reliability based durability assessment of gfrp bars … · reliability-based durability assessment...

122
RELIABILITY-BASED DURABILITY ASSESSMENT OF GFRP BARS FOR REINFORCED CONCRETE Nicole D. Jackson Thesis submitted to the Faculty of the Virginia Polytechnic Institute and State University in partial fulfillment of the requirements for the degree of Master of Science in Engineering Mechanics Scott W. Case, Chair John J. Lesko Carin L. Roberts-Wollmann 12 November 2007 Blacksburg, Virginia Keywords: Glass fiber reinforced polymer (GFRP), Monte Carlo, Concrete, Reliability Copyright 2007 by Nicole D. Jackson

Upload: others

Post on 19-Mar-2020

7 views

Category:

Documents


1 download

TRANSCRIPT

RELIABILITY-BASED DURABILITY ASSESSMENT OF GFRP BARS FOR REINFORCED CONCRETE

Nicole D. Jackson

Thesis submitted to the Faculty of the Virginia Polytechnic Institute and State University

in partial fulfillment of the requirements for the degree of

Master of Science in

Engineering Mechanics

Scott W. Case, Chair John J. Lesko

Carin L. Roberts-Wollmann

12 November 2007 Blacksburg, Virginia

Keywords: Glass fiber reinforced polymer (GFRP), Monte Carlo, Concrete, Reliability

Copyright 2007 by Nicole D. Jackson

RELIABILITY-BASED DURABILITY ASSESSMENT OF GFRP BARS FOR

REINFORCED CONCRETE

Nicole D. Jackson

ABSTRACT

The American Concrete Institute (ACI) has developed guidelines for the design of fiber

reinforced polymer (FRP) reinforced concrete structures. Current guidelines require the application

of environmental and flexural strength reduction factors, which have minimal experimental

validation. Our goal in this research is the development of a Monte Carlo simulation to assess the

durability of glass fiber reinforced polymer (GFRP) reinforced concrete designed for flexure. The

results of this simulation can be used to determine appropriate flexural strength reduction factors.

Prior to conducting the simulation, long-term GFRP tensile strength values needed to be

ascertained. Existing FRP tensile strength models are limited to short-term predictions. This study

successfully developed a power law based-FRP tensile strength retention model using currently

available tensile strength data for GFRP exposed to variable temperatures and relative humidity.

GFRP tensile strength retention results are projected at 0, 1, 3, 10, 30, and 60-year intervals. The

Monte Carlo simulation technique is then used to assess the influence beam geometry, concrete

strength, fractions of balanced reinforcement ratio, reinforcing bar tensile strength, and

environmental reduction factors on the flexural capacity of GFRP reinforced concrete beams.

Reliability analysis was successfully used to determine an environmental reduction factor of

0.5 for concrete exposed to earth and weather. For simulations with higher GFRP bar tensile

strength as well as larger beam geometry and fractions of the balanced reinforcement ratio, larger

moment capacities were produced. A strength reduction factor of approximately 0.8 is calculated for

all fractions of balanced reinforcement ratio. The inclusion of more long-term moisture data for

GFRP is necessary to develop a more cohesive tensile strength retention model. It is also

recommended that longer life cycles of the GFRP reinforced concrete beams be simulated.

This research was conducted thanks to support from the National Science Foundation Division of Graduate Education’s Interdisciplinary Graduate Education Research and Traineeship (Award # DGE-0114342) Note: The opinions expressed herein are the views of the authors and should not be interpreted as the views of the National Science Foundation.

iii

ACKNOWLEDGEMENTS

I would like to thank my advisor and committee chairperson, Dr. Scott Case, for his

guidance and advice. Without his assistance, much of the thesis would not be possible. I would also

like to thank Dr. Carin Roberts-Wollmann and Dr. Jack Lesko for serving on my thesis committee.

