

(Virgalitte et al. 1995).

Other supplementary cementitious materials are silica fume, rice husk ash, and recycled ground glass, but their availability is limited, unlike fly ash, so their costs are much higher (Glavind 2012).

Geopolymer concrete is another alternative where an alkali activated aluminosilicate material is used instead of traditional cement binders (Huiskes et al. 2016).

Geopolymers generally have a lower embodied CO2-e emissions than cement but are much more expensive to produce (Berndt et al. 2013). 126

Meanwhile, there are some barriers to implementing these newer types of materials to achieve lightweight and/or geopolymer concrete such as regulatory, technical, supply chain, and the cost of geopolymer concrete (Cabeza et al. 2013; Duxson & Provis

2008; Van Deventer et al. 2012). Several research programs are currently trying to remove these barriers to allow for a wider application of geopolymer and/or lightweight concrete.

Aggregates affect the physical properties of concrete (grade, moisture absorption, thermal conductivity, etc.), but they can also be reused as raw materials in the concrete at the end of life (Gravitt 2013). The actual choice of aggregates is very much related to the local supply chain because quarries with enough natural aggregates are rapidly being depleted in some regions and countries, which means the tendency to use crushed and manufactured aggregates is increasing (Rao et al. 2007). From an emissions point of view, a distinction must be made between natural and crushed aggregate; natural aggregates such as sand and gravel are the result of weathering and erosion and do not require any processing other than collection and transportation, whereas crushed aggregates such as manufactured sand, are mined from quarries and require mechanical crushing. Flower and Sanjayan (Flower & Sanjayan 2007) showed that granite/hornfels as a crushed aggregate has GHG emissions of 45.9 kg CO2- e/tonne and basalt as natural aggregates has GHG emissions of 35.7 kg CO2-e /tonne

(Flower & Sanjayan 2007) .

The demand for water for concrete depends on the design of the mixture and the use of plasticising additives. Water in concrete leads to minimal embodied CO2-e emissions, which leaves the cement, the coarse and fine aggregates, and GGBFS and fly ash as having the main environmental impact. 127

Previous studies into the environmental impact of producing cementitious materials and aggregates has already yielded several estimates of the embodied CO2-e emissions per tonne of concrete (Flower & Sanjayan 2007; Malhotra 2006; Mehta 2002; O’Brien et al. 2009). The embodied CO2-e emissions is calculated by multiplying the embodied

CO2-e emissions coefficients from the proposed databases (Alcorn 2003; Crawford

2011; eTool 2014; Hammond et al. 2011) for each grade of concrete by the quantity of concrete. However, individual concrete components have not received enough attention, which means the GHG emissions associated with each individual component of concrete must be properly investigated (Flower & Sanjayan 2007). Furthermore, the relationship between embodied CO2-e emissions, thermal conductivity, and alternative cementitious materials has not been determined well enough, which is why the primary objective of this Chapter is to identify the relationship between low embodied CO2-e emissions and low thermal conductivity for a large number of concrete mix designs. This Chapter analyses different concrete mix designs and compares the results while sourcing inputs from a number of available inventory databases.

4.5 Methodology

4.5.1 Materials and mix designs

This Chapter investigates 90 different concrete mix designs where the two primary performance variables are the grade and density. These concrete mix designs have been collected from 8 published journal papers and databases (Berndt 2015; CCAA

2015; Damdelen et al. 2015; Marinkovic et al. 2010; O'Moore & O'Brien 2009; Tošić et al. 2015; Wu et al. 2015; Yun et al. 2013; Zhang & Poon 2015); they represent some 128

conventional (normal weight) and some advanced methods of concrete admixture

(Marinkovic et al. 2010; Wu et al. 2015; Yun et al. 2013; Zhang & Poon 2015) that result in lightweight and ultra-lightweight concrete. Table 4-2 summarises the concrete grades and the 90 mix cases of different batches of concrete that are analysed in this

Chapter. Novel forms of concrete admixture (such as Mix 27-41) are included in this

Chapter to illustrate their thermal properties and environmental impact because they are not covered in mainstream studies. These concrete grades range from 28 MPa to

87 MPa; their concrete mix designs and ingredients are shown in Appendix A.

Table 4-2 Summary of concrete batches

Concrete Composition of mix Mix Grade Source No. Binder Aggregates Admixture Water (MPa) Portland natural aggregates, Potable Mix cement, recycled aggregates, water, CCAA 32,40 Water reducing 1-3 GGBFS, fly manufactured sand, Reclaimed (2015) ash fine natural river sand water natural aggregates, manufactured sand, Zhang and Mix Portland Potable 31.6-42.7 Lightweight Water reducing Poon 4-9 cement water aggregate*, Furnace (2015) bottom ash Portland Mix natural aggregates and Potable Berndt 32,35,40 cement, ------10-26 manufactured sand water (2015) fly ash Superplasticiser, shrinkage Portland reduction, Mix cement, natural aggregates and Potable Wu et al. 33-69.4 Viscosity modify 27-41 cenosphere, manufactured sand water (2015) agent, silica fume Polyethylene fibers, Silane Portland cement, Damdelen Mix natural aggregates, Potable 38-55 GGBFS, fly ------et al. 42-57 recycled aggregates, water ash, silica (2015) fume natural aggregates, Portland Mix Lightweight Potable Yun et al. 23-43.9 cement, ------58-69 aggregate*, Glass water (2013) fly ash bubble

129

Marinkovi Mix Portland Potable 33.6-48.6 natural aggregates ------c et al. 70-75 cement, water (2010)

Mix Portland natural aggregates, Potable Tošić et 41.5-44.2 ------76-79 cement, recycled aggregates, water al. (2015)

Portland O'Moore Mix cement, Potable and 32 natural aggregates ------80-90 GGBFS, fly water O'Brien ash, (2009) *Lightweight aggregate consists of manufactured aggregate (shale, slate and clay) and Glass bubble. This Chapter considers each individual component of concrete in order to estimate the equivalent greenhouse gas emissions and thermal conductivity of a particular design mix. The embodied CO2-e emissions for a variety of concrete mix designs is quantified by collecting the relative embodied CO2-e emissions coefficients for each individual component of concrete from existing studies (ADAA 2016; Flower & Sanjayan 2007;

McRobert 2010; Rouwette 2012).

The estimated emissions coefficient for each material is then multiplied by the respective quantity of material, and then the resulting embodied CO2-e emissions is summed up for each mix design. This comparison includes the results obtained from this Chapter that were compared to six publicly available Australian and England embodied CO2-e emissions data inventories, namely; Crawford (Crawford 2011),

Alcorn (Alcorn 2003), eTool (eTool 2014) and BPIC (an average industrial practice database) (BPIC 2014) and AusLCI (AusLCI 2016), ICE (Hammond et al. 2011) from

England. Since the study by Crawford covers embodied energy rather than embodied

CO2-e emissions, a conservative coefficient of 10% (based on the ratio used in eTool database) is used to convert data into embodied CO2-e emissions (kg CO2-e emissions). A linear interpolation is used in the Crawford databases to estimate the coefficient of embodied CO2-e emissions for every grade of concrete proposed in the

130

concrete mix data of this Chapter. With the ICE database, a linear interpolation is used to estimate the embodied CO2-e emissions coefficient when different percentages of cement are replaced with slag and/or fly ash. Calculating the thermal conductivity of each design mix is done by following the ACI122R (2014) guidelines which propose that the thermal conductivity of a concrete mixture is based on the individual material properties of which the mixture consists (aggregate), and the oven dry density of the mixture (kg/m3).

4.5.2 Embodied carbon dioxide equivalent emissions

The emissions factors for binders, aggregates, and admixtures are from Flower and

Sanjayan (2007) and are based on the Australian greenhouse office factors and method workbook (AGO 2004). The emissions factor for recycled aggregates comes from the

ARRB Group report (McRobert 2010). The embodied emissions associated with manufactured aggregates are considered to be the same as natural aggregates with regards to the upstream stage of the production process in Australia (Chandra &

Berntsson 2003). The emissions associated with potable water and captured water are based on the results of Rouwette (2012). The boundary of the system for calculating the total embodied CO2-e emissions is depicted in Figure 4-2. This Chapter considers the embodied CO2-e emissions associated with concrete and concrete materials from cradle to gate; this means it includes all the steps from extraction of raw materials, transport to the concrete plant, mixing, and the production of concrete by considering the relevant consumed energy (Diesel fuel, LPG fuel and electricity). The process of transportation and the placement of concrete is excluded from this Chapter. Table 4-3

131

summarises the final embodied CO2-e emissions coefficients that are related to individual concrete components based on Australian data.

Figure 4-2 Concrete embodied CO2-e emissions system diagram

Table 4-3 Final embodied CO2-e emissions coefficients

Activity Material Emissions coefficient References Binder Type of Portland cement 0.820 (ADAA 2016; (t CO2-e/ Ground Slag ; Ground Granulated 0.143 Flower & tonne) blast furnace Sanjayan 2007) Fly ash or pulverized fuel ash 0.027 Furnace bottom ash 0.027 cenosphere 0.027 Aggregates Natural aggregates 0.0459 (Flower &

(t CO2-e/ Recycled aggregates 0.004 Sanjayan tonne) Manufactured Sand 0.0139 2007; McRobert Fine natural river sand 0.0139 2010) Admixture Water reducing admixture 2.2 × 10−6 (Flower & (t CO2-e / L) Superplasticiser 5.2 × 10−6 Sanjayan 2007) Water Potable water 7 × 10−4 (Rouwette

(t CO2-e / 2012) tonne) Captured/ Reclaimed water 7 × 10−5

132

4.6 Results and discussion

4.6.1 Embodied emissions

The resulting cradle to gate life cycle embodied CO2-e emissions for the 90 concrete mixtures are shown in Figure 3. The quantities of embodied CO2-e emissions relate to

3 1 m of concrete, and as the results in Figure 4-3 show, the amount of embodied CO2- e emissions is influenced by variations in the concrete mixture.

Figure 4-3 Embodied CO2-e emissions for different grades of Concrete

3 Figure 4 shows the variation of embodied CO2-e emissions per m of concrete for the

32-35 MPa and 38-42 MPa groups of concrete. This data is categorised into common standardised grades of 32 and 40 MPa due to variability in the expected concrete strength (Neville 2012), and because these two categories are popular in the structural design of buildings. The results of graphically depicted embodied CO2-e emissions show the variations and the different mix designs for the two selected groups. The statistical distribution of data shows interquartile ranges between 72.9 and 103.1 Kg

3 CO2-e /m for group 32-35MPa and 38-42 MPa respectively (Figure 4-4).

133

Figure 4-4 Variation in embodied CO2-e emissions for 32 MPa and 40 MPa concrete mixes

Concrete with a grade of 32-35MPa has an embodied CO2-e emissions range from

3 3 187.2 to 417.5 kg CO2-e/m by a central tendency of 277 kg CO2-e/m . The results shown in Figure 4-5 indicate that a mix number 13 and mix number 32 achieved the lowest and highest embodied CO2-e emissions respectively, compared to the other mixes. For mix design number 13, 65% of the binder is blast furnace slag and 35% is general Portland cement. The mix with the lowest emissions (mix design number 32) has 58% general Portland cement, 37% cenosphere, and 5% silica fume.

For group 38-42 MPa, the embodied CO2-e emissions was calculated to vary from 211

3 3 to 509 kg CO2-e /m by a median value of 311 kg CO2-e /m as shown in Figure 4-6.

Mix number 22 and 36 produced the lowest and highest amount of embodied CO2-e emissions per m3 of concrete, respectively. The mix 22 binder contains 35% Portland cement and 65% blast furnace slag. Mix 36 consists of 55% Portland cement, 40% cenosphere, and 5% of silica fumes.

134

Figure 4-5 Embodied CO2-e emissions for 32-35 MPa

Figure 4-6 Embodied CO2-e emissions for 38-42 MPa concrete

Various methods have been proposed to reduce the embodied CO2-e emissions of

Portland cement (Berndt et al. 2013; Damtoft et al. 2008; Gartner 2004; Worrell 2008).

For instance, producing cement can be more efficient by reducing the proportion of clinker and replacing it with ground granulated blast furnace slag (GGBFS). Moreover, supplementary cementitious and pozzolanic materials such as GGBFS, fly ash, silica fumes, rice husk ash, and metakaolin have been considered as a replacement for

Portland cement (Berndt et al. 2013; Srinivasreddy et al. 2013; Whiting et al. 2012).

This Chapter quantifies the effect of replacing portions of Portland cement with fly ash and GGBFS. The results show that concrete mixes with fly ash have 8% to 30% less embodied CO2-e emissions compared to a mix with 100% Portland cement (mix 80-

85), and GGBFS can reduce embodied CO2-e emissions by 15.5% in a concrete 135

mixture (mix 86-90). Moreover, the emissions associated with producing concrete are related to parameters such as the availability of raw materials in the region and the amount of emissions produced during transportation. This Chapter considered the embodied CO2-e emissions associated with concrete and concrete materials from cradle to gate, so parameters such as transportation, region, etc., are not taken into account.

4.6.2 Variations in embodied CO2-e emissions coefficient

The estimated embodied CO2-e emissions for the two selected grade groups of concrete are compared between the Crawford, ICE, Alcorn, eTool, BPIC and AusLCI inventory embodied CO2-e emissions databases. Figure 4-7 and Figure 4-8 illustrate the embodied CO2-e emissions across mixture designs for 32-35 MPa and 38-42 MPa grades.

* Include Silica fume.

Figure 4-7 Embodied CO2-e emissions across inventory databases for 32 MPa concrete

136

* Include Silica fume.

Figure 4-8 Embodied CO2-e emissions across inventory databases for 40 MPa concrete

This comparison shows that the amount of embodied CO2-e emissions for grade 32

3 MPa can vary from 62.8 to 495.9 kg CO2-e /m of concrete depending on the type of mix design and inventory database. Similarly, a much different embodied CO2-e

3 emissions for grade 40MPa concrete came from (from 70.3 to 616.3 kg CO2-e /m of concrete) across the different mix designs and databases. The resulting embodied CO2- e emissions based on Crawford, eTool and BPIC databases have treated concrete as one specific product and have proposed an individual coefficient for each grade of concrete regardless of the mix of ingredients. The minor changes (less than 4%) in each database, including BPIC, eTool, and Crawford are due to changes in the density of concrete mix designs and the embodied CO2-e emissions coefficients that are a function of concrete density. Alternatively, the comparison of concrete mixes from the

ICE database and this Chapter (using the coefficients of Table 4-3) show that mix designs 13 for 32 MPa concrete and 22 for 40 MPa concrete have the lowest embodied

CO2-e emissions due to replacing 65% of cement with blast furnace slag. As expected, 137

the maximum embodied CO2-e emissions was recorded for mix 32 and mix 36 for groups 32 and 40 MPa, respectively because no supplementary cementitious materials are used (i.e. 100% Portland cement was used).

The data in Figures 4-7 and 4-8 show that the results based on AusLCI and Alcorn analysis differ by less than 4%, while both databases can show variations between the mix designs. Like the results of this Chapter, the highest embodied energy are for mix designs 36 and 32 for a grade of 32 and 40 MPa, respectively. The lowest embodied emissions are archived through mix designs 13 and 22.

The current databases cannot address the effect of silica fume and cenosphere very well because alternative cementitious materials are used in concrete mix designs 32,

36, and 49 (as shown in Figure 4-7 and Figure 4-8). However, it is reasonable to assume there is no environmental impact associated with silica fumes because they are aby-product of the production of metallurgical grade silicon (Crossin 2012).

Furthermore, since the embodied CO2-e emissions associated with the cenosphere is similar to the CO2-e emissions of fly ash, it was assumed to be the same as fly ash in the paper because both materials are by-products from the production of power within coal fired power stations (ADAA 2016).

The embodied CO2-e emissions from using different inventory databases are summarised in Figures 4-9 and 4-10. The embodied CO2-e emissions values across the

3 Alcorn, Crawford, and eTool databases vary from 255 to 540 kg CO2-e /m for group

3 32-35 MPa and from 290 to 590 kg CO2-e /m for group 38-42 MPa. The differences could be due to variations in the method used to analyse each database, to different system boundaries, and to sources of data and the quality of input in calculating the upstream process (Dixit et al. 2010). 138

The embodied CO2-e emissions factor from ICE database varies for each different mix design, other than mix designs 32, 36, 39, which include silica fume and cenosphere because this database considers different proportions of cement and cementitious material such a slag and fly ash in the concrete. In terms of the maximum proportion of slag in mix designs 13 and 25, the ICE embodied CO2-e emissions coefficients are

3 62.8 and 70.3 kg CO2-e /m , but for specific mix designs, the ICE results match closely with those from Crawford (mix 1, 3, 68, 79) and Alcorn (mix 43, 47,61). For mix designs 6 and 9, the ICE results are the same as the results from BPIC.

A comparative analysis between AusLCI, Alcorn, and the current study reveals many variations of embodied CO2-e emissions in concrete mix designs. The average differences are 16 % and 7% for grade 32 and 40 MPa; these differences are due to variations in the embodied CO2-e emissions coefficients for general purpose cement,

GGBFS, fly ash, and natural and manufactured aggregates. For instance, AusLCI proposes a factor of 0.994 (tonne CO2-e emissions) for producing an average of 1 tonne of GP cement in Australia, which is 18% higher than the coefficients proposed in

Crossin (Crossin 2012) and Flower (Flower & Sanjayan 2007) studies (used in this

Chapter). Similarly, AusLCI proposes a higher emissions factor for manufacturing

GGBFS and recycled aggregates and lower embodied CO2-e emissions for producing fly ash than this Chapter (based on (Flower & Sanjayan 2007)). The embodied CO2-e emissions associated with the production of natural aggregates is not reported directly in AusLCI, whereas ARRB gives a value of 3.97 kg CO2-e emissions per tonne of materials (McRobert 2010). Moreover, Alcorn’s database does not properly address the embodied CO2-e emissions associated with alternative cementitious materials such as fly ash and GGBFS. 139

Figure 4-9 Variations of embodied CO2-e emissions for different databases (32- 35 MPa)

Figure 4-10 Variations of embodied CO2-e emissions for different databases (38- 42 MPa)

In summary, these variations in the embodied CO2-e emissions of different concrete mix designs could affect the overall life cycle assessment of a building and building materials because the results indicate that one product might be attributed with a lower embodied CO2-e emissions than another product in one database while the same product in another database could be attributed with the same or higher emissions. For example, the results based on AusLCI, Alcorn, ICE and those produced from the additional cases in our study show that mix designs 13, 18, 22, and 26 have the lowest amount of embodied CO2-e emissions in the 90 mix designs because they replaced

65% of the cement binder with GGBFS. However, the eTool, BPIC, and Crawford databases do not show these differences of embodied CO2-e emissions across the different concrete mixes. Furthermore, variations associated with production, 140

manufacturing techniques, the type of fuel used, and the source of raw materials and transportation distance across different geographic locations must be considered because they can be quite significant between areas within the same country (Crawford

2011; Crawford 2008). These differences between the databases point to a need for transparency with regard to their ability to analyse individual concrete components.

Meanwhile, a summary of the results (Figures 4-9 and 4-10) quantifies the variations which could lead to better comparisons for research that uses these databases.

4.6.3 Thermal conductivity of concrete mix design

A comparative assessment was used to estimate the thermal conductivity of all 90 mixes from ACI122R (ACI122R 2014), while experimental data for mixes 27 to 57 are reported in the relevant published articles (Damdelen et al. 2015; Wu et al. 2015) obtained. The proposed ACI values are from Table 3.a of ACI122R-2014 and are based on practical thermal conductivity design values for normal weight (2240 to 2400 kg/m³), light weight, and ultra-lightweight concrete (less than 1840 kg/m³). Figure 4-

11 shows the theoretical and experimental thermal conductivity for all 90 concrete mix designs. This Chapter uses the data from the ACI122R method to ensure consistent comparisons across all mix designs where, as expected, the thermal conductivity is influenced by variations in the concrete mixtures.

141

Figure 4-11 Thermal conductivity of concrete mix designs

This Chapter shows that the type of cement and aggregate affected the density and thermal conductivity, while replacing normal aggregate with lightweight aggregate reduces the density and thermal conductivity of concrete. The data indicates that using lightweight aggregate instead of natural coarse aggregate changes the density of concrete from 2320 to 1727 kg/m3. Moreover, the thermal conductivity of concrete also decreased when lightweight aggregates are introduced into the mix designs. For example, the results of mix 4 and mix 9 indicate that as the proportion of aggregates in these mix designs decreased, the thermal conductivity decreased from 1.96 to 1.16

(W/mK).

Figure 4-12 shows the variation in thermal conductivity per m3 of concrete across mix design groups 32-35 MPa and 38-42 MPa. These variations are the result of changes in the proportion of normal and lightweight aggregates in the concrete mixture. For example, mix designs 32 and 36 have the lowest thermal conductivity and a lower density than all other mix designs in groups 32-35 MPa and 38-42 MPa, respectively.

142

6.0

5.0

4.0

3.0

2.0 (W/mK)

1.0 Thermal Thermal conductivity 0.0 32-35 Mpa 38-42 Mpa Grade of concrete (MPa) Figure 4-12 Variations in thermal conductivity between concrete mix designs

A brief review of previously published values shows that the estimated thermal conductivity for grade 32-35 MPa and 38-42 MPa concrete mixes could vary from 3.1

W/(m.K) to 0.36 W/(m.K). For a grade of 32-35 MPa, the lowest and highest thermal conductivity occurs in mix designs 32 and 82, and for 38-42 MPa, the lowest thermal conductivity (0.31 W/(m.K)) occurs in mix design 36.

A comparison of all the embodied CO2-e emissions obtained from Table 4-3 and the thermal conductivity of mix designs show different correlations between two variables. Figure 4-13 shows the changes of embodied CO2-e emissions against the thermal conductivity of concrete mix designs; they are also shown in Appendix A.

Figure 4-13 Embodied CO2-e emissions versus thermal conductivity across all the concrete mix designs

143

For mix designs 27-41, the results show a positive gradient between changes of thermal conductivity and embodied CO2-e emissions; this means the amount of embodied CO2- e emissions increases as the thermal conductivity of concrete increases. Note that the rate of changes of embodied CO2-e emissions and thermal conductivity for mixes 27-

41 are much higher than the other mixes; this is due to the high proportion of Portland cement and low-density aggregates in mixes 27-41. However, the results from several other mix designs show a great deal of scatter in thermal conductivity without changing the embodied CO2-e emissions values, and vice versa; this is seen in mix designs 4 to 9, where the changes in thermal conductivities range up to 41% with only a 17% change in embodied CO2-e emissions.

Figure 4-14 Embodied CO2-e emissions against the thermal conductivity for Grade 32-35 MPa and 38-42 MPa

Figure 4-14 is a comparison between the thermal conductivity and embodied CO2-e emissions of the 32-35MPa and 38-42MPa groups of concrete. Note that mixes 27-41 have the lowest thermal conductivity and the highest embodied CO2-e emissions, whereas mix designs 10-26 have the lowest amount of embodied CO2-e emissions and the highest thermal conductivity in both groups. In group 38-42 MPa, the thermal

144

conductivities associated with mix design 10-26 do not vary much whereas the

3 embodied CO2-e emissions can range from approximately 200 to 400 kg CO2-e /m .

As discussed previously, the variations shown in Figures 4-11 to 4-14 are associated with changes in the quantity and type of aggregate, and the binder materials in the concrete mix designs. Moreover, a lower thermal conductivity suppresses the energy charging /discharging rates (Fan & Khodadadi 2011), which may have a positive potential effect on the overall energy performance of buildings compared to traditional concrete. Concrete with a low thermal conductivity results has a higher thermal resistance than conventional concrete, which can slow down the heat gain and energy losses for periods of time (Kim, HK et al. 2012; Zhou et al. 2012). However, the optimal range for thermal conductivity of a concrete mix must be able to reduce or escape from passive heating before being absorbed or re-released as a stored heat before the colder night (DIIS 2013). It is therefore essential to consider the environmental impact of concrete mix designs in a more holistic way during the structural design of buildings, and also include the estimated impact on energy performance during the operational phase and end of life (life cycle) of a building.

Future research will quantify the potential effect of conventional and novel concrete materials on the thermal performance of buildings.

4.7 Conclusion

A great deal of effort is being put into compiling reliable methodologies for quantifying the environmental impact of concrete production. Some of the embodied emissions databases (eTool, Crawford, BPIC) currently available propose using an individual embodied CO2-e emissions coefficient for each grade of concrete without

145

considering variations across different mix designs. The findings from this Chapter are consistent with the common literature and confirm that significant reductions in embodied CO2-e emissions can be achieved by using supplementary cementitious materials such as fly ash, and GGBFS. However, depending on the percentage of cement replaced, fly ash can typically help to reduce the embodied CO2-e emissions of concrete by 10 to 15% compared to Portland cement. Moreover, GGBFS can also reduce embodied CO2-e emissions by 15.5% compared to common Portland cement.

The analyses of embodied CO2-e emissions has shown variations across the different inventory databases which are due to different methods of analysis, the sources of data, and the quality of input data (related to upstream process) in the calculations. This highlights the need for transparency within existing and future databases and imposes a need to extend their capability to model concrete mix designs based on individual components.

When using the ICE database, the results for embodied CO2-e emissions were sensitive to the concrete mix design because the embodied CO2-e emissions coefficients in ICE varied according to the different percentages of cement, fly ash, and GGBFS. Indeed the analysis showed that the embodied CO2-e emissions of a mix design decreased by increasing the proportion of fly ash and GGBFS in the concrete binder. The ICE database was limited in that it could not account for the effects of silica fume and cenosphere in concrete admixture mixes, even though they can be accounted for by including the cenosphere as additional fly ash and considering silica fume as a zero contribution.

The inventory databases from Crawford and eTool use the same embodied CO2-e emissions coefficients for each grade of concrete without accounting for the effects of 146

each different concrete component. The calculated of embodied CO2-e emissions from the BPIC database used average industry values, which resulted in lower embodied

CO2-e emissions than those calculated by the Crawford and eTool databases.

However, the analysis based on AusLCI, Alcorn’s analysis, and embodied CO2-e emissions coefficients (Table 4-3) that were compiled for this study considered the detailed effects of materials in the concrete mix design; it revealed many variations of embodied CO2-e emissions in the concrete mix designs. However, there are some discrepancies between the results of this study results and the AusLCI analysis due to differences in the embodied CO2-e emissions factor for Portland cement, fly ash,

GGBFS, and the types of aggregates (recycled, natural and manufactured).

This Chapter also showed that the thermal conductivity of concrete is strongly related to the properties of the concrete mixes and the proportions of its constituents. In general, the thermal conductivity of a design mix increases with increasing density and replacing normal aggregates with lightweight aggregates reduces the thermal conductivity of concrete. These lower density concrete mixes with a low thermal conductivity could be beneficial in terms of energy saving during the operational phase of buildings, but lower density concrete mix designs could have higher embodied CO2- e emissions. Hence, it is crucial to understand and consider the thermal and environmental impacts associated with concrete mix designs in an integrated way and at the design stage of building.

The results of this Chapter can be used to consider reducing the environmental impact and improving the thermal conductivity of concrete while maintaining the desired strength during the early stages of building projects.

147

The uncertainty associated with calculating the life cycle environmental impact of a builing is addressed in the following Chapter (Chapter 5). Chapter 6 considers the potential impact of concrete mix designs on thermal mass and hence on the energy performance of a building over its operating phase.

148

Chapter 5

A method of uncertainty analysis for embodied CO2-e emissions impact of building materials in Australia

Contribution of the candidate to the Published Work

The contribution of the candidate in the published paper was 90%. He co-authored them with his main and co-supervisors. The candidate analysed and designed a case study building, analysis uncertainties associated with CO2-e emissions and wrote the paper. Other authors reviewed the papers and provided useful feedback for improvement.

Citation: Published

Robati, M, McCarthy, TJ & Kokogiannakis, G 2017, A method of uncertainty analysis for life cycle embodied CO2-e emissions of building materials in Australia, 28th Biennial National Conference of the Concrete Institute of Australia 2017, 22-25 October 2017.

149

5.1 Introduction

Chapter 4 presents the environmental impact analysis associated with different structural materials (different concrete mix designs); it revealed that the potential environmental impacts associated with the selection of structural systems and materials can be quantified by using different inventory databases and online tools.

The results in Chapter 4 identify some inconsistencies in the calculation of embodied

CO2-e emissions associated with concrete mix designs. This inconsistency between the tools when estimating the environmental impact can lead to large differences in the results of CO2-e emissions and thus the overall sustainability of buildings. The GHG emissions associated with each individual component of concrete and its production needs greater refinement. This Chapter quantifies the uncertainties associated with calculating the embodied CO2-e emissions in buildings. To achieve this end, the study assessed the impact of different structural materials and construction forms used in a typical 15-storey office building in Australia, as proposed by the National Standards

Development Organisation (NS11401.1 2014). This approach uses to quantify the uncertainties associated with calculating CO2-e emissions.

5.2 The need for an embodied CO2-e emissions analysis in construction

The construction industry is a major consumer of renewable and non-renewable natural resources. The construction of new buildings has substantial environmental costs; it is estimated that worldwide, the construction industry consumes 40% of total primary energy, 40% of natural materials, 15% of the world’s freshwater resources, and generates 25% of all wastes and 40–50% of greenhouse gas emissions (GHG)

(Ding 2014; Mokhlesian & Holmén 2012; Ramesh et al. 2010).

150

The use of low CO2-e emissions building materials to minimise the industry’s environmental impact has received increasing research and development attention. A holistic approach to the selection of low CO2-e emissions building materials should consider the entire material life cycle, including building performance and embodied energy (Berge 2009; Franzoni 2011). The material life cycle includes the extraction of raw materials, manufacturing processes, transportation to the construction site, construction processes, the operational phase, and the end of life recycling and potential for reuse (Ding 2014).

As buildings become more energy efficient, the operational phase of a life cycle assessment will make an increasingly smaller contribution to the total environmental impact, on the other hand material selection will become relatively more important

(Thormark 2006). However, selecting low CO2-e emissions building materials is a challenging task (Saghafi & Teshnizi 2011), because it requires an analysis of building materials embodying CO2-e emissions impact at all stages of the life cycle, as well as an environmental performance of the material as part of operational buildings. This is an ongoing area of research due to a large number of variables and the uncertainty involved in the assessment process.

As shown in section 2.11, there are a number of researchers attempting to quantify the uncertainties associated with the whole life energy performance of buildings

(Crawford 2011; Dixit et al. 2010), but there is still a significant gap in current research related to the uncertainty with the embodied energy of materials in the processing, manufacturing, and construction of buildings, relative to operational impacts and uncertainty.

151

Low CO2-e emissions building design has become the mainstream of research and development in order to minimise the industry’s environmental impact. Previous researchers have highlighted the key points of the sustainable selection of building materials and systems by considering the performance specifications and lowest GHG over the life cycle perspective of buildings at an early stage of design (Berge 2009;

Franzoni 2011). The building life cycle embraces the extraction of raw materials, manufacturing process, transportation to the construction site, construction process, operational phase, and the end of life recycling and reuse (Ding 2014).

5.3 Environmental life cycle analysis

An Environmental Life Cycle Analysis (LCA) is a method for identifying and evaluating the environmental aspects of a product through all its activities during its life (ISO14040 2006); this method assesses the materials used and energy released by the system into the environment. Applying a life cycle assessment to the building sector is a particularly complex LCA problem (Ortiz et al. 2009; Taborianski & Prado

2004) due in part to the complexity, size, and intensive use of natural resources in all stages of building (Sharma et al. 2011). The following factors introduce further complexity to LCA in this sector:

• Buildings have a particularly long lifetime, often more than half a century, so

it is difficult to predict the whole of lifetime effects of the project from cradle-

to-grave (Cabeza et al. 2014);

• The majority of environmental impacts associated with buildings traditionally

occurs during the operational phase, and these impacts can be reduced by good

design and material selection (Kibert 2012); 152

• During the lifetime of a project, the building may undergo many changes in

terms of form and function, changes which can be as significant as the original

construction (Stephan & Crawford 2014). Future changes can potentially be

considered at an early stage of design to minimise the environmental effects of

changes (Crawford 2011);

• There are many stakeholders and shareholders involved in the building

industry.

British Standard BS EN15978: Sustainability of construction works. Assessment of environmental performance of buildings. Calculation method proposes a number of methods for assessing the environmental performance of buildings. The standard proposed calculation method involves four stages in a life cycle assessment of buildings, these include: the product stage (raw materials, transport and manufacture); the construction process (transport, construction and insulation process); the use stage

(use, maintenance, repair, replacement and refurbishment), and the end of life

(deconstruction, demolition, transport, waste processing and disposal) (EN15978

2011). This system boundary includes the extraction of raw materials, production processes, transportation, and use and disposal.

A number of studies have found that the use stage (operational energy) accounts for

80% to 85% of a building’s life cycle energy consumption (Richman et al. 2009;

Sharma et al. 2011). The energy inputs for the production of building products, the extraction and processing of raw materials, and manufacturing and transportation to construction sites are responsible for the remaining 15% to 20% of whole life cycle energy usage of a building (Asdrubali et al. 2013). The contribution made by

153

construction work, transportation, and final demolition and disposal at the end of life is deemed as almost negligible, at level of approximately 1% (Ruuska & Häkkinen

2015; Sartori & Hestnes 2007).

To understand the role that building materials have on an energy efficient design, the operational and embodied energy implications of building design options must be investigated. Since the operational energy offers most opportunities for energy efficiency, the majority of previous research has focused on reducing it rather than considering all the stages of a building’s life cycle.

Several literature reviews have also highlighted the significance of building materials and embodied energy in a lifetime energy analysis of buildings (Asdrubali et al. 2013;

Asif et al. 2007; Dixit et al. 2010; Esin 2007; Kellenberger & Althaus 2009; Rincón et al. 2013; Vieira & Horvath 2008; Wu et al. 2005). The authors found that an appropriate choice of construction and building materials can reduce the embodied energy and embodied CO2-e emissions by 17 % and 30 %, respectively, over the lifetime of buildings (González & García Navarro 2006; Thormark 2006). Asif et al.

(2007) studied the life cycle embodied energy and air emissions associated with five commonly used materials (glass, aluminium, wood, ceramic tiles, and concrete) in a

Scottish residential house. Concrete was responsible for 60% of the total embodied energy in those buildings. Similarly, Ximenes and Grant (2013) used the life cycle assessment method to determine the GHG associated with several building materials in Australia and found that structural elements consisting of concrete and bricks are responsible for up to 31% and 17% of the total greenhouse impact, respectively. The use of timber in the sub-floor resulted in between 31% and 56% reductions in embodied GHG emissions. Aye et al. (2012) undertook LCA on three forms of 154

building construction common in Australian and showed that steel structured buildings reduce the consumption of material by almost 78% by mass compared to a concrete structure. However, the steel structure resulted in a 50% increase in embodied energy compared to the concrete structure. They concluded that an efficient use of materials could result in energy savings of up to 81% of embodied energy, and materials (51% by mass).

