156
University of Calgary PRISM: University of Calgary's Digital Repository Graduate Studies The Vault: Electronic Theses and Dissertations 2017 Reliability Assessment of Historical Masonry Structures Seyedain Boroujeni, Setare Seyedain Boroujeni, S. (2017). Reliability Assessment of Historical Masonry Structures (Unpublished master's thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/27613 http://hdl.handle.net/11023/3855 master thesis University of Calgary graduate students retain copyright ownership and moral rights for their thesis. You may use this material in any way that is permitted by the Copyright Act or through licensing that has been assigned to the document. For uses that are not allowable under copyright legislation or licensing, you are required to seek permission. Downloaded from PRISM: https://prism.ucalgary.ca

Reliability Assessment of Historical Masonry Structures

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Reliability Assessment of Historical Masonry Structures

University of Calgary

PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2017

Reliability Assessment of Historical Masonry

Structures

Seyedain Boroujeni, Setare

Seyedain Boroujeni, S. (2017). Reliability Assessment of Historical Masonry Structures

(Unpublished master's thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/27613

http://hdl.handle.net/11023/3855

master thesis

University of Calgary graduate students retain copyright ownership and moral rights for their

thesis. You may use this material in any way that is permitted by the Copyright Act or through

licensing that has been assigned to the document. For uses that are not allowable under

copyright legislation or licensing, you are required to seek permission.

Downloaded from PRISM: https://prism.ucalgary.ca

Page 2: Reliability Assessment of Historical Masonry Structures

UNIVERSITY OF CALGARY

Reliability Assessment of Historical Masonry Structures

by

Setare Seyedain Boroujeni

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF MASTER OF SCIENCE

GRADUATE PROGRAM IN CIVIL ENGINEERING

CALGARY, ALBERTA

MAY, 2017

© Setare Seyedain Boroujeni 2017

Page 3: Reliability Assessment of Historical Masonry Structures

ii

Abstract

Despite the high level of vulnerability of unreinforced masonry structures under applied loads and

the importance of their reliability evaluation, there is no formal methodology to assess the

reliability of historic masonry structures. To develop an appropriate methodology, estimations of

probabilistic models of structural resistance and load effects are required to formulate a limit state

function. The stochastic characteristics of materials play key roles in the determination of

probabilistic models of structural resistance. Therefore, in current study, methodologies for

estimating the statistical characteristics of historic masonry materials through non-destructive tests

are described. Best fit probabilistic models for load effects are also presented. Target reliability

index and different approaches for calculating suitable targets for historic structures are described

as well. Evaluation of the reliability level of historical structures through the recommended

procedure would lead to more realistic and accurate levels of reliability estimation without

requiring degradation of the historic structure.

Page 4: Reliability Assessment of Historical Masonry Structures

iii

Acknowledgements

First and foremost, I owe my sincere gratitude to Dr. Nigel Shrive, my supervisor, for his

continuous and unconditional support and encouragement during my master program. He has

always been a constant source of inspiration. This work would not have been possible without his

tireless guidance, patience, scholarly inputs, and insightful comments.

My parents, my mentors Mr. Fakhroddin Seyedain and Mrs. Ahraf Seyedain, I cannot thank you

enough for your continuous love, support and encouragement throughout every single step of my

academic and personal life.

I would like to thank my husband, Danial Arab who has always loved me unconditionally and

whose good example has taught me to work hard for the things that I aspire to achieve.

Last but not least, this research study was carried out at the Civil Engineering Department of the

University of Calgary with funding provided by the Natural Science and Engineering Research

Council (NSERC) and Canadian Masonry Design Center, whose support is appreciatively

acknowledged.

Page 5: Reliability Assessment of Historical Masonry Structures

iv

Dedication

This thesis is dedicated with respect and love to:

My wonderful parents who never stop sacrificing themselves for us,

My love and my life, Danial who has made the happiest moments of my life,

My lovely brothers, Sahand and Sepehr who never left their little sister side,

I could not have done it without your faith, support and constant encouragement.

Page 6: Reliability Assessment of Historical Masonry Structures

v

Table of Content

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

Acknowledgements ........................................................................................................................ iii

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

Table of Content ............................................................................................................................. v

List of Tables ............................................................................................................................... viii

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

List of Symbols .............................................................................................................................. xi

Chapter I: Overview ........................................................................................................................ 1

1.1 Introduction ...................................................................................................................... 1

1.2 Motivation ........................................................................................................................ 2

1.3 Contributions .................................................................................................................... 3

1.4 Thesis Summery ............................................................................................................... 5

Chapter II: Previous Research Work .............................................................................................. 8

2.1 Introduction ...................................................................................................................... 8

2.2 Literature Review ............................................................................................................. 8

Chapter III: Structural Reliability Analysis .................................................................................. 15

3.1 Introduction .................................................................................................................... 15

3.1.1 Uncertainties ........................................................................................................... 15

3.1.2 Heritage ................................................................................................................... 17

3.1.3 Past Performance .................................................................................................... 17

3.1.4 Economics ............................................................................................................... 17

3.1.5 Change in Design Practice ...................................................................................... 18

3.1.6 Service Life ............................................................................................................. 18

3.2 Limit States .................................................................................................................... 18

3.3 Limit State Function ....................................................................................................... 19

3.4 Basic Theory of Reliability Assessment ........................................................................ 20

3.5 Structural Reliability Techniques ................................................................................... 26

3.5.1 Asymptotic Techniques .......................................................................................... 26

First Order Reliability Method (FORM) ......................................................... 27

Second Order Reliability Method (SORM) ..................................................... 30

First Order/Second Order Reliability in Original Space (FOROS/SOROS) ... 30

Page 7: Reliability Assessment of Historical Masonry Structures

vi

3.5.2 Simulation Based Techniques ................................................................................. 31

3.5.3 Response Surface Method....................................................................................... 34

3.5.4 Artificial Neural Networks (ANNs)........................................................................ 34

3.6 Reliability Analysis and Software .................................................................................. 34

3.7 Deterioration................................................................................................................... 35

3.7.1 Modeling of Deterioration ...................................................................................... 37

3.8 Bayes’ Theorem ............................................................................................................. 38

3.9 Framework for Reliability Assessment .......................................................................... 39

3.10 Reliability Assessment under Seismic Loading ......................................................... 41

Chapter IV: Stochastic Modeling of Applied Load ...................................................................... 44

4.1 Introduction .................................................................................................................... 44

4.2 Dead Load ...................................................................................................................... 45

4.3 Live Load ....................................................................................................................... 47

4.3.1 Total Live Load....................................................................................................... 49

4.3.2 Point-in-time Live Load .......................................................................................... 55

4.3.3 Distribution ............................................................................................................. 55

4.4 Wind Load ...................................................................................................................... 57

4.4.1 Point-in-time Wind Load ........................................................................................ 63

4.4.2 Distribution ............................................................................................................. 64

4.5 Snow Load...................................................................................................................... 65

4.5.1 Ground Snow Load ................................................................................................. 66

4.5.2 Snow Related Loads ............................................................................................... 73

4.5.3 Point-in-time Snow Load ........................................................................................ 74

4.5.4 Distribution ............................................................................................................. 74

Chapter V: Stochastic Modeling of Historic Masonry Materials ................................................. 76

5.1 Introduction .................................................................................................................... 76

5.2 Evaluation of Structural Performance ............................................................................ 77

5.3 Modulus of Elasticity ..................................................................................................... 83

5.4 Compressive Strength .................................................................................................... 84

5.5 Shear Strength ................................................................................................................ 98

5.6 Tensile Strength............................................................................................................ 100

5.7 Cohesion and Friction Coefficient ............................................................................... 104

Page 8: Reliability Assessment of Historical Masonry Structures

vii

Chapter VI: Reliability Assessment Methodology ..................................................................... 106

6.1 Introduction .................................................................................................................. 106

6.2 Reliability Assessment Procedure ................................................................................ 106

6.3 Limitations ................................................................................................................... 112

Chapter VII: Conclusion ............................................................................................................. 114

7.1 Thesis Contributions .................................................................................................... 114

7.2 Future Work ................................................................................................................. 116

References ................................................................................................................................... 118

Appendix ..................................................................................................................................... 132

Page 9: Reliability Assessment of Historical Masonry Structures

viii

List of Tables

Table 3-1 Influential factors on the target probability of failure (Schueremans, 2001) 24

Table 3-2 Target reliability and consequences class according to GruSiBau (1981) 25

Table 3-3 Tentative target reliability values 𝛽𝑇 (𝑃𝑓𝑇) (Diamantidis, 1999, 2001) 25

Table 3-4 Required target reliability according to JCSS (2001) for a 50-year observation period

26

Table 3-5 Integration of software for structural reliability analysis 35

Table 3-6 Deterioration mechanisms 36

Table 3-7 Deterioration modeling (Maes et al., 1999) 37

Table 3-8 Degradation functions (Mori et al., 1993) 37

Table 4-1 Summary of statistical parameters for dead load 47

Table 4-2 Parameters related to probabilistic modeling of live load (JCSS, 2001) 52

Table 4-3 Statistical parameters for live load recommended by Bartlett et al. (2003) 54

Table 4-4 Summary of recommended stochastic parameters provided by different authors 54

Table 4-5 Stochastic characteristics of point-in-time live load 56

Table 4-6 Summary of recommendations for distribution type 55

Table 4-7 Stochastic characteristics for maximum wind velocity in 50-year (Bartlett et al., 2003)

59

Table 4-8 Summery of statistical parameters transformation factors reported by different authors

60

Table 4-9 Statistical characteristics of variables being influential on wind load (JCSS, 2001) 63

Table 4-10 Statistical parameters of maximum 3-hour wind velocity in Canada 64

Table 4-11 Summery of the best fit distribution for parameters involved in wind load estimation

65

Table 4-12 Characteristics of the annual maximum depth in Canada (Bartlett et al., 2003) 67

Table 4-13 Characteristics of the 50-year maximum depth in Canada (Bartlett et al., 2003) 67

Table 4-14 Statistical characteristics of snow density (Kariyawasam et al., 1997) 68

Table 4-15 Statistical characteristics of transformation factor (Taylor & Allen, 2000) 70

Table 4-16 The exposure coefficient (𝐶𝑒) and shape factor (ƞ𝑎) (JCSS, 2001) 71

Table 4-17 Summary of stochastic characteristics of snow load variables (JCSS, 2001) 73

Table 4-18 Statistical characteristics of point-in-time snow (Bartlett et al. 2003) 74

Page 10: Reliability Assessment of Historical Masonry Structures

ix

Table 4-19 Summery of recommended distributions for the snow load random variables 75

Table 5-1 Nondestructive testing techniques and in-situ evaluation approaches 79

Table 5-2 Summary of standards and codes including various non-destructive methods 81

Table 5-3 Classification of testing techniques appropriate for evaluation of masonry (Harvey JR.

& Schuller, 2010) 82

Table 5-4 Values of 𝑐1 (Schubert, 2010) 83

Table 5-5 Masonry compressive strength determination from the strengths of its components 86

Table 5-6 Statistical characteristics of 𝑌 involved in probablistic model of compressive strength

JCSS (2011) 86

Table 5-7 Summary of relations between 𝐸 and 𝑓𝑚′ found in literatures 93

Table 5-8 Values of the CoV of compressive strength reported by different researchers 95

Table 5-9 Recommended distributions found in the literature 95

Table 5-10 Values of compressive strength found in the literature (TLM is Thin-Layer Mortar) 96

Table 5-11 Flexural tensile strength of masonry (Schubert, 2010) 101

Table 5-12 Statistical characteristics of 𝑌 involved in probablistic model of tensile strength

(JCSS, 2011) 103

Table 5-13 Statistical characteristics of 𝑌 involved in probablistic model of cohesion

(JCSS, 2011) 104

Table 5-14 Statistical characteristics of 𝑌 involved in probablistic model of friction coefficient

(JCSS, 2011) 105

Table A-1 Probabilistic model of dead load 136

Table A-2 Probabilistic model of live load 137

Table A-3 Probabilistic model of wind load 138

Table A-4 Considered load combination 138

Table A-5 Probabilistic model of modulus of elasticity 139

Table A-6 Probabilistic model of compressive strength 139

Page 11: Reliability Assessment of Historical Masonry Structures

x

List of Figures

Figure 3-1 Flowchart for general evaluation procedure of existing structures. ............................ 16

Figure 3-2 Load effect and resistance probability density functions. ........................................... 22

Figure 3-3 Limit state function and probability of failure. ........................................................... 28

Figure 3-4 Standard normal space related to FORM and SORM. ................................................ 33

Figure 3-5 FOROS illustration...................................................................................................... 33

Figure 3-6 Example of the updating compressive strength of a masonry unit using Bayes’

theorem (Glowienka, 2007) .......................................................................................................... 38

Figure 3-7 Reliability assessment framework............................................................................... 40

Figure 4-1 Representation of live load. ......................................................................................... 48

Figure 4-2 Redistribution of snow load applied on a duopitch roof (JCSS, 2001). ...................... 71

Figure 4-3 𝐶𝑟𝑜 as a function of the roof angle. ............................................................................. 72

Figure 5-1 Ultrasonic Pulse Method ............................................................................................. 84

Figure 5-2 Flatjack technique ....................................................................................................... 90

Figure 5-3 Schematic of flatjack technique. ................................................................................. 91

Figure 5-4 Distribution of the coefficient k for clay and concrete. .............................................. 92

Figure 5-5 Cumulative probability of coefficient k. ..................................................................... 94

Figure 6-1 Flowchart for reliability assessment of structures. .................................................... 110

Figure 6-2 Recommended approaches for estimation of the probabilistic models of historic

masonry materials ....................................................................................................................... 111

Figure A-1 Clay brick masonry shear wall ................................................................................. 132

Figure A-2 General compression failure ..................................................................................... 133

Figure A-3 Flexural failure ......................................................................................................... 134

Figure A-4 Sliding failure .......................................................................................................... 134

Figure A-5 Diagonal shear failure .............................................................................................. 135

Page 12: Reliability Assessment of Historical Masonry Structures

xi

List of Symbols

Symbol Definition

Chapter III

𝐿

Load effect

𝑅 Structural resistance

𝑓 Probability density function

𝐹 Cumulative distribution function

𝑃𝑓 Probability of failure

𝑋 Random variable

𝜃 Limit state model uncertainty

CoV Coefficient of Variation

𝑔(𝑥) Limit state function

β Reliability index

𝛽𝑇 Target reliability index

𝑝𝑓𝑇 Target failure probability

𝑡𝐿 Preset reference period

𝑅𝑓 Probability of survival

A Analysis model uncertainty

A Effective area

A Desired peak acceleration

𝐴𝐼 Influence area

𝐴𝑐 Activity factor

𝐴0 Reference area

A1 Independent area 1

A2 Independent area 2

Z Safety margin

𝑚 Mean value

𝜎 Standard deviation

ф Cumulative distribution function of the variable with a standard normal

distribution

∅ Probability density function of a standard normal variable

𝛼 Sensitivity factor

𝑡𝐿 Residual service life

𝑛𝑝 Number of endangered lives

𝑆𝑐 Social criterion factor

W Warning factor

𝐶𝑓 Cost factor

𝑥𝑖∗ Design point on the limit state function

Page 13: Reliability Assessment of Historical Masonry Structures

xii

𝑥′∗ Vector of the transformed random variables involved in design point

∇𝑙(𝑥∗) Gradient vector of the log-likelihood

𝑃(𝑥∗) Projection matrix

𝑃∗(𝑥∗) Density matrix minus the projection matrix

𝐻𝑙(𝑥∗) Hessian matrix of likelihood

𝑇 Transpose and matrix determinant

det Transpose and matrix determinant

𝑋 Vector of basic random variables

𝑁 Total number of samples

𝑁𝑓 Total number of failures

𝑓𝑑𝑒𝑔 Degradation function

𝑡

t

Time

Envelope function

Sf Unit power spectrum

fp Predominant frequency

TD Duration

𝑎𝑘(t) K time histories

𝐹𝑋𝑖𝑗(𝑥𝑖𝑗/𝑎)

Conditional cumulative distribution function for 𝑖th critical variable and the

𝑗th member (𝑋𝑖𝑗)

𝑅𝑖𝑗 Reliability conditional on 𝑖th critical value for the 𝑗th member

𝑦𝑖𝑗 Capacity of 𝑗th member for 𝑖th critical value

𝑓𝐴(𝑎) Probability distribution function of A during any one year

𝐹𝑋𝑖,𝑛 Cumulative demand distribution in n-year life time

𝑓𝑌𝑖(𝑧)

Chapter IV

Density function of the capacity of a critical variable

𝑊 Load intensity (psf)

𝑌 Random variable modeling the mean of load on the floor

є (𝑥, 𝑦) Stochastic process with zero mean associated with the deviations from the

average

𝑉 Random variable (with mean zero) accounting for the variation of the load

𝑈(𝑥, 𝑦) Random field concerning the spatial variation related to the load

𝑞 Equivalent load with uniform distribution

𝑈 Unit load

𝐸 Expected value operator

Var Variance

µ Mean of all unit floor loads associated with office buildings

𝜎2 Variance in individual floor

𝜎𝑠 An experimental constant

I(x,y) Influence function

Page 14: Reliability Assessment of Historical Masonry Structures

xiii

𝐿𝑚𝑎𝑥 Maximum sustained uniformly distributed load

λ Occurrence rate of sustained load changes in [1/year]

�̅�𝑚𝑎𝑥 Maximum equivalent uniformly distributed live load during a 50-year

reference period

𝐿 Nominal uniformly distributed live load

𝐵 Tributary area

𝜎𝐿𝑚𝑎𝑥

2 Variance of maximum live load during 50 years

𝑝 Wind pressure against the surface of structures

𝑞 Reference velocity pressure

𝐶𝑒 Exposure factor

𝐶𝑝 External pressure factor

𝐶𝑔 Gust factor

𝑞𝑇 1 in 𝑇 year reference wind pressure

𝑉𝑇 1 in 𝑇 year wind velocity

𝜌 Density of air

�̅�50 Mean maximum velocity likely in a 50-year design life

𝑉50 1 in 50-year velocity

𝐶𝑜𝑉𝑎 CoV of the maximum annual wind velocity

𝐶𝑎 Aerodynamic shape factor

𝐶𝑔 Gust factor

𝐶𝑟 Roughness factor

𝐶𝑑 Dynamic factor

𝑆 Snow load on a roof of a structure

𝐼𝑠 Importance factor for snow load

𝑆𝑠 Ground snow load with the probability of exceedance of 1 in 50 per year

𝑆𝑟 Associated rain load

𝑆𝑔 Ground snow load

𝐶𝑏 Basic roof snow load

𝐶𝑤 Wind exposure factor

𝐶𝑠 Slope factor

𝐶𝑎 Accumulation factor

𝐶𝑔𝑟 Overall combination of transformation factors

𝐶𝑒 Exposure coefficient

𝐶𝑡 Thermal coefficient

𝐶 Ground to roof transformation factor

𝑑 Snow depth

𝛾 Unit weight of snow

𝜆𝑔 Mean of 𝑙𝑛 𝑆𝑔

𝜉𝑔2 Variance of 𝑙𝑛 𝑆𝑔

Page 15: Reliability Assessment of Historical Masonry Structures

xiv

𝛾(𝑑) Average weight density of the snow

𝛾∞ Unit weight at 𝑡 = ∞

𝛾0 Unit weight at 𝑡 = 0

𝐸 Wind exposure factor

T Thermal characteristic factor

𝜀 Error term

𝑟 Transformation factor of snow load on ground to snow load on roofs

ℎ Altitude of the site of the building

ℎ𝑟 Reference altitude

𝑘 Coefficient accounts for the type of region in which structure is located

ƞ𝑎 Shape coefficient

𝐶𝑟 Redistribution coefficient

𝑢(𝐻) Averaged wind speed during one-week period at roof level of H

𝑚𝑔 Sample mean

𝑠𝑔

Chapter V

Sample standard deviation

𝐸𝑚 Mean of modulus elasticity of masonry

𝑐1 Coefficient involved in probabilistic model of modulus of elasticity

𝑐2 Coefficient involved in tensile strength perpendicular to the units

𝑐3 Coefficient involved in tensile strength parallel to the units

𝑓𝑚′ Compressive strength of masonry

𝑌2 Log-normal variable

𝐸𝑚,𝑗 Stochastic modeling of modulus elasticity of masonry

𝑓𝑏 Mean value of the unit compressive strengths

𝑓𝑚𝑜 Mean value of mortar compressive strengths

𝑘 Coefficient

𝑌 Variable related to uncertainties

𝑓𝑚,𝑗 Stochastic modeling of compressive strength

𝛼 Coefficient

𝛽 Coefficient

𝐸𝐿 Modulus of elasticity under Long-term loading

𝐸𝑚 Modulus of elasticity under short-term loading

𝑓𝑏𝑡,𝑙 Mean of tensile strength perpendicular to units

𝑓𝑏𝑡,𝑠 Mean of tensile strength parallel to the units

𝑓𝑏𝑡,𝑙,𝑗 Stochastic modeling of tensile strength perpendicular to units

𝑓𝑏𝑡,𝑠,𝑗 Stochastic modeling of tensile strength parallel to the units

𝑓𝑣,𝑗 Stochastic modeling of cohesion

𝑓𝑣,𝑚 Mean of cohesion

𝜇𝑚 Mean of friction coefficient

Page 16: Reliability Assessment of Historical Masonry Structures

xv

𝜇𝑗 Stochastic modeling of Friction coefficient

Abbreviations

ANNs Artificial Neural Networks

ULS Ultimate Limit State

SLS Serviceability Limit State

PLS Phenomenological Limit State

MVFOSM Mean Value First Order Moment

FORM First Order Reliability Method

SORM Second Order Reliability Method

FOROS/SOROS First Order/Second Order Reliability in Original Space

MCS Monte Carlo Simulation

IS Importance Sampling

FEM Finite Element Model

RSM Response Surface Method

PGA Peak Ground Acceleration

GPM General Purpose Mortar

TLM Thin Layer Mortar

UPM Ultrasonic Pulse Method

CS Calcium Silicate

AAC Autoclave Aerated Concrete

LC Lightweight Concrete

CB Clay Brick

Page 17: Reliability Assessment of Historical Masonry Structures

1

Chapter I: Overview

There is an undeniable fact that uncertainty is an inseparable part of every scientific field. This

fact highlights the need for probabilistic treatment of problems in different fields of study. In order

to avoid complex solutions and calculations, deterministic approaches are commonly prevalent in

structural engineering. Semi-probabilistic methods are also used as simplified methods in order to

combine deterministic and probabilistic approaches. However, the most accurate and efficient

solutions can be achieved through fully-probabilistic approaches which guarantee comprehensive

reliability assessment of structures.

Historic structures are of great importance as they reflect the history and culture of a people and

their society. The conservation and maintenance of heritage structures requires considerable care.

Evaluation of the reliability of historic masonry structures has gained significant importance, since

the consequences of structural failure could be serious and irreparable. Heritage masonry structures

are typically unreinforced and have low and limited load bearing capacity. Therefore, assessment

of the reliability of these structures is of special importance (D’Ayala and Speranza, 2003).

Modern codes have been developed mostly for the design of new buildings based on modern

materials, whereas historical structures were designed according to the rules and experiences

available at the time. The expected service life of a historical structure was more than that of a new

structure designed by current codes, because historical structures were expected to be preserved

for future generations. These buildings have withstood all the applied loads during their service

life to date which provides some evidence as to the serviceability and safety of these structures.

Based on the age of a structure and the type of the applied loads, different information can be

obtained from its past performance. As an example, in the case of a 30-100 year-old building which

has experienced earthquakes with ground motions less than the design event during its service-life

to date, one may conclude that the structure has presented satisfactory performance regarding the

dead and other variable loads but not potential seismic ones. In contrast, a historical structure

Page 18: Reliability Assessment of Historical Masonry Structures

2

dating back a number of centuries, has demonstrated satisfactory performance under a much bigger

variety of loads and hazards.

As these heritage structures were constructed before modern codes and standards were written,

current codes will not necessarily provide satisfactorily accurate estimates of load or resistance

when it comes to evaluation of the historical materials or the structure. This inaccuracy may lead

to inefficient and even destructive strengthening modifications. Due to the uncertain nature of the

applied loads as well as unexpected structural responses under different loading conditions,

deterministic evaluations may result in inaccurate evaluations and unnecessary strengthening.

Probabilistic approaches are used to get more accurate evaluations and consequently efficient

modifications. Through probabilistic reliability assessment of masonry structures, the necessity

and level of upgrading and rehabilitation can be defined, which should result in more appropriate

and efficient interventions (Lagomarsino and Cattari, 2015).

When “upgrading” or strengthening an existing structure, it could be more expensive to intervene

to meet a certain criterion compared to providing structural safety in the design stage of a new

structure. Therefore, determining which criteria should be satisfied and subsequently, the level of

upgrading is of special importance. Preservation of the values of a historical building plays a key

role in upgrading. Preservation should minimize the destruction of the originality of a historical

structure in terms of both materials and architecture.

Identification of vulnerable historical structures and determination of a reasonable criterion for

upgrading and strengthening them are major challenges in structural engineering and cultural

heritage preservation. To date, considerable research has been allocated to structural steel,

concrete and timber but only a few studies have been focused on the reliability assessment of

masonry structures. Despite the high level of vulnerability of unreinforced masonry structures

under applied loads and the importance of the reliability evaluation of historic masonry structures,

there is no specified methodology to assess the reliability of them.

Page 19: Reliability Assessment of Historical Masonry Structures

3

Moreover, there are difficulties to apply the methodologies recommended by the National Building

Code of Canada (NBCC, 2010) and Canadian Standard Association (CSA, 2014) standards with

respect to the structural evaluation of existing structures (Allen,1991a), including:

In the case of historical structures, there are structural systems, components and materials which

are not addressed by the NBCC (2010) or CSA (2014). Structural evaluation faces difficulties due

to the lack of information on their structural properties.

Although historical structures do not mostly follow the instructions of the code, they have

performed satisfactorily over the years. Also, a number of structural characteristics such as dead

load, strength, deterioration, fatigue and creep can be measured directly. These constitute some of

the information not considered by the NBCC’s and CSA’s design criteria.

Most of the requirements are focused on specifying the required and economical percentage of

strengthening materials and their related arrangement during construction. But, it is uneconomical

to apply these strengthening modifications to an existing structure. Therefore, alternative

requirements need to be defined.

Consequently, as an initial step towards satisfying this goal, the main focus of the research

described in this thesis is to develop a determinate, step-by-step methodology for assessing the

reliability level of historical masonry structures under applied loads. The presented methodology

can be used in a code to estimate the reliability level with some modifications and generalization.

To achieve the goal, the relevant literature is reviewed, followed by a description of reliability

analysis. Methods to obtain the necessary structural information for reliability analysis are

proposed allowing a logical assessment method to be described.

Despite the high level of vulnerability of historic structures to the applied loads and the necessity

of reliability assessment of them, there is still no formal approach for reliability assessment of

historical masonry structures. The current thesis contributes to the development of a step-by-step

reliability-based assessment procedure being appropriate and applicable for historical masonry

structures. The main contributions of this study are:

Page 20: Reliability Assessment of Historical Masonry Structures

4

1) A comprehensive literature review was done to understand various aspects of reliability

assessment of masonry structures specifically historical masonry ones. The theory of reliability

assessments of structures, reliability assessment techniques and suitable combination among finite

element structural analysis, reliability analysis methods and reliability software are reviewed

precisely.

2) Values for the target reliability index or target probability of failure (as decision criteria) being

appropriate for historic masonry structures are determined. Different values of target probability

of failure or target reliability index are recommended by different codes and standards. The

necessity of meeting target probability of failure values or the target reliability index recommended

by different codes and standards is still a controversial issue in the case of historic masonry

structures. This is due to that fact that the recommended values are determined for new masonry

materials, structural configurations and construction methodologies and may not be appropriate

for historical masonry structures having specific criteria and requirements. A number of

calculation approaches and target reliability index and failure probability values showed to be

appropriate for historic structures are discussed here.

3) Deterioration as a prevalent destructive process happening in historic structures may result in a

change in the structural resistance of masonry, the loading conditions or analysis model. In this

thesis, approaches for integrating deterioration into reliability assessment are investigated and

degradation functions are reported.

4) Reliability assessment of structures subjected to earthquakes is still challenging and complex.

This is due to the fact that different uncertainties including those in the nature of ground motions

and those in nonlinear behaviour of structures are involved in reliability assessment under seismic

loading. Reliability assessment of existing structures under seismic loading is studied and a

specific methodology is described in this regard.

5) To formulate the limit state function as the base of reliability assessment, probabilistic models

of applied loads have to be determined. In this study, the best stochastic models of applied loads

are presented.

6) The statistical characteristics of materials play key roles in the estimation of the structural

resistance probabilistic model. Various techniques can be used to determine stochastic

Page 21: Reliability Assessment of Historical Masonry Structures

5

characteristics of historic masonry materials and to get more realistic information about the

material properties. However, destructive testing methods are not recommended as such tests may

result in irreparable damage to these valuable structures. Moreover, some techniques are not

capable of determining the variability of material properties (over an element or structural system).

As estimation of the variability is essential for reliability assessment, these techniques may be

considered inappropriate with respect to structural reliability assessment. Therefore, the

applicability and capabilities of current testing techniques in order to estimate the statistical

characteristics of historic masonry materials are studied in this research.

7) Compressive strength is the most important material property influencing the structural

resistance of historic masonry structures and consequently, influences the reliability assessment of

these structures. Codes are necessarily conservative and are also generally aimed at design or

assessment with modern masonry materials. Therefore, the use of code values for historical

structures may result in inaccurate structural reliability assessment. Various techniques can

estimate the mean value of compressive strength. However, testing techniques which damage the

structure (for example, by removing samples of the masonry) are not recommended for historical

buildings because of their destructive nature. Moreover, there are restrictions in the application of

some destructive techniques for estimation of the variability of compressive strength (over an

element or structural system). Here, a procedure is proposed to determine probabilistic model of

historic masonry compressive strength using non-destructive testing techniques.

Chapter I introduces the necessities of reliability assessment of historical masonry structures. to

do so, the most important difficulties, concerns and restrictions in the application of the new

reliability assessment methods for historic masonry structures are presented and the research

motivations are expressed.

A brief review of the literature focusing on the reliability assessment of historic masonry structures

were discussed in Chapter II.

Chapter III explains structural reliability analysis. Reliability assessment and the concerns

associated with the reliability assessment of existing structures are discussed. Basic theory and the

Page 22: Reliability Assessment of Historical Masonry Structures

6

required procedure of reliability assessment are explained. Target reliability index and failure

probability values and different approaches for calculating suitable targets for historic structures

are described. Different limit state functions are introduced. Available techniques for reliability

assessment of structures are briefly discussed. Deterioration, as a common physicochemical

process resulting in gradual alteration in characteristics of historical masonry, is introduced and

the integration of deterioration into reliability analysis is studied. Finally, as the reliability

assessment of structures under seismic loading is complicated compared to the reliability

assessment under other loading conditions, the appropriate methodology for reliability assessment

of masonry structures subjected to earthquake is investigated.

