. _ .. ...2!3.:§ 5...: f. u. .72? ,5 21Wmuazwmflfl. I? 333‘ . an? m .ihifiz . ioltazt 13.. 1a....hnul‘xl... .3 :2. .. I. .. 4." l Ibiza)... , «tub .1! Ju. 4 ‘5’... S . ‘. Human 3.15...“ ) 5!". :11: xgr ‘ 1.. h! qfiqfim. ; . 11...??? .- uhfldfifiz . .215, .5: .a 5.!!!" t §th-.VBIVI-.§. LI." r .57)! .IV 1w. ”'3 LIBRARY Michigan State University This is to certify that the dissertation entitled CAPACITY REDUCTION AND FIRE LOAD FACTORS FOR LRFD OF STEEL MEMBERS EXPOSED TO FIRE presented by SHAHID IQBAL has been accepted towards fulfillment of the requirements for the Ph.D degree in Civil Engineering Wake— Major Professor’s Signature 5%”? :27, 42°10 Date MSU is an Affirmative Action/Equal Opportunity Employer PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 KIProj/AccaPrelelRC/Dateoue.indd CAPACITY REDUCTION AND FIRE LOAD FACTORS FOR LRFD OF STEEL MEMBERS EXPOSED TO FIRE By Shahid Iqbal A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Civil Engineering 2010 ABSTRACT CAPACITY REDUCTION AND FIRE LOAD FACTORS FOR LRFD OF ' STEEL MEMBERS EXPOSED TO FIRE By Shahid Iqbal Until recently, structural steel members exposed to fire were designed using prescriptive approaches that do not account for actual loading conditions and real fire scenarios. The last decade has seen the promotion of performance-based codes which allow more rational engineering approaches for the design of steel members exposed to fire. For example, Appendix 4 of the 2005 AISC Specifications (referred to hereafter as “AISC Specifications”) now allows steel members to be designed against fire using room temperature design specifications and reduced material properties. The AISC Specifications allow using the same capacity reduction factors for fire design as those used for room temperature design. For example, a capacity reduction factor of 0.9 is suggested for steel beams and columns. Most other codes suggest that a capacity reduction factor of 1.0 be used. This recommendation is based on arguments that the probability of fire occurrence and the strength falling below the design value simultaneously is very small, and that fire design is based on the most likely expected strength. The Commentary to the AISC Specifications also states that the fire load may be reduced by up to 60 percent if a sprinkler system is installed in the building. Automatic sprinklers reduce the probability of occurrence of a severe fire. The reduction in fire load should be based on proper reliability analysis that includes the effect of sprinklers on the occurrence of a severe fire, and correspondingly on the probability of failure of structural steel members. However, limited work has been done to develop capacity reduction and fire load factors based on reliability analysis. In this study, a general reliability-based methodology is proposed for developing capacity reduction and fire load factors for design of steel members exposed to fire. The effect of active fire protection systems (e.g., sprinklers, fire brigade, etc.) in reducing the probability of occurrence of a severe fire is included. The design parameters that significantly affect the fire design of steel members such as fire load, opening factor, thermal absorptivity of compartment boundaries, thickness, density and thermal conductivity of insulation are taken as random variables. Raw experimental data published in the literature was analyzed to obtain the statistics of parameters for which no statistical information was available in the literature. Model errors associated with the thermal analysis models are also characterized based on experimental data. It is found that uncertainty associated with the fire design parameters is significantly higher than that of room temperature design parameters. To illustrate the proposed methodology, capacity reduction and fire load factors are developed for simply supported steel beams and axially loaded steel columns in US. office buildings. It is found that the capacity reduction and fire load factors should not be constant for all design situations as suggested in current design specifications, but should vary depending on the presence of active fire protection systems in a building. In addition, a simplified method is proposed for computing inelastic deflections of simply supported steel beams exposed to fire, and it is shown that the fire design of most beams is goverened by the strength limit state. DEDICATION I dedicate this dissertation to my mother and father. Without their support and encouragement, I would have never been in graduate school pursuing a PhD. degree. My late father worked hard and sacrificed days and nights of his life to support my education throughout my academic life. I owe this degree to him. I also dedicate this dissertation to my sons (Hassan and Ans) who did not get the time and attention they needed from me in the last 4-5 years because of my studies, and had to participate in many events, activities and celebrations without me. iv ACKNOWLEDGMENTS I express my most sincere gratitude to my advisor Dr. Ronald S. Harichandran, Professor and Chair of the Department of Civil and Environmental Engineering, Michigan State University, for the encouragement and guidance received through the course of this study. In my academic and professional life, I have worked with many people, but, no one ever influenced my life more than Professor Harichandran. I will always remember his humbleness, care, and moral and emotional support he provided in these 4.5 years. It is impossible to appreciate, in few words, his patience and perseverance which made my graduate studies a great learning experience. I would like to thank Professor Venkatesh Kodur, Assisstant Professor Lajnef Nizar and Associate Professor Sasha Kravchenko for serving on my Ph.D committee, and for the valuable advice and discussions we had throughout my stay at Michigan State University. I would also like to thank all other faculty members who taught me, especially Dr. Rigoberto Burguefio who helped me in attaining the requisites skills and training for pursuing a Ph.D. degree. The material I learned in the courses I took with Dr. Rigoberto Burguefio, and my one-to-one academic discussions with him were very fruitful, useful, and really helped me during my Ph.D. work. For many reasons that I cannot count, I am thankful to my friends Monther Dwaikat, Mahmud Dwaikat, and Mahmoodul Haq. They are great friends, who were always available and willing to help whenever I needed it. I would like to thank the many other students in the civil and environmental engineering department at Michigan State University for their help and support during my Ph.D. study. I appreciate the help provided by Laura Taylor, Mary Mroz, Margaret Conner, Joseph Nguyen, Brooke H.M. Stokdyk, and Mary M. Gebbia-Portice, in solving my administrative problems, and visa and passport issues. I am extremely grateful to my brother, my wife, my uncles and aunts, and my cousins who are always happy for me and for my successes. I would also take this opportunity to say thanks to my friends Sher Malik, Abdul—ur- Rehman, Arif Aalam, Nabeel Anwar Cheema, Naeem Olakh, Arif Mehmood, Attiq-ur— Rehman, Adnan Qadir, Khurram, and Naveed, for being happy on my selection to pursue my Ph.D. in the US, and for their prayers. I am extremely grateful to Brigadier Gulfam and Major General Mansha for their support for pursuing my PhD studies. I am thankful to my teachers Mr. Durez Sahib, Mr. Mehar Khan Sahib, Mr. Akseer Sahib, Mr. Muhammad Yousaf Sahib, Mr. Malik Nawaz Sahib, Mr. Roshan Ali Sahib, Mr. Hasnain Shah Sahib, Mr. Tariq Mehmood Sahib, Mr. Bashir Sahib, Mr. Mushtaq Sahib, Mr. Noor Khan Sahib, Mr. Asim, Brigadier Khaleeq Kayani, Brigadier Gulfam, Brigadier Jamil, and Brigadier Abeer, for making my academic career a success. Finally, I thank the E-in-C Branch, Pakistan Army, and Ministry of Defence for financially supporting my Masters and Ph.D. studies, and I am grateful to the staff members working at these organizations for making this Ph.D. program a success. vi TABLE OF CONTENTS LIST OF TABLES xi LIST OF FIGURES xii CHAPTER 1 1 INTRODUCTION 1 1.1 BACKGROUND ............................................................................................................ l 1.2 DESIGN FOR FIRE HAZARD ......................................................................................... 3 1.3 FIRE AND STEEL MEMBERS ........................................................................................ 5 1.3.] F ire-Resistant Design of Steel Members ............................................................ 5 1.3.2 Capacity Reduction Factor for Fire Design ...................................................... 8 1.3.3 Reduction in Fire load ....................................................................................... 8 1.3.4 Uncertainties in Fire Design of Steel Members ................................................. 9 1.3.5 Deflections of Steel Beams Exposed to Fire .................................................... 10 1.4 RESEARCH OBJECTTVES ............................................................................................ 1 1 1.5 SCOPE ...................................................................................................................... 12 1.6 ORGANIZATION OF THE DISSERTATION .................................................................... 13 CHAPTER 2 15 STATE-OF-THE-ART REVIEW 15 2.1 INTRODUCTION ......................................................................................................... 15 2.2 APPROACHES FOR EVALUATING FIRE RESISTANCE OF STEEL MEMBERS ................. 16 2.2.1 Prescriptive Approach ..................................................................................... 1 7 2.2.1.1 Determination of Fire Resistance Rating .......................................... 17 2.2.1.2 Standard Fire Test ................................................................................ 18 2.2.2 Performance-Based Approach ......................................................................... 20 2.2.3 Detailed Calculation Methods ......................................................................... 21 2.2.3.1 Fire Temperature .................................................................................. 23 2.2.3.2 Temperature of Structural Member .................................................... 23 2.2.3.3 Strength Calculations ........................................................................... 24 2.3.3.4 Critical Factors in Fire Resistance Calculations .............................. 24 2.2.4 Fire Scenarios .................................................................................................. 25 2.2.4.1 Room Fires ............................................................................................. 25 2.2.4.2 Standard Fire ......................................................................................... 27 2.2.4.3 Design (or Real) Fires .......................................................................... 29 2.2.5 Loading Conditions .......................................................................................... 31 2.2.6 Failure Criteria ................................................................................................ 32 2.3 USE OF RELIABILITY THEORY IN THE DEVELOPMENT OF DESIGN SPECIFICATIONS .. 34 2.3.1 Background ...................................................................................................... 38 2.3.2 F irst—Order Second-Moment ( F OSM ) Analysis ............................................... 41 2.3.3 Hasofer-Lind Reliability Index ........................................................................ 45 2.3.4 Approximate Methods for Including Information on Distributions ................. 49 Vii 2.3.6 Selection of Load and Capacity Reduction Factors ........................................ 54 2.4 STUDIES RELATED TO RELIABILITY ANALYSES OF STRUCTURAL MEMBERS EXPOSED To FIRE .......................................................................................................................... 61 2.4.1 ECSC Study, 2001 ............................................................................................ 62 2.4.2 Magnusson and Pettersson, I981 .................................................................... 63 2.4.3 Beck, 1985 ........................................................................................................ 70 2.4.4 Li and Fitzgerald, 1996 .................................................................................... 72 2.4.5 Ellingwood, 2005 ............................................................................................. 73 2.4.6 CIB W14, 1983 and 1986 ................................................................................. 74 2.4. 7 Ramachandran, 1995a ..................................................................................... 75 2.4.8 Ramachandran, 1995b ..................................................................................... 76 2.4.9 Harmathy and Mehaffey, 1985 ......................................................................... 77 2.4.10 Fellinger, 2000 ............................................................................................... 77 2.4.11 Janes, 1995 .................................................................................................... 78 2.5 SUMMARY ................................................................................................................ 78 CHAPTER 3 81 RELIABILITY-BASED METHODOLOGY FOR DEVELOPING CAPACITY REDUCTION AND FIRE LOAD FACTORS 81 3.1 GENERAL ................................................................................................................. 81 3.2 ENGINEERING APPROACH FOR DESIGNING STEEL MEMBERS EXPOSED To FIRE ...... 82 3.3 FIRE TEMPERATURE IN THE COMPARTMENT ............................................................ 84 3.4 CALCULATING THE TEMPERATURE OF STEEL MEMBERS .......................................... 86 3.5 METHODOLOGY FOR DEVELOPING CAPACITY REDUCTION AND FIRE LOAD FACTORS ....................................................................................................................................... 89 3.5.1 Performance Function for Reliability Analysis ............................................... 89 3.5.1.1 Applied Loads ........................................................................................ 89 3.5.1.2 Capacity of Steel Members ................................................................. 90 3.5.1.3 Limit State Equation ............................................................................. 92 3.5.2 Statistics of Random Parameters ..................................................................... 92 3.5.3 Probability of Failure and Target Reliability Index ........................................ 92 3.5.4 Reliability Analysis .......................................................................................... 97 3.5.5 Determination of Partial Safety Factors .......................................................... 97 3.5.6 Combination of Partial Safety Factors into Single Capacity Reduction Factor ................................................................................................................................... 98 3.5.7 Optimal Capacity Reduction and Fire Load Factor ...................................... 100 3.6 SUMMARY .............................................................................................................. 101 CHAPTER 4 103 STATISTICS OF RANDOM PARAMETERS AND MODEL ERRORS .............. 103 4.1 STATISTICS OF RANDOM PARAMETERS .................................................................. 104 4.1.1 Fire Load Density .......................................................................................... 106 4.1.2 Ratio of Floor Area to Total Surface Area of the Compartment ................... 111 4.1.3 Opening Factor .............................................................................................. 112 4.1.4 Thermal Absorptivity of Compartment Enclosure ......................................... 113 4.1.5 Thickness of Insulation or Fire Protection Materials .................................... 117 viii 4.1.6 Thermal Conductivity and Density of Insulation ........................................... 1 18 4.2 MODEL ERRORS FOR THERMAL MODELS ............................................................... 120 4.2.1 Model Error for Steel and Fire Temperatures ............................................... 120 4.2.2 Model Error for Reduction Factors ............................................................... 122 4.3 SUMMARY .............................................................................................................. 124 CHAPTER 5 126 CAPACITY REDUCTION AND FIRE LOAD FACTORS FOR STEEL BEAMS EXPOSED TO FIRE 126 5.1 DERIVATION 0F CAPACITY REDUCTION AND FIRE LOAD FACTORS ........................ 127 5.1.1 Performance Functions for Reliability Analysis ............................................ 127 5.1.1.1 Applied Moment under Fire ............................................................... 127 5.1.1.2 Moment Capacities of Beams under Fire ........................................ 128 5.1.1.3 Performnace Function for Reliability Analysis ................................ 133 5.1.2 Model Error (Professional Factor) for Moment Capacity Equations ........... 133 5.1.3 Probability of Failure and Target Reliability Index ...................................... 134 5.1.4 Reliability Analyses ........................................................................................ 134 5.1.4.1 Reliability Analyses of Laterally Restrained Beams ...................... 138 5.1.4.2 Reliability Analyses of Laterally Unrestrained Beams .................. 139 5.2 RESULTS ................................................................................................................ 139 5.2.1 Capacity Reduction Factor ............................................................................ I39 5.2.2Fire Load Factor ............................................................................................ 141 5.2.3 Validity of Capacity Reduction and Fire Load Factors for Multiple Fire Scenarios ................................................................................................................. 142 5.2.4 Comparison of F ire Load Factors with those Based on the ECSC Method.. 147 5.2.5 Insulation Thicknesses from Performance-Based and Prescriptive Design Approaches ............................................................................................................. 149 5.2.6 Reliability Inherent in AISC Fire Design Provisions ..................................... 151 5.3 DESIGN EXAMPLES ................................................................................................ 151 5.3.1 Design Example (Laterally Restrained Beam) ............................................... 152 5.3.2 Design Example (Laterally Unrestrained Beam) .......................................... 155 5.4 SUMMARY ........................................................................................... . ................... 160 CHAPTER 6 162 CAPACITY REDUCTION AND FIRE LOAD FACTORS FOR STEEL COLUMNS EXPOSED TO FIRE 162 6.1 DERIVATION OF CAPACITY REDUCTION AND FIRE LOAD FACTORS ........................ 162 6.1.1 Performance Functions for Reliability Analysis ............................................ 162 6.1.1.1 Applied Axial Loads under Fire ......................................................... 162 6.1.1.2 Axial Capacity of Columns under Fire ............................................. 163 6.1.2 Professional Factor (Model Error) for Axial Capacity of Columns .............. 165 6.1.3 Probability of Failure and Target Reliability Index ...................................... 166 6.1.4 Reliability Analysis ........................................................................................ 166 6.2 RESULTS ................................................................................................................ 171 6.2.] Capacity Reduction Factor ............................................................................ 171 6.2.2 Fire Load Factor ............................................................................................ I 72 6.2.3 Validity of Capacity Reduction and Fire Load Factors for Difi’erent Slendemess Ratios .................................................................................................. 1 72 6.2.4 Validity of Capacity Reduction and Fire Load Factors for Different Live to Dead Load Ratios ................................................................................................... 173 6.2.5 Validity of Capacity Reduction and Fire Load F actors for Multiple Fire Scenarios ................................................................................................................. I 75 6.3 DESIGN EXAMPLE .................................................................................................. 177 6.4 SUMMARY .............................................................................................................. 182 CHAPTER 7 183 DEFLECTIONS OF SIMPLY SUPPORTED STEEL BEAMS EXPOSED TO FIRE 183 7.1 BACKGROUND ........................................................................................................ 183 7.2 DEVELOPENT OF SIMPLIFIED METHOD FOR ESTIMATING DEFLECTIONS OF SIMPLY SUPPORTED BEAMS ...................................................................................................... 184 7. 2. 1 Approach ........................................................................................................ 184 7.2.1.1 Equivalent Flexural Rigidity ............................................................... 186 7.2.1.2 Finite Element Model .......................................................................... 189 7.2.1.3 Parametric Analysis ............... . ............................................................ 194 7.2.2 Validation of Simplified Method .................................................................... 198 7.2.2.1 Effect of Heating Rate of Steel ......................................................... 200 7.2.2.2 Effect of Load Ratio ............................................................................ 200 7.2.2.3 Effect of Sectional Geometry and Span Length ............................. 202 7.2.2.4 Effect of Load Configuration .............................................................. 202 7.2.3 Comparison with Test Results ........................................................................ 203 7.3 DEFLECTIONS OF SIMPLY SUPPORTED BEAMS ....................................................... 205 7.4 SUMMARY .............................................................................................................. 209 CHAPTER 8 210 CONCLUSIONS AND RECOMMENDATIONS 210 8.1 SUMMARY .............................................................................................................. 210 8.2 CONCLUSIONS ........................................................................................................ 212 8.3 RECOMMENDATIONS FOR FUTURE RESEARCH ........................................................ 214 REFERENCES 217 LIST OF TABLES Table 3.1 - Range of target reliability index values based on the the effect of active fire protection systems in reducing the probability of occurrence of a severe fire ................. 96 Table 4.1 - Mean, COV and distributions of design parameters related to loads ........... 104 Table 4.2 - Mean, COV, distributions and nominal values of fire design parameters... 105 Table 4.3 - Comparison of statistics of fire load density (MJ/m2 of floor area) in offices ......................................................................................................................................... 109 Table 4.4 - Comparison of fire load density in offices (Yii , 2000) ............................... 109 Table 5.1 — Mean and COV of room temperature design parameters ............................. 129 Table 5.2 — Properties of laterally restrained simply supported beams used for reliability analyses ........................................................................................................................... 136 Table 5.3 — Properties of laterally unrestrained simply supported beams used for reliability analyses .......................................................................................................... 137 Table 5.4 - Combinations of b and F v used for nine fire scenarios ................................. 144 Table 5.5 — Values of b, F v and qt used for nine fire scenarios corresponding to ,8, =1.5 ......................................................................................................................................... 145 Table 5.6 - Comparison of insulation thickness .............................................................. 150 Table 6.1 - Properties of axially loaded steel columns used for reliability analyses ..... 168 Table 6.2 - Properties of columns used for validation .................................................... 175 Table 7.1 - Comparison of fire resistance time predicted by the simplified method with test results at the limiting deflection value of L/30 ......................................................... 205 Table 7.2 — Comparison of deflections of laterally restrained simply supported with the limiting values of U30 or U20 ...................................................................................... 207 Table 7.3 — Comparison of deflections of laterally unrestrained simply supported with the limiting values of U30 or U20 ....................................................................................... 207 xi LIST OF FIGURES Figure 1.1 — Comparison of a design (or real) and a standard fire ...................................... 6 Figure 2.1 - Time-temperature curve for ASTM E119 standard fire and two design fire exposures ........................................................................................................................... 19 Figure 2.2 — Degradation of strength of steel beam exposed to fire ................................. 21 Figure 2.3 — Flowchart for calculating the fire resistance of a structural system (Reporduced from Kodur 2007) ........................................................................................ 22 Figure 2.4 - Various stages of fire ................................................................ . ................... 27 Figure 2.5 - Fundamentals of risk-based design evaluations ........................................... 35 Figure 2.6 - Probability of failure, pp (Reproduced from Harichandran 2005) ................ 39 Figure 2.7 - Interpretation of the reliability index ,8 (Reproduced from Harichandran 2005) ................................................................................................................................. 41 Figure 2.8 - Linear approximation to non-linear function (Reproduced from Harichandran 2005) ................................................................................................................................. 43 Figure 2.9 - Illustration of failure boundary and design point in x- and z-spaces for two variables problem (Reproduced from Harichandran 2005) .............................................. 47 Figure 2.10- Illustration of failure boundary adjustment to achieve target reliability index (Reproduced from Harichandran 2005) ............................................................................ 56 Figure 4.1 - Comparison of yield strength and modulus of elasticity reduction models with test data ................................................................................................................... 123 Figure 5.1 - Capacity reduction and fire load factors for steel beams ............................ 140 Figure 5.2 - Nine fire scenarios for target reliability index value of 1.5 ........................ 145 Figure 5.3 - Comparison of computed and target reliability index values ...................... 146 Figure 5.4 - Comparison of fire load factors with those obtained by the ECSC method 149 Figure 5.5 - Fire and steel temperatures VS. time used in computing fire resistance for beam in example 2 .......................................................................................................... 160 Figure 6.1 - Ratio of test capacity to nominal capacity of columns for different Slendemess ratios ............................................................................................................ 166 xii Figure 6.2 - Capacity reduction and fire load factors vs. target reliability index for columns ........................................................................................................................... 171 Figure 6.3 Computed and target reliability index values for different Slendemess ratios ......................................................................................................................................... 173 Figure 6.4 - Computed and target reliability index values for different live to dead load ratios ................................................................................................................................ 174 Figure 6.5 - Computed and target reliability index values for columns for different fire scenarios .......................................................................................................................... 176 Figure 6.6 - Computed and target reliability index values for columns with reduced - capacity reduction factors ............................................................................................... 177 Figure 7.1 - Typical moment—curvature diagram for a steel section ............................... 186 Figure 7.2 - Variation of ultimate tensile strength of steel ............................................. 188 Figure 7.3 — Mechanical properties for structural steel at elevated temperature ............. 191 Figure 7.4 - Validation of high-temperature creep model .............................................. 192 Figure 7.5 - Comparison of deflections of Steel beams predicted by ANSYS with the test data for (a) protected beam, and (b) an unprotected beam ............................................. 193 Figure 7.6 - Variation of normalized flexural rigidity with normalized inelastic moment for different load and Slendemess ratios ......................................................................... 196 Figure 7.7 - Deflections predicted by ANSYS (broken lines) and simplified method (solid lines) for different heating rates. Load ratio = 0.5, load type = UDL ............................. 201 Figure 7.8 - Deflections predicted by AN SYS (broken lines) and simplified method (solid lines) for different load ratios. HR = 45°F/min, load ratio = 0.5, load type = UDL ....... 201 Figure 7.9 - Deflections predicted by ANSYS (broken lines) and simplified method (solid lines) for different sections and span lengths. Load ratio = 0.5, HR = 45 °F/min, load type = UDL ............................................................................................................................. 202 Figure 7.10 - Deflections predicted by ANSYS (broken lines) and simplified method (solid lines) for different load configurations. Load ratio=0.5, HR: 45 °F/min ............. 203 xiii Chapter 1 Introduction 1. 1 Background Steel is widely used as a construction material and finds a wide range of applications in buildings, bridges, and other infrastructure. It competes well with other construction materials because of its light weight, high strength, reliability and speed of construction. Steel framing is lighter and thinner compared to its counterparts such as concrete and wood and thus results in saving of space and cost. Although steel is non-combustible and adds no fuel in a fire, it rapidly loses its strength and stiffness at elevated temperatures, which is a cause of concern for fire engineers. Over the past few decades, there has been significant research on steel structures exposed to fire. Today, the properties and behavior of steel structures exposed to fire is better known compared to many other construction materials. A wide variety of fire protection materials are used to enhance the performance of steel under fire conditions. In recent years, fire-resistant steel is being used extensively in countries like Japan. However, in the US, most steel buildings are still designed against fire using fire protection materials. Fire represents one of the severe hazards to which a structure might be subjected during its lifetime and the performance of steel buildings under fire is often questionable. Thus, the provision of appropriate fire safety measures is an important aspect of building design for ensuring the safety of occupants, and to protect property and the environment. Fire safety measures have two components: active measures such as automatic sprinklers to control and suppress the fire, and passive measures such as the provision of adequate fire resistance to the structural and non-structural components of the building so that they do not collapse during the fire duration specified in codes. Active systems include automatic sprinklers, smoke detectors, heat detectors, fire alarms and fire brigades. Passive systems control the fire or fire effects by the systems that are built into the structure or fabric of the building, for example, structural and non-structural components like beams, columns, and walls of the building. These building components help control the spread of fire and also ensure the structural integrity of the building, if provided with adequate fire resistance. Passive fire protection can be achieved by protecting structural members in a variety of ways, e.g., by applying spray applied materials (sprayed mineral fiber, vermiculate plaster etc.), using intumescent coating, or using board materials (e.g., gypsum board) as insulation. Fire resistance is the time duration during which a structural member or system exhibits resistance with respect to structural integrity, stability and temperature transmission (Buchanan 2001). Although, there are different ways to provide fire resistance to building components, for Structural Steel members, fire resistance is generally provided using fire protection materials. The significance of the fire resistance can be attributed to the fact that when other measures such as active fire protection systems fail to control the fire, the fire resistance of building elements is the last line of defense for the safety of people, property and the environment. Fire resistance can play a crucial role on the performance of a structure in the event of fire as seen in the collapse of the WTC Twin Towers (FEMA 2002). 1.2 Design for Fire Hazard Fires cause significant loss of life and property damage throughout the world. In 2007, fires in the US. caused about 3430 deaths, 97,800 injuries and property damage worth $14 billion. The total cost of fire (including direct and indirect losses) was estimated to be around $70 billion in 2007 (Karter 2008). Design for fire safety has been improved significantly in the last century. This improvement focused mainly on reducing the loss of life by introducing active fire safety systems (which are self-triggered in the event of fire) such as sprinklers and fire detections systems. In 1913, the death rate due to fire in the United States was 9.1% per 100,000 people. However, the use of fire resistant material and improvements in design (including smoke detectors) reduced that rate to 2% per 100,000 people in 1988 (Williams 2004). Generally, the major cause of fatalities and injuries is toxic smoke evolving during a fire event. However, some of these deaths and injuries, particularly among first responders, were due to structural collapse/damage under fire conditions (FEMA 2008). The following realistic examples illustrate the catastrophic consequences that may result from fire incidents: On Oct. 26, 1986, a serious fire occurred in the lS-story Alexis Nihon Hotel in Montreal, Canada (Inser 1988). This was a steel framed building with a composite steel beams and deck floor. The fire continued for 14 hours and resulted in the partial collapse of the 11th floor. All walls, decks and beams on the four fire floors had to be replaced and the repair cost was $80 million. On Feb. 23, 1991, a serious fire occurred in the 38-story Meridian Plaza office building in Philadelphia, U.S. (Routley et al. 1991). This was a steel framed building with composite steel beams and deck floors. The fire burned for 18 hours causing significant structural damage to 9 floors. Three fire fighters died. On May 4, 1988, a fire occurred on the 12th floor of the First Interstate Bank in Los Angles, U.S (Klem 1988). This was a 62 story office building having steel framing with composite beams and deck floors. The fire lasted for 3.5 hours and caused major damage to four floors. On Sep. 11, 2001, the 47-Story WTC 7 office building totally collapsed due to fire. The above examples clearly illustrate the need for structural fire safety for minimizing the catastrophic collapse of buildings and the built-infrastructure. To minimize the possibility of structural failure during fire, building codes such as the International Building Code, specify fire resistance requirements for structural systems in buildings. The required fire resistance varies based on the importance of occupancy, number of stories and floor area of the building, type of the structural members (such as Slab, beam or column), and also on the availability of active fire protection systems (e. g., sprinklers) in the building. As an illustration, the fire resistance requirement for a steel beam in a typical building can vary from 1 to 3 hours. 1.3 Fire and Steel Members 1.3.1 Fire-Resistant Design of Steel Members The fire resistance requirement for structural members (i.e., the time during which a structural member should perform its intended function) is specified in building codes based on factors such as the type of occupancy, height of the building, and importance of the structural member. Until now, the required fire resistance for Steel members (e.g., beams, columns, etc.) was provided using prescriptive approaches, and the fire resistance rating of a large number of steel sections is included in fire resistance directories such as the Underwriter Laboratory Fire Resistance Directory (UL 2004). The designer selects the required thickness of the fire protection material from the listed steel sections by ensuring that the selected thickness of fire protection material yields a fire resistance more than that required by the building codes. The fire resistance rating of listed steel sections is determined from tests conducted in a furnace under a standard fire. If the section being designed is not included in the fire resistance directory, the thickness of the fire protection material is calculated using empirically derived correlations (Ruddy et a1. 2003). The prescriptive approach described above has yielded adequate performance, but is considered too conservative and simplistic because it does not account for actual loading conditions and real fire scenarios. The temperature of a typical structural steel member exposed to a real fire (or design fire) and the standard fire used in the prescriptive approach is compared in Figure 1.1. The two fires are quite different and are likely to yield different performances for structural members. Over the last decade, performance-based codes which allow more rational engineering approaches for the fire design of steel members are being promoted. For example, Appendix 4 of the 2005 AISC Specifications (referred to hereafter as “AISC Specifications”) now allows steel members to be designed against fire using room temperature design specifications and reduced material properties. Similar provisions were developed by the European Convention for Constructional Steel work (ECCS 2001) and are included in the Eurocode 3 (EN 2005). Using this engineering approach, the verification of design for strength during fire requires that the resistance (capacity) is greater than the load effects. This leads to satisfying the design equation ¢fRn,f 2Qu,f (1'1) 1200 . standard fire typical real fire (or design fire) rooo — / a / Q, 800 . g - o- 600 -r E 8 E 400 1 '1’ steel temperature, real fire 200 ~ Steel temperature, standard fire fire 0 f r I l r r O 20 40 60 80 100 120 Time (minutes) Figure 1.1 - Comparison of a design (or real) and a standard fire where Quf is the load effect at the time of fire, Rnf is the nominal capacity at the time of fire, and ¢f is the capacity reduction factor. The AISC Specifications (AISC 2005a) allow using the same capacity reduction factors for fire design as those used for room temperature design. For example, ¢f = 0.9 is suggested for steel beams and columns. Most other codes suggest that a capacity reduction factor of 1.0 be used (e.g., in the Eurocode 3 (EN 2005), the partial safety factor TM is 1.0 for fire design). When a steel member is exposed to fire. the temperature of the steel increases and the strength and stiffness of the steel are reduced, leading to significant loss of capacity. The temperature attained by a steel member depends onthe area of steel exposed to fire, the amount of protection applied, and on the duration and severity of a design fire. The AISC Specifications (AISC 2005a) require that while calculating the steel and fire temperatures for fire design, due consideration should be given to the effectiveness of all active fire protection systems (sprinklers, smoke and heat detectors, etc.). The commentary to the AISC Specifications (AISC 2005b) states that the fire load may be reduced by up to 60 percent if a sprinkler system is installed in the building. The fire design time, during which Equation 1-1 should be satisfied, is not precisely defined. Design specifications in various countries use this time in different ways for different occupancies. For example, single Storey buildings may be designed to protect the escape routes and to remain standing only long enough for occupants to escape the building. On the other hand, tall and high rise buildings where people cannot escape are designed to prevent the spread of fire, or collapse of the building. 