A number of previous studies identified the variation and inconsistency in the measurements of embodied energy (Buchanan & Honey 1994; Crawford 2013; Dixit et al. 2010; Huang et al. 2010; Langston & Langston 2008). Dixit et al. (2010) found these sources of uncertainty to be: variations in the method of analysis used in each assessment; different system boundaries; and the quality of data sources and input in the calculation of upstream processes. Accordingly, it is important to use methods to quantify the uncertainties associated with the life cycle analysis of buildings and construction materials. This Chapter therefore aims to quantify the uncertainty associated with the embodied CO2-e emissions of a case study building (stages A, B4 and C1-C2 as shown in section 2.8). The analysis of these uncertainties may reveal which inputs are responsible for variations in the embodied CO2-e emissions of buildings, and whether the uncertainties are significant when considering the whole of life environmental impact of buildings.

Following this introduction, the methodology section describes the tools used to analyse the uncertainty associated with the embodied CO2-e emissions of an office building in Australia. A detailed description in section 5.4 illustrates the embodied

CO2-e emissions intensity for different building materials and products, followed by a

155

discussion of the key role of four important materials selected in the overall embodied

CO2-e emissions of a building.

5.4 Methodology

This Chapter uses a probabilistic based method as the variability of dependent parameters (output variables) as determined by the uncertainty of the independent parameters (input variables). The following paragraphs summarise the workflow and methodology used to quantify the uncertainty associated with a lifetime environmental assessment of the structural design of a building.

This Chapter assesses the uncertainty associated with the embodied CO2-e emissions of a typical 15 storey reinforced concrete office building in Australia (NS11401.1

2014). Four commonly used structural floor systems were compared in terms of their embodied CO2-e emissions. The four structural frames are a Flat slab, Flat plate, Post- tensioned slab, and Waffle slab that were designed to meet two main AS3600 (AS3600

2009) considerations: strength and serviceability. The boundary for the CO2-e emissions environmental assessment of these four structural systems includes the embodied CO2-e emissions associated with construction materials from production, construction, and at end life activities (as shown in Figure 5-1).

156

Figure 5-1 Boundary study of Chapter 5

The sensitivity analysis can be grouped into screening, local, and global methods

(Heiselberg et al. 2009); where screening methods are used to evaluate a large number of design parameters; local methods are based on an OAT approach (One parameter

At a Time) which is useful for evaluating the relative importance of various design elements, and global sensitivity analysis is where all the parameters are varied at the same time the effect of range and probability density function is considered

(Heiselberg et al. 2009; Silva & Ghisi 2014). This Chapter uses global sensitivity and uncertainty analysis via Microsoft Excel function. Based on the literature reviewed, the mean and standard deviation are determined to generate subjective probability distributions for each individual building materials. The cumulative probability distributions as a result of Monte Carlo analysis reveals the variance associated with the embodied CO2-e emissions of the materials used in the designed buildings.

157

This Chapter has five stages (as shown in Figure 5-2).

1- List the input parameters and define their probable density functions;

2- Select the probability density function of the input parameters: the relative

embodied CO2-e emissions, lifetime and the transport distance were

extracted from published literature;

3- Perform a random sampling via Microsoft Excel normal distribution

function: the input parameters (embodied CO2-e emissions, lifetime and

transport distance) associated with each building material (Table 5-1) were

randomly generated 1000 times to achieve more accurate results (Inyim et

al. 2016).

4- Perform an uncertainty analysis: for each 1000 sample data, equation 5-1

was used to generate the probability distribution of all the input parameters.

The total result presents the global uncertainty analysis associated with

each of the four structural systems.

5- Perform a sensitivity analysis to quantify the magnitude of the change in

the estimated embodied CO2-e emissions of the building materials. In the

last step, the variability of each construction material was quantified and

compared against the total embodied CO2-e emissions of the building. This

stage quantified the relative importance of each building material by

considering their relative impact at each individual iteration over the total

iterations (1000). The results of this stage provide the magnitude variations

of building materials and their impacts on determining CO2-e emissions of

the building.

158

Figure 5-2 summarises the workflow and methodology used to quantify the uncertainty associated with a lifetime embodied CO2-e emissions assessment of the benchmark building.

159

Figure 5-2 the workflow and methodology used in this Chapter

160

The total embodied CO2-e emissions is calculated by adding the magnitude of each parameter through the use of Equation (5-1). Equation (5-1) represents lifetime (Cradle to Grave) CO2-e emissions associated with selection of the building materials and construction systems (Farrance & Frenkel 2014).

16 퐿푇 푄𝑖 푇퐶푂2−푒 = ∑ ( × ((푄푖 × 퐼푖) + ( × 퐼푡 × 퐷푖 ))) (5-1) 푖=1 퐿𝑖 퐶푇

퐿 Condition: 푇 ≥ 1 퐿𝑖

2 • 푇퐶푂2−푒 is the total embodied CO -e emissions associated with the building

materials (Kg CO2-e emissions); 퐿푇 represents the total lifetime of the

building, assumed to be 50 years (AS3600 2009);

th • 퐿푖 characterise lifetime associated to the i building material (number of

years); for a materials lifespan higher than 50 years (such as concrete, steel

퐿 reinformance, timber), the lifetime ratio ( 푇 ) is equal to 1; 퐿𝑖

th • 푄푖 represents the quantity of the i building material (Table 5-1);

th • 퐼푖 is the embodied CO2-e emissions associated with the i building material

(kg CO2-e /unit of material);

• 퐶푇 is related to the truck capacity, which can carry a 20ft container (volume

39 m3);

• 퐼푡 is the embodied CO2-e emissions associated with the truck used to

transport materials (excluding concrete) isassumed as 0.07155 (kg CO2-e

/tonne per km) (Moussavi Nadoushani & Akbarnezhad 2015).

th • 퐷푖 is the i distance the building material travels from the supplier to the

counstruction site (Kilometres). 161

Stage 2 of methodology considers the variations associated with the lifetime of materials, embodied CO2-e emissions, and the travel distance. The amount of variation is calculated based on collecting data from published literature to represent the mean and standard deviation values. The variations in the material’s lifespan came from published literature (Cabeza et al. 2014; Ding 2004; Ding 2008; eTool 2014; Furuta et al. 2014; Thormark 2006). The embodied CO2-e emissions coefficient associated with the building materials came from six inventory databases: BPIC (BPIC 2014), ICE

(Hammond et al. 2011), eTools (eTool 2014), Alcon (Alcorn 2003), AusLCI (AusLCI

2016), Crawford (2011), and other published literature (Moussavi Nadoushani &

Akbarnezhad 2015; Robati et al. 2016). The variations associated with the travel distance from material suppliers to the construction site are estimated by using Google maps(Poinssot et al. 2014). It is assumed that the building in this Chapter is located in the central business district of Sydney.

The spread of random numbers in stage 3 was predetermined by its specified mean and its specified standard deviation from stage 2. A normal distribution is recommended for modelling the variations associated with each input variable because the maximum and minimum CO2-e emissions values were not clear enough to define them (Inyim et al. 2016; Peña-Mora et al. 2009). It was therefore assumed that all the parameters

(lifetime, embodied CO2-e emissions and travel distance) associated with the building materials are distributed normally along the standard deviation (SD). So, the lifetime, the embodied CO2-e emissions of materials, and the travel distance between the material suppliers to the construction sites are distributed separately because each variable comes from different sources of data. A normal distribution is used because

162

when that other distribution (rectangular, triangular) is combined it often yields a net distribution which is close to normal (Farrance & Frenkel 2014).

The quantities of the parameters are related to the materials used in the building, and and the quantities of materials used in the building are derived from NS 11401.1

(NS11401.1 2014) as shown in Table 5-1. For the Flat plate, Post-tensioned, and

Waffle slab, the estimated quantities of structural materials (concrete and steel reinforcement) are based on a detailed structural analysis where the Excel spreadsheet and JMP (statistical software) (JMP 2017) are used for the uncertainty analysis.

163

Table 5-1 Uncertainty of physical parameters

Unit Quantity Distribution

Post Lifetime Embodied CO2-e Distance

reference building building reference (W.S) Waffle slab

Flat slab for the for Flat slab Flat plate (F.P) Flat plate (years) emissions (km)

-

tensioned

(F.S.R) (kg/unit of Building materials (i) material)

Mean Mean Mean

(P.T)

SD SD SD

Concrete Concrete on ground m3 250 250 250 250 85 47 210 137 25 16 1 (N20) Suspended concrete m3 2,775 3,005 2,994 2,002 85 47 258 122 25 16 2 (N32) Suspended concrete m3 124 124 124 124 85 47 298 130 25 16 3 (N40) Steel reinforcement 4 Steel in concrete tonnes 411 630 376 636 95 48 1.22 1.11 25 16 5 Tender tonnes ...... 1.4 ...... 95 48 1.22 1.11 25 16 Formwork 6 Timber formwork tonnes 261 11 11 ...... 90 46 1.87 0.29 25 46 7 Steel formwork tonnes 11.16 ...... 100 66 4.33 2.21 25 66 Plastic based tonnes ...... 4.6a 33 14 4.32 2.21 25 14 8 formwork 9 Plastic duct tonnes ...... 1.6b ...... 33 14 4.32 2.21 25 14 Roof 7 Timber battens tonnes 20.1 90 46 1.54 0.29 90 46 Roof plumbing tonnes 0.21 100 47 1.54 1 10 47 8 Steel eaves gutter 0 tonnes 0.1 100 47 1.54 1 10 47 9 Steel ridge flashing 0 1 tonnes 0.64 100 47 1.54 1 10 47 Steel fascia 0 0 1 tonnes 2.25 100 47 1.54 1 10 47 Steel downpipe 1 0 Ceiling and wall lining 1 tonnes 205.92 33 9 0.39 0.21 33 9 13 mm plasterboard 2 Insulation Roof and Ceiling tonnes 1.41 60 16 1.02 0.21 60 16 1 insulation: glass wool 3 celling batts 1 Wall insulation: glass tonnes 3.57 60 16 1.02 0.21 60 16 4 wool wall batts Vehicular doors 1 Steel doors and tonnes 0.18 100 47 1.95 1.02 10 47 5 mechanisms 0 Windows, doors and glazing Glass window (10 m2 4320.1 29 12 189.5 49.60 29 12 1 mm heat absorbing 6 float glass) F.S.R: Flat slab for the reference building in proposed span distance of 5.27 m and slab thickness of 185 mm (NS11401.1 2014); F.P: Flat plate in the reference building with span distance of 5.27 m and slab thickness of 200 mm; P.T: Post-tensioned slab in the reference building with span distance of 5.27 m and slab thickness of 175 mm. W.S: Waffle slab in the reference building with span distance of 5.27 m and a slab thickness of 250 mm (50mm topping + 200 stem depth). a. The High-Density Polyethylene weight was considered as 4.83 kg per each formwork item; b. The Plastic duct weight was assumed as 0.98 kg/m.

164

5.5 Results and discussion

The Monte Carlo uncertainty analysis represents embodied CO2-e emissions associated with different structural forms. These results are based on 1000 values of each parameter generated independently from normal distributions, with the mean and standard deviations as shown in Table 5-1. The standard deviation is estimated based on the variations associated with the Chapter parameters (lifetime, embodied CO2-e emissions and travel distance). For instance, Figure 5-3 presents the variations associated with the embodied CO2-e emissions coefficients for two grades of concrete

(N32 and N40); for concrete N32 and N40, the standard deviation and mean values are obtained from the 203 and 175 datasets, respectively.

Figure 5-3 Embodied CO2-e emissions variations for two grades of concrete (N32 and N40)

165

Figure 5-4 summarises the global uncertainties associated with the four different structural systems in this Chapter. Note that there are differences across the spread of data; the Waffle slab and Post-tensioned slab have the lowest mean embodied CO2-e emissions values compared to the flat plate and Flat slab, for the reference benchmark building. By adding up the embodied CO2-e emissions of the building materials, the total embodied CO2-e emissions associated with the proposed benchmark building

(shown as F.S.R in Figure 5-4) ranges from 2,106 to 9,664 tonne CO2-e emissions, with a mean value of 5,185 tonne CO2-e emissions. The results of different structural forms show that the Flat slab (for reference building benchmark building) has the highest mean embodied CO2-e emissions due to the thicker slab and higher volume of concrete.

166

F.S.R: Flat slab for the reference building in proposed span distance of 5.27 m and slab thickness of 185 mm (NS11401.1 2014); F.P: Flat plate in the reference building with span distance of 5.27 m and slab thickness of 200 mm; P.T: Post-tensioned slab in the reference building with span distance of 5.27 m and slab thickness of 175 mm. W.S: Waffle slab in the reference building with span distance of 5.27 m and slab thickness of 250 mm (50mm topping + 200 stem depth).

Figure 5-4 Probability distributions of predicted embodied CO2-e emissions for the four slabs

167

2 The variance associated with environmental impact (embodied CO2-e /m ) of each structural system is shown in Figure 5-5; it shows the small differences between the

2 2 embodied CO2-e /m of the four structural systems. The median embodied CO2-e /m of the buildings varies by up to 24% across the design alternatives. Reductions in embodied CO2-e emissions can be achieved in most cases by selecting the Waffle slab systems because Waffle slab provides a lighter and stiffer slab than an equivalent flat plate slab (CC 2016; CCAA 2003). Furthermore, the post-tensioned slab (P.T) reduces

2 the mean embodied CO2-e emissions by 51 kg CO2-e /m compared to a flat plate slab.

The post-tensioned slab is thinner and has a lower volume of concrete and steel reinforcement compared to a conventionally reinforced slab (F.P and F.S.R). The maximum embodied CO2-e emissions did not vary much between W.S and P.T, although the embodied CO2-e emissions associated with W.S was 2% (equivalent to

2 the 10 Kg CO2-e /m ) higher than the P.T slab.

W.S

P.T

F.P Structural forms Structural

F.S.R

0 100 200 300 400 500 600 700

Embodied CO2e associated with different structural forms (kg CO -e/m2) 2

Figure 5-5 Embodied CO2-e emissions associated with different structural

forms

168

The sensitivity analysis quantified the magnitude of the changes in the estimated embodied CO2-e emissions of the building materials (variables) across the global uncertainties. The magnitude of variations associated with each building material was estimated by variations of their embodied CO2-e emissions to the total embodied CO2- e emissions for each 1000 iterations. Figure 5-6 provides the probability associated with the environmental impact (embodied CO2-e emissions) of the materials used in the buildings. The cumulative probability gives a reasonable impression of the overall impact of the variations in embodied CO2-e emissions calculations. Note that the structural materials (Concrete and steel reinforcement) and architectural elements

(windows with aluminium frame) account for highest proportions of embodied CO2-e emissions.

Suspended concrete (N32)

Window

Steel reinforcement

Concrete on Ground (N20) Variables Timber formwork

Plasterboard

Suspended concrete (N40)

0% 10% 20% 30% 40% 50% 60% 70% 80% 90%

Probility of output data variable i

Figure 5-6 Probability variation associated with the environmental impact

(embodied CO2-e emissions) of the materials used in the designed buildings.

The variations of cumulative probability of the output data show significant variations in the amount of embodied CO2-e emissions associated with types of materials. By 169

considering all four structural systems, the level of uncertainty associated with different types of concrete ranges from 6% to 78% of the total embodied CO2-e emissions of the building. For example, the statistical distribution of the data associated with concrete (N32) has interquartile ranges between 42% and 59% with a mean value of 51%. Note there is a considerable variation in the level of uncertainty associated with the embodied CO2-e emissions of the double-glazed aluminium framed window, so the amount of uncertainty associated with embodied CO2-e emissions from the double-glazed aluminium window ranged from 16% to 25%, with central tendency mean uncertainty of 20%.

For concrete, steel reinforcement, and aluminium framed windows, the uncertainties are mainly from variations of the embodied CO2–e coefficient proposed by different inventory databases. For instance, the amount of embodied CO2-e emissions for

3 concrete changed from 62 to 562 (kg CO2-e /m material) in different inventory databases (eTool 2014; Hammond & Jones 2008; Robati et al. 2016). The CO2-e emissions impact associated with windows was sourced from databases as 216 to 279

2 (kg CO2-e emissions)/m (eTool 2014; Hammond et al. 2011). The previous studies confirm that steel, concrete, and glass are the most intensive sectors in the production of typical Australian office buildings due to the various methods and quantities of materials in the upstream process of production as well as large amount of fossil fuels used in upstream process of concrete (Crawford 2011; Yu et al. 2017).

Regardless of the structural types, the proper selection of concrete and steel reinforcement can have a huge influence on the overall embodied CO2-e emissions of buildings.

170

In terms of variations in embodied CO2-e emissions of concrete, previous studies found that Portland cement is a major source of greenhouse emissions during the production of concrete (Kajaste & Hurme 2016). The concrete industry is addressing some of the concerns about environmental problems by adding or replacing the cement with other components related to high embodied CO2-e emissions. Previous studies have shown the results of replacing cement in the concrete with alternative cementitious materials (de Castro & de Brito 2013; Ingrao et al. 2014; Jacoby &

Pelisser 2015; Robati et al. 2016). The most commonly used cementitious materials are Ground Granulated Blast Furnace Slag (GGBFS) and coal combustion fly ash.

GGBFS is a by-product of iron and steel making and fly ash is a by-product of burning coal, mainly for generating electricity. The production process of fly ash and GGBFS involves less greenhouse gas emissions than ordinary Portland cement (Robati et al.

2016; Van den Heede & De Belie 2012). For the structural design of buildings, substituting a portion of Portland cement with GGBFS could substantially reduce the negative environmental impact of concrete, but the production of GGFS is limited and therefore the effects of further shipping must be considered in the calculation of embodied CO2-e emissions.

5.6 Conclusion

A Monte Carlo simulation method has been used to examine and predict the impact of uncertainties on the embodied CO2-e emissions of a proposed benchmark building.

The probability distributions of the most effective building materials (input data) were obtained to estimate the mean (expected) embodied CO2-e emissions value associated with each of the structural design systems. The results highlight the contribution and variation of each construction material and structural system from the extraction of 171

raw materials to the construction site and end of life building. The differences between the embodied CO2-e emissions of different structural systems are not significant. The

Waffle slab (W.S) has the lowest amount of embodied CO2-e emissions per floor area

2 and the Flat plate (F.P) system has the highest amount of CO2-e /m . However, once a structural frame has been chosen, there are significant opportunities for reducing the

CO2-e emissions impact of the building by optimising the quantities of structural materials (concrete and steel reinforcement). Concrete, double glazed windows with an aluminium frame and steel reinforcement have the maximum proportion of embodied CO2-e emissions over a 50-year lifetime of a building. The maximum percentage contributed by suspended concrete (N32 used for flooring), windows and steel reinforcement on the overall embodied CO2-e emissions of designed building structures as high as 78%, 57%, 48%, respectively. An important reason for the variations in these results is the inconsistency in the inventory of data. The results of this Chapter can be used as a guideline and reference point for future comparisons of the environmental impact associated with buildings, materials, and systems. The potential impact of the structural system on the operational energy and whole life environmental costs is addressed in Chapter 6 and Chapter 7, respectively.

172

Chapter 6

Impact of structural design solutions on the energy and thermal performance of an Australian office building

Contribution of the candidate to the Published Work

The contribution of the candidate in the published paper was 90%. He co-authored them with his main and co-supervisors. The candidate modelled, simulated a case study building for energy analysis and wrote the paper. Other authors reviewed the papers and provided useful feedback for improvement.

Citation: Published

Robati, M, Kokogiannakis, G & McCarthy, TJ 2017, ‘Impact of structural design solutions on the energy and thermal performance of an Australian office building’, Building and Environment, vol. 124, no. Supplement C, pp. 258-82.

173

6.1 Introduction

The previous Chapter used Monte Carlo analysis to quantify the uncertainties associated with calculating the CO2-e emissions from extracting and transport raw materials to the construction site and end of life activities. This current Chapter investigates the impact of selecting structural materials (Normal weight and Ultra- lightweight concrete) and construction forms during the operational phase of a building.

Concrete is a heavyweight construction material whose high thermal mass could increase the thermal storage capacity of a building envelope and in turn affect indoor thermal comfort. Selecting an appropriate method for concrete construction and form could also affect the total energy performance and thermal comfort of a building, a fact that is often overlooked by structural engineers. This Chapter presents the results of energy simulations of the potential impact that concrete construction forms and structural materials have on the energy consumption of archetypal commercial office buildings in five major Australian cities (Sydney, Melbourne, Canberra, Brisbane and

Darwin). This Chapter has three stages: 1) a structural analysis of two construction forms (Flat and Waffle slab); 2) the selection of two types of structural concrete

(conventional Normal weight concrete and novel Ultra-lightweight concrete); 3) a comparative analysis to quantify the magnitude of the change in predicted annual energy consumption due to changes in the form of construction and the type of structural concrete. This Chapter provides the underlying approach and results of the first simulation-based assessment that ULWC has on the energy and indoor comfort of commercial buildings.

174

6.2 The role of structural design in the energy performance of buildings

The structural design of buildings is traditionally limited to material specifications and structural efficiency, whereas structural engineering research often attempts to provide structural efficiency by reducing the materials and resources used while increasing the longevity of structures through design. However, with the aim continuous innovation in the structural design of buildings a new model provides a framework to integrate the long-term effects of materials and systems into the design process; indeed, modern integrated structural design could utilise life cycle assessment tools to determine the whole life environmental performance of building design because life cycle energy assessments promote a more efficient use of materials and energy.

The appropriate choice of construction and building materials can potentially reduce the life cycle energy of buildings because materials with low thermal conductivity help to reduce the demand for energy as well as the associated greenhouse gases (GHG)

(Torgal & Jalali 2011). For instance, concrete is one of the main construction materials with the ability to absorb and retain energy for a long period of time; action that reduces energy consumption by storing heat in a natural daily cycle (thermal mass).

The mass components reduce temperature fluctuations in building spaces and thus reduce the associated peak heating or cooling loads (Torgal & Jalali 2011). Previous studies indicate that the thermal conductivity of concrete varies across Normal,

Lightweight, and Ultra-lightweight concrete (Marinkovic et al. 2010; Robati et al.

2016; Wu et al. 2015; Yun et al. 2013; Zhang & Poon 2015); this variation in density stems from changes in the proportion and type of aggregates, and the cementitious materials in the concrete mixture.

175

Normal weight concrete with a density between 2,200 to 2,600 kg/m3 includes cement, normal weight aggregates, and water, whereas lightweight concrete (1,350 to 1,900 kg/m3) is produced by replacing some of the solid materials in the mix with air voids

(Neville 2012). There are three possible locations for the air voids, inside the particles of aggregate, inside the cement paste, and between the coarse aggregate particles

(Neville 2012). The potential for substituting ordinary Portland cement with geopolymer materials in Lightweight concrete has been studied extensively by researchers (Abdullah et al. 2012; Robati et al. 2016). Geopolymer concrete is synthesised by mixing aluminosilicate material, alkali solutions, and water (Nuaklong et al. 2016). Also, the potential use of Lightweight hollow spheres in the design mix is a technique for producing Ultra-lightweight concrete (1,154 to 1,471 kg/m3); in fact, ultra-lightweight concrete consists mainly of lightweight hollow spheres (cenosphere materials), water, and a binder (it also includes silica fume and Portland cement)

(Robati et al. 2016; Wu et al. 2015).

The thermal properties of a concrete mix are influenced by the thermal properties of ingredients such as cement, aggregates, and the moisture existing in the mix (ACI122R

2014). The replacement of normal aggregate with lightweight aggregates reduces the density and thermal conductivity of concrete. A brief review of previously published values (Table 6-1) shows that the estimated thermal conductivity of Normal,

Lightweight, and Ultra-lightweight concrete could vary from 3.1 W/mK to 0.28

W/mK (Berndt 2015; CCAA 2015; Damdelen et al. 2015; Marinkovic et al. 2010;

O'Moore & O'Brien 2009; Robati et al. 2016; Tošić et al. 2015; Wu et al. 2015; Yun et al. 2013; Zhang & Poon 2015).

176

Table 6-1 Thermo-physical and structural properties of concrete classes as reported in the literature

Thermal Compressive References Density (kg/m3) conductivity Type of concrete strength (MPa) (W/mK)

Normal, Lightweight Wu et al. (2015) 966 – 2,251 33 – 69.4 0.28 – 1.98 and Ultra- Lightweight Blanco et al. 1,090 – 1,510 5.04 – 33.03 0.46 – 0.69 Lightweight (2000) Uysal et al. Normal and 1,329 – 2,270 NA 0.77 – 1.45 (2004) Lightweight Topçu and Uygunoğlu 880* – 1,500 3* – 9* 0.13 – 0.52 Lightweight (2007) Gül et al. (2007) 1,773 – 1,984 11.3 – 25.1 0.81 – 1.22 Lightweight Mounanga et al. 728 – 2,109 1.4 – 24.3 0.22 – 1.49 Lightweight (2008) Tandiroglu 1,798 – 1,883 60 – 80 1.46* – 1.76* Lightweight (2010) Sengul et al. 392 – 1,937 0.1 – 28.8 0.13 – 0.6 Lightweight (2011) Normal and Kim et al. (2012) 1200* – 2,350* 9* – 40* 0.32* – 0.72* Lightweight Wang and Meyer 1560-1980 18*-36.5* 0.27 – 0.61 Lightweight (2012) Huang et al. 1649 - 2001 23.33* – 48* 0.29 – 0.37 Lightweight (2013) Yu et al. (2013) 1280 - 1490 23.3 – 27.5 0.49 – 0.85 Lightweight Normal and Gao et al. (2014) 950* - 2,063* 7.67* – 62.78* 0.23* – 1.97* Lightweight Normal and Yun et al. (2013) 17,44 – 2,370 23 – 43.9 1.30* – 2.25* Lightweight *Extracted from graphs

These studies find that lower density concrete has a lower thermal conductivity, so modern concrete such as Lightweight and Ultra-lightweight concrete has better thermal buffering than traditional concrete (Normal weight concrete), as shown in

Figure 6-1.

177

2.5 R² = 0.6172 2

1.5

1

0.5 Thermal conductivity (W/mk) Thermalconductivity 0 0 500 1000 1500 2000 2500 Density (kg/m3)

Wu et al. (2015) Blanco et al. (2000) Uysal et al. (2004) Gül et al. (2007) Topçu and Uygunoğlu (2007) Mounanga et al. (2008) Tandiroglu (2010) Sengul et al. (2011) Wang and Meyer (2012) Huang et al. (2013) Yu et al. (2013) Yun et al. (2013) Gao et al. (2014)

Figure 6-1 Relationship between thermal conductivity and density

Several other studies have shown that buildings with a high thermal mass require more time to heat up and cool down, which might influence thermal comfort and demand more energy for heating and cooling (Hoes et al. 2011; Roberz et al. 2017).

Moreover, the ongoing development of more novel construction materials such as

Ultra-lightweight concrete (Huiskes et al. 2016; Roberz et al. 2017; Yu, QL et al.

2015) raises a question about their potential impact on the thermal mass of a building and hence on the overall energy performance of a real building during its operational phase.

Therefore, the primary objective of this Chapter is to indicate how the selection of concrete as a construction material affects the overall energy performance of a building. This Chapter explores a benchmarking method to evaluate the potential effects of conventional (Normal weight) and novel concrete materials (Ultra- lightweight) on the thermal performance of typical office buildings in Australia. A benchmark building serves as a framework to compare design alternatives in terms of

178

their energy performance. The benchmarking system in this study considers the different climate zones in Australia, the forms of construction (Flat and Waffle slabs), and the structural materials (conventional and novel concrete).

This research is organised as follows. Section 6.3 summarises the method used to design the structure and simulate the thermal performance of the benchmark office building. Section 6.4.1 provides the structural design and analysis results; Sections

6.4.2 and 6.4.3 compare the results of the energy performance for different structural materials and construction forms, and Section 6.5 reports the key findings of this

Chapter.

6.3 Methodology

6.3.1 Description of base building

This Chapter assesses the thermal performance of concrete materials (Normal weight and Ultra-lightweight concrete) and structural forms (lightweight and heavyweight) for a benchmark office building in Australia. This 15 storey office building is one of four benchmarking buildings proposed by the National Standard Organization

(NSDO) in Australia (NS11401.1 2014); This particular15 storey office building is a typical concrete structure (NS11401.1 2014), with a square plan shape, a total floor area of 1000m2, and an average 3.3 m height per storey, as shown in Table 6-2.

179

Table 6-2 Overall specifications of the benchmark building

Parameter Unit Specification

Basement dimensions m 31.62 × 31.62

Number of Stories --- 15

Concrete slab on ground mm 200

Concrete suspended slab mm 175

Average elevation per floor m 3.3

Total floor Area (including parking, Stairs & Verandas) m2 15,000

Total habitable area (external dimensions) m2 8,807.1

Total habitable area (internal dimensions) m2 962.4

No of floors above ground level --- 11

No of rooms --- 176

This building has two parts; the first three underground storeys are parking and storage areas, while the remaining twelve storeys are open plan office areas. The building has non-openable windows, with a thermal transmittance (U value) for the base case of

5.7 W/m2K and a Solar Heat Gain Coefficient of 0.6 (BZE 2013). A sketch of this office building is shown in Figures 6-2 and 6-3.

180

Figure 6-2 Plan of case study Figure 6-3 Section view of case study building (NS11401.1 2014) building (NS11401.1 2014)

6.3.2 Structural design parameters

In terms of structural analysis and design, a concrete structure design is considered to account for lightweight and heavyweight structures if they follow the Australian

Standards Concrete structures (AS3600 2009); the lightweight structure is designed as a Waffle slab and the heavyweight structure is based on a Flat slab. Flat slabs are very adaptable elements that are generally used to provide minimum depth and flexible column grids in construction, whereas Waffle slabs are a lighter and stiffer slab than the equivalent Flat slab. A Waffle slab has a thin topping and narrow ribs spanning in both directions between the column heads and/or beam band. The strength and serviceability aspects of the code were utilised during the design of this building. The process for structural analysis is summarised in Figure 6-4.

181

Figure 6-4 Structural analysis & design flow

The amount of live load comes from the Australian and New Zealand Standard for imposed actions (AS/NZ1170.0 2002). The live load for the office storage and parking areas was 5kPa and 3kPa for the work rooms. The dead load for concrete elements

(columns, shear walls, slabs and staircase) was obtained by multiplying the volume of the member by the unit weight of concrete. Wind loads on the building were determined in accordance with Australian and New Zealand standard wind actions

(AS/NZ1170.2 2011). The magnitude of wind pressure on the structure was calculated based on its height above ground, its size, importance, and location. The level of importance is level 3, because the consequence of failure is deemed to be high (based 182

on occupancy and by using AS 1170 (AS/NZ1170.0 2002)). For ultimate limit states and structural serviceability, the annual probability exceedance comes from AS 1170

(AS/NZ1170.0 2002), table 3.1 for a design working life of 50 years in a cyclone zone in Australia. To calculate the wind load, zone D was considered to be enough strength in the structure as well as validating the practicality of building in other zones. With the loading conditions, a combinations of action loads were used to check the serviceability and strength of the building in accordance with clause 4.2.1 and 4.2.2 of the AS1170 (AS/NZ1170.0 2002), as shown in Table 6-3. The Computer Aid

Design package Etabs, Safe and Microsoft Excel spreadsheet were used to verify the minimum requirements of the concrete design code. A summary of the structural analysis is shown in Appendix B.

Table 6-3 Loading conditions for design the building

Load Type of load (kPa) Live load-Office storage and parking area 5 Live load-Work rooms 3 Dead Load 4.3 Ultimate limit states 6.6 Wind Load- Windward Serviceability limit states 5.4 Ultimate limit states 4.1 Wind Load- Leeward Serviceability limit states 3.4 Ultimate limit states 1.3 Wind Load- Sidewall Serviceability limit states 1.1 Load combinations for Load combinations for Ultimate state design serviceability state design 1.35G 1.25G+1.5Q G+ Ψl Q 1.25G+1.5ΨlQ G+ Ψs Q 1.2G+Wu+ΨcQ G+ ΨsQ + Ws 0.9G+Wu G: permanent action (dead load); Q: Imposed action (Live load); Wu: ultimate load action; Ws: serviceability wind action; Ψl: Factor for determining quasi-permanent values (long term) of actions; Ψs: Factor for determining quasi-permanent values (long term) of actions; Ψc: Combination factor for imposed action; 183

6.3.3 Structural materials

This Chapter analyses the effects choices of concrete (normal and low-density) have on the thermal performance of a heavyweight and lightweight office structure. For the purpose of this Chapter, the types of concrete mixes were collected from previously published journal papers and databases (CCAA 2015; O'Moore & O'Brien 2009; Wu et al. 2015; Yun et al. 2013). These designs represent conventional (Normal weight) and some advanced methods of concrete admixture that give Ultra-lightweight concrete. Table 6-4 summarises the properties and grade of the concrete analysed in this Chapter. Novel forms of concrete admixture (such as Ultra-lightweight) are included in this Chapter to point out their potential effects on the thermal impacts of the building; they have not yet been covered in the mainstream of previous studies.

Table 6-4 Properties of selected concrete

Thermal Specific Grade Density Type of Concrete conductivity heat Source (MPa) (Kg/m3) (W/mK) kJ/(kg.k) N40- Normal weight 1 40 2393 1.96 0.88 (CCAA 2015)

N40- Ultra-lightweight 1 40 1400 0.31 0.88 (Wu et al. 2015)

(O'Moore & O'Brien N32- Normal weight 2 32 2470 2.10 0.88 2009) N32- Ultra-lightweight 2 32 1164 0.28 0.88 (Wu et al. 2015)

N20- Normal weight 3 20 1483 1.38 0.88 (Yun et al. 2013) 1. Grade N40 used in the vertical structural elements such as columns and shear walls. 2. Grade N32 used in the slabs (Waffle and Flat). 3. Grade N20 used in the other concrete element (staircase).

6.3.4 Operational energy analysis

Heavyweight (Flat slab) and lightweight (Waffle slab) structures were modelled and compared for their impact on the energy performance of the building by using the

DesignBuilder energy simulation software. DesignBuilder is a user interface for the

184

EnergyPlus dynamic thermal simulation engine and requires hourly weather data as inputs. The weather data used for each city in this Chapter was extracted from the

EnergyPlus weather database (EnergyPlus 2017). The weather data are in RMY format, they are a set of weather files developed to comply with the Building Code of

Australia (EnergyPlus 2017).

The equipment and occupancy schedules were extracted from the Building Code of

Australia (ABCB 2015). The schedules assume 10% of office equipment and 10% of lights remain on during unoccupied hours. The HVAC system was modelled using a variable air volume system (VAV) with the autosize routine in DesignBuilder’s

“simple” HVAC description (DesignBuilder 2017). Table 6-5 summarises the main assumptions used for the simulations.