Much of my time would not have transpired as smoothly without the endless help of Beverly

Williams and Joyce Smith. Many thanks go out to other the students in the Materials Response

Group. A special thanks goes out to Dr. Christine Fiori and Dr. John Popovics, who were my

undergraduate mentors and opened my eyes to the wonderful world of research. I wouldn’t be here

without either one of you.

iv

TABLE OF CONTENTS

LIST OF FIGURES..........................................................................................................................................................VI LIST OF TABLES.........................................................................................................................................................VIII CHAPTER 1 INTRODUCTION...................................................................................................................................... 1

1.1 BACKGROUND ............................................................................................................................................................ 1 1.2 OBJECTIVES ................................................................................................................................................................ 2

1.2.1 Primary Goal ..................................................................................................................................................... 2 1.2.2 Secondary Goals................................................................................................................................................ 2

CHAPTER 2 LITERATURE REVIEW ......................................................................................................................... 3 2.1 GENERAL .................................................................................................................................................................... 3 2.2 APPLICATIONS FOR CIVIL INFRASTRUCTURE............................................................................................................. 3 2.3 LIFE CYCLE COSTS OF FRP BRIDGE DECKS ............................................................................................................... 4 2.4 DURABILITY OF GFRP ............................................................................................................................................... 5 2.5 DEGRADATION MECHANISMS .................................................................................................................................... 6

2.5.1 Moisture degradation........................................................................................................................................ 6 2.5.2 Hygrothermal effects ......................................................................................................................................... 8 2.5.3 Alkali attack ....................................................................................................................................................... 9

2.6 DESIGN CODE PROVISIONS ........................................................................................................................................ 9 2.6.1 Flexural strength of GFRP ............................................................................................................................... 9 2.6.2 Environmental reduction factors.................................................................................................................... 11 2.6.3 Application of load and resistance factor design (LRFD) to FRP............................................................... 14

2.7 FRP TENSILE STRENGTH PREDICTIONS .................................................................................................................. 14 2.7.1 Regression-based tensile strength prediction................................................................................................ 15 2.7.2 Diffusion based tensile strength prediction ................................................................................................... 15 2.7.3 Litherland Method........................................................................................................................................... 16

2.8 EXPERIMENTAL STRENGTH DEGRADATION DATA OF GFRP BARS......................................................................... 16 2.8.1 Experimental Procedure ................................................................................................................................. 17

2.8.1.1 Conditioning of tensile test specimens .......................................................................................................................17 2.8.1.2 Tensile Testing Procedure...........................................................................................................................................18 2.8.1.3 Moisture absorption experiments................................................................................................................................20

2.8.2 Experimental Results....................................................................................................................................... 21 2.8.2.1 Tensile test results........................................................................................................................................................21 2.8.2.2 Moisture absorption results .........................................................................................................................................25

2.9 CONCLUSION ............................................................................................................................................................ 26 CHAPTER 3 COMPUTATIONAL PROCEDURE................................................................................................... 28

3.1 TENSILE STRENGTH RETENTION MODELS .............................................................................................................. 28 3.1.1 Existing Tensile Strength Models ................................................................................................................... 28

3.1.1.1 Durability prediction via short-term accelerated aging .............................................................................................28 3.1.1.2 Deterioration prediction from alkali attack ................................................................................................................34

3.1.2 Modeling existing tensile data ........................................................................................................................ 38 3.1.2.1 Tensile strength with variable conditioning temperatures ........................................................................................39 3.1.2.2 Tensile strength with variable relative humidity conditions .....................................................................................43 3.1.2.3 Tensile strength with combined relative humidity and temperature effects ............................................................49

3.2 MONTE CARLO SIMULATIONS ................................................................................................................................. 55 3.2.1 Random number generation ........................................................................................................................... 56 3.2.2 Parameterized elements .................................................................................................................................. 56

3.2.2.1 GFRP tensile strength..................................................................................................................................................56 3.2.2.2 Concrete Strength ........................................................................................................................................................58 3.2.2.3 Concrete beam and slab dimensions...........................................................................................................................59 3.2.2.4 Summary of parameterized elements .........................................................................................................................60

3.2.3 Flexural design of FRP reinforced concrete beams ..................................................................................... 61

v

3.2.3.1 Input values ..................................................................................................................................................................61 3.2.3.2 GFRP reinforcement....................................................................................................................................................62 3.2.3.3 Strain-compatibility analysis.......................................................................................................................................65 3.2.3.4 Nominal flexural strength............................................................................................................................................66