1.3.2 Capacity Reduction Factor for Fire Design The capacity reduction factor, ¢ f, accounts for the uncertainties in the material properties, variations in dimensional and sectional properties, and, approximations and assumptions used in the design and analysis procedures. As mentioned earlier, for fire design of steel members, ¢f= 0.9 is suggested in the AISC Specifications (AISC 2005a), but most other codes suggest that a capacity reduction factor of 1.0 be used (e.g., in the Eurocode 3 (EN 2005), the partial safety factor 1’ M is 1.0 for fire design). This recommendation is based on arguments that the probability of fire occurrence and the strength falling below the design value simultaneously is very small, and that fire design is based on the most likely expected strength (Buchanan 2001). Also, it is expected that live loads under fire conditions are likely to be smaller than those at room temperature conditions and hence there will be enough reserve strength available, especially for members that have been designed for load combinations such as wind, snow or earthquakes, or for members proportioned for deflection control or for architectural reasons (Buchanan 2001). However, limited work has been done to develop capacity reduction factors based on reliability analysis (Magnusson and Pettersson 1981), and neither ¢ f = 0.9, nor ¢f = 1.0 are based on reliability considerations. 1.3.3 Reduction in Fire load As mentioned previously, the AISC Specifications (AISC 2005a) suggest that while describing the design fire, due consideration should be given to the effectiveness of all active fire protection systems (sprinklers, smoke and heat detectors, etc.). The Commentary to the 2005 AISC Specifications (AISC 2005b) states that while describing the design fire, the fire load may be reduced by up to 60 percent if a sprinkler system is installed in the building. Automatic sprinklers reduce the probability of occurrence of a severe fire. The reduction in fire load should be based on proper reliability analysis that includes the effect of sprinklers on the occurrence of a severe fire, and correspondingly on the probability of failure of structural steel members. Recently, a study was conducted in Europe through a research project of the European Coal and Steel Community (ECSC) (hereinafter referred to as the ECSC study) to develop fire load factors by taking into account the variability of the fire load and the effect of active fire protection systems (ECSC 2001). However, the fire load factors were obtained using simplified assumptions instead of rigorous reliability analysis. It is not apparent whether rigorous reliability analysis would yield results similar to those of the ECSC Study. 1.3.4 Uncertainties in Fire Design of Steel Members The design of structural members exposed to fire is quite complex because both thermal and structural analyses must be performed. Compared to room temperature design, there are many more parameters involved in fire design (e.g., fire load, ventilation or opening factor, thermal properties of compartment enclosures, thickness and thermal properties of insulation or fire protection materials, and thermal properties of steel). The assumptions and approximations used in thermal analyses add further uncertainty to the design process. In the ECSC study, the variability of fire load was accounted for while developing the fire load factors, but the uncertainty associated with other fire design parameters was not explicitly included. As was done for room temperature design, the capacity reduction and fire load factors for fire design should be based on rigorous reliability analysis that includes the uncertainty associated with all design parameters. In CIB design guides (CIB 1983 and CIB 1986), a procedure is given for deveIOping the load and resistance factors for fire design. Ellingwood (2005) has recommended load factors and load combinations for the fire design and these recommendations were incorporated in the ASCE Specifications (ASCE 07 2005) and Appendix 4 of the AISC Specifications (AISC 2005a). However, not much work has been done to develop capacity reduction and fire load factors based on rigorous reliability analysis. 1.3.5 Deflections of Steel Beams Exposed to Fire Large deformations are expected under severe fire conditions, so deflections are not normally computed unless they affect the structural performance (Buchanan 2001). The AISC Specifications (AISC 2005a) specifies that for structural elements, e.g., beams and columns, the governing limit state is loss of load-bearing capacity. However, the AISC Specifications also specify that excessive deformations are not acceptable if these damage the integrity of the fire compartment. If not limited, beam deformations may damage the slabs and walls of the compartment to the extent that these components cannot control the horizontal and vertical spread of fire even though the beam may still have sufficient load bearing capacity. Additionally, deformations need to be limited because long beams may fail due to an abrupt increase of their deflection despite the fact that they may have not reached their ultimate load bearing capacity (Skowronski 1990). Furthermore, beams designed with a large margin of safety against collapse in fire, can lose that margin due to the abrupt growth of strains caused by creep at elevated temperatures even when stresses are not critical (Skowronski 1990). 10 Since design under fire may be governed either by the strength or deflection limit states, both should be considered. Design specifications such as the AISC Specifications (AISC 2005a) and Eurocode 3 (EN 2005) provide simple design equations for the strength limit state of steel beams which can be used for estimating the fire resistance. Researchers such as Yin and Wang (2004), Skowronski (1990), Burgess, El-Rimawi and Plank (1990), and Wang and Yin (2006) provide detailed method for tracing the time-deflection behavior of steel beams, but, these approaches are quite complex. Thus no simple deflection equations which can be quickly implemented in a design specification are available for computing the fire resistance of steel beams. 1.4 Research Objectives From the above discussion, it is clear that no substantial work has been done on developing capacity reduction factors and fire load factors for steel members subjected to fire conditions. ¢f = 0.9, nor ¢f = 1.0 are based on subjective judgment and not on reliability analysis. The effect of active fire protection systems such as automatic sprinklers in reducing the probability of occurrence of a severe fire was not explicitly accounted for while specifying a reduction in fire load in the Commentary to the AISC Specifications. In the ECSC study, the variability of the fire load and effects of active fire protection systems were accounted for while developing the fire load factors. However, the fire load factors were obtained using simplified assumptions and the study also did not account for variability in other fire design parameters. Therefore, there is a need to develop capacity reduction and fire load factors based on reliability analyses for use in the fire design of steel members instead of using the current ones that are primarily based ll on experience and subjective judgment. In addition, although it is recognized that excessive deflections of steel members need to be limited, no simplified method is given in the AISC Specifications for estimating the inelastic deflections of steel members exposed to fire. The objectives of this research are: 0 Propose a general reliability-based methodology for deriving capacity reduction and fire load factors for the LRFD of steel members exposed to fire. 0 Characterize uncertainties associated with the fire design parameters for deriving capacity reduction and fire load factors. 0 Develop capacity reduction and fire load factors for axially unrestrained steel beams exposed to fire. 0 Develop capacity reduction and fire load factors for axially loaded steel columns exposed to fire. 0 Suggest appropriate design/nominal values of the fire design parameters for the performance-based design of steel members exposed to fire. 0 Develop a simplified method for predicting the deflections (or alternatively fire resistance time) of simply supported steel beams exposed to fire, and check the deflections of beams designed per strength limit state. 1.5 Scope The fire load for different types of occupancies varies from one country to another and even varies significantly for different types of occupancies within the same country (CIB W14 1986). The probability of fire occurrence is also not the same for all types of occupancies (ECSC 2001). Therefore, it is not possible to develop one set of capacity 12 reduction and fire load factors for all types of occupancies. Accordingly, the CIB W14 Design Guide (1986) suggests that one set of factors may be developed for any type of occupancy (say office buildings), and then differentiated factors may be developed to modify these basic factors for other type of occupancies. The fire load factors developed in this study will be applicable to US. office buildings. Capacity reduction factors developed herein are specific to Simply supported steel beams and to axially loaded steel columns for typical fire compartments in US. office buildings. However, the methodology presented in this study is general and can be used for deriving capacity reduction factors for other types of structural members (e.g., restrained beams) by using appropriate performance functions (or design equations), corresponding to the fire load factors developed in this study. 1.6 Organization of the Dissertation This dissertation has eight chapters as outlined below. Chapter 1: Provides background information leading to this research and also presents the objectives and scope. Chapter 2: Provides a state-of—the-art review of the methods used to evaluate the fire resistance of steel members exposed to fire. The chapter also includes a brief presentation about the use of reliability theory in the development of design specifications. Based on the literature review, previous studies related to the reliability analysis of steel Structures exposed to fire are presented and their drawbacks are briefly discussed. Chapter 3: Presents a general reliability-based methodology for developing capacity reduction and fire load factors for steel members exposed to fire. 13 Chapter 4: Characterizes the fire design parameters that significantly affect the design of steel members. Statistics obtained from the literature and those developed from the experimental data are presented. Model errors associated with the thermal analysis are also described in this chapter. Chapter 5: Presents the development of capacity reduction and fire load factors for laterally restrained and unrestrained Simply supported steel beams. Example design problems also are given. Chapter 6: Presents the development of capacity reduction and fire load factors for axially loaded steel columns. Example design problems also are given. Chapter 7: Proposes a simplified method for predicting deflections of simply supported steel beams exposed to fire conditions. Also compares the deflections of the beams against the limiting values of deflections for the beams designed per Strength limit state. Chapter 8: Summarizes the main findings of the current study and provides recommendations for further research. 14 Chapter 2 State-of-the-Art Review 2. 1 Introduction In the literature, no explicit work has been reported on the derivation of capacity reduction factors for fire design of steel members except for one case (Magnusson and Pettersson 1981). Similarly, only one study has been reported in which fire load factors were developed using simplified assumptions instead of rigorous reliability analysis (ECSC 2001). However, during the last few decades, reliability analyses were performed for estimating the probability of failure and computing reliability indices for steel members exposed to fire. In some other studies, methodologies were suggested to develop reliability-based fire safety codes. Some of the important studies are summarized later in this chapter. Most of the studies related to reliability analysis were performed for 15 Steel members that were designed using the prescriptive design approaches instead Of the performance-based design approaches being promoted now. Therefore, for better understanding of the reliability Studies, these design approaches (prescriptive and performance-based) will be reviewed first. Some important information (e.g., fire scenarios, loading conditions, failure criterion, etc.) related to the fire design of steel members is also reviewed in this chapter. In addition, the use of reliability theory in the development of design specifications is also briefly described in this chapter. 2.2 Approaches for Evaluating Fire Resistance of Steel Members The fire resistance rating requirements for structural members are specified in building codes (e.g., IBC 2003) based on factors such as the type of occupancy, height of the building, and importance of the structural member. The fire resistance ratings are given as a Specified amount of time the building’s structural members are required to withstand exposure to a standard fire. For example, in high rise building, a steel beam and a steel column may be required to have a fire resistance rating of 3 and 4 hours, respectively. In the US, IBC (2003) allows both prescriptive and performance-based fire-resistant designs, although its current emphasis is clearly on the former (Ruddy et a1. 2003). In the US, structural steel members (e.g., beams, columns, etc.) are designed in accordance with the AISC Specifications (2005). Before 2005, AISC had a separate design guide (Ruddy et al. 2003) for the fire-resistant design of steel members which was based on the applications of prescriptive provisions of International Building Code. However, presently, Appendix 4 of the AISC Specifications (2005) allows steel members to be designed using either engineering analysis (performance—based approach) or by 16 qualification testing (prescriptive approach). According to the engineering approach, individual members can be designed against fire using room temperature design provisions and reduced material properties. This design approach allows using one- dimensional heat transfer analysis and a uniform temperature distribution across the steel section. In the method of qualification testing, the AISC Specifications require that structural members and components be qualified for the rating period in conformance ' with ASTM E 119 (2008). Eurocode 3 (EN 2005) is one of the most widely accepted design document for fire design. This code gives a choice of advanced or simplified calculation methods for designing steel members for fire. As an alternative to designing using calculations, the code allows that the fire design may be based on the results of fire tests, or on fire tests in combination with calculations. The simplified method in Eurocode 3 is mainly based on a sectional analysis approach. This method may give a more economical design and/or high fire resistance periods for some fire scenarios. Eurocode 3 also allows advanced methods where detailed thermal and mechanical analyses can be performed for designing members for fire. 2.2.1 Prescriptive Approach 2.2.1.1 Determination of Fire Resistance Rating Until now, the required fire resistance for steel members is provided using prescriptive approaches, and the fire resistance rating of a large number of steel sections is included in building codes, standards, test reports and special directories of testing laboratories (IBC 2000, IBC 2003, ASCE/SFPE 29, NFPA 2002, UL 2003). Most Of the information about 17 fire resistance ratings is given in the form of charts and tables. The designer selects the required thickness of the insulation for any steel section from a chart or a table by ensuring that the selected thickness of the insulation yields a fire resistance rating more than that required by the building codes. If the section being designed is not included in the charts and tables, the thickness of the insulation can be determined using empirically derived correlations (Ruddy et al. 2003). The fire resistance rating mentioned in these charts and tables is either taken directly from experiments conducted on steel members in a furnace or extended to untested members using empirically derived correlations. Most members are tested under a standard fire which is described in the next sub-section. Some members are tested under loading, while in other cases the failure criterion is the achievement of a temperature at which steel is believed to lose most of its structural strength. 2.2.1.2 Standard Fire Test The fire resistance rating of steel members is generally evaluated by subjecting the structural member to fire in a specially constructed furnace. The purpose of the standard fire test is to determine the failure time at which the structure loses the ability to withstand fire exposure while maintaining function as a load-bearing element and as a barrier to the spread of fire. The standard fire test is a comparative test that does not reflect the actual performance of the member. Standard fire resistance tests are generally carried out on building elements such as walls, floors, or columns in accordance with national standards such as ASTM E119a (2008) and ISO 834 (1975). The standards require test specimens be constructed in a similar manner as the building elements in practice. ASTM E119a specifies the dimensions of 18 the test specimen and the size of the furnace to be used for the standard fire test. Further, the furnace chamber is heated by liquid fuel or gas such that the furnace follows the specified standard ASTM E119 fire time-temperature curve shown in Figure 2.1. During the fire test, load bearing members, such as steel beams and columns, are generally loaded at service load levels (dead load + live load) that are about 50% of the room temperature capacities of the tested members. 1500 1200 3 900 I _ ‘ —ASTME119fire , + Severe fire 600 T —a— Moderate fire 300 i ' O I I I I ----------- - fl ] I F O 30 60 90 120 150 180 210 240 Time (min) Figure 2.1 - Time-temperature curve for ASTM E119 standard fire and two design fire exposures The test continues until a prescribed failure criterion is exceeded. The fire resistance of the assembly is then recorded as the duration of fire exposure until this failure. The limiting criterion used for evaluating fire resistance depends on the structural member in question. As an illustration, for a steel column, the fire resistance rating is the time at 19 which the average temperature at any level increases above 538°C, or the temperature at any one measuring point increases above 649°C (Ruddy et a1. 2003), or the column fails to sustain the applied load (ASTM E119a 2008). Different failure limit states are specified for columns, beams and slabs. The prescriptive approach described above has yielded adequate performance, but is considered too conservative and simplistic because it does not account for actual loading conditions and real fire scenarios. 2.2.2 Performance-Based Approach Over the last decade, performance-based codes which allow more rational engineering approaches for the fire design of steel members are being promoted. For example, Appendix 4 of the 2005 AISC Specifications now allows steel members to be designed against fire using room temperature design specifications and reduced material properties. Using this engineering approach, the verification of design for strength during fire requires that the resistance of the structure is greater than the load effects. This leads to satisfying the design equation ¢mef Z me (2-1) where Quf is the load effect at the time of fire, Rnf is the nominal capacity at the time of fire, and ¢f is the capacity reduction factor. The AISC Specifications (AISC 2005a) state that capacity reduction factors are the same as those used for room temperature design. Most other codes suggest that a capacity reduction factor of 1.0 be used. At elevated temperatures, the strength and stiffness of Steel reduces significantly, and if unprotected, steel members fail within a short time. The degradation of the strength of a 20 simply supported steel beam is illustrated in Figure 2.2. The fire resistance time of the beam should be greater than the required fire resistance time specified in the building codes. To reduce the degradation of strength and stiffness, members are generally protected by insulation to slow down the rise of the steel temperature. The required thickness of the insulation can be determined from Equation 2—1 using an iterative procedure. i r. 1 Load (b) Simply Supported Beam (momeqt) Strength of the beam Failure point / Applied load Fire resistance .-IL--------- Time (a) Strength Degradation of Beam Figure 2.2 — Degradation of strength of steel beam exposed to fire 2.2.3 Detailed Calculation Methods Whereas fire resistance provisions in the past were primarily based on full-scale tests, the recent trend is moving toward calculation methods. This is mainly because standard fire 21 tests are very expensive and time consuming. Advances in numerical models are facilitating the application of calculation methodologies for evaluating fire resistance. These calculation methods are used both in the prescriptive and performance-based design approaches. In recent years, several mathematical models to calculate the fire resistance of structural members in buildings have been developed. The flowchart in Figure 2.3 illustrates the general calculation procedure employed in such methods (Kodur 2007, Buchanan 2001). Increment time 1 Fire temperature: Calculation of fire temperature 1 Thermal analysis: Calculation of member temperature Thermal properties I Mechanical Strength Analysis: '1. Calculations of strains, stresses, prope '83 strength and deflection Check failure Figure 2.3 - Flowchart for calculating the fire resistance of a structural system (Reporduced from Kodur 2007) The fire resistance calculation is performed in three steps: 0 Calculation of the fire temperature. 0 Calculation of the temperatures in the fire-exposed structural member. 22 0 Calculation of the strength of the member during exposure to fire, including an analysis of the stress and strain distributions. 2.2.3.1 Fire Temperature For a performance—based design approach, the real or design fire temperature can be computed using any suitable mathematical model from the literature (EN 2002. SFPE 2000 and 2004, and Buchanan 2001, etc.). In the prescriptive design approach, the fire temperature is assumed to follow the standard time-temperature curve specified in ASTM E119a (2008). 2.2.3.2 Temperature of Structural Member The next step in the analysis is the calculation of the temperatures of the fire-exposed member. These temperatures are generally calculated using a finite difference or finite element method (Buchanan 2001, Kodur 2007). In these methods, the cross-section Of the member is divided into a number of elemental regions, which may have various shapes such as squares, triangles or layers, depending on the geometry of the member (Kodur and Harmathy 2008). For each element or layer, a heat balance equation can be derived. By solving the heat balance equations for each element or layer, the temperature history of the member can be calculated. The Steel temperature can be calculated using any advanced finite element software (e.g., ANSYS 2007, SAFIR 2003, etc.). However, most design specifications such as the AISC Specifications (AISC 2005a) and Eurocode 3 (EN 2005) allow the steel temperature to be calculated using simple thermal analysis methods such as the lumped heat capacity method. The lumped heat capacity method assumes that the steel section is a lumped 23 mass at uniform temperature. The heat balance differential equations for protected and unprotected steel members are provided in the design specifications and literature (AISC 2005a, EN 2005, Buchanan 2001, etc.). 2.2.3.3 Strength Calculations In the third step, a Stress-Strain analysis is conducted to determine the strength of the member during exposure to fire. This strength decreases with increasing temperature and duration Of fire exposure. The fire resistance can be derived by determining the time at which the strength of the member reaches the load to which the member is subjected. 2.3.3.4 Critical Factors in Fire Resistance Calculations To achieve a reliable fire resistance estimate through detailed calculations, the following factors must be accounted for (Kodur et al. 2009): 0 High temperature material properties: The temperature-dependent thermal and mechanical properties that are important for establishing the fire response of steel structures should be used. These properties vary as a function of temperature, and are crucial for modeling the fire response of steel structural members. 0 Strain components: In addition to mechanical and thermal strains, creep strains must be considered in computing steel strains at elevated temperatures. Creep strain which are often ignored, may have a significant influence on the fire response of steel structural members. 0 Fire scenarios: In most fire scenarios, there always exist a growth phase followed by a decay (cooling) phase. Further, the fire scenario is a function of compartment characteristics such as fire load density, ventilation conditions and thermal 24 absorptivity of compartment boundaries. Therefore, it is essential, in modeling the fire response of steel structural members, to account for realistic fire scenario including the decay phase. 0 Restraint effects: Fire induced restraint is believed to influence the fire response of structural Steel members. Thus, restraint effects must be accounted for in tracing the response of structural steel members. 0 Geometric nonlinearity: At elevated temperature, the structural members generally experience large deformations because of the deterioration in strength and stiffness of the members. It is important to account for geometric nonlinearity in tracing the fire response of structural steel members. 0 Failure limit states: Currently, the failure of structural steel members under fire conditions is generally evaluated based on thermal and strength failure limit states. Other failure criteria such as deflection and rate of deflection may be critical under some conditions. For evaluating the fire resistance of steel members, especially through a performance- based approach, three main factors must be considered, namely: (1) fire scenario, (2) loading conditions, and (3) failure criteria. These three factors are elaborated in the following sections. 2.2.4 Fire Scenarios 2.2.4.1 Room Fires The fire growth inside a typical room, often called a fire compartment, depends on a number of factors including: 25 0 Fire load density that is related to the quantity of combustible materials. 0 Opening factor that is related to the degree of ventilation. 0 Dimensions of the compartment. 0 Thermal properties of the lining materials. 0 Thermal properties of compartment boundaries. 0 The amount of volatile combustibles released within the room per unit time. 0 The nature Of the fuel (of cellulose or thermoplastic nature). Cellulosic materials are generally charting materials where a layer of char is created at the fire- material interface. The char layer protects the inner core of the material and slows down the burning process. Thermoplastic materials do not generally experience charring when they are subjected to fire. Figure 2.4 Shows a typical time-temperature curve for the complete process of fire development inside a fire compartment, assuming no fire suppression by sprinklers and fire fighters. Most fires spread slowly at first and then more rapidly as the fire grows. When the temperature of the upper layer reaches about 600°C, the flashover stage occurs where the burning rate increases, fire engulfs most of the fuel items, and temperatures become very high. The severity and duration of a fire depends on the amount of combustibles present in the room, the ventilation conditions, and the type of bounding surfaces of the room. After some time, either the fuel burns out or there is a lack of oxygen and the temperature drops during the decay phase (Buchanan 2001). After the flashover stage, the fire is known as a post-flashover fire which is considered structurally significant and is used for the fire-resistant design of steel members. Characterization of an appropriate post-flashover fire is difficult because of numerous 26 uncertainties arising from the multiplicity of fire scenarios and the complexity of room fire behavior. For ease of calculation, most standard use a standard time temperature curve to represent a room fire scenario (ASTM E119a 2008, ASTM 1529 1993, EN 2002, ISO 834 1975). A Intense burning . g Growth Decay E ‘ ' 8 .5. I'- Ignition Flashover Time Figure 2.4 — Various stages of fire 2.2.4.2 Standard Fire As mentioned earlier, at present most countries rely on fire resistance tests to assess the performance of structural steel elements. The time-temperature curve used in these fire resistance tests is called “a standard fire.” The most widely used test specifications for a standard fire are the ASTM E119 and ISO 834. .In these standard test specifications, the fire temperature, Tf, in °C is calculated as: ISO-834 specification: Tf = 34510g(8t "l' 1) '1' T0 (2-2) ASTM E119 specification: 27 Tf = 7500— exp(—3.79553\/1;))+ 170.41); + To (2-3) where t is the time in minutes, 1;, is the time in hours and T0 is ambient temperature in °C. The fire temperature obtained from these two specifications is independent Of the amount of fire load, ventilation conditions and properties of the compartment boundaries. In reality, however, these parameters have a significant effect on the severity of the fire temperatures, and also on the duration of the fire in any compartment. Figure 2.1 illustrates the time-temperature curves for a standard and two real or design fire scenarios. In the standard fire (ASTM E119a 2008), the fire size is the same (irrespective of compartment characteristics), temperature increases with time throughout the fire duration, and there is no decay (or cooling) phase. In real fires, the fire size is a function of compartment characteristics, and there is a decay phase as shown in Figure 2.1 (severe and moderate design fires). Although, the ASTM E119 and ISO 834 standard time-temperature curves may roughly predict the growth and burning stages of room fires, they do not consider the decay phase which always exists for real room fires, and which has a significant influence on the fire resistance of structural members. In addition, these standard time-temperature curves represent a limited range of possible scenarios involving fire load, ventilation conditions and the bounding surfaces (Bwalya et al. 2003). As mentioned earlier, such standard fires are used to define the temperature profile that is generally used in standard fire tests. While standard fire resistance tests are useful benchmarks to establish the relative performance of different structural members under the standard fire condition, they should not be relied upon to determine the survival time of structural members under realistic fire scenarios. Nor does the standard heating 28 condition bear any relation to the Often less severe heating environments encountered in real fires. Due to drawbacks of the standard time-temperature curves, most design specifications now require use of real or design fires for the design of structural members. For example, Appendix 4 of the 2005 AISC Specifications suggests that the design of load bearing structural members be based on a design (or real) fire when using an engineering /performance-based approach. 2.2.4.3 Design (or Real) Fires The fire temperature, Tf, can be estimated using a suitable mathematical model from the literature (e.g., SFPE 2000 and 2004, EN 2002, Buchanan 2001, etc.). In most of these models, the temperature of the design fire is a function of the opening factor, F v, fire load density, (1,, and thermal absorptivity, b. In this study, the Eurocode parametric fire model modified by Feasey and Buchanan (2002) is used to estimate the fire temperature under design fire scenarios, and this model is briefly described below. For the burning phase, the Eurocode equation for temperature Tf (°C) is given as * * * Tf =1325(1— 0.324e—0‘2t - 0.204e‘1-7t — 0.472e—19" )+ T0 (2-4) where t* is a fictious time in hours given as t = Ft ; tis the time in hours; l.._(Fv/Fref)2 (b/bref)2 (2-5) 29 Fv = 1‘1,”le / A, is the ventilation or Opening factor; Av is the area of ventilation openings; HV is the height of ventilation openings; A, is the total area of the bounding surfaces of the fire compartment; 1) = kpcp (WSO'S/mZK) is the square root of the thermal inertia of the bounding surfaces of the fire compartment; k, ,0 and CI) are the thermal conductivity, density and specific heat of the bounding surfaces of the fire . . . . 0.5 . compartment; F ref IS the reference value of the ventilation factor given as 0.04 (m ) In Eurocode 1; and bref is the reference value given as 1160 WsO'S/mzK in Eurocode 1. Feasey and Buchanan (2000) have shown that that the temperatures given by the Eurocode 1 (EN 2002) formula are often too low, and recommended that a value of bref= 1900 WSO'SlmzK be used instead of 1160 ‘WsO'S/mZK. Using bref= 1900 (WsO‘S/mZK) and F v =0 .04 mO'S, the burning phase of the design fire can be obtained from Equation 2-4 by using 2 F = (F, / 0.04)2 (2_ 6) (b/1900) The duration of the burning period td in hours is then given as 1 — 0 00013 ‘1' d _ ° F.— (2-7) V where q, is the fire load density (MJ/m2 ) of the bounding surfaces. 30 For the cooling phase, Eurocode 1 gives a reference decay rate (dT/dt)ref = 625 °C/hr for fires having a burning period of less than half an hour, decreasing to 250 °C/hr for fires with a burning period greater than 2 hours. For incorporating the effect of thermal insulation and opening factor, Feasey and Buchanan (2002) suggested that the reference decay rate given by the Eurocode be modified to dT _[dT] .lF,/0.04 —— — —— 2-8 dt dt ,6, 719/1900 ( ) Two design fires shown in Figure 2.1 were obtained through Equations 2-4 to 2-8. 2.2.5 Loading Conditions The prescriptive provisions in current codes of practice for evaluating fire resistance through standard fire tests are generally based on a load ratio of about 50%. The load ratio is defined as the ratio of the applied load on the structural member under fire conditions to the resistance of the member at room temperature. The load ratio depends on many factors including the type of occupancy of the building, the dead load to live load ratio, and the safety factors (load and resistance factors) used both for room temperature and fire condition designs. However, the AISC Specifications require that while designing steel members using the performance-based approach (engineering analysis), the applied load effects (e.g., bending moment, shear force, axial force, etc.) should be determined from the gravity load combination given as: U = 1.21),, + 0.5L" + 0.25,, + T (2—9) 31 where D", Ln and Sn are nominal dead, live and snow loads, respectively, and T includes loads induced by the fire itself, especially, due to restraint from the surrounding structures preventing thermal expansions, or due to a flexural member becoming a tension member after large deformations occur. ASCE- 07 (2005) also provides the load combination similar to Equation 2-9 for the design of structural members subjected to extra-ordinary events (e.g., fire). Based on ASCE- 07 (2005) and AISC Specifications (AISC 2005a), and for typical dead-load-to— live-load ratios (in the range of 2 to 3), the load ratio ranges from 40% to 70%. The load level has a significant influence on the fire resistance of structural steel members. Thus, for innovative, realistic and cost effective fire safety design, it is important to evaluate the fire resistance of structural steel members based on actual load levels instead of the typical load ratio of 50% often used in the prescriptive approach. 2.2.6 Failure Criteria The conventional approach of evaluating fire resistance (defining failure under fire conditions) is based on thermal and strength failure criteria as specified in ASTM E119a (2008). Accordingly, thermal failure of steel members is said to occur when their temperature exceeds the critical temperature limit of 538°C, and strength failure is said to occur when the steel member is unable to resist the applied load during the fire design time. Although, most codes, such as Eurocode 3 (EN 2005), allow the fire resistance of steel members to be evaluated using the critical temperature limit state, load bearing structural steel members may not fail even when their temperatures reach the limiting value of 32 538°C. This is especially true for members that have low load ratios, because they were likely designed for load combinations such as earthquake or wind, or for deflection limits. Such members will have considerable reserve strength because of low load ratios and will not fail even when their temperature reaches the limiting value of 538°C. Therefore, it is more rational to design steel members using the strength limit state, and now most codes recognize and allow this. The AISC Specifications (2005) specify that for structural elements, e.g., beams and columns, the governing limit state is loss of load-bearing capacity. However, they also specify that excessive deformations are not acceptable if these damage the integrity of the fire compartment. If not limited, deformations may damage the slabs and walls of the compartment to the extent that these components cannot control the horizontal and vertical spread of fire even though the beam may still have sufficient load bearing capacity. Additionally, deformations need to be limited because long beams may fail due to an abrupt increase of their deflection despite the fact that they may have not reached their ultimate load bearing capacity (Skowronski 1990). Furthermore, beams designed with a large margin of safety against collapse in fire, can lose that margin due to the abrupt growth of strains caused by creep at elevated temperatures even when stresses are not critical (Skowronski 1990). The deflection and rate of deflection can play a crucial role on the fire response of structural steel members exposed to fire. In fact, a deflection limit is Often applied to check the status of steel beams at ambient conditions (serviceability limit state). This criterion should be considered to determine failure under fire conditions. The resulting deflections in fire scenarios are generally higher than those at room temperature due to 33 deterioration of member stiffness and also due to temperature-induced creep. The British Standard (BS 476 1987) contains deflection and rate of deflection criteria for defining failure of a steel beam tested in a furnace. Although these deflection limit states might have been set to limit damage to the furnace during fire tests, deflection and rate of deflection can be important under certain fire conditions. This is because the integrity of the structural member cannot be guaranteed with excessive deformations. Moreover, fire resistance based on limiting deflection will help to facilitate the safety of fire fighters and also to safely evacuate occupants prior to structural collapse (Kodur and Dwaikat 2008). According to BS 476 (1987), failure is assumed to occur when: 0 The maximum deflection of the beam exceeds L/20 at any fire exposure time, or 0 The rate of deflection exceeds L2 9000d (mm/min) (2- 10) where L = span length of the beam (mm), and d = effective depth of the beam (mm). The other failure limit state that is relevant for fire resistance evaluation is the loss of Shear strength, which should also be considered. 2.3 Use of Reliability Theory in the Development of Design Specifications In general, engineering design of structures consists of proportioning the elements of a structural system so that they satisfy various criteria of performance, safety, serviceability, and durability under various demands. For example, a structure should be designed so that its capacity or resistance is greater than the effects of applied loads. 