Table 6-5 Simulated assumptions for benchmark building

Parameters Key variables References Lighting power density 9 (W/m2) (ABCB 2015) Occupancy density 10 (m2/person) (ABCB 2015) Equipment load 15 (W/m2) (ABCB 2015) Domestic hot water 0.4 (L/m2) (ABCB 2015) Infiltration 0.28 (ACH) (Egan 2011) Ventilation requirements 10 (L/s/person) (ABCB 2015) HVAC set point 18°C (heating) - (ABCB 2015) 26°C (cooling) *The schedules were extracted from Building Code of Australia (ABCB 2015)

This Chapter used the Building Code of Australia (BCA) “deemed to satisfy” approach to define the envelope construction of the modelled building (as shown in Table 6-6).

To understand the relative magnitude of the change in predicting energy consumption due to changes in the form of construction and type of structural concrete, the office building was modelled in four different ways: 1) as a Flat slab with Normal weight; 185

2) as a Flat slab with Ultra-lightweight concrete; 3) a Waffle slab with Normal weight

concrete; and 4) a Waffle slab with Ultra-lightweight concrete. The vertical elements

(columns and shear walls) consist of concrete with grade N40, the slabs (Waffle and

Flat) contain N32 and the other elements (staircase) are made of N20. The modelling

results for all four buildings revealed the total energy usage as well as the heating and

cooling loads across different input parameters (design alternatives). The total energy

consumption was compared to national and state averages determined from real world

data from Australian office buildings to ensure the results are within reasonable ranges

of the published and predicted energy consumption values (Pitt&Sherry 2012).

Table 6-6 Physical properties of benchmark building

Thermal resistance requirements and values and thermal mass values R-values Elements Item description References (m2.K/W) 1.Indoor air film (still air) Ground 2.Solid concrete (150 mm, (ABCB 1.25 3 floor 2400 kg/m ) 2015) 3.Ground thermal resistance 1.Indoor air film (still air) 2.Solid concrete (Study parameters) a. Flat with Normal weight a.1.25 concrete Intermediate b.1.81 b. Flat with Ultra- (ABCB floors c.1.22 lightweight concrete 2015) d. 1.63 c. Waffle slab with Normal weight concrete d. Waffle slab with Ultra- lightweight concrete 3.Outdoor air film (7 m/s) 1.Outdoor air film 2. Roof Water Proofing Membrane 3.Solid concrete, (Study parameters) a.4.20 a. Flat with Normal weight Based on b.4.84 concrete BCA c.4.17 b. Flat with Ultra- Roof requirments d. 4.58 lightweight concrete (ICANZ c. Waffle slab with Normal 2010) weight concrete d. Waffle slab with Ultra- lightweight concrete 4,5. Reflective Insulation Material R value 6. Reflective Air Space 186

7. Ceiling Insulation (125 mm) 8. 10mm Plasterboard 9. Indoor Air-Film (Non- Reflective Surface) 1.Outdoor air film 2. 8mm Compressed Fibre Cement Sheet 3. Reflective Insulation R- value 4. Unventilated 90mm Air Based on Space BCA External 5. Bulk Insulation Wall Batt 3.42 requirments Wall (90mm) (ICANZ 6. Reflective Insulation 2010) Material R-value 7. Unventilated Air Space 8. 110mm Brickwork 9. 10mm Plasterboard 10. Indoor Air-Film (Non- Reflective Surface) U value was taken as 5.80 W/m2K from the published literature (BZE 2013; Daly et al. 2014; Guan 2009) for single 6 mm clear glass, which is a common glass type for Window office buildings in Australia.

6.4 Results and discussion

6.4.1 Structural analysis and design

The office benchmark building has been structurally designed based on Australian

Standards in order to verify whether it can be used for realistic comparisons. The structural design specified heavyweight and lightweight alternatives for the Flat slab and Waffle slab construction. The structural analysis and design quantified the minimum size of the slab and column for each form of construction. The columns were classified into five (5) different groups based on their cross section and reinforcement details (Appendix B). The columns at the lower level have a larger cross sectional area and a higher ratio of steel than the upper columns. The dynamic lateral forces (earthquake) are excluded from the scope of this Chapter because the wind pressure loads are much more critical than earthquakes in most parts of Australia. The structural design is summarised in Table 6-7 (the structural design is shown in

Appendix B). 187

Table 6-7 Summary of the structural design

Construction form Flat slab Waffle slab Column span distance (L) 5.27 m 5.27 m Slab thickness (D) 200 mm 250 mm N20 250 250 Concrete N32 3,005 2,002 quantities (m3) N40 124 124 Steel quantities (Tonne) 753 679

Cross section

6.4.2 Energy performance of the building (Energy consumption)

Five major locations were selected for five major Australian cities and the heating and cooling hours are shown in Figure 6-5. The heating and cooling hours are calculated based on the differences between the outside weather temperature and a reference temperature which considered less than 18 degrees Celsius for heating and more than

24 degrees Celsius for cooling (BOM 2011). Darwin is located in climate zone 1, so it has a perennially hot climate with the highest number of cooling hours (Hot humid summer & warm winter). Brisbane has the second highest cooling degree hours and is

(climate zone 2) having a subtropical climate with warm, humid summers and mild winters. Sydney’s climate is influenced by abundant sunshine over the summer and a mild winter (climate zone 5) that results in higher heating degree hours than Brisbane.

Melbourne and Canberra have high heating demand compared to the other cities.

Melbourne has a temperate climate with changeable weather conditions in the spring and summer seasons (climate zone 6). Canberra is a cool temperate climate zone, with the highest heating degree hours over a year of the five climates examined in this

Chapter.

188

200 Darwin

180

) C

° 160

140 Hundreds 120 100

80 degreehours (CDHs)

temperature ≥26 temperature 60 Melbourne 40 Canberra Brisbane

base Sydney (

Cooling 20 0 0 200 400 600 Heating degree hours (HDHs) Hundreds (base temperature ≤ 18°C)

Climate zones: Darwin (1); Brisbane (2); Sydney (5); Melbourne (6); Canberra (7)

Figure 6-5 Summary of the annual heating and cooling degree-hours

The simulated annual energy consumption is compared and verified with the average national energy usage across the five major climate zones, as shown in Figure 6-6.

The Australian national average for commercial building energy consumption is

272±17 [kWh/m²], with a standard deviation of 128 [kWh/m²] per year (Bannister

2004; Daly et al. 2014), and the simulated outputs from this Chapter are within these ranges. The results of the simulated building energy performance showed that in this type of highly glazed office buildings, the cooling load is much higher than the heating load across all five climates studied.

189

National annual energy Waffle.low 200.low Waffle.normal 200.normal consumption intensity

Darwin [kWh/m2]

Brisbane

Sydney Majorcities Canbera National average: 270.6 [kWh/m2]

Melbourne

0 50 100 150 200 250 300 350 Energy consumption intensity [kWh/m2 per annum]

After Bannister (2004) Waffle.low: lightweight structure (Waffle slab) with Ultra-lightweight concrete; Waffle.normal:

lightweight structure (Waffle slab) with Normal weight concrete; 200.low: heavyweight structure

(200mm Flat slab) with Ultra-lightweight concrete; 200.Normal: heavyweight structure (200mm Flat

slab) with Normal weight concrete.

Figure 6-6 Predicted annual energy consumptions and national energy average usage across five major climate zones

The energy consumption across all five climates shows that the lightweight office building (called Waffle.low) with a lower thermal conductivity concrete (Ultra- lightweight concrete) demanded more energy than the other buildings because its fast response to temperature and heat flux excitations caused it to overheat for most of the year. The energy consumption predicted for the heavier type of office building (Flat slab using Normal weight concrete) was consistently lower than the buildings with

Ultra-lightweight concrete (Waffl.low and 200.low). Figure 6-7 shows a comparison between the cooling energy requirements of the building with different constructions

(Flat and Waffle slab) and different types of concrete. Note that the cooling energy requirements of the buildings were affected by the quantity (lightweight and heavyweight structure) and type of concrete (Normal weight and Ultra-lightweight) used in the building. Ultra-lightweight concrete had a greater effect on the demand for

190

cooling energy than for colder climates; for example, the office buildings with Ultra- lightweight concrete in Melbourne demanded up to 14% more cooling energy than the heavyweight structure (Flat slab) with Normal weight concrete.

When Normal weight concrete was used there was no great difference between the demand for cooling energy by buildings with heavyweight and lightweight structures.

However, the simulations for the building with Ultra-lightweight concrete showed that the cooling energy needed by the heavyweight structure (200.low - Flat slab) were less than the lightweight structure (Waffle.low - Waffle slab) across all five climates, albeit the differences were only between 2-3 kWh/m2 per annum.

Waffle.low: lightweight structure (Waffle slab) with Ultra-lightweight concrete; Waffle.normal: lightweight structure (Waffle slab) with Normal weight concrete; 200.low: heavyweight structure (200mm Flat slab) with Ultra-lightweight concrete; 200.Normal: heavyweight structure (200mm Flat slab) with Normal weight concrete.

Figure 6-7 Comparison between the annual energy requirements, structural forms and construction materials of the office buildings

191

6.4.3 Analysis of thermal performance

The results of sub-hourly dynamic simulations were analysed with no active heating/cooling system being used (free-floating conditions) in order to compare the effects of the different building models in terms of indoor temperature during the summer and winter seasons. To reduce the quantity of data for this Chapter, representative periods taken from the set of simulations were analysed with reference to winter and summer seasons (as shown in Table 6-8). In Australia, the summer and winter seasons are defined from December to February (for climate zone 1, the hottest season starts from mid-November) and June to August, respectively.

Table 6-8 Summer and Winter design weeks for the climate zones (DesignBuilder 2017)

City (Climate Zone) Winter design week Summer design week

Darwin (1) 10 to 16 Jun 19 to 25 November

Brisbane (2) 3 to 9 August 17 to 23 February

Sydney (5) 20 to 26 July 3 to 9 February

Melbourne (6) 6 to 12 July 27 January to 2 February

Canberra (7) 8 to 15 July 1 to 8 January

The indoor air temperature simulated hourly for the top floor was plotted against the hourly outdoor temperature to compare the indoor thermal performance across the different types of construction (as shown in Figure 6-8 and Appendix C). Indoor air temperatures were plotted against outdoor air temperatures for all four types of construction types, and show that those buildings with Normal weight concrete had a lower slope of regression in response to fluctuations in the outdoor air temperature,

192

whereas the buildings with Ultra-lightweight concrete had a higher regression coefficient.

Figure 6-8 Hourly air room temperature plotted against the hourly outdoor temperate for the Waffle.low and 200.normal in the climate zone 2 (Brisbane).

The hourly free floating analysis for the buildings with selected constructions shows how the structural mass and type of concrete affected the daily peak indoor temperatures. Table 6-9 summarises the differences in the peak daily indoor air temperature between the highest and lowest structural mass and concrete density 193

(200.normal and Waffle.low, respectively). Note that the peak indoor temperatures are higher in those building with Ultra-lightweight concrete and lower structural mass

(Waffle.low). For instance, the mean differences in the peak indoor air temperature between the Waffle.low and 200.normal cases (both located in climate 2) in summer and winter are 1.1 and 1.0°C respectively.

Table 6-9 Differences in the peak daily indoor air temperature between Waffle.low and 200.normal

Year Summer season Winter season

Annual Number of Samples (days) Annual Number of Samples (days) Annual

Number of Samples mean of Standard Deviation mean of Standard Deviation mean of Standard Deviation peak daily peak daily peak daily

(days) indoor air indoor air indoor air

temperature temperature temperature difference difference difference

[°C] [°C] [°C]

Canberra

0.9 0.64 365 1.2 0.75 90 0.7 0.52 90

Melbourne

0.8 0.75 365 1.3 0.88 90 0.5 0.51 90

Sy

dney 1.0 0.75 365 1.1 0.65 90 0.8 0.54 90

Brisbane

1.0 0.75 365 1.1 0.46 90 1.0 0.39 90

Darwin 1.0 0.75 365 1.0 1.22 90 1.1 0.18 90

194

Figures 6-9 to 6-13 show the hourly indoor air temperatures during the summer and winter seasons; note that the building with a lower structural mass (thermal mass) and lower concrete density (Ultra-lightweight) is more sensitive to changes in the outdoor temperatures.

Figure 6-9 Summer and Winter free floating temperature in climate zone 1 (Darwin)

Figure 6-10 Summer and Winter free floating temperature in climate zone 2 (Brisbane)

Figure 6-11 Summer and Winter free floating temperature in climate zone 5 (Sydney) 195

Figure 6-12 Summer and Winter free floating temperature in climate zone 6 (Melbourne)

Figure 6-13 Summer and Winter free floating temperature in climate zone 7 (Canberra)

6.4.4 Design week-free floating analysis

Figure 6-14 plots the frequency of indoor air temperature during the summer and winter design weeks by considering the heavyweight building with Normal concrete

(200.normal) and the lightweight building with Ultra-lightweight concrete

(Waffle.low); the indoor air temperature of both buildings and across all climates was outside the desired air setpoint ranges (18 to 26°C) most of the time, accompanied by consistent overheating (air temperatures higher than 26°C).

196

Canberra design week analysis (summer) Canberra design week analysis (winter) 20% 200.normal 20% Typical comfort Typical comfort 200.normal zone (18-26°C) waffle.low 15% 15% zone waffle.low (18-26°C) 10% 10%

5% 5%

Percentage hoursof

Percentage hoursof during the designweeks

during the designweeks 0% 0% 26 28 30 32 34 36 38 40 42 44 46 1416182022242628303234363840424446 Indoor Air Tempreture [°C] Indoor Air Tempreture [°C] Melbourne design week analysis (winter) Melbourne design week analysis (summer) 25% 25% 200.normal Typical 200.normal Typical 20% 20% comfort zone comfort zone waffle.low (18-26°C) waffle.low (18-26°C) 15% 15% 10% 10%

5% 5%

Percentage hoursof during the designweeks Percentage hoursof 0% 0% during the designweeks 22 24 26 28 30 32 34 36 38 40 42 44 14 16 18 20 22 24 26 28 30 32 34 36 38 40 42 44 Indoor Air Tempreture [°C] Indoor Air Tempreture [°C] Sydney design week analysis (summer) Sydney design week analysis (winter) 30% 200.normal 30% Typical 200.normal 25% waffle.low 25% comfort zone waffle.low 20% 20% (18-26°C) 15% 15% 10% 10%

5% 5% Percentage hoursof

0% Percentage hoursof 0% during the designweeks 28 30 32 34 36 38 40 42 44 46 during the designweeks 20 22 24 26 28 30 32 34 36 38 40 42 44 46 Indoor Air Tempreture [°C] Indoor Air Tempreture [°C] Brisbane Design week analysis (summer) Brisbane Design week analysis (winter) 30% 200.normal 30% 25% 25% Typical Typical comfort waffle.low 200.normal 20% zone(18-26°C) comfort zone 20% (18-26°C) 15% 15% waffle.low 10% 10% 5% 5% 0%

0% Percentage hoursof Percentage hoursof 26 28 30 32 34 36 38 40

20 22 24 26 28 30 32 34 36 38 40 42 44

during the designweeks during the designweeks Indoor Air Tempreture [°C] Indoor Air Tempreture [°C] Darwin Design week analysis (summer) 200.normal Darwin Design week analysis (winter) 40% waffle.low 30% Typical 200.normal 30% comfort waffle.low zone 20% 20% (18-26°C)

10% 10%

0% Percentage hoursof

Percentage hoursof 0% during the designweeks during the designweeks 32 34 36 38 40 42 44 46 26 28 30 32 34 36 38 40 42 44 46 Indoor Air Tempreture [°C] Indoor Air Tempreture [°C] Figure 6-14 Probability of indoor air temperature during the summer and winter design weeks for 200.normal and Waffle.low

197

Those structures with higher thermal conductivity concrete (200.normal) had lower peak indoor air temperatures than the low thermal conductivity concrete structures

(Waffle.low); for example, the variations of indoor and outdoor air temperature for the designed buildings in two climate zones (1 and 6) during the winter design week are shown in Figure 6-15. They indicate that the concrete structure with a lower thermal conductivity had a substantial increase of peak indoor air temperature by

1.2°C and 2°C in hot and cold climate zones, respectively (as shown in Figure 6-15 and in Appendix D).

Figure 6-15 Analysis of Winter design week free-floating for climate zones 1 (Darwin) and 6 (Melbourne)

In the summer design week, the resulting temperature patterns show that lighter buildings characterised by Ultra-lightweight concrete (Waffle.low) experienced a

198

higher daily oscillation than the other types of construction (as shown in Figure 6-16 and in Appendix D), where the building with higher mass and Normal weight concrete

(200.normal) structures had lower indoor air temperatures in general and a peak indoor air temperature that was 1.6-2.4°C lower than the lighter construction types. However, those structures in the hot dominated climate zone (Darwin) built with Ultra- lightweight materials lost heat quickly and cooled down faster during the night than the other buildings.

Figure 6-16 Analysis of Summer design week free-floating for climate zones 2 (Brisbane) and 7 (Canberra)

Table 6-10 shows the indoor thermal comfort conditions during operative hours (7 am to 9 pm) in the summer and winter design week. The accumulated degrees Celsius by which the hourly indoor air temperature was higher or lower than the desired comfort temperature (26 and 18°C, respectively in this case) is defined here as discomfort degree hours (DDH) (Ferrari & Zanotto 2015). The results here show that the DDH were almost 5% higher in the building constructed from Ultra-lightweight concrete 199

across all climates during the summer design week. The heavy buildings with Normal weight concrete reached a lower DDH (up to 50%) than the Ultra-lightweight concrete in cold climates (zones 6 and 7) during the winter design week. Note that those buildings with same type of concrete had a similar performance during the summer and winter design weeks.

Table 6-10 Summary of discomfort degree hours during the design weeks

Summer design week Winter design week Major cities 200.nor Waffle.nor Waffle.lo 200.norm 200.lo Waffle.norm Waffle.lo (climate) 200.low mal mal w al w al w

Canberra Canberra N.DDH 112 112 112 112 57 57 57 57

DDH 1,033 1,064 1,032 1,064 136 167 146 171

(7) M.DDH 5.44 5.60 5.43 5.60 0.71 0.87 0.76 0.89

N.DDH 98 97 97 96 10 13 10 11

Melbourne (6) Melbourne

DDH 645 660 629 648 6 12 7 12

M.DDH 3.84 3.93 3.74 3.86 0.04 0.07 0.04 0.07

N.DDH 98 98 98 98 53 54 51 52

Sydney (5) Sydney

DDH 1,257 1,295 1,256 1,299 106 137 109 138

M.DDH 7.48 7.71 7.48 7.73 0.63 0.81 0.65 0.82

Brisbane (2) Brisbane N.DDH 98 98 98 98 63 62 60 60

DDH 982 1,008 969 1,002 196 226 189 221

M.DDH 5.85 6.00 5.77 5.96 1.17 1.35 1.13 1.31

N.DDH 98 98 98 98 98 98 98 98

Darwin (1) Darwin

DDH 1,618 1,584 1,587 1,574 1,166 1,185 1,151 1,177

M.DDH 9.63 9.43 9.44 9.37 6.94 7.06 6.85 7.01

N.DDH: Number of discomfort hours during the design weeks (summer and winter); DDH: discomfort degree hours; M.DDH: Mean discomfort degree hours...... Heating load required; ..... Cooling load required.

The discomfort degree hours indicated that the 200.normal and Waffle.normal construction types would have a lower overheating peak (i.e lower DDH in Table 6- 200

10) in summer and winter conditions than the 200.low and Waffle.low types across the five major cities studied. A good example of the different discomfort degrees hours (DDH) between the four construction types is given in Figure 6-17 for climates

1 and 6. Note that the effects of structural materials (types of concrete) and construction forms (Flat and Waffle slabs) on indoor thermal conditions are slightly more noticeable in cold and moderate climates than hot and warm climates. The results of the other four climate zones are provided in Appendix E.

Figure 6-17 Discomfort degree hours during summer design week for climate zones 1 (Darwin) and 6 (Melbourne)

Figures 6-8 to 6-17 show that the thermal properties of structural concrete have more influence on the thermal performance of a building than the weight of the structure; in

201

fact the thermal properties of concrete (i.e. Ultra-lightweight versus Normal weight concrete) have a greater effect on the indoor air temperatures as the outside air temperature increases, and the differences between indoor air temperatures due to different structures (i.e. Flat slab versus Waffle slab) are more visible in moderate and cold climates.

6.5 Conclusion

This Chapter aimed to evaluate the impact that alternative concrete floor designs have on the energy performance of a typical office building. This Chapter used a benchmarking method to measure the thermal energy performance of a building using two forms of construction (Flat slab and Waffle slab) and two types of concrete

(conventional Normal weight and novel Ultra-lightweight). The structural design analysis provided the maximum and minimum building mass for the Flat slab and

Waffle slab respectively, which were then used to simulate the energy performance for whole buildings.

This Chapter analysis revealed how well structures with a higher thermal mass could moderate fluctuations between inside and outside air temperatures; those buildings with a higher concrete mass (thermal mass) stored more heat which then reduced the peak indoor air temperatures. Moreover, when Flat and Waffle slab structures are constructed from Normal weight concrete they have a similar energy performance, whereas Ultra-lightweight concrete resulted in indoor temperatures that are more sensitive to fluctuations in external air temperatures, so the building requires more energy to achieve the desired indoor temperature range.

This comparative analysis also revealed that choosing the appropriate type of concrete and construction form could reduce the annual cooling energy demand by a highly- 202

glazed office building by 14% in the colder climate zones and by 3% in warmer and hot climates.

The hourly free-floating simulation showed that a building with Ultra-lightweight concrete would experience higher daily peak indoor air temperatures during daytime, while the Lightweight building with Novel Ultra-lightweight experienced large increases of peak indoor air temperatures during the design weeks (Summer and

Winter) by 1.2°C to 2.4°C in the hot and cold climate zones, respectively; in fact this type of highly glazed office building risked overheating during the summer and winter periods.

These indoor thermal conditions confirm that buildings where conventional Normal weight concrete is used for the structural elements (slabs, columns and shear walls) had less discomfort degree hours during the design weeks than the novel Ultra- lightweight concrete.

Finally, an appropriate structural design in which the energy performance is also considered could lead to reductions in the thermal energy demand for office buildings.

Moreover, the magnitude of the impact of a structural system on the thermal performance of a structure could be changed across different alternatives. This Chapter highlights how important it is to look beyond the designed structural system and evaluate its impact with a holistic analysis. The whole life cycle environmental impact and costs associated with structural design alternatives are addressed in Chapter 7.

203

Chapter 7

Integrated life cycle cost method for low CO2-e emissions structural design by focusing on a benchmark office building in Australia

Contribution of the candidate to the Published Work

The contribution of the candidate in the published paper was 90%. He co-authored them with his main and co-supervisors. The candidate modelled, simulated a case study building for energy analysis and wrote the paper. Other authors reviewed the papers and provided useful feedback for improvement.

Citation: Accepted (Manuscript number: ENB_2017_2644) Robati, M, McCarthy, TJ & Kokogiannakis, G 2017, ‘Integrated life cycle cost method for low CO2-e emissions structural design by focusing on a benchmark office building in Australia’, Energy and Buildings.

204

7.1 Introduction

Chapter 6 investigated the possible effects that conventional and novel concrete materials would have on the thermal performance of typical office buildings in

Australia. The combination of lifetime CO2-e emissions impacts and energy performance in economic terms provides a framework to compare and evaluate the sustainability of the construction solutions because cost is a key factor that must be considered in the decision making process. The recent awareness of environmental problems has highlighted the need to include possible environmental impacts and building costs into the decision making process, but thus far this combination has received scant attention. This Chapter therefore proposes to integrate the CO2-e emissions and building costs into the structural design process using a method which includes the costs associated with building materials, construction methods and the amount of embodied carbon emissions during the life cycle of buildings. This Chapter analyses the effects that two construction systems (Flat slab and Waffle slab) and two structural materials (normal concrete and ultra-lightweight concrete) have on the overall costs of a typical high rise concrete structure (15-story office building) in

Australia (NS11401.1 2014).

7.2 CO2-e emissions impacts and life cycle cost analysis

The construction and building industry is responsible for a large part of the environmental burden because the Australian building sector, for example, uses almost 20% of Australia’s annual energy consumption and produces 23% of the GHG emissions (ABCB 2016).This situation will become even more critical due to the increasing number of houses (more than 3.3 million) resulting from the fast growth of population (NHSC 2011). A reduction in GHG is a vital need for Australia because 205

the nation committed to cope with carbon mitigation by 26-28% below the 2005 level by 2030 during the Paris UN Climate Conference (DEE 2015)

These growing pressures for CO2-e emissions accountability have led to greater efforts to improve the sustainability of the building industry in Australia (Akbarnezhad et al.

2014; Crawford 2011; Ding & Forsythe 2013; Miller et al. 2013; Moussavi

Nadoushani & Akbarnezhad 2015; Robati et al. 2016). Moreover, the need to assess the energy performance of buildings has extended from simply calculating the energy consumption during the operational phase to assessing sustainability over whole life cycle of buildings (Tian 2013; Tian & de Wilde 2011; Zuo et al. 2017).

Sustainability is now categorised into the environmental, social and economic aspects, so now most studies focus on the environmental and economic aspects of buildings by utilising Life Cycle Environmental Assessment (LCEA) and Life Cycle Cost

Assessment (LCCA) (Cabeza et al. 2014; Ding 2014; Shahrestani et al. 2013; Sharma et al. 2011; Silva & Ghisi 2014; Singh et al. 2011; Zuo et al. 2017).

Some studies tried to optimise the structural design of buildings by considering its economic and environmental aspects; they have proposed a conceptual framework

(called “Eco- efficiency method”) to measure sustainability by selecting the optimal product design and simultaneously considering the environmental impact and costs of the products (Cha et al. 2008; Hahn et al. 2010; Ji et al. 2014; Saling et al. 2002).

Others propose to evaluate the environmental and economic impacts by converting embodied CO2-e emissions into a momentary term over various stages of the building life cycle (Chou & Yeh 2015; Gu et al. 2008; Hong et al. 2013; Itsubo & Inaba 2003;

Ji et al. 2014; Kim, J et al. 2012; Silvestre et al. 2014). Hong et al. (2012) proposed to integrate the cost and CO2-e emissions associated with building structural design

206

(using different grades of concrete 21-30 MPa), while Chou and Yeh (2015) studied the environmental impacts associated with the life cycle impacts of two construction methods ( and cast in place) by focusing on the consumption of fuel, electricity, and water.

Several researches quantified the effect that the structural materials have on the whole life energy performance of buildings (Anderson & Silman 2009; Appleby 2012; DIIS

2013; Lemay & Leed 2011; Torgal & Jalali 2011); their studies have shown that basic design decisions about structural elements (type of floor, shape of core servers, arrangements of columns, and heights of beams) have a direct impact on the energy consumption of buildings. For example, Aye et al. (2012) studied the life cycle energy usage for three forms of building construction in Melbourne, Australia; they considered an eight-story multi-residential building with three different construction systems: modular prefabricated timber, conventional concrete construction, and modular prefabricated steel (Aye et al. 2012) and showed that a steel structure caused a 50% increase in embodied energy compared to a concrete structure, but the steel structure reduced material consumption up to 78% by mass compared to the concrete structure.

Hajdukiewicz et al. (2015) studied the structural and environmental performance of operating a building; they used a monitoring method for educational buildings that were mainly built with in situ and precast concrete, and showed there was a distinct lag between the outdoor and indoor air temperature in the monitored elements. They also pointed out the positive role that Ground Granulated Blast furnace Slag (GGBS) had in reducing the internal peak temperatures, and concluded by showing that the

207

thermal mass used in the floor systems slowed the flow of heat through the elements and caused a temperature lag in the system.

Previous studies have highlighted the relative impact that decision making has on energy consumption, environmental performance, and life cycle cost of buildings, but there is still no study which uses all these aspects to determine the impact of structural design on energy performance, life cycle CO2-e emissions, and life cycle costs in an integrated context for commercial buildings. Therefore, this Chapter proposes to integrate the CO2-e emissions of a building structure over its lifetime as an environmental cost in order to provide a quantitative value for evaluating global CO2- e emissions impacts made by different building structures in five Australian climate zones.

This Chapter is divided into different parts. The first part is the methodology which describes the method used to assess the integrated life cycle and analyse the whole of life costs associated with the research parameters. Section 7.3 describes the calculations of CO2-e emissions associated with the building structure, energy modelling, and life cycle cost analysis. The results and discussion about the key findings associated with the research parameters are summarised in section 7.4.

7.3 Methodology

Commercial and office buildings are built to last for several decades, but over such a long period of time a building utilises a wide range of resources and energy intensive processes. Cost effectiveness is a key component of structural design at the initial stage of projects, so this Chapter analyses the life cycle cost and CO2-e emissions impacts associated with a typical benchmark office building in Australia. This benchmark

208

building is one of four benchmarking buildings proposed by the National standard

Organization (NSDO) in Australia (NS11401.1 2014).

Figure 7-1 summarises all costs associated with a product or project over its lifetime, including the concept and definition, design and development, manufacturing and installation, operation and maintenance, and the disposal costs (AS/NZ4536 2014).

This Chapter quantifies the life cycle costs associated with the office building by considering alternative structural materials and forms of construction; the life cycle costs include the initial costs, and the operational and maintenance costs (as shown in

Equation 7-1).

퐿퐶퐶 = 푃푉퐶푃.퐶 + 푃푉푂푃.퐶 + 푃푉푅푒푝.퐶 (7-1)

LCC = Life Cycle Cost

PVCP.C = Present value of capital costs (initial costs)

PVOP.C = Present value of operational costs (over 50 year lifetime)

PVRep.C = Present value of replacement costs (over 50 year lifetime)

Figure 7-1 Life cycle cost of building

The initial cost includes materials and construction expenses; the operation and maintenance costs consist of utilities costs and repair costs that occur only every several years over a 50-year service life. The costs associated with a feasibility study

209

(concept and definition), and the design, development, and discarding (disposal) are excluded from the scope of this Chapter.

The CO2-e emissions impact associated with the benchmark building was derived in terms of CO2-e emissions because they contribute to more climate change than other

GHGs (methane, nitrous oxide, hydrofluorocarbons, perfluorocarbons and sulphur hexafluoride) (IPCC 2014; UNFCCC 2008). The environmental cost represents the cost of CO2-e emissions per tonne derived from the lifetime carbon emissions of buildings. Equation 7-2 is a method for estimating the costs associated with CO2-e emissions at each stage of a building’s life.

퐸퐼푀퐶푂2−푒 = 푃푉퐼퐶.퐶푂2 + 푃푉푂푃퐶.퐶푂2 (7-2)

EIMCO2-e emissions = Environmental impact (CO2-e emissions) costs

PVIC.CO2-e emissions = Present value of embodied CO2-e emissions cost

PVOPC.CO2-e emissions = Present value of operational CO2-e emissions costs (over 50 year lifetime)

This comparison framework is used to integrate the life cycle costs and CO2-e emissions impact of several structural design alternatives and then choose the best alternatives. The environmental impact costs are estimated from the total GHGs emitted over the building’s lifetime and the present carbon value (as shown in Figure

7-2). In this Chapter the total GHGs emitted by the building consist of the state of the product, the construction stage, and use stage.

210

Figure 7-2 Environmental cost analysis method

Figure 7-3 shows how the global assessment framework provides a method to combine the life cycle and environmental impact costs over the lifetime of the building. The economic value includes the life cycle cost (initial costs, and operational and maintenance costs) and environmental costs which includes the total equivalent CO2- e emissions cost, as a method to choose the most suitable structural design. This method compares the global costs across alternative building designs and delivers an outcome of cost results as a single global assessment parameter. This comparison framework integrates the results of life cycle costs and the environmental impacts of several structural design alternatives to help select the best alternatives.

211

Figure 7-3 Global assessment framework

7.3.1 Initial cost assessment method

The initial cost of the building includes the materials, transportation, and construction process. The quantities of building materials are from NS11401.1 (2014), the quantity of concrete and steel reinforcement comes from the detailed structural design (shown in Appendix B), and the input data to estimate these costs comes from the commonly used Australian construction cost guides and published literature (Cordell 2016;

Rawlinsons 2016). The base year taken was 2016, and the input climate data were classified based on five different climate zones across Australia: Darwin (climate zone

1); Brisbane (climate zone 2); Sydney (climate zone 5); Melbourne (climate zone 6);

Canberra (climate zone 7) (NS11401.1 2014). The cost of Ultra-lightweight concrete was calculated based on its unique mix design (proposed as mixes 32 and 36 in (Robati et al. 2016)), and the relative cost was collected from supplier price lists such as

Eastchem (2017) (supplier of hollow fly ash cenosphere) and Boral (2017) (supplier of other components). A summary of the building materials and associated costs and quantities are provided in Table 7-1. The quantities of materials used in the building

(items 8 to 16) are derived from NS 11401.1 (NS11401.1 2014), and the estimated

212

quantities of structural materials (concrete and steel reinforcement) for the Flat slab and Waffle slab are based on a detailed structural analysis.

Table 7-1 Summary of building materials and unit costs

Materials quantities Construction Materials base cost forms

Melbourne

Canberra

Building materials Brisbane Unit Sydney Da F.S W.S Cost/unit rwin

Concrete on 1 m3 250 250 $/m3 192 354 275 275 200 Ground (N20) Suspended 2,77 2 concrete m3 2002 $/m3 285 312 241 231 292 5 (N32)* Suspended 3 concrete m3 124 83 $/m3 265 372 295 304 265 (N40)* Steel in 4 tonne 411 636 $/tonne 2,350 2,220 2,205 2,205 2,660 concrete equivalent to 261 tonne Timber 5 m2 36,250 $/m2 138 (12mm thickness and 600 formwork kg/m3 density) Steel equivalent to 11.6 tonne 6 m2 11.16 $/m2 130 formwork (density:10 kg/m2) Plastic based 7 Number 218 $/number 269 ------formwork Timber equivalent to 20.1 tonne 8 m2 20.1 $/m2 136 battens (density:2 kg/m2) equivalent to 0.21 tonne (4.5 Steel eaves 9 m2 0.21 $/m2 37.4 mm thickness and 0.27 kg/m3 gutter density) equivalent to 0.1 tonne (4.5 Steel ridge 10 m2 0.1 $/m2 37.4 mm thickness and 0.27 kg/m3 flashing density) equivalent to 0.64 tonne (4.5 11 Steel fascia m2 0.64 $/m2 37.4 mm thickness and 0.27 kg/m3 density) equivalent to 2.25 tonne (4.5 Steel 12 m2 2.25 $/m2 37.4 mm thickness and 0.27 kg/m3 downpipe density) equivalent to 205.92 tonne (13 13 Plasterboard m2 205.92 $/m2 45.89 mm thickness and 668 kg/m3 density) Roof and equivalent to 1.41 tonne (90 14 celling m2 1.41 $/m2 16.28 mm thickness and 10 kg/m3 insulation density) Wall equivalent to 3.57 tonne 15 m2 3.57 $/m2 20.23 insulation (density:2 kg/m2) Steel doors 16 and m2 20 $/m2 208 ------mechanisms 17 Glass window m2 ------$/m2 475 ------

213

F.S: Flat slab; W.S: Waffle slab; *The cost associated with Ultra-lightweight concrete (N32 and N40) was 49% higher than normal concrete (the price is shown above). -The base materials cost for item 5 to 17 was extracted from NSW (Sydney) database (Cordell 2016) and it was assumed the same for the other cities.