An overview on probabilistic modeling of applied loads as one of the essential parameters in the

determination of the limit state function is provided in Chapter IV. The stochastic models of dead

load, live load, wind load, snow load are described. The recommendations of different researchers

and codes with respect to probabilistic models and the stochastic characteristics of applied loads

are studied and compared together. Point-in-time components of live load, wind load and snow

load were also studied.

Chapter V covers the stochastic characteristics and probabilistic models of historic masonry

materials. Masonry materials and their associated stochastic characteristics play key roles in the

determination of probabilistic models of structural resistance. A summary of the techniques which

are used to evaluate the characteristics of masonry materials is presented. The advantages and

disadvantages of these testing techniques as well as their feasibilities and capabilities in the

determination of the stochastic characteristics of historic masonry materials are studied. As codes

of practice mostly recommend calculation procedures, values and the best fit distributions for new

masonry materials, the accuracy, applicability and accordance of these recommendations to

historical masonry materials are investigated and discussed. Compressive strength is usually key

information in the determination of structural resistance of historic masonry structures. Therefore,

special attention is given in this thesis to the determination of the stochastic characteristics and

probabilistic model of historical masonry compressive strength using non-destructive techniques.

An approach is proposed here to estimate the compressive strength of ancient masonry structures

and its associated variability without any degradation to the historical value.

Page 23: Reliability Assessment of Historical Masonry Structures

7

Chapter VI provides a summary of the proposed procedure for reliability assessment of historic

masonry structures as well as the limitations.

Chapter VII offers concluding remarks of this research including thesis contributions and

recommendations for future work.

Page 24: Reliability Assessment of Historical Masonry Structures

8

Chapter II: Previous Research Work

Structural evaluation plays a key role in the preservation and use of structural and cultural

heritages. Traditionally, variables used in structural evaluation approaches are treated as

deterministic values. The reality that applied loads and structural resistances are uncertain and

parameters are random necessitates the application of probabilistic approaches for structural

evaluation leading to structural reliability assessment. Historical structures are among the most

complicated structures in terms of reliability assessment. Different material properties with various

stochastic characteristics, complicated configurations, shortage of available construction

documents and limitations in the application of destructive testing techniques are the main

concerns and difficulties in reliability assessment of historic structures. A brief review of previous

research work focusing on different aspects of reliability analysis of historic structure is presented

here.

Reliability assessment of structures is a developing field of civil engineering. In 1926, Mayer

suggested considering deviations in design procedure through stochastic techniques to lead to safer

and more reliable designs (reported by Brehm, 2011). Streletzki (1974) and Wierzbicki (1936) also

introduced load and resistance parameters as random variables and consequently, a probability of

failure for each structure. Although probabilistic approaches are obviously beneficial, they have

not been the focus of much research due to the computational complexity. The theory of reliability

assessment of structures was developed significantly by Freudenthal (1947: 1956). As the

formulations involving convolution functions were difficult to be solved by hand, it was not

possible to apply reliability analysis in practice until the work of Cornell and Lind during the late

1960s and early 1970s. A second moment reliability index was proposed by Cornell (1967). A

description of a format-invariant reliability index was provided by Hasofer and Lind (1974). Then,

Rackwitz and Fiessler formulated an efficient numerical procedure to calculate the reliability index

Page 25: Reliability Assessment of Historical Masonry Structures

9

in 1978. To do reliability analysis, Melchers (1999) and Van Dyck (1995) proposed the use of

basic reliability problems to outline the generalized reliability problem.

The Joint Committee on Structural Safety (JCSS) was founded in 1976. Since then, this committee

has played a considerable role in support of the development and improvement of probabilistic

approaches and reliability methods. Moreover, the JCSS has provided and enlarged the available

database. The “Probabilistic model code” (JCSS, 2003) is the outcome of all the committee’s

efforts and findings which provides users with probabilistic models related to different variables

in the field of civil engineering. The more recent version (JCSS, 2011) covers stochastic models

of random variables of different structures categorized based on their construction materials,

including concrete structures, steel structures, and masonry structures. However, this version

(JCSS, 2011) includes statistical characteristics of new construction materials in the case of

masonry structures which is not the focus of this research. Therefore, there is still a lack of

knowledge in the guidelines recommended by JCSS (2011) regarding the statistical characteristics

of historic masonry materials.

Several researchers have developed general procedures and guidelines for reliability assessment

of structures. Ellingwood (1996) investigated the condition assessment of existing structures using

reliability assessment. His research mostly focused on general and somehow qualitative

explanations regarding the procedures for probabilistic assessment of existing buildings. Faber

(2000) reviewed the general philosophy, theoretical concepts and tools being important to be able

to do reliability assessment of existing structures. General comments regarding the application of

reliability analysis in the assessment of steel structures and concrete structures were presented.

Rackwitz (2001) also reviewed theory and methods of structural reliability. The cases for which

these methods are not applicable were explained and new fields having potential application were

discussed. The importance of sampling schemes to update probability estimates were outlined.

A report prepared by BOMEL Limited for the Health and Safety Executive (2001) reviewed the

developments of structural design methods including the earliest building methods, limit state

design methods and probabilistic analysis. The report focused on the application of probabilistic

methods in the design and assessment of pressure systems. Guidelines regarding risk analysis in

conjunction with reliability analysis were also prepared. The report was mainly for industry use.

The Federal Institute of Materials Research and Testing (BAM) (2006) prepared a guideline for

Page 26: Reliability Assessment of Historical Masonry Structures

10

the assessment of existing structures (F08a). The report contains a chapter titled “probabilistic

assessment routine”. However, the chapter only includes general explanations and

recommendations regarding structural reliability assessment and procedures rather than more

specific guidelines. ISO 13822 (2010) provided general guidelines regarding the assessment of

existing structures.

Ditlevsen and Madsen (1996) published a book regarding the structural reliability. This book

includes both structural reliability philosophy and methods in details. Reliability of existing

structures and system reliability analysis are also included in this book. This book mostly covers

the reliability assessment of new structures rather than historical ones.

Reliability assessment of an existing building subjected to earthquakes has not been the focus of

many researchers due to the complexity. Among the research which has been done in this field,

the work of Shah and Dong (1984) is the one of the more comprehensive. They (Shah & Dong,

1984) presented a methodology to assess the structural reliability of existing buildings under

probabilistic seismic loadings. Through the presented methodology, the intensity and frequency

content of the ground motion are represented using a peak power spectrum obtained through

seismic hazard evaluation for the site. A set of time histories are then generated by Monte-Carlo

simulation (MCS) with the assumption of a Gaussian model for seismic time histories. Regarding

the response parameter being considered, a statistical analysis is done. A Gumbel distribution

(Generalized Extreme Value distribution Type 1) is fitted to the data. A probability distribution is

attributed to the ultimate ductility capacity and finally, the reliability level of the structure is

assessed through comparison of the distributions of demand and capacity.

Damage and deterioration are unavoidable aspects of existing structures. An analytical

investigation was carried out to examine the influence of damage and redundancy deterioration on

structural reliability (with a focus on trusses and bridges) by Frangopol et al. (1987). They showed

that probabilistic assessment of damaged structures should be done considering system reliability

methods rather than a single element one. Moreover, it was found that deterministic assessment

methods (rather than probabilistic approaches) can be reasonably accurate to estimate the influence

of deterioration on structural reliability.

Some researchers worked on different aspects of reliability assessment of masonry structures.

Schueremans and Van Gemert (1998) developed a probabilistic methodology to evaluate the

Page 27: Reliability Assessment of Historical Masonry Structures

11

reliability of masonry structures with a focus on theory and calculations of probability of local

failure in the case of masonry shear walls. The study was part of a program involving structural

strengthening of ancient masonry considering risk analysis, grouting and non-destructive tests.

Two case studies also were investigated and the reliability of each structure was assessed including

a masonry sewer system and the facade of the St.-Amandus chapel at Erembodegem in Belgium.

Maes et al. (1999) developed a general limit state framework concerning loads, resistances and

stresses. Different structural limit states were introduced including structural limit states (ultimate

limit states and serviceability limit states) and phenomenological limit states (cultural value limit

state, appearance limit state and sustainability limit state). Besides performance requirements

being common between new buildings and old buildings, other requirements which are influential

on reliability-based assessment of existing buildings were presented. Economics, heritage value,

uncertainties, past performance, changing design practice and service life are among the most

important ones. Different deterioration mechanisms and their incorporation into reliability

assessment were discussed.

A procedure to calculate structural reliability of common masonry walls under vertical bending

was presented by Stewart and Lawrence (2002). The procedure was applied to a masonry column

and the probability of failure was calculated according to an ultimate limit state (crushing of the

brick masonry) and an appearance limit state (first-cracking). Structural reliability was found to

be sensitive to wall width, workmanship and the difference between the designed and the

constructed thickness.

Lawrence and Stewart (2009) also examined the reliability of masonry walls which were designed

to be under compression according to AS3700. The probabilistic models of behaviour were

obtained using the results of many tests done on full-scale masonry walls. Then, the structural

reliability of unreinforced masonry walls under concentric vertical loading was determined

considering different variants including unit compressive strength, mortar type, tributary area and

live-to-dead load ratios. They (Lawrence & Stewart, 2009) compared the calculated reliabilities to

a target reliability index specified in AS5104 and suggested an increment in the capacity reduction

factor (∅) from 0.45 to 0.75 for 𝐻/𝑡 < 30 where 𝐻 stands for height and 𝑡 is thickness. Lawrence

and Stewart (2011) also worked on the development of a methodology for reliability assessment

of unreinforced masonry walls under vertical bending. Using reliability analysis, they calibrated

Page 28: Reliability Assessment of Historical Masonry Structures

12

the reduction factor (partial safety factor) for bending. Model error, wall height and discretization

of wall thickness were investigated as influential parameters on the reliability of masonry walls

under vertical bending. The probabilistic model of model error was determined through the

stochastic analysis of results obtained from testing of 118 walls. Statistical characteristics of

flexural tensile strength were derived based on the stochastic analysis of a database containing

7332 individual measurements of bond strength. Through the comparison of calculated reliabilities

with their associated target reliability index, the capacity reduction factor (∅) reported by AS3700

was recommended to decrease from 0.6 to 0.47. However, they (Lawrence & Stewart, 2011)

suggested further study be done on the influence of wall length and mortar type to see if these

perameters are influential on the structural reliability and if the behaviour model needs to be

improved.

Graubner and Glowienka (2008) provided the main statistical parameters required for the

reliability assessment of big size masonry. They also quantified model uncertainties which have

to be considered in the probabilistic calculations. They (Graubner and Glowienka, 2008) believed

that the statistical characteristics of big sized masonry showed small scatter compared to other

kinds of masonry.

An investigation was done on reliability assessment of masonry arch bridges considering both

serviceability and ultimate limit state functions by Casas (2010). Various failure modes as well as

the structural behaviour of the masonry arch bridges under low and high load levels were studied.

A methodology for reliability assessment of masonry arch bridges was developed. This

methodology was then applied to an existing arch bridge. However, due to the lack of experimental

data, the experimental application of the proposed methodology should be considered as

preliminary. Moreover, there is a need to update the preliminary proposed models with more

laboratory and full scale testing results. It was mentioned that probabilistic assessment of masonry

arch bridges is not common as the determination of their failure criteria is difficult. This is due to

that fact that these kinds of structures usually show a global nature of failure rather than based on

the failure of components.

Schueremans and Verstrynge (2008) worked on the use of reliability assessment techniques to

evaluate the safety of the Romanesque city wall of Leuven. Reliability analysis was done

considering three limit states including rotation equilibrium, stress in the masonry and stress in the

Page 29: Reliability Assessment of Historical Masonry Structures

13

subsoil. The geometry and material properties were estimated through survey. The procedures of

reliability analysis and safety assessment are briefly explained. Brehm and Lissel (2012) studied

the reliability of unreinforced masonry bracing wall subjected to wind loading. In this study, the

walls were designed according to the German code to satisfy the provided level of reliability. The

shear capacity of the walls was determined by both analytical models and test data, and were

compared to determine the most realistic model.

Al-habahbeh & Stewart (2015) investigated the stochastic reliability of unreinforced masonry

(URM) walls under blast loading. Finite element modeling (FEM) was combined with MCS in this

study. They concluded that URM walls are highly reliable in the case of moderate blast. Wall

thickness, blast pressure, bond strength and material strength are the most influential parameters

on the reliability of URM walls subjected to blast. Mojsilovic´ and Stewart (2015) studied the

reliability of mortar joint thickness in load bearing masonry walls. Four different sites in

Switzerland were chosen to collect data on the thickness of the mortar joints in clay block masonry

walls. Using reliability analysis of the masonry under compression, it was concluded that

probabilistic modeling of bed joint thickness leads to a higher reliability index than that calculated

from a deterministic value of bed joint thickness. Different levels of reliability were observed at

the different sites. Moreover, thicker joints resulted in lower reliability indices.

The realistic reliability analysis of complex structural systems involving different materials and

different structural elements and systems has become an important challenge in the field of

structural reliability assessment and analysis. The finite-element method is a powerful tool being

used in the analysis of complicated structures, and has been combined with reliability analysis in

attempts to capture the desirable aspects of these two methods (reliability analysis and the finite

element method). The outcome of these attempts was the stochastic finite element method (SFEM)

developed by several researchers (Haldar and Nee, 1989; Haldar and Gao, 1997; Haldar and

Mahadevan, 2000).

Quite a few researchers have studied the structural reliability of historical masonry structures.

Typically, the research studies were specifically aimed at performing a reliability assessment of a

masonry element under determined loading conditions and did not propose a step-by-step

methodology for reliability assessment of historic masonry structures using appropriate

approaches and formulae.

Page 30: Reliability Assessment of Historical Masonry Structures

14

The literature review which contributes to the content of individual thesis chapters, is mentioned

in the relevant sections.

Page 31: Reliability Assessment of Historical Masonry Structures

15

Chapter III: Structural Reliability Analysis

Reliability assessment is a procedure for evaluating the reliability of a historic building during a

performance period, based on information gathered about the building, from material properties to

structural layout. In other words, reliability assessment includes estimation of the reliability level

of a structure considering certain limit states during its service life. Figure 3-1 shows the flowchart

for a general evaluation procedure of existing structures (ISO 13822, 2003). Human safety, human

comfort, building function, damage and economics are the fundamental structural safety and

performance requirements. The basics of safety and performance requirements are similar in cases

of both new and historic structures. However, there are some concerns regarding the assessment

of historic structures: these concerns include the six factors discussed below (Maes, 1999).

3.1.1 Uncertainties

Uncertainty refers to a situation associated with imperfect and/or unknown information.

Uncertainty is categorized as aleatory uncertainty (inherent natural variability) and epistemic

uncertainty (model uncertainty, statistical uncertainty and measurement error). There are many

sources of uncertainty, including material heterogeneity and lack of construction details. At the

design stage, load and resistance factors are representative of the uncertainties in loads and

resistance. At the evaluation stage, existing structures may experience either an increase (e.g. due

to deterioration) or a decrease (e.g. measuring properties by test) in load and resistance

uncertainties in comparison with the design stage. Therefore, the range of uncertainties is broader

in the evaluation stage which necessitates identification of key components and details of the

structure (Allen, 1991a). Monitoring structures during the evaluation procedure can provide

information to update the uncertainties.

Page 32: Reliability Assessment of Historical Masonry Structures

16

Figure 3-1 Flowchart for general evaluation procedure of existing structures (ISO 13822, 2010)

Requests/Needs

Identification of the evaluation

Scenarios

Preliminary evaluation

Investigation of documents and other evidence

Initial inspection

Initial check

Decisions on immediate plans

Suggestions for detailed evaluation

Doubt about safety?

Detailed evaluation?

Detailed review and search for documents

Detailed inspection and material examination

Determination of plans

Determination of structural properties

Analysis of structures

Verification

Additional inspection?

Reporting evaluation outcomes

Decision

Reliability sufficient?

Intervention

Construction

Strengthening

Demolish, build new

Operation

Monitoring

Change or restriction

in use

Maintenance

Periodical inspection

Yes

Yes

No

No

No

No

Yes

Yes

Page 33: Reliability Assessment of Historical Masonry Structures

17

3.1.2 Heritage

In the case of heritage structures, preserving existing material is of high value, implying the

application of less intervention. Also, it is important to use materials which are already there

instead of a new structure as a supporter. In brief, heritage criteria require the minimum destruction

be applied to the heritage value of existing materials and systems in any renovation and

strengthening process (Allen, 1991a).

3.1.3 Past Performance

Historical buildings have withstood all the loads applied during their service life to date which

provides some evidence as to the serviceability and safety of these structures. Based on different

factors including the age of a structure and the type of the applied loads, different information can

be obtained from its past performance. As an example, in the case of a building aging less than

100 years which has experienced earthquakes with ground motions less than the design event

during its service-life to date, one may conclude that the structure has presented satisfactory

performance regarding the dead and other variable loads but not potential seismic ones. In contrast,

a historical structure dating back a number of centuries, has demonstrated satisfactory performance

under a much bigger variety of loads and hazards (Allen,1991a). For example, the possible gradual

loss in load bearing capacity during past earthquakes may lead to alteration of dominant failure

modes and therefore difficulties in accurate structural evaluation. Structural integrity and the

ability to absorb local failure without extensive structural collapse are key properties of existing

masonry structures. However, these properties cannot easily be quantified and are mostly based

on engineering judgment.

3.1.4 Economics

In the design stage, the cost of providing a high degree of safety is low, so that reducing safety

factors in specific situations is not beneficial in terms of saving money. Therefore, it is more

suitable practically to use general criteria, presented by CSA/NBCC, which are conservatively

applicable to all situations (Diamantidis, 1999; Ellingwood, 1996). In the evaluation stage, the cost

Page 34: Reliability Assessment of Historical Masonry Structures

18

of not meeting a criterion and, consequently, the requirement of upgrading can be large. Therefore,

due to economic considerations, determination of criteria for each situation according to the

fundamental requirements of life safety, comfort, function, and economics is necessary for

upgrading (Allen, 1991a). Moreover, Maes et al. (1999) proved that implementation of materials

and arrangements specified by some current requirements are economical and feasible during the

primary construction phase but uneconomical and infeasible after construction.

3.1.5 Change in Design Practice

Historical structures were designed and constructed many years ago based on the engineering

principles, from rules of thumb and tradition in engineering experience, which were available that

time. Therefore, they are mostly designed for compression leading to vulnerability when subject

to the lateral loads or eccentricity of vertical loads.

3.1.6 Service Life

Historic structures need to be preserved for future generation. Therefore, a new level of

performance may be considered for protected structures by the preservation authority, which may

lead to a new service life higher than that specified in modern design codes. The new expected

service life has to be satisfied with minimum level of upgrading and disturbance. Therefore, there

is a need to evaluate historic structures accurately before upgrading.

The calculation of a probability of failure necessitates the definition of failure. To do so, limit

states have to be defined. Limit states can be applied to all parts of a structure from the entire

structure to a single member. Commonly, several limit states are considered for a single structure.

There are several known limit state categories being used in structural engineering (Maes et al,

1999):

Page 35: Reliability Assessment of Historical Masonry Structures

19

Ultimate limit state (ULS)

The ultimate limit state refers to load bearing capacity, overturning and overall stability, i.e., limit

state exceedance denotes failure of structure or member probably leading to serious injury or

human fatality. Therefore, ULS is based on the concept of safety.

Serviceability limit state (SLS)

Human life is not in danger through exceedance of a serviceability limit state. In general,

serviceability limit states refer to gradual deterioration, residents’ comfort or maintenance cost.

Regarding masonry structures, SLS refers to the control of cracking, the control of deflections and

inclinations, local damage, inappropriate function of non-bearing structural elements.

Phenomenological limit state (PLS)

PLS refers to different limit states including the value limit state (the influence of

engineering/architectural changes on historical/architectural value), the appearance limit state

(improper discoloration, corrosion, local crumbling and graffiti) and the sustainability limit state

(waste and recycling assessment). Regarding the limit state value, it should be noted that it may

be based on the summation of different values, including antiquity value, historical value, symbolic

value, heuristic value and so on (Baldioli, 1998).

The limit state function is determined based on the input variables and failure modes. Codes of practice

present formulae to calculate the load effects on a member as well as its capacity (using material

properties) under different failure modes. Therefore, when the desired limit state (e.g. ULS) is selected

and the applied loads on each structural element are determined, the probable failure modes of each

member should be estimated. The formulae of the structural resistance and the load effects associated

with each structural member under the determined probable failure modes can be derived from codes

of practice. Limit state functions associated with different failure modes of a member can be calculated

Page 36: Reliability Assessment of Historical Masonry Structures

20

through the subtraction of the recommended formulae for load effect from the formulae of the

resistance.

The term reliability refers to the complement of the probability of failure. Structural safety is a

function of the applied loads and the structural capacity of the members. Reliability analysis of a

structure involves the calculation of its probability of failure under the applied loads. The general

reliability problem can be determined starting from the primary reliability problem. In the primary

reliability problem, only load effects (𝐿) and structural resistance (𝑅) are considered, both as

random variables with known probability density functions (𝑓𝐿(𝑙) and 𝑓𝑅(𝑟)). The probability of

𝑅 − 𝐿 <0 equals the probability of failure (𝑃𝑓). In order to do reliability analysis, a procedure

should be followed:

I) The hazard and failure modes under the considered hazard have to be determined.

II) A limit state function involving the variables from stage I have to be determined in order to

calculate the probability of failure. In structural engineering, the formulation of limit state function

is commonly in the basic form as follows:

𝑔(𝑅, 𝐿) = 𝑅 − 𝐿 (3-1)

where 𝑅 is the resistance and 𝐿 is a load effect. The safe and failure performances can be

conventionally defined as:

𝑔(𝑅, 𝐿) < 0 → 𝐹𝑎𝑖𝑙𝑢𝑟𝑒 (3-2)

𝑔(𝑅, 𝐿) = 0 → 𝐶𝑟𝑖𝑡𝑖𝑐𝑎𝑙 𝑆𝑖𝑡𝑢𝑎𝑡𝑖𝑜𝑛 (𝐶𝑜𝑛𝑠𝑖𝑑𝑒𝑟𝑑 𝑆𝑎𝑓𝑒) (3-3)

𝑔(𝑅, 𝐿) > 0 → 𝑆𝑎𝑓𝑒 (3-4)

Page 37: Reliability Assessment of Historical Masonry Structures

21

III) The random variables (X) which influence the structural behaviour including structural

resistance (𝑅) and load effect (𝐿) have to be identified. Then, the probabilistic characteristics of

these random variables should be determined.

IV) The probability of failure (𝑃𝑓) or reliability index (𝛽) as a reliability measure are calculated.

In general, 𝑅 and 𝐿 are a function of time and as a result 𝑃𝑓 is also a function of time. But due to

the complexity of the mathematical calculation of 𝑃𝑓, a preset reference period (𝑡𝐿) (design service

life) can be considered for reliability analysis, leading to transformation of the probability density

functions into time-invariant probability density functions. However, considering 𝑡𝐿 is possible

but it is not necessary. The generalized reliability analysis is as follows (Schueremans &

Verstrynge, 2008):

𝑃𝑓 = 𝑃(𝑔(𝑅, 𝐿) < 0) = 𝑃(𝑅 < 𝐿) = ∬ 𝑓𝑅,𝐿(𝑟, 𝑙)𝑑𝑟𝑑𝑙

𝑟<𝑙

(3-5)

Consequently, the probability of survival (𝑅𝑓) is

𝑅𝑓 = 1 − 𝑃𝑓 (3-6)

In most cases 𝑅 (resistance) and 𝐿 (load effect) are independent having continuous densities

(𝑓𝑅(𝑟), 𝑓𝐿(𝑙)). Therefore, the probability of failure can be numerically computed where Equation

3-5 can be rewritten as a convolution integral (Brehm, 2011), see Figure 3-2.

𝑃𝑓 = ∬ 𝑓𝑅(𝑟). 𝑓𝐿(𝑙)𝑑𝑟𝑑𝑙 = ∫ 𝐹𝑅(𝑙)𝑓𝐿(𝑙)𝑑𝑙

+∞

−∞

+∞

−∞

(3-7)

where 𝐹𝑅 denotes the cumulative distribution function of the resistance 𝑅, 𝑓𝐿 denotes the

probability density function of load effects 𝐿.

Page 38: Reliability Assessment of Historical Masonry Structures

22

Figure 3-2 Load effect and resistance probability density functions.

Equation 3-7 can only be solved in closed form for a small number of cases. Special analysis

methods are required to solve the integral of Equation 3-7 in most cases. The most common

reliability analysis methods are briefly discussed in section 3-5.

V) Target failure probabilities (𝑝𝑓𝑇) or a target reliability index (𝛽𝑇) are determined. The safety

assessment of structures necessitates definition of target safety levels as decision criteria. Codes

of practice present different values as the target probability of failure, e.g. 𝑃𝑓𝑇 = 5. 10−4,

according to Eurocode (EN1990, 2002). Regarding historical structures, the necessity of meeting

the target probability of failure values reported by codes is still a controversial issue. This is due

to that fact that these values are reported for new structures, not historical ones with specific criteria

and requirements. Several studies have advocated widening the discussion to develop a more

accurate target probability of failure (Ditlevsen, 1982; EC1, 1994; ISO2394, 1998; Melchers,

1999; NEN 6700, 1991; Van Dyck, 1995). Other performance criteria are also considered in these

studies in addition to those mentioned in Section 3-1 regarding the assessment of historical

buildings. Some of these criteria and parameters include possible damage (fatality, environmental

damage, economic damage and socio-cultural damage), risk level (public buildings, bridges, off-

shore structures and so on) and warning level (steady failure with a visible clue and sudden failure

without a clue).

Target probability of failure values can be calculated according to proposed experimental formulas

(Melchers 1999; Van Dyck 1995). The following formula was proposed by the Construction

Page 39: Reliability Assessment of Historical Masonry Structures

23

Industry Research and Information Association (CIRIA, 1977) in order to calculate the probability

of failure considering a preset design service life (𝑡𝐿):

𝑃𝑓𝑇 = 10−4𝑆𝑐

𝑡𝐿

𝑛𝑝 (3-8)

where 𝑛𝑝 denotes the average number of building residents or the building’s immediate

neighbourhood residents. 𝑆𝑐 is the social criterion factor, with values shown in Table 3-1. Allen

(1991b) developed another empirical formula:

𝑃𝑓𝑇 = 10−5𝑡𝐿

√𝑛𝑝

𝐴𝑐

𝑊 (3-9)

in which 𝐴𝑐 is activity factor and W is warning factor, see Table 3-1. Both formula have

advantages. The first one includes 𝑆𝑐 accounting the importance of historical buildings or their

preservation value. The second one accounts for more differentiation. But none of them considered

the number of injuries and the economic cost of failure. A formula has been proposed to calculate

the nominal target failure probabilities for historical structures accounting for the cost factor and

reassigning the social factor (Schueremans, 2001):

𝑃𝑓𝑇 =10−4𝑆𝑐𝑡𝐿𝐴𝑐𝐶𝑓

𝑛𝑝𝑊 (3-10)

where 𝐶𝑓 denotes the cost factor according to (CEB, 1976, 1978), see Table 3-1. As this formula

includes all the above mentioned factors in addition to the 𝑆𝑐 which specifically accounts the level

of importance of historical structures, it seems to be more compatible with historic structures.

Therefore, it is recommended to use Equation 3-10 to estimate the target probability failure in case

of historic structures.

GruSiBau (1981) categorized failure consequences based on defined consequence classes. The

reliability index was determined for each class and consequence for a 50-year observation period,

see Table 3-2. Diamantidis (1999, 2001) also presented final tentative target reliability values as

Page 40: Reliability Assessment of Historical Masonry Structures

24

listed in Table 3-3. JCSS (2001) developed a different approach to account for risk to both human

life and investment. Through this approach, the target reliability links to the relative cost of

reliability enhancement of structures. Table 3-4 presents the required target reliability according

to JCSS (2001).

Table 3-1 Influential factors on the target probability of failure (Schueremans, 2001).

Residual Service Life/years (𝑡𝐿)

Number of Endangered Lives (𝑛𝑝)

Economic Factor (𝐶𝑓)

Not significant 10

Significant 1

Very significant 0.1

Warning Factor (W)

Fail-Safe condition 0.01

Steady failure with some warning clue 0.1

Steady failure hidden from view 0.3

Unexpected failure with no warning 1.0

Activity Factor (𝐴𝑐)

Post-disaster activity 0.3

Normal activities 1.0

- Building 1.0

- Bridges 3.0

- High exposure structures (offshore structures) 10.0

Social Criterion Factor (𝑆𝑐)

Historical structures of great importance (e.g. listed by UNESCO) 0.005

Historical structures listed as nationally important 0.05

Historical structures listed as regionally important 0.5

Not-listed historical structures 5.0

Page 41: Reliability Assessment of Historical Masonry Structures

25

Table 3-2 Target reliability and consequences class according to GruSiBau (1981).

Consequence Class Ultimate Limit State Serviceability Limit State

1

Safe for human life

Small economic influence

Small economic impact

Small interference with use

𝛽 = 4.2 𝛽 = 2.5

2

Dangerous for human life

Significant economic influence

Significant economic influence

Considerable interference with use

𝛽 = 4.7 𝛽 = 3.0

3

Very dangerous for human life

Large economic influence

Large economic influence

Large interference with use

𝛽 = 5.2 𝛽 = 3.5

Table 3-3 Tentative target reliability values 𝛽𝑇 (𝑃𝑓𝑇) (Diamantidis, 1999, 2001).

Costs of Safety

Measures

SLS

(Permanent)

ULS - Failure Consequences

Low Moderate Significant

High 1.0 (0.2) 2.8 (3. 10−3) 3.3 (5. 10−4) 3.8 (7. 10−5)

Moderate 1.5 (7. 10−2) 3.3 (5. 10−4) 3.8 (7. 10−5) 4.3 (8. 10−6)

Low 2.0 (2. 10−2) 3.8 (7. 10−5) 4.3 (8. 10−6) 4.8 (8. 10−7)

Page 42: Reliability Assessment of Historical Masonry Structures

26

Table 3-4 Required target reliability according to JCSS (2001) for a 50-year observation period.