34 However, there are numerous sources of uncertainty in parameters that contribute to load and capacity. A 11:20) 10(0) ka Qn- Rn mR R,Q Figure 2.5 - Fundamentals of risk-based design evaluations In the presence of uncertainty, it is not simple to satisfy the basic design requirements. Figure 2.5 shows a simple case considering two variables i.e., load Q on the structure and capacity R of the structure. Both Q and R are random variables, and their randomness is characterized by their means mg and mR, standard deviations (IQ and OR, and corresponding probability density functions fR(r) and fQ(q), respectively, as shown in the Figure 2.5. The nominal values of these parameters, Q" and R", used in the conventional safety factor based design approach, are also shown in Figure 2.5. The concept of risk- based design was introduced by Freudenthal (1956). Design guidelines using the LRFD 35 concept are essentially based on the risk based design format. The concept of risk-based design is described by Haldar and Mahadevan (2000) and is summarized below. Referring to Figure 2.5, in deterministic design, design safety is achieved by requiring that R" be greater than Q" with a nominal safety factor given by The nominal capacity Rn is usually a conservative value that is one or more standard deviations below the mean value. The nominal load Q" is also a conservative value typically several standard deviations above the mean value. The safety factor accounts for many factors such as the uncertainty in the load and resistance and how conservatively the nominal load and resistance values are selected. However, the nominal safety factor may not be an indicator of the actual margin of safety in the design. In a deterministic design, the nominal safety factor can be applied to the resistance, to the load, or to both. In allowable stress design, the safety factor is used to compute the allowable stresses in the member from the ultimate stress, and a satisfactory design assumes that the stresses caused by the nominal values of the loads do not exceed the allowable stresses; i.e., Q" < Rn/SF. In the ultimate strength design method, the loads are multiplied by load factors to determine the ultimate loads; i.e., referring to Figure 2.5, Rn>an, where y is load factor. In steel design using the load and resistance factor design (LRFD) concept, the resistance factor (generally less than 1.0) and load factors (generally greater than 1.0) are used to achieve the same objective, and the safety factors 36 are used to modify both the resistance and the loads; i.e., ¢Rn > an, where (b is the resistance factor. The intent of these conventional approaches can be explained by considering the area of overlap between the two curves (the Shaded region in Figure 2.5), which provides a qualitative measure of the probability of failure. The area of overlap depends on three factors: (1) The relative position of the two curves: As the distance between the two curves increases the overlapped area reduces and the probability of failure decreases. The positions of the curves may be represented by the means (mQ and rim) of the two variables. (2) The dispersion of the two curves: If the two curves are narrow, then the area of overlap and the probability of failure are small. The dispersion may be characterized by the standard deviations (0Q and OR) of the two variables. (3) The shape of the two curves: The Shapes are represented by the probability density functions fR(r) and fQ(q) and affect the area of overlap and the probability of failure. In deterministic design procedures safety is achieved by selecting the design variables in such a way that the overlap between the two curves is small. This is commonly accomplished by shifting the positions of the curves through the use of safety factors. A more rational approach would be to compute the risk by accounting for all three overlap factors, and selecting the design variables so that an acceptable risk of failure is achieved. This is the foundation of the risk-based concept. This approach, however, requires 37 knowledge of the probability density functions of the resistance and loads (as in Figure 2.5), which is usually difficult to obtain. Often engineers must formulate an acceptable design methodology using only information on means and standard deviations. Probability theory and structural reliability methods allow safety factors and design variables that are consistent with a desired level of performance (acceptably low probability of failure) to be selected by using information on the means and standard deviations of the resistance and load parameters. The probability density functions of the resistance and loads variables can also be used, if available. Reliability-based design approaches are able to yield more uniform safety in structures and in some cases where design are conservative may also yield a reduction in costs. Structural reliability theory is briefly described in this section. 2.3.1 Background In strength design, the resistance, R (e.g., bending, shear or axial strength) of a member or connection is chosen to withstand the load effect, Q (e.g., the maximum bending moment, shear force or axial force caused by external loads). Resistances and load effects are random variables because of the various uncertainties associated with each. Safety is often assessed in terms of the safety margin, R — Q, or the logarithm of the factor of safety, In (R/Q). The probability of failure in terms of these measures is PF = P[(R - Q) < 01 = Plln (R/Q) < 0] (33-12) To obtain pp the probability density functions (PDF’S) of R and Q must be known. The PDF’S of R— Q or In (R/Q) can then be derived from these, and the areas under these PDF’S corresponding to pp are shown in Figure 2.6. 38 Two common probability distributions used for R and Q are normal and log'norrnal distributions. If R and Q are normally distributed, then R — Q is also normally distributed. If R and Q are lognonnally distributed, then ln (R/Q) = In R —- ln Q is normally distributed. For these distributions, it is useful to transform the random variable R — Q or In (R/Q) into a standard normal variate, U, with zero mean and unit standard deviation: U : (R-Q)—mR_Q aR—Q (2-13) U _ 111(R/Q) — Inln(R/Q) 0'In(R/Q) (2-14) in which m and 0 denote the mean and standard deviation of the subscripted quantity. Hence, m _ m _ pp=P[(R—Q)— R Q (2-15) UR-Q aR-Q f(R_. QIA fin (RlQ)? PF Pi R IQ I In (R70) Figure 2.6 - Probability of failure, pp (Reproduced from Harichandran 2005) 39 m m [7,: = P[1n(R/Q)<0]=P U «M =( ) is the standard normal cumulative distribution function (CDF). A reliability (or safety) index ,6 is defined for independent normally distributed R and Q as mR_ mR —m 5N = Q = 2 Q2 (2-17) (IR—Q 0R +UQ while for independent lognorrnally distributed R and Q with COV less than about 30% it is mln(R/Q) ~ ln(mR /mQ) pm 2 2 2 (2-18) alum/Q) 1le + VQ In terms of the reliability index, the probability of failure is pF = P[U < —,6’J = <1>(—,6’) (2-19) The meaning of the reliability indices are illustrated in Figure 2.7. They measure the distance of the critical point R — Q = 0 or In (R/Q) = 0 (i.e., the origin) from mR _ Q or mm (mg) in units of GR _ Q or O'jn (Is/Q). Equation 2—17 was the basis of an early recommendation for a probability-based structural design code (Cornell 1969), while Equation 2-18 was the basis for the development of the load and resistance factor design criteria for steel structures (Ravindra and Galambos 1978). 40 11(le % Tfl mu=0 U Standard normal CDF Standard normal PDF f (R - Q) A f It. In (R/ Q) flL N 0' I R/Q I I E n( ) | I I I I l I l ' I PF 5 : flN 0' R-Q: E I l _llllllllllll;§+- y ' a mR- Q R-Q 0 min (R/Q) In (R/Q) PDF of (R-O) PDF of ln(R/O) Figure 2.7 - Interpretation of the reliability index ,6 (Reproduced from Harichandran 2005) 2.3.2 First-Order Second-Moment (FOSM) Analysis The capacity R and load effect Q are usually related to several basic random variables that characterize the problem. Thus the means and variances of R and Q must be 41 determined from the statistics of these other random variables. When the relationship between R and Q and the basic random variables is non-linear, it is common to estimate the means and variances of R and Q from those of the basic random variables by linearizing the non-linear relation. Consider a random variable Y derived from another random variable X through Y = g(X). The mean and variance (first two moments) of Y can be obtained exactly only if g( ) is linear or if the PDF of X is known. If g( ) is non-linear and only the mean and variance of X are known, then the mean and variance of Y can be estimated using a first order approximation based on linearizing g(x) about some point x*. Figure 2.8 illustrates this approximation. The equation of the linear function is dg y g( ) x* (x " 35*) (2-20) 451 where, dx I — * x* _ g (x )= slope of the straight line. Equation 2—20 represents a first—order Taylor series expansion of g( ) at x*. For the multi-dimensional case where the random variable Y is a non-linear function of several basic random variables, Y = g(X 1, . . . , X“), the linearization takes the form 42 / / - Linear approximation / / / :1: > x x Figure 2.8 - Linear approximation to non-linear function (Reproduced from Harichandran 2005) n a n y = g(x1*, ..... ,xn*)+ Zilxfixi —x,-*) = a0 +Zai(xi —x,-*) (2-21) i=1 '= The mean of Y based on Equation 2-21 is my 2: a0 + Zai(mxi —.7Cl'*) . (222) 1:1 and the variance is n ~Ziai2 O'X +ZZal-aj Cov(X,-,Xj ) (23) i1: i=1j=1 43 If the Xi are uncorrelated with each other then the covariances COV (X i, Xj) = 0. Equations 2-22 and 2-23 are exact for the special case when g( ) is a linear function of the X ,2 Often the linearization point (x.*,..., xn*) is chosen as the mean point (le ’ ' ° ' ’ an ) . In this case the method is referred to as a mean-value first- order second-moment method, for which Equation 2-22 simplifies mY 2:00 :g(mX1,...,an) (2-24) and for uncorrelated random variables Xi’s, the Equation 2-23 simplifies to 1/2 2 BY 02 8X,- Xi (2-25) 0T” Z: mXi The extent to which Equation 2-24 and 2-25 are accurate depends on the effect of neglecting higher order terms in Equation 2-21 and the magnitudes of the coefficient of variations (COV’S) in Xi. If g( ) is linear and the variables are uncorrelated, Equations 2- 24 and 2-25 are exact. The reliability index, ,6 (in some studies ,8 is termed the safety index), is defined by fl=mflmr coo which is the reciprocal of the estimated COV of Y. B is the distance from my to the origin in standard deviation units. As such, ,0 is a measure of the probability that g ( ) will be less than zero. This is illustrated in Figure 2.6 which shows the densities of Y for two alternate representations of the simple two- variable problem Y = g(R, Q) = 0 discussed in 44 the previous section. Figure 2.6a shows the probability density function (generally unknown) for Y = R - Q. The shaded area to the left of zero is equal to the probability of failure. Observe that if 0R_Q remains constant, a positive shift in mR_Q will move the density to the right, reducing the failure probability. Thus an increase in ,6 leads to an increase in reliability (lower pp). 2.3.3 Hasofer-Lind Reliability Index When R and Q are related non—linearly to the basic random variables X ,- characterizing the problem, use of the reliability indices defined in Equations 2-17 and 2-18, with means and variances of R and Q computed by mean-value FOSM analysis, has the following undesirable features in reliability analysis: Significant errors may be introduced at increasing distances from the linearization point by neglecting hi gher-order terms in the Taylor series expansion. The reliability index is not invariant to different but mechanically equivalent formulations of the same problem (e.g., whether ZxFy — M < 0 or ZxFy - M/Z,C < 0 are used to describe failure in the flexural design of a continuously laterally supported beam). In effect, this means that )6 depends on how the failure criterion is formulated. The invariant reliability index proposed by Hasofer and Lind (1974) is the basis of advanced reliability methods. Failure is expressed in terms of the failure criterion Y: g(xl, xn) < 0, in which x,- is the value taken by the basic random variable X). The function g(xl, xn) is often called the response surface and may be taken to be any one 45 of R — Q, In (R/Q), R/Q — 1, etc. Linearization of g( ) is performed at a special point (x1*, xn*) on the failure surface g(xl, xn) = 0, often called the design (or checking) point. The mean and variance of Y are then computed using Equations 2—22 and 2-23. The reliability index is then defined as the distance of my from the critical value of zero (at which failure occurs) in units of O'y, i.e., flHL = my/Gy (2-27) Computation of the Hasofer-Lind reliability index is performed using the following steps: (1) The basic random variables X) are transformed to uncorrelated standard random variables Zi which have zero mean and unit variance. Z,- = ' (2—28) This may be Viewed as a coordinate transformation from x-space to z-Space. (2) The failure surface g(xl, ..., xn) =0 in x-space is transformed to the corresponding failure surface h(z 1, zn) = 0 in z-space as shown in Figure 2.9 with failure occurring when It ( ) < 0. (3) The reliability index ,6 is defined as the shortest distance between the surface 120 = 0 and the origin. ,BHL is the distance to (21*, zn*) from the origin. 46 Failure boundary _ obtained by linearizing x2 Fallure boundary F '1 b d 7'2 at the design point A, g(x,,x2)=0 a1 ure oun ary A ,.’ h(z,,zz) = 0 / . Design point / Transformation * I x2 9: _ __ _ _ ZZ 9: Inverse l ,3 Transformation l I ' a: T’ x; Z1 Z1 x-space z-space Figure 2.9 - Illustration of failure boundary and design point in x- and z-spaces for two variables problem (Reproduced from Harichandran 2005) (4) The design point in z-space, (31*, zn*), is the point on the failure surface h(z 1, Zn): 0 that is closest to the origin. A numerical optimization algorithm is usually required to find this point. This point must be determined by solving the system of equations (NBS 577 1980) a _ Bh/az, (2 29) Z," — : ,2 _ [flak/azazl Zi* = —0!Zl. ,6 (2—30) h(Z1*, Z2 *, ..... , Zn*) = 0 (2-31) searching for the drrectron cosrnes Z,- whrch minimize ,6. The derivatives are evaluated at the point (zl*, zn*). Note this procedure is equivalent to 47 linearizing the limit state equation in reduced variables at the point (z1*, 211*), and computing the reliability associated with the linearized rather than the original limit state. This is identical to locating the design point (x1*, xn*)in x-space using the inverse of the transformation in step 1, and using FOSM analysis with linearization ofg(x1, xn) at (x1*, xn*). (5) In the original variable space , the design point variables are given by X,*=mxi(l-a’ziflin) (2-32) g (X,*,X2*, . . .,X,,*) = 0 (2-33) Details of this procedure can be found in Madsen, et al. (1986). Note that if g(xl, ..., xn) is linear, the Hasofer-Lind reliability index is identical to that in Equation 2-17. If necessary, load and capacity reduction factors yi for design corresponding to a prescribed reliability index ,6 may then be determined through (Ellingwood et a1. 1980) 71' = Xi*/Xn,i (2-34) In which Xn,i is the nominal or design value of the load and capacity parameter specified in the building standard. This may be the load corresponding to a mean recurrence interval of N years, the mean maximum load during the reference period of T years, or any one of a number of other formulations. Thus the load and resistance factors depends on the way the nominal load and resistances are specified. 48 2.3.4 Approximate Methods for Including Information on Distributions If the basic random variables X,- are normally distributed and the response surface g(xl, xn) is linear, the probability of failure is (—,6’), with ,6 given by Equation 2-18. If the X,- are normally distributed but the response surface is non-linear, then pg: 2 <1>(—,6HL). Many structural problems involve random variables which are clearly non- norrnal. As examples, instantaneous live loads appear to be modeled more appropriately by gamma distributions, at least for relatively small loaded areas (Corotis and Doshi 1977). Studies of extreme wind load (Simiu and Filliben 1975) have Shown that that the annual extreme wind speed due to extra tropical storms has an Extreme Value Type I distribution. It seems appropriate that this information be used in determining ,BHL in order to have szq)(-flHL). The so-called Rackwitz-Fiessler method is often used to in- clude information on distributions of non-normal random variables in the computation of the reliability index (Madsen, et a1. 1986). Non—normal random variables are transformed to equivalent normal random variables prior to computing flHL- The basic idea is to transform the non-normal variables into equivalent normal variables prior to the solution of Equations 2-29 to 2-31. This transformation may be accomplished by approximating the true distribution of variable Xi by a normal distribution at the value 49 X i* corresponding to a point on the failure surface. The justification for this is that if the normalization takes place at the point close to that where failure is most likely, (i.e., min ,6), the estimate of the failure probability obtained by the approximate procedure should approximate the true (but unknown) failure probability quite closely. The mean and standard deviation of the equivalent normal variable are determined such that at the value Xi*, the cumulative probability and probability density of actual and approximating normal variable are equal (Rackwitz and Fiessler 1976). Thus, 0N = ¢I“IEJ) l fi(x:) (2-35) N _ * —ll * JUN mX, ‘Xi ”(D FAXI) 1' (2-36) In which Pi and fi = non-normal distribution and density functions of Xi, and ¢( ) is the N density function for the standard normal variate. Having determined 0,: and mN f X ,- ° the equivalent normal random variable, the solution proceeds exactly as described in Equations 2-9 to 2-31. In as much as the design point variable Xi* changes with each iteration, the parameters O'iN and m? must be recomputed during each iteration cycle. 1 However, since all calculations are performed by computer, this does not materially add to the complexity of the reliability analysis described earlier. The following summarizes the procedure which is used to compute the reliability index ,6 associated with the particular design or, conversely, a design parameter (such as section 50 modulus) for a prescribed ,6, probability distributions, and set of means and standard distributions (or COV) (Ellingwood et al. 1980): (1) Define the appropriate performance function. (2) Make an initial guess at the reliability index ,6 (or design parameter). (3) Set the initial design point values X i* = m Xi for all i. (4) Compute the mean and standard deviations of the equivalent normal distribution for those variables that are non-normal according to Equations 2-35 and 2-36. (5) Compute partial derivatives ah / 0X ,- evaluated at the design point Xi*. (6) Compute the direction cosines a,- as ai = f ah/az, (2-37) [Z(Bh/8z,~)2l/2 zi* = -az, ,3 (2-38) h(z1*, Z2 *, ..... , zn*) = 0 (2-39) (7) Compute new value of X,* using Xi* = mil, " Ctr-flail, (2-40) and repeat steps 4 through 7 until the estimate of a, stabilizes. (8) Compute the value of 6 necessary for, g (X1*,X2*, - - ., Xn*) = 0 2-41) 51 and repeat steps 4 through 8 until the values of 6 in successive iterations differ by some small tolerance (say 0.05). Normally convergence is obtained within 5 cycles or less, depending on the non-linearity of the limit state function. 2.3.5 LRFD Design Criteria The basic design requirement is that the reliability index 6 associated with the failure criterion R — Q < 0 or In (R/Q) < 0 should equal some assigned target value 6). Although different design formats are in use in various countries to accomplish this, the structural design format that is used in the US. is the load and resistance factor design (LRFD) format given by ¢Rn Z 2 10°an (2242) i in which (0 is called a capacity reduction factor and the y,- are called load factors and Rn and Qn,i are the nominal capacity and load effects. The (25 and y,- faetors account for the variability of R and the Qi’s. For example, for dead load (Dn), lifetime maximum wind load (Wu) and arbitrary point-in-time live load (Lmapt), the format is ¢Rn 2 70D" + llL,apth,apt + I/an (2‘43) while for dead load, lifetime maximum live load (Ln) and arbitrary point-in-time wind load (WWW), the format is ¢Rn Z yDDn + yLLn + yW,apth,apt (2'44) 52 For ease of use, instead of using two different live and wind loads, Equations 2-43 and 2- 44 are written as ¢Rn 2 yD D" + ”L" + yWWn (2-45) and, 0R" 2 yDDn + yLLn + yW/Wn (2-46) in which 7L’ = 7L,apth,apr/Ln and 7W’ 2 mflptwnflpt/Wn. Thus, while load factors such as yL, 7L.apt’ etc. are expected to be greater than unity, yLI and yWI are less than unity due to the scaling factors Ln.apt/Ln and Wmapt/Wn, respectively. While several levels of sophistication for reliability—based design exist, two of the common levels are referred to as Level I and Level II methods. Level II methods are based on the computation of a reliability index 6 as described in Sections 2.3.3 and 2.3.4, but 6 must be computed for each design situation (i.e., for each Ln/Dn, Wn/Dn, etc.). Level 1 methods involve the selection of one set of load factors to be applied to all design situations, regardless of Ln/Dn, Wn/Dn, etc., and a resistance factor which depends on the material and limit state. Levels 1 and II can be made equivalent if the load and resistance factors in the Level I format are allowed to vary for each design situation. For operational convenience, practical design criteria in the US. will be of the Level I type in the foreseeable future, with constant load and resistance factors for all design situations. This necessarily means that 6 will be slightly different for each design situation. 53 2.3.6 Selection of Load and Capacity Reduction Factors In order to have a constant 6 for all design situations in LRFD, the factors o and )4 must depend on the particular load combination, strength, and on the mean, variance and distribution of all variables in the limit state equation. If a constant set of ¢ and Vs are prescribed, the associated 6 will deviate from the target value for certain design situations. However, it is possible to select one set of load factors (ys) that minimize the extent of this deviation when considered over all likely combinations of load. While the resistance factors (¢’s) will depend on the material and limit state of interest, the load factors will be independent of these considerations. For any given design situation and load combination, the resistance factor «>11 and load factors y,” that achieve the target reliability 60 are determined using Level II reliability analysis as follows: (1) The failure criterion is defined. e. g., g(R,Qi. . . .. on) = R - 01-. . .- Qn =R -ZQ,- <2-47) i (2) The reliability index and design point associated with Equation 2—47 are determined as described in Sections 2.3.3 and 2.3.4. The basic variables R and Q) (in x-space) are transformed to normalized variables zR and ZQI (in z-space). The values of the normalized variables at the desrgn pornt are zR* and ZQi and the reliability index is 6. 54 (3) The failure criterion is adjusted by uniformly scaling it in the uncorrelated z-space as shown in Figure 2.10, such that the reliability index corresponding to the adjusted criterion is 60. For uncorrelated, normally distributed R and Qi, the factors a” and y," which achieve the target reliability index 60 by matching Equation 2-42 with the adjusted failure criterion are: 11 RI mR =——=—1+a V (248 (15 Rn Rn ( RflO R) ) I m . and, y,” = Q = Q’ (1+aQi’BOVQi) (2—49) Qn.i Qn,i in which 01R and CZQ, are direction cosines associated with the design point in z- space and are * z * (1R — ZR and CIQ. = Q1 (2-50) A ' A respectively, with A = \/(zR*)2 + 2kg, "‘)2 (231) i being the distance from the origin to the design point. In most cases (when mean and nominal values are close) OLR would be negative yielding (1)” < 1, while aQi would be positive yielding n" > 1. Note that R’ and Q,’ are the coordinates of the design point associated with the adjusted failure criterion in x-space as shown in Figure 2.10. 55 Design point Design point Failure boundary , , associated Z Q adjusted to achieve target 3515,03?th .wcrith target Q with target A eliability index re ‘3 ‘ "Y 1“ 6" + reliability h flu Z ’30 Z 0 Failleure boundary ’ index ’ — m— = , = 0 / / / / fl .5 Q g( Q) / / / Inverse / r 4‘ Transformation Q / Z :1: — " — 9: l/ '30 Q Failure boundary Q , l )9 >9 h(zR,zQ) = 0 L ' > r I ZR * ZR R Figure 2.10- Illustration of failure boundary adjustment to achieve target reliability index (Reproduced from Harichandran 2005) The required nominal capacity based on the Level II reliability analysis is therefore 1 R: = W Z 71119,...- (232) II 11 where 1/ i and ¢i are the load and capacity reduction factors for each design situation obtained from the reliability analysis. R: will vary for each design situation (i.e., for each Ln/Dn, Wn/Dn, etc.). On the other hand, a Level I design format which prescribes constant load and capacity reduction factors for all load ratios leads to a nominal resistance 1 R5 = 32 ViQn,z (2-53) I The constant ¢ and yi are then selected in order to minimize an error criterion such as 56 7 60>) = 21le ,- — RI, ,1- p,- (254) overall design situations indexed by j, and pj = weight assigned to the jth design situation (e.g., jth set of load ratios Ln/Dn and Wn/Dn for a load combination involving D, L and W loads). The weights pj are chosen to stress more likely load ratios over less likely ones, and vary for different construction materials. For example, concrete structures typically have more dead load than steel structures, so the Ln/Dn ratios prevalent for concrete structures are smaller than those prevalent for Steel structures. Hence the weight pi assigned for a specific Ln/Dn ratio for a concrete beam would be different from the weight assigned for the same Ln/Dn ratio for a steel beam. Typical weights that have been used to derive the load and resistance factors in the ANSI A58 Standard are given in Ellingwood et al. (1980). In developing the ANSI A58 load and resistance factors (Ellingwood et a1. 1980), subjective judgment was used in setting some of the load factors, and the selection process was performed in stages. For example, for gravity loads involving D + L and D + S load combinations, the optimum value of 70 was about 1.10 for a target reliability of 6) = 3.0. However, this is lower than the dead load factor used in all previous design specifications and it was felt that the profession would not accept such a low value. The minimization of Equation 2-54 was then repeated with m = 1.2 to select the optimum (ti, )1 and ys values under this constraint. Since the n and )3 values resulting from this 57 analysis were similar, the process was repeated yet again, this time taking )1 = )4; in addition to n; = 1.2 in order to simplify the final load criteria. The final gravity load case was specified as U = 1.2 0,, + 1.6 (Ln or 3,) (2—55) The resistance factors resulting from the analysis ranged from 0.78—0.79 for steel and 0.81—0.86 for reinforced concrete. The ANSI A58 specifications, however, were limited to specifying load factors, and the specification of resistance factors were left to the different material specification writing groups. 2.3.7 Capacity Reduction Factors Compatible with Selected Load Factors Once load factors are specified independent of the construction material, the material specification writing bodies are expected to decide on their own target reliabilities 60, and arrive at suitable resistance factors by minimizing Equation 2-54 with respect to ¢, with the )7 held fixed. 2.3.8 Design Criteria with Partial Safety Factors Design criteria in other countries sometimes differ from the LRFD format used in the US. A general but complex format is one that uses partial safety factors for each random variable in the problem. Rather than formulating the problem in terms of a single resistance variable R and several load effect variables Qi as in Equation 2-47, let the overall resistance be given by 58 R =fR(r1,r3, . . ., rm) (2-56) in which the r,- are subvariables that affect the overall resistance, and let the overall load effect be given by Q = ZfQ, <41, 42, ,---,q./,.) (2-57) 1 in which fQ, is the load effect due to the ith load type (e.g., dead, live, etc.), and the qki are subvariables that affect each load effect. A design criterion based on partial safety factors is fR (¢lrn,l r¢2rn,2’ ° ° " ¢mrm,n) Z ZfQi (I’ll-$1.1,- ’ I’2iqn,2i ’ ' ° '9 yni qn,l,-) (2'58) i in which the (t),- are partial factors related to the resistance subvariables, the yki are partial factors related to the load effect subvariables, and r", j and qui are nominal resistance and load values used in the design. The partial safety factors are computed using a Level II reliability analysis with the rj and qki as basic random variables (in x-space). The basic variables are transformed to normalized variables er and qu- (in z-space) and their values at the design point (i.e., I z: and Zjlk- ) are determined. For normally distributed basic variables, the partial safety 1 factors required to obtain a target reliability index 6) can then be computed through m _ ’1' "J 59 m qk- and, yk, = ' (1+a,,k_ 60qu_) (2-60) qn,ki I I in which 05,]. and aqk. are direction cosines associated with the design point in z— 1 space and are Zr. * zqk. or, = J and “ku =———'- (2-61) respectively, with 2 2 2 6=\/Z(z,j*) +:;(zqki*) +....+§(zqk1*) (2-62) J being the distance from the origin to the design point. Again in most cases (when mean and nominal values are close), the a . would be negative yieldin - < l, and a r} g I (Iki would be positive yielding 7k,- > 1.0. For non-norrnal basic random variables, the mean and COV of the random variables cannot be used directly in Equations 2-59 and 2-60. The “normal tail approximation” must be performed at the design point to obtain the mean and COV of the “equivalent normal distribution,” and these must be used in Equations 2-59 and 2-60. The equivalent normal distribution is one whose CDF and PDF match the CDF and PDF of the basic random variable, Xi, i.e., . *— . . * —- . FX. (x,*) = (bf—hi] and fx. (x,*) = —1-¢a[§-'—fl'-] (2-63) I 0- l 0". i ,- 0' where m,- and o,- are the mean and standard deviation of the equivalent normal distribution. The solutions of the above two equations are: 60 a. = ¢(¢“‘)> I fX. (xi*) 811d m,- : Xi * T0i¢—1(FX,- (xi*)) (2-64) m,- and V,- = mi/oi are used in Equation 2—59 and 2-60 to obtain the partial safety factors. As with the LRFD format, the partial safety factors in Equation 2-58 would change with each design situation (i.e., for each Ln/Dn, Wn/Dn, etc.), and an optimization procedure similar to that described earlier is necessary to arrive at constant partial safety factors for a Level 1 design specification. If desired, Equation 2-58 can be converted to the LRFD format in Equation 2-42 by computing a resistance factor 4’ and load factors y,- associated with the overall resistance and load effects, respectively: : fR(¢1r,,,1,¢2rn‘2, ........... ¢mrmm) (4 (2-65) fR(r,,’1, 01,29 .............. rm") and y. _ fQi (ylianli’yzianI’ """""" I’miani) (2-66) ' fQi (9n,1,-,61n,2,-, .............. an, ) 2.4 Studies Related to Reliability Analyses of Structural Members Exposed to Fire As mentioned at the beginning of this chapter, only one study has been reported in which fire load factors were developed using simplified assumptions instead of rigorous reliability analysis (ECSC 2001). Similarly, no explicit work has been reported on the 61 derivation of resistance factors for fire design of steel members except in one case (Magnusson and Pettersson 1981). However, during the last few decades, reliability analyses were performed for estimating the probability of failure and computing reliability indices for steel members exposed to fire. Some of the important studies are summarized in this section. 2.4.1 ECSC Study, 2001 In the ECSC study (ECSC 2001), fire load factors were developed based on the variability of the fire load and by taking into account the effect of active fire protection systems, in reducing the probability of occurrence of a severe fire. Initially, base fire load factors were derived based on the presence of some basic active fire protection systems such as a public fire brigade, and the effect of occupants in reducing the fire. Thereafter, differentiated factors were derived to modify the base fire load factors depending on the presence of other advanced active fire protection systems such as automatic sprinklers, smoke and heat detectors and special fire brigades, etc. The fire load was assumed to be described by the Gumbel distribution, and the fire load factors were calculated through I-fivq [0.577 + ln(—ln cp(0.9p,))]} 71' yq =1.05{ (2-67) {I—qu[0.577+1n(—1n(p))]} where Vq = COV of the fire load, (D = cumulative standard normal distribution function, and p = percentile used for obtaining the characteristic or nominal fire load. If the nominal value is taken as the 80th percentile, then p = 0.8. The value of the target 62 reliability index, 6,, depends on the effectiveness of active fire protection systems present in a building. Fire load factors were obtained by using simplified assumptions instead of detailed reliability analysis (ECSC 2001). For example, the value of 0.9 in the numerator in Equation 2-67 was assumed for the direction cosine of the fire load corresponding to the failure surface instead of obtaining this from reliability analysis (see Equations 2-28 to 2- 34). In addition to the variability of the fire load, there is significant uncertainty associated with other fire design parameters (such as the thickness of insulation, thermal properties (density, specific heat, and thermal conductivity) of insulation, cross-sectional (e.g., plastic modulus, cross-sectional area, etc.), mechanical (yield strength, elastic modulus, etc.) and thermal (density, specific heat, and thermal conductivity) properties of steel, thermal properties (density, specific heat, and thermal conductivity) of compartment enclosures, and opening factors that was not accounted for in the ECSC study. 2.4.2 Magnusson and Pettersson, 1981 Thus far, Magnusson and Pettersson (1981) are the only one who explicitly focused on the derivation of capacity reduction factors for the design of structural members exposed to fire. Magnusson and Pettersson (1981) derived load and capacity reduction factors for an insulated simply supported steel beam that was assumed to be used in an office building. They considered following parameters as random variables: (1) Fire load density, qt 63 (2) (3) (4) (5) (6) Thickness of insulation, (1) Thermal conductivity of insulation, k,- Insulation parameter, x = (Ai/V) x ((1)/kg). A,/ V is the ratio of exposed surface area of insulation to the volume of steel and was considered a deterministic parameter in the reliability study. The mean and COV of K were obtained using the first order Taylor series approximation from the means and COV’s of di and kg. Maximum steel temperature, Ts,max, given as Ts,max = n+Tl+T2+T3 (2’68) where, Tn = the deterministic value of maximum steel temperature given by design curves (graphs) for the nominal values of fire load density, opening factor, (Ai/V) and (di/kl'). T1 = the uncertainty due to variation in insulation parameter, rc. T2 = the uncertainty reflecting the prediction error in the theory of compartment fires and heat flow analysis. T3 = the correction term reflecting the difference between a natural fire in the lab and a real fire in a room. The true capacity, Rf (7) Rf: (Rn, fI-R 1+R2)Mm (2-69) where Raf: nominal value of the capacity calculated according to creep deflection theory as a function of T". R1 = uncertainty measured by a comparison between the theoretical value of Rn,f and laboratory tests. R2 = uncertainty due to difference between laboratory tests and in site fire exposure. Mm = random factor, expressing uncertainty in material strength. Expressing the load effects as Wf= EL(DL+LL) where EL is the random variable expressing the dispersion in I load effects prediction. DL and LL describe the basic variability of the live and dead loads, respectively. The room temperature design of the beam was performed according to allowable stress design methods given in the Swedish Codes. The load and capacity reduction factors were derived for an arbitrarily chosen target reliability index of 2.16 and for a dead to live load ratio of 1.0. The resulting factors were: 0 Capacity reduction factor = 0.77 0 Live load factor = 1.88 0 Dead load factor = 1.14 65 Load factors were to be applied to the mean values of dead and live loads. The value of the capacity reduction factor found from this study is less than 1.0 suggested in most of the fire design specifications such as Eurocode 3 (EN 2005). In Magnusson and Pettersson’s work, the statistics of thickness and thermal conductivity of insulation were assumed due to lack of data. In the US. the test data related to thickness and thermal conductivity of insulating materials is available which should be characterized and incorporated in the reliability studies. Statistics of some model errors (e. g., T3 and R2, etc.) associated with the steel temperature and capacity of the beam were assumed in this study. The fire temperature used in Magnusson and Pettersson’s work were obtained from the curves (graphs) given in the Fire Engineering Design of Steel Structures (Pettersson et al. 1976). These curves give time vs. fire temperatures corresponding to certain values of fire load density and opening factor for a compartment that is assumed to have boundaries made of a material having relatively high values of density and thermal conductivity (brick, concrete). The nominal steel temperature, T", can also be calculated from the graphs given in the Fire Engineering Design of Steel Structures (Pettersson et al. 1976) for a given value of fire load density, opening factor, di/ki and Ai/V. If the compartment boundaries are made of materials other than brick or concrete for which graphs are given, then the manual gives a translation factor, kf, which should be multiplied to the nominal values of fire load density and Opening factor to obtain their effective values. These effective values of fire load density and opening factor should then be used to obtain nominal values of steel temperature from the graphs. In 66 Magnusson and Pettersson’s work, they recognized that kf should be treated as a random variable, but assumed it was deterministic stating that the inaccuracies introduced by this assumption were negligible. Buchanan (2001) showed that thermal properties of bounding surfaces have a significant effect on the fire temperature. The thermal properties of different bounding surface materials differ significantly and their variability should be modeled. The opening factor was also considered to be deterministic in Magnusson and Pettersson’s study. The curves used by Magnusson and Pettersson (1981) for estimating the fire and steel temperatures differ from the new mathematical models (EN 2002, Feasey and Buchanan 2002) being suggested for use in the performance-based approaches. These models enable fire temperature to be estimated through equations for any value of fire load density, opening factor and thermal inertia of compartment boundaries. For reliability analysis, it is convenient and more rational to use the steel and fire temperatures in the performance functions in terms of basic random variables such as fire load density, opening factor, and thermal inertia of compartment boundaries. Magnusson and Pettersson performed the reliability analysis using both Monte Carlo simulations and the first order reliability method based on the Taylor series approximation. Monte Carlo simulation was considered the preferred choice but they stated that it was difficult to check the accuracy of the method (Magnusson and Pettersson 1981). For the first order reliability methods, they developed a linear expressions for the maximum steel temperature, Tsmax, in terms of the basic variables (fire load density, thickness and thermal conductivity of insulation), although the relationship between temperature and the basic variables is non-linear. The capacity of 67 the beam was nonlinearly related to the maximum steel temperature, Twmvc. Magnusson and Pettersson first obtained the mean and COV of TS, max using the first order Taylor series approximation. They then used the first order Taylor series approximation again to obtain the mean and COV of the capacity this time using the mean and coefficient of T5,,nax. This may result in significant errors because of the nonlinear relationship between the steel temperature and the capacity of the beam (Ellingwood et al. 1980). New mathematical models for estimating the fire and steel temperatures and reliability software afford an opportunity to use the steel temperature in the performance function in terms of basic random variables for more accurate reliability analysis. In Magnusson and Pettersson’s study, the effect of active fire protection systems in reducing the probability of occurrence of a severe fire was not included explicitly. Instead, resistance and load factors were developed for an arbitrarily chosen target reliability index of 2.16. Recent studies (Ellingwood and Corotis 1991, ECSC 2001) provide a framework for selecting a target reliability index by incorporating the effect of active fire protection systems in reducing the probability of occurrence of severe fires and correspondingly in reducing the probability of failure under fire. These studies enable the selection of more rational target reliability indices which are based on the effectiveness of active fire protection systems present in a building. The effectiveness and reliability of active fire protection systems is also now better known than three decades ago and can be used for selecting more rational target reliability indices. In Magnusson and Pettersson’s work, the information on the distribution of random variables was not used when the first order analyses were performed. Most of the fire 68 design parameters are non—normal (e.g., the fire load density is best described by the Gumbel distribution (ECSC 2001)), and it is appropriate that information on distributions be incorporated into the reliability analysis. The statistics of the fire loads used by Magnusson and Pettersson were specific to typical Swedish office buildings and differ from those found from the survey of US. office buildings performed by Culver (1976). The statistics of other parameters such as thickness, density, thermal conductivity of insulation and bounding surfaces also differ from country to country, and these parameters need to be characterized for conditions specific to the US. wherever possible. The engineering approach used for the design of steel beams and corresponding design equations used for reliability analysis were also different than those suggested in the AISC Specifications (AISC 2005a). The load factors of 1.0 for dead load and 1.4 for the sustained part of the live load that were used to estimate the applied load effects for fire design are different from the gravity load combination of 1.20,, + 0.5L" suggested in the AISC Specifications for the design of steel members exposed to fire. These different loads will result in different resistance factors. In addition, in the US, the load factors for fire design have been proposed by Ellingwood (2005), and have already been incorporated into the AISC Specifications, and there is a need to develop resistance factors corresponding to these specific dead and live factors. The fire load is a major parameter in fire design, and uncertainty associated with it has a significant effect on the safety of the design. As mentioned in the introduction, the AISC Specifications recommend that the fire load be reduced by 60% if a reliable sprinkler system is installed. The fire load can be reduced appropriately if a separate safety factor 69 is used for the fire load. In Magnusson and Pettersson’s work, a combined resistance factor was developed for all the resistance parameters. Because Steel temperature was not included in the performance function in terms of basic variables, it was not easy to develop a separate safety factor for the fire load. 2.4.3 Beck, 1985 Beck (1985) performed reliability analyses of the flexural strength of Simply supported steel beams that were protected by lightweight insulation material. The insulation thicknesses for beams were obtained using the prescriptive approach instead of engineering approaches currently being promoted. Fire load, opening or ventilation factor, thickness and thermal conductivity of insulation, exposed area of insulation per unit length, volume of steel per unit length, steel temperature, yield strength of steel, and, dead and live loads were taken as random variables. Statistical models were developed for all the random parameters. The probability of failures and corresponding reliability indices were computed for various values of opening factors and for different fire resistance times. Beck found that values of the reliability index varied from zero to 5.0, and concluded that there is considerable variation in the probabilities of failure and the corresponding reliability indices for beams designed according to prescriptive approaches and there is a need to adopt a more rational design approach for ensuring uniform safety in fire design. Magnusson and Pettersson (1974) presented graphical results for the maximum steel temperature, and using these graphical results Beck derived an equation for predicting the maximum steel temperature, Ts,max, which is: 70 Tsmax = 25 + 0.6237", JP — 1500(F, — 0.08)(1 — [0:028]! ) (270) where q, = fire load density, F v = opening factor, and K = insulation parameter = (Al-l V) x (di/ki). Ai/ V is the ratio of exposed surface area of insulation to volume of steel, d, is the thickness of insulation, and k, is the thermal conductivity of the insulation. Beck used the fire load density applicable to a compartment, having an area of 120 to 400m2, from the survey results of US. office buildings performed by Culver (1976). The author first developed the statistical models for basic variables qt, F v and K. The variation of most of the basic parameters was assumed based on judgment and experience due to lack of data. However, now it is possible to characterize the variation based on information available in the literature. The mean and COV of the maximum steel temperature were then obtained through Equation 2-70 by first order approximation (see Equations 2-22 and 2- 23) using the means and COV’S of the basic random variables. This approximation may result in significant errors as the relation between the basic variables and the maximum steel temperature is nonlinear and most of the basic variables are non-normal (e.g., fire load is best described by the Gumbel distribution (ECSC 2001)). The steel and fire temperature models used in Beck’s work were different then the advanced mathematical models being suggested for use in the performance-based approaches. As per new fire models (EN 2002, Feasey and Buchanan 2002), the fire temperature significantly depends on the type of compartment boundaries (Buchanan 2001), but in Beck’s work, the thermal properties (density, specific heat, thermal 71 conductivity) of compartment boundaries were not explicitly included as random variables. In Beck’s work, the information on the distribution of random variables was not used. Most of the fire design parameters are non-normal, and it is more appropriate that the information on distributions be incorporated in the reliability analysis. Beck found the probability of failure and corresponding reliability indices but did not develop resistance factors that incorporated the effect of active fire protection systems. 