7.3.2 Operation and maintenance costs assessment method

The operating costs of the building is based on the energy consumed over a 50-year service life, while the operating energy over the life cycle of buildings is calculated based on the simulated annual heating and cooling load. The estimated annual energy used was multiplied by the relative energy market forecasts up to 2040 (Economics

2015; Jacobs 2016) and the future method of calculation to extend the estimated costs of energy from 2040 to 2066 (a 50 year lifetime). Equation 7-3 was used to determine the present operational costs over a 50 year, although the future costs were then discounted by 7% as a nominal per year (Lawania & Biswas 2016).

50 퐹퐸퐶 푃푉푂푃.푐표푠푡 = ∑ 푛 (7-3) 푦=1 (1+푑)

PVOP.cost = Present value of operational costs

FEC = future energy costs (based on the market forecast and future costs analysis) d = discounted rate per year n = the appropriate number of years

The present value associated with maintenance (replacement) costs is estimated for glass windows (25 years) and plasterboard (35 years), both of which have a shorter lifetime than the building (50 years) (Rauf & Crawford 2012). The costs of ongoing repair, replacement, refurbishment, and external site development are excluded from this Chapter.

214

7.3.3 Environmental costs estimation method

The most common category for environmental impact used in life cycle assessment is global warming, so this Chapter used CO2-e emissions as a key method for assessing the environmental impact of various structural design alternatives. Here, the calculated

CO2-e emissions at each stage of the designed buildings (production, construction, end of life demolition and use) were included. For the production and construction phases, the CO2-e emissions emitted while manufacturing and transporting the construction materials were estimated. The embodied CO2-e emissions associated with construction work, transportation, and final demolition at the end of life is estimated at a level of about 1% (Ruuska & Häkkinen 2015; Sartori & Hestnes 2007).

During the operational stage, energy conversion results in the release of greenhouse gas emissions which was estimated by using the national emissions factor proposed by the Australian National Greenhouse Accounts (DEE 2016). The emissions factor used to convert the consumption of operational energy into CO2-e emissions is a function of the electricity purchased and consumed in 2015 (Lawania & Biswas 2016).

Electricity is a dominant energy source in the benchmark building to provide the required cooling load; Other energy sources do not contribute to total energy used

(Robati et al. 2017). It must be considered that the emissions projections are inherently uncertain, and this uncertainty increases into the projected future emissions. Based on

Australia’s report into emissions projections (DEE 2016), future emissions from the combustion of fuels to generate electricity are predicted at level of 186 Mt CO2-e emissions in 2030, which is roughly equivalent to 2015 levels. As such, this Chapter uses a base year of 2015 to estimate the CO2-e emissions associated with the energy usage stage of a building.

215

The values related to the embodied CO2-e emissions of materials is extracted from the accessible literature (Alcorn 2003; AusLCI 2016; Crawford 2011; eTool 2014;

Hammond et al. 2011; Moussavi Nadoushani & Akbarnezhad 2015; Robati et al.

2016). The mean distance from manufacturing companies to the site (the central business district for each city) is measured using the Google map tools (Poinssot et al.

2014). In the last stage, the environmental impact is converted into costs. The price of

CO2-e emissions is based on the Adams et al. (2014) method and the Australia

Emissions Trading Scheme (Combet 2012) with an inflation rate of 3% per year (RBA

2016). Future CO2-e emissions is discounted at 7% as a nominal rate per year

(Lawania & Biswas 2016). Equation 7-4 provides a present value for the costs of CO2- e emissions over the lifetime of the buildings.

50 푛 퐶푃퐶푂2−푒 ∗ (1+푎) 푃푉퐶푂2−푒 푐표푠푡 = ∑ 푛 (7-4) 푦=1 (1+푑)

PVCO2-e emissions = Present value of CO2-e emissions costs

CPCO2-e emissions = Current CO2-e emissions price a = is the expected increase in price per year (inflation rate) d = discounted rate per year n = the appropriate number of years

7.3.4 Building design and construction

The system for constructing the proposed 15 story office building is a mid-rise concrete structure (NS11401.1 2014); it has a square plan shape with a gross area of

1000 m2, and consists of twelve 3.30 metre high stories and three floors of parking (as

216

shown in Figure 7-4). The geometry and construction materials are summarised in

Table 7-2.

Table 7-2 Overall specification for the benchmark building

Parameter Unit Specification Basement dimensions m 31.62 × 31.62 Number of Stories --- 15 Concrete slab on ground mm 200 Concrete suspended slab mm 175 Average elevation per floor m 3.3 Total floor Area (including parking, Stairs & m2 15,000 Verandas) Total habitable area (external dimensions) m2 8,807.1 Number of floors above ground level --- 11 Number of rooms --- 176

Figure 7-4 Template of 15-storey commercial office building as visualised in DesignBuilder

217

The structural system is designed to meet the minimum needed to satisfy the national construction code (ABCB 2015; AS3600 2009; AS/NZ1170.0 2002) (as shown in

Table 7-3).

Table 7-3 Summary of the structural design

Construction form Flat slab (F.S) Waffle slab (W.S) Column span distance (L) 5.27 m 5.27 m Slab thickness (D) 200 mm 250 mm Concrete N20 250 250 quantities N32 3,005 2,002 (m3) N40 124 124 Steel quantities (Tonne) 753 679

Cross section

This current Chapter evaluated the possible effects of normal and low-density concrete with a higher weight (Flat slab) and lower weight (Waffle slab) office structure when the most common (Normal Weight) and novel (ultra-lightweight) concrete materials are used. The types of concrete mix designs were extracted from previously published journal papers and databases (CCAA 2015; O'Moore & O'Brien 2009; Robati et al.

2016; Wu et al. 2015; Yun et al. 2013) (shown in Table 7-3).

7.3.5 Energy modelling

To analyse the energy over the lifetime of the project, this office building is modelled in DesignBuilder (energy simulation software) so that the effects of alternative structural systems on the energy consumption could be assessed. This Chapter used the Building Code of Australia (BCA) “deemed to satisfy” approach to define the

218

envelope construction of this building (as shown in Table 7-3). The concrete thermal resistance and thermal mass, as two of the more prominent aspects of an energy analysis of building, are presented in Table 7-4 below.

Table 7-4 Benchmark building physical properties

Thermal resistance requirements and values and thermal mass values R-values Elements Item description References (m2.K/W)

Ground 1.25 Solid concrete*1 (150 mm, 2400 kg/m3) (ABCB 2015) floor Solid concrete (Study parameters) a. Flat slab with Normal Weight concrete*1 a.1.250 b. Flat slab with Ultra-lightweight b.1.81 concrete*2 Slabs c.1.216 (ABCB 2015) c. Waffle slab with Normal Weight d. 1.627 concrete*1 d. Waffle slab with Ultra-lightweight concrete*2

a.4.203 Solid concrete, (Study parameters) *1 b.4.836 a. Flat slab with Normal Weight concrete *2 Roof c.4.169 b. Flat slab with Ultra-lightweight concrete (ABCB 2015) *1 d. 4.581 c. Waffle slab with Normal Weight concrete d. Waffle slab with Ultra-lightweight concrete*2 3.42 Wall 2. 125 mm minimum solid reinforced concrete*1 (ABCB 2015)

U value was taken as 5.80 with SHG=0.81 for all climates (Daly et al. 2014; Window Guan 2009) *2Ultra- *1 Normal Weight lightweight concrete concrete Grade (MPa) 40(a) 32(b) 20(c) 40(a) 32(b) Density (Kg/m3) 2393 2470 1744 1400 1164 (CCAA 2015; Thermal conductivity(W/mK) 1.96 2.10 1.18 0.31 0.28 Robati et al. 2016) Specific heat (kJ/(kg.k)) 0.88 0.88 0.88 0.88 0.88 a. Grade N40 used in the vertical structural elements such as columns and shear walls. b. Grade N32 used in the slabs (Waffle and Flat). c. Grade N23 used in the other concrete element (staircase). 219

The internal energy loads in the office building form a large portion of energy usage and are significant input parameters in the energy analysis. Table 7-5 summarises the assumptions made to analyse the energy of the benchmark building. The schedules associated with the building are extracted from ABCB (2015).

Table 7-5 Benchmark building simulation assumptions

Parameters Key variables References Lighting power density and schedule 9 (W/m2) (ABCB 2015) Occupancy density and schedule 10 (m2/person) (ABCB 2015) Equipment load and schedule 15 (W/m2) (ABCB 2015) Domestic hot water 0.4 (L/m2) (ABCB 2015) Infiltration 0.28 (ACH) (Egan 2011) Ventilation requirements and schedule 10 (L/s/person) (ABCB 2015) HVAC set point 18-26 °C (ABCB 2015) HVAC DesignBuilder (CCAA 2015; Robati et Simple HVAC, al. 2016) Auto size

The results of modelling the benchmark building are shown in terms of total energy usage across different design alternatives. Here the total energy consumption is compared with national and states average energy usage to ensure that the predicted energy consumption is realistic (Pitt&Sherry 2012), and then, at the final stage, the total energy consumption of the benchmark building across four climates is converted into the equivalent energy costs and environmental costs (equivalent CO2-e emissions).

220

7.4 Results and Discussion

7.4.1 Lifetime environmental impacts

Figure 7-5 is a comparison between whole of life CO2-e emissions for the benchmark

office building across five major climate zones. Note that the region dominated by

heat such as Darwin has higher CO2-e emissions than the colder climate zones

(Canberra and Melbourne) due to the high cooling load during the operational phase

of the building in the hot climate where total CO2-e emissions are much higher than

cold climates. Moreover, CO2-e emissions associated with these buildings reveal that

Ultra-lightweight concrete released more CO2-e emissions than conventional concrete,

and the heavier building with Ultra-lightweight concrete (200.low) produced highest

2 carbon dioxide emissions (CO2-e /m ) across all five major cities. The NABERS

(National Australian Built Environment Rating System) rating tool revealed that the

environmental impact of these different design alternatives scored from 1 star (poor

performance) to 3.5 stars (above average performance). The buildings located in

Darwin and Canberra had the lowest (1 star) and highest (3.5 star) environmental

impact rating, respectively, but more importantly, these ratings changed across various

design alternatives in some regions; in Melbourne for instance, the lighter weight

building made from novel concrete had a lower rating (2.5 star) than the others, which

means that the selection of structural materials and the form of construction influences

the overall environmental ratings.

221

200.low 200.normal Waffle.low

Waffle.normal NABERS (3.5 star) NABERS (3 star)

NABERS (2.5 star) NABERS (1 star)

300

250

per year) per 200

2 e /m e

- 150 2

100

50 GHG (Kg CO (Kg GHG

0 Canberra Melbourne Sydney Brisbane Darwin Major cities

Source of the reference points NABERS (1.5, 2.5, 3 and 3.5 star rating) (Flores 2015)

Waffle.low: Lightweight structure (Waffle slab) with Ultra-lightweight concrete; Waffle.normal: Lightweight structure (Waffle slab) with Normal Weight concrete; 200.low: heavyweight structure (200mm Flat slab) with Ultra-lightweight concrete; 200.Normal: heavyweight structure (200mm Flat slab) with Normal Weight concrete.

2 Figure 7-5 Annual GHG (CO2-e emissions) normalised by net internal area (m )

Figure 7-6 shows the CO2-e emissions intensity associated with different phases of

2 life cycle for two cities with the highest and lowest amount of CO2-e /m . The first bar shows the CO2-e emissions related to the production phase, the construction phase and the end of life (demolition) phase of the building; the second bar shows the CO2-e emissions from operational phase of the building over a 50 year lifetime, and the last bar is the whole life CO2-e emissions for each alternative design. These results reveal how the range of CO2-e emissions for the buildings is influenced by changes in the type of concrete and type of construction. For instance, the Lightweight structures designed with Ultra-lightweight concrete had higher CO2-e emissions (5% in Canberra and 2% in Darwin) than the other design alternatives, whereas the Lightweight structure made from Normal weight concrete (Waffle.Normal) has the lowest CO2-e

222

emissions across both cities. This trend can be seen in the other three main cities (as shown in Appendix F).

Canbera Darwin Pre-use and construction phase Production, constuction, end of life Operation (50 years) phase Operation (50 years) phase Total Lifetime Total Lifetime 7,000 14,000

) 12,000

) 6,000

2 2

5,000 10,000 (kg /m (kg 4,000 8,000

3,000 6,000

emissions emissions (kg /m (kg emissions

2 4,000

2 2,000 CO CO 1,000 2,000 0 0

Design alternatives Design alternatives

2 2 *ACT average state CO2-e emissions (kg/m ) intensity: *NT average state CO2-e emissions (kg/m ) 12,140 (BZE 2013; Pitt&Sherry 2012) intensity: 9,421 (BZE 2013; Pitt&Sherry 2012)

Waffle.low: Lightweight structure (Waffle slab) with Ultra-lightweight concrete; Waffle.normal: Lightweight structure (Waffle slab) with Normal Weight concrete; 200.low: heavyweight structure (200mm Flat slab) with Ultra-lightweight concrete; 200.Normal: heavyweight structure (200mm Flat slab) with Normal Weight concrete.

Figure 7-6 Life cycle CO2-e emissions normalised by the gross floor area and separated by the type of concrete and method of construction.

7.4.2 Present value environmental costs of buildings

The environmental life cycle cost includes the total cost associated with CO2-e emissions over the whole life (50 years) of the office buildings. This Chapter estimated the CO2-e emissions at the production (peruse), construction, use stage, and end of life demolition. The CO2-e emissions of the final phase was considered at a 1% level by previous studies (Ruuska & Häkkinen 2015; Sartori & Hestnes 2007). Figure 7-7 shows the present environmental life cycle cost for four design scenarios (different forms of construction and structural materials), and the present cost of CO2-e emissions (Australian dollars) per m2 net internal area.

223

Environmental costs of the buildings 280

260 )

2) 240

/m 2 220 200 180 160

140 ($AUD/(Tonne CO ($AUD/(Tonne Present Enviromental cost Enviromental Present 120 100 Canberra Melbourne Sydney Brisbane Darwin 200.low $143.85 $175.77 $171.28 $204.32 $262.32 200.normal $140.33 $171.86 $166.52 $198.82 $260.16 Waffle.low $146.13 $179.45 $173.88 $206.97 $265.42 Waffle.normal $138.88 $170.32 $164.96 $197.50 $259.29

Waffle.low: Lightweight structure (Waffle slab) with Ultra-lightweight concrete; Waffle.normal: Lightweight structure (Waffle slab) with Normal Weight concrete; 200.low: heavyweight structure (200mm Flat slab) with Ultra-lightweight concrete;

200.Normal: heavyweight structure (200mm Flat slab) with Normal Weight concrete.

Figure 7-7 Environmental costs of the buildings

Figure 7-7 shows that the Lightweight building (Waffle slab) constructed with Ultra- lightweight concrete has the highest amount of carbon emissions costs ($AUD) per

2 normalised CO2-e emissions for the net settlement area (Tonne CO2-e /m ) over the lifetime of the buildings, while the Waffle slab with Normal weight concrete has the lowest carbon emissions costs. This shows that Ultra-lightweight concrete can cost up to 5% more over the whole of life environmental costs than Normal weight

(conventional) concrete, which means the type of concrete used is a large part of the

2 total CO2-e emissions. For Melbourne, the total carbon cost per m (net internal area)

2 of the building changed from 170 to 180 ($AUD/ (Tonne CO2-e /m )) when Ultra- lightweight is used as the main structural material. This change in cost of CO2-e emissions also occurs when the heavier building consists of 200mm thick Flat slabs

(200.normal and 200.low). 224

7.4.3 Life cycle cost analysis

7. Capital costs

This Chapter evaluated the capital costs associated with Flat slab and Waffle slab construction methods and Normal and Ultra-lightweight concrete across five regions

(as shown in Figure 7-8).

200.low 200.normal Waffle.low Waffle.normal

1,880 ) 2 1,860 1,840 1,820 1,800 1,780 1,760

1,740 Initial captal cost ($/m cost captal Initial 1,720 1,700 Melbourne Canbera Sydney Brisbane Darwin

Figure 7-8 Initial capital costs (construction) of the building

The results show that the average cost of a Lighter weight structure (Waffle Slab) with

Normal Weight concrete is less than the heavier structure (Flat slab). For example, the initial cost of the building in Melbourne with Waffle slab and Normal Weight concrete is 6% lower than the Flat slab with normal concrete (200.normal), however, Ultra- lightweight concrete in the structure resulted in higher capital cost in all five climate zones. The initial cost of the building with a Flat slab with Ultra-lightweight concrete

(200.low) is higher than the cost of the construction systems. As the literature highlights, the availability of supplementary cementitious materials used in Ultra- lightweight concrete is limited compared to Normal Weight concrete, so the costs are higher (Glavind 2012). For instance, the cost of Ultra-lightweight concrete used in this 225

Chapter is affected mainly by the price of Cenosphere (hollow particles from the production of fly ash) which is higher than the other concrete components (Eastchem

2017). Apart from the costs, there are still obstacles to the use of Lightweight and/or

Ultra-lightweight concrete, i.e., regulatory, technical, and supply chain (Cabeza et al.

2013; Duxson & Provis 2008; Van Deventer et al. 2012). There are several research programs currently aiming to remove these obstacles to allow for a wider use of

Lightweight and/or Ultra-lightweight (Huiskes et al. 2016; Yu, R et al.

2015), so it is worth considering when Ultra-lightweight concrete may become more available.

8. Operating and maintenance costs of the buildings

The present values associated with the operating expenses are derived from forecasts of energy consumption and energy prices; these simulations were used to determine the operation costs over the lifetime of buildings (50 years). The maintenance costs are compared to the present value and the cost of replacing materials with shorter lifespans than buildings (50 years), such as glass windows (25 years) and plasterboard

(35 years) (Rauf & Crawford 2012). Figure 7-9 shows the costs associated with Energy consumption and the materials used across five cities in Australia. The cost analysis shows that Darwin with its warm winter and hot summer had the highest energy consumption, while Melbourne, with its mild temperature had lower energy consumption than the other cities; however the operational costs are much higher than the replacement costs over the lifetime of the buildings. The energy performance was influenced by the methods of construction (Flat slab and Waffle slab) and the types of structural materials (concrete density: Normal Weight and Ultra-lightweight). The results show that selecting the right forms of construction and type of concrete could 226

save 2.4% of the running costs (during lifetime) for the building in Darwin and 5% for the other cities (Brisbane, Sydney, Melbourne, and Canberra). The Lightweight office building (Waffle) with Normal Weight concrete (Waffle.normal) has lower running costs than the alternatives, whereas the operating and replacement costs associated with the analysis reveal a consistently higher expenditure for the Waffle slab made from low density concrete (Ultra-lightweight).

Present value of operating & maintenance costs (PV AUD $/m2)

Darwin Brisbane Sydney Melbourne Canberra

20 30 40 50 60 70 80 90 Canberra Melbourne Sydney Brisbane Darwin Waffle.normal $35.02 $32.23 $41.93 $53.62 $84.35 Waffle.low $36.79 $33.92 $44.14 $56.13 $86.21 200.normal $34.99 $32.23 $41.93 $53.56 $84.13 200.low $35.72 $32.85 $42.98 $54.89 $84.59

Figure 7-9 Present value of operational and replacement costs of the building

7.4.4 Combined life cycle environmental and cost net present value

Table 7-6 summarises the whole life cycle cost assessment of the office building across five major cities in Australia; these costs are presented as environmental costs and life cycle cost (a combination of capital cost, operating costs, and maintenance costs) per net internal area. These results indicate that the energy demand at the operational phase and capital phase are the highest proportion of costs over the 50 year lifetime of the building. The equivalent cost of CO2-e emissions from production, construction, use, and demolition (end of life) can be up to 5% of the total cost of the buildings across all cities and design alternatives. The overall cost (LCCA+ 227

Environmental costs) of the office building in Sydney ranged from to 4,017 ($

AUD/m2) for the Waffle slab with normal concrete (Waffle.normal) to 4,189 ($

AUD/m2) for the Flat slab with Ultra-lightweight concrete (200.low).

Table 7-6 Whole life environmental and life cycle cost assessment of the building

PV.Operational

PV. (%) Comparison PV. Initial and LCCA Environmental Total costs Type of cost replacement Location cost Building cost ($ ($ ($ ($ AUD/m2) ($ AUD/m2) AUD/m2) AUD/m2) AUD/m2)

200.normal 1,819 1,611 3,431 172 3,603 +2%

200.low 1,870 1,642 3,512 176 3,688 +5%

Waffle.normal 1,743 1,612 3,354 170 3,525 ---- Melbourne Waffle.low 1,779 1,696 3,475 180 3,655 +4%

200.normal 1,830 1,750 3,580 140 3,720 +2%

200.low 1,853 1,786 3,639 144 3,783 +4%

Waffle.normal 1,754 1,751 3,505 139 3,644 ---- Canberra Waffle.low 1,770 1,840 3,610 146 3,756 +3% 200.normal 1,836 2,096 3,932 167 4,099 +2%

200.low 1,868 2,149 4,017 171 4,189 +4%

Waffle.normal 1,755 2,096 3,852 165 4,017 ---- Sydney Waffle.low 1,779 2,207 3,986 174 4,160 +4%

200.normal 1,817 2,678 4,495 199 4,694 +2%

200.low 1,870 2,744 4,614 204 4,818 +4%

Waffle.normal 1,741 2,681 4,423 198 4,620 ---- Brisbane Waffle.low 1,779 2,806 4,585 207 4,792 +4% 200.normal 1,841 4,206 6,047 260 6,307 +1%

200.low 1,864 4,229 6,093 262 6,356 +2%

Waffle.normal 1,769 4,218 5,987 259 6,246 ---- Darwin Waffle.low 1,786 4,311 6,096 265 6,362 +2% PV: Present value Waffle.low: Lightweight structure (Waffle slab) with Ultra-lightweight concrete; Waffle.normal: Lightweight structure (Waffle slab) with Normal Weight concrete; 200.low: heavyweight structure (200mm Flat slab) with Ultra-lightweight concrete; 200.Normal: heavyweight structure (200mm Flat slab) with Normal Weight concrete. This result indicates that a Waffle slab with an appropriate type of concrete (Normal

Weight) can save up to $156 per m2 (average value across all cities) in the total life 228

cycle cost of the building across all five major cities, whereas the use Ultra-light weight concrete in the Flat slab increased the total costs of the building by almost 3% compared to Normal weight (conventional) concrete. A comparison between the

Waffle slab and Flat slab, including Normal weight concrete, reveals that a lightweight structure with Normal weight concrete (Waffle.normal) can consistently save up to

2% in the total cost of the building, whereas the total cost associated with buildings constructed from Ultra-lightweight concrete across a lightweight structure

(Waffle.low) and heavyweight structure (200.low) has not changed, and moreover, of the methods of construction and structural materials are more tangible in colder climates than hotter climates such as Melbourne and Darwin.

7.5 Conclusion

This current Chapter assessed how different structural designs and construction systems affected the environmental impact and life cycle costs of a typical office building in Australia. The two main parameters in this Chapter are the method of construction (Flat slab and Waffle slab) and the type of concrete (Normal Weight and

Ultra-lightweight) used as structural materials. This Chapter finds that the total life cycle cost of buildings is influenced by the selection of structural materials and system of construction, indeed it has been shown that an appropriate building design can save almost 7% of the cost of material consumption, 5% of the total energy consumption expense, and 5% of the CO2-e emissions. A Lightweight building with a Waffle slab and Normal density costs less than the other buildings (LCCA and environmental costs) across the five main climate zones; the heavyweight building with a Flat slab and Normal weight concrete is the second best design alternative, saving almost 3% in total costs compared to the other buildings. 229

The analysis of CO2-e emissions costs shows that the use phase of the building is responsible for most CO2-e emissions, while the other building phase accounts for almost 5% of total CO2-e emissions. The present value of CO2-e emissions varies from

6 to 9.5 $AUD/m2 depending on the type of concrete, method of construction, and the climate of the city. The operational phase shows that CO2-e emissions due to energy consumption is strongly influenced by the weight of construction and type of concrete

(Normal Weight and Ultra-lightweight) used in the building. In general, a Lightweight building with normal density concrete uses less energy than the alternatives, and therefore produces less CO2-e emissions during the operating phase.

The findings of this Chapter show that selecting the optimal structural design based on a single-phase life cycle assessment might make it difficult to choose the ideal design alternatives. This is why all the stages of life cycle assessment must be considered when selecting alternative designs to achieve more environmentally friendly buildings.

The findings of this Chapter might be used as a guideline to optimise the performance of concrete structures by considering the efficiency of the structural materials and construction systems, but further studies must consider the potential impact of other structural forms (timber, steel and Post-tensioned) on the whole life cycle of buildings.

A summary of key findings and contribution of this research are provided in Chapter

8.

230

Chapter 8

Conclusion

8.1 Introduction

This Chapter draws together a series of publications leading to thesis by publication work. Chapters 4-7 have each been published or accepted for publication and combined give a coherent narrative of this research.

Chapter 4 incorporates embodied CO2-e emissions and thermal properties associated with the selection of concrete mix designs. The inconsistency in results of embodied

CO2-e emissions (as shown in Section 4.6) lead Chapter 5 to take into the account the uncertainties associated with the lifetime embodied CO2-e emissions of building materials and their impact on buildings. Chapter 5 takes into the account the uncertainties associated with the lifetime embodied CO2-e emissions of building materials and their impact on four different structural systems.

The question about potential impact of novel construction materials (as shown in

Section 4.7) is addressed in Chapter 6. The Chapter assesses the impact that structural design alternatives have on the energy demands and thermal performance of an office building in Australia.

Finally, Chapter 7 combines lifetime CO2-e emissions impacts and energy performance in economic terms to compare and evaluate the effects that two construction systems (Flat slab and Waffle slab) and two structural materials (normal concrete and ultra-lightweight concrete) have on the overall costs and sustainability of a typical high rise concrete structure in Australia. Chapter 7 proposes a framework to

231

demonstrate how different structural alternatives affect the lifetime energy consumption, the materials used, the CO2-e emissions, and the costs in a building.

The literature review and my publications that make up this thesis explore the life cycle CO2-e emissions necessary for the structural engineering design of multi-storey office buildings. This research highlighted a specific gap in the existing knowledge in the structural design of buildings. A number of research studies have considered the environmental impact that structural design has on different phases of a building’s life cycle, but no detailed work, until now, that assessed the CO2-e emissions and costs associated with selecting structural materials and forms over the lifetime of buildings.

Hence, this thesis examines the impact that structural design alternatives have over the lifetime of a building by using existing measurement tools and simulation methods.

This PhD identifies and quantifies the potential impact of structural design decisions and construction practices in an Australian context. The outcomes of this current PhD contribute to our knowledge by addressing the effects of structural design on lifetime costs, environmental impacts by focusing on CO2-e emissions, energy usage, and thermal performances of a building in Australia.

8.2 Review of aims and objectives

The primary aim of this PhD was to explain the impact that structural design decisions have on the life cycle impacts of buildings by achieving the key objectives which is shown in section 1.5.

This research was carried out in a number of distinct steps to address the objectives presented in Chapter 1 (section 1.5). The methodology used in this research (see

Chapter 3) was formed on the existing multi-assessment method to investigate the research questions from various perspectives. The key methodology consists of: 232

(1) Determining the embodied CO2-e emissions impacts associated with various

concrete design mix alternatives based on the method used in sections 3.6, 3.7

and 4.5. This stage considered the strategy proposed in the reviewed literature

(sections 1.2, 2.3-2.5) to quantify CO2-e emissions associated with various

types of concrete mixes.

(2) Implementing an uncertainty analysis to quantify the variations associated

with lifetime CO2-e emissions of the design alternatives (as discussed in

sections 1.2, 2.1 and 3.8). This stage of the study examines the uncertainties

caused by selecting various types of building materials in four different

construction systems (As shown in sections 3.8 and 5.4).

(3) Utilising an energy analysis method (sections 3.7 and 6.3.4) to identify the

potential impact of selecting a conventional and novel type of concrete (as

shown in section 2.9 and 6.1) on energy usage and thermal performance of

the benchmark office building (as shown in 6.3.1), while considering the two

commonly used constructional systems including: Flat slab and Waffle slab

(as discussed in sections 2.12, 3.4, 6.3.2 and 6.3.3).

(4) Developing a framework to describe a lifetime of CO2-e emissions and

construction costs (as discussed in sections 1.2.3, 2.8, 2.9) associated with

various design alternatives at the initial stage of the decision making process

(as shown in Chapter7). This stage of the study used a method (sections 3.10

and 7.3) to include the costs associated with building materials, construction

methods and the amount of embodied carbon dioxide emissions during the

life cycle of buildings. This was conducted by considering variations and

focusing on various types of structural concrete (conventional and novel

233

concrete) that were used in two different construction systems (Waffle slab

and Flat slab).

An overview of the key research findings is presented in the following section.

8.3 Research Findings Pertinent to the Research Objectives

1. Review existing literature in sustainability and sustainable building design and then identify the main key strategies, parameters, and tools associated with structural design; The review of existing literature pertaining to sustainability and sustainable building design identified the key factors associated with the sustainable design of the building

(Sections 2.2 to 2.5). The results indicate that the main strategies (as shown in Section

2.3) and parameters (a shown in Section 2.4) could have a considerable effect on the life cycle impact of buildings. Several multi-criterion rating and benchmarking systems, as well as tools and databases, describe the environmental impact associated with the production and consumption of different types of building materials (section

2.6). Much of the published work is based on limited system boundaries e.g. cradle to gate, or is focussed on the impact of material production only or on one stage of the life cycle. They do not consider the whole of life impacts (sections 2.8 and 2.9).

From the perspective of sustainable structural design, researchers have attempted to address the impact of structural systems on the environment by analysing a specific stage of a building’s life cycle (pre-use, use, or end of life) (section 2.8) (Asdrubali et al. 2013; Dixit et al. 2012; Ruuska & Häkkinen 2015; Saghafi & Teshnizi 2011; Torgal

& Jalali 2011); only limited attention has been given to the comprehensive life cycle analyses (Cradle to Grave) needed to determine the impact of structural design during the lifetime performance of buildings. This literature review has indicated there is still a lack of information quantifying the CO2-e emissions impact that the form of 234

construction and the properties of structural materials has over a building’s lifespan.

As such, a research methodology (Chapter 3) was developed to investigate the potential impact that structural solutions would have on the environmental and economic aspects of the lifetime impact of a building structure at the initial stage of decision making.

2. Investigate the relative impact that properties of selected construction materials will have on the environment at the initial stage of structural design; The appropriate selection of structural materials has a significant impact on the environment. The results of Chapter 4 reveals that the environmental impact, as measured by embodied CO2-e emissions, associated with various concrete mix designs as a structural material is systematically influenced by the cementitious materials used in Section 4.6.1. The analysis of thermal conductivity associated with concrete mix designs showed a direct relationship with the increasing density of concrete that is consistent with the relevant literature, as shown in section 4.6.3.

Moreover, Section 4.6.3 also shows that lower density concrete mixes could have a higher CO2-e emissions while providing a low thermal conductivity. This could be useful for energy saving during the operational stage of buildings but detrimental to the carbon footprint of construction.

Key findings:

- The embodied CO2-e emissions associated with common structural grades of

3 concrete (38-42 MPa) ranges from 211 to 509 kg CO2-e /m . This is directly

related to the proportion of cement used in the mix design (section 4.6.1). The

addition of readily available supplementary cementitious material can reduce

235

embodied CO2-e emissions (kg CO2-e emissions) by up to 16% compared to

general practice.

- The embodied CO2-e emissions for a selected grade of concrete (32MPa) can

3 vary from 63 kg to 496 kg CO2-e /m of concrete, but this depends on the type

of mix design and inventory database (4.6.2). The comparison shows some

inconsistencies in the calculation of embodied CO2-e emissions across the

different databases that are attributed to variations in the embodied CO2-e

emissions coefficients.

- Some of the embodied emissions databases currently available treated concrete

as only one specific product, and thus proposed an individual embodied CO2-

e emissions factor for each grade of concrete, regardless of the mix of

ingredients. This is an over simplification that leads to variations in estimating

the environmental impact at the initial stage of life cycle analysis of buildings.

Section 4.6.2 highlighted the need for transparency across the existing and

future databases. A summary of these findings will determine the variations

which could be used for comparison purposes for studies that used these

databases.

3. Investigate the uncertainty involved in the environmental impact that each construction material has on a typical high-rise office building; The results of section 4.6.2 show there are many variations in the selection of structural materials. The variations associated with a life cycle assessment of the structural solutions for a benchmark office building were considered in Chapter 5. The uncertainty analysis (section 5.5) revealed a specific range for the probability and reliability of the input parameters. The cumulative probability distribution curve revealed a range of embodied CO2-e emissions for each construction material during 236

the lifetime of buildings from extraction of materials to the end of life demolition

(section 5.4). The range of embodied CO2-e emissions for each construction material shows the uncertainties associated with variations in input parameters and their influence on the overall life cycle environmental impact of buildings.

The results of Chapter 5 established a baseline for the proposed office building that can be used for the designing alternatives at the early stage of decision making to predict the embodied CO2-e emissions impact over its whole life. Chapter 5 also provides a tool for assessing the long term impact of these early design decisions.

Key findings:

- The Waffle slab and Post-tensioned slab have the lowest total mean embodied

CO2-e emissions values compared to the Flat plate and Flat slab for the

benchmark building (Table 8-1).

Table 8-8-1 Embodied CO2-e emissions associated with different structural forms

2 Overall embodied CO2e (kg CO2-e /m )

Type of structural system Value F.S.R F.P P.T W.S Min 145 127 102 119 Mean 359 354 298 292 Max 732 764 646 590 Standard 83 94 81 71 deviation F.S.R: Flat slab for the reference building in proposed span distance of 5.27 m and slab thickness of 185 mm (NS11401.1 2014); F.P: Flat plate in the reference building with span distance of 5.27 m and slab thickness of 200 mm; P.T: Post-tensioned slab in the reference building with span distance of 5.27 m and slab thickness of 175 mm. W.S: Waffle slab in the reference building with span distance of 5.27 m and slab thickness of 250 mm (50mm topping + 200 stem depth).