Relative Cost

of Reliability

Enhancement

Failure Consequences

Minor Average Major

e.g. agricultural building e.g. residential building e.g. high-rise building

Large

𝛽 = 1.7 𝛽 = 2.0 𝛽 = 2.6

𝑃𝑓 ≈ 5. 10−2 𝑃𝑓 ≈ 3. 10−2 𝑃𝑓 ≈ 5. 10−3

Medium

𝛽 = 2.6 𝛽 = 3.2 𝛽 = 3.5

𝑃𝑓 ≈ 5. 10−3 𝑃𝑓 ≈ 7. 10−4 𝑃𝑓 ≈ 3. 10−4

Small

𝛽 = 3.2 𝛽 = 3.5 𝛽 = 3.8

𝑃𝑓 ≈ 7. 10−4 𝑃𝑓 ≈ 3. 10−4 𝑃𝑓 ≈ 10−5

Structural reliability techniques that have been developed can be classified into two general

categories including asymptotic techniques and simulation based approaches. A discussion on the

applicability of each technique in the estimation of structural reliability, as well as a brief overview

of their related limit state functions (explicit or implicit) is presented in the following sections.

3.5.1 Asymptotic Techniques

Early developments regarding asymptotic techniques were only focused on the functions of the

first moment (mean) and the second moment (variance). Numerous methods have been developed

based on second moment formulation using the reliability index (𝛽) proposed by Cornell (1969)

and Ang et al. (1975) including Mean Value First Order Moment (MVFOSM) (Cornell, 1969),

Generalized Safety Index (Hasofer & Lind, 1974), First Order Reliability Method (FORM)

(Rackwitz & Flessler, 1978) and Second Order Reliability Method (SORM) (Breitung, 1991).

Page 43: Reliability Assessment of Historical Masonry Structures

27

FORM and SORM are the most developed structural reliability techniques and are discussed

further in the following sections. First Order/Second Order Reliability in Original Space

(FOROS/SOROS) as a modified version of FORM and SORM are explained as well.

First Order Reliability Method (FORM)

The first order reliability method (FORM) is one of the reliable analytical methods which make

use of first and second moments of random variables. The name comes from the fact that FORM

linearizes the limit state function using the first-order of the Taylor series approximation.

Cornell (1969) proposed the original formulation of the FORM in which the limit state function is

linearized at the mean values of the random variables using the first-order Taylor series

approximation. In this formulation both 𝑅 and 𝐿 are assumed to be independent and normally

distributed. The limit state function (𝑔(𝑅, 𝐿)) is consequently normally distributed and therefore,

the moments can be calculated according to the following expressions:

𝑔(𝑅, 𝐿) = 𝑅 − 𝐿 (3-11)

𝑚𝑔 = 𝑚𝑅 − 𝑚𝐿 (3-12)

𝜎𝑔 = √𝜎𝑅2 + 𝜎𝐿

2 (3-13)

where 𝑚𝑔 is mean value of 𝑔(𝑅, 𝐿) and 𝜎𝑔 is standard deviation of 𝑔(𝑅, 𝐿). As 𝑔(𝑅, 𝐿) is normally

distributed, the probability of failure can be calculated using the formula of cumulative distribution

function of normal distribution (𝐹𝑔) as follows

𝑃𝑓 = 𝑃 (𝑔(𝑅, 𝐿) < 0) = 𝐹𝑔(0) = ɸ (𝑥 − 𝑚𝑔

𝜎𝑔) = ɸ (

0 − 𝑚𝑔

𝜎𝑔) = ɸ (

−𝑚𝑔

𝜎𝑔) (3-14)

Page 44: Reliability Assessment of Historical Masonry Structures

28

Cornell (1969) used the term in the parenthesis (ɸ (−𝑚𝑔

𝜎𝑔)) to define the reliability index as follows,

see Figure 3-3.

𝛽 =𝑚𝑔

𝜎𝑔= −ф−1(𝑃𝑓) (3-15)

𝛽 =1

𝐶𝑜𝑉𝑔 (3-16)

where ф−1 is the inverse of the cumulative distribution function of the variable with a standard

(Gaussian) normal distribution.

Figure 3-3 Limit state function and probability of failure (Glowienka, 2007).

The definition of reliability index by Cornell (1969) has several advantages. 𝛽 is independent from

the distribution type of 𝑔. The stochastic moments of the variables (mean 𝑚𝑖 and standard

deviation 𝜎𝑖) are only required to calculate the approximate structural reliability. However, the

Page 45: Reliability Assessment of Historical Masonry Structures

29

missing parameter in the calculation of reliability index, i.e. distribution type, can influence the

estimate of reliability significantly especially in the case of structural engineering problems

(Brehm, 2011). Madsen et al. (1986) reported the formulation of the limit state as the more

important shortcoming related to this formulation and also discussed the mathematical treatment

of the limit state function as the reason for the shortcoming.

A modified formulation was developed by Hasofer & Lind (1974) so that basic variables are

transformed into standard normal space at the design point on the limit state function. The

reliability index refers to the minimum distance between the origin of the transformed random

variable space to the most likely failure point (design point). Distribution type of 𝑔 is considered

using this advanced FORM.

In this method, an initial design point (𝑥𝑖∗) is determined by assuming values for n-1 of random

variables (𝑋𝑖) (usually mean values). By solving the limit state function (𝑔 = 0), the remaining

random variable is obtained to ensure the design point is located on the failure boundary. Reduced

variables are obtained as follows

𝑥𝑖′∗ =

𝑥𝑖∗ − 𝜇𝑋𝑖

𝜎𝑋𝑖

(3-17)

where 𝑥′∗ is the vector of the transformed random variable. The reliability index is the shortest

distance between the design point and the centre of the transformed random variables. An iterative

procedure is used to estimate the reliability index which is discussed in more detail in Nowak &

Collins (2000). The reliability index and probability of failure are obtained by:

𝛽 = √(𝑥′∗)𝑇(𝑥′∗) (3-18)

𝑃𝑓 = ɸ(−𝛽) (3-19)

As shown in Figure 3-4, in the computation of 𝑃𝑓, the limit state approximation is done linearly at

the design point. However, the original and/or transformed limit state has been proven to be

Page 46: Reliability Assessment of Historical Masonry Structures

30

nonlinear in the case of practical applications. Therefore, no exact reliability index value can be

calculated by FORM. SORM was introduced to overcome this imperfection of FORM.

Second Order Reliability Method (SORM)

Second order approximations are introduced by SORM regarding the nonlinear limit state, leading

to a more accurate reliability index, see Figure 3-4. In SORM, a quadric surface at the design point

is considered to approximate the failure surface (Breitung, 1984; Der Kiureghian et al., 1987). The

failure surface is also approximated by a set of tangent hyperplanes by Ditlevsen (1984) and

Madsen et al. (1986) during the same period. Analysis utilized asymptotic approximation as

proposed by Breitung (1984):

𝑃𝑓𝑆𝑂𝑅𝑀≈ ɸ(−𝛽) ∏(1 + 𝛽к𝑖)

−12⁄

𝑛−1

𝑖=1

(3-20)

where к𝒊 is the principal curvature of the limit state related to the minimum distance point, and 𝛽

denotes the reliability index obtained by FORM.

The reliability index determined through the procedure of SORM is expected to be more accurate

compared to that obtained by FORM. As the formulation of SORM is based on second-order partial

derivatives of the limit state function on the transformed space, SORM is more complicated than

FORM.

First Order/Second Order Reliability in Original Space (FOROS/SOROS)

Breitung (1994) and Geyikens (1993) proposed other asymptotic techniques, namely First

Order/Second Order Reliability in Original Space (FOROS/SOROS). Through these methods,

there is no need to transform random variables to a standard normal space. The analysis is done in

the original space of the random variables. In other words, in order to evaluate the probability of

failure in the original space, the full model does not need to be transformed to the normal space,

see Figure 3-5. There are some advantages with this method compared to both FORM and SORM.

Page 47: Reliability Assessment of Historical Masonry Structures

31

As an example, through application of either FORM or SORM, a simple linear limit state is

changed to nonlinear, probably leading to complexity and lack of accuracy. However, it remains

linear in either FOROS or SOROS. The probability of failure in these methods can be determined

as:

𝑃𝑓 =(2𝜋)

(𝑛−1)2⁄ 𝑓(𝑥∗)

‖𝛻𝑙(𝑥∗)‖√‖𝑑 𝑒𝑡 (𝑃∗𝑇(𝑥∗)𝐻𝑙(𝑥∗)𝑃∗(𝑥∗) − 𝑃(𝑥∗))‖

(3-21)

where 𝑥∗ and ∇𝑙(𝑥∗) denote the PLM (Point of Maximum Likelihood) and the gradient vector of

the log-likelihood, respectively. 𝑃(𝑥∗) is the projection matrix and 𝑃∗(𝑥∗) is density matrix minus

the projection matrix. 𝐻𝑙(𝑥∗) refers to the Hessian matrix of likelihood. T is the transpose and

matrix determinant is shown by det.

3.5.2 Simulation Based Techniques

Enrico Fermi (1930’s) used Monte Carlo Simulation (MCS) in order to calculate neutron diffusion

but he did not publish any studies related to it. MCS is executed in a way in which joint probability

density functions are used to generate a number of independent samples (N) of the vector of

random variables (X). The limit state function (g(X)) of each sample (𝑥𝑖) is then evaluated, and

the probability of failure is calculated as:

𝑃𝑓𝑠𝑖𝑚𝑢𝑙𝑎𝑡𝑖𝑜𝑛≈

𝑁𝑓

𝑁 (3-22)

where 𝑁 and 𝑁𝑓 denote the total number of samples and the total number of failures (performance

function less than zero), respectively. The major deficiency of MCS is related to the large number

of analyses required to obtain accurate results; i.e. even in the case of simple problems, a large

number of simulations have to be executed. Importance Sampling (IS) (Harbitz, 1986) and

Adaptive Sampling (Karamchandani et al., 1989) are examples of other approaches that have been

Page 48: Reliability Assessment of Historical Masonry Structures

32

proposed based on simulation. In the case of both explicit and implicit limit state functions,

simulation-based approaches can be utilized. However, a particular algorithm needs to be

employed to set the simulation method and the FEM regarding the implicit limit state functions.

Simulation-based approaches may lose their practicality in cases of time consuming structural

analysis such as nonlinear analysis, dynamic analysis and complex systems of structures.

Page 49: Reliability Assessment of Historical Masonry Structures

33

Figure 3-4 Standard normal space related to FORM and SORM (Hassanien, 2006).

Figure 3-5 FOROS illustration (Hassanien, 2006).

Page 50: Reliability Assessment of Historical Masonry Structures

34

3.5.3 Response Surface Method

A relationship between a collection of input variables and the output response of a system is

designed to be found by a set of statistical and mathematical approaches, defined as the Response

Surface Method (RSM) (Myers & Montgomery, 1995; Khuri & Cornell, 1996). The main goal of

the application of RSM in structural reliability analysis is defined as the approximation of the

original implicit limit state function (Faravelli, 1989; Rajashekhar et. al., 1993; Abdelatif et al.,

1999). The Lagrange interpolation method and the regression method, developed by Abdelatif et

al. (1998) are among the most important response surface methods utilized for solving reliability

problems. The collection process of input data in addition to the design of experiments are factors

which RSM mainly depends on. The main concern regarding the design of experiments is the level

of accuracy in locating the points in the failure point vicinity, given that the failure point is not yet

determined.

3.5.4 Artificial Neural Networks (ANNs)

In structural reliability, limit state functions can be approximated by Artificial Neural Networks.

The application of ANNs in structural reliability has only been investigated slightly (Hurdato et

al., 2001; Gomes, et al., 2004; Deng et al., 2005). An ANN is commonly established based on

groups called layers. There are two typical layers in this model: an input layer and an output layer.

The input layer is responsible for providing data to the network (Murotsu, 1993) and the output

layer is responsible for obtaining the network response to a specific input. To estimate the weights

of the input variables related to the output response, a learning process is considered.

Structural analysis and reliability methods need to be appropriately integrated in order to enable

reliability assessment in real situations. Integrated software developments have been achieved

through several research programs. Table 3-5 presents the list of suitable combinations among

finite element structural analysis, reliability analysis methods and reliability software

(Schueremans and Verstrynge, 2008).

Page 51: Reliability Assessment of Historical Masonry Structures

35

Table 3-5 Integration of software for structural reliability analysis.

Software Tool Structural

Analysis (FEM)

Reliability Analysis Method

MC FORM/SORM Use of RS

NASREL NASCOM

COMREL -

SYSREL -

PERMAS-RA PERMAS

PROBAB DIANA

NESSUS ABAQUS/ANSYS

SBRA Included

OPTIMUS NASTRAN

Deterioration refers to physicochemical processes having environmental sources and introduces

the concept of time dependency into reliability analysis due to its gradual influence on the

behaviour and the appearance of buildings. Deterioration results in a reduction in the defined

standard level of performance (Franke & Schuman, 1998).

One or both groups of basic variables (resistance (𝑅) and load effect (𝐿)) can be influenced by

deterioration as well as the analysis model. In other words, deterioration may change masonry

resistance, but also may alter the loading conditions or might affect the analysis model. To be able

to describe the influence of deterioration on reliability assessment fully, all these influences on one

or more of the basic variables in addition to the analysis model need to be examined. Deterioration

processes impacting material strength (and consequently, structural limit states) and those not

directly impacting structural limit states (but influencing the phenomenological limit states) cannot

be clearly differentiated. For example, as residual strength is influenced by freeze-thaw cycles,

Page 52: Reliability Assessment of Historical Masonry Structures

36

they can be placed in the first category. On the other hand, building appearance is also influenced

by spalling (Maes, 1999).

Existing masonry structures, especially the ones with a service life of more than 50 years, which

is proposed as the standard design service-life, experience many deterioration processes (Maes,

1999). The most well-known deterioration processes related to masonry structures are presented

in Table 3-6. Frank & Schuman (1998) provided a more detailed overview in this regard.

Table 3-6 Deterioration mechanisms (Maes et al., 1999).

Deterioration Type Cause Consequences Imposed to

Freeze-thaw cycles -Temperature change

- Moisture content - Spalling

- Section resistance

- Appearance

Chemicals

(e.g. sulfates,

chemical spill,

acids, road salt)

- Diffusion

- Absorption,

- Capillary suction

- Moisture content

- Concentration difference

- Chemical reaction

- Disintegration

- Disfiguration

- Efflorescence

- Salt crystallization

- Discoloration

- Material resistance

- Appearance

Abrasion - Particulate - Material Erosion - Material resistance

- Appearance

Pollution

(𝐶𝑂2, 𝑂2, 𝑁𝑂𝑥) - Chemical agents

- Disintegration

- Discoloration

- Material resistance

- Appearance

Solar radiation - UV, IR - Disintegration

- Discoloration

- Material resistance

- Appearance

Biological and

microbiological

agents

- Fungi

- Bird droppings

- Biological deposits

- Vegetation or plants

- Disintegration

- Spalling

- Material resistance

- Appearance

Page 53: Reliability Assessment of Historical Masonry Structures

37

3.7.1 Modeling of Deterioration

In order to integrate deterioration into reliability analysis, one or more of the basic variables related

to either resistance or load effects, analysis model uncertainty (𝐴) or limit state model uncertainty

(𝜃) have to account for their time-dependency, see Table 3-7. Mori et al. (1993) proposed several

possible degradation functions presented in Table 3-8.

Table 3-7 Deterioration modeling (Maes et al., 1999).

Basic Variables Time Dependency

Load effects (𝐿) 𝐿(𝑡) = 𝑆0. 𝑓𝑑𝑒𝑔,𝐿(𝑡)

Resistance (𝑅) 𝑅(𝑡) = 𝑅0. 𝑓𝑑𝑒𝑔,𝑅(𝑡)

Analysis model parameters (A) 𝐴(𝑡) = 𝐴0. 𝑓𝑑𝑒𝑔,𝐴(𝑡)

LS model uncertainty (𝜃) 𝜃(𝑡) = 𝜃0. 𝑓𝑑𝑒𝑔,𝜃(𝑡)

𝐿0, 𝑅0, 𝐴0, 𝜃0 are random variables reported at a reference time

Table 3-8 Degradation functions (Mori et al., 1993).

Model Formula

Linear 𝑓𝑑𝑒𝑔(𝑡) = 1 − 𝑐. 𝑡

Parabolic 𝑓𝑑𝑒𝑔(𝑡) = 1 − 𝑐. 𝑡2

Square Root 𝑓𝑑𝑒𝑔(𝑡) = 1 − 𝑐. √𝑡

c is a constant value

Page 54: Reliability Assessment of Historical Masonry Structures

38

Lack of data is one of the most important problems in assessing historical structures. This

necessitates finding methods to overcome this shortcoming. Available prior information obtained

by either expert opinion or testing results can be used to increase the sample size. This enlargement

can consequently result in better estimation of the probabilistic models of materials. Bayes (1763)

developed Bayes’ theorem as a data updating method which is discussed briefly as follows.

In this method, the stochastic distribution of the prior parameters can be updated using new data

(e.g. test data). The same type of distribution is usually considered for both the prior and posterior

distributions. The distribution of the compressive strength of a masonry unit before and after the

updates besides the associated likelihood distribution was illustrated by Glowienka (2007), see

Figure 3-6.

Figure 3-6 Example of the updating compressive strength of a masonry unit using Bayes’

theorem (Glowienka, 2007).

Bayes’ theorem is expressed as the following relationship

𝑓𝜃(𝜃|𝑥) =𝑓𝑋(𝑥|𝜃). 𝑓𝜃(𝜃)

𝑓𝑋(𝑥)=

𝑓𝑋(𝑥|𝜃). 𝑓𝜃(𝜃)

∫ 𝑓𝑋(𝑥|𝜃). 𝑓𝜃(𝜃)𝑑𝜃 (3-23)

Page 55: Reliability Assessment of Historical Masonry Structures

39

where 𝑓𝑋(𝑥|𝜃) is the probability density function of a random variable dependent on 𝜃. 𝑓𝜃(𝜃) is

the prior probability distribution function of the corresponding vector of parameters and 𝑓𝜃(𝜃|𝑥)

is the posterior probability distribution function of the vector of parameters. 𝑓𝑋(𝑥|𝜃) can be

estimated using the likelihood distribution of measured data. In the case of uncorrelated data,

𝑓𝑋(𝑥|𝜃) can be derived as follows

𝑓𝑋(𝑥|𝜃) = 𝑓𝑋(𝑥1|𝜃). … . 𝑓𝑋(𝑥𝑛|𝜃) = ∏ 𝑓𝑋(𝑥𝑖|𝜃) = 𝐿(𝜃|𝑥) 𝑛𝑖=1 (3-24)

The integral in Equation 3-23 sets the area under the probability distribution function to one, in

order to convert 𝑓𝜃(𝜃|𝑥) to a true probability of failure function. Therefore, the integral can be

considered as a constant coefficient. The posterior probability distribution function is commonly

formulated as

𝑓𝜃(𝜃|𝑥) = 𝑘 . 𝐿(𝜃|𝑥) . 𝑓𝜃(𝜃) (3-25)

The possibility of continuous application is one of the most important advantages of Bayes’

theorem as the updating can be done as often as required. For the next updating cycle, the estimated

posterior distribution serves as the prior one. A larger sample size would lead to more realistic

approximation of the distribution of 𝑋. However, updating will become inefficient with

enlargement of the sample size. Therefore, an evaluation is needed beforehand to determine if the

updating will be effective.

Considering all of the concepts mentioned above, the process of reliability analysis can be

summarized in a flowchart as shown in Figure 3-7.

Page 56: Reliability Assessment of Historical Masonry Structures

40

Figure 3-7 Reliability assessment framework.

Limit State Function

- Resistance (R)

- Load Effect (L)

Random Variables

- Asymptotic Techniques

- Simulation-based Techniques

- Response Surface Method

- Artificial Neural Networks

(ANNs)

Reliability Analysis

𝛽 & 𝑃𝑓

𝛽 > 𝛽𝑇

𝑃𝑓 < 𝑃𝑓𝑇

Deterioration

𝛽 < 𝛽𝑇

𝑃𝑓 > 𝑃𝑓𝑇

Reliable Non-reliable

Page 57: Reliability Assessment of Historical Masonry Structures

41

Reliability assessment of existing structures subjected to seismic loading is one of the most

challenging and complex assessments involving different aspects and concepts in the field of

structural engineering. In addition to the non-linear behaviour of the structure, there are

uncertainties in the nature of possible ground motions. Due to the complexity, new influential

parameters and almost different reliability assessment procedures, reliability assessment under

seismic loading is discussed separately in this section. Shah & Dong (1984) developed a method

to assess the reliability of existing structures subjected to probabilistic earthquake loadings. Based

on their studies, Monte-Carlo simulation can be used to obtain the distribution of the maximum

response, given the characteristics of an earthquake of strong motion. The procedure to do so is:

1) A unit power spectrum (Sf) and time envelope function (t) are constructed based on the

predominant frequency (fp) and duration (TD).

2) Monte-Carlo simulation is used to generate a set of K time histories (𝑎𝑘(t), k = 1, 2,…, K).

3) A desired peak acceleration (A = a) is selected and 𝑎𝑘(𝑡) is scaled using this acceleration.

4) A structural model is built and hysteretic characteristics selected.

5) An inelastic dynamic analysis is done for each sample 𝑎𝑘(𝑡) to estimate the maximum response

of the 𝑖th critical value for the 𝑗th member (𝑋𝑖𝑗,𝑘) for each time history (k).

6) Data (𝑋𝑖𝑗,𝑘(𝑘 = 1, 2, … , 𝐾)) are fitted to a probability distribution function. In other words, the

cumulative distribution function being conditional on 𝑎 (𝐹𝑋𝑖𝑗(𝑥𝑖𝑗/𝑎)) for 𝑖th critical variable and

the 𝑗th member (𝑋𝑖𝑗) is calculated in a way to fit the obtained data appropriately. The Gumbel

Extreme Value distribution is proved to fit the data well:

𝐹𝑋𝑖𝑗(𝑥𝑖𝑗|𝑎) = 𝑒𝑥𝑝 {−exp (−

𝑥𝑖𝑗 − 𝑏𝑖𝑗(𝑎)

𝑐𝑖𝑗(𝑎))} (3-26)

Given A = a, 𝑏𝑖𝑗(𝑎) and 𝑐𝑖𝑗(𝑎) are estimated by data fitting to the distribution. Given the capacity

of 𝑗th member for 𝑖th critical value (𝑦𝑖𝑗), the conditional reliability (𝑅𝑖𝑗) can be calculated by:

Rij = Reliability of the jth member regarding the ith critical value

Page 58: Reliability Assessment of Historical Masonry Structures

42

𝑅𝑖𝑗 = 𝑃 [𝑋𝑖𝑗 < 𝑦𝑖𝑗] = 𝐹𝑋𝑖𝑗(𝑦𝑖𝑗) = 𝑒𝑥𝑝 {−𝑒𝑥𝑝 (−

𝑦𝑖𝑗 − 𝑏𝑖𝑗(𝑎)

𝑐𝑖𝑗(𝑎))} (3-27)

Considering the values obtained related to 𝑅𝑖𝑗, failure is anticipated to occur in members with low

reliability. The boundary of structural reliability is:

min(𝑅𝑗) ≥ 𝑅 ≥ ∏ 𝑅𝑗

𝑗

(3-28)

As there is only a slight difference between the two bounds, as the large inelastic deformation

generally concentrates on only weak members of a structure subjected to seismic loads, only the

lower bound is used as structural reliability:

𝑅𝑇𝑜𝑡𝑎𝑙 = ∏ 𝑅𝑖𝑗

𝑖𝑗

(3-29)

It should be considered that the reliability estimation given in the previous steps is conditional on

the PGA (A = a) and the capacity (𝑌𝑖𝑗 = 𝑦𝑖𝑗). The final probability of failure can be calculated by

knowing the probability distribution function of A and 𝑌𝑖, as follows. Two sequential events with

the same characteristics (A, fp and TD) show maximum load effect and demand independently.

Therefore, the cumulative demand distribution in n-year life time can be expressed as:

𝐹𝑋𝑖,𝑛= [𝐹𝑋𝑖

(𝑥𝑖)]𝑛 (3-30)

The seismic demand is considered only as a function of the peak ground acceleration A, in almost

all applied seismic hazard analysis. In other words, determinist estimation of 𝑓𝑝 and 𝑇𝐷 are

expressed by the local soil conditions and past experience. Therefore, in Equation 3-31 we have:

Page 59: Reliability Assessment of Historical Masonry Structures

43

𝐹𝑋𝑖(𝑥𝑖) = ∫ 𝐹𝑋𝑖

(𝑥𝑖|𝑎) 𝑓𝐴(𝑎) 𝑑𝑎 (3-31)

where 𝑓𝐴(𝑎) denotes the probability distribution function of A during any one year. The reliability

of 𝑗th member in a lifetime of n years is:

𝑅𝑗 = ∫ 𝐹𝑋𝑖,𝑛(𝑧) 𝑓𝑌𝑖

(𝑧) 𝑑𝑧 (3-32)

in which 𝑓𝑌𝑖(𝑧) is the density function of the capacity of a critical variable.

It is obvious that identification of weak members by this method and their strengthening would

lead to improvement of the overall reliability of the structure. As either of the loading parameters

including intensity, peak acceleration, duration or frequency, cannot effectively be the

representative of total load effects or responses, estimation of structural responses or load effects

encounter important uncertainties. Uncertainties stem from three different sources including:

I) Seismic loads

II) Structural capacity

III) Load effects due to inadequate representation of load and consequently, calculation of

load effects

According to Shah & Dong (1984), the uncertainty which dominates the overall uncertainty in

final reliability analysis is the uncertainty in the ground motion. Therefore, to achieve more

accurate reliability analysis results in the case of structures subjected to seismic loads, estimation

of the loads or evaluation of the seismic hazard for the site should be done with great care.

Page 60: Reliability Assessment of Historical Masonry Structures

44

Chapter IV: Stochastic Modeling of Applied Load

Structural members have to be designed to be able to carry their applied loads during their service

life. The term “load” denotes the actions leading to stress development in the members. Moreover,

effects that may be influential on the capacity of a member are also considered as loads (e.g.

corrosion). In this thesis, the term loads refer to the applied forces on members (e.g. dead load or

snow load). Effects like corrosion are not considered.

As explained in Chapter 3, in order to calculate reliability of a structure, limit state functions have

to be estimated. Limit state functions take into account both loads and resistance. Therefore, to be

able to calculate an accurate probability of failure and reliability index, stochastic modeling of the

applied loads needs to be as realistic as stochastic modeling of the resistance.

In this chapter, stochastic models of the most common applied loads on historical structures are

presented.

Generally, loads vary over time and space and, therefore, can be expressed as random variables. It

is complex and not efficient to assess the load stochastically for every design case. Therefore,

general stochastic models for different loads have been derived to be used in the estimation of the

probability of failure and the reliability index. Load actions have been modeled based on the kind

of the load. There are four categories commonly used to classify loads:

Dead Load: dead loads usually refer to the self-weight of the structure. They apply permanently

on a member and therefore, there is minor scatter around the mean - the variation is low.

Live Load: live loads refer to the loads from occupancy. These kinds of loads are time dependent

and vary significantly with time leading to higher coefficients of variation compared to other types

of loads.

Wind and Snow loads: these loads are also time dependent and change considerably with time.

Page 61: Reliability Assessment of Historical Masonry Structures

45

Accidental loads: accidental loads refer to those with large quantity but short duration of

occurrence. The probability of occurrence of these kinds of loads is small during the observation

period. Earthquakes, explosions and avalanches are typical examples of accidental loads.

Dead load refers to the weight of structural members and their related components in addition to

the partitions and permanent equipment. Typically, dead load contributes about 70% of the total

load for ordinary residential structures and offices, which makes dead loads influential on the

evaluation of structural behaviours. Dead load has been shown to be less influential on the

probability of failure of concentrically compressed masonry members because of the small scatter

and sensitivity value. Shear failure occurs mostly under minimum axial load. Therefore, shear

walls are expected to be more sensitive to dead load and accordingly dead load is required to be

assessed in more detail. Dead loads are assumed to be constant over the service life. However, this

assumption is uncertain. Dimensional tolerances and uncertainty in the unit weight of materials

are the main causes of dead load uncertainty. In addition, the process of converting the load into

load effects leads to uncertainty. This process can be completed accurately in case of indeterminate

structures only if the sequence of construction is known.

Bartlett et al. (2003) presented a summary of the statistical parameters for loads which have been

used to calibrate the loads and load combination criteria for the 2005 National Building Code of

Canada (NBCC). Based on their literature review, the bias coefficient (the ratio of mean to nominal

values) and the coefficient of variation (CoV) (the ratio of standard deviation to mean value) of

dead load has been reported to be from 1.00 to 1.05 and 0.06 to 0.09, respectively. CoVs of

modeling and analysis which are usually assumed to be unbiased, are reported to be in the range

of 0.03 to 0.07. Therefore, the CoV of the dead load effect increases to a range of 0.05 to 0.10. A

bias of 1.05 and a CoV of 0.10 (reported by Ellingwood et al., 1980) were adopted by different

organizations (Standards Association of Australia, 1985, South African Bureau of Standards, 1989

and European Committee for Standardization, all reported by Kemp et al., 1998; Tabsh, 1997;

Ellingwood, 1999), see Table 4-1. In the case of counteraction of the dead load effects with the

effects of other loads, a normal distribution with a bias of 1.00 and a CoV of 0.10 was assumed

for dead load.

Page 62: Reliability Assessment of Historical Masonry Structures

46

They (Bartlett et al., 2003) also expressed that there was a common assumption of normal

distribution of dead loads, perhaps due to the tendency of tolerances to be distributed normally.

However, there are no actual data available to verify this assumption. To some extent, the statistical

characteristics of dead loads depend on the type of construction materials and the size of the

structure. In other words, in the case of large structural components, weight is not very sensitive

to absolute dimensional tolerances which results in a reduction in dead load bias coefficients and

coefficients of variation. All the above-mentioned studies focused on steel and concrete structures.

Therefore, accordance of the proposed statistical parameters with masonry structures, especially

historic ones, is in doubt.

Brehm (2011) determined the stochastic model appropriate for the dead load of a masonry bracing

wall. He expressed that determination of dead load is based on the density of material and the

volume of the member. The density of a material varies over the member (e.g. concentration of

aggregate). However, members are usually considered to be homogeneous. Therefore, the

influence of the scatter of material density on dead load is considered to be negligible and

consequently dead load is assumed to distribute uniformly over a member. Mortar density is

assumed to be similar to the masonry unit density. Therefore, mortar joints, even the thicker one

(general purpose mortar, GPM), were expressed to have negligible effect on the scatter of dead

load. As the research was based on new masonry units produced in factories and plants, the scatter

of unit dimension was also considered to be negligible. However, the volume of a masonry wall

was assumed to be variable due to workmanship and provided a coefficient of variation regarding

geometry. Brehm (2011) also introduced the self-weight of the concrete slabs as the most

influential parameter on self-weight of the structure and reported influence of the coefficients of

variation of the concrete density and concrete slab thicknesses on the scatter of the slab self-weight.