2.4.4 Li and Fitzgerald, 1996 Li and Fitzgerald investigated the relative safety of a structural steel beam protected by gypsum board insulation. The beam was assumed to be protected by two 5/8 inch layers of gypsum board that would normally provide a two hours fire resistance rating according to the prescriptive design approach. The beam was assumed to be used in a typical U.S. office buildings’ compartment with dimensions of 5 m x 5 m x 3 m. Only the fire load and the opening factor were considered as random variables, and the remaining parameters were assumed to be deterministic. The mean and standard deviation of the fire load and the opening factor were taken from a survey of US. office buildings (Culver 1976). The influence of the opening factor (ranging in values from 0.01m“2 to 0.15m1/2) on the reliability indices was studied for a simply supported beam and it was concluded . 1/2 . . that an opening factor of 0.025m rs most severe for the assumed compartment srze. Reliability indices were also obtained for various values of fire loads using the adverse values of thermal inertia and opening factors by assuming different support conditions. A 72 beam fixed at both ends gave higher values of reliability indices compared to a simply supported beam. This study was limited in scope and considered just two random variables. It was not reported how the reliability analyses were performed (e.g., what was the performance function), however, the structural analysis was based on the procedures published in Fire Engineering Design of Steel Structures (Pettersson et al. 1976). Although the study was limited in scope, it did help to demonstrate that a smaller value of the opening factor (0.025m1/2 in this case) is likely to yield a lower reliability index and thus a higher probability of failure. 2.4.5 Ellingwood, 2005 This work is related to the loads and load combinations to be considered in fire design. Ellingwood showed showed that the probability of coincidence of a fire with maximum values of live load, Ln, roof live load, Lr, snow load, Sn, wind, or earthquake loads at or near the magnitudes given in ASCE Standard 7—02 (2005) is very small, and a structure is likely to be loaded to only a fraction of the design loads when the fire occurs. Therefore, these load combinations need not be considered for fire design. The companion action of 1.2Dn, 0.5L", 0.5L,, and 0.25,, represent the most probable values of load on the- structure at the time of fire. These recommendations were included in the AISC Specifications and in the ASCE Standard 7-02 (2005). Ellingwood suggested that the capacity reduction factors should be developed corresponding to these values of loads. 73 2.4.6 CIB W14, 1983 and 1986 The CIB design guides (CIB W14 1983 and CIB W14 1986) recognized the need for a rational design approach for fire design and provided a complete methodology for the development of load and resistance factors for structural fire design. Some of the recommendations were: 0 The load and resistance factors can be derived in the time domain, temperature domain and strength domain. However, it was suggested that for load bearing members it is more appropriate to develop the safety factors in the strength domain using real fire scenarios. 0 Since fire scenarios vary according to the type of occupancy, height of the building, etc., it was suggested that separate factors be derived for different types and categories of occupancies. 0 Partial safety factors depend on the chosen target reliability index, which in turn reflects the target probability of failure. CIB W 14 (1986) suggests that while specifying safety factors for fire design, the rare occurrence of a severe fire (a fire that can cause structural damage) should be, accounted for. Thus the target reliability index is usually smaller than the index used for deriving safety factors at room temperature. The target reliability index should vary depending on the presence of active fire protection systems such as automatic sprinklers, fire brigade, etc., because these measures affect the occurrence of a structurally significant fire. If the target reliability index is varied, the resulting load and resistance factors also will vary. These guidelines suggests that basic safety 74 factors such as resistance factor may be modified using differentiated factors based on the presence of active fire safety measures. 2.4.7 Ramachandran, 1995a Ramachandran (1995) suggested that the performance of a building and installed fire protection systems may be assessed in terms of the probable area of damage for property protection and probability of death or fatality rate per fire for life safety. Acceptable minimum levels specified for area damage and fatality rate will provide design values for escape facilities, size of compartments, and fire resistance for structural and non- structural members. Subject to these levels, the design of a building can be altered, depending on the presence of active fire protection systems such as sprinklers, detectors, and smoke control systems. Occurrence of the flashover stage during a fire is undesirable as this would adversely affect the stability of structural elements. Ramachandran performed reliability analysis to calculate the probable area of damage and probability of flashover in a sprinklered and unsprinklered building and concluded that for the occupancy type considered, sprinklers reduce the probability of flashover by a factor of 3.0. He suggested that this factor may be used to deterrninine the reduction in the fire resistance requirements for a sprinklered compartment. This study is not related to the derivation of resistance factors for steel members exposed to fire, but, does suggest that sprinklers have a significant influence in reducing the probability of occurrence of a severe fire, and this effect should be considered in evaluating the fire resistance of steel members exposed to fire. 75 2.4.8 Ramachandran, 1995b Ramachandran, Elms and Buchanan pursed the approach proposed in the CIB Design Guides (1983 and 1986) for the development of probability-based designs for fire safety. The authors suggested comparing the fire resistance, RF, and fire severity, SF, in the time domain. The fire severity is the time to fire exposure, and the fire resistance is the duration of time a building element can resist fire. Fire resistance, Rp, and fire severity, Sp, were assumed to follow the normal distribution. The authors suggested that if data is not available to determine the standard deviations, O'R F and O'SF , of Rp and Sp, respectively, a coefficient of variation, r, may be assigned for both Rp and Sp so that OR = rmR and O'SF = 177151, ,where me and mSF , are the means of Rp and F F Sp, respectively. Reliability indices were estimated for different m R F /m S F values by assuming r = 0.15. The authors suggested reducing the requirement of fire resistance in sprinklered buildings according to the sprinkler factor. A procedure for calculating the sprinkler factor was described in the paper. The authors suggested that the actual distribution of Rp and Sp may be close to lognormal but they used the normal distribution for convenience. This study was again based on the prescriptive design approach, and the fire resistance was compared in the time domain, contrary to the suggestion in the fire literature that the 'fire resistance of load bearing members should be compared in the strength domain. However, it shows that sprinklers have a significant effect in reducing the probability of 76 occurrence of fire, and this effect should be accounted for while determining the fire resistance of structural members. 2.4.9 Harmathy and Mehaffey, 1985 In this study, two techniques were suggested for calculating the fire resistance of buildings in such a way that the failure probability remains the same in all cases. Fire load and experimentally determined fire resistance (fire resistance determined through the prescriptive approach) were considered as random variables and the remaining variables were taken to be deterministic. It was proposed that a deterministic value of Opening factor that would result in worst case fire scenarios be used and a method was suggested to determine that value of the opening factor. This was done on the basis that no information was available about the variation of the opening factor. Based on a previous study, the COV of 0.10 to 0.15 was suggested for experimentally determined fire resistance. This study is based on the prescriptive approach that has been frequently used in the past. However, in recent years the use of performance-based approaches for the fire design of steel members has been recommended, which has a different format than that used in this study. 2.4.10 Fellinger, 2000 In this study, a reliability analysis of a concrete beam and slab was done and it was found that comparing the fire resistance in the time domain results in different levels of safety for different types of structures. It was suggested that the fire resistance be determined using the strength domain. 77 2.4.11 Janes, 1995 In this work, a statistical analysis was performed to develop an adoption factor to be used in the analytical models proposed in Eurocode 3 (EN 2005) for determining the limiting temperature of structural members so that the models give the same failure temperature as found in fire tests. In the statistical analysis of steel beams and columns, a target reliability index of 2.0 was used instead of 3.0 used in developing reliability-based codes for room temperature design of steel members. The study was done for a different purpose but suggests that use of a reduced target reliability index of 2.0 is appropriate in the fire design, which is consistent with the recommendations of other studies (Ellingwood and Corotis 1991, ECSC 2001). 2.5 Summary This chapter presented an overview of the: (1) fire resistance evaluation methods for structural steel members, (2) use of reliability theory in developing design specifications, and (3) reliability studies related to steel members exposed to fire. Until now, most steel members were designed using prescriptive design approaches and consequently most of the reliability studies were also performed on steel members designed by prescriptive approaches. However, in the new design specifications, there is a shift toward the use of performance-based design approaches, which are essentially based on the risk-based design format, wherein load and resistance factors are used for minimizing the risk of structural failure. Thus there is a need to develop these safety factors through reliability analysis. There is very limited work that was done explicitly for deriving the capacity reduction and fire load factors for the design of steel members exposed to fire. Only work related to 78 the derivation of fire load factors was performed using simplified assumptions instead of rigorous reliability analysis, and it is not clear whether detailed reliability analysis will yield similar results. In studies related to the derivation of capacity reduction factors, the variation of many parameters was assumed due to lack of data, and also the parameters were mostly specific to European conditions. In the first order reliability methods used in those studies, the means and COV’S of the response variables were determined from the basic variables using the first order Taylor series approximation, which may result in significant errors due to the nonlinear relationship between the response and basic variables. Some of the parameters (e.g., thermal absorptivity or thermal inertia of bounding surfaces which has a significant influence on the fire temperature) were not explicitly treated as random variables in these studies. The fire and steel temperature models used in the earlier reliability studies were different than the current models used by many researchers. These new fire models and reliability-based software afford an opportunity to perform more accurate reliability analysis of steel structures exposed to fire. Over the last few decades, fire research has advanced significantly, and more tests data is available that can be used to better characterize the random parameters. The effectiveness of active fire protection systems in reducing the probability of occurrence of severe fires has been well recognized in the fire literature, and has been better characterized in the last few decades. In earlier studies related to the derivation of resistance factors of steel members, the effect of these active fire protection systems was not explicitly included. There is no reported work related to the derivation of capacity reduction and fire load factors specific to US. conditions for the LRFD of steel members exposed to fire. The 79 2005 AISC Specifications allow the use of the LRFD format for the design of steel members exposed to fire. The load factors for the LRFD of steel members have been proposed in one study, and are included in the AISC Specifications. At present, the AISC Specifications and the commentary to the AISC specifications Specify a capacity reduction factor and a fire load factor specific to one kind of active fire protection system (sprinklers), which are based primarily on subjective judgment. There is a need to develop capacity reduction factors and fire load factors that can account for the effect of all fire protection systems using the US. data wherever possible. 80 Chapter 3 Reliability-Based Methodology for Developing Capacity Reduction and Fire Load Factors 3. 1 General In this chapter, a general reliability-based methodology is presented for developing capacity reduction and fire load factors for the LRFD of steel members exposed to fire. As discussed in Chapter 2, in previous studies, the effect of active fire protection systems (e.g., sprinklers, smoke and heat detectors, fire brigade, etc.) in reducing the probability of occurrence of a severe fire was not explicitly considered. In this chapter, based on the recommendations of some earlier works (ECSC 2001, Ellingwood and Corotis 1991), an approach is suggested for developing capacity reduction and fire load factors corresponding to a preselected target reliability index that accounts for the effect of active 81 fire protection systems (e.g., sprinklers, smoke and heat detectors, etc.) in reducing the probability of occurrence of a severe fire. In the previous studies, the mean and variation of steel temperature was obtained using first order Tylor series approximations, which may result in significant errors. In the methodology proposed in this chapter, the steel tempaerture is included in the performance function in terms of the basic variables such as fire load, thermal absorpitivity, opening factor, etc. For incorporation of the steel temperature into the performance function in terms of its basic variables, an analytical solution of the heat balance differential equation of steel tempaerture was developed. The methodology developed in this chapter is then used in Chapters 5 and 6 to develop capacity reduction and fire load factors for simply supported steel beams and columns. To better understand the performance functions used in the reliability, the engineering approach for designing steel members subjected to fire conditions is described first. 3.2 Engineering Approach for Designing Steel Members Exposed to Fire In the engineering approach, the nominal capacity of steel members exposed to fire, Rnf, is a function of fabrication parameters, F i, and reduced material properties, kj(Ts)Mj, and may be expressed as R,,,=fR(F,, ....... F,,k,(TS)M1, ....... kk(Ts)Mk) (3-1) where the F i are dimensional and sectional properties (e.g., depth of section, cross— sectional area, etc.), and M j are the material properties at room temperature (e. g., yield 82 strength, etc.). kj(TS) are factors that account for reduction in strength and stiffness of steel at elevated temperature, and their values at different values of steel temperature, TS , are specified in the AISC Specifications (AISC 2005a) and Eurocode 3 (EN 2005). According to the AISC Specifications, the ultimate design action (applied axial force, bending moment, or shear force, etc.), Quf, is determined from the load combination given by U = 1.21),, + 0.51., + 0.25,, + T (32) where D", Ln and 5,, are nominal dead, live and snow loads, respectively, and T includes loads or loads effects induced by the fire itself (such as additional bending moment induced due to thermal expansions being restrained by the surrounding structure). The magnitude of the term T in Equation 3-2 will depend both on the type of restraint and on the steel temperature, TS. For a simply supported beam, the term T in Equation 3-2 will be zero because thermal expansion is free to occur. The governing design equation for fire design of steel members can then be expressed as (if R... f 2 QM <3-3) Under fire conditions, both the nominal capacity, Rnf, estimated through Equation 3-1 and the applied load or load effect, T, in Equation 3-2, depend on the steel temperature, Ts, which in turn depends on the design fire (or time-fire temperature curve). The design fire depends on many factors such as ventilation conditions, thermal properties of the boundaries and the fire load (representative of combustible materials present), etc. As 83 mentioned earlier, the Commentary to the AISC Specifications (AISC 2005b) states that while describing the design fire, the fire load may be reduced by up to 60 percent if a sprinkler system is installed in the building. In a similar vein, Eurocode 1 (EN 2002) suggests a reduction in the fire load, while the ECSC study recommends either a reduction or increase in the fire load depending on the intended reliability. This reduction or increase (called fire load factor, yq, in this study) is to be applied to the fire load used in describing the design fire, and will affect the nominal capacity of all members and the applied load effect, T, in Equation 3-2 for restrained members. At elevated temperatures, the strength and stiffness of steel reduces significantly, and if unprotected, steel members fail within a short time. Therefore, steel members are generally protected by insulation or fire protection material to slow down the rise of the steel temperature. The required thickness of insulation can be determined from Equation 3-3 using an iterative procedure, and the fire temperature in the compartment and the steel temperature of the member required for this procedure can be estimated as described in the next two sections. 3.3 Fire Temperature in the Compartment The fire temperature, Tf, can be estimated using a suitable mathematical model from the literature (e.g., SFPE 2000 and 2004, EN 2002, and, Feasey and Buchanan 2002, etc.). In this study, the Eurocode parametric fire model modified by Feasey and Buchanan (2002) was used to estimate the fire temperature under real fire scenarios. The modified Eurocode fire model is briefly described below. The fire temperature in °C during the burning phase is given by 84 T, =1325(1—0.324e_0'2’ —0.204e“-7t -0.472e‘1°t )+20 (3-4) where r‘ is a fictious time in hours related to the real time t through ti = Ft , __(F,,/0.04)2 r 2 (b/1900) (3-5) Fv = opening factor (ml/2), and b = thermal absorptivity of the bounding surfaces of the compartment (WSO'S/mzK). The duration of the burning phase, tab in hours is given by td = 000013-18;— (3-6) V where q, = fire load (MJ/m2 of the total surface area of bounding surfaces). For the cooling phase, the Eurocode gives a reference decay rate (dT/ dtlref = 625°C/hr for fires having a burning period of less than half an hour, decreasing to 250°C/hr for fires with a burning period greater than 2 hours. To incorporate the effect of thermal insulation and the opening factor, Feasey and Buchanan (2002) suggested that the Eurocode reference decay rate be modified to dT _(dT] .lF,/0.04 '37 7:— ,ef ,lb/1900 (3‘7) The fire temperature, Tf, is thus found to be dependent upon: (1) the amount of ventilation (represented by the opening factor, F v); (2) the amount of combustibles in the compartment (represented by the fire load, (It)? and 85 (3) the thermal absorptivity, b, of the compartment boundaries (representing the amount of heat absorbed by the compartment boundaries). 3.4 Calculating the Temperature of Steel Members Once the fire temperature variation with time is known, the temperature of steel elements can be estimated through thermal analysis. The steel temperature can be calculated using any advanced finite element software such as SAFIR (Franssen et al. 2004), ANSYS (2007), etc. However, most design specifications such as the AISC Specifications (AISC 2005a) and Eurocode 3 (EN 2005), allow the steel temperature to be calculated using simple thermal analysis methods such as the lumped heat capacity method. The lumped heat capacity method assumes that the steel section is a lumped mass at uniform temperature. The heat balance differential equation for steel members protected by insulation can then be written as (Buchanan 2001) dT F kl pscs — = [—j —— (Tr ‘ Ts) (3-8) dt V dipscs pscs +0.5(F/V)di,0,-Ci where dT/dt = rate of change of steel temperature, F = surface area of unit length of the member (m2) ,V = volume of steel per unit length of the member (m3), p5 = density of steel (kg/m3), Cs = specific heat of steel (J/kg.K), p,- = density of the insulation (kg/m3), c,’ = specific heat of insulation (J/kg.K), cl, = thickness of insulation (m), k; = thermal conductivity of insulation (W/m.K), T, = steel temperature (°C), and Tf= fire temperature (°C). 86 Equation 3-8 can be written in finite difference form and the steel temperature can then be calculated at any time using a finite difference method that can be implemented in a spreadsheet. However, for incorporation into performance functions used in reliability analysis, a closed-form expression for calculating the maximum steel temperature, Tmmx, is convenient. Therefore, Equation 3-8 was solved analytically to obtain the solution Ts,max : Tf.max — ’16“: _1) + ’15”! + C4e-aI (3-9) where a, _ (£)[ ki ][ pscs ] - (3-10) V dipscs pscs +0.5(F/V)d,-,0,-c,- 625 for td < 0.5 hours xi. = 625 — 250(td — 0.5) for 0.5 S ta, S 2 hours (3_11) 250 for rd > 2 hours 15 z=—-l-In[-:C—§), C4 = emd (stUd) “Tf.max "3). (3-12) a 4 5 ,lF, /0.04 = 3-13 Jib/1900 ( l The steel temperature at the end of the burning phase of the fire is st (1d) = 1345+ CI(€_wd - e—O'an ) + C2(e_wd - fund )+ C3(e'wd — e—lgnd ) -1325e'wd (3-14) 87 __ 429.36! _ 27030, C _ 625.405 h , C- , .——-—, 3— were 1 a_0.2[‘ r a—1.7F 05—191“ (3-15) The maximum temperature of the fire, Tfi max, can be calculated through Equation 3-4 using I = Id. The temperature attained by a steel section protected by insulation, Ts, is thus found to be dependent upon: (1) cl, = thickness, ki = thermal conductivity, Ci 2 specific heat, and p,- = density of the fire protection material or insulation; (2) p5: density of steel , and cs = specific heat of steel; and (3) F/V = section factor (ratio of exposed surface area of insulation to volume of steel per unit length of the member). Equation 3-8 is used to estimate the temperature of steel members protected by insulation. The temperature of unprotected steel members can be estimated through a similar equation (Buchanan 2001). The heat balance differential equation for unprotected steel members is not presented since in the US. steel columns are always protected and steel beams are almost always protected. The codes allow using the lumped mass method, but caution that this method may be overly conservative for certain situations such as for a composite steel beam with a concrete slab on top in which a significant thermal gradient can occur through the depth. In this study, the lumped mass method is used because it is convenient within a reliability-based framework. The error arising from this method because of the assumption of a uniform temperature distribution is accounted for through a model error (or professional factor) that is presented later in this dissertation. 88 3.5 Methodology for Developing Capacity Reduction and Fire Load Factors Capacity reduction and fire load factors can be developed by performing the following steps: (1) Select an appropriate performance function (design equation) for the reliability analyses of a structural member, and identify the random design parameters. (2) Characterize model errors (i.e., the professional factor) to account for the differences in the capacity calculated from the design equation and that measured in fire tests. (3) Obtain statistical parameters such as the mean, coefficient of variation (COV) and distributions of random design parameters (e.g., yield strength of steel, cross- sectional area of steel, fire load, opening factor, etc.). (4) Select a target reliability index, 6,, which reflects the target probability of failure and is a relative measure of safety. (5) Calculate the capacity reduction and fire load factor through reliability analysis. In the succeeding paragraphs, the above steps are elaborated in the context of steel members exposed to fire. 3.5.1 Performance Function for Reliability Analysis 3.5.1.1 Applied Loads Ellingwood (2005) showed that the probability of coincidence of a fire with maximum values of live load, roof live load, snow, wind, or earthquake loads is negligible, and a 89 structure is likely to be loaded to only a fraction of the design load when a fire occurs. Therefore, it is appropriate to use the combination of dead and arbitrary-point-in-time live load for reliability analysis under fire conditions. This is consistent with Beck’s (1985) recommendation. Therefore, the load effect, Wf, for reliability analysis may be calculated as Wf= E(cDAD + CLBLapt) (3-16) where cD and cL = deterministic influence coefficients that transform the load intensities to load effects (e.g., moment, shear, and axial force), A and B = random variables reflecting the uncertainties in the transformation of loads into load effects, E = a random variable representing the uncertainties in structural analysis, and D and Lap, = random variables representing dead and arbitrary-point-in-time live load. The statistics of random variables A, B, E, D and Lap, are presented in the next chapter. 3.5.1 .2. Capacity of Steel Members The actual capacity of steel members under fire can be obtained by modifying the nominal capacity given by Equation 3-1 to R, = P.fR(f,F,, ....... f,F,,k,(t,T,)m,M,, ....... k, (t,T,)m,M,,) (3-17) where P, f), mj, and ts are the following non-dimensional random variables: P= “Professional factor” reflecting uncertainties in the assumptions used to determine the capacity from design equations. These uncertainties may result from using approximations in place of exact theoretical formulas, and 90 from assumptions such as perfect elasto-plastic behavior and a uniform temperature across the section. f,- = Random variable that characterizes the uncertainties in “fabrication.” m,- = Random variable that characterizes uncertainties in “material properties.” ts: Random correction factor that accounts for differences between the steel temperature obtained from models and that measured in actual tests. It is assumed that the random variables fi and m,- are the same as those used for developing LRFD specifications for ambient temperature conditions and their statistics are available in the literature. The statistics of P are specific to each design equation, cannot be generalized, and can be obtained from a comparison between the predicted capacity and test results. The statistics of P specific to design equations used in this study for simply supported steel beams and axially loaded steel columns are presented in Chapters 5 and 6, respectively. The statistics of ts are characterized in Chapter 4. The steel temperature, TS, in Equation 3-17 is a function of many basic parameters (e.g., fire load, opening factor, thermal absorptivity, thickness of insulation, thermal properties (density, specific heat, thermal conductivity) of insulation, and thermal properties (density, specific heat, thermal conductivity) of steel, etc.). The statistics of these basic parameters which significantly affect the steel temperature are developed in the next chapter. The steel tempaerture, TS, can be estimated as described in the previous section. 91 In this study, the steel tempaerture, Ts, is incorporated in Equation 3-17 in terms of basic variables using Equations 3-4 to 3-15. 3.5.1.3 Limit State Equation Using Equations 3-16 and 3-17, the limit state equation for reliability analysis under fire conditions may be written as g(X)=Rf- Wf (3-18) where X denotes a vector containing all the random variables. The probability of failure, pp, of a steel element under fire is Pr = P1800 < 0] (3-19) 3.5.2 Statistics of Random Parameters The design parameters that significantly affect the fire design of steel members were identified in the previous section (Section 3.5.1), and their means, coefficients of variation (COV), and distribution types are characterized in the next chapter. 3.5.3 Probability of Failure and Target Reliability Index Partial safety factors depend on the selected target reliability index, which in turn reflects the target probability of failure. CIB W 14 (1986) suggests that the rare occurrence of a severe fire should be taken into account and therefore the partial safety factors are generally lower than those used at room temperature. In a similar vein, the target reliability index should vary depending on the presence of active fire safety measures such as automatic sprinklers, fire brigade, etc., because these measures affect the probability of occurrence of a severe fire. Correspondingly, the target reliability index is 92 usually smaller than that used for deriving safety factors at room temperature. The probability of failure under fire conditions can be written as (AISC 2005a, ECSC 2001, and CIB W 14 1986) PF = PF,fi Psf (3-20) where pp = target probability of failure (which is the same as that used for room temperature LRFD) pp, fi = probability of failure under fire The probability of occurrence of a severe fire can be expressed as (ECSC 2001, CIB W 14 1986) p5; = 1911921031944; (3-21) where p1 = probability of fire occurrence including the effect of occupants and standard fire brigade (per m2 of floor area and per year) p2 = reduction factor depending on fire brigade type and on time between alarm and firemen intervention (p2 is also the probability Of failure of the fire brigade in stopping the fire) p3: reduction factor if automatic fire detection (by smoke or heat) and! or automatic transmission of the alarm are present p4 = reduction factor if sprinkler system is present (p4 is also the probability of failure of the sprinkler in stopping the fire) 93 Af= floor area of the fire compartment By rearranging Equation 3-20, the probability of failure under fire becomes PF pF,r:_ _ f pg, (3 22) The reliability index, 6, is related to the probability of failure through pp = (IX-6) (323) Therefore, for fire conditions, the target reliability index can be written as _ —1 PF :6: — 1” 1‘— (3-24) psf Since the probability of occurrence of a severe fire varies depending on the presence of active fire protection systems, the target reliability index varies for different design situations. Consequently, the capacity reduction factor and fire load factor will vary for each design situation. The framework of determining the target reliability index from Equations 3-20 to 3-24 was used to develop safety factors on the fire load (ECSC 2001). Using the values for p1, p2, p3 and p4 suggested in the ECSC study, the values of the target reliability index were calculated using Equation 3-24 for typical fire compartments of an office building and are shown in Table 3.1. Under normal conditions, the Eurocode requires a maximum target probability of failure of 7.23x10-5 for building life, which corresponds to a target reliability index of 3.8 (ECSC 2001). We used this value to calculate the target reliability index values shown in Table 3.1. Lower target reliability index values do not mean that the target probability of failure has increased, but rather reflects the presence of active fire protection systems that reduce the 94 probability of occurrence of a severe fire and accordingly reduce the probability of failure under fire. Th e AISC Specifications allow a reduction in fire load for sprinkler systems only. If we consider only the effect of sprinklers in reducing the probability of occurrence of a severe fire, then the target reliability index will depend on the effectiveness/reliability of sprinklers and on the size of the compartment. In the ECSC study, sprinklers were considered highly effective with effectiveness ranging from 95-99.5%. Bukowski et al. (1999) performed a statistical analysis of the Operational reliability of sprinklers and concluded that the reliability of sprinklers lies within 88 and 98%. Based on the combined operational effectiveness and performance effectiveness data published by NFPA, William (2005) assessed the overall reliability of automatic sprinkler systems to be 91%. Column 4 of Table 3.1 shows the target reliability index values calculated from. Equation 3-24 assuming a conservative value of 70% for the reliability/effectiveness of sprinklers. The maximum value of target reliability index is around 2.0. Since sprinklers are installed in most US. office buildings, it is reasonable to use target reliability index values ranging from zero to 2.0 for deriving capacity reduction and fire load factors. . To account for the reduced probability of occurrence of a severe fire, a similar approach to the one presented in the ECSC study was suggested by Ellingwood and Corotis (1991) for fire resistant structural design. They suggested a probability of failure for fire situations that corresponds to 6, of about 1.5, which falls within the range of zero to 2.0 found from the ECSC study. In this study, the capacity reduction and fire load factors are developed for a range of reliability index values (zero to 2.0), both for simply supported steel beams and columns. 95 .Eom 5:: $2 a 5?: E322 Combs 05 523 E 30:35 82865 Tm 033. 5 «DR 33% of. 6qu $6 of 3d mm; we; SN 3d Sud 00m who we; emd a: 2: R..— de 3N owe wed om; mmd S ._ 3: ma; mmd ovd cow de mm; mod .2: we.“ cw; Ed mmd Omm 2 .0 mm; 3.0 :1; an; aflm 8m m: 5.0 No; :4 -.N omm 3.0 mod 34 oo._ 2:0. oom Ed wmd mm; we; _o.m ow— vvd 2.. mm; em; o2 Ed .36 mo; :4 mm 3.0 wed GA on 2 .o «3 mm .3: mom—.33 88338»? 85239:». SEE-Soc a.” E5982. 5:35 5895 8.25 .3 use: .3 .833 N .8 S 3. exe we we Carma ”New“: 35:0 3.83 .858qu 559809 _ ..SEEEm ...oifiam ...oifiam ehWoofl—“M—a ANEV 3.5 .82.,— on: bans—om «swung. 9:.— Page a we 35:32. me mime—395 2: $535.. =_ 3:39? £59395 PE 953 he «echo 2: 2: :e tome: 8:3.» 5?: DEA—«=9. 833 we oueam . fin 035. 96 3.5.4 Reliability Analysis When the performance function (Equation 3-18) is chosen, all the random parameters and model errors (e.g., ts and P in Equation 3-17) are characterized (i.e., their means, COV’s, and distribution types have been established), and a target reliability index is calculated through Equation 3—24, reliability analyses can be performed as detailed in Section 2.2. In this study, the maximum steel temperature obtained from Equation 3—9 is incorporated in the performance function (Equation 3-17) in terms of the basic variables and the reliability analysis is then performed as described in Sections 2.3.2 and 2.3.3. The information about the distribution of the basic random variables is included as described in Section 2.3.4. Any structural reliability analysis software can be used to perform the reliability analyses. In this study, the FERUM (Finite Element Reliability Using Matlab) software (Der Kiureghian 2006) is used to perform the reliability analyses. FERUM is a general purpose structural reliability software written using Matlab. It can be used to perform reliability analysis using different methods, including the first order reliability method (FORM). The output of FORM analysis for a particular design situation includes the reliability index, the probability of failure, the values of all design parameters at the failure/design point, and the direction cosines of the design point for each design parameter. 3.5.5 Determination of Partial Safety Factors When reliability analysis has been performed, the partial safety factors for each of the basic random parameter can be obtained using the direction cosine of the design point for 97 each design parameter. The detailed discussion about the partial safety factors is provided in Sections 2.3.6 and 2.3.8. For a normally distributed random design parameter, X, the partial safety factor resulting from the first-order reliability method (FORM) is given by mx ¢X=X (1+ ax fl,Vx ) (3—25) ’1 where ax , mx, Vx and X" are the direction cosine of the “design point,” mean, COV and nominal value of X, respectively, and B, is the target reliability index. For non-normal random parameters, an explicit equation can be derived for the partial safety factor if the PDF and CDF are simple functions (e. g., a Gumbel distribution). The safety factor can be determined from Equation 3-25 using the mean and standard deviation of the equivalent normal variable at the design point if the PDF and CDF are complex functions (e.g., a lognormal distribution). The procedure for estimating the mean and standard deviation of the equivalent normal variable at the design point is explained in Sections 2.3.4 and 2.3.8. 