237

- The level of uncertainty associated with different types of concrete (N40, N32

and N20) ranged from 6% to 78% of the total embodied CO2-e emissions of

the building. This level of variation, as highlighted in the previous Chapter

(section 4.6.2), comes from the significant difference in embodied CO2-e

emissions coefficients for each different concrete mix design. From the

structural design perspective, selecting an appropriate type of concrete could

substantially reduce the negative environmental impact of concrete.

- The structural materials (Concrete and steel reinforcement) and architectural

elements (windows with aluminium frame) accounted for the highest

proportions of embodied CO2-e emissions (as shown in Table 6-2).

Table 8-2 Variations in embodied CO2-e emissions of the materials used in the buildings

Embodied CO2-e emissions for the building materials

Materials Min Mean Max

Suspended concrete 7.79% 52.28% 78.73% (N32)1 Suspended concrete 0.10% 1.86% 6.02% Concrete (N40)2 Concrete on Ground 0.16% 4.54% 16.63% (N20)3 Window 3.78% 19.48% 48.68% Steel reinforcement 0.48% 13.48% 44.48% Timber formwork 0.17% 0.47% 1.36% Plasterboard 0.41% 2.20% 9.23% The variations associated with the other building materials were less than 1%. 1. Grade N40 used in the vertical structural elements such as columns and shear walls. 2. Grade N32 used in the slabs. 3. Grade N20 used in the other concrete element (staircase).

4. Investigate the relative contribution that types of structural concrete and types of slab have on the lifetime energy demand and thermal performance of office buildings;

238

The results of the energy simulation (section 6.4.2) showed that the thermal energy performance of the building was influenced by different forms of structural materials, and the thermal capacity of concrete construction forms can be utilised to shift thermal loads (sections 6.4.3 and 6.4.4), reduce peak demand, and reduce operational energy consumption. The selection of an appropriate concrete type was more important in terms of energy performance in the coldest (Melbourne and

Canberra) and hottest (Darwin) climate zones of this PhD.

Key findings

- Energy consumption across all five climate zones shows that a lightweight

office building (called Waffle.low) with a lower thermal conductivity concrete

(Ultra-lightweight concrete) required 5% more energy (up to 9.6 kWh/m2 per

year) than the other buildings (section 6.4.2).

- The variations in indoor and outdoor air temperatures for the different building

designs during the winter design week indicated that the Ultra-lightweight

concrete structure had a substantially higher peak indoor air temperature

(almost 2°C) which is undesirable in hot or mild climate zones (section 6.4.4).

- The discomfort degree hours (DDH) were almost 5% higher in the building

characterised with Ultra-lightweight concrete across all climates during the

summer design week (Table 6-10).

- Buildings made from Ultra-lightweight concrete respond to temperature and

heat flux stimulation faster, which causes overheating for most of the year.

- The thermal performance of the studied building have been more effected by

type of concrete than the construction system (Figure 6-8 to 6-17), and use of

use of Ultra-lightweight concrete in construction systems may require 239

alternative methods to mitigate the decreased thermal performance of

buildings.

5. Develop a life cycle assessment model of several office buildings as benchmark buildings. Propose a global assessment parameter as the results of whole life, CO2-e emissions, and structural costs during the lifetime of buildings. The developed life cycle assessment method was applied to the benchmark building to identify the lifetime CO2-e emissions costs associated with two construction systems (Flat slab and Waffle slab) and two structural materials (Normal concrete and

Ultra-lightweight concrete). The results showed that the office building designed with lightweight construction method (Waffle slab) and normal concrete (Normal weight) had a lower life cycle cost (50-year lifespan) than the other design alternatives across the five climate zones in Australia. It was found that selecting the appropriate construction forms and type of concrete can save on the total costs (life cycle cost and

CO2-e emissions cost) of the building across all five major cities. It was also found that an appropriate building design can save up to 7% in material use, 5% of total energy consumption, and 5% in life cycle CO2-e emissions.

Key findings

- The Lightweight structures designed with Ultra-lightweight concrete had

higher lifetime CO2-e emissions and environmental costs (up to 5.5%) than the

other design alternatives, whereas the Lightweight structure made from

Normal weight concrete (Waffle.Normal) had the lowest CO2-e emissions and

environmental costs (sections 7.4.1 and 7.4.2).

240

- The results in section 7.4.3 showed that selecting the appropriate forms of

construction and type of concrete could save 2.4% of the running costs (during

lifetime) for the building in Darwin and 5% for the other cities (Brisbane,

Sydney, Melbourne, and Canberra).

- The developed global assessment parameter (whole life environmental and life

cycle cost assessment) revealed that the Waffle slab with a right type of

concrete (Normal Weight) can save up to $156 per m2 (average value across

all cities) in the total life cycle cost of the building across all five cities.

- The global assessment parameter (whole life cost of the buildings) showed that

using Ultra-light weight concrete in the Flat slab increased the total cost of the

building by almost 3% compared to Normal weight (conventional) concrete.

The lightweight structure consisting of Ultra-lightweight concrete had a higher

initial capital cost and a higher demand for more operational energy (as shown

in 7.4.3) during the lifetime of the building and as a result, had a higher

environmental impact than the other design alternatives. From a structural

design perspective, it is worth looking beyond the structural system to

incorporate the potential long term impact that the selection of construction

materials has on the decision making process.

- The comparison between the type of structure (Waffle slab and Flat slab),

which included Normal weight concrete, revealed that a lightweight structure

with Normal weight concrete (Waffle.normal) can consistently save up to 2%

in the total cost of a building. While both systems have similar operating

energy costs, the Waffle slab has a lower environmental and initial capital costs

due to using less concrete in the structural system.

241

8.4 Contributions of the Research Findings

From a low CO2-e emissions structural design perspective, it is useful to understand how structural systems and materials can influence the lifetime performance of a building. Several databases (section 2.9), tools (section 2.6.3) and standards (sections

2.6.1, 2.6.2 and 2.6.3) exist to quantify the life cycle impact of building materials and systems, but until now, there has been almost no comprehensive life cycle assessment of the structural design and construction materials; this research therefore addresses that knowledge gap by achieving the following goals.

(i) The potential environmental impact and thermal performance of concrete.

The findings of Chapter 4 add to our understanding of the embodied CO2-

e emissions associated with different concrete design mixes by considering

the emissions for each ingredient. The table compiled for this study (Table

4.3) addresses the effects of concrete materials in an Australian context, so

now structural engineers can select an appropriate concrete mix to

minimise its embodied CO2-e emissions in Australia.

This research concludes that the embodied carbon assessment associated

with the properties of concrete varies for different inventory databases. The

summarised results of section 4.2 (Figure 4.9 and Figure 4.10) provide a

quantified comparisons across different databases that can be used as a

research tool where these databases are used.

(ii) An uncertainty analysis of the life cycle environmental impact of building

materials

242

The literature review highlighted (sections 2.10 and 5.3) the complexity

associated with applying a life cycle assessment to the building sector

because buildings have a very long lifetime (more than 50 years); it is

therefore a challenging task to predict the whole of lifetime impact of a

project from “Cradle to Grave”. Chapter 5 provides an uncertainty analysis

method to model the variations associated with the lifetime CO2-e

emissions of a building, the results of this which will provide the

contribution and magnitude of changes associated with the CO2-e

emissions of the most intensive building material over a building’s

lifetime. It was found that the selection of concrete, the type of windows

and the amount of steel reinforcement had the maximum impact on the

overall embodied CO2-e emissions across all four building structures.

Moreover, the established data in Section 5.5 provides a reference point

and guidelines for future studies to compare how their system performance

compared to the benchmark office building.

(iii) Potential impact of structural materials on the energy and thermal

performance of buildings

The ongoing development of more novel construction materials such as

Ultra-lightweight concrete, which combines moderate thermal insulation

with the load-bearing capacity raises a question about their potential

impact on the thermal mass of a building. Hence the effect of new materials

on the overall energy performance of a real building during its operational

phase merits study. This part of the work points out the importance of a

243

holistic approach to look beyond the designed structural system and

evaluate its impact over the lifetime of a building.

One of the major contributions of this research is the subsequent

development and application of a simulation based assessment that Ultra-

lightweight concrete has on the overall energy performance of a

benchmark building. Energy analysis and thermal performance was used

to investigate the impact of selecting the type of concrete (Normal and

Ultra-lightweight) and structural systems on the overall indoor comfort and

energy performance of a benchmark office building in Australia. The

energy analysis estimated the effect that the form of construction and

structural materials had on the thermal energy performance of the building.

The association between thermal capacity and concrete structure can be used

to decrease the operational energy demand, reduce the peak load, and shift the

thermal loads. Selecting the appropriate type of concrete was more important

in terms of energy performance in the coldest (Melbourne and Canberra) and

hottest (Darwin) climate zones of this study.

(iv) Develop an assessing tool to quantify the lifetime CO2-e emissions

associated with different structural design alternatives of buildings.

This research develops a framework for the life cycle CO2-e emissions of

the structural design of buildings and a method to demonstrate how

different structural solutions and materials affect lifetime energy

consumption, CO2-e emissions, and costs in a building. This framework

addresses the lifetime effects of structural decision as a single a single

economic value. It was shown that an appropriate structural design can 244

reduce the cost of material usage (by 7%), the operational energy demand

(by 5%), and the CO2-e emissions impacts (by 5%). This framework can

be used as a supporting decision making tool to select the most efficient

structural materials and methods of construction over the lifetime of the

building.

The research undertaken in this thesis provides a quantitative method that will enable structural engineers to integrate life cycle building performance and emissions awareness into the decision making process; it also proposes a framework to assess the efficiency of structural alternatives over a lifetime whilst simultaneously considering the environmental impact of the initial design. During this research, some observations pointed out the need for further research in this field which would enhance our knowledge of the low CO2-e emissions structural design of buildings.

8.5 Recommendations for further research

The previous sections investigated the potential impact that structural materials and types of construction have on the life cycle CO2-e emissions of an office building.

This investigation used a benchmarking method to compare the impact of different design alternatives (various types of concrete and construction systems), but the potential impact of structural design on other aspects of sustainable building design were outside the scope of this PhD.

There are also several areas for further development and applications for the work undertaken in this thesis. The framework developed in this PhD to assess lifetime environmental impacts and costs associated with different structural solutions was applied to a benchmark office in Australia, but it could usefully be applied to assess other types of buildings. This PhD examined two aspects of structural design over the 245

whole life cycle of a building, the floor systems (slab thickness) in various systems of construction, and different types of concrete (low density and high density). A similar approach could be used to assess the potential effect of structural design on various types of buildings by considering alternative framing systems and materials such as a cross laminated timber system.

The benchmark office building in this research is one of four benchmark buildings proposed the by National Standard Development Organization (NS11401.1 2014), of which far too little attention has been given to the other three. Considering the potential effect that structural systems have on the lifetime performance of various buildings can help in understanding possible future development in their sustainable structural design. It is recommended that the impact of structural design solutions be considered on the other archetype benchmark buildings, as proposed by NS11401.1

(2014), in order to enhance our understanding of how structural engineering promotes the demonstrated advantages of sustainable tools.

The levels of uncertainty associated with estimating the life cycle performance of buildings might be further investigated by incorporating the cost uncertainty associated with the structural design of buildings. The cost uncertainty incorporates the variations associated with structural solutions, rehabilitation measures, and the expected losses that might happen during the lifetime of a building for each defined limit states (Tsimplokoukou et al. 2014). The uncertainties associated with structural design can include hazard uncertainties (such as wind loads), structural uncertainties

(such a structural capacity, materials properties and stiffness), and interaction mechanism uncertainties (such as the duration of pressure levels (Tsimplokoukou et al. 2014). Integrating these developed life cycle assessment methods (as proposed in

246

Chapter 7) with structural performance assessment method could improve the sustainability aspects of building design.

Finally, the continuing development of more novel construction materials raises a concern about their potential impact on the lifetime performance of buildings. This

PhD has shown the potential impact of novel construction materials (Ultra-lightweight concrete) on the lifetime performance of a benchmark building; it also provides a base for developing a labelling model to compare the total CO2-e emissions of materials and to demonstrate its comparative performance against the benchmark data; a labelling method can also help structural engineers realise the ecological impact and potential lifetime performance of a material in a simple and clear manner.

8.6 Summary

This thesis set out to evaluate the impact that a structural system and construction materials have on the sustainability of buildings. A core part of this thesis was based on the sequence of papers (Chapters 4-7) already published or accepted for publication. This research mainly emphasised the life cycle analysis of a building while measuring the effects that design solutions have on the environment over its whole lifetime, from Cradle to Grave.

Returning to the question posed at the beginning of this PhD, it is now possible to state that structural engineering can contribute to the sustainability of buildings by understanding the long term characteristics and limitations of alternative designs at a global scale before selecting an appropriate design.

247

References

ABCB 2015, National construction code, volume one Australian Building Codes Board, Canberra. ABCB 2016, Energy efficiency provisions, National counstruction code volume two, Australian Building Codes Board, Canberra. Abdalqader, AF, Jin, F & Al-Tabbaa, A 2016, ‘Development of greener alkali-activated cement: utilisation of sodium carbonate for activating slag and fly ash mixtures’, Journal of Cleaner Production, vol. 113, no. Supplement C, pp. 66-75. Abdullah, MMAB, Hussin, K, Bnhussain, M, Ismail, KN, Yahya, Z & Abdul Razak, R 2012, ‘Fly ash-based geopolymer lightweight concrete using foaming agent’, International journal of molecular sciences, vol. 13, no. 6, pp. 7186-98. Abrahamsen, RB & Malo, KA 2014, ‘Structural design and assembly of “treet”-a 14-storey timber residential building in norway’, in World Conference on Timber Engineering 2014, p. 8. ACI122R 2014, Guide to thermal properties of concrete and masonry systems, American Concrete Institute, 9780870319716, https://books.google.com.au/books?id=- 1fdrQEACAAJ. ADAA 2012, Fly ash technical note No 11, http://www.adaa.asn.au/uploads/default/files/adaa_technical_note_11.pdf. ADAA 2016, Ash Development Association of Australian15/03/2016, . Adalberth, K 1997, ‘Energy use during the life cycle of buildings: a method’, Building and Environment, vol. 32, no. 4, pp. 317-20. Adams, PD, Parmenter, BR & Verikios, G 2014, ‘An emissions trading scheme for Australia: national and regional impacts’, Economic Record, vol. 90, no. 290, pp. 316-44. AGO 2004, Australian Green house office factors and method workbook. Akbarnezhad, A, Ong, KCG & Chandra, LR 2014, ‘Economic and environmental assessment of deconstruction strategies using building information modeling’, Automation in Construction, vol. 37, pp. 131-44. Aktas, CB & Bilec, MM 2012, ‘Impact of lifetime on US residential building LCA results’, The International Journal of Life Cycle Assessment, vol. 17, no. 3, pp. 337-49. Al-Sanea, SA & Zedan, MF 2011, ‘Improving thermal performance of building walls by optimizing insulation layer distribution and thickness for same thermal mass’, Applied Energy, vol. 88, no. 9, pp. 3113-24. Alcorn, A 2003, Embodied energy and CO2 coefficients for NZ building materials, Centre for Building Performance Research, Wellington, New Zealand. Ali, MM & Moon, KS 2007, ‘Structural developments in tall buildings: current trends and future prospects’, Architectural science review, vol. 50, no. 3, pp. 205-23. Alimoradi, A 2014, ‘Sustainable structural design: Outlook and potentials’, Journal of Structural Engineering, vol. 141, no. 3, p. B2514001. Anastasiou, EK, Liapis, A & Papayianni, I 2015, ‘Comparative life cycle assessment of concrete road pavements using industrial by-products as alternative materials’, Resources, Conservation and Recycling, vol. 101, no. Supplement C, pp. 1-8. Anderson, JE & Silman, R 2009, ‘A life cycle inventory of structural engineering design strategies for greenhouse gas reduction’, Structural Engineering International, vol. 19, no. 3, pp. 283-8. Appleby, P 2012, Integrated sustainable design of buildings, Routledge. AS3582.1 2016, Supplementary cementitious materials - Fly ash, Standards Australia, Sydney. 248

AS3600 2001, Concrete structures, Standards Australia International Ltd, Sydney, Australia. AS3600 2009, Concrete structures, Standards Australia International Ltd, Sydney, Australia. AS/NZ1170.0 2002, Part0: General principles structure design actions, Standards Australia International Ltd, Sydney, Australia. AS/NZ1170.1 2002, Part 1: Permanent, imposed and other actions, Australian/ New Zealand Standard, Standards Australia International Ltd. AS/NZ1170.2 2011, Part 2: Wind actions, Standards Australia International Ltd, Sydney, Australia. AS/NZ4536 2014, Life cycle costing—An application guide, Standards Australia International Ltd, Sydney, Australia., Sydney, Australia. Asadi, E, Da Silva, MG, Antunes, CH & Dias, L 2012, ‘Multi-objective optimization for building retrofit strategies: a model and an application’, Energy and Buildings, vol. 44, pp. 81-7. Asdrubali, F, Baldassarri, C & Fthenakis, V 2013, ‘Life cycle analysis in the construction sector: Guiding the optimization of conventional Italian buildings’, Energy and Buildings, vol. 64, no. 0, pp. 73-89. Asif, M, Muneer, T & Kelley, R 2007, ‘Life cycle assessment: A case study of a dwelling home in Scotland’, Building and Environment, vol. 42, no. 3, pp. 1391-4. AusLCI 2016, The Australian national life cycle inventory database, http://alcas.asn.au/. Ayaz, E & Yang, F 2009, ‘Zero carbon isn’t really zero: why embodied carbon in materials can’t be ignored’, Design Intelligence. Aye, L, Ngo, T, Crawford, RH, Gammampila, R & Mendis, P 2012, ‘Life cycle greenhouse gas emissions and energy analysis of prefabricated reusable building modules’, Energy and Buildings, vol. 47, pp. 159-68. Bannister, P 2004, Australian building codes board: Class 5 benchmarking, Canberra, Australia. Bengtsson, J & Howard, N 2010, A life cycle impact assessment Part 1: Classification and characterisation. Berardi, U 2012, ‘Sustainability assessment in the construction sector: Rating stems and rated buildings’, Sustainable development, vol. 20, no. 6, pp. 411-24. Berge, B 2009, The ecology of building materials, Routledge. Berndt, ML 2015, ‘Influence of concrete mix design on CO2 emissions for large wind turbine foundations’, Renewable Energy, vol. 83, pp. 608-14. Berndt, ML, Sanjayan, J, Foster, S & Castel, A 2013, Pathways for overcoming barriers to implementation of low CO2 concrete, CRC for Low Carbon Living. Bisinella, V, Conradsen, K, Christensen, TH & Astrup, TF 2016, ‘A global approach for sparse representation of uncertainty in Life Cycle Assessments of waste management systems’, The International Journal of Life Cycle Assessment, vol. 21, no. 3, pp. 378-94. Blankendaal, T, Schuur, P & Voordijk, H 2014, ‘Reducing the environmental impact of concrete and asphalt: a scenario approach’, Journal of Cleaner Production, vol. 66, no. Supplement C, pp. 27-36. Blengini, GA & Di Carlo, T 2010, ‘The changing role of life cycle phases, subsystems and materials in the LCA of low energy buildings’, Energy and Buildings, vol. 42, no. 6, pp. 869-80. Bojacá, CR & Schrevens, E 2010, ‘Parameter uncertainty in LCA: stochastic sampling under correlation’, The International Journal of Life Cycle Assessment, vol. 15, no. 3, pp. 238-46. BOM 2011, Mean monthly and mean annual heating and cooling degree days data (base climatological data sets), Bureau of Meteorology, viewed 17/5/2017, .

249

Boral 2017, Building and construction materials, viewed 25/2/2017, . Bosteels, T, Tipping, N, Botten, C & Tippet, M 2010, ‘Sustainability benchmarking toolkit for commercial buildings: Principles for best practice’, Better Buildings Partnership. BPIC 2014, Building product life cycle inventory, Building Products Innovation Council, viewed 8/8/2015, . Buchanan, AH & Honey, BG 1994, ‘Energy and carbon dioxide implications of building construction’, Energy and Buildings, vol. 20, no. 3, pp. 205-17. BZE 2013, Zero carbon Australia buildings plan, Beyond Zero Emissions, Melbourne Energy Institute, University of Melbourne, Melbourne, Australia. Cabeza, LF, Barreneche, C, Miró, L, Morera, JM, Bartolí, E & Inés Fernández, A 2013, ‘Low carbon and low embodied energy materials in buildings: A review’, Renewable and Sustainable Energy Reviews, vol. 23, pp. 536-42. Cabeza, LF, Rincón, L, Vilariño, V, Pérez, G & Castell, A 2014, ‘Life cycle assessment (LCA) and life cycle energy analysis (LCEA) of buildings and the building sector: A review’, Renewable and Sustainable Energy Reviews, vol. 29, pp. 394-416. CASBEE 2016, The assessment method employed by CASBEE, Japan Green Build Council and Japan Sustainable Building Consortium 28/4/2016, . CC 2016, Concrete framed buildings: a guide to design and construction, Concrete Centre, London. CCAA 2003, Guide to long-span concrete floors, Cement and Concrete Association of Australia, Sydney, Australia. CCAA 2015, Green Star mat-4 concrete credit user guide, Sydney, Australia, http://www.concrete.net.au/iMIS_Prod. Celik, K, Meral, C, Petek Gursel, A, Mehta, PK, Horvath, A & Monteiro, PJM 2015, ‘Mechanical properties, durability, and life-cycle assessment of self-consolidating concrete mixtures made with blended portland cements containing fly ash and limestone powder’, Cement and Concrete Composites, vol. 56, pp. 59-72. Cha, K, Lim, S & Hur, T 2008, ‘Eco-efficiency approach for global warming in the context of Kyoto Mechanism’, Ecological Economics, vol. 67, no. 2, pp. 274-80. Chandra, S & Berntsson, L 2003, Lightweight aggregate concrete, Norwich, N.J. : Noyes Publications. Chang, Y, Ries, RJ & Lei, S 2012, ‘The embodied energy and emissions of a high-rise education building: A quantification using process-based hybrid life cycle inventory model’, Energy and Buildings, vol. 55, pp. 790-8. Chen, C, Habert, G, Bouzidi, Y, Jullien, A & Ventura, A 2010, ‘LCA allocation procedure used as an incitative method for waste recycling: An application to mineral additions in concrete’, Resources, Conservation and Recycling, vol. 54, no. 12, pp. 1231-40. Cho, H-W, Roh, S-G, Byun, Y-M & Yom, K-S 2004, ‘Structural quantity analysis of tall buildings’, in CTBUH Conference: tall buildings in historical cities–culture & technology for sustainable cities, pp. 10-3. Cho, YS, Kim, JH, Hong, SU & Kim, Y 2012, ‘LCA application in the optimum design of high rise steel structures’, Renewable and Sustainable Energy Reviews, vol. 16, no. 5, pp. 3146-53. Chou, J-S & Yeh, K-C 2015, ‘Life cycle carbon dioxide emissions simulation and environmental cost analysis for building construction’, Journal of Cleaner Production, vol. 101, pp. 137-47. Chowdhury, AA, Rasul, M & Khan, MMK 2008, ‘Thermal-comfort analysis and simulation for various low-energy cooling-technologies applied to an office building in a subtropical climate’, Applied Energy, vol. 85, no. 6, pp. 449-62.

250

Chowdhury, R, Apul, D & Fry, T 2010, ‘A life cycle based environmental impacts assessment of construction materials used in road construction’, Resources, Conservation and Recycling, vol. 54, no. 4, pp. 250-5. CIBSE 2006, Environmental design : CIBSE guide A, Chartered Institution of Building Services Engineers London, London, England. Ciroth, A, Fleischer, G & Steinbach, J 2004, ‘Uncertainty calculation in life cycle assessments’, The International Journal of Life Cycle Assessment, vol. 9, no. 4, pp. 216-26. Coakley, D, Raftery, P & Keane, M 2014, ‘A review of methods to match building energy simulation models to measured data’, Renewable and Sustainable Energy Reviews, vol. 37, pp. 123-41. Combet, G 2012, Securing a clean energy future : implementing the Australian Government's climate change plan, The Australian Government Department of Climate Change and Energy Efficiency, Canberra, Australia, ISBN 978-1-922003- 44-7. Cordell 2016, Building cost guide : Commercial industrial - NSW, Cordell Building Publications, Australia. Crawford, R 2011, Life Cycle Assessment in the Built Environment, Taylor & Francis. Crawford, R & Treloar, G 2005, ‘An assessment of the energy and water embodied in commercial building construction’, in Australian Life Cycle Assessment Conference, pp. 1-10. Crawford, RH 2008, ‘Validation of a hybrid life-cycle inventory analysis method’, Journal of Environmental Management, vol. 88, no. 3, pp. 496-506. Crawford, RH 2013, ‘Post-occupancy life cycle energy assessment of a residential building in Australia’, Architectural Science Review, vol. 57, no. 2, pp. 114-24. Crawford, RH & Treloar, GJ 2003, ‘Validation of the use of Australian input-output data for building embodied energy simulation’, in IBPSA 2003: Proceedings of the Eighth International Building Performance Simulation Association Conference on Building Simulation: For better Building Design, pp. 235-42. Crawley, DB, Hand, JW, Kummert, M & Griffith, BT 2008, ‘Contrasting the capabilities of building energy performance simulation programs’, Building and Environment, vol. 43, no. 4, pp. 661-73. Crossin, E 2012, ‘Comparative life cycle assessment of concrete blends’, Centre for Design, RMIT University: Melbourne, Australia. Crossin, E 2015, ‘The greenhouse gas implications of using ground granulated blast furnace slag as a cement substitute’, Journal of Cleaner Production, vol. 95, no. Supplement C, pp. 101-8. Čuček, L, Klemeš, JJ & Kravanja, Z 2012, ‘A review of footprint analysis tools for monitoring impacts on sustainability’, Journal of Cleaner Production, vol. 34, pp. 9- 20. Daly, D 2015, ‘Improving energy efficiency in lower quality commercial buildings: Simulation, retrofit optimisation and uncertainty’, Doctor of Philosophy thesis, University of Wollongong. Daly, D, Cooper, P & Ma, Z 2014, ‘Implications of global warming for commercial building retrofitting in Australian cities’, Building and Environment, vol. 74, pp. 86-95. Daly, D, Cooper, P & Ma, Z 2014, ‘Understanding the risks and uncertainties introduced by common assumptions in energy simulations for Australian commercial buildings’, Energy and Buildings, vol. 75, pp. 382-93. Damdelen, O, Georgopoulos, C & Limbachiya, M 2014, ‘Measuring thermal mass of sustainable concrete mixes’, in Computing in Civil and Building Engineering (2014), pp. 1554-61.

251

Damdelen, O, Georgopoulos, C & Limbachiya, M 2015, Measuring thermal Mass of sustainable concrete mixes, Scholars' Press. Damtoft, J, Lukasik, J, Herfort, D, Sorrentino, D & Gartner, E 2008, ‘Sustainable development and climate change initiatives’, Cement and concrete research, vol. 38, no. 2, pp. 115-27. Danatzko, JM & Sezen, H 2011, ‘Sustainable structural design methodologies’, Practice Periodical on Structural Design and Construction, vol. 16, no. 4, pp. 186-90. de Castro, S & de Brito, J 2013, ‘Evaluation of the durability of concrete made with crushed glass aggregates’, Journal of Cleaner Production, vol. 41, pp. 7-14. De Schepper, M, Van den Heede, P, Van Driessche, I & De Belie, N 2014, ‘Life cycle assessment of completely recyclable concrete’, Materials, vol. 7, no. 8, pp. 6010-27. De Wolf, C, Pomponi, F & Moncaster, A 2017, ‘Measuring embodied carbon dioxide equivalent of buildings: A review and critique of current industry practice’, Energy and Buildings, vol. 140, pp. 68-80. DEE 2013, National waste reporting 2013, Department of Environment and Energy, viewed 21/4/2017, . DEE 2015, Australia’s 2030 emissions reduction target, Department of Environment and Energy, Canberra, Australia, http://www.environment.gov.au. DEE 2016, Australia’s emissions projections 2016, Department of the Environment and Energy, Canberra, Australia, http://www.environment.gov.au. DEE 2016, National greenhouse accounts factors, Department of the Environment and Energy, Canberra, Australia, http://www.environment.gov.au. DEHP 2014, Waste—everyone’s responsibility: Queensland waste avoidance and resource productivity strategy (2014–2024). December edn, Department of Environment and Heritage Protection, Queensland Government, Brisbane, Australia, 21/4/2017, www.ehp.qld.gov.au. DesignBuilder 2017, HVAC Model Option, DesignBuilder Software Ltd, viewed 1/02/2017, . Dhir, R 2006, ‘The potential of fly ash: the future looks bright’, Concrete, vol. 40, no. 6, pp. 68-70. DIIS 2013, Your home : Australia's guide to environmentally sustainable homes, Department of Industry, Innovation and Science, Canberra, Australia. Dimoudi, A & Tompa, C 2008, ‘Energy and environmental indicators related to construction of office buildings’, Resources, Conservation and Recycling, vol. 53, no. 1, pp. 86- 95. Dimoudi, A & Tompa, C 2008, ‘Energy and environmental indicators related to construction of office buildings’, Resources, Conservation and Recycling, vol. 53, no. 1-2, pp. 86-95. Ding, G 2004, ‘The development of a multi-criteria approach for the measurement of sustainable performance for built projects and facilities’, Doctor of Philosophy thesis, University of Technology Sydney. Ding, G 2008, ‘Sustainable construction-The role of environmental assessment tools’, Journal of Environmental Management, vol. 86, no. 3, pp. 451-64. Ding, G 2014, ‘Life cycle assessment (LCA) of sustainable building materials: an overview’, pp. 38-62. Ding, G & Forsythe, PJ 2013, ‘Sustainable construction: life cycle energy analysis of construction on sloping sites for residential buildings’, Construction Management and Economics, vol. 31, no. 3, pp. 254-65. Dixit, MK 2017, ‘Embodied energy analysis of building materials: An improved IO-based hybrid method using sectoral disaggregation’, Energy, vol. 124, no. Supplement C, pp. 46-58.

252

Dixit, MK, Fernández-Solís, JL, Lavy, S & Culp, CH 2010, ‘Identification of parameters for embodied energy measurement: A literature review’, Energy and Buildings, vol. 42, no. 8, pp. 1238-47. Dixit, MK, Fernández-Solís, JL, Lavy, S & Culp, CH 2012, ‘Need for an embodied energy measurement protocol for buildings: A review paper’, Renewable and Sustainable Energy Reviews, vol. 16, no. 6, pp. 3730-43. Dossche, C, Boel, V, De Corte, W, Van den Heede, P & De Belie, N 2016, ‘A plant based LCA of high-strength elements and the assessment of a practical ecological variant’, Cement and Concrete Composites, vol. 73, no. Supplement C, pp. 192-202. Duxson, P & Provis, JL 2008, ‘Low CO2 concrete: are we making any progress’, Environment design guide, Australian Institute of Architects. Eastchem 2017, Chemical products manufacturers, viewed 25/2/2017, . EC 2013, Use of common methods to measure and communicate the life cycle environmental performance of products and organisations, European Commission, http://eur- lex.europa.eu/legal-content/EN/TXT/PDF/?uri=CELEX:32013H0179&from=EN. EC 2014, Resource Efficiency Opportunities in the Building Sector., European Commission, Brussels, http://ec.europa.eu/environment/eussd/pdf/SustainableBuildingsCommunication.pdf. Ecoinvent 2013, Introduction to ecoinvent Version 3.1, viewed 28th of August 2014, . Economics, F 2015, Electricity market forecasts: 2015, Frontier Economics, Sydney, http://www.frontier-economics.com. Egan, AM 2011, ‘Air tightness of Australian offices buildings: reality versus typical assumption used in energy performance simulation’, in 12th Conference of International Building Performance Simulation Association, Sydney, Australia. Eisenhower, B, O'Neill, Z, Fonoberov, VA & Mezić, I 2012, ‘Uncertainty and sensitivity decomposition of building energy models’, Journal of Building Performance Simulation, vol. 5, no. 3, pp. 171-84. ELCD 2007, European Reference Life Cycle Database Directorate General Joint Research Centre (JRC), viewed 20/4/2015, . EN15978 2011, Sustainability of construction works : assessment of environmental performance of buildings : calculation method, British Standards Institution, United Kingdom. EnergyPlus 2017, Weather data sources, U.S. Department of Energy’s (DOE) Building Technologies Office (BTO), viewed 1/02/2017, . Esin, T 2007, ‘A study regarding the environmental impact analysis of the building materials production process (in Turkey)’, Building and Environment, vol. 42, no. 11, pp. 3860-71. eTool 2014, Life cycle assessment online tool, viewed 23/6/2015, . Faggianelli, GA, Mora, L & Merheb, R 2017, ‘Uncertainty quantification for Energy Savings Performance Contracting: Application to an office building’, Energy and Buildings, vol. 152, no. Supplement C, pp. 61-72. Fan, L & Khodadadi, JM 2011, ‘Thermal conductivity enhancement of phase change materials for thermal energy storage: A review’, Renewable and Sustainable Energy Reviews, vol. 15, no. 1, pp. 24-46. Farrance, I & Frenkel, R 2014, ‘Uncertainty in measurement: a review of Monte Carlo simulation using Microsoft Excel for the calculation of uncertainties through

253

functional relationships, including uncertainties in empirically derived constants’, The Clinical Biochemist Reviews, vol. 35, no. 1, p. 37. Fernando, D, Adey, B & Walbridge, S 2012, ‘Sustainable design of steel girders for a short- span bridge’, in First International Conference on Performance-Based and Life- Cycle Structural Engineering, pp. 1-10. Ferrari, S & Zanotto, V 2015, Building energy performance assessment in Southern Europe, Springer. Finkbeiner, M, Schau, EM, Lehmann, A & Traverso, M 2010, ‘Towards life cycle sustainability assessment’, Sustainability, vol. 2, no. 10, pp. 3309-22. Finnveden, G, Hauschild, MZ, Ekvall, T, Guinee, J, Heijungs, R, Hellweg, S, Koehler, A, Pennington, D & Suh, S 2009, ‘Recent developments in Life Cycle Assessment’, J Environ Manage, vol. 91, no. 1, pp. 1-21. Fitch, P & Cooper, JS 2005, ‘Life cycle modeling for adaptive and variant design part 2: case study’, Research in Engineering Design, vol. 15, no. 4, pp. 229-41. Flores, C 2015, How Australian buildings have achieved unprecedented energy savings, The National Australian Built Environment Rating System (NABERS), https://www.iea.org/media/workshops/2015/eeuevents/behave1103/S3flores.pdf. Flower, DM & Sanjayan, J 2007, ‘Green house gas emissions due to concrete manufacture’, The International Journal of Life Cycle Assessment, vol. 12, no. 5, pp. 282-8. Foraboschi, P, Mercanzin, M & Trabucco, D 2014, ‘Sustainable structural design of tall buildings based on embodied energy’, Energy and Buildings, vol. 68, pp. 254-69. Forsberg, A & von Malmborg, F 2004, ‘Tools for environmental assessment of the built environment’, Building and Environment, vol. 39, no. 2, pp. 223-8. Foucquier, A, Robert, S, Suard, F, Stéphan, L & Jay, A 2013, ‘State of the art in building modelling and energy performances prediction: A review’, Renewable and Sustainable Energy Reviews, vol. 23, no. Supplement C, pp. 272-88. Franzoni, E 2011, ‘Materials selection for green buildings: which tools for engineers and architects?’, Procedia Engineering, vol. 21, pp. 883-90. Frischknecht, R, Jungbluth, N, Althaus, H-J, Doka, G, Dones, R, Heck, T, Hellweg, S, Hischier, R, Nemecek, T & Rebitzer, G 2005, ‘The ecoinvent database: Overview and methodological framework (7 pp)’, The International Journal of Life Cycle Assessment, vol. 10, no. 1, pp. 3-9. Frischknecht, R, Jungbluth, N, Althaus, H-J, Hischier, R, Doka, G, Bauer, C, Dones, R, Nemecek, T, Hellweg, S & Humbert, S 2007, Implementation of life cycle impact assessment methods. Data v2. 0 (2007). Ecoinvent report No. 3, Ecoinvent Centre, Swiss Federal Laboratories for Materials Testing and Research (EMPA), Duebendorf (Switzerland). Fuller, S 2010, ‘Life-cycle cost analysis (LCCA)’, National Institute of Building Sciences, An Authoritative Source of Innovative Solutions for the Built Environment, vol. 1090. Fuller, S 2016, Life-Cycle Cost Analysis (LCCA), Whole Building Design Guide (WHDG)6/11/2017, . Furuta, H, Frangopol, DM & Akiyama, M 2014, Life-cycle of structural systems: Design, assessment, maintenance and management, CRC Press. Galli, A, Wiedmann, T, Ercin, E, Knoblauch, D, Ewing, B & Giljum, S 2012, ‘Integrating ecological, carbon and water footprint into a “footprint family” of indicators: definition and role in tracking human pressure on the planet’, Ecological indicators, vol. 16, pp. 100-12. Gan, VJ, Cheng, JC, Lo, IM & Chan, CM 2017, ‘Developing a CO 2-e accounting method for quantification and analysis of embodied carbon in high-rise buildings’, Journal of Cleaner Production, vol. 141, pp. 825-36.