He suggested a bias factor of 1.0 and coefficient of variation of 0.06 for dead load with normal

distribution (log-normal distribution in some cases), see Table 4-1.

The Joint Committee on Structural Safety (JCSS) probabilistic model code (2001) assumes self-

weight as a time independent variable with a probability of occurrence at an arbitrary point-in-

time almost equal to one. The uncertainties in the magnitude of self-weight are described as small

compared to other kinds of loads. A Gaussian distribution is suggested to be considered for the

weight density and dimensions of a structural member. To make the calculation of self-weight

Page 63: Reliability Assessment of Historical Masonry Structures

47

simple, a Gaussian distribution is also assumed as the distribution type of self-weight. The JCSS

code presented the CoV of masonry weight density equal to 5% and mean values and standard

deviations for deviation of cross-section dimensions of masonry members from their nominal

values as 0.02 and 0.04 for unplastered and 0.02 and 0.02 for plastered, respectively. JCSS states

their reported values are derived from a large population from various sources.

Table 4-1 Summary of statistical parameters for dead load.

References Bias COV Distribution Type

Ellingwood et al. (1980) 1.05 0.1 -

Brehm (2011) 1.0 0.06 Normal/Log-normal

Bartlett et al. (2003) 1.0 0.10 Normal

The load associated with use and building occupancy, including the weight of people, equipment,

furniture and materials in storage, is referred to as live load. Live load varies over time and

location. Total live load is the sum of contributions categorized as permanent live load and

transient live load (Ferry Borges & Castanheta, 1971). Permanent live load is the component of

live load remaining constant for a period of time (usually the duration of the tenancy) but still

removable, e. g. furnishings, while transient live load is associated with a short or instantaneous

duration, e. g. group of people. The total live load is sum of these two components, see Figure 4-

1. Due to the large variation associated with live load, especially as a result of spatial variation, it

is complex to determine a stochastic model for the live load. Presentation of a detailed description

regarding the theory of live load modeling is beyond of scope of this thesis.

Page 64: Reliability Assessment of Historical Masonry Structures

48

Figure 4-1 Representation of live load (Brehm, 2011).

A simplified approach to model live load probabilistically is (McGuire & Cornell, 1974; Pier &

Cornell, 1973)

𝑊(𝑥, 𝑦) = 𝑌 + є (𝑥, 𝑦) (4-1)

where W refers to load intensity (psf) applied on a floor, Y is a random variable modeling the

mean of load on the floor and є (𝑥, 𝑦) is a stochastic process with zero mean associated with the

deviations from the average. CIB (1989) outlined a similar model regarding live load modeling as

follows:

𝑊(𝑥, 𝑦) = 𝑚 + 𝑉 + 𝑈(𝑥, 𝑦)

(4-2)

where 𝑚 is the deterministic mean of live load depending on the usage type of building (e. g.

residential, office and so on). 𝑉 represents a random variable (with mean zero) accounting for the

Page 65: Reliability Assessment of Historical Masonry Structures

49

variation of the load associated with two independent areas on the same floor or between floors

(A1 and A2). 𝑈(𝑥, 𝑦) stands for a random field concerning the spatial variation related to the load.

It is too sophisticated to determine the structural response by evaluating of the increment of every

single load based on Eq. 4-2. Therefore, a methodology was introduced by which an equivalent

load with uniform distribution (q) leading to the same load effect is determined as a preferable

alternative solution.

4.3.1 Total Live Load

Ellingwood & Culver (1977) presented mean and variance of unit load (U) as

𝐸 [𝑈] = µ (4-3)

Var [U] = 𝜎2 +1

𝐴∫ ∫ ∫ ∫ 𝐶𝑜𝑉 [є (𝑥, 𝑦), є (𝑢, 𝑣)] 𝑑𝑥𝑑𝑦𝑑𝑢𝑑𝑣 … (4-4)

where µ is mean of all unit floor loads associated with office buildings and 𝜎2 is variance in

individual floor.

They then simplified Eq. 4-4 as

Var [U] = 𝜎2 +𝜎𝑠

2

𝐴 (4-5)

in which 𝜎𝑠 denotes an experimental constant.

The mean and variance of the equivalent load with uniform distribution leading to same effect

(𝑞) are

E [q] = µ (4-6)

Page 66: Reliability Assessment of Historical Masonry Structures

50

𝑉𝑎𝑟 [𝑞] = 𝜎2 +𝜎𝑠

2

𝐴𝜅 (4-7)

𝜅 =∬ 𝐼2(𝑥, 𝑦) 𝑑𝑥𝑑𝑦

𝐴

[∬ 𝐼(𝑥, 𝑦) 𝑑𝑥𝑑𝑦𝐴

]2 (4-8)

In which 𝐼(𝑥, 𝑦) is the influence function for the response quality sought. A is the influence area

over which I(x,y) assumes nonzero values. Considering multi-storey column loads, there is a slight

difference in the first term of Eq. 4-1 from that of floor loads (McGuire & Cornell, 1974). In

practice, the difference is negligible and may be ignored.

Ellingwood & Culver (1977) also determined the mean and variance of the maximum sustained

uniformly distributed load (𝐿𝑚𝑎𝑥 ) based on the National Bureau of Standards (NBS) load survey

results (office buildings) in the U.S.

𝐸[𝐿𝑚𝑎𝑥] = 0.924 + 1.853/√𝐴 𝑘𝑁/𝑚2 (4-9)

𝑉𝑎𝑟[𝐿𝑚𝑎𝑥] = 0.033 + 4.024/𝐴 𝑘𝑁/𝑚4 (4-10)

A similar analysis done earlier by McGuire & Cornell (1974) using U.K. data resulted in

𝐸[𝐿𝑚𝑎𝑥] = 0.867 + 2.101/√𝐴 𝑘𝑁/𝑚2 (4-11)

𝑉𝑎𝑟[𝐿𝑚𝑎𝑥] = 0.026 + 3.194/𝐴 𝑘𝑁/𝑚4 (4-12)

It has been shown that there is close agreement between 𝐸[𝐿𝑚𝑎𝑥] values based on the NBS and

U.K. survey results (Ellingwood & Culver, 1977). Regarding the maximum total equivalent

Page 67: Reliability Assessment of Historical Masonry Structures

51

uniformly distributed load (q), the mean and variance were determined using the NBS survey

results (Ellingwood & Culver, 1977) as

𝐸[𝑞] = 0.895 + 7.588/√𝐴 𝑘𝑁/𝑚2 (4-13)

𝑉𝑎𝑟[𝑞] = 0.032 + 4.024/𝐴 𝑘𝑁/𝑚4 (4-14)

And the analysis of McGuire & Cornell (1974) using U.K. data resulted in

𝐸[𝑞] = 0.713 + 11.135/√𝐴 𝑘𝑁/𝑚2 (4-15)

𝑉𝑎𝑟[𝑞] = 0.026 + 3.194/𝐴 𝑘𝑁/𝑚4 (4-16)

Ellingwood & Culver (1977) also showed that there is a significant difference between the mean

values of the two expressions in terms of small influence areas. However, the difference is not

related to the difference in the NBS and U.K. load survey results. It is attributed to the parameters

which are used in the extraordinary load model.

JCSS (2001) based its suggested procedure regarding the determination of the stochastic

characteristics of permanent live load on the formula derived by Rackwitz (1995) who generalized

the stochastic moments of live load (similar to Ellingwood & Culver, 1977) as follows

𝐸[𝑞] = 𝐸[𝑊(𝑥, 𝑦)] = 𝑚𝑞 (4-17)

𝑉𝐴𝑅[𝑞] ≈ 𝜎𝑉2 + 𝜎𝑈

2 ∙ 𝐴0

𝐴 ∙ 𝜅 (4-18)

Page 68: Reliability Assessment of Historical Masonry Structures

52

𝜅 ≈ ∫ 𝑖2(𝑥, 𝑦)𝑑𝐴

(∫ 𝑖(𝑥, 𝑦)𝑑𝐴)2 (4-19)

where 𝑚𝑞 refers to the mean of the live load. 𝜎𝑉 and 𝜎𝑈 are the standard deviations of V and U,

respectively. A refers to the effective area and 𝐴0 is the reference area of the load measure. 𝜅 is

defined based on the required stress resultants and the structural system. Values of 𝜅 are

recommended by Melchers (1999), JCSS (2001) and Hausmann (2007) (reported by Brehm

(2011)).

Table 4-2 Parameters related to probabilistic modeling of live load (JCSS, 2001).

Type of

Use

Ref.

area Permanent Load Transient Load

𝐴0

(𝑚2)

𝑚𝑝𝑒𝑟𝑚

(𝑘𝑁/𝑚2)

𝜎𝑉 (𝑘𝑁/𝑚2)

𝜎𝑈 (𝑘𝑁/𝑚2)

1/λ

(a) 𝑚𝑡𝑟𝑎𝑛

(𝑘𝑁/𝑚2) 𝜎𝑈

(𝑘𝑁/𝑚2)

1/v

(a)

dp

(d)

Office 20 0.5 0.30 0.60 5 0.20 0.40 0.3 1-3

Lobby 20 0.2 0.15 0.30 10 0.40 0.60 1.0 1-3

Residential 20 0.3 0.15 0.30 7 0.30 0.40 1.0 1-3

Hotel 20 0.3 0.05 0.10 10 0.20 0.40 0.1 1-3

Hospital 20 0.4 0.30 0.60 5-10 0.20 0.40 1.0 1-3

Laboratory 20 0.7 0.40 0.80 5-10 - - - -

Library 20 1.7 0.50 1.00 >10 - - - -

Classroom 100 0.6 0.15 0.40 >10 0.50 1.40 0.3 1-5

Sales room 100 0.9 0.60 1.60 1-5 0.40 1.10 1.1 1-14

Industrial

Light 100 1.0 1.00 2.8 5-10 - - - -

Heavy 100 3.0 1.50 4.10 5-10 - - - - aload fluctuation rate baverage load duration

JCSS (2001) suggested a similar method in order to calculate the transient live loads. As a result

of the large scatter of transient live loads, a stochastic field is used to model them. Therefore 𝜎𝑉 is

set to be zero. The importance of transient live load is more obvious in the case of small values of

the influence area (A) such as balconies or stairs. A summary of required parameters for the live

load model based on JCSS (2001) is presented in Table 4-2 where dp is duration of transient load

(year), λ is occurrence rate of permanent live load changes in [1/year] and v is occurrence rate of

transient live load changes in [1/year].

Page 69: Reliability Assessment of Historical Masonry Structures

53

Bartlett et al. (2003) considered Eq. (4-20) derived by simulation results (Ellingwood and Culver,

1977) to calculate the expected value of maximum equivalent uniformly distributed live load

during a 50-year reference period (�̅�𝑚𝑎𝑥 (𝑘𝑁/𝑚2)).

�̅�𝑚𝑎𝑥 = 0.895 + 7.6/√𝐴𝐼 (4-20)

in which 𝐴𝐼 denotes the influence area (𝑚2). NBCC (1999) presented the corresponding nominal

uniformly distributed live load (𝐿 (𝑘𝑁/𝑚2)) as follows

𝐿 = 2.4 (0.3 + √9.8/𝐵) (4-21)

where 2.4 refers to the specified live load (𝑘𝑁/𝑚2) and B is the tributary area (𝑚2). ASCE7- 98

(ASCE 2000) considered the influence area to be equal to the tributary area regarding two-way

slabs and twice and four times of the tributary area for interior beams and columns, respectively.

Based on the above-mentioned formula, the bias can be calculated easily by dividing Eq. (4-20)

by Eq. (4-21) depending on the element type and its related influence area (Bartlett et al., 2003).

It was shown that with reduction of the nominal live load through increasing influence area, the

bias suggested by ASCE (2000) changes in accordance with ASCE-98 instead of Eq. (4-20)

(Bartlett et al., 2003).

The simulation results for office buildings done by Ellingwood and Culver (1977) was also used

to calculate the CoV of the maximum equivalent uniformly distributed live load regarding an office

floor over a 50-year reference period. The variance (𝜎𝐿𝑚𝑎𝑥

2 (𝑘𝑁/𝑚2)) of maximum live load during

50 years was reported to be

𝜎𝐿𝑚𝑎𝑥

2 = 0.033 + 4.025/√𝐴𝐼 (4-22)

By dividing the square root of Eq. (4-22) by Eq. (4-20), the CoV of the live load is calculated.

Bartlett et al. (2003) considered a bias of 0.9 and a CoV of 0.17 for a 50-year maximum live load.

Page 70: Reliability Assessment of Historical Masonry Structures

54

Moreover, they found out that the CoV values related to beams reported by Kariyawasam et al.

(1997) are smaller than those reported for columns. Also, the CoVs are almost insensitive to the

influence area (Bartlett et al., 2003).

In order to transform the live load into a live load effect, two factors including modeling and

analysis were considered by Bartlett et al. (2003). The modeling factor was used to address the

simplification of the actual live load as an equivalent uniformly distributed load. This factor was

reported to be unbiased with a CoV of 0.20 (Ellingwood et al., 1980). These values were assumed

to be related to columns and were decreased to a CoV of 0.1 related to beams by Kariyawasam et

al. (1997). Analysis factors were used to transform the live load to load effects. Ellingwood et al.

(1980) suggested an assumption of unbiased with a CoV of 0.05 regarding this factor. Bartlet et

al. (2003) determined the combined effect of the modeling and analysis factors having a bias of

1.0 and a CoV of 0.206 and assumed them to be time-independent, see Table 4-3.

Table 4-3 Statistical parameters for live load recommended by Bartlett et al. (2003).

Type Bias CoV

50-year maximum live load 0.9 0.17

Transformation to load effect 1.0 0.206

Table 4-4 Summary of recommended stochastic parameters provided by different authors.

Reference Classification Mean CoV Bias q

CIB (1989)

Office 2.64 0.19 - 2.42

Residential 1.73 0.20 - 1.57

Classroom 1.63 0.12 - 1.63

Rackwitz (1995)

Office 1.81 0.20 - 1.54

Residential 1.52 0.29 - 1.32

Classroom 2.65 0.36 - 2.23

Glowienka (2007)

Office 2.51 0.37 - 2.09

Residential 1.81 0.28 - 1.59

Classroom 3.61 0.22 - 3.49

Brehm (2011) Residential - 0.2 1.1 -

Page 71: Reliability Assessment of Historical Masonry Structures

55

There are other recommendations regarding the stochastic parameters of live load determined by

various researchers. A summary of recommended stochastic parameters provided by different

authors is presented in Table 4-4.

4.3.2 Point-in-time Live Load

Some studies have been directed to statistical analysis of point-in-time live load (Ellingwood &

Culver, 1977; Chalk & Corotis, 1980; Ellingwood et al., 1980). To model the load, a sustained

component (2-8 years duration) and an extraordinary component (one or twice per year) may be

considered. Bartlett et al. assumed a duration of 6 months for both components and derived point-

in-time live load parameters with a procedure presented by Wen (1990). The bias and the CoV

were reported to be 0.273 and 0.674, respectively. Transformation factors were assumed to be

same as those for a 50-year maximum live load. A Weibull distribution has been assumed for

point-in-time live load (Wen, 1990). The stochastic characteristics of point-in-time live load are

summarized in Table 4-5.

Table 4-5 Stochastic characteristics of point-in-time live load.

Load Type Bias CoV

Point-in-time Load 0.273 0.674

4.3.3 Distribution

It is recommended that the permanent component of live load be modeled by a Gamma distribution

(Pier & Cornel, 1973; Chalk & Corotis, 1980; Ellingwood, 1980; JCSS, 2001). Although there is

currently a lack of accurate data obtained from load measurements regarding transient live load,

the mean and standard deviation of the load was stated to be of almost equal quantity. Therefore,

an exponential distribution such as a Gamma distribution was also suggested to be used for the

transient component of live load (Pier & Cornel, 1973; Chalk & Corotis, 1980; Ellingwood, 1980;

Rackwitz, 1995). A Gamma distribution was recommended to be used to represent the Live load

Page 72: Reliability Assessment of Historical Masonry Structures

56

effect (modeling and analysis factors) by Ellingwood et al. (1980) and Kariyawasam et al. (1997).

However, Bartlett et al. (2003) assumed a normal distribution for the overall effect of modeling

and analysis. The reason for this assumption was stated to be the dominance in the overall effect

uncertainty of a single factor (the modeling uncertainty) as well as the lack of data associated with

the modeling uncertainty.

Table 4-6 Summary of recommendations for distribution type.

Variable Type Distribution Type Reference

Permanent Live Load Gamma distribution

Pier & Cornel (1973); Chalk

& Corotis (1980); Ellingwood

(1980); JCSS (2001)

Transient Live Load Gamma distribution

(Pier & Cornel (1973); Chalk

& Corotis (1980); Ellingwood

(1980); Rackwitz (1995)

Modeling and Analysis Factors Gamma distribution

Ellingwood et al. (1980);

Kariyawasam et al. (1997)

Normal distribution Bartlett et al. (2003)

Total Live Load Gumbel distribution

Allen (1975); MacGregor

(1976); Ellingwood et al.

(1980); European Committee

for Standardization (1994);

Kariyawasam et al. (1997);

Tabsh (1997); Ellingwood

(1999); Brehm (2011)

Point-in-time Live Load Weibull distribution Wen (1990)

The maximum live load over the lifetime of a structure has been assumed to be represented by a

time-independent random variable using a Gumbel distribution (Allen, 1975; MacGregor, 1976;

Ellingwood et al., 1980; European Committee for Standardization, 1994; Kariyawasam et al.,

Page 73: Reliability Assessment of Historical Masonry Structures

57

1997; Tabsh, 1997; Ellingwood, 1999; Brehm, 2011). Considering the live load as a Poisson

process may lead to obtaining different reliabilities in some cases in comparison with using an

equivalent Gumbel distribution. However, the assumption of representing the maximum live load

by a Gumbel distribution has been accepted in practice. A summary of the recommended

distribution types for different components and parameters as well as total live load is presented

in Table 4-6.

Wind is a natural phenomenon created by the differences in atmospheric temperature leading to

differences in air-pressure. Wind loads are one of the main horizontal loads on structures and

depend on many parameters. Among the parameters wind velocity (𝜐) and gust intensity,

estimated based on annual extremes, are the main influential ones. The shape of the

member/structure, the surrounding roughness, the dynamic behaviour of the structure and its

altitude above ground are other influential parameters. It is worth mentioning that to account for

the influence of altitude above ground, the wind load at an altitude of 10 m above ground is

considered as a reference value. In structural analysis, wind loads refer to the stresses or forces on

members due to the applied wind. Various studies have examined the uncertain nature of wind

load and estimated its related stochastic parameters, but as wind load parameters are strongly

dependent on the geographical region of the structure, the NBCC procedure for wind load

calculation as well as studies with respect to the regions of Canada are mostly presented here.

A simple gust factor approach was adopted by the 1995 NBCC to calculate wind load on structures.

This simplified procedure estimated the wind pressure (𝑝) against the surface of structures as

𝑝 = 𝑞𝐶𝑒𝐶𝑝𝐶𝑔 (4-23)

where 𝑞 is the reference velocity pressure, 𝐶𝑒 is the exposure factor, 𝐶𝑝 is the external pressure

factor and 𝐶𝑔 is the gust factor. The reference wind pressure is influenced by the climatic condition

of the region. With the assumption of constant climatic condition, the maximum distribution of the

wind pressure during the service life of a structure can be estimated by the annual maximum

Page 74: Reliability Assessment of Historical Masonry Structures

58

pressure distribution. The reference wind pressure is normally derived using Bernoulli’s rule and

the wind velocity. The 1 in 𝑇 year reference wind pressure (𝑞𝑇) is estimated based on the 1 in 𝑇

year wind velocity (𝑉𝑇) (Appendix C of the 1995 NBCC) as follows

𝑞𝑇 = 1

2𝜌𝑉𝑇

2 (4-24)

where 𝜌 is the density of air which strongly depends on the air temperature (e.g. 1.2929 𝑘𝑔/𝑚3

considering dry air at 0 ºC and standard pressure of atmosphere). The CoV of air density was

estimated to be roughly 2.5%-5.0% (Ellingwood et al. 1980; MacGregor et al. 1997) and is

significantly less than the CoV of 𝑉𝑇2. Therefore, air density is assumed to be deterministic. NBCC

specifies values for each of the variables used in the calculation of wind load. However, they are

uncertain.

1 h wind velocity data related to over 100 sites (mostly airports) with 10-22 years of record are

used to derive the reference velocity pressures presented in Appendix C of the 1995 NBCC.

Locations which are listed in Appendix C are the centres of urban areas and, therefore, mostly

differ from the locations from where the data were collected. Interpolation was used to determine

the 1 in 30-year wind velocities and dispersion parameters related to these urban centres from the

collected data. Moreover, for each site, 1 in 10-year and 1 in 100-year velocities were derived as

well, and finally, 1 in 10-year, 1 in 30-year and 1 in 100-year reference pressures were calculated

by means of Eq. 4-24. It should be mentioned that there are no available documents including the

transformation factors used to calculate the velocities and dispersion parameters of the locations

listed in Appendix C from the sites where data were collected. The CoVs of maximum annual

wind velocities related to data collection sites were lower than the corresponding CoVs back-

calculated from the wind pressures of sites listed in Appendix C (Bartlett et al., 2003).

Bartlett et al. (2003) investigated the stochastic characteristics of wind load (1 in 50 year specified)

in three different regions of Canada (Regina, Rivière-du-Loup and Halifax). Those characteristics

have been accepted for calibration of the load and load combination criteria of the NBCC (2005).

The annual maximum wind velocity of 311 sites were provided by the Engineering Climatology

Section of the Canadian Meteorological Centre in Downsview, Ontario for this investigation. 223

Page 75: Reliability Assessment of Historical Masonry Structures

59

out of the 311 sites had a data record of at least 10 years. Bartlett et al. (2003) calculated the bias

as the ratio of the mean maximum velocity likely in a 50-year design life (�̅�50) to the specified 1

in 50-year velocity (𝑉50) using

�̅�50

𝑉50=

1 + 3.050 𝐶𝑜𝑉𝑎

1 + 2.592 𝐶𝑜𝑉𝑎 (4-25)

where 𝐶𝑜𝑉𝑎 is the CoV of the maximum annual wind velocity. The stochastic characteristics

reported for the maximum wind velocity in 50-year period are summarized in Table 4-7.

Table 4-7 Stochastic characteristics for maximum wind velocity in 50-year (Bartlett et al., 2003).

Site 𝐶𝑜𝑉𝑎 Bias CoV

Regina 0.108 1.039 0.081

Rivière-du-Loup 0.170 1.054 0.112

Halifax 0.150 1.049 0.103

As mentioned above, in order to transfer the reference velocity pressure to the pressure developed

on the surface of structures, some coefficients including gust, exposure and pressure coefficients

must be considered. These coefficients are assumed to be time-independent and are considered to

fit a log-normal distribution (based on the Central Limit Theorem). Table 4-8 presents a summary

of the statistical parameters for these transformation factors reported by different authors.

Page 76: Reliability Assessment of Historical Masonry Structures

60

Table 4-8 Summery of statistical parameters of transformation factors reported by different

authors (Bartlett et al., 2003).

Reference Bias CoV Comment

𝐶𝑒

Ellingwood et al. (1980) 1.0 0.16

Davenport (1981) 1.0 0.10 Reported by Kariyawasam et al. (1997)

Davenport (1982) 0.80 0.16 Low building

Ellingwood and Tekie (1999) 0.93 0.129 Low Building

Davenport (2000) 0.80 0.20 Tall building (simplified methods)

Bartlett et al. (2003) 0.80 0.16

𝐶𝑔

Ellingwood et al. (1980) 1.0 0.11

Davenport (1981) 1.0 0.05 Reported by Kariyawasam et al. (1997)

Ellingwood and Tekie (1999) 0.97 0.093

𝐶𝑝

Allen (1975) 1.0 0.10 Added a model error coefficient of 0.85

Ellingwood et al. (1980) 1.0 0.12

Davenport (1981) 1.0 0.10 Reported by Kariyawasam et al. (1997)

Ellingwood and Tekie (1999) 0.88 0.067

𝐶𝑒𝐶𝑔

Allen (1975) 0.85 0.13

𝐶𝑔𝐶𝑝

Ellingwood et al. (1980) 1.0 0.16

Davenport (1981) 1.0 0.11 Reported by Kariyawasam et al. (1997)

Davenport (1982) 0.80 0.15 Low buildings

Ellingwood and Tekie (1999) 0.85 0.115

Davenport (2000) 0.80 0.21 Tall building (simplified method)

Bartlett et al. 2003 0.85 0.15

𝐶𝑒𝐶𝑔𝐶𝑝

Allen 1975 0.72 0.174

Ellingwood et al. (1980) 0.85 0.239

Davenport (1981) 0.85 0.150 Reported by Kariyawasam et al. (1997)

Davenport (1982) 0.54 0.325 Low buildings, only external pressures

Rosowsky and Cheng (1999) 0.73 0.264 Low buildings, high wind regions

Ellingwood and Tekie (1999) 0.70 0.208 Normal distribution for all factors

Davenport (2000)

0.57

0.71

1.00

0.250

0.150

0.087

Simple method

Detailed method

Wind tunnel and meteorological investigation

Bartlett et al. (2003) 0.68 0.22 Log-normal distribution

Page 77: Reliability Assessment of Historical Masonry Structures

61

The NBCC differs significantly from ASCE 2000 in terms of the transformation factors. For

example, the NBCC has reduced the gust and pressure coefficients for the design of low-rise

buildings to account for directionality and other factors while ASCE-98 (2000) suggests a separate

coefficient equal to 0.85 in this regard. The difference in pressure coefficients reported by the

NBCC and those reported by ASCE7-98 regarding taller buildings can be mentioned as another

example of the inconsistency between these two codes. This is due to the fact that the NBCC

pressure coefficients permit the designer to calculate the leeward suction at the mid-height of the

structure instead of at the top of the structure. Therefore, the pressure coefficients reported by the

NBCC are less severe than those reported by ASCE7-98.

JCSS (2001) estimated the wind load applied per unit area of a structure using the following

relations

i. Small rigid structures

𝑊 = 𝑞𝐶𝑎𝐶𝑔𝐶𝑟 (4-26)

ii. Large rigid structures sensitive to dynamic effects (natural frequency less than 1 Hz)

𝑊 = 𝑞𝐶𝑑𝐶𝑎𝐶𝑔𝐶𝑟 (4-27)

where q is the reference velocity pressure, 𝐶𝑎 is an aerodynamic shape factor, 𝐶𝑔 is the gust factor,

𝐶𝑟 is the roughness factor and 𝐶𝑑 is the dynamic factor. It is clear that the factors recommended

by JCSS are different from those recommended by the NBCC, except the gust coefficient and

therefore, their stochastic characteristics differ from each other. The mean and CoV of wind load

using the mean and CoV of the influential random variables assumed to be uncorrelated are as

follows

𝐸(𝑤) = 𝐸(𝐶𝑔)𝐸(𝐶𝑎)𝐸(𝐶𝑟)𝐸(𝑞) (4-28)

Page 78: Reliability Assessment of Historical Masonry Structures

62

𝐶𝑜𝑉𝑊2 = 𝐶𝑜𝑉𝐶𝑔

2 + 𝐶𝑜𝑉𝐶𝑎

2 + 𝐶𝑜𝑉𝐶𝑟

2 + 𝐶𝑜𝑉𝑞2 (4-29)

𝐸(𝑤) = 𝐸(𝐶𝑑)𝐸(𝐶𝑔)𝐸(𝐶𝑎)𝐸(𝐶𝑟)𝐸(𝑞) (4-30)

𝐶𝑜𝑉𝑊2 = 𝐶𝑜𝑉𝐶𝑑

2 + 𝐶𝑜𝑉𝐶𝑔

2 + 𝐶𝑜𝑉𝐶𝑎

2 + 𝐶𝑜𝑉𝐶𝑟

2 + 𝐶𝑜𝑉𝑞2 (4-31)

A summary of the statistical characteristics of the above random variables as suggested by JCSS

(2001) is presented in Table 4-9. The coefficient of variation of the resulting wind pressure is

shown to be almost double the coefficient of variation of the wind velocity according to JCSS

(2001). Therefore, the CoV of wind pressure can be estimated directly from the CoV of the wind

velocity. However, it is clear that it would be an approximation and may lead to a less accurate

estimation compared with using the formula above and the stochastic characteristics of the

influential factors directly.

Generally, the characteristic value of the reference wind velocity and consequently the reference

wind pressure have been specified based on each code criteria and may differ from those of

another. In case of the NBCC, the reference wind velocity and reference wind pressure have been

determined based on the probability of being exceeded 1 in 50 years. In other words, these

reference values have commonly a 50-year return period. From a statistical point of view, there is

an obvious limitation regarding the number of extreme values that can be measured during an

observation period of 50 years. However, a relatively large database can be gained for an

observation period of 1 year. MCS can then be used to generate extreme values for large

observation periods like 50 year.

Page 79: Reliability Assessment of Historical Masonry Structures

63

Table 4-9 Statistical characteristics of variables being influential on wind load (JCSS, 2001).

Reference Variable Bias CoV

Davenport (1987)

𝑞 0.8 0.2-0.3

𝐶𝑎

-Pressure coefficient

-Force coefficient

1.0

1.0

0.1-0.3

0.1-0.15

𝐶𝑔 1.0 0.1-0.15

𝐶𝑟 0.8 0.1-0.3

𝐶𝑑 1.0 0.1-0.2

Vanmarcke (1992)

Structure period

-Small amplitude

-Large amplitude

0.85

1.15

0.3-0.35

0.3-0.35

Structure damping

-Small amplitude

-Large amplitude

0.8

1.2

0.4-0.6

0.4-0.6

4.4.1 Point-in-time Wind Load

The statistical characteristics of point-in-time wind load which have been adopted for calibration

of load and load combination of the NBCC (2005) were estimated based on several assumptions.

Wind velocity and consequently, the resultant pressure are considered to be stochastic and the

transformation factors are time independent variables. 3 h is assumed to be the duration of each

wind pulse. Statistical parameters related to the maximum wind velocity in 3 h periods were

reported for three different regions of Canada (basis of calibration of load and load combination

of the NBCC (2005)) is presented in Table 4-10 (Bartlett et al., 2003). The transformation factors

which are used to transform wind velocity to wind pressure applied on the surface of structures are

the same for both maximum wind load over T years and the point-in-time wind load. Therefore,

Page 80: Reliability Assessment of Historical Masonry Structures

64

the previously mentioned transformation factors and their related stochastic characteristics can be

used to estimate stochastic modeling of point-in-time wind load as well.

Table 4-10 Statistical parameters of maximum 3-hour wind velocity in Canada (Bartlett, 2003).