3.5.6 Combination of Partial Safety Factors into Single Capacity Reduction Factor For convenience in design, the variability of all design parameters except for the fire load was accounted for through a combined capacity reduction factor instead of using a separate partial safety factor for each design parameter. The detailed discussion about the combination of partial safety factors into a single capacity reduction factor is provided in Section 2.3.8. 98 The partial safety factors, q); obtained through Equation 3-25 for each individual design parameter except the fire load can be combined into a single capacity reduction factor, (gr, through _ fR(Xn,q9¢an,1a¢2Xn.2, ........... ¢an,m) _ 3-26 fR(Xn,q9Xn,l,Xn,2, .............. Xan) ( ) ¢f where, d,- and Xn,i are the partial safety factors and nominal values for all parameters related to the capacity of steel members except the fire load, respectively, and qu is the nominal value of the fire load. Fire load is a major parameter in fire design, and uncertainty associated with the fire load has a significant effect on the safety of the design. Therefore, the variability of the fire load on overall safety is accounted for through the specific partial safety factor on the fire load. The use of a separate safety factor for the fire load also enables the fire load to be reduced for certain target reliability index values as demonstrated later in Chapters 5 and 6. As mentioned in the introduction, the Commentary to the AISC Specifications (AISC 2005b) states that the fire load be reduced by 60% if a reliable sprinkler system is installed. In a similar vein, Eurocode 1 (EN 2002) suggests a reduction in the fire load, while the ECSC study recommends either a reduction or an increase in the fire load depending on the intended reliability. These recommendations also motivated use of a separate safety factor for fire load. Using this approach, the effect of active fire protection systems (e.g., automatic sprinklers) in reducing the probability of occurrence of a severe fire is accounted for through reliability analysis and often results in a reduction in the fire 99 load as demonstrated later in Chapters 5 and 6. This reduction or increase (called the fire load factor, yq, in this study) is to be applied to the fire load used in describing the design fire. 3.5.7 Optimal Capacity Reduction and Fire Load Factor The capacity reduction factor obtained through Equation 3-26 and the fire load factor obtained through Equation 3-25 will vary for each design situation. For ease of design, it is desirable to have a single optimal capacity reduction factor applicable to all design situations. In addition, for fire design of steel members, the AISC Specifications recommend using dead and live load factors of 1.2 and 0.5, respectively. Therefore, the optimization procedure described in Section 2.3.6, and summarized below is used to develop optimal capacity reduction and fire load factors corresponding to dead and live load factors of 1.2 and 0.5, respectively. The required nominal capacity (e. g., moment, shear, or axial capacity) based on the first order reliability method (FORM) is 1 Rig] = W Z YiHQn,i (3-27) II 1’ . . . . . where 7,- and ¢i are the load and capacrty reduction factors for each desrgn srtuatron obtained from Equations 3-25 and 3—26, respectively. The nominal capacity corresponding to constant dead and live load factors is 1 1 Rn =—ZYiQn.i (3-28) ¢f i 100 where the 7,- are 1.2 and 0.5 for dead and live load, respectively. The value of gt in Equation 3-28 can be selected by minimizing the objective function 8(¢f ) = ZlRfij - R1. ,- 12 P j (329) where 171' is the weight assigned to the jth design situation. In this study, each design situation was assigned the same weight. 3.6 Summary In this chapter, a general reliability-based methodology is proposed for developing capacity reduction and fire load factors for the design of steel members exposed to fire. As discussed in Chapter 2, active fire protection systems, especially sprinklers, have a significant effect on the occurrence of a severe fire, and this effect should be accolmted for while calaculating the fire resistance of steel members. In the methodology proposed in this chapter, the capacity reduction and fire load factors correspond to a preselected target reliability index that accounts for the effect of active fire protection systems (e.g., sprinklers, smoke and heat detectors, etc.) in reducing the probability of occurrence of a severe fire. Contrary to previous studies (Magnusson and Pettersson 1981, and Beck 1985), instead of approximating the mean and variance of the steel temperature, the steel temperature was expressed in terms of the basic variables in the performance function, and the information on the distributions of random parameters was used. In the ECSC study, fire load factors were obtained using simplified assumptions, whereas in this work the fire load factors were derived through rigorous reliability analysis. 101 To illustrate the proposed methodology, in Chapters 5 and 6, capacity reduction and fire load factors are derived for simply supported steel beams and columns in US. office buildings exposed to fire. 102 Chapter 4 Statistics of Random Parameters and Model Errors In Chapter 3, the design parameters that affect the fire design of steel members were identified. In this chapter, the design parameters that significantly affect the fire design of steel members are chosen as random variables, and their statistics are obtained. In Chapter 2, it was concluded that there is not much information available about the varation of random fire design parameters based on US. test data. In this chapter, the information related to US. conditions is used wherever possible. Chapter 3 descibed two types of models that are used for the fire design of steel members: (1) the thermal models used to estimate the temperature of steel member, and (2) the structural models (or design equations). The thermal model for estimating the steel temperature was presented in Section 3.4. The model error associated with this model is characterized in this chapter based on experimental data. The errors associated with the structural models are specific to design equations and cannot be generalized and are 103 characterized for simply supported steel beams and columns in Chapters 5 and 6, respectively. 4. 1 Statistics of Random Parameters The design parameters that significantly affect the fire design of steel members were chosen as random variables, and their means, coefficients of variation (COV), and distribution types are summarized in Tables 4.1 and 4.2. The statistics of the arbitrary- point—in-time live load, fire load and ratio of floor area to total surface area of the fire compartment are specific to US. office buildings. The remaining parameters are general and apply to steel buildings of all use categories. The statistics of the dead and arbitrary- point-in-time live loads in Table 4.1 were reported by Ellingwood (2005) and Ravindra and Galambos (1978). The statistics of the remaining parameters in Table 4.1 were reported by Ravindra and Galambos (1978), and were assumed to have normal distributions. Table 4.1 - Mean, COV and distributions of design parameters related to loads Variable Characterizes variation in Mean COV Distribution A the transformation of dead loads 1.00 004 normal into load effects the transformation of live loads B into load effects 1'00 0'20 normal E structural analysis 1.00 0.05 normal D dead loads l.05*nominal 0. 10 normal L apt arbitrary-point-in-time live load 0.24*nominal variable Gamma Note: The COV of the arbitrary-point-in-time live load depends on the tributary area (Ravindra and Galambos 1978) and is given as: 0.82[1-0.00113(A7—56)] for 56 _<_ ATS 336 square feet 0.56[l-0.0001865(A7—336)] for AT> 336 square feet 104 Table 4.2 - Mean, COV, distributions and nominal values of fire design parameters No. of Dis tri- Nominal Variable Data Mean COV . , button Pomts . * 2 90th Frre load, q f 564 MJ/m 0.62 Gumbel percentile Ratio of floor area to total area, 1610 0.192 0.23 lognormal mean Af/AI Opening factor, Fv # 1*nominal 0.05 normal mean Thermal conductivity of normal weight concrete (NWC), k 18 1.747 W/m.K 0.171 normal mean Specific heat of WC’ cp 15 856 J/kg.K 0.062 normal mean Density of NWC, p 12 2258 kg/m3 0.069 normal mean Thermal conductivity of 105 lightweight concrete (LWC), k 0.372 W/m.K 0.199 Gumbel mean Specific heat of LWC, cp ‘3 826 J/kg.K 0.062 Gumbel mean Density of LWC, p 22 1344 kg/m3 0.069 normal mean Thermal absorptivity of NWC, @ 0 5 2 1830 Ws ”/m K 0.094 normal mean b/vwc Thermal absorptivity of LWC, @ 0 5 2 640 Ws ' /m K 0. 107 normal mean wac Thermal absorptivity of gypsum 16 0 5 2 423 Ws "/m K 0.09 normal mean board, bg Thermal absorptivity of a compartment having a 50/50 mix $ 0 5 2 of NWC and gypsum board as 1127 WS .. /m K 0-10 normal mean boundaries, bmix Thickness of fire protection materials (1' & nominal+lll6 0.20 lognormal ’ l . _ # inch 0.05 normal mean ( l) spray applied materials nominal mean (2) gypsum board systems Density of fire protection 3 materials, pi 10 307 kg/m 0.29 normal , , 1 8 3 0.07 lognormal mean (1) spray applied materials 745 kg/m mean (2) gypsum board systems Thermal conductivity of fire P'O‘w'on Tig‘236ocki' at 24 0.187 W/m. K 0.24 lognormal temPe'a‘U’e ° ,' . 40 0.159 W/m. K 0.28 lognormal mean (1) spray applied materials mean (2) gypsum board systems Note: * based on the survey of 23 office buildings in US. and # assumed @ derived from the statistics of density (p), thermal conductivity (k) and specific heat (cp) for respective types of concretes $ derived using bNWC and bg (see Equation 4-8) & Obtained from Carino et al. (2005) 105 Raw experimental data was analyzed as discussed below to obtain the statistics of all parameters in Table 4.2 except for the fire load. To determine if data is reasonably fitted by a particular distribution goodness-of-fitness test was performed using Kolmogorov- Smimov (K-S) test. A significance level of 5% was used while performing K-S test. Normal, lognormal, Extreme Values and Gamma distributions were considered and the the one which best fitted the data per K-S test was selected. 4.1.1 Fire Load Density The fire load is based on the quantity of combustible materials present in a fire compartment, and is, a measure of the total energy released in a fire. For calculating the fire temperature, the fire load is generally expressed as the fire load density (ECSC 2001) l , CIf=-A—Z(ll/ixmiXHc,iXMi) (4-1) f where qf = fire load density (MJ/m2 of the floor area), Af = floor area of the fire compartment (m2), 111,- = de—rating factor for assessing the protected amount of fire load of the material i, mi = factor between 0 and l describing the combustion behavior of material i, He, i = net calorific value of material 1' (MJ/kg), and M,- = mass of material i (kg). HQ,- x M,- represents the total amount of energy released from material i assuming complete combustion. Complete combustion of some materials does not occur in a fire and this is accounted for by using a factor m. However, because there is a lack of experimentally substantiated and verified values, the factor m may be neglected 106 (Pettersson et al. 1976). De-rating factors are applied to account for the quantity of enclosed combustibles (such as paper in metal filing cabinets) that will not burn in a fire, and the resultant fire load is known as the de-rated fire load. Fire loads have been historically established by surveys of typical buildings in various use categories (e.g., office buildings, residential buildings). Culver (1976) reported statistics of the total and de-rated fire load for 23 typical U.S. office buildings. Culver reported the fire load as lb/ft2 of floor area that was converted into fire load density using HG, i = 17.5 MJ/kg (calorific value of wood) and m = 1.0. The obtained statistics of fire load density are: (1) Total fire load density 0 Mean = 624 MJ/m2 of floor area 0 COV = 0.60 (2) De-rated fire load density 0 Mean = 564 MJ/m2 of floor area 0 COV = 0.62 Another survey of two office buildings was done in the US. in 1995 (Caro and Milke 1995). Both, compartmented and open types of offices, were surveyed and the de-rated fire loads were reported as lb/ft2 of floor area. These were converted into fire load density using Hc,i = 17.5 MJ/kg (calorific value of wood) and m = 1.0. The obtained statistics of fire load density are: (1) Fire load density-open plan offices 107 0 Mean = 1243.4 MJ/m2 of floor area 0 COV = 0.39 (2) Fire load density-compartmented offices 0 Mean = 1135.7 MJ/m2 of floor area 0 COV = 0.11 The statistics of the fire load density from the two surveys mentioned above (Culver 1976, Caro and Milky 1995) are compared in Table 4.3 with fire load densities reported in the literature. Table 4.4 shows the fire load densities obtained from another reference (Yii 2000). Tables 4.3 and 4.4 show a significant scatter in the statistics of fire load densities for office occupancies obtained from surveys carried out in different countries. The large scatter can be attributed primarily to the different furnishings used in different countries, different surveying techniques, and use of different values of combustion and de-rating factors. Even within the US, the three fire load surveys reported completely different values. The mean fire load reported by Culver (1976) is much smaller than that reported by Caro and Milke (1995). On the contrary, the COV of the fire load reported by Culver is much higher than that reported by Caro and Milke. For this study, statistics of the de—rated fire loads reported by Culver (1976) were chosen instead of those given in the more recent survey of Caro and Milke (1995) for the following reasons: 0 In the 1995 survey, the sample size was very small (2 buildings) compared to the 1976 survey (23 buildings). 108 Table 4.3 - Comparison of statistics of fire load density (MJ/m2 of floor area) in offices Fire load density Reference Country/year 2 Remarks Mean ( MJ/m ) COV NCES (2001) 420 0.30 Harmothy (1985) 434 0.35 Switzerland 800 Bukowski (1986) US. 932 (l942&l957) 624 0.60 total fire load 0‘1"” (1976) ”‘8' (1976) 563 0.62 de-rated fire load , l 136 0.1 1 compartmented offices lk 9 . . 1995 Caro and M1 e (l 95) U S ( ) 1243 0.39 open plan offices JCSS (2001) 600 0.30 Kumar (1997) India (1997) 328 0.75 Japan (1964) 1085 0.17 Netherland .. (1963) 263 Yn (2000) 385 Newzeland 664 0.33 998 Table 4.4 - Comparison of fire load density in offices (Yii , 2000) Reference Fire Load Density Mean Standard Deviation COV (MJ/mz) (MJ/mz) Mabin.( 1994) 21 10 1149 0.54 Barnett( 1984) 224 - Narayanan (1994) 476 233 0.49 F.E.D.G (1994)** 800 , - Petterson (l976)** 600 - B.S.F.S.E. (1993) 420 - Swedish data** 41 1 334 0.81 US. data* 555 625 1.13 European data* 420 370 0.88 Swiss data* 750 - Duch data* 410 330 0.85 French data* 330 400 1.21 Notes: *CIB W14(l983),Table A 1.3.4, ** Swiss data 109 The COV of 0.62 reported by Culver is significantly higher than the COV of 0.39 (open plan offices) and 0.11 (compartmented offices) reported by Caro and Milke. The larger COV is statistically significant and will cater for the large variation in fire loads. In a recent study in Europe (ECSC 2001), a mean of 420 MJ/m2 and a COV of 0.3 for fire load density were used for developing the safety factor on fire load. The combustion factor m = 0.8 used in the ECSC study (ECSC 2001) for reducing the fire load is not used in this study. This will yield conservatism and will account for the higher values of mean fire load reported by Caro and Milke. One of the authors involved in the 1995 survey was contacted to ascertain the reasons for the significantly higher values of mean fire load compared to those reported by Culver. The author was of the opinion that the survey carried out in 1995 had a very small sample size and suggested that it is more reasonable to use the values reported by Culver for reliability analyses. Fire engineering guideline documents, including the [PEG (International Fire Engineering Guideline 2005) caution that mean values should not be used for design purposes since half of the buildings would be expected to exceed this value and recommended use of the 90th or 95th percentile values for design purposes (Bukowski 2006). Buchanan (2001) also recommended that the design fire load should be that which has less than 10% probability of being exceeded in the 50 year life of a building. For the present study, the 90th percentile of the fire load was chosen as the nominal value. The fire load was assumed to follow the Gumbel distribution (ECSC 2001). 110 4.1.2 Ratio of Floor Area to Total Surface Area of the Compartment Culver (1976) reported the fire load per unit floor area of the compartment. For calculating the fire temperature, the fire load needs to be converted to per unit area of the total surfaces of the compartment. This conversion can be done using Ar qt : q f E— (4‘2) where q, = fire load per unit total surface area of the compartment, (1f: fire load per unit floor area of the compartment, A f: floor area of the compartment, and A, = total surface area of the compartment. The ratio Af/A, varies for each compartment and should therefore be treated as a random variable in the reliability analysis. No statistical infonnation is available in the literature about this ratio. Culver (1976) reported the range of floor areas for 23 office buildings in the US, but did not explicitly report the height of the rooms. Therefore, the height of the rooms was assumed to be 12 feet to establish the statistical parameters of the ratio Af/At. Additionally, structural and architectural drawings of three representative office buildings in Detroit were examined to establish the mean, COV and distribution of the ratio Af/At. The statistics of the ratio Af /At obtained for the three office buildings were combined with those obtained from the data reported by Culver (1976). The combined mean, COV and distribution of the ratio Af/A, are: 0 Mean 2 0.192 111 0 COV = 0.23 0 Distribution = lognormal 4.1.3 Opening Factor The opening factor represents the ventilation conditions present in a fire compartment. The duration and severity of the fire depends on the value of the opening factor, which in turn depends on the sizes of windows and doors in a compartment, and can be calculated through F. = Av\/Hv / A. <4-3) where, Av = area of the openings and Hv = height of the openings. A building and its structural components are first” designed for room temperature conditions and then for fire. The values of the opening factor for a fire compartment can be accurately estimated from the architectural drawings of a building and is not likely to be significantly different from the design or nominal values. Therefore, it is reasonable to treat the opening factor similar to the dead load in reliability analysis. For the opening factor we assumed the nominal values to be the mean values, a COV of 0.05, and a normal distribution. Typical low, medium and high values of opening factors in actual building compartments are 0.04 mm, 0.08 ml/2 and 0.12 ml/2 (Beck 1985). To determine the range of nominal values of the opening factor the architectural drawings of three typical office buildings in Detroit were analyzed. In general, the opening factor was between 0.02 ml,2 and 0.3 1/2 m . 112 4.1.4 Thermal Absorptivity of Compartment Enclosure The thermal absorptivity, b, of the compartment boundaries is a measure of the amount of heat absorbed by the compartment boundaries and may be calculated through b=\/k—fib—p <4-4) where, k, p and cp are thermal conductivity, density and specific heat of the bounding material, respectively. The thermal absorptivity is a function of temperature, but the Eurocode 3 allows room temperature properties to be used for design. We performed a detailed analysis to study the effect of thetemperature variation of thermal absorptivity on the steel temperature, and the results indicated that it is reasonable to use room temperature values of thermal absorptivity. There is no information available in the literature about the variability of the thermal absorptivity of bounding materials. However, some researchers have reported thermal properties of some commonly used bounding materials such as normal and lightweight concretes and gypsum board. These reported room temperature thermal properties were used to characterize the variability of thermal absorptivity of normal and lightweight concretes and gypsum board as described below. Thermal properties (density, thermal conductivity and specific heat) of gypsum boards reported by different researchers (Carino et al. 2005, Manzello et a1. 2003, Mehaffey et al. 1994, Thomas 2002, and Wullschleger and Wakili 2008) were first used to obtain the thermal absorptivity, bg, through Equation 4-4. The statistics of bg based on these calculated values are: . Mean of bg = 423.5 WSO'SlmZK 113 0 COV of bg = 0.23 0 Distribution of bg = lognormal The mean value is close to the value of 410 WsO.5/m2K reported by Buchanan (2001) for gypsum board. In case of normal and lightweight concretes, all three corresponding thermal properties were not available for a particular tested specimen. Therefore, first the statistics of density (p), thermal conductivity (k) and specific heat (cp) for both types of concretes were obtained using test data. Thermal properties were reported by Harmathy and Allen (1973), Lie and Kodur (1996), Shin et al. (2002), Schneider et al. (1981), and Whiting et al. (1978) for normal weight concrete, and Harmathy and Allen (1973), Stukes et al. (1986), and Whiting et al. (1978) for lightweight concrete. For normal weight concrete, Schneider et al. (1981) included test data obtained in six other studies which was also used for characterizing thermal properties of normal weight concrete. The statistics derived from these test data are also shown in Table 4-2. The mean and variance of thermal absorptivity for both concretes were then estimated analytically as shown below assuming that p, k and cp are independent: Elbl = EL/kpc, l= Tfif. (krdk Tfifprmdp 10.]? f.,, (c, >ch = Eli; l>< Elx/El>< EL]; l (4-5) Elf]: ENE)? = E[k]>< E[p]x Elcp] (4-6) 114 2 equaled—(150]) (.7) where E [.] and Var L] are the mean and variance. The statistics of thermal absorptivity obtained for normal and lightweight concretes are: (1) Thermal absorptivity of normal weight concrete, bNWC 0 Mean of bNWC = 1830 WsO'SlmZK . Cov of wac = 0.094 0 Distribution of wac = normal (2) Thermal absorptivity of normal weight concrete, wac . Mean of wac = 640 WSO'S/mZK 0 COV of bLWC = 0.107 0 Distribution of wac = normal The mean value of wac = 1830 WsO'S/mzK compares well with the value of 1900 0.5 2 0.5 2 Ws /m K reported by Buchanan (2001). The mean value of wac = 640 Ws /m K compares well with the value of 660.6 Wso°5/m2K reported by Kirby et al. (1994) for lightweight concrete blocks. The distribution of thermal absorptivity for both types of concrete was obtained through Monte Carlo simulations. Buchanan (2001) studied the effect of two types of bounding materials (normal weight concrete having b = 1900 WsO'SlmzK and gypsum board having b = 410 WsO'S/mzK) on the fire temperature. A typical commercial office building constructed from a mixture of these materials on the walls and ceiling would give values of fire temperature inbetween 115 those obtained by using either of the individual materials (Buchanan 2001). Therefore, statistics of thermal absorptivity, b, were also obtained for a compartment assuming that 50% of total surface area was constructed of normal weight concrete and the other 50% of gypsum board. The total thermal absorptivity of this compartment can be expressed as 0.5Ab +0.5A,bNWC 173+ch = t g (4-8) A, The mean and variance of bg+ch for this mixed compartment were estimated as E[bg+ ~wc] = 0.5E[bg 1+ 0.5E[bNWC] (4-9) Var[bg+ we] = 0.52(Var[bg D2 + 0.52(Var[bNWC])2 (4-10) and yielded . Mean of 1.8,ch = 1127 wSO'S/mZK . cov of 1.8,ch = 0.10 0 Distribution of bg+NwC = normal To study the fire and steel temperatures likely to occur in real fire scenarios, Kirby et al. ( 1994) conducted 9 fire tests using different materials (such as lightweight concrete blocks, autoclaved aerated concrete slabs, fluid sand, ceramic fiber, and fireline plasterboard) as walls, roof and floor of the compartment. The combined value of thermal absorptivity in all tests ranged from 350-755 WsO'SlmzK. The statistics of thermal absorptivity developed in this study effectively cover the values of thermal absorptivity used by Kirby et a1. (1994). 116 4.1.5 Thickness of Insulation or Fire Protection Materials Steel members may be protected either using spray applied fire protection materials or board systems. Carino et al. (2005) studied the variation of thickness of spray applied fire protection materials used in the World Trade Center (WTC). They observed that the average thickness is generally higher than the specified thickness and that the thickness is distributed lognormally. Their results were used for the COV and distribution type for insulation thickness. Because the thicknesses of fire protection materials used in the WTC were determined using prescriptive approaches, the mean of the insulation thickness was not taken from this study. Instead, based on the analysis in the report and conversation with a fire proofing expert (Ferguson 2008), the mean was taken to be 1/16- inch higher than the thickness required using performance—based design. There is no information available on the variability of thickness of board materials, but since they are produced under controlled conditions, the nominal thickness was taken as the mean value and the COV was assumed to be 0.05. Insulation may come off when a member is exposed to fire. The probability of insulation getting damaged or falling off under fire conditions may be determined by performing the tests on protected steel members. The events of insulation falling off during different tests may then be used for estimating the probability of insulation getting damaged or falling off during the fire conditions which could then be included in structural reliability study. At present, there is no information about the probability of falling off for insulation, therefore, in this study; it is assumed that insulation is not damaged under fire conditions. 117 4.1.6 Thermal Conductivity and Density of Insulation Bruls et al. (1988) studied the variation of thermal conductivity at different temperatures. Although, thermal conductivity varies with temperature, they concluded that since the failure of structural steel elements generally occurs at a temperature of 400 to 600°C, the v thermal conductivity corresponding to a critical temperature of 500°C can be used in design. Statistical analysis of the thermal conductivity in the temperature range of 400-600°C for eight representative materials used in the US. (five reported by Bentz and Prasad 2007, two reported by Carino et al. 2005, and one tested at Michigan State University) was performed. The mean, COV and distribution type established from this statistical analysis are:' 0 Mean of thermal conductivity of spray applied fire protection materials 0. 187 W/m. K 0 COV of thermal conductivity of Spray applied fire protection materials 0.10 0 Distribution of thermal conductivity of spray applied fire protection materials = lognormal Room temperature values of density reported in the literature for different fire protection materials (Bentz and Prasad 2007, and Carino et al. 2005) were used to obtain its statistics. The mean and COV of the density of spray applied fire protection materials are given in Table 4.2. Due to insufficient data, it was not possible to estimate a distribution and a normal distribution was assumed. To reiterate: 0 Mean of density of spray applied fire protection materials = 0.187 W/m. K 118 0 COV of density of spray applied fire protection materials = 0.10 0 Distribution of density of spray applied fire protection materials = lognormal Different types of board materials can be used as fire protection materials (e.g., fiber- silicate or fiber calcium silicate boards, and gypsum plaster (Buchanan 2001). Thermal properties of all of these boards are generally not easily available because of their proprietary nature. However, thermal properties of fire rated type X gypsum boards have ‘ been reported by various researchers and were used to obtain the statistics of thermal conductivity and density. Statistics of thermal conductivity of gypsum board materials in the temperature range of 400-600°C were obtained using test data (Bentz and Prasad 2007, Carino et al. 2005, Manzello et al. 2003, Mehaffey et al. 1994, Sultan 1996, and Thomas 2002) and are: 0 Mean of thermal conductivity of board materials = 0.187 W/m. K o COV of thermal conductivity of board materials = 0.10 0 Distribution of thermal conductivity of board materials = lognormal Room temperature values of the density of fire rated type X gypsum board reported by different researchers (Carino et al. 2005, Mehaffey et al. 1994, Samuel et al.2003, Thomas 2002,Tsantaridis et al. 1999, and Wullschleger and Wakili 2008) were used to obtain the statistics that are shown in Table 4.2. Buchanan (2001) reported that typical values of thermal conductivity are 0.15 W/m.K and 0.20 W/m.K, respectively, and typical values of densities are 600 kg/m3 and 800kg/m3, respectively for two types of board materials (i.e., fiber-silicate or fiber 119 calcium silicate boards, and gypsum plaster). The mean density of 745 kg/m3, and mean thermal conductivity of 0.16 W/m.K, fall within the range of reported values. Therefore, although the statistics of density and thermal conductivity were obtained using test data of gypsum boards only, they should adequately represent other types of board materials as well. 4.2 Model Errors for Thermal Models 4.2.1 Model Error for Steel and Fire Temperatures The maximum temperature of steel sections estimated using Equations 3-4 to 3-8 differs from that measured in actual fire tests. Reasons for the differences include: o The actual heat absorbed by the compartment boundaries may be different than that used in the model through use of the thermal absorptivity, b. 0 The actual ventilation conditions may be different than those accounted for through use of the opening factor, F v. 0 Differences in the actual duration of burning of the fire from that predicted using Equation 3-6. 0 The approximation and assumptions used in the models for estimating fire and steel temperatures. To account for the differences in calculated and measured steel temperatures, the model error was characterized as described below, both for steel beams (three sided exposures) and steel columns (four sided exposure). The experimental temperature of steel elements has been reported by many researchers but most of these tests were carried out under standard fires instead of real fires, and thus 120 cannot be used to estimate the error arising from the fire models. Kirby et a1. (1994) carried out a series of nine real fire tests and recorded the temperature of protected and unprotected steel elements. The tests were performed for a range of fire loads (380 — 760 2 . . . . 1/2 MJ/m of floor area), for different opening condrtrons (F v = 0.0029 — 0.062 m ), and various types of materials were used as compartment boundaries in order to represent all possible real fire scenarios. Foster et al. (2006) reported the temperature of four protected steel columns. In this test, the fire load was 720 MJ/m2 of the floor area, and the opening [/2 factor was 0.043 m . The model error for the temperature of steel beams, tsb, was characterized using the test data reported by Kirby et al. (1994), and the model error for the temperature of steel columns, tsc. was characterized using the test data reported by Kirby et al. (1994) and Foster et al. (2006). tsb has a mean of 0.98 and COV of 0.11, and tsc has a mean of 1.05 and COV of 0.13. Both, tsb and tsc were best described by the Gumbel distribution. In the last decade, many real fire tests were carried out all over the world, especially in the UK. In most of these tests the steel beams were unprotected, and therefore, reported steel temperatures cannot be used for characterizing the model error. In almost all of these tests, steel columns were protected but various parameters (e. g., type, thickness and properties of fire protection material, type, size and thermal properties of bounding materials) required as input data for estimating the temperature of columns were not 121 explicitly reported. Therefore, the experimental temperatures recorded in these tests could not be used. 4.2.2 Model Error for Reduction Factors The strength and stiffness of steel reduces significantly at elevated temperatures. Most design specifications give values of the reduction factors for strength and stiffness at different steel temperatures. Values of the yield strength reduction factor, ky, and the modulus of elasticity reduction factor, kE, given in the AISC Specifications (AISC 2005a) at different steel temperatures, T5, are plotted and compared with test data reported by different researchers (Chen et al. 2006, Gayle et al. 2005, Kelly and Sha 1999, Kirby and Preston 1988, Li et al. 2003, Makelainen et al. 1998, and Outinen and Makelainen 2004) in Figure 4.1. There is some scatter in the reported values of ky and k5, even though the steel material is generally very similar. This may have more to do with different definitions of the yield strength and different techniques of testing than differences in the material. However, to account for the scatter in the data, model errors were established for the reduction factors by comparing the values given in the AISC Specifications with the test data. Only values at temperatures between 400°C and 600°C were used for characterizing the model errors for the reasons described in Section 4.1.6. The model error for k6,), has a mean of 1.03 and a COV of 0.12 and is best described by the Gumbel distribution. The model error for ke, E has a mean of 1.13 and a COV of 0.15 and is normally distributed. 122 1 0.8 - Ill 8 0.6 ~ 0.4 ~ 0.2 k5: EscrsyEs o l I I f X l. 0 200 400 600 800 1000 0 200 400 600 800 1000 Temperature (°C) Temperature (°C) (a) Yield strength reduction factor (ky) (b) MOdUIUS 0f elasticity reduction factor (k5) Figure 4.1 - Comparison of yield strength and modulus of elasticity reduction models with test data ke,y and ke,E are correction factors which account for the differences between the nominal reduction factors (e.g., yield strength reduction factor, k =Fy(Ts)/Fy, or modulus of elasticity reduction factor, kE=ES(TS)/ES) specified in the codes and those measured in tests. In fire tests, actual ratios of Fy(TS)/Fy and ES(TS)/ES are used and the resulting measured capacity is based on these. When measured capacity is compared with the design capacity based on the nominal values of F),(TS)/Fy and ES(TS)/Es, the resulting difference includes the effect of both the uncertainties associated with the design equation and that arising from the difference of actual and nominal ratios of Fy(Ts)/Fy and E S( Ts)/E 3. Therefore, in such situations the uncertainty associated with reduction factors is implicitly included in the professional factor P. Herein, the uncertainty associated with 123 reduction factors was characterized separately. Through probability concepts, the uncertainty associated with the reduction factors and that associated with the design equation can be separated if the design equation is simple (e.g., for laterally restrained beams). However, separating the two model errors is difficult when the design equation is complex as for laterally unrestrained beams. Therefore, the uncertainty associated with the professional factor, P, for both types of beams (laterally restrained and unrestrained) and steel columns was characterized as described later in Chapters 5 and 6, respectively, that implicitly includes the uncertainty associated with reduction factors. Accordingly, errors associated with reduction factors are not used as a separate factor in this study. 4.3 Summary This chapter presents the statistics of random parameters obtained from the literature and analysis of raw experimental data. Information applicable to US. conditions is used wherever possible. All random parameters characterized in this chapter except for the fire load, ratio Af/At, and arbitrary-point-in-time live load, are applicable for the reliability analysis of buildings of all use categories. The statistics of the fire load and arbitrary- point-in-time load for different types of occupancies are reported in the literature. However, not much information is available about the ratio Af/At, and these should be obtained from architectural drawings of buildings. Model errors associated with thermal models were also characterized based on the experimental data. This chapter presents more complete information for the variables and model errors compared to previous studies.The statistics of the random variables and model errors derived in this chapter form part of the input data for deriving capacity 124 reduction and fire load factors for simply supported beams and columns in Chapters 5 and 6. 125 Chapter 5 Capacity Reduction and Fire Load Factors for Steel Beams Exposed to Fire In this chapter, capacity reduction and fire load factors are developed for simply supported steel beams for the flexural limit state using the methodology described in Chapter 3 and using the statistics of random parameters and model errors presented in Chapter 4. The bending capacity equations for laterally restrained simply supported beams given in the AISC Specifications (AISC 2005a) and those for laterally unrestrained simply supported beams given by Takagi and Deierlein (2007) are used. Predictions of structural capacity under fire are still relatively new and evolving. With improved understanding of structural behavior under fire, performance equations may change and future design 126 refinements may be necessary. However, the methodology for deriving capacity reduction and fire load factors described in Chapter 3 and the random parameters and model errors characterized in Chapter 4 are general and may be used for deriving capacity reduction and fire load factors using improved design equations. 5. 1 Derivation of Capacity Reduction and Fire Load Factors 5.1.1 Performance Functions for Reliability Analysis 5.1.1.1 Applied Moment under Fire According to the AISC Specifications, the applied moment, Muf, is determined from the load combination given by Mufz 1.2MD + 0.5ML + 0.2M5 + T (5-1) where, MD, ML and M5 are nominal dead, live and snow load moments, respectively, and T includes moments induced by the fire itself (such as additional bending moment induced due to thermal expansions being restrained by the surrounding structure). In Section 3.5.1.1, it was concluded that it is appropriate to use the combination of dead and arbitrary-point-in-time live load for reliability analysis under fire conditions. Thus, for reliability analyses, the applied moment, Maj, under fire conditions may be expressed in terms of basic variables as Maj: E(cDAD + cLBLapt) (5-2) 127 where CD and cL = deterministic influence coefficients that transform the load intensities to moments, A and B = random variables reflecting the uncertainties in the transformation of loads into load effects, E = a random variable representing the uncertainties in structural analysis, and D and Lap, = random variables representing dead and arbitrary- point-in-time live load. The statistics of A, B, E, D and Lapt are given in Table 4-1. 