254

Gartner, E 2004, ‘Industrially interesting approaches to “low-CO 2” cements’, Cement and Concrete research, vol. 34, no. 9, pp. 1489-98. GBCA 2014, Green Star-Rating Tools, Green Building Council of Australia (GBCA), viewed 16/8/2014, . Gilbert, BP, Hancock, SB, Bailleres, H & Hjiaj, M 2014, ‘Thin-walled timber structures: An investigation’, Construction and Building Materials, vol. 73, no. Supplement C, pp. 311-9. Glavind, M 2012, Guidelines for green concrete structures: Guide to good practice, Germany, 9782883941076, https://books.google.com.au/books?id=i0rkAwAAQBAJ. Goggins, J, Keane, T & Kelly, A 2010, ‘The assessment of embodied energy in typical reinforced concrete building structures in Ireland’, Energy and Buildings, vol. 42, no. 5, pp. 735-44. González, MJ & García Navarro, J 2006, ‘Assessment of the decrease of CO2 emissions in the construction field through the selection of materials: Practical case study of three houses of low environmental impact’, Building and Environment, vol. 41, no. 7, pp. 902-9. Grant, A, Ries, R & Thompson, C 2016, ‘Quantitative approaches in life cycle assessment— part 2—multivariate correlation and regression analysis’, The International Journal of Life Cycle Assessment, vol. 21, no. 6, pp. 912-9. Gravitt, D 2013, ‘Eco-efficient construction and building materials’, Construction Management and Economics, vol. 31, no. 11, pp. 1164-5. Gu, L, Lin, B, Zhu, Y, Gu, D, Huang, M & Gai, J 2008, ‘Integrated assessment method for building life cycle environmental and economic performance’, in Building Simulation, vol. 1, pp. 169-77. Guan, L 2009, ‘Implication of global warming on air-conditioned office buildings in Australia’, Building Research & Information, vol. 37, no. 1, pp. 43-54. Guggemos, A & Horvath, A 2005, ‘Comparison of environmental effects of steel- and concrete-framed buildings’, Journal of Infrastructure Systems, vol. 11, no. 2, pp. 93- 101. Gursel, AP, Maryman, H & Ostertag, C 2016, ‘A life-cycle approach to environmental, mechanical, and durability properties of “green” concrete mixes with rice husk ash’, Journal of Cleaner Production, vol. 112, pp. 823-36. Gustavsson, L, Joelsson, A & Sathre, R 2010, ‘Life cycle primary energy use and carbon emission of an eight-storey wood-framed apartment building’, Energy and Buildings, vol. 42, no. 2, pp. 230-42. Haapio, A & Viitaniemi, P 2008, ‘A critical review of building environmental assessment tools’, Environmental Impact Assessment Review, vol. 28, no. 7, pp. 469-82. Habert, G, d’Espinose de Lacaillerie, JB & Roussel, N 2011, ‘An environmental evaluation of geopolymer based concrete production: reviewing current research trends’, Journal of Cleaner Production, vol. 19, no. 11, pp. 1229-38. Habert, G & Roussel, N 2009, ‘Study of two concrete mix-design strategies to reach carbon mitigation objectives’, Cement and Concrete Composites, vol. 31, no. 6, pp. 397- 402. Hahn, T, Figge, F, Liesen, A & Barkemeyer, R 2010, ‘Opportunity cost based analysis of corporate eco-efficiency: A methodology and its application to the CO2-efficiency of German companies’, Journal of Environmental Management, vol. 91, no. 10, pp. 1997-2007. Hajdukiewicz, M, Byrne, D, Keane, MM & Goggins, J 2015, ‘Real-time monitoring framework to investigate the environmental and structural performance of buildings’, Building and Environment, vol. 86, pp. 1-16.

255

Hájek, P, Fiala, C & Kynčlová, M 2011, ‘Life cycle assessments of concrete structures–a step towards environmental savings’, Structural Concrete, vol. 12, no. 1, pp. 13-22. Hammond, G, Jones, C, Lowrie, F & Tse, P 2011, Embodied carbon: the inventory of carbon and energy (ICE), BSRIA. Hammond, GP & Jones, CI 2008, ‘Embodied energy and carbon in construction materials’, Proceedings of the Institution of Civil Engineers - Energy, vol. 161, no. 2, pp. 87- 98. Hammond, GP & Jones, CI 2010, ‘Embodied carbon: The concealed impact of residential construction’, in Global Warming, Springer, pp. 367-84. Hancock, GJ 2007, Design of Cold-formed Steel Structures: To Australian/New Zealand Standard AS/NZS 4600: 2005, Australian Steel Institute. Harish, VSKV & Kumar, A 2016, ‘A review on modeling and simulation of building energy systems’, Renewable and Sustainable Energy Reviews, vol. 56, no. Supplement C, pp. 1272-92. Harris, D, Harris, D, Fitzgerald, L & Fitzgerald, L 2017, ‘Life-cycle cost analysis (LCCA): a comparison of commercial flooring’, Facilities, vol. 35, no. 5/6, pp. 303-18. Heiselberg, P, Brohus, H, Hesselholt, A, Rasmussen, H, Seinre, E & Thomas, S 2009, ‘Application of sensitivity analysis in design of sustainable buildings’, Renewable Energy, vol. 34, no. 9, pp. 2030-6. Hensen, JL & Lamberts, R 2012, Building performance simulation for design and operation, Routledge. Hoes, P, Trcka, M, Hensen, JLM & Hoekstra Bonnema, B 2011, ‘Investigating the potential of a novel low-energy house concept with hybrid adaptable thermal storage’, Energy Conversion and Management, vol. 52, no. 6, pp. 2442-7. Hong, J, Shen, GQ, Peng, Y, Feng, Y & Mao, C 2016, ‘Uncertainty analysis for measuring greenhouse gas emissions in the building construction phase: a case study in China’, Journal of Cleaner Production, vol. 129, pp. 183-95. Hong, T, Ji, C & Park, H 2012, ‘Integrated model for assessing the cost and CO2 emission (IMACC) for sustainable structural design in ready-mix concrete’, J Environ Manage, vol. 103, pp. 1-8. Hong, T, Koo, C & Kwak, T 2013, ‘Framework for the implementation of a new renewable energy system in an educational facility’, Applied Energy, vol. 103, pp. 539-51. Hossain, MU, Poon, Chi S, Dong, Ya H & Xuan, D 2018, ‘Evaluation of environmental impact distribution methods for supplementary cementitious materials’, Renewable and Sustainable Energy Reviews, vol. 82, no. Part 1, pp. 597-608. Hossain, MU, Poon, CS, Lo, IM & Cheng, JC 2016, ‘Evaluation of environmental friendliness of concrete paving eco-blocks using LCA approach’, The International Journal of Life Cycle Assessment, vol. 21, no. 1, pp. 70-84. Huang, Y, Bird, R & Heidrich, O 2009, ‘Development of a life cycle assessment tool for construction and maintenance of asphalt pavements’, Journal of Cleaner Production, vol. 17, no. 2, pp. 283-96. Huang, Z, Ding, X, Sun, H & Liu, S 2010, ‘Identification of main influencing factors of life cycle CO2 emissions from the integrated steelworks using sensitivity analysis’, Journal of Cleaner Production, vol. 18, no. 10-11, pp. 1052-8. Huiskes, DMA, Keulen, A, Yu, QL & Brouwers, HJH 2016, ‘Design and performance evaluation of ultra-lightweight geopolymer concrete’, Materials & Design, vol. 89, pp. 516-26. Hunkeler, D, Lichtenvort, K & Rebitzer, G 2008, Environmental life cycle costing, CRC Press. Ibn-Mohammed, T, Greenough, R, Taylor, S, Ozawa-Meida, L & Acquaye, A 2013, ‘Operational vs. embodied emissions in buildings—A review of current trends’, Energy and Buildings, vol. 66, no. Supplement C, pp. 232-45. 256

ICANZ 2010, Insulation handbook part 1: Thermal Performance, 11/2010 edn, Insulation Council of Australia and New Zealand, 20/4/2017, http://icanz.org.au. Ingrao, C, Giudice, AL, Tricase, C, Mbohwa, C & Rana, R 2014, ‘The use of basalt aggregates in the production of concrete for the prefabrication industry: Environmental impact assessment, interpretation and improvement’, Journal of Cleaner Production, vol. 75, pp. 195-204. Inyim, P, Zhu, Y & Orabi, W 2016, ‘Analysis of Time, Cost, and Environmental Impact Relationships at the Building-Material Level’, Journal of Management in Engineering, vol. 32, no. 4, p. 04016005. IPCC 2014, Climate Change 2014–Impacts, Adaptation and Vulnerability: Regional Aspects, Cambridge University Press, Intergovernmental Panel on Climate Change. Ishak, SA & Hashim, H 2015, ‘Low carbon measures for cement plant – a review’, Journal of Cleaner Production, vol. 103, pp. 260-74. Islam, H, Jollands, M & Setunge, S 2015, ‘Life cycle assessment and life cycle cost implication of residential buildings—A review’, Renewable and Sustainable Energy Reviews, vol. 42, pp. 129-40. Islam, H, Jollands, M, Setunge, S & Bhuiyan, MA 2015, ‘Optimization approach of balancing life cycle cost and environmental impacts on residential building design’, Energy and Buildings, vol. 87, pp. 282-92. ISO14040 2006, Environmental management–life cycle assessment–principles and framework, International Organization for Standardization, Switzerland. ISO14067 2013, Greenhouse gases - Carbon footprint of products - Requirements and guidelines for quantification and communication, International Organization for Standardization, Switzerland. ISO15392 2008, Sustainability in building construction - General principles, International Organization for Standardization, Switzerland. ISO15673 2005, Guidelines for the simplified design of structural reinforced concrete for buildings, The International Organization for Standardization, Switzerland. ISO21929-1 2016, Sustainability in building construction – sustainability indicators – Part 1: Framework for the development of indicators for buildings, International Organization for Standardization, Switzerland. ISO21930 2007, Sustainability in building construction- Environmental declaration of building products, International Organization for Standardization, Switzerland. ISO21931-1 2010, Sustainability in building construction --Framework for methods of assessment of the environmental performance of construction works - Part 1: Buildings, International Organization for Standardization, Switzerland. Itsubo, N & Inaba, A 2003, ‘A new LCIA method: LIME has been completed’, The International Journal of Life Cycle Assessment, vol. 8, no. 5, p. 305. Iyer-Raniga, U, Moore, T & Wasiluk, K 2014, Residential building sustainability rating tools in Australia, Environment design guide, Australian Institute of Architects. Jacobs 2016, Retail electricity price history and projections - Public, Jacobs Australia Pty Limited, Melbourne, Australia. Jacoby, PC & Pelisser, F 2015, ‘Pozzolanic effect of porcelain polishing residue in Portland cement’, Journal of Cleaner Production, vol. 100, pp. 84-8. Jain, N, Lala, CR, Chopra, M, Ramallo-González, AP & Natarajan, S 2015, ‘Impacts of uncertainty in energy modelling widely used in aggressive energy efficiency regulations’. Javed, MF, Hafizah, N, Memon, SA, Jameel, M & Aslam, M 2017, ‘Recent research on cold-formed steel beams and columns subjected to elevated temperature: A review’, Construction and Building Materials, vol. 144, no. Supplement C, pp. 686-701.

257

Ji, C, Hong, T & Park, HS 2014, ‘Comparative analysis of decision-making methods for integrating cost and CO2 emission – focus on building structural design –’, Energy and Buildings, vol. 72, pp. 186-94. Jiao, Y, Lloyd, CR & Wakes, SJ 2012, ‘The relationship between total embodied energy and cost of commercial buildings’, Energy and Buildings, vol. 52, pp. 20-7. JMP 2017, statistical software, SAS Campus Drive, viewed 8/8/2016, . Joseph, P & Tretsiakova-McNally, S 2010, ‘Sustainable non-metallic building materials’, Sustainability, vol. 2, no. 2, pp. 400-27. Junnila, S & Horvath, A 2003, ‘Life-Cycle environmental effects of an office building’, Journal of Infrastructure Systems, vol. 9, no. 4, pp. 157-66. Kajaste, R & Hurme, M 2016, ‘Cement industry greenhouse gas emissions – management options and abatement cost’, Journal of Cleaner Production, vol. 112, pp. 4041-52. Kang, GS & Kren, A 2007, Structural engineering strategies towards sustainable design, http://seaonc.org. Karimpour, M, Belusko, M, Xing, K & Bruno, F 2014, ‘Minimising the life cycle energy of buildings: Review and analysis’, Building and Environment, vol. 73, pp. 106-14. Kellenberger, D & Althaus, H-J 2009, ‘Relevance of simplifications in LCA of building components’, Building and Environment, vol. 44, no. 4, pp. 818-25. Kelly, A 2017, On shaky ground: Divergent industry demand trends have led to division revenue fluctuations, IBISWorld, IBISWorld Industry Report E: Construction in Australia, www.ibisworld.com.au. Kestner, DM, Goupil, J & Lorenz, E 2010, Sustainability guidelines for the structural engineer, American Society of Civil Engineers (ASCE). Kibert, CJ 1994, ‘Establishing principles and a model for sustainable construction’, in Proceedings of the First International Conference on Sustainable Construction, pp. 6-9. Kibert, CJ 2012, Sustainable construction: green building design and delivery: green building design and delivery, John Wiley & Sons. Kim, HK, Jeon, JH & Lee, HK 2012, ‘Workability, and mechanical, acoustic and thermal properties of lightweight aggregate concrete with a high volume of entrained air’, Construction and Building Materials, vol. 29, pp. 193-200. Kim, J, Hong, T & Koo, C-W 2012, ‘Economic and environmental evaluation model for selecting the optimum design of green roof systems in elementary schools’, Environmental science & technology, vol. 46, no. 15, pp. 8475-83. Kim, TH & Chae, CU 2016, ‘Environmental Impact Analysis of Acidification and Eutrophication Due to Emissions from the Production of Concrete’, Sustainability, vol. 8, no. 6, p. 578. Kim, TH & Tae, SH 2016, ‘Proposal of Environmental Impact Assessment Method for Concrete in South Korea: An Application in LCA (Life Cycle Assessment)’, International journal of environmental research and public health, vol. 13, no. 11, p. 1074. Kloepffer, W 2008, ‘Life cycle sustainability assessment of products’, The International Journal of Life Cycle Assessment, vol. 13, no. 2, pp. 89-95. Klöpffer, W & Grahl, B 2014, Life Cycle Assessment (LCA), John Wiley & Sons. Knoeri, C, Sanyé-Mengual, E & Althaus, H-J 2013, ‘Comparative LCA of recycled and conventional concrete for structural applications’, The International Journal of Life Cycle Assessment, vol. 18, no. 5, pp. 909-18. Kofoworola, OF & Gheewala, SH 2008, ‘Environmental life cycle assessment of a commercial office building in Thailand’, The International Journal of Life Cycle Assessment, vol. 13, no. 6, p. 498. Kofoworola, OF & Gheewala, SH 2009, ‘Life cycle energy assessment of a typical office building in Thailand’, Energy and Buildings, vol. 41, no. 10, pp. 1076-83. 258

Kourmpanis, B, Papadopoulos, A, Moustakas, K, Kourmoussis, F, Stylianou, M & Loizidou, M 2008, ‘An integrated approach for the management of demolition waste in Cyprus’, Waste Management & Research, vol. 26, no. 6, pp. 573-81. Lamnatou, C, Notton, G, Chemisana, D & Cristofari, C 2014, ‘Life cycle analysis of a building-integrated solar thermal collector, based on embodied energy and embodied carbon methodologies’, Energy and Buildings, vol. 84, pp. 378-87. Langston, YL & Langston, CA 2008, ‘Reliability of building embodied energy modelling: an analysis of 30 Melbourne case studies’, Construction Management and Economics, vol. 26, no. 2, pp. 147-60. Lawania, K, Sarker, P & Biswas, W 2015, ‘Global warming implications of the use of by- products and recycled materials in Western Australia’s housing sector’, Materials, vol. 8, no. 10, pp. 6909-25. Lawania, KK & Biswas, WK 2016, ‘Cost-effective GHG mitigation strategies for Western Australia’s housing sector: a life cycle management approach’, Clean Technologies and Environmental Policy, vol. 18, no. 8, pp. 2419-28. Lee, M & White, G 2008, Embodied Energy LCA Assessment of replacement options for windows and traditional buildings. Crichton Carbon Centre. LEED 2012, What is LEED?, US Green Building Council,, Washington, DC, US. LEED 2016, LEED 2009 for new construction and major renovations, U.S. Green Building Council, Washington, DC, US http://www.usgbc.org/. Lemay, L & Leed, A 2011, Life cycle assessment of concrete buildings, www.nrmca.org. Lenzen, M 2000, ‘Errors in conventional and Input‐Output—based Life—Cycle inventories’, Journal of Industrial Ecology, vol. 4, no. 4, pp. 127-48. Lenzen, M & Crawford, R 2009, ‘The path exchange method for hybrid LCA’, Environmental science & technology, vol. 43, no. 21, pp. 8251-6. Lenzen, M & Dey, C 2000, ‘Truncation error in embodied energy analyses of basic iron and steel products’, Energy, vol. 25, no. 6, pp. 577-85. Leung, W, Noble, B, Gunn, J & Jaeger, JA 2015, ‘A review of uncertainty research in impact assessment’, Environmental Impact Assessment Review, vol. 50, pp. 116-23. Li, Y, Liu, Y, Gong, X, Nie, Z, Cui, S, Wang, Z & Chen, W 2016, ‘Environmental impact analysis of blast furnace slag applied to ordinary Portland cement production’, Journal of Cleaner Production, vol. 120, no. Supplement C, pp. 221-30. Liu, M, Zhang, Y & Wang, Y 2015, ‘Influence of environment conditions on carbon emission of fly ash concrete’, Materials Research Innovations, vol. 19, no. sup9, pp. S9-216-S9-9. Mainey, AJ, Gilbert, BP, Fernando, D & Bailleres, H 2015, ‘Thin-walled timber and FRP- timber veneer composite CEE-sections’, in International Conference on Performance-based and Life-cycle Structural Engineering, pp. 1443-52. Majeau-Bettez, G, Strømman, AH & Hertwich, EG 2011, ‘Evaluation of process-and input– output-based life cycle inventory data with regard to truncation and aggregation issues’, Environmental science & technology, vol. 45, no. 23, pp. 10170-7. Malhotra, V & Mehta, P 2002, ‘High-performance, high-volume fly ash concrete’, Concrete International, vol. 24, no. 7, pp. 30-4. Malhotra, VM 2006, ‘Reducing CO2 emissions’, Concrete international, vol. 28, no. 9, pp. 42-5. Marinković, S, Malešev, M & Ignjatović, I 2014, ‘Life cycle assessment (LCA) of concrete made using recycled concrete or natural aggregates-11’. Marinkovic, S, Radonjanin, V, Malesev, M & Ignjatovic, I 2010, ‘Comparative environmental assessment of natural and recycled aggregate concrete’, Waste Manag, vol. 30, no. 11, pp. 2255-64.

259

Marinković, S, Radonjanin, V, Malešev, M & Ignjatović, I 2010, ‘Comparative environmental assessment of natural and recycled aggregate concrete’, Waste Management, vol. 30, no. 11, pp. 2255-64. Marsh, B 2003, ‘High volume fly ash concrete’, Concrete, vol. 37, no. 4, pp. 54-5. Maydl, P 2004, ‘Sustainable engineering: state-of-the-art and prospects’, Structural Engineering International, vol. 14, no. 3, pp. 176-80. McLellan, B, Corder, G, Giurco, D & Ishihara, K 2012, ‘Renewable energy in the minerals industry: a review of global potential’, Journal of Cleaner Production, vol. 32, pp. 32-44. McLeod, R 2005, ‘Ordinary portland cement-with extraordinarily high CO2 emissions’, BFF Autumn, pp. 30-3. McRobert, J 2010, Sustainable aggregates – CO2 emission factor study, ARRB Group, August 2010, http://www.sustainableaggregates.com.au/docs/SASA_C02_Emission_Factor_Study .pdf. Mearig, T, Coffee, N & Morgan, M 1999, Life cycle cost analysis handbook:Draft, State of Alaska, Department of Education and Early Development, Education Support Services/Facilities. Mehta, PK 2002, ‘Greeningof the Concrete Industry for Sustainable Development’, Concrete international, vol. 23. Menzies, GF, Turan, S & Banfill, PF 2007, ‘Life-cycle assessment and embodied energy: a review’, Proceedings of the Institution of Civil Engineers-Construction Materials, vol. 160, no. 4, pp. 135-44. Michael, B 2013, ‘Assessment and adaptation of an appropriate green building rating system for Nigeria’, Journal of Environment and Earth Science, vol. 3, no. 1, pp. 1-10. Miller, A & Ip, K 2013, ‘Sustainable construction materials’, in Design and Management of Sustainable Built Environments, Springer, pp. 341-58. Miller, D, Doh, J-H, Guan, H, Mulvey, M, Fragomeni, S, McCarthy, T & Peters, T 2013, Environmental impact assessment of post tensioned and reinforced concrete slab construction, The 22nd Australasian Conference on the Mechanics of Structures and Materials - From Materials to Structures: Advancement through Innovation, Taylor & Francis Group. Miller, SA, Monteiro, PJM, Ostertag, CP & Horvath, A 2016, ‘Concrete mixture proportioning for desired strength and reduced global warming potential’, Construction and Building Materials, vol. 128, no. Supplement C, pp. 410-21. Mitchell, LM 2010, ‘Green Star and NABERS: learning from the Australian experience with green building rating tools’, Energy Efficient Cities: Assessment Tools and Benchmarking Practices, pp. 93-124. Mohammadi, J & South, W 2017, ‘Life cycle assessment (LCA) of benchmark concrete products in Australia’, The International Journal of Life Cycle Assessment, pp. 1-21. Mokhlesian, S & Holmén, M 2012, ‘Business model changes and green construction processes’, Construction Management and Economics, vol. 30, no. 9, pp. 761-75. Mokhtar, SN, Mahmood, NZ, Che Hassan, CR, Masudi, AF & Sulaiman, NM 2011, ‘Factors that contribute to the generation of construction waste at sites’, Advanced Materials Research, vol. 163, pp. 4501-7. Moncaster, AM & Symons, KE 2013, ‘A method and tool for ‘cradle to grave’ embodied carbon and energy impacts of UK buildings in compliance with the new TC350 standards’, Energy and Buildings, vol. 66, pp. 514-23. Moon, KS 2008, ‘Sustainable structural engineering strategies for tall buildings’, The Structural Design of Tall and Special Buildings, vol. 17, no. 5, pp. 895-914. Moussavi Nadoushani, ZS & Akbarnezhad, A 2015, ‘Effects of structural system on the life cycle carbon footprint of buildings’, Energy and Buildings, vol. 102, pp. 337-46. 260

Mujagic, JU, Dolan, JD, Ekwueme, CG, Fanella, D & LaBoube, RA 2012, Structural Design of Low-rise Buildings in Cold-formed Steel, Reinforced Masonry, and Structural Timber, McGraw Hill Professional. Nadoushani, ZM & Akbarnezhad, A 2015, ‘Comparative analysis of embodied carbon associated with alternative structural systems’, in ISARC. Proceedings of the International Symposium on Automation and Robotics in Construction, vol. 32, p. 1. Neville, AM 2012, Properties of concrete, Prentice Hall, London, UK. Newlands, M, Jones, M, McCarthy, M & Zheng, L 2012, ‘Using fly ash to achieve low embodied CO2 concrete’, in Proceedings of EUROCOALASH 2012 Conference. Nguyen, A-T, Reiter, S & Rigo, P 2014, ‘A review on simulation-based optimization methods applied to building performance analysis’, Applied Energy, vol. 113, no. Supplement C, pp. 1043-58. Nguyen, BK & Altan, H 2011, ‘Comparative review of five sustainable rating systems’, Procedia Engineering, vol. 21, no. 0, pp. 376-86. NHSC 2011, National Housing Supply Council: State of Supply Report, Key findings of the 2011, National Housing Supply Council (NHSP), Canberra. Norman, J, MacLean, H & Kennedy, C 2006, ‘Comparing high and low residential density: Life-cycle analysis of energy use and greenhouse gas emissions’, Journal of Urban Planning and Development, vol. 132, no. 1, pp. 10-21. NS11401.0 2014, Labelling and declaration of environmental attributes of building products- Type III environmental declarations, National Standards Development Organisation Limited. NS11401.1 2014, Labelling and declaration of environmental attributes of building products- Type III environmental declarations, National Standards Development Organisation Limited, Australia. Nuaklong, P, Sata, V & Chindaprasirt, P 2016, ‘Influence of recycled aggregate on fly ash geopolymer concrete properties’, Journal of Cleaner Production, vol. 112, pp. 2300-7. O'Moore, LM & O'Brien, KR 2009, ‘Impact of supplementary cementitious material content and transportation distance on greenhouse gas emissions embodied in concrete’, in 24th Biennial Conference of Concrete Institute of Australia (Concrete 09), pp. 1-9. O’Brien, KR, Ménaché, J & O’Moore, LM 2009, ‘Impact of fly ash content and fly ash transportation distance on embodied greenhouse gas emissions and water consumption in concrete’, The International Journal of Life Cycle Assessment, vol. 14, no. 7, pp. 621-9. O’Brien, M, Doig, A & Clift, R 1996, ‘Social and environmental life cycle assessment (SELCA)’, The International Journal of Life Cycle Assessment, vol. 1, no. 4, pp. 231-7. Obuzor, GN, Kinuthia, JM & Robinson, RB 2011, ‘Enhancing the durability of flooded low- capacity soils by utilizing lime-activated ground granulated blastfurnace slag (GGBS)’, Engineering Geology, vol. 123, no. 3, pp. 179-86. Oh, BK, Park, JS, Choi, SW & Park, HS 2016, ‘Design model for analysis of relationships among CO2 emissions, cost, and structural parameters in green building construction with composite columns’, Energy and Buildings, vol. 118, pp. 301-15. Onat, NC, Kucukvar, M & Tatari, O 2014, ‘Scope-based carbon footprint analysis of US residential and commercial buildings: an input–output hybrid life cycle assessment approach’, Building and Environment, vol. 72, pp. 53-62. Optis, M & Wild, P 2010, ‘Inadequate documentation in published life cycle energy reports on buildings’, The International Journal of Life Cycle Assessment, vol. 15, no. 7, pp. 644-51.

261

Ortiz, O, Castells, F & Sonnemann, G 2009, ‘Sustainability in the construction industry: A review of recent developments based on LCA’, Construction and Building Materials, vol. 23, no. 1, pp. 28-39. Ozoemena, M, Cheung, WM & Hasan, R 2017, ‘Improving uncertainty analysis of embodied energy and embodied carbon in wind turbine design’, The International Journal of Advanced Manufacturing Technology, pp. 1-13. Pachauri, RK, Allen, MR, Barros, VR, Broome, J, Cramer, W, Christ, R, Church, JA, Clarke, L, Dahe, Q & Dasgupta, P 2014, Climate change 2014: synthesis report. Contribution of Working Groups I, II and III to the fifth assessment report of the Intergovernmental Panel on Climate Change, IPCC. Panesar, DK, Seto, KE & Churchill, CJ 2017, ‘Impact of the selection of functional unit on the life cycle assessment of green concrete’, The International Journal of Life Cycle Assessment, pp. 1-18. Park, J, Tae, S & Kim, T 2012, ‘Life cycle CO2 assessment of concrete by compressive strength on construction site in Korea’, Renewable and Sustainable Energy Reviews, vol. 16, no. 5, pp. 2940-6. PAS2050 2011, Specifi cation for the assessment of the life cycle greenhouse gas emissions of goods and services, British Standards Institution. PCA 2015, Global cement consumption on the rise, Portland Cement Association, . Peereboom, EC, Kleijn, R, Lemkowitz, S & Lundie, S 1998, ‘Influence of inventory data sets on life‐cycle assessment results: a case study on PVC’, Journal of Industrial Ecology, vol. 2, no. 3, pp. 109-30. Peña-Mora, F, Ahn, C, Golparvar-Fard, M, Hajibabai, L, Shiftehfar, S, An, S & Aziz, Z 2009, ‘A framework for managing emissions from construction processes’, in Proc., Int. Conf. & Workshop on Sustainable Green Bldg. Design & Construction, National Science Foundation. Pérez-Lombard, L, Ortiz, J, Coronel, JF & Maestre, IR 2011, ‘A review of HVAC systems requirements in building energy regulations’, Energy and Buildings, vol. 43, no. 2-3, pp. 255-68. Pérez-Lombard, L, Ortiz, J, González, R & Maestre, IR 2009, ‘A review of benchmarking, rating and labelling concepts within the framework of building energy certification schemes’, Energy and Buildings, vol. 41, no. 3, pp. 272-8. Pineda, P, García-Martínez, A & Castizo-Morales, D 2017, ‘Environmental and structural analysis of cement-based vs. natural material-based grouting mortars. Results from the assessment of strengthening works’, Construction and Building Materials, vol. 138, no. Supplement C, pp. 528-47. Pitt&Sherry 2012, Baseline energy consumption and greenhouse gas emissions In commercial buildings in Australia, Council of Australian Governments (COAG) National Strategy on Energy Efficiency. Poinssot, C, Bourg, S, Ouvrier, N, Combernoux, N, Rostaing, C, Vargas-Gonzalez, M & Bruno, J 2014, ‘Assessment of the environmental footprint of nuclear energy systems. Comparison between closed and open fuel cycles’, Energy, vol. 69, pp. 199-211. Ponechal, R & Staffenova, D 2017, ‘Impact of external wall insulation thickness on internal surface temperature behaviour’, in MATEC Web of Conferences, vol. 117, p. 00140. Pongiglione, M & Calderini, C 2016, ‘Sustainable structural design: Comprehensive literature review’, Journal of Structural Engineering, vol. 142, no. 12, p. 04016139. Poon, C-S, Yu, AT & Jaillon, L 2004, ‘Reducing building waste at construction sites in Hong Kong’, Construction Management and Economics, vol. 22, no. 5, pp. 461-70.

262

Proske, T, Hainer, S, Rezvani, M & Graubner, C-A 2013, ‘Eco-friendly concretes with reduced water and cement contents — Mix design principles and laboratory tests’, Cement and Concrete Research, vol. 51, no. Supplement C, pp. 38-46. Protocol, GG 2011, Product life cycle accounting and reporting standard. Pushkar, S & Verbitsky, O 2016, ‘Effects of different allocation approaches for modeling mineral additives in blended cements on environmental damage from five concrete mixtures in Israel’, Materials and Structures, vol. 49, no. 10, pp. 4401-15. Radu, AL, Scrieciu, MA & Caracota, DM 2013, ‘Carbon footprint analysis: towards a projects evaluation model for promoting sustainable development’, Procedia Economics and Finance, vol. 6, pp. 353-63. Rahman, A, Rasul, M, Khan, MMK & Sharma, S 2015, ‘Recent development on the uses of alternative fuels in cement manufacturing process’, Fuel, vol. 145, pp. 84-99. Rahman, MM, Rasul, M & Khan, MMK 2010, ‘Energy conservation measures in an institutional building in sub-tropical climate in Australia’, Applied Energy, vol. 87, no. 10, pp. 2994-3004. Rahman, MM, Rasul, M & Khan, MMK 2011, ‘Feasibility of thermal energy storage systems in an institutional building in subtropical climates in Australia’, Applied Thermal Engineering, vol. 31, no. 14, pp. 2943-50. Rajagopalan, N, Bilec, MM & Landis, AE 2012, ‘Life cycle assessment evaluation of green product labeling systems for residential construction’, The International Journal of Life Cycle Assessment, vol. 17, no. 6, pp. 753-63. Ramage, M, Foster, R, Smith, S, Flanagan, K & Bakker, R 2017, ‘Super Tall Timber: design research for the next generation of natural structure’, The Journal of Architecture, vol. 22, no. 1, pp. 104-22. Ramesh, T, Prakash, R & Shukla, KK 2010, ‘Life cycle energy analysis of buildings: An overview’, Energy and Buildings, vol. 42, no. 10, pp. 1592-600. Randl, N, Steiner, T, Ofner, S, Baumgartner, E & Mészöly, T 2014, ‘Development of UHPC mixtures from an ecological point of view’, Construction and Building Materials, vol. 67, no. Part C, pp. 373-8. Rao, A, Jha, KN & Misra, S 2007, ‘Use of aggregates from recycled construction and demolition waste in concrete’, Resources, conservation and Recycling, vol. 50, no. 1, pp. 71-81. Rauf, A & Crawford, RH 2012, ‘The effect of material service life on the life cycle energy of residential buildings’, in ASA2012: The 46th Annual Conference of the Architectural Science Association (formerly ANZAScA)–Building on Knowledge: Theory and Practice. Rauf, A & Crawford, RH 2015, ‘Building service life and its effect on the life cycle embodied energy of buildings’, Energy, vol. 79, no. Supplement C, pp. 140-8. Rawlinsons 2016, Rawlinsons construction cost guide 2016. RBA 2016, Reserve Bank of Australia—Inflation target., viewed 20/20/2017, . Reap, J, Roman, F, Duncan, S & Bras, B 2008, ‘A survey of unresolved problems in life cycle assessment’, The International Journal of Life Cycle Assessment, vol. 13, no. 5, p. 374. Rebitzer, G, Ekvall, T, Frischknecht, R, Hunkeler, D, Norris, G, Rydberg, T, Schmidt, W-P, Suh, S, Weidema, BP & Pennington, D 2004, ‘Life cycle assessment: Part 1: Framework, goal and scope definition, inventory analysis, and applications’, Environment international, vol. 30, no. 5, pp. 701-20. Renouf, MA, Grant, T, Sevenster, M, Ridoutt, B, Ximenes, F, Logie, J, Bengtsson, J, Cowie, A & Lane, J 2015, Best practice guide to Life Cycle Impact Assessment (LCIA) in Australia.