Site Bias CoV

Regina 0.156 0.716

Rivière-du-Loup 0.064 0.149

Halifax 0.084 1.001

One may note that as the bias represents the ratio of the mean to the characteristic value suggested

by the relative code, this ratio can be used a basis for understanding the appropriateness of the

characteristic values reported by the codes. As the bias factor was shown to be near to 1 for wind

velocity during the longer observation period (50 years) as presented in Table 4-7, this proves that

the characteristic values of the NBCC code that Bartlett et al. (2003) used were appropriate.

However, bias factors reported for point-in-time wind velocities of the three different regions in

Canada as represented in Table 4-10 show around 90% difference between the actual mean value

and the NBCC characteristic value. This may be due to the fact that Bartlett et al (2003) used the

characteristic values related to a 50-year return period to calculate the point-in-time wind load bias

ratio leading to this considerable difference.

4.4.2 Distribution

The best-fit distribution for a 50-year maximum wind velocity has been reported to be a Gumbel

distribution (JCSS, 2001; Bartlett et al, 2003) and for point-in-time velocity the best fit distribution

has been reported to be Weibull (Bartlett et al., 2003). The log-normal distribution is suggested for

each of the transformation factors, (JCSS, 2001) as well as their overall combination (Bartlett et

al., 2003). However, Ellingwood and Tekie (1999) recommended normal distribution for all

factors. The overall wind load pressure is suggested to be modeled by a Gumbel distribution (JCSS,

2001). However, this distribution usually misses the upper limit. Therefore, a Gumbel distribution

may not be the best choice although it is commonly used in different studies as its application is

Page 81: Reliability Assessment of Historical Masonry Structures

65

simple and it needs only a few parameters. Kasperski (2000) suggested the Weibull distribution as

a better fit for wind load and to prevent ignorance of the upper limit. This is due to considering a

third parameter 𝜏, referred to as the shape parameter, besides the moments in the Weibull

distribution defining the upper limit. A summary of the recommended distributions for each

parameter as explained is presented in Table 4-11.

Table 4-11 Summery of the best fit distribution for parameters involved in wind load estimation.

Parameter Distribution Reference

Maximum velocity (50-year return period) Gumbel Bartlett et al. (2003); JCSS

(2003)

Point-in-time velocity Weibull Bartlett et al. (2003)

Each transformation factor

Normal Ellingwood and Tekie (1999)

Log-normal Bartlett et al. (2003); JCSS

(2003)

Overall transformation factor Log-normal Bartlett et al. (2003)

Wind pressure Gumbel JCSS (2001)

Weibull Kasperski (2000), Brehm (2011)

Snow loads are prevalent in mountainous and cold regions all over the word. There are

uncertainties in the nature of the occurrence of extreme snowfalls and consequently, the duration

and intensity of the resultant snow loads on structures. Therefore, snow loads should be treated in

a probabilistic manner. In order to calculate the actual snow load on a roof accurately, the

difference in the quantity of snow or rain being accumulated and that of the snow or rain being

revoked by the wind, melting or evaporation should be calculated (Ellingwood and O’Rourke,

1985). However, it is not possible to quantify snow in this manner as data are not available. As the

snow load and consequently its probabilistic parameters depend on the location of the structure

Page 82: Reliability Assessment of Historical Masonry Structures

66

(like wind load), here only the snow load model and the stochastic parameters reported in the

NBCC and for Canada are presented. The NBCC presents a model to estimate the snow load on a

roof of a structure as follows

𝑆 = 𝐼𝑠[(𝐶𝑏𝐶𝑤𝐶𝑠𝐶𝑎)𝑆𝑠 + 𝑆𝑟] (4-32)

𝐶𝑔𝑟 = 𝐶𝑏𝐶𝑤𝐶𝑠𝐶𝑎 (4-33)

where 𝐼𝑠 is importance factor for snow load, 𝑆𝑠 is the ground snow load with the probability of

exceedance of 1 in 50 per year, 𝑆𝑟 is the associated rain load, 𝐶𝑏 refers to the basic roof snow load

of a value of 0.8, 𝐶𝑤, 𝐶𝑠 and 𝐶𝑎 are the wind exposure factor, the slope factor and the accumulation

factor, respectively. Based on the NBCC, 𝑆𝑟 has to be taken less than 𝐶𝑔𝑟𝑆𝑟.

4.5.1 Ground Snow Load

Ground snow loads are considered as the basis of estimation of roof snow load in Canada. The

snow depth (𝑑) and the unit weight of snow (𝛾) are the parameters used in the calculation of

ground snow load. In other words, data are usually reported as depth of snow by weather stations.

Snow depth then transfers to ground snow load by means of the relation between depth and density

of snow. Appendix C of the 2010 NBCC includes the values of ground snow load for selected

locations in Canada. In order to calculate these recommended values, Gumbel distributions were

fitted to the reported data of the maximum annual accumulated depth related to 1618 stations

having 7-38 years of record (Newark et al., 1989; NBCC, 1995). Then, the values for the 1 in 30

years were determined. Snow densities of various geographical regions having common climatic

characteristics (manifested by forest type) were estimated and site elevations were considered

through normalization of the loads. The next step was preparation of smoothed contour maps.

Through interpolation and rectification accounting for the elevation, the final loads related to the

different geographical regions listed in Appendix C were computed. It should be mentioned that

there are no documents available regarding the weighting factors used in interpolating the ground

Page 83: Reliability Assessment of Historical Masonry Structures

67

snow values corresponding to the data collection sites to calculate the values concerning the sites

listed in Appendix C.

Bartlett et al. (2003) also studied the stochastic characteristics of snow load in Canada. They used

the data related to annual maximum snow depth collected from 1278 sites provided by the

Engineering Climatology Section of the Canadian Metrological Center. Table 4-12 presents the

characteristics of the annual maximum snow depth in Canada. Then, 50-year maximum snow

depths were calculated based on annual data, see Table 4-13.

Table 4-12 Characteristics of the annual maximum depth in Canada (Bartlett et al., 2003).

Type CoV Mean Median

Annual maximum snow depth 0.08-1.34* 0.491 0.47

* 95% of the CoV values are in the range of 0.22-0.95.

Table 4-13 Characteristics of the 50-year maximum depth in Canada (Bartlett et al., 2003).

Type Bias CoV

50-year maximum snow depth 1.1 0.2

Snow density is the other influential parameter on ground snow load. Therefore, determination of

stochastic characteristics of snow density is necessary to calculate the stochastic characteristics of

ground snow load and consequently those of the snow load on the roof of a structure. Statistical

characteristics of snow density have been summarized and reported by Kariyawasam et al. (1997)

based on the study of Newark (1984), see Table 4-14. The values were reported based on the forest

type in the region of the sites considered.

Taylor and Allen (2000) investigated the statistical characteristics the ground-to-roof

transformation factor in consistency with the definition of ground-to-roof snow load factor

recommended by 1995 NBCC. The database of 112 roofs in four Canadian cities during 13 years

Page 84: Reliability Assessment of Historical Masonry Structures

68

were used to estimate the statistical values of the ground-to-roof transformation factor as

summarized in Table 4-15.

Table 4-14 Statistical characteristics of snow density (Kariyawasam et al., 1997).

Forest Mean

𝑘𝑔/𝑚3

Standard Deviation

𝑘𝑔/𝑚3 CoV Sites

Acadian 220 50 0.23 Fredericton, Halifax, St. John’s, Charlottetown

Aspen grove 220 40 0.18 Edmonton, Saskatoon, Winnipeg, Thunder Bay

Boreal 190 60 0.32 Northern Prairies, mid-northern Ontario and

Quebec

Coast 430 25 0.06 Coastal British Columbia

Columbia 360 35 0.10 Southeastern British Columbia

Great Lakes 220 60 0.27 Southern and central Ontario, southern Quebec

Montane 260 25 0.10 Interior British Columbia and Yukon

Prairie 210 40 0.19 Southern Prairies, Regina

Subalpine 360 30 0.08 Vancouver and Fraser Valley

Tundra 300 80 0.27 Arctic

Taiga 200 80 0.40 Subarctic, Yellowknife

Ellingwood and O’Rourke (1985) did a study regarding the calculation of snow loads applied to

structures using probabilistic models. They accepted the general formula of roof snow calculation

(NBCC, 1977; ISO 4355, 1981; ANSI A58.1, 1982) as a product of ground snow load and a

transformation factor as follows:

𝑆𝑟 = 𝐶𝑆𝑔 (4-34)

Page 85: Reliability Assessment of Historical Masonry Structures

69

in which 𝑆𝑔 is the ground snow load and 𝐶 is the ground to roof transformation factor. Ellingwood

and O’Rourke (1985) estimated 𝑆𝑔 corresponding to the 98 percentile value over a 50-year return

period using Log-normal and Gumbel distribution functions based on the statistical study done by

Ellingwood and Redfield (1983), indicating that either the Log-normal or the Gumbel distribution

may be suitable for 𝑆𝑔

Log-normal distribution:

𝑆𝑔 = exp(𝜆𝑔 + 2.054𝜉𝑔) (4-35)

Gumbel distribution:

𝑆𝑔 = 𝜇𝑔 + 3.902/𝛼𝑔 (4-36)

𝜇𝑔 ≈ 𝑚𝑔 − 0.5772/𝛼𝑔 (4-37)

𝛼𝑔 ≈ 1.283/𝑠𝑔 (4-38)

where 𝜆𝑔 and 𝜉𝑔2 refer to the mean and variance of 𝑙𝑛 𝑆𝑔. 𝑚𝑔 and 𝑠𝑔 are the sample mean and

standard deviation.

O’Rourke et al. (1982) defined a relationship for determination of the overall combination of

transformation factors (𝐶𝑔𝑟) as follows

𝐶𝑔𝑟 = 0.47𝐸𝑇𝜀 (4-39)

in which 𝐸 is a wind exposure factor (from 0.9 to 1.3); T is thermal characteristic factor (from 0.1

to 1.2); and 𝜀 is the error term with a log-normal distribution, a mean value of 1.0 and a CoV of

Page 86: Reliability Assessment of Historical Masonry Structures

70

0.44. It is worth mentioning that the O’Rourke et al. (1982) formula leads to transformation factor

values which are from 0.53 to 0.91 times the 1995 NBCC basic roof snow load factor equal to 08.

Table 4-15 Statistical characteristics of transformation factor (Taylor & Allen, 2000).

Location Bias CoV

Sheltered Location

Halifax

Chicoutimi

Saskatoon

Recommendation

0.61

0.71

0.30

0.60

0.45

0.44

0.11

0.42

Exposed locations

Halifax

Ottawa

Saskatoon

Recommendation

0.47

0.44

0.33

0.5

0.46

0.41

0.22

0.42

Drift locations

Chicoutimi

Ottawa

Saskatoon

Recommendation

0.86

0.60

0.61

0.60

0.30

0.38

0.43

0.42

JCSS (2001) also presents a probabilistic model for snow loads. According to JCSS (2001), the

snow load on a roof can be computed by the following relation

𝑆𝑟 = 𝑆𝑔𝑟𝑘ℎ/ℎ𝑟 (4-40)

𝑟 = ƞ𝑎𝐶𝑒𝐶𝑡 + 𝐶𝑟 (4-41)

Page 87: Reliability Assessment of Historical Masonry Structures

71

in which 𝑆𝑔 accounts for the snow load on the ground at the weather station; 𝑟 is the transformation

factor of snow load on ground to snow load on roofs; ℎ is the altitude of the site of the building;

ℎ𝑟 is the reference altitude (300 m); 𝑘 is a coefficient accounts for the type of region in which

structure is located (𝑘 = 1.25 for coastal regions and 𝑘 = 1.5 for inland mountainous regions; ƞ𝑎

is a shape coefficient, (Table 4-16); 𝐶𝑒 is an exposure coefficient (Table 4-16); 𝐶𝑡 is a thermal

coefficient; and 𝐶𝑟 is a redistribution coefficient resulting from wind (neglected for monopitch

roofs, equal to ±𝐶𝑟𝑜 for symmetrical duopitch roofs, Figure 4-2, Figure 4-3).

Table 4-16 The exposure coefficient (𝐶𝑒) and shape factor (ƞ𝑎) (JCSS, 2001).

Slope of The Roof ƞ𝑎𝐶𝑒

𝛼 = 0° 0.4 + 0.6 exp (−0,1 𝑢(𝐻))*

𝛼 = 25° 0.7 + 0.3 exp (−0,1 𝑢(𝐻))

𝛼 = 60° 0

* 𝑢(𝐻) is the averaged wind speed during one-week period at roof level of H

Figure 4-2 Redistribution of snow load applied on a duopitch roof (JCSS, 2001).

Page 88: Reliability Assessment of Historical Masonry Structures

72

Figure 4-3 𝐶𝑟𝑜 as a function of the roof angle (JCSS, 2001).

𝑆𝑔 is considered to be time dependant. However, it is space independent within a region with

almost same altitude and same climatic condition. 𝑆𝑔 is determined by either the water equivalent

of snow or snow depth. The reported values are used directly in the determination of ground snow

load in the first case. In the second case, snow depth values have to be transformed to snow load

using the following relation

𝑆𝑔 = 𝑑𝛾(𝑑) (4-42)

𝛾(𝑑) =𝜆𝛾∞

𝑑ln {1 +

𝛾0

𝛾∞[exp (

𝑑

𝜆) − 1]} (4-43)

where 𝑑 is the snow depth; 𝛾(𝑑) is the accounts for the average weight density of the snow; 𝛾∞ is

unit weight at 𝑡 = ∞ (equal to 5 𝐾𝑁/𝑚3); and 𝛾0 is unit weight at 𝑡 = 0 (equals to 1.70 𝐾𝑁/𝑚3).

The stochastic characteristics of snow load variables recommended by JCSS (2001) are

summarized in Table 4-17.

Page 89: Reliability Assessment of Historical Masonry Structures

73

Table 4-17 Summary of stochastic characteristics of snow load variables (JCSS, 2001).

Symbol mean CoV

𝑆𝑔 Observation Observation

𝑑𝑔 Observation Observation

𝑘 1.25/1.5 Deterministic

ℎ𝑟 300 m Deterministic

𝛾0 1.7 Deterministic

𝛾∞ 5.0 Deterministic

𝜆 0.85 Deterministic

ƞ𝑎𝐶𝑒 Table 4-16 0.15

𝐶𝑡 0.8 – 1.0 Deterministic

𝐶𝑟𝑜 Figure 4-2 & Figure 4-3 1.0

4.5.2 Snow Related Loads

Associated rain load is one of the main snow-related loads accounting the weight of rain which

may fall over a snowpack. The values of associated rain load reported in the NBCC (Appendix A)

have been calculated based on the database related to winter rainfall according to the study of

Taylor and Allen (2000). The data being used in the determination of the probability models for

ground snow load are usually taken from daily measurements of the water equivalent ground load.

Therefore, as heavy rain can commonly penetrate down through the snowpack and drain away, it

is hard to claim the percentage of maximum rain on snow being captured by daily measurements

(Ellingwood & O’Rourke, 1985). Moreover, the available data of associated winter rainfall should

be filtered carefully to remove those on non-existent snowpacks. In other words, as there are no

available, reliable and clear data regarding the associated rain load, derivation of its statistical

parameters is difficult. Colbeck (1977) worked on roof loads resulting from rain-in-snow.

Page 90: Reliability Assessment of Historical Masonry Structures

74

However, there is still a lack of a general and acceptable approach for analysis of associated rain

load.

Icing on eaves and the development of ice dams are two other associated snow loads which have

been shown occasionally to lead to structural distress (MacKinley, 1983). Roof overhangs and

moisture-resistant roofing members are substantially endangered with these effects. Critically,

warm structures being purely insulated are threatened with ice formation as well. However, there

is a of comprehensive and quantitative study in this regard.

4.5.3 Point-in-time Snow Load

The statistical characteristics of point-in-time snow load have not been investigated broadly and

comprehensively. Bartlett et al. (2003) just studied the point-in-time snow load in Canada. There

are no available documents regarding the details of their study and some general assumptions were

pointed out in their related publication. Snow accumulation was assumed to be a stochastic process

and transformation factors (converting depth to the load and the ground snow load the roof one)

were assumed to be time dependent. A 14-day period over 3 months of a year was considered as

the point-in-time pulse (i.e. 6 events per year). Table 18 presents the statistical characteristics of

point-in-time snow depth and transformation factor to load effect consistent with Canadian regions

and the 1995 NBCC criteria recommended by Bartlett et al. (2003).

Table 4-18 Statistical characteristics of point-in-time snow (Bartlett et al., 2003)

Parameter Bias CoV

Point-in-time depth 0.196 0.882

Transformation factor 0.600 0.420

4.5.4 Distribution

Different researchers have introduced almost separate coefficients and parameters as random

variables in the determination of snow loads. It is clear that each of these parameters have their

own distribution, and therefore differ from one reference to another. Here, the best fit distributions

for the random variables and parameters of the NBCC and JCSS, which are the two references

Page 91: Reliability Assessment of Historical Masonry Structures

75

being explained above, are presented. Considering the NBCC criteria for the calculation of snow

load, snow densities were recommended to have a Normal distribution; 50 year-maximum depth

and point-in-time depth were fitted to Gumbel and Weibull distributions, respectively.

Transformation factors were allocated Log-normal distribution (Bartlett et al., 2003). JCSS (2003)

recommended a Gamma distribution for snow depth on the ground. Density has been considered

deterministic as well as the insulation parameter (one of the transformation factors). The shape and

redistribution coefficients (two other transformation factors) have been fitted to Beta distributions.

All other parameters were assumed to be deterministic, see Table 4-17. Ellingwood and O’Rourke

(1985) considered both Log-normal and Gumbel distributions for the ground snow load based on

the study done by Ellingwood and Redfield (1983) and recommended a Log-normal distribution

for the transformation factor. Ellingwood and Redfield (1983) proposed the Log-normal

distribution for regions where the snow cover shows intermittency over the snow season and the

annual extremes happen commonly with a severe winter storm. Moreover, large variabilities have

been shown to exist in annual extreme snow loads (Eurocode, 1994). The Log-normal distribution

has a longer upper tail compared to the Gumbel distribution to be able to represent these

variabilities. The recommended distributions regarding snow load random variables are

summarized in Table 4-19.

Table 4-19 Summery of recommended distributions for the snow load random variables.

Variable Distribution Comment Reference

Density Normal - Bartlett et al. (2003)

Snow depth

Gumbel 50-year maximum depth,

reported for 2005 NBCC Bartlett et al. (2003)

Gamma - JCSS (2001)

Weibull Point-in-time Bartlett et al. (2003)

Transformation factor

Log-normal Overall factor, based on

1995 NBCC Bartlett et al. (2003)

Beta Shape coefficient &

redistribution coefficient JCSS (2003)

Log-normal - Ellingwood &

O’Rouke (1985)

Page 92: Reliability Assessment of Historical Masonry Structures

76

Chapter V: Stochastic Modeling of Historic Masonry Materials

Masonry has been used in construction for thousands of years, with many structures still standing.

In other words, masonry is among the most durable construction materials and masonry structures

usually outlive their design or expected service life. In the past, natural units like cut stones were

commonly used. They have been replaced with manufactured units as the primary building block

because of efficient industrial manufacturing. Hydraulic lime mortars have also been substituted

by Portland cement mortars. Therefore, a wide variety of materials in terms of units and mortar

have been used throughout construction history. Masonry is mostly identified as a homogenous

material in structural engineering. It leads to consideration of the characteristics of the units and

mortars after combination in order to estimate masonry properties (Brehm 2011).

The resistance of structures is mainly based on the strength of the materials from which the

structure is constructed. Therefore, determination of the material properties which are influential

on the structural resistance is a key step in the estimation of structural reliability. Historical

structures often contain multiple construction materials (different bricks or stones for example)

with various characteristics leading to more complexity in the identification and determination of

material properties. Additionally, there are usually no construction documents (even as-built

documents in modern construction often fail to provide enough information). In this situation,

there is a need for supplementary in-situ and laboratory tests to quantify material properties

(FEMA 365). As a result of possible damage to the structure’s fabric or even structural instability,

destructive techniques should not be used in order to get information about the type of materials

in the construction and their properties. All of these information shortages and concerns make

analysis and understanding of the behaviour of a historical structure more complicated but

necessary.

The main focus here is to present a possible procedure in order to get an estimate of the material

properties and their variability without applying destructive tests to the historical structure. There

are various formulas presented by standards and previous researchers in order to calculate the

Page 93: Reliability Assessment of Historical Masonry Structures

77

resistance of a structure under its probable failure modes. The most frequently used and thus

influential material properties in these formulas are modulus of elasticity, compressive strength,

tensile strength, cohesion and coefficient of friction. Since compressive strength is most

fundamental for masonry, the main focus of this thesis is the estimation of the probabilistic model

related to compressive strength. Useful techniques in order to get information about other material

properties of historical structures and the stochastic modeling of them are also explained.

Evaluation of the performance of historical masonry structures and gathering information about

them would enhance understanding of their current conditions and capacity and would minimize

the modifications and additions applied to them. There is a wide variety of evaluation methods

available, showing different accuracy and feasibility. Some of these methods have been

specifically developed to evaluate masonry structures. However, some of them are borrowed from

other fields of science (e.g. archeology and aerospace). Investigation procedures usually include

several complementary evaluation methods. Different levels of damage or deconstruction can be

incurred by masonry by using masonry evaluation techniques. In case of historic structures, even

small damage can be lead to expensive and difficult repairs. Typically, investigation approaches

and relative evaluation methods should be collected carefully to have minimum disruption to

historical structures. A list of nondestructive testing techniques, in-situ evaluation approaches and

their related purposes is summarized in Table 5.1. Standards and codes which govern various

nondestructive testing methods are reported in Table 5-2. In order to choose an appropriate testing

technique, several criteria should be taken into account. The type of required information,

structural condition, cost and complexity are among the most important criteria in selection of

efficient and feasible testing techniques. Table 5-3 categorizes testing techniques being applicable

to evaluate or investigate different masonry conditions based on the above-mentioned criteria. An

effectiveness grade is given to each method based on the accuracy of the information that the

method can provide. This classification considering the cost, complexity and effectiveness grades

is based on the experiences of different authors who worked with these methods and is expressed

here only as a general guideline. In other words, all these parameters and consequently the

classification may differ based on various factors (e.g. location, accessibility and schedule). In the

Page 94: Reliability Assessment of Historical Masonry Structures

78

case of load testing, a service level is usually considered so that no significant damage is applied

to the structure. If it is required to increase the load to reach a level near failure, considerable repair

may be needed. Moreover, in visual testing, slight repair may be required in the case of a borescope

and other small probe observations. However, larger openings for visual testing result in more

significant repairs.

Codes have not considered more detailed structural characteristics like cracks and fatigue in the

evaluation of structural resistance. However, in the case of more detailed reliability assessment,

there is a need to involve more detailed characteristics of structures in the estimation of resistance.

Cracks, spalls, efflorescence and surface erosion are among the most well-known signs of masonry

distress. Visual observation is a significant method in the evaluation of masonry distress. An

experienced investigator can conclude the reason of many crack patterns and surface erosion

patterns from only surface observation. Sometimes visual observation should be accompanied with

other subsurface investigations including non-destructive evaluation, probe openings and

borescope observations to estimate the subsurface conditions more accurately.

Cracks, spalls and surface erosion can be evaluated using a number of nondestructive techniques.

Ultrasonic Pulse Method (UPM) is used to estimate crack depth and extent in place as well as

surface condition. The impact echo method is also useful in getting information about cracks.

Petrography in the laboratory can be used to get information about distress mechanism where it is

possible to take samples from materials.

In masonry, the corrosion of embedded metal elements may lead to cracking. To investigate the

probability of active corrosion in subsurface layers, a non-destructive method namely half-cell

potential is used.

The appropriate non-destructive technique is selected based on the situation. However, it is

recommended to use visual observation (done by an expert) at the beginning of the performance

evaluation procedure. Generally, to do more accurate reliability evaluation, determination of more

exact limit state functions and consequently more detailed structural resistance evaluation is

necessary. Estimation of more exact limit state functions considering more detailed structural

characteristics is out of scope of current thesis. Here only the most important material properties

being used in code-based structural resistance evaluation and their stochastic characteristics are

discussed.

Page 95: Reliability Assessment of Historical Masonry Structures

79

Table 5-1 Nondestructive testing techniques and in-situ evaluation approaches (Hussain &

Akhtar, 2017).

Testing Method Measured parameters

Stone masonry Brick masonry

Visual

Visual inspection Macroscopic flaws, cracks

and deformation

Macroscopic flaws, cracks

and deformation

Borescope Macroscopic flaws, cracks

and deformation

(underneath the surface)

Macroscopic flaws, cracks

and deformation

(underneath the surface)

Acoustic

Ultrasonic Pulse Method

(UPM)* Flaws, Crack, deterioration,

strength and modulus of

elasticity

Flaws, Crack, deterioration,

hardness of surface, strength

and modulus of elasticity Acoustic emission

Impact-Echo (IE)*

Mechanical Pulse Velocity

(MPV)

Voids and discontinuities of

masonry

Voids and discontinuities of

masonry

Physical methods

Rebound hammer*

Strength Hardness of surface and

strength Waitzmann hammer

Flatjack

Push or shove test _

Shear strength of brick

assembly

Bond wrench test _

Flexural bond strength

(Tensile strength)

Radar Method

Ground Penetrating Radar

(GPR)

Flaws, cracks, voids and

moisture content

Flaws, cracks, voids and

moisture content

Page 96: Reliability Assessment of Historical Masonry Structures

80

Testing Method Measured parameters

Stone masonry Brick masonry

Penetrating radiation

methods

Radiography Surface voids, crack, path

density

_

Radiometry _

X-Ray Methods

X-Ray tomography Chemical composition

(stone), chemical attack,

inner structure

_

X-Ray diffraction _

X-Ray fluorescence _

Thermal imaging

Infrared thermography Crack, voids, delamination,

damage to thermal

insulation, air circulation

Crack, voids, delamination,

air circulation

Corrosion diagnose

Phenolphthalein indicator

test

Diagnosis of corrosion

Diagnosis of corrosion

Chloride penetration test

Half-cell potential test

Rapid chloride test

Quantab test

Volhard test

*intended to be used for concrete but can be applied or can be adopted for masonry as

well.

Page 97: Reliability Assessment of Historical Masonry Structures

81

Table 5-2 Summary of standards and codes including various non-destructive methods.

Testing Methods Standards

Visual ASTM C823/C823M

NDIS 3418:1993

Ultrasonic Pulse Method (UPM) ASTM C597-97

NDIS 2416-1993

IS 13311 (Part 1):1992

BS 1881: Part 203: 1986

BS 4408: Part 5

BS 12504-4, Part 4 2004

ISO/DIS 8047, C-26-72

COST 17624

Impact-Echo (IE) ASTM C1383-15

Acoustic emission ASTM E2983

Rebound Hammer ASTM C805-97

ASTM D5873

IS 13311 (Part 2):1992

BS 1881: Part 202: 1986

EDIN EN 12398 (1996)

ISO/CD 8045

Flatjack ASTM C1196-14a

Ground Penetrating Radar (GPR) ASTM D6087-08

ASTM D6432-11

Radiography ASTM E1742/E1742M

NDIS 1401-1992

BS 1881: Part 205: 1970

BS 4408: Part 3

IR thermography ASTM D4788-88, D 4788-03 (2013)

Phenolphthalein indicator test ASTM C1202

ASTM C114

AASHTO T277

Half-cell potential test ASTM C876-91

Volhard test ASTM 1411-09

NT 208

BS 1881-Part 6

DS 423.28

NS 3671

Page 98: Reliability Assessment of Historical Masonry Structures

82

Table 5-3 Classification of testing techniques appropriate for evaluation of masonry (Harvey JR.

& Schuller, 2010).

Destructive Non-destructive with

minor repairs Non-destructive

Sam

pli

ng (

core

tes

ting

& p

rism

tes

tin

g)

Bo

nd w

ren

ch t

est

In-s

itu

lo

ad t

esti

ng

Pet

rog

raph

y

Fla

tjac

k/S

hea

rjac

k

Gro

und

Pen

etra

tion

Rad

ar (

GP

R)

Ult

raso

nic

Pu

lse

Met

ho

d (

UP

M)

Imp

act

Ech

o (

IE)

Rad

iog

rap

hic

Infr

ared

Th

erm

og

raph

y (

IRT

)

Hal

f-ce

ll p

ote

nti

al

Reb

oun

d h

amm

er/S

urf

ace

test

s

Met

al d

etec

tio

n (

Ind

uct

ion

)

Vis

ual

Characteristics*

In-place strength G G E A E A A A

In-place stress E

In-place uniformity G G G G G G G G A

In-place deformability E E A A A

Location of crack E E G A

Movement of crack G

Rebar information E G E A E

Corrosion of rebar G A

Location of anchor and ties E E E

Voids in grout E A A E G A

Voids in masonry E A A E G A

Durability E E E G A

Performance under applied loads E

Cost**

M M H H M H H H H H M S S S

Complexity**

L M H H M H H H M M L L L L

* E = Excellent information, G= Good information, slightly variable or unclear results, A = Approximation, extremely variable

** L = Low, M = Medium, H = High

Page 99: Reliability Assessment of Historical Masonry Structures

83

The modulus of elasticity is the most important material property for determining the deformation

of structures. JCSS (2011) as a probabilistic model code relates modulus of elasticity to the

compressive strength of masonry in order to estimate the stochastic modeling as follows

𝐸𝑚 = 𝑐1. 𝑓𝑚′ (5-1)

𝐸𝑚,𝑗 = 𝑌2 . 𝐸𝑚 (5-2)

where 𝐸𝑚 is mean of modulus elasticity of masonry, 𝑐1 is a coefficient according to Schubert

(2010) summarized in Table 5-4 and 𝑌2 is log-normal variable with a mean of 1.0 and coefficient

of 25% for all types of unit and mortar. It should be considered that 𝑐1 is claimed to be a prior

value needing updating using test data. As can be seen, the reported values of 𝑐1 are mostly

presented for new materials not historical ones. Therefore, using this formula to estimate the

modulus of elasticity of ancient building materials and understand their deformation behaviours

may lead to inaccuracy and unreliable estimation.

Table 5-4 Values of 𝑐1 involved in probabilistic model of modulus of elasticity (Schubert, 2010).