5.1.1.2 Moment Capacities of Beams under Fire Moment Capacity of Laterally Restrained Beams The nominal moment capacity of a simply supported, laterally restrained steel beam exposed to fire can be expressed as where Z).C = plastic section modulus, F v = yield strength of steel at room temperature, and ky(TS) = yield strength reduction factor that depends on the temperature, T3, of the steel member. The actual moment capacity can be obtained by modifying Equation 5-3 to M f = P b lflzxky(tsts)mlF y (5'4) where f1 and m1 are non-dimensional random variables that characterize uncertainty in ZyC and Fy, respectively, and their statistics are given in Table 5.1. tsb is the model error for steel temperature that was characterized in Section 4.2.1. Steel temperature, Ts is function 128 of many parameters (see Equations 3-4 to 3-8) and theirs statistics are given in Table 4.2. Pbl is the professional factor (model error) that is characterized in the next section. Table 5.1 - Mean and COV of room temperature design parameters Variable Characterizes variation in Mean COV m1 Yield strength, F y 1.03 0.063 m2 Modulus of elasticity, E 5 1.04 0.045 f 1 Plastic section modulus, Zx 1.03 0.034 f2 Elastic section modulus, SI 1.02 0.035 f3 Radius of gyration, ry 1.00 0.016 f4 Moment of inertia, IV 1.02 0.050 f5 Cross-sectional area , A s 1.03 0.031 Moment capacity of laterally unrestrained beams Takagi and Deierlein (2007) compared the design provisions of AISC with finite element simulations for laterally unrestrained beams exposed to fire and reported that the AISC specifications predict strengths 80—100% higher than the simulated results. Therefore, the use of current AISC fire design provisions for reliability analysis was considered inappropriate, and the alternative equations proposed by Takagi and Deierlein (2007) which have a format similar to that of the AISC design provisions were used. Takagi and Deierlein proposed equation for inelastic lateral torsional buckling only, and maintained the current AISC equation for elastic lateral torsional buckling of beams. Using these proposed equations, the nominal moment capacity of laterally restrained steel beams exposed to fire can be expressed as 129 For Il 34,01): A CX (Ts) M", = M,(TS)+{M,,(TS)—M,(Ts) 1— (5-5) MT.) For A > /l,(Ts): at? (T ) 2 7r Mmf = :1— ES(TS)IyG(TS)J + Iwa[ 2 S J (5-6) ry ry where 2 2, (T5) = l \[E5(TS)G(TS)JA 1 1+ 1+ 4 Cw S, F21 (Ts) (5-7) SJ, 2 FL(TS) 1y G(TS)J . M p (T3) = Z xF y (Ts ) = reduced plastic moment M, (T5) = SxFL (TS ) = elastic moment at the onset of yielding F L (TS ) = F p (T, ) - F, (Ts ) = temperature dependent initial yield stress F p (T3) = k p (T3 )F y = temperature dependent proportional limit F, (T5) = ky (TS)F, F r = residual stress at ambient temperature which is specified in the AISC Specifications as 69 MPa, TS < CX (TS) = 0.6 + 250 _ 3.0 (5-8) 130 ,1 = I/ry = Slendemess ratio, L = span length, ry = radius of gyration about minor axis, Sx = elastic section modulus, Z = plastic section modulus, Iy = moment of inertia about minor axis, J = torsional constant, Cw = warping constant, A = cross-sectional area, E S(TS) = kEEs 2 reduced elastic modulus at temperature TS, E s = elastic modulus of steel, G(TS) 2 reduced shear modulus. ky, kp and k]; are the yield strength, proportional limit and elastic modulus reduction factors, respectively, and their values for different steel temperatures are given in design specifications (AISC 2005a and EN 2005). The strength of laterally unrestrained beams predicted by Takagi and Deierlein’s equations is quite similar to the Eurocode 3 design provisions. However, Dharma and Tan (2007) reported that the Eurocode 3 design provisions are too conservative, especially for temperatures less than 500°C. The Eurocode 3 strength design equations for beams were proposed by Vila Real et al. (2003) based on both numerical simulations and experimental investigations of beam strength. Later, Vila Real et a1. (2004) also found that their equations were overly conservative for loading other than uniform bending, and suggested using a moment distribution adjustment factor similar to the one used in Eurocode l for the room temperature design of laterally unrestrained beams. Although similar to Eurocode 1, the AISC provisions for room temperature design include a lateral torsional buckling modification factor, Cb, for non-uniform moment diagrams, but Takagi and Deierlein did not explicitly report whether or not the C], factor should be used with their equations. Beam strengths predicted by Takagi and Deierlein’s 131 equations using Cb were compared with experimental data (Mesquita et al. 2005) and were found to be conservative even if the Cb factors are used. Therefore, Takagi and Deierlein’s equations were used with the C b factors. The actual moment capacity of laterally unrestrained beams can be obtained by modifying Equations 5-5 to 5-8 to For xi. S 2., (tSst): 21 Cx (tsts) Mf : Pb2 Mr(t.sts)+{Mp(tsts) _Mr(tsts){l—m] (5'9) r Sb 3 For Ii. > 1,.(lsts); 7f MfZPbZ llf3ry 2 72m E t T r m2Es(tsts)f4IyG(tsts)J+f4Iwa[ 2 S( Sb 3)] (5‘10) 3 y where 2 A, ('3'st) = 7r JmZES (tSbTS)G(TS )JfSAS 1 1+ 1+4 CW ( fZSx ] F2L(tsts) (5'11) fZSx 2 FL(Ts) f4ly C(Ts)‘, M p (tst5) = le xml F y (2‘5st ) 2 reduced plastic moment M ,(tst5) = fszFLUSbTS ) = elastic moment at the onset of yielding F L (tst5) = F p (tsts) — F, (tst5) = temperature dependent initial yield stress F p (t Sst) = k1,,(tstS)m1Fv 2 temperature dependent proportional limit Fr(tsts) = ky(tsts)Fr 132 F r = residual stress at ambient temperature which is specified in the AISC Specifications as 69 MPa C(tsts) = m1 Es (tsts ) /(1 + U) , where v is the poission’s ratio, and t T C tsTs =0.6+—-S-9—S-S3.0 5-12 x( b ) 250 ( ) where fi and mi are non-dimensional random variables that characterize uncertainty in fabrication and material parameters as given in Table 5.1. PM is the professional factor (model error) that is characterized in the next section. 5.1.1.3 Performnace Function for Reliability Analysis The performance function for reliability analysis of a beam can be expressed as g(X) = M;— Ma,f ‘ (543) where X denotes a vector containing all the random design parameters. The probability of failure, pp , of a steel beam under fire is PF = P1800 < 01 (5-14) 5.1.2 Model Error (Professional Factor) for Moment Capacity Equafions To account for the difference in the measured capacity of a laterally restrained beam in a laboratory and that predicted by Equation 5-3, the model error, Pblr was characterized using the test results presented by Kruppa (1979) and Wainman (1992). Kruppa (1979) reported results for ten tests conducted in Germany and France and six tests conducted in 133 Japan. Wainman (1992) reported the applied loads and failure temperature of two beams. Pbl has a mean of 0.99 and a COV of 0.11, and is best described by the lognormal distribution. To account for the difference in the measured capacity of laterally unrestrained beams in a laboratory and that predicted by Equation 5—5 or 5-6, the model error, sz, was characterized using the test results presented by Mesquita et al. (2005). They tested 5 beams of varying span lengths subjected to different applied loads. Each beam was replicated three times yielding 15 test results. They reported both the failure moments and failure temperatures which were used for characterizing, PM. PM has a mean of 1.67 and a COV of 0.19, and is best described by the Gumbel distribution. sz'has a significantly high mean because the model for calculating the nominal capacity of laterally unrestrained beams is very conservative. 5.1.3 Probability of Failure and Target Reliability Index In Section 3.5.3, it was shown that for typical U.S. office buildings the target reliability index is likely to vary from zero to 2.0 depending on the size of the fire compartment and on the effectiveness/reliability of active fire protection systems. Therefore, in the next section, capacity reduction and fire load factors are developed for this range of target reliability index values. 5.1.4 Reliability Analyses Twenty five steel beams (10 laterally restrained and 15 laterally unrestrained) ranging in length from 3 m (10 ft) to 13.7 m (45 ft) with live loads ranging from 2.4 kPa (50 psf) to 134 4.8 kPa (100 psf) were selected for the reliability study. The AISC design specifications were used to first design the beams for ambient temperature conditions. The same beams were then designed for fire exposure (b = 640 Wso'5/m2K and F v = 0.02 mm) using the engineering approach described in Section 3.2, and the required thickness of insulation to withstand the design fire was determined from Equations 5-1, 5-3, 5-5 and 5-6 using an iterative procedure. As suggested in most codes, a capacity reduction factor of 1.0 was used for the initial design for fire. The beams were assumed to be protected by spray applied fire protection materials, which is generally the case in the US. The spans of the beams, dead and live loads, insulation thicknesses, and the section factors for laterally restrained and laterally unrestrained beams are given in Tables 5.2 and 5.3, respectively. The rate of temperature rise of a protected or unprotected structural steel member exposed to fire depends on the section factor, or massivity factor, which is a measure of the ratio of the heated perimeter to the area or mass of the section (Buchanan 2001). The section factor is an important parameter because the rate of heat input is directly proportional to the area exposed to the fire, and the subsequent rate of temperature increase is inversely proportional to the heat capacity of the member (equal to the product of the specific heat and mass of the steel segment). Generally a section with a lower section factor will experience lower temperatures. Therefore, in Tables 5.2 and 5.3, sections with lower section factors have smaller insulation thicknesses. 135 83w: Rees 2.35 5.3 328 A35 :38 838 sweat $3.8 2-5 Beam 8 9: m2 , 2: mom a: o8 emm SN RN .-s 5:26. she so: so: an: 5.: 2w: ABS $0.: God and :5 32.0.2... 8.2 2.3. 3.2. 2.2. ens. See enmm meow 8.? QR 5:. 5:235 2 be 6m. : 82: 3.9 see see 52: 5.8 82: see 8e: 33 35 92: em. _ N mm... 3.2 M3: ene 33 3.2 2 .w 3: .520. 6on 2 _.: EL: 3:: 5.8 62: see 6:: see :2: £3 .33 one: 3.2 8.2 93 mm: ea: 86 mod med one 3.2 520. a: as so 50 88 as as as a: a: at Sam 9: 2.2 e; is 3.0 is mes 2e Re 8a a flomwuem 2:23 chess 86:5 one; 323 30.33 8.23 85:.» 2ow 2ow seem in an we 5 3. mm x. B S 3 3:5 3353.... mama—55 5:33.?— Lou now: 2.52. Btoansm 39:? 105958.. 33.53. mo 83.5.3.5 I Nd 033—. 136 23$ €38 83$ 33% 838 2-5 2: we «3 S; omm TE 3.3; 558m 8w: 3.8 an: 85,: ans :5 $25.2... $3 3.3 one. 2.2 one 55 5:23:— 6de A38 808 3.9 fine 83 2.3 25 w? 35 w? Maw 26 use. 3.8 $08 €08 ave £08 EB 33 can Ed :.m 9% m3 mom E5 83 as as G: a: ct 5mm Ed N3. 03 In: 84“ E 3395 $wa 55;» 33m? 8:56 5:25 58: m5 Sn 2: «E 2: 3:5 238.23 Sod; 323 5.03 320 $3.: 63$ $38 5.3 338 5.3: 2-5 :L 2: 2: we a; 02 mmm .8 ma OR 75 Stan—Euam 8mg am: 3va and :w: 5.: :3; fiod 6&2 five :5 $25.25 3.3 3.2 9.: 2% mm? 3.3 3.5. mm: 8.8 3.: E5 5:235 :2: an: so? 3on 8on $0.8 81: £58 Snow 8on E): .5335 8.2 3. _ m m? ct: me: E o, 3% 3.2 cum 3: E72 $9 9.: EA: $2: :2: 6:: $0.9 $0.9 $0.9 8on E3 .5335 :2 8.: 3.0 m2: 3.: NE: 0; Ga 3.0 mo: 6%. G: 53 as 62 88 a: as 88 a: a: 2: 5mm ~52 2.2 Ed 3.0 Ed 3d 8s 26 Re 8a E eta; 8133 0223 93.3 3.923 $ow 85;» 85;» 2ow 233 Escomeaom in an an E 3 E E 8 S 5 3E: Sevens; 89:95 5:35.“: ..8 com: 253: .5239; 39:6 355.533: b.303— uo 35.5%.:— I «em 035. 137 5.1.4.1 Reliability Analyses of Laterally Restrained Beams FORM analysis was performed for each design situation (each of the 10 laterally restrained beams) using the performance function given in Equation 5-13. Using the direction cosines obtained from the reliability analysis, the partial safety factors for each design parameter were obtained through Equation 3-25 for each design situation. These individual partial safety factors, except for the fire load, were then combined into a single capacity reduction factor using the procedure described in Sections 2.3.8 and 3.5.6: .2 ($1211),((zrhszmgFy Mn.f ¢f (545) where ¢,- are the partial safety factors for each design parameter obtained through Equation 3-25, and Mnf is the nominal bending capacity of a laterally restrained beam calculated through Equation 5-3. Thus 10 different capacity reduction factors (one for each beam) were obtained. Thereafter, a single optimized capacity reduction factor, (bf, corresponding to dead and live load factors of 1.2 and 0.5, respectively, was obtained using the optimization procedure described in Section 3.5.7. Euqal weights were used for all beams. The dead and live load factors of 1.2 and 0.5, respectively, are proposed in AISC Specifications (AISC 2005a) (see Euqation 5-1). Since the capacity reduction factors were obtained for ,8, values ranging from O to 2, this procedure was repeated for each [3, value. A similar procedure was used to obtain the fire load factors corresponding to each value of ,6,. 138 5.1.4.2 Reliability Analyses of Laterally Unrestrained Beams A procedure similar to that used for the reliability analyses of laterally restrained beams was followed for the unrestrained beams, and the capacity reduction factors were derived for fit ranging from O to 2. As mentioned earlier, the equations proposed by Takagi and Deierlein (2007) for calculating the moment capacity of laterally unrestrained beams are very conservative. The mean of the experimentally obtained capacity was found to be 1.67 times the nominal capacity. Since the capacity reduction factor depends significantly on the conservatism inherited in the design equation, it is generally not possible to have the same capacity reduction factor for laterally restrained and unrestrained beams. However, for ease of use, the same capacity reduction factors as those for laterally restrained beams were maintained for unrestrained beams, and to cater for the conservatism in the design equations, an additional factor, referred to as the capacity adjustment factor is proposed. ' A constant capacity adjustment factor, (1)0 = 1.67 yielded uniform reliability for all beams. In other words, the design capacity given by Equations 5-5 and 5-6 is to be multiplied both by grand ¢a. 5.2 Results 5.2.1 Capacity Reduction Factor The plot of the capacity reduction factor, gr, vs. the target reliability index is shown in Figure 5.1. The value of @c for a given target reliability index, ,6,, is 139 1.0 for ,6, s 1.25 f = (5-16) 1.5 — 0.4,6, for 1.25 S ,6, S 2.0 1.6 - 1.2 Q >3 0.8a "B § é 04< 7q 0 l l 7 l l Figure 5.1 - Capacity reduction and fire load factors for steel beams According to the AISC Specifications, the capacity reduction factor for fire design is 0.9 and is the same as that for room temperature design. However, most other codes suggest that a capacity reduction factor of 1.0 be used (e.g., in the Eurocode 3, the partial safety factor m is 1.0 for fire design). Results obtained in this study indicate that the nominal strength need not be reduced if the target reliability index is less than 1.25, which in turns depends on the effectiveness of active fire protection systems in reducing the probability of occurrence of a severe fire. Since most office buildings in the US. are required to have automatic sprinklers, the target reliability index is not likely to exceed 1.25 (see Table 3.1 and Section 3.5.3) even if automatic sprinklers are not as effective as commonly assumed. Therefore, using a capacity reduction factor of 1.0 is considered reasonable for most design situations. 140 5.2.2 Fire Load Factor The plot of the fire load factor, yq, vs. the target reliability index is shown in Figure 5.1. The nominal value of the fire load was taken as the 90th percentile (Buchanan 2001, Bukowski 2006). The value of yq for a given target reliability index, ,6,, is O.4+0.4,B, for ,6, £1.25 yq = (5-17) 0.15 + 0.6,6, for 1.25 S ,6, S 2.0 The yq is to be applied to the fire load which is then used in describing the design fire (time-fire temperature curve) using Equations 3-4 to 3-7. When the target reliability index is less than 1.42, the fire load factor given by Equation 5-17 is less than 1.0 indicating that the fire load can be reduced as suggested in the ECSC study and Eurocode 1 (EN 2002). The commentary to the AISC Specifications states that the'fire load may be reduced by up to 60% if a sprinkler system is installed. The maximUm reduction should be considered only when the automatic sprinkler system is considered to be of the highest reliability, i.e., having reliable and adequate water supply, supervision of control valves, and regular schedule for maintenance in accordance with NFPA recommendations (NFPA 2002). The reduction in fire load specified in Figure 5.1 depends on the target reliability index, which in turn depends on the effectiveness of active fire protection systems in reducing the probability of occurrence of a severe fire. The proposed approach is more general and enables the reduction in fire load to be specified for sprinkler systems of all categories i.e., having low, high or medium reliability, as well as for other active fire protection systems. 141 In the ECSC study, sprinklers were given very high effectiveness ranging from 95- 99.5%. Bukowski et a1. (1999) carried out statistical analysis of the operational reliability of sprinklers and concluded that it lies between 88% and 98% with a mean value of 93%. Based on the combined operational effectiveness and performance effectiveness data published by NFPA, William (2005) assessed the overall reliability of automatic sprinkler systems to be 91%. Based on these studies, sprinklers of low and medium reliability may be assigned reliability/effectiveness values ranging from 88% to 93%, and then the target reliability index can be obtained using the methodology described in Section 3.5.3. Using these target reliability index values, the fire load factors can be obtained from Equation 5- 17 (or from Figure 5.1). Thus the proposed approach enables the reduction in fire load to be specified for sprinkler systems having lower effectiveness in reducing the probability of occurrence of a severe fire. For example, for a fire compartment size of 200 m2, the following fire load factors are obtained for sprinklers with varying effectiveness: 0 Sprinklers of high (98%) effectiveness: ,8, = 0, n, = 0.4 0 Sprinklers of medium (93%) effectiveness: ,8, = 0.73, yq = 0.69 o Sprinklers of low (88%) effectiveness: fit 2 1.09, yq = 0.84 5.2.3 Validity of Capacity Reduction and Fire Load Factors for Multiple Fire Scenarios The capacity reduction and fire load factors shown in Figure 5.1 were developed for a fire compartment assumed to be constructed of lightweight concrete blocks having b = 640 0.5 2 . . . . . . . Ws /m K and havrng relatively poor ventrlatron conditions represented by an openrng 142 1/2 . . . factor, F v = 0.02 m . In reality, the compartments may be constructed usrng drfferent bounding materials such as gypsum board, lightweight concrete blocks, a combination of gypsum board and normal weight concrete, or by using other types of lining materials on walls and roofs. Depending on the type of bounding materials of the compartment, the thermal absorptivity, b, may have different values. Kirby et al. (1994) used different types and combinations of lining materials in nine natural fire tests. The tests were carried out primarily for studying equivalency between standard fire and realistic fire . . . . 0.5 2 scenarios. The value of thermal absorptrvrty in all tests ranged from 350-755 Ws /m K. The statistics of thermal absorptivity derived in this study (values given in Table 4.2) effectively cover the range of thermal absorptivity values used by Kirby et al. (1994), and are used in the succeeding analyses. Similarly, the ventilation conditions in different compartments may vary considerably. Opening factors of 0.04 mm, 0.08 ml,2 and 0.12 m“2 are typical low, medium and high values in actual building compartments (Beck 1985). Based on the architectural drawings of three typical office buildings in Detroit, it was found that in general the opening factor may have values between 0.02 m”2 and 0.3 mm. Therefore, opening factor values of 0.02 mm, 0.08 ml,2 and 0.12 ml,2 were used in the reliability analyses. As shown later, higher opening factor values generally yield higher reliability index values, so opening factor values higher than 0.12 m1/2 were not considered. After establishing the range of thermal absorptivity, b, and opening factor, F v, to be used for reliability analysis, nine fire scenarios obtained from different combinations of 143 opening factors and thermal absorptivity (see Table 5.4) were selected to validate the capacity reduction and fire load factors derived above. For these nine fire scenarios, six beams, 2 laterally restrained (B2 and B8 in Table 5.2) and 4 laterally unrestrained (B3, B7, B13 and B15 in Table 5.3), were designed for fire conditions using the capacity reduction and fire load factors shown in Figure 5.1. For laterally unrestrained beams, the capacity adjustment factor of 1.67 was also used. Each beam was designed for target reliability indices of 0, 0.5, 1.0, 1.5 and 2.0 for all nine fire scenarios. The 9 fire scenarios for the target reliability index of 1.5 are shown in Figure 5.2, and the values of fire load, opening factor, and thermal absorpitivity used for describing each of these fires are given in Table 5.5. Thus, for each target reliability index value, there were 54 design situations (9 fire scenarios x 6 beams), and a total of 270 design situations for the five target reliability index values. Reliability analysis was then performed and the computed reliability index values for the six beams are compared with the target reliability index values in Figure 5.3. Table 5.4 - Combinations of b and Fv used for nine fire scenarios b (WsogmrK) 423 640 1127 F 0.02 x x x 132 0.08 x x x (m 0. 12 x x x 144 1 400 1200 . a 1000 2. 2 800 3 E o 600 D. .5. ,_ 400 200 o 50 100 Time (mins) 150 -<>- Fire 1 —°— Fire 2 + Fire 3 + Fire 4 + Fire 5 —°— Fire 6 -+— Fire 7 —— Fire 8 -- Fire 9 200 Figure 5.2 - Nine fire scenarios for target reliability index value of 1.5 250 Table 5.5 - Values of b, Fv and qt used for nine fire scenarios corresponding to ,6; =l.5 Parameter Unit Fire Fire Fire Fire Fire Fire Fire Fire Fire 1 2 3 4 5 6 7 8 9 0.5 2 b WS [m K 423 423 423 640 640 640 1127 1127 1127 F ml/2 0.02 0.08 0.12 0.02 0.08 0.12 0.02 0.08 0.12 v 2 Qt MJ/m 206 206 206 206 206 206 206 206 206 145 3.0 2.5 — 1.04 I 0.5 1| 0.0 l l I l l O 0.5 1 1.5 2 2.5 3 fit Figure 5.3 - Comparison of computed and target reliability index values The computed reliability index values compare very well with the target values, ,6,, indicating that the derived capacity reduction and fire load factors work for all design situations considered. The reliability index values are conservative for target reliability index values less than about 1.25, especially for the higher opening factors. This . conservatism is believed to be primarily due to three reasons: (1) For target reliability indices of less than 1.25 (see Figure 5.1), the capacity reduction factor, 9c, found from reliability analysis is greater than 1.0, and the nominal strength can be increased. However, since Wis generally always less than or equal to 1.0 in LRFD specifications, we restrained the gcfor fire design to also not exceed 1.0. Designed beams will therefore have reserve strength. This can be compensated for by lowering the fire load factor. In Figure 5.1, we lowered the 146 fire load factor in a conservative manner so that some reserve strength is still present. Because of this inherent conservatism, the reliability index values are higher than the target reliability index. (2) For reliability analyses, the mean value of insulation thickness is taken to be 1/ 16 inch higher than the nominal value. For the lower target reliability index values, especially for higher opening factors, the nominal insulation thicknesses are very small and the addition of 1/16 inch to the nominal values represents a large percentage increase and yields higher reliability indices. (3) The norrrinal insulation thicknesses are very small for the lower target reliability index values, and thus the steel temperature is quite sensitive to variation in the random parameters during the reliability analysis. 5.2.4 Comparison of Fire Load Factors with those Based on the ECSC Method As mentioned in Section 2.4.1, a study was conducted in Europe to develop fire load factors (ECSC 2001). The fire load factor in the ECSC study was obtained using simplified assumptions instead of rigorous reliability calculations, and was specified for any ,8, value through {1 — [73% [0.577 + ln(— ln (13(0916, D1} = 1.05 J— yq 6 (5-18) {I — 7V q [0.577 + ln(— 1n( pm} 147 where Vq = COV of the fire load, (D = cumulative standard normal distribution function, and p = percentile used for obtaining the characteristic or nominal fire load. If the nominal value is taken as the 90th percentile, then p = 0.9. In Figure 5.4, the yq obtained in this study is compared with that obtained using Equation 5-18 for US. fire load statistics, taking the nominal value of fire load to be the 90th percentile. yq obtained from the ECSC method is greater than that derived in the this study for ,6, values smaller than about 1.5, and is almost the same for ,6, values greater than 1.5. As shown earlier, yq derived in this study yields conservative ,6 values for fit values less than about 1.5, and hence the yq obtained according to the ECSC approach will yield even more conservative results. For ,6, values greater than 1.5, the beams designed as proposed herein yield the intended safety level or higher (see Figure 5.2) because yq is used in combination with a q that is less than 1.0 (see Figure 5.1). 148 1.2 -. 0.8 ~ w >~ O 4 3! q (present study) - y q (ECSC method) 0 l l l l l 0 0.5 1 1.5 2 2.5 fit Figure 5.4 - Comparison of fire load factors with those obtained by the ECSC method 5.2.5 Insulation Thicknesses from Performance-Based and Prescriptive Design Approaches In Table 5.6, the thicknesses of insulation computed according to the engineering approach using the capacity reduction and fire load factors developed in this study for target reliability index values of 0.5, 1.0 and 1.5 are compared with those obtained from the prescriptive approach for two beams (B2 and B8 in Table 5.2). Insulation thicknesses from the prescriptive approach were obtained using Design No. N708 given in Underwriters Laboratories (UL 2004). The nine fire scenarios described in Table 5.4 were considered. The thicknesses obtained from the performance-based approach are normalized with respect to the corresponding prescriptive thickness for a 2-hr fire rating. 149 Table 5.6 - Comparison of insulation thickness Thickness Normalized thickness from performance-based approach from - - b=423 b=640 b=1127 Target * ”gags: 0.5 2 05 2 0.5 2 reliability g Pg“ m) (Ws /m K) (Ws /m K) (Ws /m K) "“1“ a Fire Rating 1/2 1/2 1/2 (hrs) Fv (m ) Fv (m ) Fv (m ) 2 3 0.02 0.08 0.12 0.02 0.08 0.12 0.02 0.08 0.12 Bl 29 42 1.46 0.79 0.64 1.25 0.74 0.64 0.91 0.61 0.55 I?! = 0'5 82 24 35 1.41 0.72 0.58 1.19 0.68 0.59 0.84 0.55 0.49 B] 29 42 1.85 0.90 0.71 1.61 0.88 0.74 1.23 0.74 0.66 18!: 1'0 BIZ 24 35 1.83 0.84 0.65 1.57 0.82 0.68 1.17 0.68 0.60 Bl 29 42 2.41 1.08 0.85 2.15 1.08 0.89 1.72 0.94 0.82 ,6! = 1-5 B2 24 35 2.45 1.02 0.78 2.15 1.02 0.82 1.68 0.88 0.76 In Table 5.6, the shaded thicknesses indicate situations where the performance-based approach yields thicknesses higher than the prescriptive approach. The performance- based approach generally yields lower insulation thicknesses, especially for lower target reliability index values, which in turn depend on the effectiveness of active fire protection systems in reducing the probability of occurrence of a severe fire. It should be noted that a direct comparison between the thicknesses of insulation computed from the performance-based approach with that obtained from the prescriptive approach isnot rational. In the performance-based approach, the required thickness of insulation depends on the real fire scenario in a compartment which in turn depends on the amount of fire load present in the compartment, ventilation conditions and on the type of the bounding materials. On the contrary, the thickness of insulation obtained from the prescriptive approach is based on fire tests conducted in a furnace under a standard fire which does not represent the real room conditions. The purpose of the comparison shown in Table 5.6 is to highlight that although the fire load factors are smaller than 1.0 for target 150 reliability index values less than 1.25, the performance-based approach still yields thicknesses that are comparable to those used in the prescriptive approach. 5.2.6 Reliability Inherent in AISC Fire Design Provisions It is of interest to determine what fl value is inherent in the AISC approach. The insulation thicknesses for the six beams and the 9 different fire scenarios described earlier (see Table 5.4) were determined using the AISC approach for two cases: (1) using 41%: 0.9 and by reducing the 90th percentile of the fire load by 60% (for sprinklers of the highest reliability) as suggested in the Commentary to the AISC Specifications; and (2) using ¢f = 0.9 and using the 90th percentile of the fire load assuming that there are no reliable sprinklers in the building. The reliability analyses were performed using these insulation thicknesses, and it was found that the reliability index varied from 0.2 to 0.5 for Case 1 and from 1.45 to 1.60 for Case 2. 5.3 Design Examples To illustrate the applicability of the proposed approach in a design situation, the method is applied to evaluate the fire resistance of two simply supported beams (one laterally restrained and the other laterally unrestrained). In this example, the value of the target reliability index, ,6,, is obtained using the procedure described in Section 3.5.3. The effectiveness of active fire protection systems in reducing the probability of occurrence of a severe fire were obtained from the ECSC study (ECSC 2001). In design specifications, the specifications authority may specify ,Bt 151 values depending on the effects of active fire protection systems, and the designer may not need to calculate it. 5.3.1 Design Example (Laterally Restrained Beam) Problem statement: A fire compartment in a lO-story open—plan office building has a floor area of 200 m2. A 9.14 m (30 ft) long steel beam is to carry a service dead load of 11.27 kN/m (0.77 kips/ft) and service live load of 11.68 kN/m (0.80 kips/ft). A room temperature design of the beam was performed using the AISC design specifications, and a W18x35 section was selected. The beam has a plastic section modulus = 1089.74 cm3 (66.5 in3). The yield strength of the steel = 34473.8 N/cm2 (50 ksi). The fire compartment is 20 m square and 3.65 m high, and is made of lightweight concrete with density = 2000 kg/m3, specific heat = 840 J/kg.K and thermal conductivity = 0.8 W/mK. The compartment has four windows each 4.0 m wide and 2.0 m high. The fire load is 196 MJ/m2 per unit surface area of the compartment. The compartment is equipped with automatic sprinklers, and smoke and heat detectors. The automatic sprinklers are assumed to be 70% effective in reducing the probability of occurrence of a severe fire. The effect of other active fire protection systems may be neglected. The beam is to be protected using a light weight insulating material with density = 300 kg/m3, specific heat = 1100 J/kg.K and thermal conductivity = 0.2 W/mK. As per the building code, the beam is required to have a 2-hours fire resistance rating. Calculate the required thickness of the insulation. 152 Solution: 1. Applied moment under fire Using Equation 5-1 Muf = (1.20 + 0.5L)(span2/8)= (1.2(1 1.27) + 0.5(1 1.68))(9. 142/8) 2 202.2 kN-m 2. Capacity of the beam Selection of target reliability index: Probability of fire occurrence in an office building 2 4 x 10'7(per m2, per year) (ECSC 2001); probability of fire occurrence for 200 m2 compartment assuming a 55-year building life, p. = 200 x 55 x 4 x 10.7 = 0.0044 (ECSC 2001); probability of failure of automatic sprinklers, p2 = 0.30; probability of occurrence of a severe fire, psf = p1.p2 - 0.00132 (Equation 3-21) ; target probability of failure for life of the building, pp 7.23x10-5 (ECSC 2001). Using Equation 3-24 and the normal density table, the target reliability index, _ P flz=¢l 1’ f =l.6 psf Selection of appropriate capacity reduction and fire load factors: Using Equations 5-16 and 5-17 with ,8, = 1.6: Fire load factor, yq = 0.15 + 0.6 ,8; = 1.11 Capacity reduction factor, (13f = 1.5 - 0.4 fl, = 0.86 153 Fire parameters: Nominal fire load = 196 MJ/mz; total surface area of the compartment, A, = 1092 m2; height of windows, Hv = 2.0 m; total area of windows, Av = 32 m2; opening factor, Fv = Fv = AM] H v / A, = 0.041 mm; density of compartment boundaries, p = 2000 kg/m3; specific heat of compartment boundaries, c = 840 J/kg.K; thermal conductivity of compartment boundaries, k = 0.8 W/mK; thermal absorptivity of compartment boundaries, 193/1906,, = 1160 Wso'slmzK. Nominal capacity of beam: As described in Section 3.2, the required thickness of insulation by the engineering approach can be computed using an iterative procedure. Iteration I : Assume thickness of insulation 2 25 mm. Fire load = yq x nominal value of fire load = 1.11 x 196 = 217.56 MJ/m2 Using this fire load and the other parameters given above, the fire temperature can be computed using Equations 3-4 to 3-7 (Section 3.3), and the maximum steel temperature can be computed using Equation 3-8 or 3-9 (Section 3.4). Maximum steel temperature, Tamax = 630.2°C occurs after 1.25 hours Yield strength reduction factor, ky(Ts) = 0.373 154 (The value of ky(Ts)was obtained from the AISC Specifications (AISC 2005a) corresponding to TS, max). Using Equation 5-3 with ¢f= 0.86, the capacity of the beam under fire is: be: ¢anf = qzxkymwy = (0.86)(1089740)(0.373)(345) = 120600545 N-mm = 120.6 kN—m Since be < Muf, the insulation thickness is inadequate. Perform next iteration by assuming a higher insulation thickness. After 4 iterations, the required thickness of insulation was obtained as 32.1 mm with a maximum steel temperature of TS, max = 547°C that occurs after 1.35 hours. This implies that the beam not only has required 2 hours fire resistance but it will not fail because steel temperature starts decreasing after 1.35 hours. The iterative solution procedure can be easily implemented in an Excel spreadsheet. Note: In this example, the required thickness of insulation was determined using an iterative procedure. Using the same approach, the failure temperature of the steel beam and the corresponding fire resistance time can also be computed for a given value of insulation thickness. 5.3.2 Design Example (Laterally Unrestrained Beam) Problem statement: 155 A fire compartment in a 4-story open-plan office building has a floor area of 100 m2. A 9.14 m (30 ft) long steel beam is to carry a service dead load of 11.27 kN/m (0.77 kips/ft) and service live load of 11.68 kN/m (0.80 kips/ft). A room temperature design of the beam was performed using the AISC design specifications, and a W10x54 section was selected. The properties of the beam are: 2 2 0 Cross sectional area, AS = 102 cm (15.8 in ) . . 3 . 3 0 Elastrc sectron modulus, S x = 983.2 cm (60m ) . . 3 . 3 0 Plastic sectron modulus, Zx = 1091.4 cm (66.6 in ) . . . . 4 . 4 - Moment of inertra about minor axrs, Iy = 4287.2 cm (103 in ) 0 Radius of gyration about minor axis, ry = 6.5 cm (2.56 in) . 6 . 6 o Warping constant, Cw = 623003 cm (23201n ) . 4 . 4 0 Torsronal constant, J = 75.75 cm (1.82 in ) 2 . 0 Yield strength of the steel, F y = 34473.8 N/ cm (50 ksi) 0 Modulus of elasticity, E s = 19994804 N/ cm2 (29000 ksi) The fire compartment is 10 m x 10 m x 3.65 m, and is constructed from concrete with density = 2300 kg/m3, specific heat = 980 J/kg.K and thermal conductivity = 1.6 W/mK. The concrete walls and ceiling are covered with 12 mm thick gypsum plaster board with 3 density = 700 kg/m , specific heat = 1700 J/kg.K, and thermal conductivity = 0.2 W/mK. 156 The compartment has one window which is 3.0 m wide and 1.5 m high. The fire load is 2 196 MJ/m per unit surface area of the compartment. The building is not eqipped with any active fire protection systems. However, public fire brigade services are available. The beam is provided with 44.5 mm (1.75 in.) thick lightweight insulating material with 3 density = 300 kg/m , specific heat = 1100 J/kg.K and thermal conductivity = 0.2 W/mK. As per the building code, the beam is required to have a 2-hours fire resistance rating. Check if the provided insulation is adequate. Solution: 1. Applied moment under fire Using Equation 5-1 Muf = (1.2D + 0.5L)(span2/8)= (1.2(11.27) + 0.5(11.68))(9.142/8) = 202.2 kN-m 2. Capacity of the beam Selection of target reliability index: Probability of fire occurrence in an office building = 4 x 10-7(per m2, per year) (ECSC 2 2001); probability of fire occurrence for 100 m compartment assuming a 55-year building life, p1 = 100 x 55 x 4 x 10.7 = 0.0022 (ECSC 2001); probability of occurrence of a severe fire, psf = p1 = 0.0022 (Equation 3-21) ; target probability of failure for life of the building, p, = 7.231110”5 (ECSC 2001). 157 Using Equation 3-24 and the normal density table, the target reliability index, .3: =¢_1[1--le]= 1.84 Sf Selection of appropriate capacity reduction and fire load factors: Using Equations 5-16 and 5-17 with fl, = 1.84: Fire load factor, )4, = 0.15 + 0.6 fl, = 1.254 Capacity reduction factor, (13c = 1.5 - 0.4 = 0.76 Fire parameters: 2 2 Nominal fire load = 196 MJ/m ; total surface area of the compartment, A, = 346 m ; . . . 2 . height of wrndows, HV 2 1.5 m; total area of wrndows, Av = 3.75 m ; openrng factor, Fv = F, = A,,/H,, IA, =0.015 mm. 1 .5 2 Thermal absorptivity of concrete boundaries, b = kfl‘p = 1900 Ws0 /m K Thermal absorptivity of gypsum plaster board, b = 1 lkpcp = 488 WsO'SImZK . . 0.5 2 Resultant thermal absorptrvrty of the compartment, b = 626 Ws /m K (The procedure for calculating b is given by Buchanan (2001)). Nominal capacity of beam: 2 Fire load = yq x nominal value of fire load = 1.254 x 196 = 245.8 MJ/m 158 Using this fire load and the other parameters given above, the fire temperature was computed using Euqtaions 3-4 to 3-7 (Section 3.3), and the steel temperature was computed using Equation 3-8 (Section 3.4) using a spreadsheet. The resulting time—fire temperature and time-steel temperature curves are shown in Figure 5.5. Steel temperature after 2 hours = 554°C Maximum steel temperature, Ts’max = 718°C occurs after 3.54 hours Yield strength reduction factor, ky(TS) = 0.613 Elastic modulus reduction factor, kE(TS) = 0.44 Proportional limit reduction factor, kp(Ts) = 0.26 (The values of ky(TS), kE(TS)and kp(TS) were obtained from the AISC Specifications (AISC 20053) and Eurocode 3 (EN 2005) corresponding to Ts = 554°C). Using Equation 5-5 with (I? = 0.76 and the capacity adjustment factor, ¢a = 1.67 (See Section 5.1.4.2), the capacity of the beam under fire is be: @147me f (with Mn,f given by Equation 5-5) = 214.2 kN-m Since be > Muf , the insulation thickness is adequate for a 2- hours fire resistnce rating. 159 1200 1 Tr g K O 800 9 1 Ts 15 3 X a. E 0 '- 400 - Ts =5 54°C 0 I it I l l l l 0 60 1 20 180 240 300 360 420 Time (mins.) Figure 5.5 - Fire and steel temperatures vs. time used in computing fire resistance for beam in example 2 5.4 Summary To illustrate the proposed methodology presented in Chapter 3, capacity reduction and fire load factors are developed for simply supported steel beams in US. office buildings exposed to fire. Fire load, ratio of floor area to the total area of fire compartment, opening factor, thermal absorptivity of compartment boundaries, thickness, density and thermal conductivity of fire protection material, dead load, and live load are taken as random variables. Mechanical and sectional properties of steel (e.g., yield strength, section modulus, etc.) are also considered to be random variables. The effect of active fire protection systems (e. g., sprinklers, smoke and heat detectors, fire brigade, etc.) in reducing the probability of occurrence of a severe fire is included. From detailed reliability analyses, it is found that the capacity reduction and fire load factors should not be constant for all design situations as suggested in 160 design specifications, and should vary depending on the presence of active fire protection systems in a building. The variation of the fire load factor based on the presence of active fire protection systems is in agreement with the Commentary to the AISC Specifications, the Eurocode 1, and the ECSC study. For most office building compartments in the US. equipped with sprinklers, use of (1%: 1.0 is reasonable, and yq is likely to lie between 0.4 and 1.0. 161 Chapter 6 Capacity Reduction and Fire Load Factors for Steel Columns Exposed to Fire In this paper, capacity reduction and fire load factors are developed for the design of axially loaded steel columns exposed to fire. 6. 1 Derivation of Capacity Reduction and Fire Load Factors 6.1.1 Performance Functions for Reliability Analysis 6.1.1.1 Applied Axial Loads under Fire According to the AISC Specifications, the ultimate applied axial load, Puf, is determined from the gravity load combination given by 162 Puf: 1.2PD + 0.5PL + 0.2135 + T (6-1) where, PD, PL and P5 are nominal dead, live and snow loads, respectively, and T includes load effects induced by the fire itself. In Section 3.5.1.1 it was concluded that it is appropriate to use the combination of dead and arbitrary-point-in-time live load for reliability analysis under fire conditions. Thus, for reliability analysis, the applied axial load, Paf, under fire conditions may be expressed in terms of basic variables as where A and B = random variables reflecting the uncertainties in the transformation of loads into load effects, E = a random variable representing the uncertainties in structural analysis, and D and Lap, = random variables representing dead and arbitrary-point-in- time live load. The statistics of A, B, E, D and Lap, are given in Table 4.1. 6.1.1.2 Axial Capacity of Columns under Fire Takagi and Deierlein (2007) compared the AISC and Eurocode 3 design specifications with finite element simulations for columns exposed to fire. They reported that the AISC Specifications (AISC 20053) are unconservative at elevated temperatures, particularly for Slendemess ratios between 40 and 100 and temperatures above 500°C. For instance, at 500°C the nominal capacities predicted by the AISC Specifications are up to 60% larger than capacities predicted by simulations. On the other hand, the Eurocode 3 column strength equations were within 20% of the simulated results. The equations proposed by 163 Takagi and Deierlein (2007) were used in this study, which have a format similar to that of the AISC Specifications and predict strengths similar to the Eurocode 3 (EN 2005) columns strength equations (see Figure 6 in Takagi and Deierlein (2007)): ' r 1ky(Ts)Fy PM 210.42 53”“ iAsky(Ts)Fy L ' J (6-3) rrzkE(TS)ES (KL/ r)2 Fears) = (6—4) Pnf = nominal capacity of column under fire, A S = cross-sectional area, KL = effective length, r = radius of gyration about the buckling axis, ES 2 elastic modulus. kE(TS) is the elastic modulus reduction factor that depends on the temperature, TS. The actual capacity of steel columns under fire can be obtained by modifying the nominal capacity given in Equations 6-3 and 6-7 to r Jky(’scTs)mle l Pf = PC1042 ' Fem) ffSAskyUscTslmlF y (6-5) 2 F (T )= 71' kE(lSCTS)m2ES f3r where fi and m,- are non-dimensional random variables that characterize the uncertainty in the fabrication and material parameters, and their statistics are given in Table 5.1. 