263

Richman, R, Pasqualini, P & Kirsh, A 2009, ‘Life-Cycle analysis of roofing insulation levels for cold storage buildings’, Journal of Architectural Engineering, vol. 15, no. 2, pp. 55-61. Rincón, L, Castell, A, Pérez, G, Solé, C, Boer, D & Cabeza, LF 2013, ‘Evaluation of the environmental impact of experimental buildings with different constructive systems using Material Flow Analysis and Life Cycle Assessment’, Applied Energy, vol. 109, no. 0, pp. 544-52. Roaf, S, Crichton, D & Nicol, F 2009, Adapting buildings and cities for climate change: a 21st century survival guide, Routledge. Robati, M, Kokogiannakis, G & McCarthy, TJ 2017, ‘Impact of structural design solutions on the energy and thermal performance of an Australian office building’, Building and Environment, vol. 124, pp. 258-82. Robati, M, McCarthy, TJ & Kokogiannakis, G 2016, ‘Incorporating environmental evaluation and thermal properties of concrete mix designs’, Construction and Building Materials, vol. 128, pp. 422-35. Robertson, AB, Lam, FC & Cole, RJ 2012, ‘A comparative cradle-to-gate life cycle assessment of mid-rise office building construction alternatives: laminated timber or reinforced concrete’, Buildings, vol. 2, no. 3, pp. 245-70. Roberz, F, Loonen, RCGM, Hoes, P & Hensen, JLM 2017, ‘Ultra-lightweight concrete: Energy and comfort performance evaluation in relation to buildings with low and high thermal mass’, Energy and Buildings, vol. 138, pp. 432-42. Robinson, H, Symonds, B, Gilbertson, B & Ilozor, B 2015, ‘Design Economics for the Built Environment Impact of Sustainability on Project Evaluation’. Rossi, B 2014, ‘Discussion on the use of stainless steel in constructions in view of sustainability’, Thin-Walled Structures, vol. 83, pp. 182-9. Rouwette, R 2012, ACM – LCA of Geopolymer Concrete (E-Crete), start2see Pty Ltd, Melbourne, Australia, http://www.acm.com.au/. Ruuska, AP & Häkkinen, TM 2015, ‘The significance of various factors for GHG emissions of buildings’, International Journal of Sustainable Engineering, vol. 8, no. 4-5, pp. 317-30. Ryan, EM & Sanquist, TF 2012, ‘Validation of building energy modeling tools under idealized and realistic conditions’, Energy and Buildings, vol. 47, pp. 375-82. Saade, MRM, Silva, MGd & Gomes, V 2015, ‘Appropriateness of environmental impact distribution methods to model blast furnace slag recycling in cement making’, Resources, Conservation and Recycling, vol. 99, no. Supplement C, pp. 40-7. Saghafi, MD & Teshnizi, ZSH 2011, ‘Recycling value of building materials in building assessment systems’, Energy and Buildings, vol. 43, no. 11, pp. 3181-8. Saling, P, Kicherer, A, Dittrich-Krämer, B, Wittlinger, R, Zombik, W, Schmidt, I, Schrott, W & Schmidt, S 2002, ‘Eco-efficiency analysis by BASF: the method’, The International Journal of Life Cycle Assessment, vol. 7, no. 4, pp. 203-18. Sarkisian, M, Hu, L & Shook, D 2012, ‘Mapping a structure's impact on the environment’, in Structures Congress 2012, pp. 898-909. Sartori, I & Hestnes, AG 2007, ‘Energy use in the life cycle of conventional and low-energy buildings: A review article’, Energy and Buildings, vol. 39, no. 3, pp. 249-57. Scheuer, C, Keoleian, GA & Reppe, P 2003, ‘Life cycle energy and environmental performance of a new university building: modeling challenges and design implications’, Energy and Buildings, vol. 35, no. 10, pp. 1049-64. Schmidt, J & Griffin, C 2013, ‘Barriers to the design and use of cross-laminated timber structures in high-rise multi-family housing in the United States’, Portugal: Guimaraes. Seto, KE 2015, ‘Life Cycle Assessment and Environmental Efficiency of Concrete Materials’, University of Toronto (Canada). 264

Shahrestani, M, Yao, R & Cook, GK 2013, ‘A review of existing building benchmarks and the development of a set of reference office buildings for England and Wales’, Intelligent Buildings International, vol. 6, no. 1, pp. 41-64. Shahrestani, M, Yao, R & Cook, GK 2014, ‘A review of existing building benchmarks and the development of a set of reference office buildings for England and Wales’, Intelligent Buildings International, vol. 6, no. 1, pp. 41-64. Sharifi, A & Murayama, A 2013, ‘A critical review of seven selected neighborhood sustainability assessment tools’, Environmental Impact Assessment Review, vol. 38, pp. 73-87. Sharma, A, Saxena, A, Sethi, M & Shree, V 2011, ‘Life cycle assessment of buildings: a review’, Renewable and Sustainable Energy Reviews, vol. 15, no. 1, pp. 871-5. Shen, W, Cao, L, Li, Q, Zhang, W, Wang, G & Li, C 2015, ‘Quantifying CO 2 emissions from China’s cement industry’, Renewable and Sustainable Energy Reviews, vol. 50, pp. 1004-12. Silva, AS & Ghisi, E 2014, ‘Uncertainty analysis of user behaviour and physical parameters in residential building performance simulation’, Energy and Buildings, vol. 76, pp. 381-91. Silvestre, J, De Brito, J & Pinheiro, MD 2014, ‘Environmental impacts and benefits of the end-of-life of building materials–calculation rules, results and contribution to a “cradle to cradle” life cycle’, Journal of Cleaner Production, vol. 66, pp. 37-45. Singh, A, Berghorn, G, Joshi, S & Syal, M 2011, ‘Review of Life-Cycle Assessment Applications in Building Construction’, Journal of Architectural Engineering, vol. 17, no. 1, pp. 15-23. Skullestad, JL, Bohne, RA & Lohne, J 2016, ‘High-rise Timber Buildings as a Climate Change Mitigation Measure – A Comparative LCA of Structural System Alternatives’, Energy Procedia, vol. 96, pp. 112-23. Srinivasreddy, AB, McCarthy, TJ & Lume, E 2013, ‘Effect of rice husk ash on workability and strength of concrete’. Stephan, A 2013, ‘Towards a comprehensive energy assessment of residential buildings: a multi-scale life cycle energy analysis framework’. Stephan, A & Crawford, RH 2014, ‘A multi-scale life-cycle energy and greenhouse-gas emissions analysis model for residential buildings’, Architectural Science Review, vol. 57, no. 1, pp. 39-48. Stephan, A & Stephan, L 2016, ‘Life cycle energy and cost analysis of embodied, operational and user-transport energy reduction measures for residential buildings’, Applied Energy, vol. 161, no. Supplement C, pp. 445-64. Storey, JB 2008, Construction materials stewardship: The status Quo in selected countries: CIB W115, Centre For Building Performance Research, Victoria University of Wellington. Suh, S, Lenzen, M, Treloar, GJ, Hondo, H, Horvath, A, Huppes, G, Jolliet, O, Klann, U, Krewitt, W & Moriguchi, Y 2004, ‘System boundary selection in life-cycle inventories using hybrid approaches’, Environmental science & technology, vol. 38, no. 3, pp. 657-64. Suh, S, Tomar, S, Leighton, M & Kneifel, J 2014, ‘Environmental performance of green building code and certification systems’, Environ Sci Technol, vol. 48, no. 5, pp. 2551-60. Survey, USG 2015, Mineral commodity summaries 2015, U.S. Geological Survey, Washington. Suzuki, M & Oka, T 1998, ‘Estimation of life cycle energy consumption and CO2 emission of office buildings in Japan’, Energy and Buildings, vol. 28, no. 1, pp. 33-41.

265

Taborianski, VM & Prado, RT 2004, ‘Comparative evaluation of the contribution of residential water heating systems to the variation of greenhouse gases stock in the atmosphere’, Building and Environment, vol. 39, no. 6, pp. 645-52. Tait, MW & Cheung, WM 2016, ‘A comparative cradle-to-gate life cycle assessment of three concrete mix designs’, The International Journal of Life Cycle Assessment, vol. 21, no. 6, pp. 847-60. Takano, A, Hughes, M & Winter, S 2014, ‘A multidisciplinary approach to sustainable building material selection: A case study in a Finnish context’, Building and Environment, vol. 82, pp. 526-35. Takano, A, Winter, S, Hughes, M & Linkosalmi, L 2014, ‘Comparison of life cycle assessment databases: A case study on building assessment’, Building and Environment, vol. 79, pp. 20-30. Tam, VWY & Tam, CM 2006, ‘A review on the viable technology for construction waste recycling’, Resources, Conservation and Recycling, vol. 47, no. 3, pp. 209-21. Taylor, M & Collins, D 2006, ‘Novel cements: low energy, low carbon cements’, Fact Sheet, vol. 12. Teixeira, ER, Mateus, R, Camões, AF, Bragança, L & Branco, FG 2016, ‘Comparative environmental life-cycle analysis of concretes using biomass and coal fly ashes as partial cement replacement material’, Journal of Cleaner Production, vol. 112, no. Part 4, pp. 2221-30. Tharumarajah, A & Grant, T 2006, ‘Australian national life cycle inventory database: moving forward’, in proceedings of the 5 th ALCAS Conference, Melbourne, Australia. Thormark, C 2002, ‘A low energy building in a life cycle—its embodied energy, energy need for operation and recycling potential’, Building and Environment, vol. 37, no. 4, pp. 429-35. Thormark, C 2006, ‘The effect of material choice on the total energy need and recycling potential of a building’, Building and Environment, vol. 41, no. 8, pp. 1019-26. Thormark, C 2007, ‘Motives for design for disassembly in building construction’, in International congress sustainable construction, materials and practices challenge of the industry for the new millennium, Lisbon. Tian, W 2013, ‘A review of sensitivity analysis methods in building energy analysis’, Renewable and Sustainable Energy Reviews, vol. 20, pp. 411-9. Tian, W & de Wilde, P 2011, ‘Uncertainty and sensitivity analysis of building performance using probabilistic climate projections: A UK case study’, Automation in Construction, vol. 20, no. 8, pp. 1096-109. Torgal, FP & Jalali, S 2011, Eco-efficient construction and building materials, London Limited: Springer Verlag. Tošić, N, Marinković, S, Dašić, T & Stanić, M 2015, ‘Multicriteria optimization of natural and recycled aggregate concrete for structural use’, Journal of Cleaner Production, vol. 87, pp. 766-76. Treloar, GJ 1997, ‘Extracting embodied energy paths from input–output tables: towards an input–output-based hybrid energy analysis method’, Economic Systems Research, vol. 9, no. 4, pp. 375-91. Tsimplokoukou, K, Lamperti, M & Negro, P 2014, ‘Building design for safety and sustainability’, JRC Science and Policy Reports. doi, vol. 10, p. 338223. Turk, J, Cotič, Z, Mladenovič, A & Šajna, A 2015, ‘Environmental evaluation of green concretes versus conventional concrete by means of LCA’, Waste Management, vol. 45, no. Supplement C, pp. 194-205. UNFCCC 2008, Kyoto Protocol Reference Manual on Accounting of Emissions and Assigned Amounts, Germany.

266

Van den Heede, P & De Belie, N 2012, ‘Environmental impact and life cycle assessment (LCA) of traditional and ‘green’ concretes: Literature review and theoretical calculations’, Cement and Concrete Composites, vol. 34, no. 4, pp. 431-42. Van den Heede, P & De Belie, N 2014, ‘A service life based global warming potential for high-volume fly ash concrete exposed to carbonation’, Construction and Building Materials, vol. 55, no. Supplement C, pp. 183-93. Van Deventer, JS, Provis, JL & Duxson, P 2012, ‘Technical and commercial progress in the adoption of geopolymer cement’, Minerals Engineering, vol. 29, pp. 89-104. Vieira, PS & Horvath, A 2008, ‘Assessing the end-of-life impacts of buildings’, Environmental science & technology, vol. 42, no. 13, pp. 4663-9. Virgalitte, SJ, Luther, MD, Rose, JH, Mather, B, Bell, LW, Ehmke, BA, Klieger, P, Roy, DM, Call, BM & Hooton, RD 1995, ‘Ground Granulated Blast-Furnace Slag as a Cementitious Constituent in Concrete’, American Concrete Institute ACI Report 233R-95. Wang, E & Shen, Z 2013, ‘A hybrid Data Quality Indicator and statistical method for improving uncertainty analysis in LCA of complex system–application to the whole- building embodied energy analysis’, Journal of Cleaner Production, vol. 43, pp. 166-73. Wang, S & Wu, H 2006, ‘Environmental-benign utilisation of fly ash as low-cost adsorbents’, Journal of Hazardous Materials, vol. 136, no. 3, pp. 482-501. Wang, S, Yan, C & Xiao, F 2012, ‘Quantitative energy performance assessment methods for existing buildings’, Energy and Buildings, vol. 55, no. 0, pp. 873-88. Wang, Y, Jones, P, Plass, N, Kaltenegger, I & Wheeler, J 2007, A Strategy for Sustainable Housing Development in Xi'an, Welsh School of Architecture. Warner, RF, Foster, SJ & Kilpatrick, AE 2010, ‘Reinforced concrete basics: analysis and design of reinforced concrete structures’. Watson, P & Jones, D 2005, ‘Redefining life cycle for a building sustainability assessment framework’, paper presented to Fourth Australian Life Cycle Assessment Conference . Sydney, Australia. Webster, MD 2004, ‘Relevance of structural engineers to sustainable design of buildings’, Structural engineering international, vol. 14, no. 3, pp. 181-5. Whitehead, B, Andrews, D, Shah, A & Maidment, G 2014, ‘Assessing the environmental impact of data centres part 1: Background, energy use and metrics’, Building and Environment, vol. 82, pp. 151-9. Whitehead, B, Andrews, D, Shah, A & Maidment, G 2014, ‘Assessing the environmental impact of data centres part 2: Building environmental assessment methods and life cycle assessment’, Building and Environment, vol. Volume 93, pp. Pages 395-405. Whiting, B, McCarthy, T & Lume, E 2012, ‘Drying shrinkage of concrete made from recycled concrete aggregate’, paper presented to 22nd Australasian Conference on the Mechanics of Structures and Materials Sydney, Australia. Wimpenny, D 2009, ‘Low carbon concrete-options for the next generation of infrastructure’, Concrete Solutions, vol. 9, pp. 41-1. Worrell, E 2008, Energy efficiency improvement and cost saving opportunities for cement making. An Energy Star Guide for Energy and Plant Managers, Lawrence Berkeley National Laboratory, Environmental Energy Technologies Division, U.S. Wu, P, Xia, B, Pienaar, J & Zhao, X 2014, ‘The past, present and future of carbon labelling for construction materials–a review’, Building and Environment, vol. 77, pp. 160-8. Wu, X, Zhang, Z & Chen, Y 2005, ‘Study of the environmental impacts based on the “green tax”—applied to several types of building materials’, Building and Environment, vol. 40, no. 2, pp. 227-37. Wu, Y, Wang, J-Y, Monteiro, PJM & Zhang, M-H 2015, ‘Development of ultra-lightweight cement composites with low thermal conductivity and high specific strength for 267

energy efficient buildings’, Construction and Building Materials, vol. 87, pp. 100- 12. Ximenes, FA & Grant, T 2013, ‘Quantifying the greenhouse benefits of the use of wood products in two popular house designs in Sydney, Australia’, The International Journal of Life Cycle Assessment, vol. 18, no. 4, pp. 891-908. Y. J. Siew, R, C. A. Balatbat, M & G. Carmichael, D 2013, ‘A review of building/infrastructure sustainability reporting tools (SRTs)’, Smart and Sustainable Built Environment, vol. 2, no. 2, pp. 106-39. Yalcintas, M 2008, ‘Energy-savings predictions for building-equipment retrofits’, Energy and Buildings, vol. 40, no. 12, pp. 2111-20. Yan, H, Shen, Q, Fan, LCH, Wang, Y & Zhang, L 2010, ‘Greenhouse gas emissions in building construction: A case study of One Peking in Hong Kong’, Building and Environment, vol. 45, no. 4, pp. 949-55. Yang, D, Fan, L, Shi, F, Liu, Q & Wang, Y 2017, ‘Comparative study of cement manufacturing with different strength grades using the coupled LCA and partial LCC methods—A case study in China’, Resources, Conservation and Recycling, vol. 119, no. Supplement C, pp. 60-8. Yepes, V, Martí, JV & García-Segura, T 2015, ‘Cost and CO2 emission optimization of precast–prestressed concrete U-beam road bridges by a hybrid glowworm swarm algorithm’, Automation in Construction, vol. 49, pp. 123-34. Yokoo, N, Terachima, T & Oka, T 2013, ‘Embodied energy and CO2 emission associated with building construction by using I/O based data and process based data in Japan’, in SB13 Graz conference, Germany, pp. 311-6. Yu, M, Wiedmann, T, Crawford, R & Tait, C 2017, ‘The Carbon Footprint of Australia's Construction Sector’, Procedia Engineering, vol. 180, pp. 211-20. Yu, QL, Spiesz, P & Brouwers, HJH 2015, ‘Ultra-lightweight concrete: Conceptual design and performance evaluation’, Cement and Concrete Composites, vol. 61, pp. 18-28. Yu, R, Spiesz, P & Brouwers, H 2015, ‘Development of an eco-friendly Ultra-High Performance Concrete (UHPC) with efficient cement and mineral admixtures uses’, Cement and Concrete Composites, vol. 55, pp. 383-94. Yun, TS, Jeong, YJ, Han, T-S & Youm, K-S 2013, ‘Evaluation of thermal conductivity for thermally insulated concretes’, Energy and Buildings, vol. 61, pp. 125-32. Zhang, B & Poon, CS 2015, ‘Use of Furnace Bottom Ash for producing lightweight aggregate concrete with thermal insulation properties’, Journal of Cleaner Production, vol. 99, pp. 94-100. Zhang, J, Cheng, JC & Lo, IM 2014, ‘Life cycle carbon footprint measurement of Portland cement and ready mix concrete for a city with local scarcity of resources like Hong Kong’, The International Journal of Life Cycle Assessment, vol. 19, no. 4, pp. 745- 57. Zhang, X, Shen, L & Zhang, L 2013, ‘Life cycle assessment of the air emissions during building construction process: A case study in Hong Kong’, Renewable and Sustainable Energy Reviews, vol. 17, no. Supplement C, pp. 160-9. Zhou, D, Zhao, CY & Tian, Y 2012, ‘Review on thermal energy storage with phase change materials (PCMs) in building applications’, Applied Energy, vol. 92, pp. 593-605. Zhu, L, Hurt, R, Correia, D & Boehm, R 2009, ‘Detailed energy saving performance analyses on thermal mass walls demonstrated in a zero energy house’, Energy and Buildings, vol. 41, no. 3, pp. 303-10. Zuo, J, Pullen, S, Rameezdeen, R, Bennetts, H, Wang, Y, Mao, G, Zhou, Z, Du, H & Duan, H 2017, ‘Green building evaluation from a life-cycle perspective in Australia: A critical review’, Renewable and Sustainable Energy Reviews, vol. 70, pp. 358-68.

268

269

Appendix A: Mix properties of different batches of concrete

This appendix contains details of 90 different concrete design mixes that were used as

part of Chapter 4. Table A-1 shows the mix properties of different batches of concrete

collected from 8 published journal papers and databases.

Table A-1 Mix properties of different batches of concrete

Binder (kg/m3) Aggregates (kg/m3) Admixture (liter) Water (g/m3)

product

)

3

3

m

e / e

-

(kg/m

aggregate

Density

2

References

bottom ash

Mix Numbers

strength (MPa)

CO

Silane

Fly ash

al al aggregates

GGBFS

Conductivity (W/mk)

silica fumesilica

cenosphere

kg kg

Potable water

Superplasticiser

Portland cement

Reclaimed water

Natur

Polyethylene fibres

Manufactured Sand

Furnace

shrinkage reduction

Recycled aggregates

Fine natural river sand Lightweight

Water reducing admixture

Viscosity modifying agent

1 32 175 120 80 ...... 546 455 270 532 ... 1.3 ...... 141 2320.3 1.84 200.9

2 32 255 85 35 ...... 549 445 245 589 ... 1.3 ...... 68 70 2342.3 1.88 260.8

3 40 260 80 125 ...... 1067 ... 90 580 ... 1.6 ...... 76 70 2349.6 1.89 286.3 (CCAA 2015)(CCAA

4 39.4 325 ...... 828 ... 1041 ...... 3.6 ...... 195 ... 2392.6 1.97 319.0

5 46.8 450 ...... 755 ... 477 1.5 ...... 175 ... 1858.5 1.35 401.4

6 42.7 450 ...... 156 ...... 566 ... 477 2 ...... 175 ... 1826.0 1.30 403.0

7 40.9 450 ...... 312 ...... 377 ... 477 1.4 ...... 175 ... 1792.4 1.25 404.6

8 34.1 450 ...... 468 ...... 189 ... 477 1.9 ...... 175 ... 1760.9 1.20 406.2 (Zhang & Poon 2015) 9 31.6 450 ...... 624 ...... 477 1.8 ...... 175 ... 1727.8 1.16 407.7

10 32 330 ...... 1093 ... 778 ...... 160 ... 2361.0 1.91 331.6

11 32 254 ... 84.5 ...... 1090 ... 787 ...... 170 ... 2385.2 1.95 271.7

12 32 168 ...... 1089 ... 774 ...... 164 ... 2362.2 1.91 222.0

13 32 116 ...... 1095 ... 780 ...... 159 ... 2366.4 1.92 187.2

14 35 370 ...... 1035 ... 801 ...... 157 ... 2362.7 1.91 362.0

15 35 280 ... 93 ...... 1054 ... 797 ...... 158 ... 2382.0 1.95 291.6

16 35 188 ...... 1039 ... 784 ...... 158 ... 2357.0 1.90 239.6

17 35 196 ...... 1053 743.4 ...... 157 ... 2345.3 1.88 203.4

18 35 131 ...... 1061 ... 780 ...... 158 ... 2373.5 1.93 202.1

19 40 400 ...... 1080 ... 710 ...... 168 ... 2358.0 1.90 387.4 (Berndt2015)

20 40 300 ... 100 ...... 1095 ... 719 ...... 164 ... 2378.0 1.94 309.0

21 40 200 ...... 1082 ... 715 ...... 166 ... 2363.0 1.91 252.2

22 40 140 ...... 1075 ... 712 ...... 167 ... 2353.8 1.90 211.2

23 40 420 ...... 1030 ... 715 ...... 168 ... 2333.0 1.86 401.6

24 40 315 ... 105 ...... 1020 ... 718 ...... 172 ... 2330.2 1.85 317.9

25 40 210 ...... 1040 ... 740 ...... 164 ... 2363.8 1.91 260.3

26 40 151 ...... 1048 ... 720 ...... 168 ... 2365.7 1.92 221.5

27 67.6 377 ...... 33 ...... 946 ... 810 ...... 5.4 ...... 172 ... 2343.4 1.98 364.9

28 69.4 836 ...... 73 ... 348 ...... 4.9 10.5 ...... 302 ... 1574.4 0.40 697.1

29 56.9 732 ...... 64 ... 402 ...... 5.2 9.8 ...... 282 ... 1495.0 0.36 613.1

30 55.9 731 ...... 64 ... 268 ...... 5.9 8.9 0.2 ...... 287 ... 1365.0 0.35 608.7

31 48.8 607 ...... 53 ... 442 ...... 5.6 9.8 ...... 282 ... 1399.4 0.33 511.3

32 33 499 ...... 43 ... 317 ...... 6.6 9.1 0.2 ...... 290 ... 1164.9 0.28 418.6

33 66.1 846 ...... 74 ... 352 ...... 5.2 ...... 305 ... 1582.2 0.39 705.6

34 69.4 836 ...... 73 ... 348 ...... 4.9 10.5 ...... 302 ... 1574.4 0.40 697.1 (Wu et al. 2015) 35 49.8 607 ...... 53 ... 442 ...... 5.6 9.8 ...... 282 ... 1399.4 0.33 511.3

36 40.9 606 ...... 53 ... 442 ...... 6.7 9.7 0.18 ...... 282 ... 1399.6 0.31 510.5

37 66.1 846 ...... 74 ... 352 ...... 5.2 ...... 305 ... 1582.2 0.39 705.6

38 66.5 775 ...... 67 ... 350 ...... 3.6 ...... 5.3 4.2 304 ... 1509.1 0.43 646.5

39 54.4 832 ...... 72 ... 346 ...... 4.3 ...... 5.7 ... 301 ... 1561.0 0.39 693.3 270

40 63.1 1355 ...... 118 ...... 1.3 14.9 ...... 499 ... 1988.2 0.84 1114.4

41 51.3 1179 ...... 103 ...... 16.9 0.91 ...... 561 ... 1860.8 0.80 969.9

42 45 345 ...... 1826 ...... 195 ... 2366.0 0.92 366.7

43 42 190 155 ...... 1826 ...... 195 ... 2366.0 0.88 261.8

44 41 295 ... 60 ...... 1802 ...... 185 ... 2342.0 0.82 326.2

45 43 275 ...... 70 ...... 1826 ...... 195 ... 2366.0 0.84 311.2

46 39 345 ...... 1447 ...... 204 ... 2361.0 0.72 350.8

47 42 190 155 ...... 1447 ...... 204 ... 2361.0 0.67 245.8

48 41 295 ... 60 ...... 1438 ...... 189 ... 2342.0 0.61 311.0

49 38 275 ...... 70 ...... 1447 ...... 204 ... 2361.0 0.65 295.3 2362.0 0.99 530.6 50 (Damdelen2015) et al. 55 557 ...... 1610 ...... 195 ...

51 51 251 306 ...... 1610 ...... 195 ... 2362.0 0.93 323.5

52 53 478 ... 120 ...... 1583 ...... 180 ... 2361.0 0.89 467.9

53 54 501 ...... 111 ...... 1610 ...... 195 ... 2417.0 0.90 487.7

54 54 583 ...... 1234 ...... 204 ... 2358.0 0.77 536.0

55 49 321 262 ...... 1234 337 ...... 204 ... 2358.0 0.73 358.7

56 48 502 ... 126 ...... 1212 331 ...... 190 ... 2361.0 0.70 472.0

57 52 466 ...... 117 ...... 1234 337 ...... 204 ... 2358.0 0.75 443.3

58 43.9 288 ... 32 ...... 1756 ...... 175 ... 2251.0 1.93 317.6

59 NA 288 ... 32 ...... 6 1730 ...... 175 ... 2231.0 1.67 316.6

60 35.3 288 ... 32 ...... 12 1602 ...... 175 ... 2109.0 1.71 310.9

61 32.1 288 ... 32 ...... 24 1364 ...... 175 ... 1883.0 1.56 300.3

62 24.6 288 ... 32 ...... 37 1097 ...... 175 ... 1629.0 1.44 288.4

63 37.5 288 ... 32 ...... 826 ...... 552 ...... 175 ... 1873.0 1.32 300.3

64 36.2 288 ... 32 ...... 12 826 ...... 409 ...... 175 ... 1742.0 1.28 294.0

65 (Yun et al. 2013) 28.1 288 ... 32 ...... 23 826 ...... 276 ...... 175 ... 1620.0 1.25 288.2

66 23 288 ... 32 ...... 35 826 ...... 127 ...... 175 ... 1483.0 1.18 281.7

67 37.7 288 ... 32 ...... 834 ...... 583 ...... 175 ... 1912.0 1.33 302.1

68 33 288 ... 32 ...... 23 826 ...... 289 ...... 175 ... 1633.0 1.30 288.8

69 36 288 ... 32 ...... 834 ...... 510 ...... 175 ... 1839.0 1.29 298.7

70 36.6 300 ...... 1902 ...... 179 ... 2381.0 1.95 333.3

71 41.8 353 ...... 1854 ...... 182 ... 2389.0 1.96 374.6

72 48.6 402 ...... 1798 ...... 188 ... 2388.0 1.96 412.2

73 33.6 300 ...... 611 ...... 179 40 1130.0 1.73 274.0

74 41.1 351 ...... 596 ...... 183 39 1169.0 1.76 315.2 (Marinković et al. 2010)

75 48.1 402 ...... 579 ...... 189 29 1199.0 1.75 356.2

76 43.7 354 ...... 1164 ...... 185 ... 1703.0 1.98 343.7

77 41.5 384 ...... 1165 ...... 201 ... 1750.0 1.90 368.4

78 44.2 354 ...... 555 555 ...... 185 20 1669.0 1.87 318.0

79 (Tošić et al. 2015) 42.5 365 ...... 1071 ...... 180 38 1654.0 1.82 303.6

80 32 324 ...... 1929 ...... 184 ... 2437.0 2.05 354.2

81 32 273 ... 510 ...... 1931 ...... 181 ... 2895.0 2.88 326.3

82 32 258 ... 660 ...... 1921 ...... 183 ... 3022.1 3.11 317.6

83 32 243 ... 81 ...... 1923 ...... 180 ... 2427.0 2.03 289.7

84 32 227 ... 96 ...... 1924 ...... 185 ... 2432.0 2.04 277.0

85 32 192 ... 128 ...... 1910 ...... 177 ... 2407.0 1.99 248.6

86 32 240 80 ...... 1910 ...... 240 ... 2470.0 2.11 295.9

87 32 220 100 ...... 1910 ...... 220 ... 2450.0 2.07 282.4 (O'Moore & 2009)O'Brien 88 32 210 110 ...... 1910 ...... 210 ... 2440.0 2.05 275.6

89 32 190 130 ...... 1910 ...... 190 ... 2420.0 2.02 262.1

90 32 180 100 ...... 1910 ...... 180 ... 2370.0 1.93 249.6

271

Appendix B: Detailed structural design

This appendix summarises the structural design of the benchmark buildings. The information in this section has been categorised into a summary of the design for Flat slab and Waffle slab (section B-1), Manual hand calculation and versifications for Flat slab (section B-2), and typical concrete column design (CAD and Excel spread sheet output) in section B-3.

Section B-1: Summary of a detailed structural design A summary of the Flat slab detailed structural design is provided in Table B-1.

Table B-1 Flat slab detailed structural design

Structure details- Flat slab

(Cross section) (mm) section) (Cross

Quantity of Concrete of Concrete Quantity

Number of Columns Number of

Total Steel ( Total Steel

Steel arrangement Steel arrangement

Total Concrete

Grade of concrete Grade of

Quantity of steel Quantity Size of

% Steel

(tonne)

element (m

Structure elements (Cross 3

)

tonne)

section)

(m

(mm) 3

)

Level 1 Interior 500×500 10 N 32 24 20 18 3% to 3 perimeter 350×350 8 N 28 24 10 11 4%

Level 4 Interior 400×400 10 N 28 24 13 13 4% to 6 perimeter 325×325 8 N 24 24 8 8 3%

Column Level 7 Interior 375×375 10 N 20 24 11 7 2% to 9 perimeter 300×300 N40 8 N 20 24 7 93 6 3% 83

Level Interior 375×375 8 N 24 24 8 8 3% 10 to 12 perimeter 300×300 8 N 20 24 5 6 3% Level Interior 275×275 6 N 16 24 6 2 3% 13 to 15 perimeter 250×250 8 N 16 24 5 4 3% Column strip & Mid span: Top- N12@150 mm; Slab Suspended floor with 200 mm Bot- N12@100 0.56 N32 2469 3,000 654 654

drop panel (depth) mm (Same for % both directions)+ Drop panel (N12@ 300 mm) N12@300 mm 200 mm Wall N40 both sides (Top & --- 31 9 4% 9 (thickness) Bottom) 15 mm N12@200 mm Staircase N20 --- 250 7 1% 7 (thickness) both directions

272

Table B-2 provides a summary of Waffle slab detailed structural design.

Table B-2 Waffle slab detailed structural design

Structure details- Waffle Slab

Quantity of Concrete of Concrete Quantity SteelTotal (tonne)

Number of Columns Number of

Grade of concrete Grade

Quantity of steelof Quantity Steel Total Concrete Size of

arrange Steel % element (tonne) ment (m

3

Structure elements (Cross ) (Cross

section) (m section)

3

(mm) )

(mm)

Level Interior 500×500 10 N 32 24 20 18 3% 1 to 3 perimeter 350×350 8 N 28 24 10 11 4% Level Interior 400×400 10 N 28 24 13 13 4% 4 to 6 perimeter 325×325 8 N 24 24 8 8 3%

Column Level Interior 375×375 10 N 20 24 11 7 2% 7 to 9 perimeter 300×300 N40 8 N 20 24 7 93 6 3% 83 Level Interior 375×375 8 N 24 24 8 8 3% 10 to 12 perimeter 300×300 8 N 20 24 5 6 3% Level Interior 275×275 6 N 16 24 6 2 3% 13 to 15 perimeter 250×250 8 N 16 24 5 4 3% 250 mm Column strip & Suspended floor (50 mm Mid span: Top- 704 thickness) N16@ 140 mm; Bot- 3 N20 for

Slab 3500×324 Drop panel each Ribs (Same 298 0.21 mm N32 2,002 580 580

for both % directions); 200×300 Spacing of Ribs Sterm 1,000 mm every 900 mm each direction N12@300 mm 200 mm Staircase N40 both sides (Top ----- 31 9 4% 9 (thickness) & Bottom) 15 mm N12@200 mm Staircase N20 ----- 250 7 1% 7 (thickness) both directions

273

Section B-2: Manual calculation and verification for flat of slab

This section provides details of the manual calculation verification for Flat slabs. All the design calculations for the reinforced slabs comply with the Australian standards, including: Concrete structure (AS3600 2009) and Structural Actions (AS/NZ1170.1

2002).