Unit Mortar 𝑐1

CS[1] GPM[4], TLM[5] 500

AAC[2] GPM 520

TLM 560

LC[3] GPM 1040

TLM 930

Perforated clay brick

GPM 1170

TLM 1190

Lightweight 1480

[1] Calcium Silicate, [2] Autoclave Aerated Concrete, [3] Lightweight concrete

[4] General Purpose Mortar, [5] Thin Layer Mortar

In order to prevent any damage and disturbance to historical structures, it is suggested to employ

non-destructive tests for these kinds of structures instead of loading techniques. The Ultrasonic

Page 100: Reliability Assessment of Historical Masonry Structures

84

Pulse Method (UPM), Figure 5-1, is a nondestructive method which has proven to be efficient in

estimating the modulus of elasticity of masonry structures (Brozovsky 2013). Consequently, to

obtain a data set related to the modulus of elasticity of historical masonry structures, UPM is

recommended

Figure 5-1 Ultrasonic Pulse Method (n.d.).

The compressive strength of the masonry is usually the most important information being desired

to estimate for structural resistance evaluation. Compressive strength can be used to estimate the

load-bearing capacity of structures. There are difficulties in the estimation of compressive strength

of existing masonry structures including variable material properties, different construction

techniques, lack of knowledge about the existing damage made throughout the life of a masonry

structure and the absence of appropriate code. In the case of ancient structures, compressive

strength was probably not considered or specified in the design procedure and any documents

regarding the compressive strength of the masonry have been destroyed or lost throughout history.

Even if some documents regarding the compressive strength of the masonry in the design stage

could be found, they are not that reliable as the present compressive strength of the masonry

obviously is different from that hundreds of years ago. Weathering, deterioration and the effects

of long-term applied load are some of the causes of this difference in material strength. Moreover,

Page 101: Reliability Assessment of Historical Masonry Structures

85

inspection, specimen removal and testing are restricted regarding heritage structures. All these

statements show the difficulties but necessities of the determination of the present compressive

strength of the historical masonry structure to be able to evaluate the structural resistance

accurately. Techniques used to estimate the compressive strength of existing structures

experimentally range from simulating the structure through testing specimens in laboratories to

removing a prism from the structure in order to do a test. Each technique has its own advantages

and disadvantages. However, an attempt is made here to find an optimum procedure that allows

estimation of the variability of the compressive strength of a structure in addition to the mean value

in order to be able to do reliability analysis. A brief review of the most well-known approaches

and testing methods for estimation of the compressive strength of masonry follows.

The compressive strength of a masonry structure is based on the compressive strength of its

components (units and mortar). Some of the methods are based on the determination of

compressive strength of whole masonry structure according to the estimation of compressive

strength of the unit and mortar individually. Table 5-5 illustrates some of the proposed expressions

for estimation of a deterministic value of compressive strength of brick masonry according to the

compressive strength of the brick and the mortar.

The JCSS (2011) probabilistic model code also proposed formulas regarding the estimation of a

stochastic model of masonry compressive strength using the compressive strengths of its

components to do reliability assessment

𝑓𝑚′ = 𝑘. 𝑓𝑏

𝛼. 𝑓𝑚𝑜𝛽 (5-3)

𝑓𝑚,𝑗 = 𝑌. 𝑓𝑚′ (5-4)

where , 𝑓𝑚′ , 𝑓𝑏, and 𝑓𝑚𝑜 represent the compressive strength of masonry, and the mean values of the

unit and mortar compressive strengths, respectively. K, 𝛼 and 𝛽 are coefficients. 𝑓𝑚,𝑗 represents

the probabilistic model for compressive strength and 𝑌 is a variable related to uncertainties

distributed lognormally, see Table 5-6.

Page 102: Reliability Assessment of Historical Masonry Structures

86

Table 5-5 Masonry compressive strength determination from the strengths of its components.

Reference Expressions

Mann (1982) f'm=0.83fb0.49fmo

0.32

Hendry & Malek (1986) f'm=0.317fb0.531fmo

0.208

Dayaratnam (1987) f'm=0.275fb0.5fmo

0.5

Eurocode 6 (1996) f'm=0.5fb0.7fmo

0.3

Bennet et al. (1997) f'm=0.3fb

ACI 530.99 (1999) f'm =2.8+0.2fb

MSJC (2002) f'm=(400+0.25fb)/145

Dymiotis & Gutlederer (2002) f'm=0.3266 fb(1-0.0027fb+0.0147fmo)

Gumaste et al. (2006) f'm=0.317fb0.866fmo

0.134

Kushik et al. (2007) f'm=0.63fb0.49

fmo0.32

Garzón-Roca et al. (2013) f'm=0.53fb+0.93fmo-10.32

Table 5-6 Statistical characteristics of 𝑌 involved in probabilistic model of compressive strength

(JCSS, 2011).

Unit Type Mortar Mean CoV Distribution

CS[1]

TLM[4]

1.0 0.2 Log-normal

CS (Large sized units) 1.0 0.2 Log-normal

AAC[2] 1.0 0.16 Log-normal

AAC (Large sized units) 1.0 0.14 Log-normal

CB[3] GPM[5] 1.0 0.17 Log-normal

[1] Calcium Silicate, [2] Autoclave Aerated Concrete, [3] Clay Brick

[4] Thin Layer Mortar, [5] General Purpose Mortar

The procedure of determination of the compressive strength of a historical masonry structure based

on that of its components seems not to be effective due to the following reasons:

Page 103: Reliability Assessment of Historical Masonry Structures

87

1) To get information about the compressive strengths of the unit and mortar, access to the

construction documents and/or application of material testing is required. In the case of

historical structures, typically construction documents are not available and removal of

units contravenes conservation guidelines.

2) There are often various types of mortars and units with different material properties in a

single historic structure. Therefore, it may not be appropriate to apply the expressions

above to the historical building.

3) The formulas above have mostly been developed for new masonry materials and therefore

may not be accurate enough for historical materials.

4) There is no standard correlation between strength of whole masonry and the strength of its

components. Different correlations have been suggested by different researchers leading to

different results.

5) Any error in measuring compressive strength of the unit and mortar would result in

propagation of the error and consequently, inaccuracy in estimation of the compressive

strength of the whole masonry (Sabri et al., 2015; Steil et al., 2001; Chunha et al., 2001)

6) As masonry constituents work together in a structure, the homogeneity of the components

is an influential factor in compressive strength of the whole masonry. Therefore, it may be

inaccurate to estimate the compressive strength of the masonry considering mortar and

units individually (Radovanović, 2015).

7) There is no standard approach to estimate the compressive strength of mortar in-situ. Some

codes (e.g. The American Standard ASTM C 780) propose testing samples which are

constructed with the same materials as those in the existing masonry structure. This method

seems not to be effective and feasible in the case of historical structures, as the strength of

the mortar has been changed over hundreds of years due to deterioration. Moreover, the

compressive strength of the mortar depends on several factors including water/cement

ratio, mix proportions, aggregate ratio and sand type (Nwofor, 2012; Appa Rao, 2001;

Neville, 1996) which give challenges to correct sampling. Workability is influential on the

strength of constructed masonry as well. Zejak (2015) showed an inconsistency between

the tests results of mortar compressive strength determined in the field with the structural

properties of the masonry. Overall, the properties of prepared mortar samples may differ

Page 104: Reliability Assessment of Historical Masonry Structures

88

from the properties of mortar in historical structures leading to inaccuracy in the estimation

of the compressive strength of the whole masonry structure.

Core testing, categorized as small diameter core testing and large diameter core testing, is another

technique used for determining the compressive strength of masonry. Small diameter core testing

is used for units and mortar individually and large diameter core testing is appropriate for whole

masonry (unit and mortar together). There are some difficulties associated with these two methods.

In the case of small core testing techniques, mortar sampling is difficult and results are the

compressive strength of unit and mortar individually. The sampling direction is influential on the

estimated compressive strength. In other words, the compressive strengths of samples taken from

the same element differ by changing the sampling directions. In the case of large core testing, the

number of samples is restricted as the extraction of many large samples may influence the strength

of the structure. Moreover, it is not possible to do large core testing for elements with small cross

sections such as columns as the elements that are extracted from them might lead to deformation

or a decrease in the bearing capacity. The compressive strength of the extracted sample shows the

compressive strength of the location where sample is taken from. As a single historical structure

might be constructed by different masons from different materials with different strengths, this

testing method may lead to inaccuracy in terms of estimation of the compressive strength of the

whole masonry. The direction of sampling is also influential in the case of large core testing.

Masonry prism testing is another sampling method used for the estimation of compressive strength

of masonry. There are two ways of prism testing, including construction of prisms with the same

units and mortar as those by which the structure build up and extraction of prisms from existing

structures using a saw-cutting machine. The key advantage of this technique is its usability for all

kinds of masonry both in terms of the material (e. g. block, clay brick, and silicate brick) and

configuration (solid units and hollow units). Using each method of prism testing raises

controversial issues. In the case of the first way, it is challenging to be able to construct the prisms

in laboratories having exactly all detailed properties of the masonry in the existing structure. This

difference in detail may lead to inaccuracy in the estimation of compressive strength. Moreover, it

is hard and sometimes impossible to find the exact materials which had been used in the original

structure due to their uniqueness. The cost of the first way of prism testing is high as all materials

Page 105: Reliability Assessment of Historical Masonry Structures

89

(unit and mortar) besides expert masons have to be provided. The problems of the second prism

testing method are the same as the large dimeter core testing method as mentioned above.

Both the core testing and prism testing are categorized as sampling techniques in Table 5-3.

Finding the appropriate sample size to remain intact during removal and transportation is very

difficult when extracting a sample from existing masonry. Any damage or fracture made to the

specimen throughout the procedure makes the sample useless. Overall, both core testing and prism

testing can easily lead to irrational and inaccurate estimation of the actual compressive strength

and methods involving sample extraction are considered as destructive testing methods which

makes them unacceptable for use on historic structures that are to be conserved.

Another testing method that has become popular in the evaluation of compressive strength is the

flat-jack testing method, See Figure 5-2. This method is an in-situ technique for evaluation of the

stress in the masonry for determining the masonry strength. The method is described as either non-

destructive with minor repairs or semi-destructive, based on different references. In this method,

thin stainless-steel bladders are placed into slots cut in masonry (in bed joints) some distance apart

– one above the other. The bladders are inflated to apply pressure and the deformation of masonry

located between them is monitored. Through the gradual increase of pressure, the stress-strain

relationship and loading cycles can be determined. The compressive strength is then estimated

based on the resultant experimental stress-stain curve together with a value of Young’s

compressive modulus. The compressive strength can be estimated more exactly if damage to the

masonry can be accepted: the maximum pressure applied by the flatjacks is more than the

compressive strength of masonry and the development of cracks is considered as a sign of failure.

Rossi (1987) and Noland et al. (1990) evaluated the accuracy of the flatjack technique. Both

showed that flatjack testing has an error ranging between 15%-20% which is reasonable and

acceptable for practical estimation of compressive strength. However, there are also some

difficulties regarding the flatjack technique besides the restrictions associated with its application

mentioned above. The section of masonry which is under testing has to be damaged and reach the

failure level to be able to estimate compressive strength. In the case of elements or construction

parts with small cross section (e.g. a narrow column), application of flatjack method may not be

possible due to the reduction of bearing capacity or total destruction of the element tested. Also, it

is difficult to interpolate results associated with unsuccessful cutting and recovered distances.

Page 106: Reliability Assessment of Historical Masonry Structures

90

Reliable interpolation of results obtained from too weak or nonhomogeneous materials may be

challenging. In addition, there is a need to repair the masonry after using this technique which

involves repointing the joint with unloaded mortar in what was originally a loaded mortar joint

(thus altering the structure locally) and finding a compatible mortar may be expensive or even

impossible in the case of historical buildings. Flatjack technique can not also predict the

compressive strength of multi-wythe walls (common walls in the historic masonry structures)

accurately due to the short length of the bladder which can not reach to the inner wythe, see Figure

5-3. Moreover, in order to assess the reliability of a historical structure, it is not feasible to estimate

the variability in the compressive strength over an element or masonry structure by this method.

This is due to the fact that to obtain a CoV of compressive strength, flatjack testing should be

repeated in many locations leading to a high level of destruction and interference to a structure

with historical value.

Figure 5-2 Flatjack technique (n.d.).

Page 107: Reliability Assessment of Historical Masonry Structures

91

Figure 5-3 Schematic of flatjack technique.

All the above statements regarding different methods of compressive strength estimation show that

there is a need for a new method to determine compressive strength which is compatible with

historical structures and able to provide the variability for reliability analysis. It is therefore

suggested that the compressive strength be estimated from the modulus of elasticity. Different

researchers and codes and standards recommend different relationships between the compressive

strength and the modulus of elasticity of masonry. A summary is presented in Table 5-7.

As with the expressions for estimation of whole masonry strength based on the strength of its

components, these formulas may also not lead to an exact value of the compressive strength

because they were developed for modern materials rather than historical ones. However, using a

non-destructive test multiple times and deriving therefrom a mean strength and the associated

variability will potentially provide more accurate values of that strength and variability than trying

to determine the means and CoVs of compressive strengths of the mortars and the units. The non-

destructive UPM can provide a measure of the variability of the materials, whist also not causing

damage to the structure under consideration. Obviously, the accuracy and applicability of this

suggested procedure should be checked experimentally in future work.

As can be seen, there is a wide range of suggestions that have been presented regarding the

relationship between E and f’m. The difference is only in the coefficients, here referred to as (k).

The distribution of the coefficients is illustrated in Figure 5-4. As can be observed, the coefficient

Page 108: Reliability Assessment of Historical Masonry Structures

92

of f’m is most frequently in the range of 700 to 900 in the case of concrete masonry. This range is

also frequently reported in the case of clay masonry in addition to the coefficient being equal to

1000.

Figure 5-4 Distribution of the coefficient k for clay and concrete.

According to Figure 5-5, the fifty percent-probable coefficients are 800 and 890 for clay and

concrete masonry respectively. Therefore, it can be concluded that the reported coefficients which

are higher than 800 in the case of clay masonry and 890 in the case of concrete masonry lead to

more conservative compressive strength estimates, while those lower than these fifty-percent

probable coefficients result in higher values of compressive strength. Therefore, the compressive

strength of historical masonry structures could be estimated according to the values of E obtained

from the non-destructive tests and the relation between E and 𝑓𝑚′ . Coefficients for stone masonry

have not been reported, so this is clearly an area where work needs to be done to aid in the

estimation of the reliability and safety of historic structures. Given that stone units vary in size as

well as source material, it is likely that more than one coefficient will be needed to provide

reasonable estimates of strength from modulus measures. The best fit statistical distributions have

to be determined as well.

0

1

2

3

4

Fre

quen

cy

Relation between E and f'm

Concrete Masonry

Clay Masonry

Page 109: Reliability Assessment of Historical Masonry Structures

93

Table 5-7 Summary of relations between 𝐸 and 𝑓𝑚′ found in literatures.

Reference Formula Type

BS 5628 (1978) 𝐸 = 900 𝑓𝑚′ Core filled concrete masonry

NZS (1990) 𝐸 = 800 𝑓𝑚′ Masonry

Eurocode 6 (1996) 𝐸 = 1000 𝑓𝑚′ Masonry

FEMA 306 (1999) 𝐸 = 550 𝑓𝑚′ Clay masonry

MSJC (2002) 𝐸 = 700 𝑓𝑚

′ Clay masonry

𝐸 = 900 𝑓𝑚′ Concrete masonry

IBC (2003) 𝐸 = 700 𝑓𝑚′ Masonry

CSA (2004) 𝐸 = 850 𝑓𝑚′ < 20000 MPa Clay and concrete masonry

NNI (2005) 𝐸 = (400 − 1000) 𝑓𝑚′ Masonry

AS3700 (1998)

𝐸𝑚∗∗ = 700 𝑓𝑚

′ 𝐸𝐿∗∗ = 450 𝑓𝑚

′ Clay units (5 MPa <f'uc<30 MPa, M2 and M3)

𝐸𝑚 = 1000𝑓𝑚′ 𝐸𝐿 = 660 𝑓𝑚

′ Clay units (f'uc>30 MPa)

𝐸𝑚 = 1000 𝑓𝑚′ 𝐸𝐿 = 500 𝑓𝑚

′ Concrete units (density>1800 kg/m3) and CS units

𝐸𝑚 = 750 𝑓𝑚′ 𝐸𝐿 = 500 𝑓𝑚

′ Concrete units (density<1800 kg/m3)

𝐸𝑚 = 1000 𝑓𝑚′ 𝐸𝐿 = 350 𝑓𝑚

′ Grouted concrete or clay masonry

𝐸𝑚 = 500 𝑓𝑚′ 𝐸𝐿 = 250 𝑓𝑚

′ AAC***

Paulay and Priestly (1992) 𝐸 = 1000 𝑓𝑚′ Masonry

Drysdale et al. (1994) 𝐸 = 210 − 1670 𝑓𝑚′ Masonry

Vermelfoort (2005) 𝐸 = 700 − 750 𝑓𝑚′ Clay brick masonry

Kaushik et al. (2007) 𝐸 = 550 𝑓𝑚′ Clay masonry

Mohamad et al. (2007) 𝐸 = 758 − 1021 𝑓𝑚′ Hollow concrete masonry

Budiwati (2009) 𝐸 = 1000 𝑓𝑚′ Clay and concrete masonry

Costigan (2015)

𝐸 = 230 𝑓𝑚′ PC-lime***

𝐸 = 130 𝑓𝑚′ Hydraulic lime***

𝐸 = 85 𝑓𝑚′ Feebly hydraulic lime***

*Eave is considered in this study **Em=Short-term loading, EL= Long-term loading, average of Em and EL is

considered in this study ***Modern material, therefore, not considered in this study and just for information and

comparison.

Page 110: Reliability Assessment of Historical Masonry Structures

94

Figure 5-5 Cumulative probability of coefficient k.

Considering the coefficient of variation of the compressive strength of masonry obtained by

previous researchers, it can be concluded that the average coefficient of variation is 20% in the

case of both concrete masonry and clay masonry, Table 5-8. According to the research of

Galambos (1982), Schueremans & Van Gemert (1999), Schueremans (2001), Holicky & Markova

(2002), Park et al. (2009), Brehm (2011) and Mojsilovic & Stewart (2014), the best fit distribution

for the compressive strength of masonry (both concrete and clay) is reported to be the log normal

distribution, Table 5-9.

Some research has been done on the estimation of masonry compressive strength and its

variability. Schueremans (2001), Bojsiljkov et al. (2004), Löring (2005), Bojsiljkov & Tomazevic

(2005), Fehling & Stürz (2006a) and Kaushik et al. (2007) worked on clay brick which is common

in historic masonry structure. Calcium silicate and autoclave aerated concrete units are modern

masonry units. There has also been some research focused on the material characteristics of

masonry built with these kinds of modern materials (Jäger & Schöps (2004), Löring (2005),

Fehling & Stürz (2006a), Fehling & Stürz (2006b), Costa (2007), Graubner & Glowienka (2008),

Schermer (2007), Magenes (2007), Höveling et al. (2009) and Gunkler et al. (2009)). A summary

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

500 700 900 1100

Cu

mu

lati

ve

Pro

bab

ilit

y

Coefficient k

Concrete Masonry Clay Masonry

~800 ~890

Page 111: Reliability Assessment of Historical Masonry Structures

95

of the results of these studies, obtained mostly through experimental tests, is presented in Table 5-

10. The data in that table give an idea about the expected range of compressive strength. The

compressive strength values obtained via other methods, such as the ones described above, can

also be compared to the values reported in Table 5-10

Table 5-8 Coefficient of variations of compressive strength reported by different researchers.

Reference CoVmean

Kirtschig & Kasten (1980) 0.17

Galambos et al. (1982) 0.18

Shah & Dong (1984) 0.25

Tschötschel (1989) 0.25

Schueremans (2001) 0.19

Holicky & Markova (2002) 0.20

Graubner & Glowienka (2008) 0.19

Table 5-9 Recommended distributions for compressive strength of masonry (both concrete and

clay) found in the literature.

Reference Distribution Comment

Galambos et al. (1982) LN -

Schueremans & Van Gemert (1999) LN Brick masonry

Schueremans (2001) LN Historical

masonry

Holicky & Markova (2002) LN -

Park et al. (2009) LN URM

Brehm (2011) LN CB, CS, AAC

Mojsilovic & Stewart (2015) LN -

Page 112: Reliability Assessment of Historical Masonry Structures

96

Table 5-10 Values of compressive strength found in the literature (TLM is Thin-Layer Mortar).

Reference Unit Mean

(MPa) COV Description

Kaushik et al. (2007) Clay Brick

4.1 0.24 Weak mortar

7.5 0.18 Strong mortar

6.6 0.2 Intermediate

mortar

Schueremans (2001)

Clay Brick

(For Historic

Masonry)

4.53 0.17 Cores Ø150,

h=300mm

4.26 0.19 Pillars,

h=360mm

6.07 0.093 Wallettes,

h=570mm

4.3 0.137 Cores Ø113,

h=300mm

3.75 - wall,

h=2000mm

Fehling & Stürz (2006a) Clay Brick 6.7 - TLM, M5

Löring (2005) Clay Brick 5.6 - MG IIa

Bojsiljkov & Tomazevic

(2005) Clay Brick 4.6 -

M5, M6, M7,

M15, LM5

Bojsiljkov et al. (2004) Clay Brick 5.45 - M5, TLM

Fehling & Stürz (2006a) Calcium Silicate Wall 15 - TLM

Schermer (2007) Calcium Silicate Wall 15 - TLM

Magenes (2007) Calcium Silicate Wall 15 - TLM

Löring (2005) Calcium Silicate Wall 15 - TLM

Jäger & Schöps (2004) Calcium Silicate Wall 7.3 - TLM

Gunkler et al. (2009) Calcium Silicate Wall 20.6 - TLM

Page 113: Reliability Assessment of Historical Masonry Structures

97

Reference Unit Mean

(MPa) COV Description

Fehling & Stürz (2006b)

Autoclave aerated

concrete

(unit-mortar

combination 4/TLM)

2.6 - TLM

Löring (2005) Autoclave aerated

concrete 2.3 - TLM

Costa (2007) Autoclave aerated

concrete 2.4 - TLM

Jäger & Schöps (2004) Autoclave aerated

concrete 3.3 - TLM

Höveling et al. (2009) Autoclave aerated

concrete 4.26 - TLM

Graubner & Glowienka (2008)

Calcium Silicate

13.5

0.16

Class 16

18.8 Class 20

27 Class 28

Autoclave aerated

concrete

2.4

0.14

Class 2

4.1 Class 4

5.5 Class 6

Page 114: Reliability Assessment of Historical Masonry Structures

98

Understanding the shear failure of masonry structures can be important in order to evaluate

structural resistance to lateral loads (e.g. wind and earthquake). It is difficult to estimate the shear

strength without large scale testing. However, in-situ testing techniques applied to single masonry

units can provide comparable information to full scale ones. These in-situ testing methods are also

less destructive and more economical and consequently better than full scale testing in the case of

historic structures. Quite a few testing methods have been reported for estimation of the shear

strength of masonry. These methods are in situ testing techniques. An early one, named the “push

test”, was developed to be used in the evaluation of the seismic resistance of older unreinforced

masonry structures (ABK, 1984). In this method, a brick unit and its associated head joints (located

on opposite ends of the determined test unit) are removed. The next step is to place a hydraulic

ram in the wall and displace the wall laterally. The mortar bed joints located straight below and

above the designated test unit are sheared. This method is only appropriate for masonry structures

or elements having strong units and weak mortar. In such structural systems, shear cracks develop

in a stair-step pattern through mortar joints or bed joints slide over each other while the units

remain uncracked. In other words, bed joint sliding or stair-step cracking diagonally along mortar

joints dominate the shear failure. The shear strength of the mortar joints is then related to the shear

strength of the masonry system using empirical formulae. This method is not applicable for new,

modern masonry structures having strong mortar, but can be applied to historical structures as they

mostly have weak mortar. It should be noted however, that although application of the push test is

possible for historical structures from a scientific and theoretical point of view, it leads to

destruction and interference of historical systems which may be expensive to repair or even be

irreparable.

JCSS (2011) has not reported any direct approaches for estimation of the shear strength of

masonry. The committee introduced the tensile strength of the units (i.e. shear failure) and the

cohesion between units and mortar (i.e. sliding failure) as the two parameters determining the shear

strength of masonry walls. The presentation of a probabilistic model for estimation of both the

tensile strength and cohesion by JCSS (2011) may be the reason of not reporting direct approaches

associated with the estimation of shear strength of masonry.

Page 115: Reliability Assessment of Historical Masonry Structures

99

A new American Society for Testing and Materials International (ASTM) standard presented three

semi- destructive test methods regarding the estimation of shear strength. The mortar joint shear

strength index is calculated by dividing the maximum recorded horizontal force by the gross area

of the upper and lower bed joints. One of these methods is the conventional approach with a

hydraulic ram as explained above. The second method involves flatjacks in order to apply

compression to the test location. The shear strength index and the coefficient of friction are the

outcomes of this method. The relationship between shear strength thickness and applied

compressive stress is obtained by doing the shear test while the compressive stress is increased.

Moreover, the coefficient of friction is the slope of the best-fit line passing through the measured

joint shear strength index and the applied compressive stress plotted versus each other. As the

friction coefficient is measured in this method instead of assumed, this method would lead to more

accurate results in terms of joint shear strength index. The other two testing procedures described

by ASTM assume a value of friction coefficient in their strength calculation procedures. The last

method is also an in-situ test which can be performed using special flatjacks. By pressurizing the

masonry horizontally between two flatjacks and monitoring the resultant deformation, the shear

strength of the masonry can be estimated. Using flatjacks rather than a hydraulic ram results in a

test that does not require the removal of masonry units.

Flatjack tests have a minor destructive effect on the masonry compared to the conventional

approach using a hydraulic hammer, so they may be more appropriate for masonry structures in

terms of destructivity. However, estimation of strength variability with flatjack tests is arguably

destructive as that would require making a large number of slots. Moreover, the flatjack method

has not shown good accuracy in cases of deep, multi-wythe and multi-layer structures as the

bladders cannot penetrate the masonry sufficiently to get the desired information (as mentioned

earlier). Therefore, there is a lack of an appropriate test method for estimation of the variability of

the shear strength of historic masonry structures required to be able to determine their reliability

under the vertical loading. Consequently, there is a potential area of research here. Here, shear

strength is proposed to be considered deterministic and to be estimated by few applications of

flatjack technique (as semi-destructive testing method) to restrict the potential destructions of

variability estimation.

Page 116: Reliability Assessment of Historical Masonry Structures

100

Quite a few research studies have been directed to the estimation of the tensile strength of historic

masonry. This may be due to the fact that the tensile strength of masonry perpendicular to the bed

joints has been proved to be negligible. Therefore, it is usually considered to be zero in the

calculation of masonry resistance. However, in some cases (e.g. slender walls) where the flexural

tensile strength may significantly influence the structural resistance, there may be a necessity to

include a probabilistic model of tensile strength in reliability analysis.

The bond-wrench test can be used to estimate the tensile strength of historic masonry, although

the test is semi-destructive. In this method, the units are removed to get the instrument in place,

wrench off some units to get the bond strength and then mortar and all units are returned to their

original places. If the test does not lead to the damage to the prims under the test, it considers semi-

destructive. However, there are limitations in the repetitive application of the wrench test to

estimate the variability of tensile strength over a historic masonry system due to the destructivity.

In other words, in order to understand how the tensile strength changes over a historic masonry

system, the wrench test should be repeated many times which leads to destruction of the historic

structure. Therefore, it is recommended to neglect tensile strength of structural members which

their resistances are not sensitive to tensile strength. However, the member which its resistance is

sensitive to tensile strength, can be tested by wrench test as it is semi-destructive (minor

destruction).

Schubert (2010) worked on the flexural tensile strength of masonry perpendicular and parallel to

bed joints. As the cohesion of thin layer mortar (TLM) is reliable, masonry with TLM was

considered in this research. It was showed that there is considerable scatter in this property and the

scatter is mostly dependent on the characteristics of the mortar. The type of head joint (filled or

unfilled) was shown to have only a small influence on the flexural tensile strength of the masonry.

The flexural tensile strengths of masonry parallel and perpendicular to the bed joints are

summarized in Table 5-11.

Page 117: Reliability Assessment of Historical Masonry Structures

101

Table 5-11 Flexural tensile strength of masonry (Schubert, 2010).

Unit Type Head Joint Number

of Tests

Mean of 𝑓𝑡

(𝑁/𝑚𝑚2)

Range of measured 𝑓𝑡

(𝑁/𝑚𝑚2)

Parallel to bed joints

CB[1] Un-filled 2 0.21 0.2, 0.22

CS[2]

non-perforated Filled 8 0.75 0.36 – 1.14

Un-filled 11 0.71 0.38 – 0.97

perforated Filled 4 0.48 0.45 – 0.51

Un-filled 4 0.25 0.29 – 0.35

ACC[3] Filled 10 0.43 0.22 – 0.64

Un-filled 6 0.2 0.16 – 0.24

Perpendicular to bed joints

CB Un-filled 3 0.28 0.26 – 0.3

CS

perforated Filled &

Un-filled 8 0.56 0.35 – 0.73

non-perforated Filled 4 0.34 0.23 – 0.48

ACC Filled &

Un-filled 23 0.4 0.25 – 0.81

Concrete Blocks Filled 5 0.33 0.22 – 0.44

[1] Clay Brick, [2] Calcium Silicate, [3] Autoclave Aerated Concrete

JCSS (2011) categorized tensile strength into longitudinal tensile strength and splitting tensile

strength and provided a probabilistic model for the tensile strength of the units as follows:

Tensile strength perpendicular to units (longitudinal tensile strength)

Page 118: Reliability Assessment of Historical Masonry Structures

102

𝑓𝑏𝑡,𝑙 = 𝑐2. 𝑓𝑏 (5-5)

Tensile strength parallel to the units (splitting tensile strength)

𝑓𝑏𝑡,𝑠 = 𝑐3. 𝑓𝑏 (5-6)

where 𝑓𝑏𝑡,𝑙 and 𝑓𝑏𝑡,𝑠 are mean of tensile strength perpendicular to units and parallel to the units,

respectively. 𝑐2 and 𝑐3 are ratios with values for different materials and perforation being presented

in detail in JCSS (2011). JCSS (2011) does not describe shear strength individually. Therefore,

although the splitting tensile strength involves shear strength in concept, it is mentioned under the

category of tensile strength based on the JCSS (2011). The tensile strength of the units has a strong

effect on the shear capacity of the masonry walls. Therefore, JCSS (2011) just defined probabilistic

models for the tensile strength of different kinds of masonry units in the direction dominating the

shear strength of the masonry. In the case of calcium silicate units and autoclaved aerated concrete

units, the shear capacity is dominated by the tensile strength in the longitudinal direction whereas

the splitting tensile strength describes the shear capacity of clay brick units. Therefore,

probabilistic modeling of the tensile strength of calcium silicate units and autoclaved aerated

concrete units is suggested only in terms of the longitudinal direction and the tensile strength of

clay brick units is recommended for the perpendicular direction as follows

Tensile strength in the longitudinal direction of calcium silicate units and autoclaved aerated

concrete units

𝑓𝑏𝑡,𝑙,𝑗 = 𝑌. 𝑓𝑏𝑡,𝑙 (5-7)

Splitting tensile strength of clay brick units

𝑓𝑏𝑡,𝑠,𝑗 = 𝑌. 𝑓𝑏𝑡,𝑠 (5-8)

Page 119: Reliability Assessment of Historical Masonry Structures

103

where 𝑌 is a lognormal variable, see Table 5-12. The large values of CoV reveal that the tensile

strength of the units shows large scatter.