164 15c is the model error for steel temperature which was characterized in Section 4.2. PC is the professional factor (model error) that is characterized in the next subsection. The performance function for reliability analysis of column can then be expressed as 801’) = Pf- Paf (6-7) where X denotes a vector containing all the random design parameters. The probability of failure, pp , of a steel column under fire is PF = P[g(X) < 0] (6-8) 6.1.2 Professional Factor (Model Error) for Axial Capacity of Columns To account for differences between axial capacities of columns measured in the laboratory and that predicted by Equation 6—3, the professional factor, PC, was characterized using the test results presented by Janss and Minne (1981) and Franssen et al. (1998). Janss and Minne (1981) reported results for eighteen columns with Slendemess ratio between 25 and 102 for which the yield strength was measured. Franssen et al. (1998) reported test results for twenty one fire tests with Slendemess ratio between 20 and 140. The nominal capacity of these tested columns was calculated through Equation 6-3. Pc (ratio of measured axial capacity to nominal capacity) is plotted against Slendemess ratios in Figure 6.1. PC has a mean of 1.10 and a COV of 0.18, and is best described by the normal distribution. 165 1.5 4 o ’ . O. :0: o: . . O O . . 0 o e 2’ 3 :2 1 a O O .0 ‘ 76 o C '9 e U) 0 is 0.5 O L Q. 0 i l l 0 50 100 150 Slendemess ratio Figure 6.1 - Ratio of test capacity to nominal capacity of columns for different slenderness ratios 6.1.3 Probability of Failure and Target Reliability Index In Section 3.5.3 it was shown that for typical U.S. office buildings the target reliability index is likely to vary from zero to 2.0, depending on the size of the fire compartment and on the effectiveness/reliability of active fire protection systems. Therefore, in the next section, capacity reduction and fire load factors were developed for target reliability index values ranging from zero to 2.0. 6.1.4 Reliability Analysis Twenty one steel columns with slenderness ratios ranging from 25 to 200 and axial load capacities ranging from 133 kN (30 kips) to 10,675 kN (2400 kips) were selected for the reliability study. Columns with smaller capacity are representative of those in upper stories and those with higher capacity are representative of those in lower stories of typical office buildings. Live to dead load ratios ranging from 0.5 to 5.0 were considered. 166 The AISC design specifications were used to first design the columns for ambient temperature conditions. The same columns were then designed for fire exposure (b = 640 0. 2 Ws 5/m K and F v = 0.02 mm) using the engineering approach described in Section 3.2 and the required thickness of insulation to withstand the design fire was determined from Equations 6-1 and 6-3 using an iterative procedure. As suggested in most codes, a capacity reduction factor of 1.0 was used for the initial design for fire. The columns were assumed to be protected by gypsum board insulation, which is generally the case in the US. Load ratios (ratio of applied load under fire to room temperature nominal capacity) ranging from 0.35 to 0.66 were considered. The applied loads, live to dead load ratios, slenderness ratios, section factors, load ratios, and the insulation thicknesses for the columns used in reliability analyses are given in Table 6.1. 167 3.3 8.3 8.9; as; as; 8.3 AS: So: So: :2: 5 N: NE N3 N: 9: mm mm mm 2: 3.8m .5sz 3N. : Gm: 8m. : am. : am. : an : 8on :68 85.8 :5 $95.2... 8.2 3.: 8.2 $3. on? 2:: F: 8.2 8.8 as 5.3.3:. woo woo :3 23 as 93 Go mm. 8; 3.2 35 mm we 8 02 we own as. o: an” 2:: 80585.8 8.2 of. 8s 8s m: of m? 8. of S can :89 some :5 :9 some 89: £5 :33 Wen: ES 2: an: m: _ a8 3 84 was 82 23 $8 7: can 69 at $3 as am: 8%: 8:: so. : av 33 as: R: mom o: o: .8 as 82 2% 2mm 7: a: a: a: as a: a: a: a: E 3: 5% SA SA ohm 03 com 85” 02 8a 2a a 8:23 EBB 523 523 523 EBB 8:23 8:23 8:23 5282838 8 no 8 5 mo ..0 8 8 5 £5 232:2: monk—£5 b=3a=oa you com: 3538 .83 12:8— »:Exa «o 8.22—2.— l fie 2A.; 168 .E .E .2: .2: .3: 2...: ...w: ...w: .35 6...: 2-..: 8.8m .2 NR 8. 8. 8. mm. mm. a. R. R. ..E 5.6% .E: .3: .8: .3: .2: ....: .8: an: .3... .2: ...5 .856... 2.3 2.3 3.2 9.: Sam 83 $2 8.2 3...“ 2.3 E... 5.3.3... a... 8... cm... :5 9.... .4... 2... mm... 3.... a... 2:: 18.. O5“.— 2. 8“ NS NS NS 0.? 3m 0.? 3..“ 4.94. 3.52.53 on. on... 8m 8m cod 8..“ 8A 8.. 3.. 8a 9.. ...N. E .3: am... .8: .33 3.: .940 .8: .8: .81.... mm. mm m: N .8“ o. w. E. N .3... MR. 88 Sam 2.. .33 m3... 8: .2. .3. ...m: .8: .B: .8: .3: .8: ..m: 9...... .33 ea... 3. 8 m»... $0 ..3 mm. 4N... Rm. ea c. .. 7... .o: .m: .o: .o: .o: .o: .N: .o: .N: .o: 8. 5% 85m 8... 8s 8..” 8a 8a oi oi 2a 8a E EGmuuvm 2.5.3 2.5.3 3.23 8.83 8.53 93:3 93:3 8.3.3 8.3.3 8.3.3 5:28 ..No a... ...u ...0 20 20 EU 20 N5 .5 3...: 23982.... 2.2.52.8: mowing 3:339. ..8 can: 2.528 $on conno— Eag he 85.2.95 1 do visa. 169 Using FERUM, FORM analysis was performed for each design situation (each of the 20 columns) using the performance function given in Equation 6-7. Using the direction cosines obtained from the reliability analysis, the partial safety factors for each design parameter were obtained through Equation 3-25 for each design situation. These individual partial safety factors, except for the fireload, were then combined into a single capacity reduction factor through . ‘/ky(¢lTs)02Fy l 40.42 PM) >¢3Asky(¢lTs)¢2Fy ¢f : L J (6—9) me 2 _ 7T kE(¢lTs )¢4Es where Fe (Ts) — KL 7- (6'10) ¢5rl Q are the partial safety factors for each design parameter obtained through Equation 3- 24, and Pnfis the nominal capacity of the column calculated through Equation 6-3. Thus 20 different capacity reduction factors (one for each column) were obtained. Thereafter, a single optimized capacity reduction factor, gr, corresponding to dead and live load factors of 1.2 and 0.5, respectively, was obtained using the optimization procedure described in Section 3.5.7. Equal weights were used for all columns. Since the capacity reduction factors were obtained for ,6, ranging from 0 to 2, this procedure was repeated for each ,8,. The fire load factors were constrained to be the same as those obtained for steel beams since it is not desirable to have different fire load factors for beams and columns. 170 6.2 Results 6.2.1 Capacity Reduction Factor The plot of the optimized capacity reduction factor, ¢f vs. ,6, is shown in Figure 6.2. The value of (bf is given by 1.0 for ,6, S 1.25 f : (6-11) 1.25—0.25, for 1.25 3 fl, s. 2.0 Again, as was the case for steel beams, the results obtained for the columns indicate that the nominal capacity need not be reduced (i.e., (1}c = 1.0) if ,6, is less than 1.25, which in turn depends on the effectiveness of active fire protection systems in reducing the probability of occurrence of a severe fire. Since most office buildings in the US. are required to have automatic sprinklers, ,6, is not likely to exceed 1.25. Therefore, using q = 1.0 is reasonable for most design situations. 1.6 — 1.21 y 3” 0.84 '3 3 ‘s‘ 04* Ki 0 l I l l l o 05 1 1.5 2 25 Figure 6.2 - Capacity reduction and fire load factors vs. target reliability index for columns 171 6.2.2 Fire Load Factor As mentioned earlier, the fire load factor, yq was constrained to be the same as that derived for steel beams and is plotted against the target reliability index in Figure 6.2. The value of yq for a given target reliability index, ,6,, is 0.4+0.4,6, for ,6, S 1.25 7., = (6-12) 0.15 +0.6,6, for 1.25 S ,8, S 2.0 The yq is to be applied to the fire load which is then used in describing the design fire ' (time-fire temperature curve) using Equations 3-4 to 3-7. 6.2.3 Validity of Capacity Reduction and Fire Load Factors for Different Slendemess Ratios Two steel columns (a light W10x19 section and a heavy W14x90 section, having radii of gyration of 22 mm and 94 mm, respectively) were chosen to study the effect of slenderness ratio on the reliability index. Slendemess ratios ranging from 25 to 200 were used for reliability analysis. As shown in the next section, a live to dead load ratio of 2.0 yields lower values of the reliability index, and therefore, in order to be conservative, this value was used. Both columns were designed using the fire load and capacity reduction factors given in Equations 6-11 and 6-12. The reliability index values computed from the reliability analysis, ,8, are compared with ,8, values in Figure 6.3. The ,8 values compare quite well with the ,8, values, indicating that the derived capacity reduction and fire load factors work well for all slenderness ratios. The comparison 172 shown in Fig. 3 is for two columns, and the ,8 values coincide for both columns for each slenderness ratio and for each ,6,. 3'5 — 0 flt=o.o ' flt=0.5 A flt=1.o " 3.0 - ..... a: 25 _ x flt=1.5 . 1: ° ’ E 2.0 . ’ ° 32‘ 1.5 x x x x X :3 1.0 A a A A A 7) . I I u I a: 0.5 0.0 9 ‘3 ‘3 ‘3 ‘3 0 50 100 150 200 Selenderness ratio Figure 6.3 Computed and target reliability index values for different slenderness ratios 6.2.4 Validity of Capacity Reduction and Fire Load Factors for Different Live to Dead Load Ratios Two steel columns, a W16x57 section and a Wl4x90 section, having radii of gyration of 40 mm and 94 mm, respectively, were chosen to study the effect of live to dead load ratio on the reliability index. Live to dead load ratios ranging from 1.0 to 5.0 were considered, and both columns were designed using the fire load and capacity reduction factors shown in Figure 6.2. Slendemess ratios of 25 and 75 were used for the Wl4x90 and W16x57 columns, respectively. These slenderness ratios yield relatively low and high reliability index values (see Figure 6.3), and span the most likely range of slenderness ratios for steel columns. The reliability index values computed from reliability analysis, ,6, are 173 compared with ,8, values in Figure 6.4. The ,8 values compare quite well with the ,6, values, indicating that the derived capacity reduction and fire load factors may be used. for all live to dead load ratios considered herein. The ,6 values are conservative (i.e., higher than the fit) for lower live to dead load ratios. As described earlier, the applied load effects under fire are computed from the load combination 1.2Dn + 0.5L", and the dead load factor is the same for fire design as for room temperature, while the live load factor for fire design is smaller than that for room temperature gravity design (0.5 instead of 1.6). Thus under fire situations, when the design is dominated by the dead load (which is the case for smaller live to dead load ratios), higher insulation thicknesses are required. When the insulation thickness is large, the column is thermally more stable, the variation of other random parameters do not have a significant effect on the reliability of the column, and the column design generally yields a higher ,8 value. 4.0 _. o [31:04) - flt=o.5 A flt=1.0 3.5 - x 3 O _ X fltzls . flt=200 0 . 'g 2.5 ’ . ' E. 2.0 3‘ . 3 it ‘ g 1.5 a 35. it ’3“ ’i g 1.0 a 3 *3 A A a? 05 5 ' ' ' ' 0.0 9 Q E? 9 9 o 1 2 3 4 5 W Ratio Figure 6.4 - Computed and target reliability index values for different live to dead load ratios 174 6.2.5 Validity of Capacity Reduction and Fire Load Factors for Multiple Fire Scenarios The capacity reduction and fire load factors shown in Figure 6.2 were developed for a fire compartment assumed to be constructed of lightweight concrete blocks having a thermal absorptivity b = 640 WsO'SImZK, and having an opening factor F v = 0.02 mm. In reality, the compartments may be constructed using different bounding materials and may have different ventilation conditions. Therefore, the capacity reduction and fire load factors were validated by considering 9 fire scenarios described in Section 5.2.3 and Table 5.4. For these nine fire scenarios, five columnswere designed for fire conditions using the capacity reduction and fire load factors shown in Figure 6.2. The steel sections used for these columns, and their room and elevated temperature capacities are given in Table 6.2. Table 6.2 - Properties of columns used for validation Parameter W10xl9 W10x30 W18x65 Wl4x90 Wl2xl90 H (m) 3.66 3.66 3.66 3.66 3.66 231 876 2497 5276 10751 I"n (KN) 104 394 1124 2 7 48 7 Pnf (KN) 3 5 3 Load ratio 0.45 0.45 0.45 0.45 0.45 For all columns, a live to dead load ratio of 2.0 was assumed. Each column was designed for ,8, values of 0, 0.5, 1.0, 1.5 and 2.0 for all nine fire scenarios. Thus, for each ,6, value there were 45 design situations yielding a total of 225 design situations. Reliability analysis was then performed and the computed reliability index values, ,6, for all five columns are compared with the ,6, values in Figure 6.5. 175 2.5 . 2.0 ~ K 1.5 - 13% below [3, R '\ 1-0 ' 24% below ,3, 0.5 0.0 . . . . . 4 o o 5 1 1.5 2 2 5 fit Figure 6.5 - Computed and target reliability index values for columns for different fire scenarios The ,6 values compare quite well with the ,6, values, indicating that the derived capacity reduction and fire load factors work well for all design situations considered. The ,6 values are conservative for ,8, values less than about 1.5. For ,8, values of less than 1.5 (see Figure 6.2), the ¢f found from reliability analysis was greater than 1.0, and the nominal capacity could be increased. However, since Qvis generally always taken to be less than or equal to 1.0 in LRFD specifications, we restrained the (If for fire design to also not exceed 1.0. Because of this inherent conservatism, the ,6 values are higher than the ,8, values. In Figure 6.5, 24% and 13% of the [3 values fall just below the target value for ,6, values of 1.5 and 2.0, respectively. To ensure that the ,8 values are always higher than the target 176 value, the derived capacity reduction factor was reduced from 0.95 to 0.90 and from 0.85 to 0.80 for ,6, values of 1.5 and 2.0, respectively. Five columns were redesigned using these reduced capacity reduction factors and yielded the ,6 values shown in Figure 6.6, which are always higher than the target values. If greater conservatism in design is desired, the reduced values of capacity reduction factors given by ¢ _ 1.0 for ,B,Sl.0 f _ 1.2—02,6, for 103,490 (6'13) can be used instead of those given by Equation 6-11 (Figure 6.2). 3.0 2.5 - 2.0 - 1.5 - 1.0 - 0.5 0.0 l l f .l l l o 0.5 1 1.5 2 2.5 3 fit Figure 6.6 - Computed and target reliability index values for columns with reduced capacity reduction factors 6.3 Design Example To illustrate the applicability of the proposed approach in a design situation, the procedure is applied to evaluate the fire resistance of an axially loaded steel column. 177 In this example, the value of target reliability index, ,6,, is obatined using the procedure described in Section 3.5.3. The effectiveness values of active fire protection systems in reducing the probability of occurrence of a severe fire are obtained from the ECSC study (ECSC 2001). In design specifications, the specifications authority may specify ,6, values depending on the effects of active fire protection systems, and designer may not need to calculate it. Problem statement: 2 A fire compartment in an 8-story open-plan office building has a floor area of 100 m . A 3.65 m (12 ft.) long interior steel column is to carry a service dead load of 178 kN (40 kips) and service live load of 360.3 kN (81 kips). A room temperature design of the column was performed using the AISC design specifications, and a W10x30 section was selected. The column has a cross-sectional area = 5703 mm , and radius of gyration = 2 34.8 mm. The yield strength of the steel = 345 N/mm and elastic modulus = 200,000 2 N/mm . The fire compartment is 10 m square and 3.65 m high, and is made of lightweight 3 . . concrete with density = 2000 kg/m , specific heat = 840 J/kg.K and thermal conductrvrty = 0.8 W/mK. The compartment has two windows each 3.0 m wide and 1.5 m high. The 2 . fire load is 196 MJ/m per unit surface area of the compartment. The compartment is equipped with automatic sprinklers, and smoke and heat detectors. The automatic 178 sprinklers are assumed to be 80% effective in reducing the probability of occurrence of a severe fire. The effect of other active fire protection systems may be neglected. The 3 column is to be protected using gypsum board with density = 745 kg/m , specific heat = 1200 J/kg.K and thermal conductivity = 0.16 W/mK. . , Calculate the required thickness of the gypsum board insulation. Solution: 1. Ultimate applied axial load under fire Using Equation 6-1: Puf = 1.2PD + 0.5PL= 1.2(178) + 0.5(360.3) = 393.75 kN 2. Capacity of the column Selection of target reliability index: The target reliability index will be determined using the procedure given in Section 3.5.3. — 2 Probability of fire occurrence in an office building = 4 x 10 7(per m , per year) (ECSC 2 . 2001); probability of fire occurrence for 100 m compartment assuming a 55-year building life, p1 = 100 x 55 x 4 x 10‘7 = 0.0022 (ECSC 2001); probability of failure of automatic sprinklers, p2 = 0.20; probability of occurrence of a severe fire, psf= p1.p2 - 0.00044 (Equation 3-21) ; target probability of failure for life of the building, pt = 7.23x10-5 (ECSC 2001). 179 Using Equation 3-24 and the normal density table, the target reliability index, a =¢‘l[l- pf]=1.0 psf Selection of appropriate capacity reduction and fire load factors: Using Equations 6-12 and 6-13 with ,6, = 1.0 Fire load factor, yq = 0.4 + 0.4 ,8, = 0.8 Capacity reduction factor, (13c: 10 Fire parameters: 2 Nominal fire load = 196 MJ/m ; total surface area of the compartment, At = 346 m2; . . 2 . height of wrndows, Hv = 1.5 m; total area of wrndows, Av = 9 m ; openrng factor, Fv = 1/2 Fv = AM/Hv / A, = 0.032 m ; density of compartment boundaries, p = 2000 kg/m3; specific heat of compartment boundaries, c = 840 J/kg.K; thermal conductivity of compartment boundaries, k = 0.8 W/mK; thermal absorptivity of compartment boundaries, b 2' kpcp =1160Wso'5/m2K Nominal capacity of column: As described in Section 3.2, the required thickness of insulation by the engineering approach can be computed using an iterative procedure. Iteration I : Assume thickness of insulation (gypsum board) = 15 mm. 180 Fire load = yq x nominal value of fire load = 0.80 x 196 = 156.8 MJ/m2 Using this fire load and the other parameters given above, the fire temperature can be computed using Euqtaions 3-4 to 3-7 (Section 3.3), and the maximum steel temperature can be computed using Equation 3-8 or 3-9 (Section 3.4). Maximum steel temperature, Twmu = 606.5°C Yield strength reduction factor, ky(TS) = 0.454 Elastic modulus reduction factor, kE(TS) = 0.298 (Values of ky(TS) and kE(Ts) were obtained from AISC Specifications (AISC 20053) corresponding to Tam“). Using Equations 6-3 and 6-4 with (13c: 1.0, the capacity of the beam under fire is Paf: (bmef (with Pnf given by Equations 6-3 and 6-4) = 201.6 kN Since Pc,f < Puf , the insulation thickness is inadequate. Perform next iteration by assuming a higher insulation thickness. After 4 iterations, the required thickness of insulation was obtained as 21.5 mm. The iterative solution procedure can be easily implemented in an Excel spreadsheet. Note: In this example, the required thickness of insulation was determined using an iterative procedure. Using the same approach, the failure temperature of the steel column 181 and the corresponding fire resistance time can also be computed for a given value of insulation thickness. 6.4 Summary Capacity reduction and fire load factors are developed for steel columns in US. office buildings exposed to fire. The effect of active fire protection systems (e.g., sprinklers, smoke and heat detectors, fire brigade, etc.) in reducing the probability of occurrence of a severe fire is included by adjusting the target reliability index appropriately. From detailed reliability analyses, it is found that the capacity reduction and fire load factors should not be constant for all design situations as suggested in design specifications, but should vary depending on the presence of active fire protection systems in a building and the compartment size. As with the AISC and Eurocode provisions, the fire load factor should be reduced for typical fire compartment sizes when active fire protection systems are present. 182 Chapter 7 Deflections of Simply Supported Steel Beams Exposed to Fire 7. 1 Background The AISC Specifications (AISC 2005a) stipulate that for structural elements, e.g., beams and columns, the governing limit state is loss of load-bearing capacity. However, they also specify that excessive deformations are not acceptable if these damage the integrity of the fire compartment. If not limited, deformations may damage the slabs and walls of the compartment to the extent that these components cannot control the horizontal and vertical spread of fire, even though the beam may still have sufficient load bearing capacity. Since design under fire may be governed either by the strength or deflection limit states, both should be considered. Therefore, the required thickness of insulation for beams based on the strength limit state can be first determined using the capacity reduction and 183 fire load factors developed in Chapter 5, and then the deflections of the beams could be checked. Design specifications such as the AISC Specifications (2005) and Eurocode 3 (EN, 2005) provide simple design equations for the strength limit state of steel beams which can be used for estimating the fire resistance. Researchers such as Yin and Wang (2004), Skowronski (1990), Burgess, El-Rimawi and Plank (1990), and Wang and Yin (2006) provide detailed methods for tracing the time-deflection behavior of steel beams, but these approaches are quite complex. No simple deflection equations which can be quickly implemented in a design specification are available for computing the deflections of steel beams. Therfore, a simplified method which can be quickly implemented in a design specification for computing the deflection of simply supported steel beams is developed in this chapter. Subsequently, the simplified method is used to compute the deflections of steel beams designed based on the strength limit state. 7.2 De velopent of Simplified Method for Estimating Deflections of Simply Supported Beams 7.2.1 Approach Design specifications such as the AISC Specifications (2005) allow steel beams exposed to fire to be designed using room temperature design specifications with reduced material properties. According to this concept, under fire conditions, the elastic deflections, Ae, of steel beams can be calculated using room temperature deflection equations and reduced material properties. For example, the elastic deflections of a uniformly loaded, simply supported beam can be calculated as: 184 __ 5wL4 ‘9 384ES(TS)I (7-1) where L = length of the beam, w = load per unit length, I = moment of inertia of the cross-section, ES(TS) = reduced elastic modulus calculated as ES(TS) = kEES , Es = elastic modulus used at room temperature, k1; = stiffness reduction factor that accounts for reduction in stiffness of steel at elevated temperature. At elevated temperature, the stiffness of the steel beams degrade rapidly, the deflections are likely to be beyond the elastic range, and Equation 7-1 cannot be used in most cases. However, in the inelastic range, Equation 7-1 may be used for calculating the plastic deflection, AP’ of steel beams by using an equivalent flexural rigidity, [E 3(TS)I]eq, instead of [ES(TS)I]. This concept is similar to the one used to calculate the deflections of reinforced concrete beams by using an equivalent moment of inertia that accounts for cracking. Therefore, it is proposed that in the inelastic range the deflection of a uniformly loaded simply supported beam be calculated as: _ 5wL4 P 384w, (2)11“, (7‘2) where [ES(TS)I]eq is the equivalent flexural rigidity at the steel temperature, T5. In the next section, a simplified equation is developed for calculating the equivalent flexural rigidity. 185 7.2.1.1 Equivalent Flexural Rigidity The proposed method for estimating the fire-induced deflection of simply supported beams is based on specifying an equivalent flexural rigidity, [ES(TS)I]eq of the steel section as a function of steel temperature. [E 5(Ts)l]eq is interpolated between the elastic flexural rigidity, [ES(TS)I]1 and the flexural rigidity at ultimate failure, [ES(TS)I]2. A typical moment curvature relationship of a steel section at any steel temperature, Ts, is shown in Figure 7.1. The curvature in the section increases linearly with increase in the applied bending moment, M, until yielding occurs at M = My} , where My} is the yield moment capacity at elevated temperature. After yielding, the increase in bending moment leads to a rapid increase in curvature, thereby resulting in larger deflections. The beam, however, continues to carry the applied bending moment until ultimate failure occurs when the maximum bending moment in the beam reaches the ultimate bending capacity A M A[E5(TS)I]1=kEEsI i" Curvature, K K" Figure 7.1 - Typical moment-curvature diagram for a steel section 186 of the section M = Maj. Figure 7.1 shows the elastic and ultimate flexural rigidities of the section, [ES(TS)I]1 and [ES(TS)I]2, respectively. The ultimate secant flexural rigidity is defined as: . Mu T [E5 (T5 )I]2 : K" (7‘3) where Ku = ultimate curvature. The ultimate curvature in Equation 7-3 is assumed to occur when the maximum strain in the section is equal to 0.2. Further, in order to account for the effect of strain hardening, the ultimate moment capacity is calculated using the ultimate strength instead of the yield strength. The ratio of ultimate strength to yield strength, F u/Fy, of steel is based on material prOperty tests (Wainman and Kirby 1987, and Luecke et al. 2005), and is plotted in Figure 7.2. Based on the experimental data, the F u/F y ratio is generally significantly higher than 1.25. However, to be conservative, the ultimate strength of steel was assumed to be 1.25 times its yield strength as suggested by the Eurocode 3 (EN 2005), i.e., F = 1.25Fy. Based on these assumptions, Ku = 0.2/(d/2), Myj = SxFy(Ts) = yield moment capacity, and Muf = F “(TS)Zx = 1.25 Fy(TS)Zx = ultimate moment capacity. Zx and Sx are the plastic and elastic section modulus, respectively, d is the depth of the section, and Fy(Ts) = kyFy 2 yield strength at steel temperature, T. Fy = nominal yield strength at 187 room temperature, and kv = yield strength reduction factor which accounts for reduction in steel strength at elevated temperatures. 2 1 OUKtest_Flange OUK test_Web ANIST_WTC 1.7 1 5 ‘ o . o o O O O u? o . 2 ‘5 1.5 ~ ‘ ll. . A ‘ A A O ‘ 0 Q ‘ : o 8 A 1.25 0 A ‘ A A ‘ A 1 l I T l 0 5 1O 15 20 No of Tests Figure 7.2 - Variation of ultimate tensile strength of steel Prior to yielding, the elastic deflection can be obtained using the elastic flexural rigidity of the beam [ES(TS)I]1 in Equation 7-1. However, when the beam enters the inelastic range, the equivalent flexural rigidity [ES(TS)I]eq is used to compute the inelastic deflection of the beam as per Equation 7-2. In order to obtain the equivalent flexural rigidity of a deflected beam in the inelastic range (M > Myf), the inelastic deflection of the beam is assumed to be equal to the elastic deflection, but with equivalent properties used instead of the elastic properties. For example, for a uniformly loaded, simply supported beam, using Equation 7-2, [E S(TS)I]eq can be written as 188 5wL4 [ES(TS)I]eq 2—3—8-4—A— P (7-4) Using Equation 7-4, the temperature-dependent relationship between [E 3(TS)I]eq and the applied moment M can be obtained for different values of applied load, w, provided the inelastic deflection, AP, of the beam is known. For developing this relationship, the inelastic deflection, AP, of beams were obtained using the ANSYS software (ANSYS, 2007) by perfomring nonlinear finite element analysis of a series of beams under fire exposure. The ANSYS software is capable of handling geometric and material nonlinearities as well as thermal analysis. 7.2.1.2 Finite Element Model Finite element analysis was used to generate the temperature-deflection history of a series of simply supported beams. The finite element model was created in ANSYS and comprised two sub-models; namely, the thermal model and the structural model. The thermal model provided the temperature distribution in the steel member which was applied as a temperature-body-load in the structural model of the steel beam. For thermal analyses, two types of elements, namely PLANE55 and SURF151, were used. The cross- section of the steel beam was meshed with PLANESS elements. Heat transferred through convection and radiation was applied on the exposed boundaries of the section using the SURF151 element. For structural analysis, the beam was discretized using 90 BEAM189 elements which account for material and geometrical nonlinearities. 189 Stress-strain curves given in ASCE (1992) and shown in Figure 7.3 were used in the ANSYS finite element model. These stress-strain curves account for strain hardening. To account for primary and secondary effects of creep in structural steel, the “implicit creep model 11” built into ANSYS was used. This creep model is given by: Atecr = e _ 62 1 c3 -511. 66 .51 primary +“Elsecondary —Clos at €Xp( T )+C5t0.s 6Xp( T ) (7‘5) S S where c1 = 5><10-6 per minute, c2 = 6.95, (‘3 = — 0.4, C4 =16500°C, c5 = 20le3 per . —5 —3. . . minute, C6 = 6x10 , C7 = 5x10 C, t = time (mm), as = steel stress (MPa), T5 = steel temperature (°C), A80 = increment of creep strain, eprimary = primary creep strain, and eprimary = secondary creep strain. This creep model has been calibrated with tests as shown in Figure 7.4. Details on the calibration and validation of the ANSYS high temperature creep model can be found elsewhere (Kodur and Dwaikat 2009). Test data reported by Thor (1973) and Wainman and Kirby (1987) were used to validate the structural finite element model in ANSYS. The tested beams were a protected HEZZOB section and an unprotected UB356x17lx67 section which were simply supported and uniformly loaded. The protected beam was loaded to 62% of its room temperature yield capacity and the unprotected beam was loaded to 35% of its room temperature ultimate capacity. The mid-span deflections predicted by AN SYS are compared with test data in Figure 7.5, and the predictions are very good. 190 -- 80 500 — T: 20°C T= 200°C 400 — “ 60 T: 400°C ,3 ._ g 3°° T= 500°C 40 3 E - m -' (D «n 3 20° T= 600°C 3; b —- 20 (D U) 100 T: 700°C T= 900°C I 0 fr: 1 l l I I l l O L o 2.5 5 7.5 10 12.5 15 17.5 20 ." Strain, % I Temperature (°F) 32 392 752 1112 . 1472 1832 E‘ 1.0 - 3. EfTs YES 9 0.8 '1 °' 0.6 - 3 J ,5 0.4 '2 a. / ‘5 . Fun-S Wu 2 0.0 . . r . . o 200 400 600 800 1000 Temperature (°C) Figure 7 .3 - Mechanical properties for structural steel at elevated temperature 191 "° A36 Steel 6 i - ANSYS creep model ' i 0‘3 1 - -o- - - Creep test data (10°Clmin. ‘ o\° . g 0.6 O = 50 MPa ‘5 o = 100 MPa 1:. g 0.4“ 0:150 MP8 0 0.2 ~ 0.0 1 . r 0 150 300 450 600 750 Steel temperature, °C 2.0 1 ,’ o = 122.6 MPa / , " Gradeso Steel 1'5 i .I ANSYS creep model 32 .' 'o - - 4 - — Creep test data (600'C) 519.: o, ' \a=98MPa 3:: ' a = 73.5 MPa 0 = 49 MPa gas .' \ ,, _, - 0.0 l —‘ - _ ’ —J¢ r 0 50 100 150 200 250 300 350 Time, mln Figure 7.4 - Validation of high-temperature creep model 192 l' Load = 62% of R.T. yield capacity 600 ~ Temp. of bottom flange -- 120 A A mmnmmmmmmmmnn _ E g 500 - .L. l . .I —- 100 z e 400 ‘ L23200 mm . __ 80 ,g 3 - o E 300 _ Temp. of ‘ -L 60 £3 a top flange .. 3 g 200 - K ._ 40 r: l- . - Test 3‘, "' 100 - . uJ" -- 20 . ,3 - -/'/ Model .9 ‘0 o - . ‘ . . . . . o 5 0 2O 40 60 80 100 1 20 140 Time (minutes) (a) 800 . Load = 35% of R.T. ultimate capacity g 160 llllllllllllllllllllllllllllllllll __ 140 A A .A. .I. - E 9 600 r L=4500 mm ‘” 120 E I: T TGSt : h emp. of bottom\A , -_ 100 o 3 flange 11':- 2 400 - ~- 80 § 8. Temp. of u .5 E top flange _- 50 D a ./ c '— fl _ 200 - .4/\ -- 4o 9., .3 - e a) / - Model 2- 20 E /‘ O . l I T l f 0 0 5 10 15 20 25 30 Time (minutes) (b) Figure 7.5 - Comparison of deflections of steel beams predicted by ANSYS with the test data for (a) protected beam, and (b) an unprotected beam 193 7.2.1.3 Parametric Analysis In order to obtain the equivalent flexural rigidity through Equation 7-4, parametric analysis was carried out using the finite element model validated above. The parameters that were varied in the analysis included different slenderness ratios (Ury = 123, 185 and 246), and different load ratios (30%, 50% and 70% of room temperature moment capacity). The analysis was carried out by applying the load gradually on the simply supported beam until the room—temperature response was obtained, and then the steel temperature was increased, simulating fire exposure, until failure occurred. Failure was defined to occur at the last time step for which the iterative analysis converged. Using the temperature-deflection history obtained from the nonlinear analysis described above, Equation 7-4 was used to compute the equivalent flexural rigidity of the beam as a function of steel temperature. Both the equivalent flexural rigidity and the bending moment capacity of the beam significantly reduce with increase in steel temperature. Since the rate of reduction of these quantities depends on many factors, such as load and geometric characteristics, both flexural rigidity and the bending moment ratio were normalized as follows: _ [Es (T5)I]eq —[Es (T5)I]2 - 7-6 5 1E, (T,)I]1 —[E_, (Tsfllz ( ) and M - M y,T w = (7-7) Mu,T —My,T where f = normalized flexural rigidity, 1,11 = normalized inelastic moment and M = applied moment. 194 Curves of normalized flexural rigidity and normalized inelastic moment are shown in Figure 7.6 for different load and slenderness ratios. The values of w from 1.0 to 1.25 correspond to the strain-hardening range. As the steel temperature increases, the bending moment capacity of the beam reduces, and this leads to an increase in the normalized inelastic moment, t//. Also, the deterioration of steel properties and the spread of plasticity in the steel beam cause further deflection in the beam, and thus reduce the normalized flexural rigidity, f, of the beam. Figure 7.6 also shows that the influence of initial load ratio on the normalized flexural rigidity is nonlinear with the worst case of reduction in 5 occurring at a load ratio of 50%. This can be explained by the fact that the normalized flexural rigidity is a measure of the change in deflection with respect to a change in moment in the beam. Both flexural rigidity and moment ratio (initial applied moment to moment capacity) vary with steel temperature. Generally, for beams with higher initial moment ratios prior to fire (higher load ratios), the spread of plasticity along the beam occurs sooner than for beams with a lower initial moment ratio (lower load ratios). Therefore, the rate of deflection for beams with higher load ratios is generally more gradual as compared to beams with lower load ratios. The nonlinear effect of load ratios on the normalized flexural rigidity also can be attributed to geometric nonlinearities at such high inelastic deflection rates and magnitudes. Both geometrical and material nonlinearities are responsible for the nonlinear degradation of the flexural rigidity of the beam. In order to account for this degradation, an approximate lower bound curve was fitted to the curves shown in Figure 7.6. This curve (model) is given by 195 1.00 0.80 0.60 0.40 0.20 Normalized rigidity, § 0.00 (7-8) — Load ratio = 30% — - Load ratio = 50% ---- Load ratio = 70% Fitted model (Equation 7-8) L/ry = 123 0.00 0.25 0.50 0.75 1.00 1.25 Normalized moment, at Figure 7.6 - Variation of normalized flexural rigidity with normalized inelastic moment for different load and slenderness ratios Equation 7-8 is also plotted in Figure 7.6. From Equations 7-6, 7-7 and 7-8, the equivalent flexural rigidity may be expressed as: [Es (Ts [Es(Ts)1]1 -[Es (T0112 1+ 1511/25 The equivalent flexural rigidity computed by Equation 7-9 can be used to calculate the fire induced inelastic deflection of simply supported steel beams. Equation 7-9 is based on a conservativ ely fitted curve to results from nonlinear finite element analysis. At M = 196 Myj the value of [ES(TS)I]eq = [ES(TS)I]1 as desired. However, when M = Mu], the value of [ES(TS)[]eq = (15[ES(TS)I]2—[ES(TS)I]1)/ 16 which is not equal to [ES(TS)I]2 as desired. Near ultimate strength failure (M = MN), the mid—span deflection of the beam is quite significant that it alters the horizontal projection of the beam length between the supports. This is reflected by a stiffened response of the beam that leads to values of [E 5(TS)I]eq that are slightly higher than the idealized [ES(TS)I]2 near failure. In Figure 7.6, the curves for normalized flexural rigidity seem to be dependent on the load ratios. However, for simplicity, an approximate lower bound curve was fitted (Equation 7-8). This was done for two reasons: (1) Most buildings have a load ratio of 0.5 or less (Buchanan, 2001) and the fitted curve matches the predicted curves well at this load ratio; (2) As shown later, for smaller load ratios (e.g. 0.35), the simplified method will yield conservative results but the conservatism is less than 10%. It is also shown in the next section that Equation 78 is proposed in such a way that it takes into consideration the influence of high temperature creep on the fire-induced deflection of beams. Creep is insignificant in structural steel at room temperature, but it becomes very significant at temperatures above 400°C and at higher loads. Despite the significant effect of creep deformations on steel structures exposed to fire, creep is usually not explicitly included in the design process under fire because of the lack of data and difficulty of calculations. The usual assumption made is that the stress-strain relationships used for fire design are ‘effective’ relationships which implicitly include creep deformations 197 (Buchanan 2001). Other researchers (Anderberg 1986 and Poh 1995) have shown how creep deformations can be explicitly included in computer models. Since the main focus of this study was to develop a simple expression for calculating deflections, instead of including creep deformations through a separate factor that is likely to be complex, the effect of creep deformations on the equivalent flexural rigidity is accounted for through the fitted equation (Equation 7-8) that is an approximate lower bound of the curves shown in Figure 7.6. In the next section, while validating the simplified method, creep deformations are explicitly included in the ANSYS analysis using the creep model given by Equation 7-5 and it is shown that the equivalent flexural rigidity expression obtained from Equation 7-9 gives very reasonable results. 7.2.2 Validation of Simplified Method For simplicity, most design specifications such as the AISC Specifications (2005) and Eurocode 3, allow the use of a uniform temperature distribution across the steel section. The uniform temperature can be calculated by simple thermal analysis methods (e.g., lumped heat capacity method) instead of sophisticated computer software. Therefore, the simplified method developed herein is based on the assumption of a uniform temperature distribution. Therefore, results from ANSYS when the temperature distribution is kept uniform are used to validate the proposed simplified method. In actual fire tests, the temperature distribution across the section is not uniform and there is a significant temperature gradient. Therefore, in the next section, the fire resistance computed from the simplified method is compared with that measured in actual fire tests, wherein, an' average temperature of the steel section is used and the deflections resulting due to the temperature gradient are accounted for through a separate term. 198 In performance based approaches currently being promoted, there is likely to be a limit on deflections to ensure the integrity of the fire compartment and to provide safe conditions for fire fighters. Until now, this deflection limit is not well defined and agreed upon. The British Standard (BS 476-20 1987) suggests a deflection limit of [/20 that has been adopted by most researchers. However, a deflection limit of U30 has also been used by many researchers. In validating the simplified method, it is therefore appropriate to compare the fire resistance time given by the simplified method with that predicted by ANSYS when the mid-span deflection of the beam reaches the limiting values of [/30 and [/20 In the succeeding sections, although the entire deflection curves are provided, the discussion focuses on the comparison of the fire resistance time between the limiting deflection values of U30 and U20, because the intent of the simplified method is to predict accurate deflections at these limits (U30 and U20). It is desirable that the proposed method be applicable to: all types of loading configurations; different fire scenarios (real fires and standard fire); varying heating rates of steel (slow or fast); all load ratios; and different sectional geometry and span lengths. This is yet another reason why ANSYS simulations rather than experimental data were used to initially validate the simplified method, because many parameters are not explicitly reported for experiments. Creep deformations are significant at slower heating rates and higher loads. Most buildings have a load ratio of 0.5 or less (Buchanan 2001). A load ratio (LR) of 0.5 is used for all comparisons except when different load ratios are explicitly mentioned. A relatively slow heating rate (HR) of 45°F/min (7°C/min) is used for most of the comparisons. However, design fires producing heating rates lower and higher than 199 45°F/min (7°C/min) are also used to validate the simplified method for different heating rates. In reality, the heating rate of steel varies with time, but in this study the heating rate refers to the average heating rate obtained by dividing the maximum steel temperature by the time taken to achieve that temperature. 