Applied load:

For the suspended office floor:

Self-weigh (SW): 25 kPa (by considering concrete as Normal weight with steel reinforcement)

Additional Dead load (ADL): 0.5 kPa (for permanent services and fixtures)

Live load (LL): 3.0 kPa (AS/NZ1170.1 (2002), Clause 3.4.1)

The strength of concrete was f’c = 32 MPa for the slabs, and the maximum size bars were 16 mm diameter.

Reinforced Flat slab

Minimum slab thickness (depth)

Based on the Australian standard (AS3600 2009), the clear cover required to meet fire resistance and durability was 35mm, that also includes a construction tolerance

(5mm). The thickness of the preliminary slab was selected by using the Cement and

Concrete Association of Australia guidelines (CCAA 2003). Although this guideline is based on an older version of AS3600 (2001), it has a good preliminary point to determine the required depth of the slab.

Span= 5.27m, LL=3.0 kPa based on the graph on page 15 (CCAA 2003) that gives the first trial D as 170 mm.

The depth of slab can be controlled by the deflection limit:

274

1 ∆ 3 ( ) 1000 퐸푐 퐿푒푓 퐿푒푓 ≤ 푘3푘4 [ ] (AS 3600, Clause 9.3.4.1) 푑 퐹푑,푒푓

Where:

D= 170mm

SW= 25*0.17=4.25 kPa

DL= 4.25+0.5=4.75 kPa

LL= 3 kPa

퐿푒푓 = 푚𝑖푛 [퐿푛 + 퐷; 퐿] = [(5270 - 500) +170; 5270] = 4940mm

(AS 3600, Clause 9.3.4.1)

푘3 = 0.95 푓표푟 푡푤표 푤푎푦 푓푙푎푡푠푙푎푏 푤𝑖푡ℎ표푢푡 푑푟표푝 푝푎푛푒푙 (AS 3600, Clause 9.3.4.1)

푘4 = 0.75 − 퐸푥푡푒푟𝑖표푟 푠푝푎푛 (AS 3600, Clause 9.3.4.1)

∆ 1 = - Total deflection (AS 3600, Table 2.3.2) 퐿푒푓 250

∆ 1 = – Incremental deflection (AS 3600, Table 2.3.2) 퐿푒푓 500

퐸퐶 = 30100 MPa

휓푠 = 0.7 (AS 1170.0:20, Table 4.1)

휓푙 =0.4 (AS 1170.0:20, Table 4.1)

휓퐶 =0.4 (AS 1170.0:20, Table 4.1)

푘퐶푠 = [2.0 - 1.2 (Asc/Ast)] ≥ 0.8 (AS 1170.0:20, Table 4.1)

= 2 (assumed Asc=0 for most critical case)

퐹푑,푒푓 = (1 + 푘퐶푠)푔 + (휓퐶 + 푘퐶푠 × 휓푙)q- Total deflection

(AS 3600, Clause 9.3.4.1)

275

= (1 + 2) × 4.75 + (0.7 + 2 × 0.4) × 3

= 18.75 kPa

퐹푑,푒푓 = (푘퐶푠 × 푔) + (휓퐶 + 푘퐶푠 × 휓푙)q- Incremental deflection

= (2 × 4.75) + (0.7 + 2 × 0.4) × 3

= 14 kPa

Total deflection:

1 ∆ 3 ( ) 1000 퐸푐 퐿푒푓 퐿푒푓 ≤ 푘3푘4 [ ] 푑 퐹푑,푒푓

1 1 3 퐿 ( ) × 1000 × 30100 푒푓 ≤ 0.95 × 1.75 × [ 250 ] 푑 18.75

4940 Then d = = 159.87 푚푚 30.90

Incremental deflection:

1 ∆ 3 ( ) 1000 퐸푐 퐿푒푓 퐿푒푓 ≤ 푘3푘4 [ ] 푑 퐹푑,푒푓

1 1 3 ( ) × 1000 × 30100 ≤ 0.95 × 1.75 × [ 500 ] 14

Then

4940 푑 = = 182.75 푚푚 27.03

Therefore, incremental deflection governs the minimum slab thickness. Since there will be steel reinforcement in both directions (X and Y), the minimum slab thickness is: 276

D = d+ cover+ ½ bar thickness (X axis) + bar thickness (Y axis).

D = 182.75 + 35 + 12 + 12= 235.75 mm 2

Revision of the design is necessary and several alternatives were considered:

1. Use a drop panel, then K3 = 1.05 (AS 3600, Clause 9.3.4.1)

2. Use Asc in the design process, then kcs= 1

Recheck the calculations to satisfy the ratio of span over thickness by considering

new assumptions

D = 200 mm

Lef = [(5270 - 500) + 200; 5270] = 4970 mm

퐹푑,푒푓 = (푘퐶푠 × 푔) + (휓퐶 + 푘퐶푠 × 휓푙)q- Incremental deflection

= (1× 4.75) + (0.7 + 2 × 0.4) × 3

= 8.75 kPa

1 ∆ 3 ( ) 1000 퐸푐 퐿푒푓 퐿푒푓 ≤ 푘3푘4 [ ] 푑 퐹푑,푒푓

1 1 3 ( ) × 1000 × 30100 ≤ 1.05 × 1.75 × [ 500 ] 8.3

4970 Then d = = 144.89 푚푚 34.30

12 D = 144.89 + 35 + + 12= 197.89 mm ≈ 200 푚푚 ∴ 푂푘 2

12 dx = 200 – 35 – = 159 mm 2

12 dy = 200 – 35 – – 12 = 147 mm 2

277

Correct the applied load:

D= 200mm

SW = 25×0.2 = 5 kPa

DL= 5 + 0.5= 5.5 kPa

LL= 3 kPa

Determine the bending moment:

Since the slab is symmetrical, only one direction was calculated to represent the process of design and analysis. Moreover, the bending moments and strip dimensions are equal in both axes (X and Y), so the bending moments were determined using the

Simplified method.

Design bending moment and reinforcement in the long span direction:

Design strips:

FigureB-1 shows the dimensions and layout of the design strips.

278

Figure B-1 the dimensions and layout of the design strips 279

Table B-3 provides the design moment factors for the Flat slab supported by columns.

Table B-3 design moment factors for the Flat slab

Location of Exterior moment Mid span moment Interior moment moment (negative moment) (positive moment) (negative moment) End span 0.25 0.5 0.75 Interior span ------0.35 0.65 Reference: AS3600 (2009), Table 6.9.5.3 , Table 6.1.4.3

TableB-4 provides the distribution of bending moments to the column strip.

Table B-4 distribution of bending moments to the column strip

Distribution of bending moments to the column strips Span number Left side/end Mid-span Right side/end 1 1.00 0.6 0.75 2 0.75 0.6 0.75 3 0.75 0.6 0.75 4 0.75 0.6 0.75 5 0.75 0.6 0.75 6 0.75 0.6 1.00 Edge Design Strip

Fd=1.2 DL+1.5 LL= 1.2×(5.5) +1.5×(3) = 11.1 kPa

Design length LT = 5.27 m

Lo= L- 0.7asupL- 0.7asup.R (AS 3600, Clause 9.3.4.1) asup=(Column size/2)+(thickness of drop panel)= (550/2)+(260-200) = 335 mm

Lo = 5270 – 0.7×335= 5035 mm ≈ 5.04 푚

Mo

Fd × L × L2 = T o (AS 3600, Clause 6.10.4.2) 8

11.1 × 5.27 × 5.042 Mo = = 185.73 kN. m 8

280

Design Moments:

TableB-3 summarises the design moments for the end span- Edge design strip

Table B-5 the design moments for the end span- Edge design strip

Exterior moment Mid span moment Interior moment (negative moment): (positive moment): (negative moment): 0.25Mo 0.5Mo 0.75Mo -46.43 kN.m 92.86 kN.m -139.29 kN.m column column Middle Middle column Middle strips strips strips strips (0) strips (0.6) strips (0.4) (1.00) (0.75) (0.25) -46.43 55.71 37.14 -104.46 -34.82 0 kN.m kN.m kN.m kN.m kN.m kN.m TableB-6 summarises the design moments for the interior span- Edge design strip

Table B-6 the design moments for the interior span- Edge design strip

Exterior moment (negative moment): Mid span moment (positive moment): 0.65Mo 0.35Mo

-120.72 kN.m 65 kN.m

column strips Middle strips column strips Middle strips (0.4) (0.75) (0.25) (0.6)

-90.54 kN.m -30.18 kN.m 39 kN.m 26 kN.m Interior Design Strip

Fd=1.2 DL+1.5 LL= 1.2×(5.5) +1.5×(3) = 11.1 kPa

Design length LT = 5.27 m

Lo= L- 0.7asupL- 0.7asup.R (AS 3600, Clause 9.3.4.1) asup=(Column size/2)+(thickness of drop panel)= (550/2)+2×(260-200) = 395 mm

Lo = 5270 – 0.7×395= 4993.5 mm ≈ 5 푚

Fd × L × L2 Mo = T o (AS 3600, Clause 6.10.4.2) 8

281

11.1 × 5.27×52 Mo = = 182.80 kN. m 8

Design Moments:

Table B-7 summarises the design moments for the end span- Edge design strip

Table B-7 the design moments for the end span- Edge design strip

Exterior moment Interior moment Mid span moment (negative moment): (negative moment): (positive moment): 0.5Mo 0.25Mo 0.75Mo -45.7 kN.m 91.4 kN.m -137.1 kN.m column Middle column Middle column Middle strips strips (0) strips (0.6) strips (0.4) strips strips (1.00) (0.75) (0.25) -45.7 kN.m 0 kN.m 54.84 kN.m 36.56 kN.m -102.82 -34.27 kN.m kN.m

Table B-8 summarises the design moments for the interior span- Edge design strip

Table B-8 summarises the design moments for the interior span- Edge design strip

Exterior moment (negative moment): Mid span moment (positive moment): 0.65Mo 0.35Mo -118.82 kN.m 63.98 kN.m

column strips Middle strips column strips (0.6) Middle strips (0.4) (0.75) (0.25)

-89.11 kN.m -29.70 kN.m 38.38 kN.m 25.59 kN.m

FigureB-2 provides the bending moments measured in kNm

282

Figure B-2 Bending moment plan for Flat slab, f'c= 32MPa 283

Reinforcement

In the previous section, the negative bending moments at the supports and positive bending moments at the mid-span were determined and then distributed into the column strip and middle strip.

The following section provides a sample calculation for the reinforcement required to resist a bending moment in the Y direction. The remaining calculation to determine the steel reinforcement in both directions was calculated by using the Excel spreadsheet and SAFE software.

Data:

For this design, N16 bars were used as the main distribution reinforcement.

2 Fsy = 500 MPa, Ast= 200 mm

At mid-span: dx= 200-35-16/2 = 157 mm

At mid-span= +M*= 55.71 kN.m

𝜌푚푖푛

2 ′ 퐷 푓푐푡.푓 = 0.24 ( 푠) ( ) (AS 3600, Clause 9.1) 푑 푓푠푦

= 0.24 (200/157)2× (3.4/500) = 00.00264

Minimum reinforcement required for shrinkage and temperature effect:

𝜌 ≥ 0.75 × 0.0035(bDs)/(bd) (AS 3600, Clause 9.4.3)

≥ 0.75 × 0.0035(200)/(157) = 0.0033

Therefore the shrinkage and temperature reinforcement governs:

2 Ast= 0.75× 0.0035× bDs= 0.75× 0.0035×1000×200 = 525 mm /m

For 2.635 m: 525 ×2.635= 1383.38 mm2

284

Check the capacity of the section with this reinforcement

∅푀푢 = 0.8 × 푓푠푦 × (0.925푑)

= (0.8 ×1383.38×500×0.925×157) = 80.36 kN.m

Using N16 has too much space between the bars so change them to N12.

Hence, N12 @ 150 mm cts.

At First Interior Column:

Dx= 200+60 (drop panel)= 260 dx= 260-35-(16/2)= 217 mm

At mid-span= +M*= -102.82 kN.m

2 ′ 퐷푠 푓푐푡.푓 𝜌푚푖푛 = 0.24 ( ) ( ) (AS 3600, Clause 9.1) 푑 푓푠푦

= 0.24 (260/217)2× (3.4/500) = 00.00234

Minimum reinforcement required for the shrinkage and temperature effects:

𝜌 ≥ 0.75 × 0.0035(bDs)/(bd) (AS 3600, Clause 9.4.3)

≥ 0.75 × 0.0035(260)/(217) = 0.0031

Therefore shrinkage and temperature reinforcement governs:

2 Ast= 0.75× 0.0035× bDs= 0.75× 0.0035×1000×250 = 656.25 mm /m

For 2.635 m: 656.25 ×2.635= 1726.21 mm2

Check the capacity of the section with this reinforcement

∅푀푢 = 0.8 × 푓푠푦 × (0.925푑)

= (0.8 ×1383.38×500×0.925×217) = 111.07 kN.m

Using N16 has too much space between the bars so change them to N12.

Hence, N12 @ 100 mm cts.

285

Punching shear design

The punching shear is checked on a critical section at a distance of dom/2 from the face of the support (AS 3600 Clause 9.2.1.1). For rectangular columns and concentrated loads, the critical area is taken as a rectangular area with the sides parallel to the sides of the columns or the point loads (AS 9.2.1.3). The following Figure B-3 represents the punching perimeters considered by AS 3600 (2009).

Figure B-3 the punching perimeters considered by AS 3600 (2009).

The shear capacity fcv was calculated based on the lower of two expressions from AS

3600, Clause 9.2.3, as shown below:

2 ′ 0.17(1 + )√푓푐 훽ℎ 푓푐푣 = 푚𝑖푛 { (AS 3600, Clause 9.2.3) ′ 0.34 √푓푐

βh is the ratio of the longest dimension to the shortest dimension of the critical section=1

2 0.17 (1 + ) √32 = 2.88 1 푓푐푣 = 푚𝑖푛 { = 1.92 0.34 √32 = 1.92

From Table 4.4 (Warner et al. 2010), the out of balance moment between two adjacent faces of the first interior column is:

∗ 푀푉 = (푐표푒푓푓𝑖푐𝑖푒푛푡 × 푀표) end span - (푐표푒푓푓𝑖푐𝑖푒푛푡 × 푀표) internal span

(Warner et al. (2010), Page 253)

286

∗ 푀푉 = (0.75 × 185.73) - (0.65 × 182.80) = 22.85 kN.m

This is equal to the out of balance moments on each side of the interior column in x direction, which is shown in Figure B-2.

∗ 푀푉 = (102.82 + 34.27 + 34.27) - (89.11 + 29.70 + 29.70) = 22.85 kN.m

The minimum bending moment to be transmitted is:

1.5퐿퐿 푀∗ = 0.06 ((1.2퐷퐿 + ) 퐿 퐿2 − 1.2퐷퐿 퐿 퐿2 ) 푉 2 푡 0 푡 표

(Warner et al. (2010), Equation 4.15)

1.5 × 3 푀∗ = 0.06 ((1.2 × 5.5 + ) 5.27 × 5.042 − 1.2 × 5.5 × 5.27 × 52) 푉 2

∗ 푀푉 = 0.06(1184.71 − 869.55)= 18.91 kN.m

The corresponding load to be transferred by the punching shear

Tributary Area = 5.27 × 5.27 = 27.77 m2

Punching force 푉∗ = 11.1 × 27.77 = 308.247 kN

푑푥 + 푑푦 푑 = 표푚 2 dy: Thickness of slab with drop panel=200+60 = 260 mm dx: dy- bar thickness (N12)= 260-12= 248 mm

260 + 248 푑 = = 254 mm 표푚 2

푉푢표 = 푢푑표푚푓푐푣 (Warner et al. (2010), Equation 4.16)

푓표푟 𝑖푛푡푒푟𝑖표푟 푐표푙푢푚푛

푑 푑표푚 푦 푎푥푖푠 푈 = 푐표푙푢푚푛 푡ℎ𝑖푐푘푛푒푠푠 + ( 표푚 푥 푎푥푖푠) × 2 + ( ) × 2 2 2

푢 = 550 + 254 × 2 =1058 mm

−3 푉푢표 = 1058 × 254 × 1.92 × 10 = 515.96 푘푁 287

∅푉푢표 ∅푉푢 = ∗ 푀푉 푎 (1 + ∗ ) 8푉 푑표푚 푢

((Warner et al. 2010), Equation 4.18)

0.7 × 515.96 ∅푉 = = 361.15 푘푁 푢 22.85 (1 + 550 + 254 ) (8 × 308.247 × 254)/1058)

∗ ∅푉푢 > 푉

361.15 푘푁 > 308.247 kN ∴ Ok

Hence, failure will not occur at the front face and the drop panel is enough to avoid the need for fitments.

Calculate Vu.min:

1.2 푉 푉 = 푢표 (AS 3600, Equation 9.2.4(2)) 푢.푚푖푛 푢 푀∗ (1 + 푉 ) (2 푉∗푎2)

1.2 × 515.96 푉 = = 619.114 푘푁 푢.푚푖푛 1058 × 22.85 (1 + ) (2 × 308.247 × (550 + 254)2)

∅푉푢.푚푖푛= 0.7× 619.11= 433.38 kN

361.15 푘푁 > 308.247 kN ∴ Ok

Hence minimum reinforcement is sufficient.

Design of closed ties:

푑 12 푦 = 푎 − min 푐표푣푒푟 − 푏 = (550 + 254) − 35 − ( ) = 763 푚푚 푙 2 2

퐴 0.2 푦 푉∗ 2 0.2 × 763 308.247 2 mm2 ( 푠푤) = 푙 ( ) = ( ) = 0.154 푆 푓푠푦.푓 ∅푉푢.푚푖푛 500 433.38 mm

퐴 mm2 ( 푠푤) = 0.154 ; if we use N12 As = 110 mm2 푆 mm 288

S= 714 mm

S≤ max (300, D) (AS 3600, Clause 9.2.6)

S≤ max (300, 260)

S= 300 mm

Then use N12 @ 300 mm cts for the torsional strips.

퐷 ∗ Also, ∅Vu max = ∅ 3 푉 √ ≥ 푉 needs to be checked. (AS 3600, Clause 9.2.4(5)) 푢.푚푖푛 푎

260 ∗ ∅Vu max = 3 × 433.38√ = 739.34 ≥ 푉 =308.247 ∴ 푂푘 (550+254)

Check the deflection of the slab (using simplified method): (AS 3600, Clause 9.3.3)

∆풕풐풕풂풍= ∆풔 + 풌풄풔∆풔.풔풖풔

퐿2 ∆풔= ( )(푀퐿 + 10푀푀 + 푀푅) 96 퐸푐퐼푒푓.푎푣푒

Deflection for short-term service load

Fs = FDead + ΨsFLive

Fs.l = FDead + ΨLFLive

FDead = 5.5 푘푃푎

FLive = 3.0 푘푃푎

Ψs = 0.7 (푠ℎ표푟푡 푡푒푟푚 푙𝑖푣푒 푙표푎푑 푓푎푐푡표푟) (AS 1170.2, Table 4.1)

ΨL = 1 (퐿표푛푔 푡푒푟푚 푙𝑖푣푒 푙표푎푑 푓푎푐푡표푟) (AS 1170.2, Table 4.1)

Fs = 5.5 + 0.7 × 3.0 = 7.6 kPa (short term)

Fs.l = 5.5 + 1.0 × 3.0 = 8.5 kPa (long term)

Moment at the exterior column:

퐹 퐿 퐿2 7.6 × 5.27 × 52 −푀 = 푠 푡 표 = = 62.58 푘푁. 푚 푠 16 16 289

Moment at the first interior column:

퐹 퐿 퐿2 7.6 × 5.27 × 52 +푀 = 푠 푡 표 = = 100.3 푘푁. 푚 푠 10 10

Moment at the end of mid-span:

퐹 퐿 퐿2 7.6 × 5.27 × 52 −푀 = 푠 푡 표 = = 71.52 푘푁. 푚 푠 14 10

Moment in the column strip (2.63m) of the design strip and based on Table 4.5

(Warner et al. 2010)

Moment at the exterior column: −푀푠 = 0.75 × 62.58 = −46.93 푘푁. 푚

Moment at the first interior column: +푀푠 = 0.6 × 100.3 = 60.18 푘푁. 푚

Moment at the end of mid-span: −푀푠 = 0.75 × 71.52 = −53.64 푘푁. 푚

Effective second moment of the area of the column strip:

Calculation for the mid-span:

푏ℎ3 2635 × 2003 퐼 = = = 1756 × 106푚푚4 푔 12 12

Calculation for the interior column:

푏ℎ3 2635 × 2603 퐼 = = = 3.85 × 109푚푚4 푔 12 12

The section crack moment Mcr at the mid-span:

′ 푀푐푟 = 푍(푓푐푡.푓 − 𝜎푐푠) (Warner et al. (2010), Equation 3.41)

2.5푝푤−0.8푝푐푤 ∗ 𝜎푐푠 = 퐸푠휀푐푠 (Warner et al. (2010), Equation 3.11) 1+50푝푤

퐴푠푡 푝푤 = (Warner et al. (2010), Equation 3.12) 푑푏푤

퐴푠푐 푝푐푤 = (Warner et al. (2010), Equation 3.13) 푑푏푤

∗ 휀푐푠: 푑푒푠𝑖푔푛 푠ℎ푟𝑖푛푘푎푔푒 푠푡푟푎𝑖푛 (AS 1170.2, Table 3.1.7.2)

290

2.5 × 0.0033 − 0.8 × 0 𝜎 = × 200 × 103 × 700 × 10−6 = 1.0푀푃푎 푐푠 1 + 50 × 0.033

Mid-span

′ 푀푐푟 = 푍(푓푐푡.푓 − 𝜎푐푠)

퐼 Z= 푔 =17.56 × 106푚푚3 퐷/2

6 푀푐푟 = 17.56 × 10 × (0.6√32 − 0.49)= 51.01 kN.m

Interior column

′ 푀푐푟 = 푍(푓푐푡.푓 − 𝜎푐푠)

퐼 Z= 푔 =29.68 × 106푚푚3 퐷/2

6 푀푐푟 = 29.68 × 10 × (0.6√32 − 0.49)= 82.21 kN.m

For interior support since Ms ≤ 푀푐푟 (53.64 ≤ 82.21)

9 9 4 퐼푒푓 = 0.6 퐼푔= 0.6 × 3.85 × 10 = 2.31× 10 푚푚

For the exterior column since Ms ≤ 푀푐푟 (62.58 ≤ 82.21)

9 9 4 퐼푒푓 = 0.6 퐼푔= 0.6 × 3.85 × 10 = 2.31× 10 푚푚

For the mid-span since Ms ≤ 푀푐푟 (60.18 ≥ 51.01)

The second moment of area for the fully cracked section Icr at the mid span is:

3 퐸푠 200×10 2 푛 = = = 6.6 , nAst= 6.6 × 525= 3465 mm 퐸푐 30100

1000×dn×dn/2= nAst× (d-dn)

2 500 dn = 3465 × (147- dn)

2 3 dn +6.912 dn – 1.01× 10 =0 dn= 28.51 mm

푏ℎ3 퐼 = + 푛퐴 (푑 − 푑 )2 푐푟 3 푠푡 푥 푛

291

2635 × 28.513 퐼 = + 3456(147 − 28.51)2 = 68.87 × 106 푚푚4 푐푟 3

The effective second moments of area is:

2 퐼푐푟 푀푐푟 퐼푒푓 = 퐼푐푟/(1 − (1 − ) ( ) ) (Warner et al. (2010), Equation 3.40) 퐼푔 푀

68.87 × 106 51.01 2 퐼 = 68.87 × 106/(1 − (1 − ) ( ) ) 푒푓 1756 × 106 60.18

6 4 퐼푒푓 = 222.36 × 10 푚푚

Thus, for the end span, the weighted average effective second moment of area is:

9 4 퐼푒푓.푎푣푒 = 1.266 × 10 푚푚

Long term deflection is given by∆푡표푡푎푙= ∆푠 + 푘푐푠∆푠.푠푢푠, where ∆푠and ∆푠.푠푢푠 are short term deflections due to the short span proportion of (0.79) and the loading combination of Fs 푎푛푑 Fs.l.

퐿2 ∆풔= ( )(푀퐿 + 10푀푀 + 푀푅) 96 퐸푐퐼푒푓.푎푣푒

52702 ∆ = ( )(−46.93 + 10 × 60.18 − 53.64) × 106 풔 96 × 30100 × 1.266 × 109

∆풔= 3.805 mm

Fs 7.6 ∆푠.푠푢푠= ∆풔= × 3.805 = 3.40푚푚 Fs.l 8.5

∆푖푛푐푟= 푘푐푠∆푠.푠푢푠= 2 × 3.40 = 6.80 푚푚

∆푡표푡푎푙= ∆푠 + 푘푐푠∆푠.푠푢푠= 3.805 + 6.80 = 10.60 푚푚

Limiting deflection

퐿푒푓 5720 ∆ = = = 11.44 푚푚 푖푛푐푟 500 500

∆푖푛푐푟= 푘푐푠∆푠.푠푢푠= 6.80 푚푚 < 11.44 푚푚 ∴ 푂푘

292

퐿 5720 ∆ = 푒푓 = = 22.88 푚푚 푡표푡푎푙 250 250

∆푡표푡푎푙= ∆푠 + 푘푐푠∆푠.푠푢푠= 10.60 푚푚 < 22.88 푚푚 ∴ 푂푘

Therefore, a 200 mm deep slab is adopted with N12 @ 150 mm centres (cts) at mid span and N12 @ 100 mm at column strips.

293

Section B-3: Concrete column design This part provides an overview of the basic assumptions, design preconditions, and some of the design parameters that affect the design of concrete columns; CAD software (Etabs) was used to calculate the required longitudinal steel reinforcement.

The following two steps were considered when designing the reinforced concrete columns:

1- Axial force-biaxial moment interaction surfaces for all the different types of

concrete sections in the model.

2- Check the Axial and Bending capacity of each column for the factored

loading combinations.

The following Tables B-9 to B-16 show the sample structural analysis for a selected column in level 1.

Table B-9 Section Properties Cover (Torsion) b (mm) h (mm) dc (mm) (mm) 350 350 64 30

Table B-10 Column Element Details (Summary) Level Element Section ID Combo ID Station Loc Length (mm) LLRF 350X350m (1.2DDL+ Story1 C2 3300 3300 0.407 m 1.5 LL)

Table B-11 Material Properties Lt.Wt Factor Ec (MPa) f'c (MPa) fy (MPa) fys (MPa) (Unitless) 32800 40 1 413.69 413.69

Table B-12 Design Code Parameters ΦT ΦC ΦVns ΦVs ΦVjoint 0.8 0.6 0.7 0.6 0.7

294

Table B-13Axial Force and Biaxial Moment Design For N* , M*2 , M*3 Minimum Minimum Design N* Design M*2 Design M*3 Area Rebar % M2 M3 kN kN-m kN-m mm² % kN-m kN-m 2657.1922 -46.5009 31.167 46.5009 46.5009 3016 2.46

Table B-14 Axial Force and Biaxial Moment Factors km Factor δb Factor δs Factor K Factor Length

Unitless Unitless Unitless Unitless mm Major 0.4 1 1 1 3300 Bend(M3) Minor 0.43439 1 1 1 3300 Bend(M2)

Table B-15 Shear Design for V*2 , V*3 Shear V* Shear ΦVuc

kN kN Major, V*2 14.228 218.077 Minor, 0.8716 218.077 V*3

Table B-16 Column axial load Interaction diagram: and bending moment P M3 Point (kN) (kNm) 1 3146.9902 0

2 2594.0073 65.5322

3 2286.5342 101.9699

4 1875.9332 130.0551

5 1441.5865 148.2139

6 856.3381 157.2899

7 626.0161 158.6059

8 255.2145 149.9586

9 -97.6088 118.7691

10 -749.8022 40.197

Figure B-3 Structural design: Interaction 11 -997.985 0 diagram for a perimeter columns level 1-3 (Based on Etabs output)

295

Excel spread sheet to calculate column:

This part shows the Excel spread sheet used to verify the designed concrete columns.

Figure B-4 Structural design: Excel spread sheet for a perimeter columns level

1-3 (reinforcement)

296

Figure B-5 Structural design: Excel spread sheet for a perimeter column level 1-3

(interaction diagram)

297

Appendix C: Hourly air room temperature distributions

This appendix provides the hourly simulated indoor air temperature for the top floor of the building plotted against the hourly outdoor temperature to compare the inside thermal performance between different types of constructions. Section 6.4.3 discusses the hourly air temperature for two selected buildings in climate zone 2 (Brisbane). The data associated with the other climate zones is provided in the following figures (Figures C-1 to C-4).

298 299

Figure C-1 hourly air room temperature plotted against the hourly outdoor temperate for the Waffle.low and 200.normal in the climate zone 1 (Darwin).

300

Figure C-2 hourly air room temperature plotted against the hourly outdoor temperate for the Waffle.low and 200.normal in the climate zone 5 (Sydney).

301

Figure C-3 hourly air room temperature plotted against the hourly outdoor temperate for the Waffle.low and 200.normal in the climate zone 6 (Melbourne).

302

Figure C-4 hourly air room temperature plotted against the hourly outdoor temperate for the Waffle.low and 200.normal in the climate zone 7 (Canberra).

303

Appendix D: Free-floating analysis during the design weeks

This appendix provides the data associated with variations of the indoor and outdoor air temperature for the designed buildings. Section 6.4.4 shows the results for two climate zones (1 and 6) during the winter design week. The free-floating analysis for the summer and winter design weeks associated with other climate zones is provided in the following section (Figures D-1 to D-8).

Figure D-1 Summer design week free-floating analysis for climate zones 1 (Darwin)

Figure D-2 Summer design week free-floating analysis for climate zones 2 (Brisbane)

304

Figure D-3 Winter design week free-floating analysis for climate zones 2 (Brisbane)

Figure D-4 Summer design week free-floating analysis for climate zones 5 (Sydney)

Figure D-5 Winter design week free-floating analysis for climate zones 5 (Sydney)

305

Figure D-6 Summer design week free-floating analysis for climate zones 6 (Melbourne)

Figure D-7 Summer design week free-floating analysis for climate zones 7 (Canberra)

Figure D-8 Winter design week free-floating analysis for climate zones 7 (Canberra)

306

Appendix E: Discomfort degree hours during the design weeks

This appendix provides the discomfort degree hours for two types of construction and structural materials (Normal and Ultra-lightweight concrete). Section 6.4.4 discusses the discomfort degree hours for summer conditions in climates 1 and 6. The analysis of discomfort degree hours associated with summer and winter conditions across the other climate zones is provided in the following section (Figure E-1 to E-8)

Figure E-1 discomfort degree hours during winter design week for climate zones 1 (Darwin)

Figure E-2 discomfort degree hours during summer design week for climate zones 2 (Brisbane)

307

Figure E-3 discomfort degree hours during winter design week for climate zones 2 (Brisbane)

Figure E-4 discomfort degree hours during summer design week for climate zones 5 (Sydney)

Figure E-5 discomfort degree hours during winter design week for climate zones 5 (Sydney)

308

Figure E-6 discomfort degree hours during winter design week for climate zones 6 (Melbourne)

Figure E-7 discomfort degree hours during summer design week for climate zones 7 (Canberra)

Figure E-8 discomfort degree hours during winter design week for climate zones 7 (Canberra)

309

Appendix F: Life cycle CO2-e emissions for the analysed buildings

This appendix provides whole lifetime CO2-e emissions normalised by gross floor area and separated by the type of concrete and method of construction. The figures associated with Darwin and Canberra were used in section 7.4 to compare the total CO2-e emissions associated with different forms of construction and types of concrete. The following figures (F-1 to F-6) provide the results for each city.

Melbourne

Pre-use and construction phase Operation (50 years) phase Total Lifetime 9,000 7,800 ) 7,468 7,638 2 8,000 7,403 7,000

(kg/m 6,000 5,000 4,000

emissions emissions 3,000

e e -

2 2,000

CO 1,000 0 200.normal 200.low Waffle.normal Waffle.low Design alternatives 2 * Victoria average state CO2-e emissions (kg/m ) intensity: 15,872

Figure F-1 Life cycle CO2-e emissions normalised by gross floor area and separated by type of concrete and construction method- Melbourne

310

Canbera

Pre-use and construction phase Operation (50 years) phase Total Lifetime 7,000 6,351

6,097 6,250 6,036 ) 2 6,000

5,000 (kg/m 4,000

3,000 emissions emissions

e e 2,000

- 2

1,000 CO 0 200.normal 200.low Waffle.normal waffle.low Design alternatives

2 * ACT average state CO2-e emissions (kg/m ) intensity: 12,140

Figure F-2 Life cycle CO2-e emissions normalised by gross floor area and separated by type of concrete and construction method- Canberra

Sydney

Pre-use and construction phase Operation (50 years) phase Total Lifetime

8,000 7,442 7,558 7,236 7,171

7,000

) 2 6,000

(kg/m 5,000 4,000

emissions emissions 3,000

e e -

2 2,000 CO 1,000 0 200.normal 200.low Waffle.normal Waffle.low Design alternatives

2 * NSW average state CO2-e emissions (kg/m ) intensity: 12,113

Figure F-3 Life cycle CO2-e emissions normalised by gross floor area and separated by type of concrete and construction method- Sydney

311

Brisbane

Pre-use and construction phase Operation (50 years) phase Total Lifetime 10,000 8,879 8,997

9,000 8,640 8,585 ) 2 8,000 7,000 (kg/m 6,000 5,000

4,000

emissions emissions e e

- 3,000 2

2,000 CO 1,000 0 200.normal 200.low Waffle.normal waffle.low Design alternatives

2 * QLD average state CO2-e emissions (kg/m ) intensity: 12,310

Figure F-4 Life cycle CO2-e emissions normalised by gross floor area and separated by type of concrete and construction method- Brisbane

Darwin

Pre-use and construction phase Operation (50 years) phase Total Lifetime 14,000 11,307 11,401 11,272 11,538

) 12,000 2

10,000 (kg/m 8,000

6,000

emissions emissions

e e -

2 4,000

CO 2,000

0 200.normal 200.low Waffle.normal Waffle.low Design alternatives 2 * NT average state CO2-e emissions (kg/m ) intensity: 9,421

Figure F-5 Life cycle CO2-e emissions normalised by gross floor area and separated by type of concrete and construction method- Darwin