3.3 RELIABILITY ASSESSMENT....................................................................................................................................... 67 3.3.1 Weibull distribution models ............................................................................................................................ 68

3.3.1.1 Background on the Weibull distribution ....................................................................................................................68 3.3.1.2 Methods for estimating parameters ............................................................................................................................70 3.3.1.3 Development of confidence intervals .........................................................................................................................71

3.3.2 Determination of environmental reduction factors ....................................................................................... 76 3.3.2.1 Monte Carlo simulations for bar strength ..................................................................................................................76 3.3.2.2 Computational procedure ............................................................................................................................................77

3.3.3 Determination of strength reduction factors ................................................................................................. 79 CHAPTER 4 RESULTS .................................................................................................................................................. 81

4.1 INTRODUCTION......................................................................................................................................................... 81 4.2 ENVIRONMENTAL REDUCTION FACTORS ................................................................................................................ 81 4.3 MONTE CARLO SIMULATION RESULTS .................................................................................................................... 85

4.3.1.1 Effect of environmental reduction factors..................................................................................................................86 4.3.1.2 Effect of beam geometry .............................................................................................................................................90 4.3.1.3 Effect of concrete strength ..........................................................................................................................................97

4.4 STRENGTH REDUCTION FACTORS........................................................................................................................... 100 CHAPTER 5 CONCLUSIONS .................................................................................................................................... 106

5.1 INTRODUCTION....................................................................................................................................................... 106 5.2 CONCLUSIONS ........................................................................................................................................................ 106

5.2.1 Environmental reduction factors.................................................................................................................. 106 5.2.2 Monte Carlo simulation parameters ............................................................................................................ 106 5.2.3 Strength reduction factors ............................................................................................................................ 107

5.3 RECOMMENDATIONS FOR FUTURE WORK.............................................................................................................. 107 REFERENCES................................................................................................................................................................ 109

vi

LIST OF FIGURES

Figure 2-1: Primary and secondary effects of moisture absorption on composite materials. ...................................... 8 Figure 2-2: Side and full view schematics of the conditioning tank used by Bhise for short-term immersion testing of GFRP bars. ........................................................................................................................................................................ 18 Figure 2-3: Layout of the specimen ready for tensile test. Anchors are used to provide better gripping in the UTM and are applied after the specimen is removed from the solution................................................................................... 19 Figure 3-1: Tensile strength retention of GFRP1 bars exposed to Solution 1 at 20, 40, and 60°C [20]..................... 30 Figure 3-2: Tensile strength retention versus exposure time for GFRP1 bars [20]...................................................... 33 Figure 3-3: Tensile strength retention versus long-term exposure time for GFRP1 bars............................................. 34 Figure 3-4: GFRP tensile strength results for specimens subject to accelerated aging at 40 C with a 1.0 mol/l aqueous NaOH solution. [21]............................................................................................................................................ 35 Figure 3-5: Experimental and projected tensile strength results for GFRP bars.......................................................... 37 Figure 3-6: Predicted tensile strength retention for GFRP based on a diffusion model for strength retention. ......... 38 Figure 3-7: Temperature-based time shift factors for GFRP versus exposure temperature. Note that the reference temperature is 30 °C........................................................................................................................................................... 40 Figure 3-8: Tensile strength retention of GFRP bars subject to variable temperatures after the temperature-based shift factors have been applied. ......................................................................................................................................... 41 Figure 3-9: Tensile strength retention versus shifted time for GFRP bars subject to variable temperatures. The power-law model for GFRP tensile strength retention is overlaid. ................................................................................ 42 Figure 3-10: Effect of a variable moisture absorbing condition on the tensile strength of GFRP plates as a function of exposure time. [24] ........................................................................................................................................................ 45 Figure 3-11: Relative humidity based time shift factors for GFRP versus exposure relative humidity. Note that the reference relative humidity is 30%.................................................................................................................................... 46 Figure 3-12: Tensile strength retention of GFRP plates subject to variable relative humidity after the relative humidity-based shift factors have been applied................................................................................................................ 47 Figure 3-13: Tensile strength retention versus shifted time for GFRP plates subject to variable relative humidity. The power-law model for the GFRP tensile strength retention is overlaid. ................................................................... 48 Figure 3-14: Natural logarithm of temperature-based shift factors versus the inverse of temperature. Note that the reference temperature is 30 °C.......................................................................................................................................... 51 Figure 3-15: Natural logarithm of relative humidity-based shift factors versus difference in relative humidity. Note that the reference relative humidity is 30%. ..................................................................................................................... 52 Figure 3-16: Tensile strength retention versus shifted time for GFRP composites with superimposed temperature and relative humidity effects. The power-law model for the GFRP tensile strength retention is overlaid................... 54 Figure 3-17: Concrete beam with dimensions.................................................................................................................. 59 Figure 3-18: The Weibull probability density function [29] ........................................................................................... 69 Figure 4-1: Comparison of Weibull-based moment capacities for Member A using projected 0 and 60 year GFRP tensile strength values ........................................................................................................................................................ 82 Figure 4-2: Comparison of Weibull-based moment capacities for Member A using both 0 year and 60-year projected GFRP tensile strength values. The 0 year revised and ACI moment capacities have incorporated environmental reduction factors of 0.5 and 0.7, respectively, to the 0-year GFRP tensile strength. ........................... 83 Figure 4-3: Environmental reduction factors for Member A designed with 4000 psi concrete as a function of time. 85