Table 5-12 Statistical characteristics of 𝑌 involved in probabilistic model of tensile strength

(JCSS, 2011).

Unit Type CoV Mean Distribution

CS[1] 0.26 1.0 Log-normal

AAC[2] 0.16 1.0 Log-normal

CB[3] 0.24 1.0 Log-normal

[1] Calcium Silicate, [2] Autoclave Aerated Concrete, [3] Clay Brick

Li et al. (2016) reported mean values and the CoV for both the tensile strength of mortar and the

tensile strength of brick. The mean and CoV of the tensile strength of mortar were reported to be

1.21 and 0.30, respectively and the mean and CoV of brick tensile strength were reported to be

3.13 and 0.34, respectively. Moreover, a lognormal distribution was recommended to be the best-

fit distribution for the tensile strength of the mortar while the Weibull distribution was considered

to describe best the distribution of the tensile strength of brick. Li et al. (2016) did not specifically

define the type of unit and mortar that they used in their studies. Therefore, there may be

inaccuracy if the reported values are generalized for other brick masonry structures.

All of the studies mentioned above and the findings regarding the tensile strength of the masonry

(Schubert, 2010; JCSS, 2011, Li et al., 2016) are focused on new masonry materials (units and

mortar) and may not be accurate for historic masonry materials. Therefore, there is still a gap in

knowledge regarding the tensile strength of historic masonry and its variability in order to do

reliability analysis accurately, specifically in the case of elements in which tensile strength plays

a key role in the determination of load bearing capacity. In other historic masonry systems, where

tensile strength does not have significant influence and as mostly historical structures are

unreinforced, the tensile strength of the masonry and its variability (due to its brittle nature) can

be neglected in reliability analysis.

Page 120: Reliability Assessment of Historical Masonry Structures

104

Cohesion and friction coefficient are the two parameters that influence the shear strength of the

masonry. However, no study could be found focusing on the evaluation of the stochastic

characteristics of the cohesion and friction coefficient of historic masonry structures.

JCSS (2011) only presents the probabilistic models related to cohesion and friction coefficient

considering new masonry materials which may not be accurate for reliability analysis of historic

masonry materials. The JCSS (2011) recommended probabilistic models for both cohesion and

friction coefficient are presented as follows:

- Cohesion

𝑓𝑣,𝑗 = 𝑌. 𝑓𝑣,𝑚 (5-9)

where 𝑓𝑣,𝑚 is the mean of cohesion and 𝑌 is a random variable according to Table 5-13.

- Friction coefficient

𝜇𝑗 = 𝑌. 𝜇𝑚

(5-10)

where a mean value of 0.8 is normally considered for the friction coefficient and 𝑌 is a random

variable given in Table 5-14.

Table 5-13 Statistical characteristics of 𝑌 involved in probabilistic model of cohesion (JCSS,

2011).

Unit Type Mean CoV Distribution

CS[1] 1.0 0.35 Log-normal

AAC[2] 1.0 0.35 Log-normal

CB[3] 1.0 0.40 Log-normal

[1] Calcium Silicate, [2] Autoclave Aerated Concrete, [3] Clay Brick

Page 121: Reliability Assessment of Historical Masonry Structures

105

Table 5-14 Statistical characteristics of 𝑌 involved in probabilistic model of friction coefficient

(JCSS, 2011).

Unit Type Mean CoV Distribution

CS[1] 1.0 0.19 Log-normal

AAC[2] 1.0 0.19 Log-normal

[1] Calcium Silicate, [2] Autoclave Aerated Concrete,

Li et al. (2016) reported the mean, CoV and best-fit distribution associated with the cohesion of

unreinforced masonry constructed with an extruded brick. They described neither the type of brick

nor the type of mortar used in their studies. The Mean and CoV of the cohesion were reported to

be 1.81 and 0.27, respectively. Moreover, the lognormal distribution was considered for this

material property. However, as they did not mention exactly the characteristics of the materials

that were studied, it may not be accurate to generalize the reported values to other unreinforced

brick masonry structures, like historic ones.

Given the information above, there is a gap in knowledge regarding the probabilistic characteristics

of cohesion and friction coefficient of masonry structures especially historical masonry structures.

Page 122: Reliability Assessment of Historical Masonry Structures

106

Chapter VI: Reliability Assessment Methodology

Historic masonry structures are constructed with various masonry materials and are mostly

unreinforced. Therefore, they have limited load bearing capacity, making them vulnerable to the

applied loading conditions. Assessing the reliability of their performance under applied loads is of

great importance. If there is a need to upgrade and strengthen historic structural systems to preserve

them, minimum interventions have to be applied to save their historical value. Reliability

assessment can therefore act as a useful tool to determine the level of required upgrading leading

to minimization of the intervention. Although there is need to analyze the reliability of a historic

masonry structure, no formal methodology has been proposed in this regard. In this chapter, an

attempt is made to initiate the development of a step-by-step methodology appropriate for

reliability assessment of historic masonry structural members considering the specific limitations

and criteria which exist in this regard. This is the first step before full structural reliability could

be determined. There is no doubt that there are still significant gaps in knowledge and limitations

in the development of a comprehensive and accurate reliability assessment methodology for

historical masonry structures. Investigation of all the limitations needs more time and research

beyond the scope of this thesis.

A general layout of the evaluation procedure of existing structures is presented as a flowchart by

Figure 3-1 in Chapter 3. To evaluate the structure, the process in the flowchart should be followed.

This evaluation involves different steps from the preliminary evaluation to the detailed one.

To start the structural evaluation, the first step is to gather information about the structure. This

includes investigation of documents, an initial inspection and initial check decisions on immediate

plans and suggestions for detailed evaluation. One of the main objectives of restoration of

historical structures is conservation. Therefore, a careful analysis of the structure is required.

Page 123: Reliability Assessment of Historical Masonry Structures

107

Specific and comprehensive information about the structure’s materials and techniques as well as

its condition need to be gathered. Architectural surveying is the main step of direct investigation

which includes geometric and morphological measurements, materials survey, and

photogrammetry. The next step of data gathering is historical investigation. Critical reading and

understanding of bibliographies of iconographies and available documents are the main parts of

historical investigation. Historical documents can provide useful information about material

properties and construction methods. The integration of the knowledge and information obtained

leads to a more precise and comprehensive understanding of the structure under evaluation. The

more carefully the primary evaluation is performed, the more accurate is the assessment achieved.

In the case of historical structures, as there are typically almost no documents available, a field

inspection and direct investigation are of special importance.

When useful information is gathered from direct and historical investigation and study, a detailed

reliability assessment may be required. The flowchart (Figure 3-1) does not include the detailed

steps which need to be followed in order to do reliability assessment. Therefore, an attempt is made

here to review the process of reliability assessment (Figure 6-1) of a historic structure considering

the previously discussed parameters. Here the focus is on reliability assessment of each member

individually. System reliability assessment can provide more accurate evaluation since it considers

the probability of failure as the probability of failure of whole system including the individual

members. However, connection behaviour and strength are also considered in system reliability

which are influential in structural resistance of historic masonry structures. System reliability

assessment is thus highly complex in a new structure, but especially so in the case of historic

masonry structures because there is no knowledge and information about connections between the

different systems and materials typically found in such a structure: system (whole structure)

reliability is thus an area ripe for further research and is outside the scope of this thesis.

The basis of reliability assessment is the limit state function, being the resistance minus the load

effect, as discussed in Chapter 3. Codes of practice recommend formulas to estimate the required

resistance and load effects (defined as R and L in Chapter 3) for different structural members under

different failure scenarios. There is general acceptance in Canada that if historic structures can

satisfy 60% of the structural resistance required in the new NBCC, they are considered to be in a

Page 124: Reliability Assessment of Historical Masonry Structures

108

safe condition. Therefore, the limit state function can be formulated for each historic structural

member to be 60% of required resistance minus the load effect recommended by NBCC code.

Material properties play a key role in the determination of structural resistance. CSA S304-14

recommends expressions for estimating the resistance of a masonry member with respect to

different failure scenarios, which include general material properties. Compressive strength, shear

strength, tensile strength, cohesion and friction coefficient are material properties which are

considered in determination of structural resistance. The stochastic characteristics of the material

properties of historic masonry structures are recommended to be estimated directly and specifically

for each structural element under evaluation due to their specific nature and characteristics using

field testing techniques (reasons were discussed in detail earlier). The available approaches,

concerns and recommendations associated with the estimation of the probabilistic models of these

material properties are discussed in detail in Chapter 5. As there is no non-destructive method to

estimate the shear strength, tensile strength, cohesion and friction coefficient more than one

approach is recommended for them here, see Figure 6-2. Shear strength is recommended to be

estimated using flatjack. The estimated value of shear strength can be considered either

deterministic to minimize the destruction caused by repeating the test or as a test data for Bayes’

theorem to update prior estimation. Tensile strength is proposed to be neglected in the case of an

ordinary member (non-sensitive to tensile strength). In the case of members which their resistance

is sensitive to their tensile strength, 4 approaches are recommended, including consideration of

tensile strength as deterministic; the use of the semi-destructive wrench test, considering the result

of wrench test as a test data for Bayes’ theorem to update available data and expand the database,

considering the results of the wrench test as a mean and using the probabilistic model of JCSS

(2011); and lastly, considering the probabilistic model of JCSS (2011) without modification.

Cohesion and the friction coefficient are also recommended to be modeled either deterministically

using recommended values, or by using the JCSS (2011) probabilistic model. It should be noted

that each approach adds its related error to the reliability assessment.

In order to determine the load effect, the probabilistic model of applied load needs to be known.

The stochastic characteristics of dead load, live load, wind load and snow load are described in

detail in Chapter 4.

Page 125: Reliability Assessment of Historical Masonry Structures

109

Knowing the limit state function and the probabilistic models of the material properties and applied

loads, the reliability index needs to be defined in next step. There are several structural reliability

techniques which are briefly explained in Chapter 3. Asymptotic techniques and simulation based

approaches are two reliability methods that have been developed and are commonly used in

structural reliability assessment. The selection of the appropriate structural reliability technique

depends several factors like time concerns, required accuracy and the limit state function.

The final step is to compare the calculated reliability index or probability of failure with the target

reliability index as a decision criterion. As discussed in Chapter 3, the necessity of meeting the

target reliability index and target probability of failure as recommended by codes of practice (for

new structures) is a controversial issue in reliability assessment of historic masonry structures. The

target probability of failure and target reliability index values reported for masonry structures as

well as some formulae for the calculation of the targets considering specific criteria associated

with historic masonry structures are reported in Chapter 3.

The process of reliability assessment structures is summarized in Figure 6-1 (see the example of

the application of the method presented in Appendix). Although the process of reliability

assessment of structures is the same in general, there are some parameters which should be

considered in reliability assessment of historic masonry structures which make the reliability

assessment procedure more specific. The necessity of considering a specific target reliability index

and target probability of failure as decision criteria, the difficulty and restriction in estimating the

material properties and related variabilities for estimating their probabilistic models and the

likelihood of inaccuracy in the available limit state functions for describing the behaviour of

historic structures are the most challenging parameters necessitating the development of a specific

methodology to assess the reliability of historic masonry structures. The first two parameters are

discussed in this thesis. However, there is still plenty of room for development. Figure 6-2 shows

the recommended approaches for estimation of the probabilistic models of historic masonry

materials.

Page 126: Reliability Assessment of Historical Masonry Structures

110

Figure 6-1 Flowchart for reliability assessment of structures.

Reliability Assessment of Historical Masonry Structures

Data Collection

(Direct and Historical Investigation)

Probabilistic Model of

Historic Masonry Material

Probabilistic Model of

Applied Loads Limit State

𝑔(𝑥) = 𝑅 − 𝐿

Structural Reliability Techniques

𝛽 & 𝑃𝑓

𝛽 > 𝛽𝑇

𝑃𝑓 < 𝑃𝑓𝑇

𝛽 < 𝛽𝑇

𝑃𝑓 > 𝑃𝑓𝑇

Reliable Non-reliable

Page 127: Reliability Assessment of Historical Masonry Structures

111

Figure 6-2 Recommended approaches for estimation of the probabilistic models of historic

masonry materials.

Wrench Test, Deterministic

Wrench Test, Mean JCSS, Distribution and CoV

JCSS Probabilistic Model

Modulus of

Elasticity Compressive

Strength Shear

Strength Tensile

Strength

Cohesion & Friction Coefficient

Deterministic,

Recommended

values

JCSS Model

Non-sensitive to

Tensile Strength

Negligible, 0

Sensitive to

Tensile Strength

UPM E= k.f'm

Clay

Masonry Concrete

Masonry

k=700-900 Median: 890

k=700-900 Median: 800

Flatjack,

Deterministic

Probabilistic Model of

Historic Masonry Material

Flatjack, Bayes’ theorem

Wrench Test, Bayes’ theorem

Page 128: Reliability Assessment of Historical Masonry Structures

112

The required resistance of historic masonry structures in the evaluation stage is thought to be 60%

of the required resistance in the design stage recommended by NBCC. This thought is unwritten

and no scientific base or detailed analysis and investigation has been reported for it. As formulation

of the limit state function is a key parameter in the outcome of reliability assessment, acceptance

and application of this general belief with respect to satisfactory structural resistance may result in

an inaccurate reliability index or probability of failure.

Moreover, the intent of formulae to calculate the resistance presented by codes of practice is

accepted here to satisfy safety. The appropriateness of these formulae for the evaluation of historic

masonry structures however, is a controversial issue as they are determined for new masonry

materials and constructions. Some research is recommended to be done to verify the

appropriateness of these formulae for historic masonry structures having specific characteristics.

Limit state functions formulated by these recommended expressions should be investigated in

terms of their accuracy in the estimation of reliability of historic structures.

Many researchers have investigated the relationship between compressive strength and modulus

of elasticity and have reported different relationships between these two material properties.

Nevertheless, this is the basis of the recommendation in this study regarding the calculation of

compressive strength, as discussed in Chapter 5. As different coefficients have been suggested for

this relationship, the median values of the coefficient are considered here as the final recommended

coefficient. However, the median value may not be the real representation of this coefficient which

may result in inaccurate estimation of compressive strength and any property derived from the

compressive strength. A study is therefore recommended to define the most accurate coefficient.

Moreover, no information was found regarding the relationship of compressive strength and

modulus of elasticity in the case of stone masonry. To be able to generalize the approach, this

relationship needs to be defined for stone masonry in its various forms.

The stochastic characteristics of some material properties of historic masonry structures cannot be

estimated directly due to the lack of non-destructive approaches. Therefore, deterministic values

of them or probabilistic models recommended for new masonry materials are used. To achieve a

Page 129: Reliability Assessment of Historical Masonry Structures

113

more accurate reliability index or probability of failure, there is a need to develop non-destructive

testing approaches to estimate the probabilistic models of relevant properties accurately.

Page 130: Reliability Assessment of Historical Masonry Structures

114

Chapter VII: Conclusion

Historical masonry structures should be preserved for future generations. Despite the high level of

vulnerability of unreinforced masonry structures under applied loads and the importance of their

reliability evaluation, there is no formal methodology to assess the reliability of historic masonry

structures. In this study, an attempt is made to initiate development of a step-by-step methodology

for assessing the reliability level of historic masonry structures. Reliability assessment requires the

limit state function to be formulated, being the structural resistance minus the load effect.

Therefore, in order to develop an appropriate determinate methodology, estimations of

probabilistic models of structural resistance and load effects are required to formulate a limit state

function. Moreover, a target reliability appropriate for historical structures has to be determined.

Currently, there is no formal method for reliability assessment of historical masonry structures.

This thesis contributes to the development of a reliability-based assessment methodology for

historical masonry structures as follows:

A comprehensive literature review was done on different aspects of reliability assessment of

masonry structures, specifically historical masonry ones, including the theory of reliability

assessment of structures, limit states, reliability assessment techniques and suitable combination

among finite element structural analysis, reliability analysis methods and reliability software.

Reliability assessment of historic masonry structures necessitate determination of target reliability

index or target probability of failure as decision criteria. Codes of practice recommended different

values as target probability of failure or target reliability index. Regarding historical structures, the

necessity of meeting the target probability of failure values or target reliability index presented by

different codes of practice is still a controversial issue. This is due to that fact that the

recommended values are reported considering new materials, structural configurations and

construction methodologies, not historical ones, with specific criteria and requirements. In this

thesis, different calculation approaches and values associated with target reliability index and

Page 131: Reliability Assessment of Historical Masonry Structures

115

failure probability being appropriate for historic structures are described. Use of one formula is

recommended.

Deterioration is a common process in historic structures which may result in a change in masonry

resistance, the loading conditions or analysis model. Appropriate approaches for integration of the

deterioration into reliability assessment were studied and degradation functions were reported.

Reliability assessment of structures under seismic loading is a challenging and complex field of

science as involves different uncertainties including those in the nature of ground motions and

those in the nonlinear behaviour of structures. The specific methodology for reliability assessment

of existing structures under seismic loading was reviewed.

The stochastic modeling of applied loads as one of the key parameters in the determination of the

limit state function including dead load, live load, wind load and snow load were reviewed.

Different literature- and code-recommended probabilistic models and stochastic characteristics for

applied loads were studied and compared. Statistical characteristics of point-in-time components

of live load, wind load and snow load were also investigated. Finally, the best fit probabilistic

models for different load effects were presented.

The stochastic characteristics of construction materials play key roles in the determination of

probabilistic models of structural resistance. There are various testing techniques which can be

used to estimate stochastic characteristics of historical masonry materials. Destructive testing of a

historic masonry structure or its components to get more realistic information about the material

properties is not recommended as such tests may lead to irreparable damage to these valuable

structures. Moreover, some of testing techniques are not capable of estimating the variability of

material properties over an element or structural system, this being essential for reliability

assessment. A summary of testing techniques being used for the evaluation of masonry materials

properties was presented. The strengths and weaknesses of each technique as well as its

applicability in estimation of stochastic characteristics of historic masonry materials properties

were investigated. The calculation procedures, values and the best fit distributions recommended

by several codes of practice regarding the estimation of the statistical characteristics of masonry

material properties were presented. As code recommendations are mostly for new masonry

materials, the accuracy, applicability and accordance of them with historical masonry materials

were discussed.

Page 132: Reliability Assessment of Historical Masonry Structures

116

Compressive strength is the most important material property influencing the structural resistance

of historic masonry systems and consequently affects the reliability assessment of these structures.

Codes are necessarily conservative and are also generally aimed at design or assessment with

modern masonry materials, so the use of code values for historical structures may lead to inaccurate

reliability assessment. There are various testing techniques which can be used in order to estimate

the mean value of compressive strength related to existing masonry structures. Testing techniques

which damage the structure (for example, by removing samples of the masonry) are not

recommended for historical buildings because of their destructive nature. Moreover, these

destructive techniques are also restricted in their ability to estimate the variability of compressive

strength over an element or structural system. Here, a proposal is made on how to determine a

probabilistic model of the masonry compressive strength from non-destructive tests.

This research focused on proposing a methodology for the estimation of a probabilistic model of

compressive strength of historic masonry structures, as the most important material characteristic

influencing the structural resistance. The idea was to get more reliable information directly without

any disturbance to the historical value of the structure. However, there are other material properties

which are influential on the structural resistance in some cases (e.g. flexural tensile strength in the

case of slender walls). There is still no procedure for estimating the stochastic characteristics of

these material properties without any disturbance to the historic masonry. Further work is needed

to assess the statistical characteristics of these material properties.

In this thesis, the procedure for estimating compressive strength was based on the relationship

between E as modulus of elasticity and 𝑓𝑚′ as compressive strength. No information was found

regarding this relationship for stone masonry. The focus has been on the development of

methodology for estimating of compressive strength of concrete masonry and clay masonry –

frequently used modern masonry materials. Therefore, determining the relationship between E and

𝑓𝑚′ for stone masonry is obviously an area where work needs to be done to aid in the estimation of

the reliability of historic masonry structures.

Page 133: Reliability Assessment of Historical Masonry Structures

117

The accuracy and applicability of the proposed methodology regarding the estimation of the

stochastic characteristics of the compressive strength of historical masonry structures should be

evaluated through several experimental tests.

In order to do reliability assessment, a limit state function needs to be determined being structural

resistance minus load effect. Codes and standards recommend different formulas to calculate

structural resistance associated with different failure modes. The capability of these formulae to

be an accurate representation of the structural resistance of historical structures is still a

controversial issue. This is due to that fact that the recommended expressions are obtained

considering new materials, structural configurations and construction methodologies rather than

historical ones with specific criteria and requirements. Therefore, more research should be done in

this area.

The resistance and behaviour of connections play a key role in the resistance and behaviour of the

structural systems especially in case of historic masonry structures which may have weak

connections. This parameter should be considered in future research using system reliability

assessment.

Page 134: Reliability Assessment of Historical Masonry Structures

118

References

Abdelatif H, Sh.S, El-Meligy M, Sidky AN. 1999. Effect of Openings on Safety of Flat Plates.

Proceedings of 3rd International Conference on Civil & Arch. Eng. March 9-11; Military

Technical College, Cairo, Egypt. pp. 111-124.

Abdelatif H, Sh.S, El-Meligy M. 1998. Reliability-based Analysis of 3D Multi Story Structure.

Scientific Bulletin, Journal of Engineering, Ain Shams Univ., Cairo, Egypt, 33(1): 51-70.

ABK .1984. Methodology for mitigation of seismic hazards in existing unreinforced masonry

buildings. Topical Report 8: The Methodology. Contract NSF-C-PFR78-19200, US National

Science Foundation, Washington, DC.

ACI. 1999. Building code requirements for masonry structure. Committee 530, Farmington Hills,

MI: American Concrete Institute.

Al-Habahbeh OM, Stewart MG. 2015. Stochastic reliability of unreinforced masonry walls

subjected to blast. The Sixth Jordanian International Civil Engineering Conference

(JICEC06). 2015 March 10-12; Amman-Jordan.

Allen DE. 1975. Limit states design — a probabilistic study. Canadian Journal of Civil

Engineering, 2: 36–49.

Allen DE. 1991b. Criteria for design safety factors and quality assurance expenditure. Structural

Safety and Reliability. Elsevier Amsterdam, pp. 667-678.

Allen, DE. 1991a. Limit states criteria for structural evaluation of existing buildings. Canadian

Journal of Civil Engineering. 18(6): 995–1004.

Ang HSA, and Tang WH. 1975. Probability Concepts in Engineering Design. Volume I-Basic

Principles. John Wiley & Sons, Inc.

ANSI A58.1. 1982. American National Standard Minimum Design Loads in Buildings and Other

Structures. American National Standard Institute. New York.

Appa Rao G. 2001. Generalization of Abrams' law for cement mortars. Cement and Concrete

Research. 31(3): 495-502.

AS-3700. 1998. Masonry structures. Australian standard. Australia.

ASCE7-98. 2000. Minimum design loads for buildings and other structures. American Society of

Civil Engineers (ASCE).

ASTM C780-02. 2002. Standard Test Method for Preconstruction and Construction Evaluation of

Mortars for Plain and Reinforced Unit Masonry. ASTM International.

Page 135: Reliability Assessment of Historical Masonry Structures

119

Baldioli A. 1998. Quantitative and qualitative criteria for priorities in conservation: Methods,

feelings and experience [Dissertation]. R. Lemaire Centre for the Conservation of the

historical towns and buildings.

Bartlett FM, Hong HP, Zhou W. 2003. Load factor calibration for the proposed 2005 edition of

the National Building Code of Canada: Companion-action load combinations. Canadian

Journal of Civil Engineering. 30(2): 440–448.

Bennett RM, Boyd KA, Flanagan RD. 1997. Compressive properties of structural clay tile prisms.

Journal of Structural Engineering: 123(7): 920–926.

Bojsiljkov V, Tomazevic M, Lutman M. 2004. Optimization of shape of masonry units and

technology of construction for earthquake resistant masonry buildings. Research report.

University of Ljubljana.

Books HSE. 2001. Probabilistic Methods: uses and abuses in structural integrity. Contract

Research Report 398. Prepared by BOMEL Limited for the Health and Safety Executive.

Brehm E. 2011. Reliability of unreinforced masonry bracing walls - probabilistic approach and

optimized target values [Dissertation]. Darmstadt. TU Darmstadt.

Brehm Eric, Lissel Shelley. 2012. Reliability of Unreinforced Masonry Bracing Walls. 15th

International Brick and Block Masonry Conference. Florianópolis. Brazil.

Breitung K. 1984. Asymptotic Approximations for multi-normal integrals. Journal of Engineering

Mechanics, ASCE, 110 (3): 357-366.

Breitung K. 1994. Asymptotic Approximations for Probability Integrals. Lecture Notes in

Mathematics (#1592). Springer Verlag. Berlin.

Breitung KW. 1991. Probability Approximations by Log Likelihood Maximization. Journal of

Engineering Mechanics, ASCE, 117 (3): 457-477.

Brozovsky J. 2013. Use of non-destructive ultrasonic pulse testing methods in evaluation of brick

parameters. In Proceedings of the 3rd International Conference on Sustainable Construction

Materials & Technologies (SCMT’13), pp. 1-10.

BS 5628. 1978. Unreinforced Masonry. British Standard: Part 1.

Budiwati IAM. 2009. Experimental compressive strength and modulus of elasticity of masonry.

Jurnal Ilmiah Teknik Sipil, 13(1): 71-81.

Casas Joan R. 2011. Reliability-based assessment of masonry arch bridges. Construction and

Building Materials 25.4, pp. 1621-1631.

Page 136: Reliability Assessment of Historical Masonry Structures

120

CEB. 1976. Common unified Rules for Different Types of Construction and Materials. Vol. 1.

CEB Bulletin No. 116.

CEB. 1978. Recommended Target probabilities of failure. Comunnauté Europeënne du Beton.

CEB model code.

CIB W81. 1989. Actions on Structures - Live Loads in Buildings. Conseil International du

Bâtiment pour la Recherche l'Etude et la Documentation (CIB). Report 116. Rotterdam.

CIRIA. 1977. Rationalisation of Safety and Serviceability factors in Structural Codes. Report No.

63. Construction Industry Research and Information Association. London.

Colbeck SC. 1977. Roof loads resulting from rain-on-snow; results of a physical model. Can. J.

Civil Eng., 4: 82-490.

Cornell CA. 1967. Bounds on the Reliability of Structural Systems. Journal of the Structural

Division (ASCE), 93 (ST1): 171-200.

Cornell CA. 1969. A Probability-based structural code. Journal of ACI, 66 (12): 974-985.

Costa A. 2007. Experimental testing of lateral capacity of masonry piers [Dissertation]. University

of Pavia.

Costigan A, Pavía S, Kinnane O. 2015. An experimental evaluation of prediction models for the

mechanical behavior of unreinforced, lime-mortar masonry under compression. Journal of

Building Engineering. 4: 283–294.

CSA S304. 2014. Design of Masonry Structures, Standard, Canadian Standards Association.

Ontario, Canada.

CSA S304.1. 2004. Design of masonry structures. Canadian Standards Association (CSA) Ontario,

Canada.

CSS. 2001. Probabilistic model code. Zurich: Joint Committee on Structural Safety.

Cunha EH, Guimarães GN, Carasek H. 2001. Influência do tipo de argamassanaresistência à

compressão da alvenariaestrutural. In: Anais do IV SimpósioBrasileiro de Tecnologia de

Argamassas. Brasília, pp. 397-408.

D’Ayala D, Speranza E. 2003. Definition of collapse mechanisms and seismic vulnerability of

historic masonry buildings. Earthquake Spectra. 19(3): 479-509.

Davenport AG. 1981. Some attributes of the wind load requirements of the National Building Code

of Canada. In Proceedings of the Canadian Conference on Wind Engineering. 13–14 April;

Vancouver, B.C, pp. 1–12.

Page 137: Reliability Assessment of Historical Masonry Structures

121

Davenport A.G. 1982. On the assessment of the reliability of wind loading on low buildings. In

Proceedings of the 5th Colloquium on Industrial Aerodynamics. 14–16 June; Aachen,

Germany.

Davenport AG. 1987. Proposed new international (ISO) wind load standard. High winds and

building codes. Proceedings of the WERC/NSF Wind engineering symposium. Kansas City,

Missouri, pp. 373-388.

Davenport, AG. 2000. A comparison of seismic and windstorm hazards. In Proceedings of the 6th

Environmental Specialty Conference of the Canadian Society for Civil Engineering. 7–10

June; London, Ont. Canadian Society for Civil Engineering, Montréal, Que, pp. 504–509.

Dayaratnam P. 1987. Brick and reinforced brick structures. Oxford & IBH.

Deng J, Gu D, Li X, Yue Z-Q. 2005. Structural reliability analysis for implicit performance

functions using ANN. Structural Safety, 27: 25-48.

Der Kiureghian A, Lin HZ, Hwang SJ. 1987. Second order reliability approximations. Journal of

Engineering Mechanics, ASCE, 113 (8): 1208- 1225.

Diamantidis D. 2001. Assessment of Existing Structures. Joint Committee of Structural Safety

JCSS. Rilem Publications S.A.R.L.,The publishing Company of Rilem. 159 pages.

Ditlevsen O. 1982. Model uncertainty in structural safety. Structural safety. Elsevier Scientific

Publishing Company. Amsterdam, 1: 72-86.

Ditlevsen Ove, Henrik O Madsen. 1996. Structural reliability methods. Vol. 178. New York:

Wiley.

Ditlevsen, O. 1984. Taylor Expansion of series system reliability. Journal of Engineering

Mechanics. ASCE, 107(EM6): 1208-1225.

Drysdale, R. G., Hamid, A. A., & Baker, L. R. 1994. Masonry structures: Behaviour and design.

Englewood Cliffs, N.J.: Prentice-Hall.

Dymiotis, C., & Gutlederer, B. M. 2002. Allowing for uncertainties in the modelling of masonry

compressive strength. Construction and Building Materials 16(8): 443–452.

Ellingwood BR, Culver C. 1977. Analysis of live loads in office buildings. ASCE Journal of the

Structural Division, 103(ST8): 1551–1560.

Ellingwood BR, Galambos TV, MacGregor JG, Cornell CA. 1980. Development of a probability

based load criterion for American National Standard A58. NBS Special Publication 577,

National Bureau of Standards, U.S. Department of Commerce, Washington, D.C.