7.2.2.1 Effect of Heating Rate of Steel Creep deformations form a significant part of the total deflection of beams at elevated temperatures and depend on the rate of heating and the load ratio. To determine whether the simplified method works for all heating rates, four fire scenarios producing different heating rates of steel were selected. Slow and fast heating rates of steel sections represent well insulated and poorly insulated (or unprotected) steel beams, respectively. The deflectionspredicted by the simplified method and ANSYS are shown in Figure 7.7. The simplified method gives deflections very close to those given by ANSYS and the maximum difference in the fire resistance time (i.e., the time when the deflection reaches the limits of U30 and U20) from. the two methods is less than 2%. 7.2.2.2 Effect of Load Ratio Creep deformations are strongly dependent on the load ratio and increase for higher load ratios. The deflections predicted by the simplified method for five load ratios are compared with the ANSYS predictions in Figure 7.8. The two methods give very similar deflections and the maximum difference in the fire resistance times is less than 7% at the deflection limits of U30 and U20. The simplified method gives a relatively conservative estimate of the deflections for low load ratios (e.g., 35%) because the fitted curve for normalized flexural rigidity (see Equation 7 -8 and Figure 7.6) is lower than that estimated 200 from finite element simulations. As mentioned earlier, most structural members will have a load ratio of about 0.5 under fire (Buchanan 2001) and the simplified method gives very accurate results for this case. . 5 2° 450 fig Deflection limit = Ll20 5 16 375 i Deflection limit = U30 c- E 300 " 12 2' 5’ HR = 85'Flmin. ; .3 .3 225 - e : HR = GO'Flmin. -' a g g 150 _ .4— HR = 50'Flmin. . 8 3 - HR=40°FImin._. ' 4 . . . o 60 80 100 120 Time, min Figure 7.7 - Deflections predicted by ANSYS (broken lines) and simplified method (solid lines) for different heating rates. Load ratio = 0.5, load type = UDL I g . .1 20 450 K ' 5 375 4 i a 243° . .5- E 300 3 ' :- LR = 0.80 .1 12 § .2 225 - LR = 0.70 , ‘6 i-l ' a) 3 LR = 0.60 , .- 8 g g 150 — LR = 0.50 ,' o 0 LR = 0.35 ,' - 4 75 — 0 ‘ t ' 0 0 20 40 60 80 100 120 140 Time,min. Figure 7.8 - Deflections predicted by ANSYS (broken lines) and simplified method (solid lines) for different load ratios. HR = 45°F/min, load ratio = 0.5, load type = UDL 201 7.2.2.3 Effect of Sectional Geometry and Span Length The deflections predicted by the simplified method are compared with the ANSYS predictions for three steel sections (light, heavy and intermediate) and for three span lengths (20 ft, 30 ft and 40 ft) in Figure,7.9. The maximum difference in the fire resistance times at the deflection limits of U30 and U20 predicted by the two methods is less than 1%. Since the three beams are of different lengths, the deflection limits of U30 and U20 will be different for each beam. In Figure 10, only the lowest U30 and highest U20 are shown. - 24 60° 3 H' tht L120 t: : I e 525 ~ 9 I. a: 20 I e 450 - ,0' - 16 - E. 375 ~ w12x1s———> . ‘5- C ,9 30° _ W18X35 ,r' l 12 g E 225 _ waxes? *‘7‘ i 3 § 0) ' d) D 150 ‘ Lowest U30 I 5" _ 4 o 75 ‘ .0...' 0 r - ' ° 0 25 50 75 100 125 Time,min. Figure 7.9 - Deflections predicted by ANSYS (broken lines) and simplified method (solid lines) for different sections and span lengths. Load ratio = 0.5, HR = 45 °F/min, load type = UDL 7.2.2.4 Effect of Load Configuration In all of the analyses and comparisons hitherto, the beam was subjected to a uniformly distributed load. It is now demonstrated that the simplified method also gives good results for other load configurations. In Figure 7.10, deflection profiles are compared for two loading configurations: a beam subjected to a single point load at mid-span and a beam 202 subjected to two-point loads at one-third spans. The mid-span deflections for these cases are PL3/ (48[ES(TS)I]eq) and 23PL3/(648[ES(TS)I]eq).The deflections predicted by the simplified method agree very well with the ANSYS simulations. 500 "l i t E 20 - . . . /‘ i 'l' g E 400 iDeflectronltmrt=L/20 :0 : 16 .E g- Deflection Ilmlt = L/30\ 2': I :1 é "3 3°°l :' .' ‘2 .8. g 3 p t 1 ”5 .3 Two point load t '0 1: c 200 : . 8 5 (U i ' . a CL 1 ;: 00 '3 100 5 4 352 E ’ One i 1 point load I o :’ : . *T’" . i J: O O 25 50 75 100 125 150 Time. min. . Figure 7.10 - Deflections predicted by ANSYS (broken lines) and simplified method (solid lines) for different load configurations. Load ratio=0.5, HR: 45 °F/min At the limiting deflection value of U30 and U20, the maximum difference in fire resistance time predicted by the two methods is less than 10% for the case of a single point load. 7.2.3 Comparison with Test Results As mentioned earlier, the temperature distribution through the depth of the steel section is not uniform under fire conditions and a significant temperature gradient exists, especially for steel beams supporting a concrete slab on the top flange. Since deflections of beams at elevated temperatures occur primarily due to a reduction in stiffness of the section and the temperature gradient across the section, both of these effects need to be considered to 203 predict deflections accurately. The contribution to deflection due to stiffness reduction is effectively accounted for by the use of an equivalent flexural rigidity. However, a separate term needs to be introduced to account for the contribution of the temperature gradient to deflections. If the thermal gradient is assumed to be linear through the beam depth, the additional elastic deflection due to the temperature gradient is L2 AT Ap,TG = a??? (7-10) where, a = coefficient of thermal expansion of steel, AT = the difference between the temperature of the top and bottom flanges, d = the section total depth, L = beam span length. The total inelastic deflection of the beam, including that due to temperature gradient, is then Ap,total = Ap + Ap,TG (7-11) In Table 7.1, the fire resistance time at the limiting deflection value of [/30 predicted by the simplified method is compared with test results for 13 beams. Beams 1 to 12 were tested in the UK. (W ainman and Kirby 1987 and Wainman 1992), and the data for beam no. 13 was reported by Thor (1973). Beams 1 to 11. were unprotected, and beams 12 and 13 were protected by fire protection material. These tests were terminated when the mid- span deflection of beams reached a limiting value of [/30. For each beam, the mid-span deflection and the mean temperatures of the bottom flange, top flange and web were reported. In the simplified method, AP is computed using the average of the top flange, 204 bottom flange, and web temperatures of the section in Equation 7-9, and by using the temperature difference between the top and bottom flanges in Equation 7- 10. Table 7.1 - Comparison of fire resistance time predicted by the simplified method with test results at the limiting deflection value of L/30 Load Fire Resistance Time (min) Test # Beam Section Length (ft) Ratio (%) Test Model Difference 1 U8254x146x43 15.00 46 23 22.4 -0.6 2 U8254x146x43 15.00 53 22 21.3 -0.7 3 U8356x171x67 14.75 56 27 24.6 -2.4 4 UB406x178x60 14.75 35 23 2.1 -0.9 5 UB305x165x54 14.75 34 22.5 21.3 -1.2 6 U8254x146x43 15.00 34 30 30.8 0.8 7 U8356x171x67 14.75 35 24.5 24.5 0.0 8 U8254x146x43 14.75 44 21 21.6 0.6 9 U8254x146x43 14.75 45 22 21.1 -0.9 10 U8356x1 71 x67 14.75 51 29 24.6 -4.4 11 UB254x146x43 14.75 36 27 24.7 -2.3 12 UB254x146x44 14.75 67 93 91.7 -1.3 13 HE220 B 10.50 55 130 124.2 -5.8 7.3 Deflections of Simply Supported Beams 10 laterally restrained and 14 laterally unrestrained beams were designed for fire using the performance—based approach. The required thickness of insulation was determined through Equations 5-1, 5-3, 5-5 and 5-6 using an iterative procedure. The capacity reduction and fire load factors given in Figure 5.1 were used. The deflection of the beams were computed through Equation 7-1 if Mu,f< My], and through Equation 7-2 if Mu,f> MyJ. E 5(Ts)l]eq was calculated through Euqation 7-9. Muf was as Muf=MD+Mlflw (7-12) 205 I-nl‘mufl ’ .. s where MD = applied moment due to dead load and Mum, = applied moment due to arbitrary—point—in-time live load. Arbitrary-point-in-time live load, [apt was assumed to be 0.24Ln (Ellingwood 2005). The computed deflections are compared with limiting deflection values of U30 and U20 in Tables 7.2 and 7.3. It was found that deflections are generally smaller than the limiting values of U30 or U20, and the strength limit state is critical in case of simply supported beams. 206 A0063 80448 50.30 80.20 50¢: 50.50 80.20 80.50 50.9 END 0:: 0N3 £232,001 0w.wwo 00.000 othv 006w 0thv 0N>$v 00.3w 0w.v0m 00.wNN $.me EE 951:: Bow: A006: 80.20 60d: 80.20 80.30 80.0: 80.8 80.8 Swév 0:0 0Q:— :ouuouov 0N3»V 0v.00v 0m.v0m 0wé0m 0w.v0m 0w.v0m 00.me 0Nm0N OWNS N042 SE 951:: Amman: 30.8 C ENC $0.9 35.0 A2 .3 $NNO Sn. 3 $0.: :5 :omugcon 0N3 m 0N2 N 006m 34.: 0m.mm mNmb 38m wfimv MVKN EE 2 5.00 Sm. C A0000 A3200 Swdv €0.00 ondv $500 00000 Aowdv 000: 0.34 3: 0m.0~ 3.3 wmé ONE w0.: End vwfi 3.2 :.w w0.: E\Zx SNOW 2 0.: £104.00 8500 €5.00 30.00 30.00 $0.00 $0.00 20.00 0030 was 08G 3.2 0N2 owd 3.: vN: mod mod ovd mvd wNQ :56. 33 8E 88 89 82 80 $8 88 G: 8: at 5&0 N59 2.9 3.0 3.0 3.0 3.0 SH 000 hmé 09m E N0 :23 0534.95 0950 >9 035:5 mmxf? #49:: >9 mmxfla 336:5 mfixmg 2ow :cfloow anm 0:— 0: mm FM on ma em M»— um um— $_:D 53:55:."— 0N: .3 ems no 339» «.515. 2: .53 03.253 ban—v. 005953.. £9.83 ae 28:00:00 he car—505.5 I Nb wish 207 839 83: 83: 839 3.8: :5 as 5:89... 5:55 has 839. 2.8% 838 838 as 83: 88.8: 83: 83: 88.9 2: 83.55858 5:55 8.93 3.88 33mm 3.84 832 as 5.: 3.9 84.: 98.: 84.9 2: 838:5 833 8.2 3.? 9R 8.: as 83.9 3.9 8M9 24.9 3.9 8%: .33 2: 33 m3 3? 33 a” 8;: 24.9 3.9 3.9 34.9 3.9 2%: 33 use 3.8 :3 9% 2.8 m3 .572 839 839 3.3 3.89 83: a: gum 8.: 83 3.9 2.2 83 3x33 93;» 258;» 333 83;, ii 5:358“. m8 Em an 2m 2m 5:5 238.3...— 839 83.9 83: 83: 83: 83: 83: 88.3: 88.9 83.: :9 83 838 88.88 833 8.3 833. 833 8.3. 93.8 8333 339 as .8383. mice: 83: 83: 83: 83: 83: 83: 88.8: 88.9 88.9 83.: :5 on: 839. 84884 93.83 938m 938 83.33 8.3.“ 83.83 932 ~32 5:. 82.89% ”:95: :3: 3.9 3.: :2: 3.9 3.9 3.9 :8: 84.: 88.9 :5 558:5 83.: 84.89 9.2: :89 33.8: $.33 5.: .34 8.8 2.9 55 3.9 am: 839 3.9 83.9 :89 84.9 3.9 8n9 83.9 3:: 332: 83.9 83.3 33 9.9 38.: 3.8 3 m 8.9 83 38.: .57: 3.9 :2: :49 39 E9 829 88.9 88.9 88.9 83.9 8.8: 332.5 :9 8.: 93 8.9 3.: 8.9 9 o 2.8 Rd 38.: :59 Am: 8: 89 89 89 89 3: 9 Am: a: a: 5% mm: 9.9 :8 3.8 3.3 3.8 33 9.8 $4. 83 E 83:5 8:33 8993 93;» 38:5 3333 88:5 “98:5 233 333 Seaman:— Sm a. w: E on ma 3. B S 3 £5 220525. can .8 as as mos—a.» w:EE= 20 53 08.8.5.8 39:6 0955.58.53 309.32 .3 8:808:00 .3 58.39.80 I m6 Bash. 208 7.4 Summary In this chapter, a simplified method was developed for computing deflectios of simply supported steel beams exposed to fire. In this method, deflections are predicted by the same expressions used for predicting the elastic deflection of beams at room temperature but using an equivalent flexural rigidity instead of the elastic flexural rigidity. At the deflection limit of U30 to U20, the simplified method predicts fire resistance times very close to those predicted by the ANSYS finite element program which accounts for thermal and creep effects. The fire resistance time predictions by the simplified method also compare very well with test data at the deflection limit of U30. The proposed method is very simple and can be included in design specifications for the fire—resistant design of simply supported steel beams. The deflections of simply supported beams designed for the strength limit states were also computed using the simplified method, and it was found that the deflections were smaller than the limiting values of U30 or U20. 209 l‘fig Chapter 8 Conclusions and Recommendations 8. 1 Summary A general reliability-based methodology is proposed for developing capacity reduction and fire load factors for LRFD of steel members exposed to fire. Statistics (mean, COV, and distribution) of a variety of parameters such as fire load, opening factor, thermal absorptivity of compartment boundaries, thickness of insulation, thermal conductivity and density of insulation were obtained from experimental data reported in the literature. Model errors associated with the thermal models were also characterized based on experimental data. Both 3-sided and 4-sided fire esposure were considered while developing thermal model errors. Experimental data applicable to US. conditions was used whenever possible. In the proposed methodology, the effect of active fire protection systems (e.g., sprinklers, smoke and heat detectors, etc.) in reducing the probability of occurrence of a severe fire is accounted for by using a reduced target reliability index for developing capacity reduction and fire load factors for steel members exposed to fire. 210 Based on the information available about the effectiveness of active fire protection systems in reducing the proability of occurrence of a severe fire, a range of target reliability index value was established for US. office compartment ranging in floor area from 25m2 to 500m2. Both compartmented and open plan office buildings were considered to establish the range of floor areas. To illustrate the proposed methodology, capacity reduction and fire load factors were derived for simply supported steel beams and axially loaded steel columns in US. office buildings exposed to fire. Capacity reduction and fire load factors were developed for a range of target reliability index values (0-2). Fire design provisions given in the AISC Specifications and the literature were used for steel beams and columns, respectively. The capacity reduction and fire load factors developed in this study can be used for the fire design of simply supported steel beams and columns in US. office buildings instead of the current factors which are primarily based on subjective judgement. For ease of use, linear relationships are developed for computing capacity reduction and fire load factors for a given value of target reliability index. Capacity reduction and fire load factors for other types of structural members (e.g., restrained beams) and for other type of occupancies may be developed using the proposed methodology, model errors, and the statistics of fire design parameters. In addition, a simplified method is proposed for computing inelastic deflections of simply supported steel beams exposed to fire, and it is shown that the fire design of most beams is goverened by the strength limit state. The proposed method is very simple and can be included in design specifications for the fire-resistant design of simply supported steel beams. 211 1F 8.2 Conclusions The main conclusions resulting from this study are as follows: The present capacity reduction and fire load factors in the AISC fire design specifications are primarily based on experience and subjective judgement, and do not consider the statistical variations of the various parameters involved in the design. There was only one study done in Sweden to to develop capacity reduction factors for the design of steel members exposed to fire. No study has been done so far to develop capacity reduction and fire load factors for LRFD of steel members in the US. The effect of active fire protection systems, especially sprinklers, in reducing the. probability of occurrence of a severe fire is well recognized in the literature, but, has not been explicitly included in developing capacity reduction factors for design of steel members exposed to fire. From the statistics of fire design parameters developed in this study, it is found that uncertainty associated with the fire design parameters is much higher than that associated with room temperature design parameters. The errors associated with thermal models are also much higher than those associated with structural models used for room temperature design. The capacity reduction and fire load factors correspond to a preselected target reliability index that accounts for the effect of active fire protection systems (e.g., sprinklers, smoke and heat detectors, etc.) in reducing the probability of 212 occurrence of a severe fire. Therfore, for uniform safety, these need to be developed for a range of target reliability index values. 0 Based on the effectiveness of active fire protection systems reported in the literature and the typical fire compartment sizes in US. office buildings, it is resonable to develop capacity reduction and fire load factors corresponding to target reliability index values ranging from zero to 2.0. 0 It is found that the fire load factor should vary depending on the presence of active fire protection systems, and the fire load may be reduced in buildings equipped with very effective and reliabile active fire protection systems, especially, sprinklers. This is in agreement with the Commentary to the AISC Specifications, the Eurocode l, and the ECSC study. 0 Capacity reduction factors for simply supported beams and axially loaded steel columns in US. office buildings should vary based on the presence of active fire protection systems in the buildings, and should not be constant as suggested in present design specifications. 0 For most office building compartments in the US. equipped with sprinklers, using capacity reduction factor value of 1.0 is reasonable, and fire load factor is likely to lie between 0.4 and 1.0. Current structural fire design provisions are still relatively new and evolving, with various remaining uncertainties and information gaps. The methodology proposed herein is an initial attempt to characterize uncertainties in current fire design provisions. It is expected that as the research in this field yields improved understanding of structural behavior under fire, future design refinements will be necessary. 213 8.3 Recommendations for Future Research While this study has advanced the state—of—the-art with respect to the development of capacity reduction and fire load factors for LRFD of steel members exposed to fire, further research is required to develop these factors for all types of structural steel members used in different types of occupancies. The following are recommendations for further research in this area: Capacity reduction factors developed herein are specific to simply supported beams for typical fire compartments in US. office buildings. However, the methodology presented in this study and the random parameters and model errors characterized herein are general and can be used for deriving capacity reduction and fire load factors for other types of structural members such as restrained beams and beam - columns. As stated in the Design Guide for Structural Fire Safety (CIB W14, 1986), the fire load varies considerably from one type of occupancy to another. Therefore, it is not possible to develop one set of capacity reduction and fire load factors for all types of occupancies. Accordingly, one set of factors may be developed for any type of occupancy (say office buildings), and then differentiated factors can be developed to modify these basic factors for other types of occupancies. All random parameters characterized in this study except for the fire load, ratio Af/At, and arbitrary-point-in-time live load, are applicable for the reliability analysis of buildings of all use categories. The statistics of the fire load and arbitrary-point- in-time load for different types of occupancies are reported in the literature. However, not much information is available about the ratio Af/At, and these 214 should be obtained from architectural drawings of buildings. Subsequently, differentiated factors for buildings of other use categories (e.g., residential buildings, industrial buildings, etc.) can be developed using the methodology described in this study. 0 The statistics of fire design parameters can be further improved by using more tests data as it becomes available. 215 Appendix A Journal and conference papers resulting from the work are: (l) Iqbal, S., and Harichandran, RS. (2009). “Capacity reduction and fire load factors for design of steel members exposed to fire.” ASCE Journal of Structural Engineering. (paper re-submitted after addressing reviewers’ comments) (2) Iqbal, S., and Harichandran, RS. (2009). “Capacity reduction and fire load factors for LRFD of steel columns exposed to fire.” Fire Safety Journal. (paper re-submitted after addressing reviewers’ comments) (3) Iqbal, S., Dwaikat, M.S., and Harichandran, RS. (2008). “A Simplified Method for Determining Fire Resistance of Steel Beams Exposed to Fire.” AISC Engineering Journal. (submitted) (4) Iqbal, S., Harichandran, RS, and Kodur, V., (2008). “Capacity reduction factor for flexural strength of steel beams exposed to fire.” Proceedings, Inaugural International Conference of the Engineering Mechanics Institute, Minnesota. (5) Iqbal, S., and Harichandran, R.S. (2008). “Reliability-based design specifications for simply supported steel beams exposed to fire.” Proceedings, 10th International Conference on Structural Safety and Reliability, Osaka, Japan. (6) Iqbal, S., Dwaikat, M.S., and Harichandran, RS. (2009). “A simple approach for calculating inelastic deflections of simply supported steel beams under fire.” Proceedings, ASCE Structure Congress 2009, Austin, Texas. (7) Iqbal, S., and Harichandran, RS. (2010). “Methodology for reliability-based design of steel members exposed to fire.” Proceedings, International Colloquium: Stability and Ductility of Steel Structures, Rio de J aneiro, Brazil. (8) Iqbal, S., and Harichandran, RS. (2010). “Capacity reduction and fire load factors for steel columns exposed to fire.” Proceedings, International Colloquium: Stability and Ductility of Steel Structures, Rio de J aneiro, Brazil. (9) Iqbal, S., and Harichandran, RS. (2010). “Reliability-Based capacity reduction and fire load factors for design of steel members exposed to fire.” Proceedings, 6th International Conference on Structures in Fire, East Lansing, United States. 216 References 7.] 10. ll. 12. Aderberg, Y. (1986). “Measured and predicted behavior of steel beams and columns in fire.” Lund Institute of Technology, Lund, Sweden. AISC (2005a). Load and resistance factor design specifications for structural steel buildings. American Institute of Steel Construction, Inc., Chicago, IL. AISC (2005b). Commentary on the load and resistance factor design specifications for structural steel buildings. American Institute of Steel Construction, Inc., Chicago, E. ANSYS Inc. (2007). ANSYS software and manual. Release ll.OSPl UP20070830, Canonsburg, PA. ASCE. (1992). Structural fire protection. Manual No.78. American Society of Civil Engineers, Structural Division, New York: ASCE Committee on Fire Protection, New York, NY. ASCE/SFPE29 (1999). Standard calculation method for structural fire protection. American Society of Civil Engineers, Reston, VA. ASCE-O7 (2005). Minimum design loads for buildings and other structures. American Society of Civil Engineers, Reston, VA. ASTM (2008). Standard methods of fire test of building construction and Materials. ASTM Test Method E119-08a, American Society for Testing and Materials, West Conshohocken, PA. ASTM (1993). Standard test methods for determining efiects of large hydrocarbon pool fires on structural members and assemblies. ASTM Test Method E1529, American Society for Testing and Materials, West Conshohocken, PA. Beck, V. R. (1985). “The prediction of probability of failure of structural steel elements under fire conditions.” Transactions of the Institution of Engineers, Australia, 27 (1), 111-118. Bentz, D. P., and Prasad, K. R. (2007). “Thermal performance of fire resistive materials 1. Characterization with respect to thermal performance models.” Report No.: NISTIR 7401 , National Institute of Standards and Technology. Bruls, A., Cajot, L. G., and Franssen, J. M. (1988). “Characterization of an insulating material with regard to ECCS recommendations for the fire safety of steel structures.” Journal of Constructional Steel Research, 9(2), 111-135. 217 13. 14. 15. 16. 17. l8. 19. 20. 21. 22. 23. 24. BS (1987). Fire tests on building materials and structures — Part 20: Method for determination of the fire resistance of elements of construction (general principles). BS 476-3:1987, London, U. K. Buchanan, A. H. (2001). Structural design for fire safety. John Wiley & Sons Ltd., New York. Bukowski, R. W. (2006). “Determining design fires for design—level and extreme events.” Proceedings of the SFPE 6th International Conference, Tokyo, Japan. Bukowski, R. W., Budnick, E. K., and Schemel, C. F. (1999). “Estimates of the operational reliability of fire protection systems.” Proceedings of International Conference on Fire Research and Engineering, Chicago, IL, 87-98. Burgess, I. W., El-Rimawi, J. A. and Plank, R. J. (1990). "Analysis of beams with non-uniform temperature due to fire exposure." Journal of Constructional Steel Research, 16, 169-192. Bwalya, A. C., Benichou, N., and Sultan, M. A. (2003). Literature review on design fires. Report No. IRC—RR-I37, Ottawa, Ontario, Canada. Carino, N. J ., Starnes, M. A., Gross, J. L., Yang, J. C., Kukuck, S. R., Prasad, K. R., and Bukowski, R. W. (2005). “Federal building and fire safety investigation of the World Trade Center disaster: Passive fire protection.” Report No.: NIST NCSTAR I-6A, National Institute of Standards and Technology, Gaithersburg, MD. Caro, T. C., and Milke, J. A. (1995). “A survey of fuel loads in contemporary office buildings.” Report No.: NIST-GCR-96-69, National Institute of Standards and Technology, Gaithersburg, MD. Chen, J ., Young, B., and Uy, B. (2006). “Behavior of high strength structural steel at elevated temperatures.” Journal of Structural Engineering, 132(12), 1948-1954. CIB W14. (1983). “A conceptual approach towards a probability based design guide on structural fire safety.” Fire safety Journal, 6(1), 1-79. CIB W14. (1986). “Design guide for structural fire safety.” Fire Safety Journal, 10(2), 81-137. Cornell, C. A. (1969). “A probability-based structural code.” Journal of the American Institute, 66(12), 974-985, 1969. 218 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. Corotis, R. B., and Doshi, V. A. (1977). “Probability models for live load survey results,” Journal of the Structural Division, 103(ST6), 1257-1274. Culver, C. G. (1976). “Survey results for fire loads and live loads in office buildings.” National Bureau of Standards, Washington, DC. Der Kiureghian, A., Haukaas, T., and Fujimura, K. (2006). “Structural reliability software at the University of California, Berkeley.” Structural Safety, 28(1-2), 44-67. Dharrna, RB, and Tan, K.H. (2007). “Proposed design methods for lateral- torsional buckling of unrestrained steel beams in fire,” Journal of Constructional Steel Research, 63, 1066-1076. Dwaikat M. B., and Kodur V. R. (in press), "Response of Restrained Concrete Beams under Design Fire Exposure." Journal of Structural Engineering, ASCE. ECCS Technical Committee 3 (2001). “Model code on fire engineering, lst edition.” Document No. 111, European Convention for Constructional Steelwork, Brussels, Belgium. ECSC. (2001). Natural fire safety concept. Valorization research project. Product-Structural Dept., ARBED-Research Centre L—4009 ESCH/ALZETTE. Ellingwood, B. Galambos, T. V., ‘MacGregor, J. G., and Cornell, C. A. (1980). “NBS special publication 577: Development of a probability based load criterion for American National Standard A58.” National Bureau of Standards, Washington, DC. Ellingwood, B. R. (2005). “Load combination requirements for fire-resistant structural design.” Journal of F ire Protection Engineering, 15(1), 43-61. Ellingwood, B. R. and Corotis, R. B. (1991). "Load combinations for buildings exposed to fires." Engineering Journal, AISC 28(1):37-44. EN (2002). Eurocode 1: Actions on structures- Part 1-2: General actions- Actions on structures exposed to fire. BS EN 1991-1-2: 2002. EN (2005). Eurocode 3: Design of steel structures- Part 1-2: General rules- Structural fire design. BS EN 1993-1-2: 2005. Feasey, R. and Buchanan, A. (2002). “Post-flashover fires for structural design.” Fire Safety Journal, 37(1), 83-105. 219 38. 39. 40. 41. 42. 43. 45. 46. 47. 48. 49. Fellinger, J. H. H. (2000). “Fire resistance; reliability vs. time analysis.” Proceedings of Composite Construction in Steel and Concrete IV Conference, Banff, Alberta, Canada. FEMA (2002). "World trade center building performance study: data collection." Preliminary Observations, and Recommendations, Federal Emergency Management Agency (FEMA), Federal Insurance and Mitigation Administration, Washington, DC. FEMA (2008). Firefighter fatalities in the United States in 2007, US. Fire Administration, FEMA, Washington DC, USA. Ferguson, D. (2008). Verbal communication. Reicherbach Corporation, Lansing, Michigan, US. Foster, 3., Chladna, M., Hsieh, C., Burgess, 1., and Plank, R. (2007). “Thermal and structural behaviour of a full-scale composite building subject to a severe compartment fire.” Fire Safety Journal, 42(3), 183-199. Franssen J. M., Kodur V. K. R., and Mason J. (2004). User manual for SAFIR. University of Liege, Belgium. Franssen, J. M., Talamona, D., Kruppa, J. and Cajot, L. G. (1998). “Stability of steel columns in case of fire: experimental evaluation.” Journal of Structural Engineering, 124(2), 158-163. Freudenthal, A. M. (1956). “Safety and the probability of structural failure.” ASCE Transactions, 121, 1337-1397. Gayle, F. W., Fields, R. J., Luecke, W. E., Banovic, S. W., Foecke, T., McCowan, C. N., Siewert, T. A., and McColskey, J. D. (2005). “Federal building and fire safety investigation of the World Trade Center disaster: Mechanical and metallurgical analysis of structural steel.” Report No.: NIST NCSTAR 1-3, National Institute of Standards and Technology, Gaithersburg, MD. Ghojel, J. I. and Wong, M. B. (2005). "Three-sided heating of I-beams in composite construction exposed to fire." Journal of Constructional Steel Research, 61, 834-844. Haldar, A., and Mahadevan, S. (2000). Probability, reliability, and statistical methods in engineering design. John Wiley & Sons, Inc., New York. Harichandran, R. S. (2005). “Class notes for CE 810: Reliability-based design.” Michigan State University, East Lansing, MI. 220 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. Harmathy, T. Z., and Allen, L. W. (1973). “Thermal properties of selected masonry unit concretes.” ACI Journal, 70(2), 132-142. Harmathy, T. Z., and Mehaffey, J. R. (1985). “Design of building for prescribed level of structural safety.” In Fire safety, Science and engineering, T. Z. Harmathy (ed.), STP 882, ASTM, Baltimore, MD. Hasofer, A. M., and Lind, N. C. (1974). “Exact and invariant second moment code format.” Journal of the Engineering Mechanics Division, ASCE, 100 (EMl),111-121. Intenational Code Council (ICC). (2000). International Building Code, 2000. Falls Church, VA. Intenational Code Council (ICC). (2003). International Building Code, 2003. Falls Church, VA. Iqbal, S. and Harichandran, R. S. (2009). “Reliability-based design specifications for simply supported steel beams exposed to fire. ” Proceedings of 10th International Conference on Structural Safety and Reliability, Osaka, Japan. . Isner, MS. (1988). “$80 Million Fire in Montreal High-Rise.” Fire Journal, Feb., 64-70. ISO. (1975). Fire resistance tests — elements of building construction. International Organization for Standardization, ISO 834-1975, Geneva, Switzerland. J anes, J. (1995). “Statistical analysis of fire tests on steel beams and columns to Eurocode 3, Part 1.2.” Journal Construction Steel Research, 33, 39-50. Janss, J. Minne, R. (1981). “Buckling of steel columns in fire conditions.” Fire Safety Journal, 4(4), 227-235. Karter M. J. (2008). "Fire loss in the United States in 2007." Fire Analysis and Research Division, National Fire Protection Association, MA, USA. Kelly, F. S., and Sha, W. (1999). “A comparison of the mechanical properties of fire-resistant and S275 structural steels.” Journal of Constructional Steel Research, 50(3), 223-233. Kirby B. R., and Preston P. R. (1988). “High temperature properties of hot- rolled structural steels for use in fire engineering design studies.” Fire Safety Journal, 13(1), 27-37. 221 63. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. Kirby, B. R., Wainman, D. E., Tomlinson, L. N., Kay, T. R., and Peacock, B. N. (1994). “Natural fires in large scale compartments—A British steel technical, fire research station collaborative project.” British Steel Technical, Swinden Laboratories, Moorgate, Rotherham. Klem, T. J. 1988. “First Interstate Bank fire, Los Angeles California, May 4, 1988.” Fire Investigation Report, National Fire Protection Association, Quincy, Massachusetts. Kodur V. R., and Dwaikat M. B. (2008). “Performance-based fire safety design of reinforced concrete beams.” Journal of Fire Protection Engineering, 17(3), 293-320. Kodur V. R. and Harmathy T. Z. (2008). “Properties of building materials.” In SFPE Handbook of Fire Protection Engineering, DiNenno P. J. (ed.), 4th edition, National Fire Protection Association, Quincy, MA, 1.167-1.l95. Kodur V. R. (2007). “Design strategies for protecting built infrastructure against fire hazard.” Proceedings of the Protect 2007 Conference, Whistler, BC, Canada, 1-1 1. Kodur V. R., Dwaikat M. B., and Raut, N. (in press). “Macroscopic FE model for tracing the fire response of reinforced concrete structures.” Engineering Structures. Kodur, V. K. R., and Dwaikat, M. M. S. (in press). “Effect of high temperature creep on the fire response of restrained steel beams.” Materials and Structures Journal. Koffel, W. E. (2005). “Reliability of automatic sprinkler systems.” (Mar. 25, 2009). Kruppa, J. (1979). “Collapse temperature of steel structures.” Journal of the Structural Division, 105(ST9), 1769-1788. . Li, Fan, Fitzgerald, R. W. (1996). “Structural safety for fire conditions.” Proceedings of the Structures Congress XIV, Chicago, IL, 352-356. Li, T. T., and Kodur, V. K. R. (1996). “Thermal and mechanical properties of steel-fibre-reinforced concrete at elevated temperatures.” Canadian Journal of Civil Engineering, 23(2), 511-517. Lie, G-Q., Jiang, S-C., Yin, Y-Z., Chen, K., and Li, M-F. (2003). “Experimental studies on the properties of constructional steel at elevated temperatures.” Journal of Structural Engineering, 129(12), 1717-1721. 222 75. 76. 77. 78. 79.‘ 80. 81. 82. 83. 84. 85. Lin T. D., and Ellingwood B. (1987). “Flexural and shear behavior of reinforced concrete beams during fire tests.” Report N0. NBS-GCR-87-536, Portland Cement Association, IL, USA. Lin T. D., Gustaferoo A. H., and Abrams M. S. (1981). "Fire endurance of continuous reinforced concrete beams." R&D Bulletin RD072.0IB, Portland Cement Association, IL, USA. Luecke, W. E., McColskey, J. D., McCowan, C. N., Banovic, S. W., Fields, R. J., Foecke, T., Siewert, T. A., and Gayle, F. W. (2005). “Federal building and fire safety investigation of the World Trade Center disaster: mechanical properties of structural steel.” Report No.: NIST NCSTAR I-3D (Draft), National Institute of Standards and Technology. Pettersson, O., Magnusson, S. E., and Thor, J. (1976). “Fire engineering design of steel structures.” Publication No. 50, Swedish Institute of Steel Construction, Stockholm, Sweden. Madsen, H. O., Krenk, S. and Lind, N. C. (1986). Methods of structural safety. Prentice-Hall, Englewood Cliffs, New Jersey. Magnusson, S. E., and Pettersson, O. (1981). “Rational design methodology for fire exposed load bearing structures.” Fire Safety Journal, 3(4), 227-241. Makelainen, P., Outinen, J., and Kesti, J. (1998). “Fire design model for structural steel S420M based upon transient-state tensile test results.” Journal of Constructional Steel Research, 48(1), 47-57. Manzello, S. L., Park, S-H., Mizukarrri, T., and Bentz, D. P. (2008). “Measurement of thermal properties of gypsum board at elevated temperatures.” Proceedings of the 5th International Conference on Structures in Fire, Singapore, 656-665. Mehaffey, J. R., Cuerrier, P., and Carisse, G. (1994). “A model for predicting heat transfer through gypsum-board/wood-stud walls exposed to fire.” Fire and Materials, 18(5), 297-305. Mesquita, L. M. R., Piloto, P. A. G., Vaz, M. A. P., and Vila Real, P. M. M. (2005). “Experimental and numerical research on the critical temperature of laterally unrestrained steel 1 beam.” Journal of Constructional Steel Research 61, 1435—1446. ' NFPA (2002). Handbook of fire protection engineering. 3rd edition.” National Fire Protection Association, Quincy, MA. 223 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. Outinen, J ., and Makelainen, P. (2004). “Mechanical properties of structural steel at elevated temperatures and after cooling down.” Fire and Materials, 28(2-4), 237-251. Poh, K. W. and B., I. D. (1995). "Analysis of structural member under elevated temperature conditions." Journal of Structural Engineering, ASCE, 12(4), 664- 675. Rackwitz, R. and Fiessler, B. (1976). “Note on discrete safety checking when using non-normal stochastic models for basic variables.” Loads Project Working Session, MIT, Cambridge. Ramachandran, G. (1995a). “Probability-based building design for fire safety; Part 1.” Fire Technology, 31(3), 265-275. Ramachandran, G. (1995b). “Probability-based building design for fire safety; Part 2.” Fire Technology, 31(4), 355-368. Ravindra, M. K. and Galambos, T. V. (1978). “Load and resistance factor design for steel.” Journal of the Structural Division, 104(ST9), 1337-1353. Routley, J.G., Jennings, C. and Chubb, M (1991). “High-rise office building fire on Meridian Plaza, Philadelphia, US.” Federal Emergency Management Agency, US. Fire Administration, National Fire Data Center. Ruddy, J. L., Marlo, J. P., Ioannides, S. A. .and Alfawakiri, F. (2003). “Fire resistance of structural steel framing.” Steel Guide No. 19, American Institute of Steel Construction, Inc., Chicago, IL. Schmidt, B. J ., and Bartlett, F. M. (2002). “Review of resistance factors for steel: data collection.” Canadian Journal of Civil Engineering, 29(1), 98-108. Schneider, U., Diederichs, U., and Ehm, C. (1981). “Effect of temperature on steel and concrete for PCRV’s.” Nuclear Engineering and Design, 67(2), 245- 258. SFPE (2000). Guide to performance-based fire protection analysis and design of buildings. Society of Fire Protection Engineers, Bethesda, MD. SFPE (2004). Engineering guide: fire exposures to structural elements. Society of Fire Protection Engineers, Bethesda, MD. Shin, K-Y., Kim, S-B., Kim, J-H., Chung, M., and Jung, P-S. (2002). “Thermo- physical properties and transient heat transfer of concrete at elevated temperatures.” Nuclear Engineering and Design, 212(1-3), 233-241. 224 99. 100. 101. 102. 103. 104. . 105. 106. 107. 108. 109. 110. Sirrriu, E., and Filliben, J. (1975). “Statistical analysis of extreme wind loads.” Technical Note No. 868, National Bureau of Standards, Washington, DC. Skowronski, W. (1990). "Load capacity and deflection of fire-resistant steel beams." Fire Technology, 310-328. Stuckes, A. D., Tinker, J. A., and Simpson, A. (1986). “Effect of moisture on the thermal conductivity of lightweight aggregate concrete.” Building Services Engineering Research and Technology, 7 (1), 27-32. Sultan, M. A. (1996). “A model for predicting heat transfer through noninsulated unloaded steel-stud gypsum board wall assemblies exposed to fire.” Fire Technology, 32(3), 239—259. Thomas, G. (2002). “Thermal properties of gypsum plasterboard at high temperatures.” Fire and Materials, 26(1), 37-45. Thor, J. (1973). “Deformations and critical loads of steel beams under fire exposure conditions.” Report No. DI6:I973, National Swedish Building Research Summaries, Stockholm. . Tsantaridis, L. D., Ostman, B. A-L., Konig, J. (1999). “Short communication: fire protection of wood by different gypsum plasterboards.” Fire and Materials, 23(1), 45-48. Underwriters Laboratories Inc. (UL). (2004). Fire tests of building construction and materials, 13th edition. Standrad No. UL 263, Northbrook, IL. Vila Real, P. M. M., Piloto, P. A. G., and Franssen, J. M. (2003). “ A new proposal of a simple model for the lateral-torsional buckling of unrestrained steel I-beams in case of fire: experimental and numerical validation,” Journal of Constructional Steel Research, 59, 179-199. Vila Real, P. M. M., Lopes, N., Simoes da Silva, L., and Franssen, J. M. (2004). “Lateral-torsional buckling of unrestrained beams under fire conditions: improvement of EC3 proposal,” Computers and Structures, 82, 1737-1744. Wainman, D. E. (1992). “The effect of load ratio on the limiting temperatures observed during two BS476: part 21 fire resistance carried out on 254x146mmx43kg/m universal beams protected with thick Vicuclad insulation board.” Report No. Sl/HED/R/SII 99/18/92/C, British Steel Technical, Swinden Laboratories, Moorgate, Rotherham. Wainman, D. E. and Kirby, B. R. (1987). “Compendium of UK. standard fire test data on unprotected structural steel.” Report No. RS/RSC/SlO328/1/87/B, British Steel Technical, Swinden Laboratories, Moorgate, Rotherham. 225 111. 112. 113. 114. .115. 116. Wang, Y. C. and Yin, Y. Z. (2006). "A simplified analysis of caternary action in steel beams in fire and implications on fire resistant design." Steel and Composite Structures, 6(5), 367-386. Whiting, D., Litvin, A., and Goodwin, S. E. (1978). “Specific heat of selected concretes.” AC1 Journal, 75 (7), 299-305. Williams B.K. (2004), “Fire performace of FRP-strengthened reinforced concrete flexural memebers”, PhD Thesis, Queen’s University, Kingston, Ontario, Canada. Wullschleger, L., and Wakili, K. M. (2008). “Numerical parameter study of the thermal behaviour of a gypsum plaster board at fire temperatures.” Fire and Materials, 32(2), 103-119. Yii, H.W. (2000). Effect of Surface Area and Thickness on Fire Loads. Fire Engineering Research Report 2000/13, University of Canterbury, Christchurch, New Zealand. Yin, Y. Z. and Wang, Y. C. (2004). "A numerical study of large deflection behavior of restrained steel beams at elevated temperatures." Journal of Constructional Steel Research, 60, 1029-1047. 226