Page 138: Reliability Assessment of Historical Masonry Structures

122

Ellingwood BR, Redfield R. 1983. Ground snow loads for structural design. J. Struct. Eng. 109

(ST4): 950-964.

Ellingwood BR, O’Rourke M. 1985. Probabilistic models of snow loads on structures. Structural

Safety, 2: 291–299.

Ellingwood BR. 1996. Reliability-based condition assessment and LRFD for existing structures.

Structural Safety, 18(2): 67-80.

Ellingwood BR, Tekie PB. 1999. Wind load statistics for probability-based structural design.

ASCE Journal of Structural Engineering, 125(4): 453–464.

EN1990. 2002. Basis for design of structures, Eurocode, European Standard.

European Committee for Standardization. 1994. ENV 1991-1: 1994 Eurocode 1: basis of design

and actions on structures — Part 1: basis of design. European Committee for Standardization,

Brussels, Belgium.

European Committee of Standardization. 1996. Design of masonry structures. Part 1-1: General

rules for buildings—Reinforced and unreinforced masonry. ENV 1996-1-1, Eurocode 6.

Brussels, Belgium.

F08a. 2006. Guideline for the assessment of existing structures. SAMCO Final Report, Federal

Institute of Materials Research and Testing (BAM), Division VII.2 Buildings and Structures

Unter den Eichen 87, 12205 Berlin, Germany.

Faber MH. 2000. Reliability based assessment of existing structures. In Progress in Structural

Engineering and Materials, 247–53.

Faravelli L. 1989. Response surface approach for reliability analysis. Journal of Engineering

Mechanics, ASCE, 115(12): 2763-2781.

Federal Institute of Materials Research and Testing (BAM). n.d. Division VII.2 Buildings and

Structures. Unter den Eichen 87, 12205 Berlin, Germany.

Fehling E, Stürz J. 2006a. Test results on the behaviour of masonry under static (monotonic and

cyclic) in plane lateral loads, Technical Report D7.1a of the collective research project

ESECMaSE.

Fehling E, Stürz J. 2006b. Experimentelle untersuchungen zum schubtragverhaltenvon

porenbetonwandscheiben. Research Report. Universität Kassel.

Page 139: Reliability Assessment of Historical Masonry Structures

123

FEMA 306. 1999. Evaluation of earthquake damaged concrete and masonry wall buildings, basic

procedures manual, ATC-43. Federal Emergency Management Agency, California: Applied

Technology Council.

FEMA 356. 2000. Prestandard and commentary for the seismic rehabilitation of buildings. Federal

Emergency Management Agency, Washington, DC, pp. 518.

Ferry Borges J, Castanheta M. 1971. Structural safety. Laboratorio Nacional de Engenharia Civil,

Lisbon.

Flatjack test, n.d. photograph, retrieved March 20, 2017

<http://www.ncrep.pt/view.php?id=14388>.

Frangopol Dan M, James P Curley. 1987. Effects of damage and redundancy on structural

reliability. Journal of Structural Engineering, 113.7: 1533-1549.

Franke L, Schuman I. 1998. Damage Atlas: Classification and analyses of damage patterns found

in brick masonry. Expert System for the Evaluation of the Deterioration of Ancient Brick

Masonry Structures, European Commission, Protection and Conservation of the European

Heritage, Project EV5V-CT92-0108, Research report 8(2), Fraunhofer IRB Verlag.

Galambos TV, Ellingwood B, MacGregor JG, Cornell CA. 1982. Probability based load criteria:

assessment of current design practice. Journal of the Structural Division, 108(5): 959–977.

Garzón-Roca J, Marco CO, Adam JM. 2013. Compressive strength of masonry made of clay bricks

and cement mortar: Estimation based on Neural Networks and Fuzzy Logic. Engineering

Structures, 48: 21–27.

Geyikens Ph. 1993. Optimization and quadrature algorithms for structural reliability analysis in

the original variable domain [Dissertation]. KV Leuven, Civil Engineering Department.

Glowienka S. 2007. Zuverlässigkeit von Mauerwerkswänden aus großformatigen Steinen

[Dissertation]. Technische Universität Darmstadt.

Gomes HM, Awruch AM. 2004. Comparisons of response surface and neural networks with other

methods for structural reliability analysis. Structural Safety, 26: 49-67.

Graubner CA, Glowienka S. 2008. Stochastic modelling of modern masonry. International Brick

& Block Masonry Conference. Vol. 14.

GruSiBau. 1981. Grundlagen zur festlegung von sicherheitsanforderungen für bauliche anlagen

(GruSiBau). NA Bau, Beuth Verlag, Berlin.

Page 140: Reliability Assessment of Historical Masonry Structures

124

Gumaste KS, Nanjunda Rao KS, Venkatarama Reddy BV, Jagadish, KS. 2006. Strength and

elasticity of brick masonry prisms and wallettes under compression. Materials and Structures

40(2): 241–253.

Gunkler E, Budelmann H, Sperbeck S. 2009. Schubtragfähigkeit von Kalksandsteinwänden aus

KS XL-PE mit geringem Überbindemaß. Research report No. 25/09.

Haldar A, Gao L. 1997. Reliability evaluation of structures using nonlinear SFEM. Uncertainty

Modeling in Finite Element, Fatigue, and Stability of Systems, World Scientific Publishing,

River Edge, N.J., pp. 23–50.

Haldar A, Mahadevan S. 2000. Reliability assessment using stochastic finite element analysis.

Wiley, New York.

Haldar A, Nee KM. 1989. ‘Elasto-plastic large deformation analysis of PR steel frames for LRFD.

Computers and Structures, 31 (5): 811– 823.

Harbitz A. 1986. An efficient sampling method for probability of failure calculation. Structural

Safety, 3(2): 109-115.

Harvey DW, Schuller MP. 2010. Nondestructive Evaluation: Structural Performance of Masonry.

APT Bulletin 41.2/3: 1-8.

Hasofer AM, Lind N. 1974. An exact and invariant first-order reliability format. Journal of

Engineering Mechanics, 100 (EMl): 111-121.

Hassanien SA. 2006. Optimization under Uncertainty for Structural Systems [Dissertation].

University of Calgary, Canada.

Hausmann G. 2007. Verformungsvorhersage vorgespannter Flachdecken unter Berücksichtigung

der stochastischen Eigenschaften [Dissertation]. Technische Universität Darmstadt.

Hendry AW, Malek M. 1986. Characteristic compressive strength of brickwork from collected test

results, Mason. International, 7: 15–24.

Holicky M, Markova J. 2002. Calibration or reliability elements for a column. JCSS workshop on

reliability based code calibration, Zurich.

Höveling H, Steinborn T, Schöps P. 2009. Schubtragfähigkeit von Mauerwerk aus Porenbeton-

Plansteinen und Porenbeton-Planelementen. Research report AiF-Nr.14642BG. Leibniz

Universität Hannover.

Hurtado JE, Alvarez D. 2001. Neural networks-based reliability analysis: A comprehensive study.

Computational methods and applied mechanics for engineering Vol. 97: 113-132.

Page 141: Reliability Assessment of Historical Masonry Structures

125

International Code Council. 2003. International building code.

Hussain A, Akhtar S. 2017. Review of Non-Destructive Tests for Evaluation of Historic Masonry

and Concrete Structures. Arabian Journal for Science and Engineering. 1; 42(3): 925-40.

ISO 13822. 2003. Bases for design of structures — Assessment of existing structures.

ISO 13822. 2010. Bases for design of structures — Assessment of existing structures.

ISO 4355. 1981. Bases for Design of Structures-Determination of Snow Loads on Roofs.

International Standard, International Organization for Standardization.

ISO2394. 1998. General principles on Reliability for Structures. International Standards

Organization.

Jäger W, Schöps P. 2004. Kosteneinsparung durch Ansatz realitätsnaher Bemessungskonzepte für

die Schubbeanspruchung von Mauerwerksbauten. Research report F 2461 of the Federal

Ministry of Transport. Berlin: Building and Urban Development.

JCSS. 2003. Probabilistic model code. Joint Committee on Structural Safety, Zurich,

www.jcss.ethz.chJ.

JCSS. 2011. Probabilistic model code. Joint Committee on Structural Safety, Zurich,

Karamchandani A, Bjerager P, Cornell AC. 1989. Adaptive importance sampling. proceedings of

international conference on structural safety and reliability (ICOSSAR), pp. 855-862.

Kariyawasam SN, Rogowsky DM, MacGregor JG. 1997. Resistance factors and companion-action

load factors for reinforced concrete building design in Canada. Canadian Journal of Civil

Engineering, 24: 593–602.

Kasperski M. 2000. Festlegung und Normung von Entwurfswindlasten. Habilitation, Ruhr-

Universität Bochum.

Kaushik HB, Rai DC, Jain SK. 2007. Stress-strain characteristics of clay brick masonry under

uniaxial compression. Journal of Materials in Civil Engineering, 19(9): 728–739.

Kemp AR, Mahachi J, Milford R. 1998. Developing LRFD codes in South Africa. In Proceedings

of the Structural Engineers World Congress. 19–23 July; San Francisco, Calif. pp. 315-3.

Khuri AI, Cornell JA. 1996. Response surfaces: design and analyses. Marcel Dekker, Inc.

Kirtschig K, Kasten D. 1980. Zur Festlegung von Mauerwerksklassen bei Ingenieurmauerwerk.

Mauerwerk-Kalender. Ernst & Sohn, Berlin.

Page 142: Reliability Assessment of Historical Masonry Structures

126

Lagomarsino S, Cattari S. 2015. PERPETUATE guidelines for seismic performance-based

assessment of cultural heritage masonry structures. Bulletin of Earthquake Engineering 13(1):

13-47.

Lawrence SJ, Stewart MG. 2009. Reliability-based calibration of the capacity reduction factor for

design of masonry in compression to AS3700. Australian Journal of Structural Engineering,

9.2: 97-110.

Lawrence SJ, Stewart MG. 2011. Model error and structural reliability for unreinforced masonry

walls in vertical bending. Journal of the International Masonry Society Masonry International

Vol 24.1.

Li J, Stewart MG, Masia MJ, Lawrence SJ. 2016. fSpatial Correlation of Material Properties and

Structural Strength of Masonry in Horizontal Bending. Journal of Structural Engineering. Jun

21; 142(11):04016112.

Löring S. 2005. Zum Tragverhalten von Mauerwerksbauten unter Erdbebeneinwirkung. Doctoral

thesis. Universität Dortmund.

MacGregor JG, Kennedy DJL, Bartlett FM, Chernenko D, Maes MA, Dunaszegi L. 1997. Design

criteria and load and resistance factors for the Confederation Bridge. Canadian Journal of Civil

Engineering, 24: 882–897.

MacGregor JG. 1976. Safety and limit states design for reinforced concrete. Canadian Journal of

Civil Engineering, 3: 484–513.

MacKinley I. 1983. The neglected hazards of snow and cold. Am. Inst. Arch. J., 68: 52-59.

Madsen HO, Krenk S, Lind NC. 1986. Methods of structural reliability. Prentice Hall, Englewood

Cliffs, New Jersey.

Maes MA, Schueremans L, Van Balen K. 1999. Reliability based assessment of existing masonry

structures. ICASP. Vol. 8.

Magenes G. 2007. Report on static-cyclic tests at University of Pavia. Presentation at ESECMaSE

Dissemination Meeting. Cologne.

Mann W. Statistical evaluation of tests on masonry by potential functions. 1982. Sixth

international brick masonry conference.

Masonry Standards Joint Committee (MSJC). 2002. Building code requirements for masonry

structures, ACI 530-02/ASCE 5-02/TMS 402-02. Detroit: American Concrete Institute,

Page 143: Reliability Assessment of Historical Masonry Structures

127

Structural Engineering Institute of the American Society of Civil Engineers, the Masonry

Society.

Mayer M. 1926. Die Sicherheit der Bauwerke, Springer, Berlin

McGuire RK, Cornell CA. 1974. Live loads effects in office buildings. J. the Struct. Div., ASCE,

100(7): 1351-1366.

Melchers RE. 1987. Structural reliability analysis and prediction. Ellis Horwood Limited.

Melchers RE. 1999. Structural reliability, analysis and prediction. Ellis Horwood Series in Civil

Engineering.

Mohamad G, Lourenço PB, Roman HR. 2007. Mechanics of hollow concrete block masonry

prisms under compression: Review and prospects. Cement and Concrete Composites, 29(3):

181–192.

Mojsilović N, MG Stewart. 2015. Probability and structural reliability assessment of mortar joint

thickness in load-bearing masonry walls. Structural Safety 52: 209–18.

Mori Y, Ellingwood BR. 1993. Reliability based service life assessment of aging concrete

structures. Journal of Structural Engineering, ASCE, 119(5): 1600-1621.

Murotsu Y, Shao S, Chiku N. 1993. Studies on assessment of structural reliability by response

surface method and neural network. Proceedings of the IFIP WG7, 5 Fifth Working

Conference on Reliability and Optimization of Structural Systems. March 16; North-Holland

Publishing Co, pp. 173-180.

Myers RH, Montgomery DC. 1995. Response surface methodology: process and product

optimization using designed experiments. Wiley Series in probability and statistics.

NBCC. 1995. National Building Code of Canada. Institute for Research in Construction, National

Research Council of Canada, Ottawa, Ont.

NBCC. 1997. National Building Code of Canada. Institute for Research in Construction, National

Research Council of Canada, Ottawa, Ont.

NBCC. 1999. National Building Code of Canada. Institute for Research in Construction, National

Research Council of Canada, Ottawa, Ont.

NBCC. 2005. National Building Code of Canada, Institute for Research in Construction, National

Research Council of Canada, Ottawa, ON, Canada.

NBCC. 2010. National Building Code of Canada, Institute for Research in Construction, National

Research Council of Canada, Ottawa, ON, Canada.

Page 144: Reliability Assessment of Historical Masonry Structures

128

NEN 6700. 1990. Technische grondslagen voor bouwconstructies (TGB) – Algemene

Neville AM. 1996. Properties of concrete. Fourth edition, John Wiley & Sons Inc., New York.

Newark MJ, Welsh LE, Morris RJ, Dnes WV. 1989. Revised ground snow loads for the 1990

National Building Code of Canada. Canadian Journal of Civil Engineering, 16: 267–278.

Newark MJ. 1984. A new look at ground snow loads in Canada. In Proceedings of the Eastern

Snow Conference, 41st Annual Meeting, Washington, D.C., 29: 37–48.

NNI. 2005. Technical principles for building structures - TGB Masonry structures - Basic

requirements and calculation methods. NEN standard 6790:2005. Delft: Nederlands

Normalisatie Instituut.

Noland JL, Atkinson RH, Schaller MP. 1990. A review of the flat-jack method for Nondestructive

evaluation. Proceedings of Conference on Nondestructive Evaluation of Civil Structures and

Materials, Boulder, USA.

Nowak AS, Collins KR. 2000. Reliability of Structures. McGraw-Hill.

Nwofor TC. 2012. Durability of block work: the effect of varying water/cement ratio of mortar

joint. Pelagia Research Library, Advances in Applied Science Research, 3 (3):1848-1853.

NZS 4230. 1990. Design of reinforced concrete structures. New Zealand standard.

O'Rourke MJ, Redfield R, Von Bradsky P. 1982. Uniform snow loads on structures. Journal of the

Structural Division, 108(12): 2781-98.

Park J, Towashiraporn P, Craig JI, Goodno BJ. 2009. Seismic fragility analysis of low-rise

unreinforced masonry structures. Engineering Structures, 31(1): 125–137.

Paulay T, Priestley MJN. 1992. Seismic design of reinforced concrete and masonry buildings. New

York: Wiley.

Peir JC, Cornell CA. 1973. Spatial and temporal variability of live loads. ASCE Journal of the

Structural Division, 99(ST5):903–922.

Rackwitz R, Fiessler B. 1978. Structural reliability under combined random load sequences.

Computers and Structures, 9: 489-494.

Rackwitz R. 1995. Live loads in buildings. Manuscript, unpublished, Munich.

Rackwitz R. 2001. Reliability Analysis—a Review and Some Perspectives. Structural Safety,

23(4): 365–95.

Page 145: Reliability Assessment of Historical Masonry Structures

129

Radovanović Ž, SinđićGrebović R, Dimovska S, Serdar N, Vatin N, Murgul V. 2015. Testing of

the mechanical properties of masonry walls – Determination of compressive strength. Applied

Mechanics and Materials, Trans Tech Publications, 725: 410-418.

Rajashekhar MR, Ellingwood B. 1993. A new look at the response surface approach for reliability

analysis. Structural Safety, 12: 489-494.

Rosowsky DV, Cheng N. 1999. Reliability of light-frame roofs in high-wind regions. ASCE

Journal of Structural Engineering, 125(7): 725–733.

Rossi PP. 1987. Recent development of the flat jack test on old masonry structures. Presented

paper at Italy-USA workshop evaluation and retrofit of masonry structure.

Sabri MM, Ulybin AV, Zubkov SV. 2015. A review on different destructive methods to determine

the Compressive Strength of an Existing Masonry. International Journal of Engineering

Research and Development, 11(4): 19–26.

Schermer D. 2007. Report on static-cyclic test at University of Munich, Presentation at

ESECMaSE Dissemination Meeting, Cologne.

Schubert P. 2010. Eigenschaftswerte von Mauerwerk, Mauersteinen und Mauermörtel,

Mauerwerk-Kalender, Ernst & Sohn, Berlin.

Schueremans L, Van Gemert D. 1998. Reliability analysis in structural masonry engineering.

IABSE Colloquium Saving Buildings in Central and Eastern Europe, Berlin.

Schueremans L, Van Gemert D. 1999. Evaluating the reliability of structural masonry elements

using the response surface technique. Durability of Building Materials and Components, 8:

1330-1342.

Schueremans L. 2001. Probabilistic evaluation of structural unreinforced masonry [Dissertation].

KULeuven, Belgium.

Schueremans, L, Verstrynge E. 2008. Use of reliability methods for evaluating safety of historic

structures. SAHC08–6th International Conference on Structural Analysis of Historical

Constructions.

Schueremans, L, Verstrynge E. 2008. Use of reliability methods for evaluating safety of historic

structures. SAHC08–6th International Conference on Structural Analysis of Historical

Constructions.

Page 146: Reliability Assessment of Historical Masonry Structures

130

Shah HC, Dong WM. 1984. Reliability assessment of existing buildings subjected to probabilistic

earthquake loadings. International Journal of Soil Dynamics and Earthquake Engineering, 3.1:

35-41.

South Africa Bureau of Standards. 1989. The general procedures and loadings to be adopted in the

design of buildings: SABS 0160:1989. South Africa Bureau of Standards, Pretoria, South

Africa.

Standards Association of Australia. 1985. Guidelines for conversion to limit states codes,

Committee BD16 — Loading of Structures. Standards Association of Australia, Canberra,

Australia.

Steil RO, Calçada LML, Oliveira AL, Martins VC, Prudêncio Jr LR. 2001. Influência do tipo de

argamassa no fator de eficiência e nadeformabilidade de alvenarias de blocos de concreto. In

Anais do IV SimpósioBrasileiro de Tecnologia de Argamassas, Brasília, 2001, p.423-434.

Stewart MG, Lawrence S. Structural reliability of masonry walls in flexure. Masonry International,

15.2 (2002): 48-52.

Streletzki NS. 1947. Statistical basis for evaluation of the structural safety factor. Stroizdat,

Moscow: State Publishing House for Buildings.

Tabsh SW. 1997. Safety of reinforced concrete members designed following ACI 318 building

code. Engineering Structures, 19(10): 843–850.

Taylor DA, Allen DE. 2000. Statistical variation and duration of snow load on flat roofs in Canada.

Institute for Research in Construction, National Research Council of Canada, Ottawa, Ont.

Unpublished study.

Tschötschel M. 1989. Zuverlässigkeitstheoretisches Konzept für Mauerwerkskonstruktionen

[Dissertation]. Technische Universität Leipzig (in German).

Ultrasonic Pulse Method, n.d. photograph, retrieved Jan 2017,

<https://www.proceq.com/compare/pull-off-testing/>

Vanmarcke EH. 1992. Local spatial variation of earthquake ground motion. Proceedings of

W.C.E.E., pp.6615-6620.

Vermeltfoort AT. 2005. Brick–mortar interaction in masonry under compression [Dissertation].

Eindhoven: Eindhoven University of Technology.

Wen YK. 1990. Structural load modelling and combination for performance and safety evaluation.

Elsevier, Amsterdam, the Netherlands.

Page 147: Reliability Assessment of Historical Masonry Structures

131

Wierzbicki W. 1936. Safety of structures as a probabilistic problem. (Technical Review) Przeglad

Techniczny, Warsaw, Poland.

Zejak D, Vatin N, Murgul V. 2015. Analysis of the masonry structure calculation with vertical

ring beams according to European Standards (2015) Applied Mechanics and Materials, 725-

726, pp. 111-117.

Page 148: Reliability Assessment of Historical Masonry Structures

132

Appendix

This appendix includes a simple example on reliability assessment of an ordinary unreinforced

shear wall (see Figure A-1) using the recommended procedure. Different failure modes of

unreinforced shear wall are explained and their associated limit state functions are determined.

However, as the recommended procedure is the same for different elements under different failure

modes, the probabilities of failure of an unreinforced shear wall under two load combinations and

in the case of two failure modes are calculated as an example.

Figure A-1 Clay brick masonry shear wall.

System Definition

The system which is considered in this example is an unreinforced masonry shear wall. The wall

was constructed with clay units (190mm x 90mm x 57mm) and general purpose mortar. The wall

is considered to be an interior 3m high, 2m long and 0.09m thick wall. The wall is assumed to be

surrounded with a slab of 2 m width and 0.03 m thickness. Self-weight of the shear wall and the

slab is 2.5 𝑘𝑁/𝑚3.

Page 149: Reliability Assessment of Historical Masonry Structures

133

Failure Modes of Masonry Shear Wall

The combination of axial and shear stress, the material properties and slenderness ratio h/lw, can

determine the failure modes of shear walls. The failure modes of a shear wall include:

General compression failure

Overturning failure/Flexural failure

Sliding failure

Diagonal shear failure

General compression failure:

This failure occurs when the axial load applied on the shear wall is high and masonry compressive

strength is low, see Figure A-2.

Figure A-2 General compression failure.

Overturning failure/Flexural failure:

This failure is due to the exceedance of the vertical stress at the toe of the wall from the

compressive strength of the masonry. Therefore, The units located in the bottom corner of the wall

crush. This failure mode is more likely to occur for slender walls, Figure A-3.

Page 150: Reliability Assessment of Historical Masonry Structures

134

Figure A-3 Flexural failure.

Sliding failure:

This failure mode is more likely for low axial load and large horizontal force (mostly in an

unreinforced wall) which result in the exceedance of shear stress from the sliding strength of the

bed joints, Figure A-4.

Figure A-4 Sliding failure.

Diagonal shear failure:

This failure mode can occur when axial load is higher compared to sliding failure mode (potentially

through units), Figure A-5.

Page 151: Reliability Assessment of Historical Masonry Structures

135

Figure A-5 Diagonal shear failure.

Limit state

Ultimate limit state is considered in this example.

Limit State Function

Flexure (axial force and moment) limit state function:

𝑔(𝑥) = 𝑃𝑓

𝐴−

𝑀𝑓𝑦

𝐼 𝛼𝐷 = 0.9

(A-1)

𝑔(𝑥) = ∅𝑚𝑓𝑚′ −

𝑃𝑓

𝐴+

𝑀𝑓𝑦

𝐼 𝛼𝐷 = 1.25

(A-2)

where 𝑀𝑓 is factored moment and 𝑃𝑓 is factored axial load at the section under consideration.

Sliding Limit state Function:

𝑔(𝑥) = 𝑉𝑟 − 𝑊 (A-3)

𝑉𝑟 = ∅𝑚𝜇𝑃2 (A-4)

𝑃2 = 0.9 𝐷 (A-5)

Page 152: Reliability Assessment of Historical Masonry Structures

136

where 𝑊 is the applied wind load on the wall, 𝜇 is friction coefficient, 𝑃2 is compression force

perpendicular to the sliding face and 𝐷 is the applied dead load.

Diagonal Shear Limit State Function:

𝑔(𝑥) = 𝑉𝑟 − 𝑊 (A-6)

𝑉𝑟 = ∅𝑚(𝑉𝑚𝑏𝑤𝑑𝑣 + 0.25𝑃𝑑)𝛾𝑔 + (0.6∅𝑠𝐴𝑣𝑓𝑦𝑑𝑣

𝑆) ≤ 0.4∅𝑚√𝑓𝑚

′ 𝑏𝑤𝑑𝑣𝛾𝑔 (A-7)

𝑉𝑚 = 0.16 (2 −𝑀𝑓

𝑉𝑓𝑑) √𝑓𝑚

′ 0.25 ≤𝑀𝑓

𝑉𝑓𝑑≤ 1

(A-8)

where 𝑊 is the applied wind load on the wall, 𝑏𝑤 width of wall, 𝑑𝑣 is effective depth, 𝛾𝑔 is

grouting factor, 𝑀𝑓 is factored moment and 𝑉𝑓 is factored shear at the section under

consideration.

Probabilistic Model of Applied Load

Bartlett et al. (2003) presented a summary of the statistical parameters for loads which have been

used to calibrate the loads and load combination criteria for the 2005 National Building Code of

Canada (NBCC). Their recommended probabilistic model is considered in this example.

Dead load:

Dead load is considered to be the self-weight of the wall in addition to the self-weight of the slab.

Bias factor of dead load is considered to be equal to 1 (Bartlett et al., 2003). The probabilistic

model considered for dead load in this example can be found in Table A-1.

Table A-1 Probabilistic model of dead load.

Mean (𝐾𝑁) CoV Distribution Type

1.65 0.10 Normal

Page 153: Reliability Assessment of Historical Masonry Structures

137

Live load:

The nominal uniformly distributed live load is calculated as follows (NBCC, 1999):

𝐿 = 2.4 (0.3 + √9.8

𝐵) = 4.48 (A-9)

𝐵 = 2 × 2 = 4 𝑚2 (A-10)

where 𝐵 is tributary area. Having the bias value of 0.9, the mean value of live load is calculated to

be 4.03. The probabilistic model of 50-year maximum live load is considered in this example

(Bartlett et al., 2003). Table A-2 presents the probabilistic model of live load.

Table A-2 Probabilistic model of live load.

Mean (𝐾𝑁) CoV Distribution Type

4.03 0.17 Gumbel

Wind load:

The shear wall is assumed to resist the wind which is applied to a 6 m × 3 m wall locating

perpendicular to the shear wall. The reference velocity pressure is derived from Appendix C of the

2010 NBCC for Halifax (𝑞 = 0.58 𝐾𝑝𝑎). Knowing the reference velocity pressure, wind velocity

is calculated from the relationship between wind pressure and wind velocity (see Chapter 4). The

bias factor associated with wind velocity in 50-year is given to be equal to 1.049 for Halifax

(Bartlett et al., 2003). Therefore mean value of wind velocity is calculated to be equal to 1 𝑚/𝑠2.

Exposure factor (𝐶𝑒) is assumed to be 0.9 and the external pressure-gust coefficient (𝐶𝑝𝐶𝑔) is

assumed to be 1.15. The bias factor of the overall transformation factor, its CoV and distribution

type is considered based on the recommendation of Bartlett (2003) (see Chapter 4). The importance

factor is estimated to be 1.15.

Page 154: Reliability Assessment of Historical Masonry Structures

138

Overall transformation factor = 𝐶𝑒𝐶𝑝𝐶𝑔 = 0.9 × 1.15 = 1.035 (A-11)

Table A-3 presents the considered probabilistic modeling for wind load parameters.

Table A-3 Probabilistic model of wind load.

Parameter Mean (𝑚/𝑠2) CoV Distribution Type

Wind velocity 1 0.17 Gumbel

Overall transformation factor 0.7 0.22 Lognormal

Considered Load combinations:

Two load combinations are selected to be considered in this example, see Table A-4.

Table A-4 Considered load combination.

Load Combination

Principal Load Companion Load

0.9D 1.4W

(0.9D or 1.25D) + 1.5L 0.4W

Probabilistic Model of Material Properties

Modulus of Elasticity:

The mean value and CoV of modulus of elasticity of clay brick masonry (with weak mortar)

reported by Kaushik et al. (2007) is considered in this example. The distribution of modulus of

elasticity is also considered to be log-normal as recommended by JCSS (2011). These assumptions

are made to be able to proceed this example. However, the modulus of elasticity and its statistical

characteristics are recommended to be estimated by UPM as a non-destructive test in the case of

real application of the procedure for reliability evaluation of historic masonry structures. Table A-

5 presents the probabilistic model of modulus of elasticity.

Page 155: Reliability Assessment of Historical Masonry Structures

139

Table A-5 Probabilistic model of modulus of elasticity.

Mean (𝑀𝑃𝑎) CoV Distribution type

2300 0.24 Log-normal

Compressive Strength:

The compressive strength of clay masonry is recommended to be estimated by modulus of

elasticity using following formula

𝐸 = 800 𝑓𝑚′ (A-12)

It should be noted that 800 is the median of the coefficient k values reported by different authors.

The probabilistic model of compressive strength is presented in Table A-6.

Table A-6 Probabilistic model of compressive strength.

Mean (𝑀𝑃𝑎) CoV Distribution type

2.88 0.24 Log-normal

Probability of Failure under Flexure

The probability of failure of the shear wall is estimated in both cases of the flexural failure mode

and diagonal shear failure mode under two load combinations. The Monte-Carlo Method is used

to calculate the probability of failure in this example. There are other reliability analysis methods

(mentioned in Chapter 3) which can be used.

Using the Mont-Carlo Method, the number of failures is estimated to be 1 out of 600 in the case

of flexural failure under the first load combination presented in Table A-4, which means

𝑃𝑓 =1

600= 0.0017 (A-13)

The target probability of failure is calculated as follows (see Chapter 3)

Page 156: Reliability Assessment of Historical Masonry Structures

140

𝑃𝑓𝑇 =10−4𝑆𝑐𝑡𝐿𝐴𝑐𝐶𝑓

𝑛𝑝𝑊=

10−4 × 0.005 × 100 × 1.0 × 0.1

10 × 1.0= 5 × 10−7 (A-14)

where 𝑡𝐿 is assumed to be 100 years and 𝑛𝑝 is assumed to be equal to 10. The values of the other

parameters are derived from Table 3-1 in Chapter 3. Therefore probability of failure is more than

target probability of failure which means the shear wall is unreliable in case of flexural failure

under the first load combination.

However, no flexural failure occurs under the second load combination (𝑃𝑓 = 0). Diagonal shear

failure has not occurred under either load combination (𝑃𝑓 = 0) as well which means the

considered shear wall is probably not going to experience diagonal shear under these two load

combinations.