vii

Figure 4-4: Weibull-based nominal flexural strength as a function of environmental reduction factor for Member A designed with 4000 psi concrete at balanced reinforcement ratio conditions. .............................................................. 87 Figure 4-5: Percentage of simulations occurring below the nominal flexural strength for GFRP reinforced concrete members as a function of the environmental reduction factor for Member A designed with 4000 psi concrete and balanced reinforcement ratio conditions. ......................................................................................................................... 88 Figure 4-6: Percentage of failures out of 10,00 simulations due to FRP bar rupture for Member A designed with 4000 psi concrete and balanced reinforcement ratio conditions. ................................................................................... 89 Figure 4-7: Total GFRP reinforcing bar area for Member A with respect to varying concrete strength and fractions of balanced reinforcement ratio. ....................................................................................................................................... 91 Figure 4-8: Total GFRP reinforcing bar area for Member C with respect to varying concrete strength and fraction of balanced reinforcement ratio. ....................................................................................................................................... 92 Figure 4-9: Nominal flexural strength for GFRP reinforced concrete members with respect to total GFRP reinforcing bar area for Members A and C. The illustrations shown are used to draw contrast to different beam geometries of the different members.................................................................................................................................. 93 Figure 4-10: Nominal flexural strength for GFRP reinforced concrete Member A with respect to environmental reduction factor and fraction of balanced reinforcement ratio. The member was designed with 4000 psi concrete. . 94 Figure 4-11: Nominal flexural strength for GFRP reinforced concrete member C with respect to environmental reduction factor and fraction of balanced reinforcement ratio. The member was designed with 4000 psi concrete. . 95 Figure 4-12: Percentage of simulation failures due to FRP bar rupture for Member A designed with 4000 psi concrete with respect to fractions of balanced reinforcement ration and environmental reduction factor. All environmental reduction factors were subject to 100% FRP bar rupture failure except for factors of 0.8 and 1.0. . 96 Figure 4-13: Percentage of simulation failures due to FRP bar rupture for Member C design with 4000 psi concrete and with respect to fractions of balanced reinforcement ration and environmental reduction factor. All environmental reduction factors were subject to 100% FRP bar rupture failure except for the factors of 0.8 and 1.0............................................................................................................................................................................................... 97 Figure 4-14: Nominal flexural strength for GFRP reinforced concrete Member A with respect to fractions of balanced reinforcement ratio and concrete strength. ...................................................................................................... 98 Figure 4-15: Percentage of simulation failures due to bar rupture with respect to fraction of balanced reinforcement ratio for Member A designed with 4000, 5000, and 6000 psi concrete. Values of 100% indicate all 10,000 simulations experienced this failure mode........................................................................................................................ 99 Figure 4-16: ACI-based flexural strength reduction factors with respect to fractions of the balanced reinforcement ratio. 1.5 and 1.6 times the balanced reinforcement ratio are also presented to verify the strength reduction factor of 0.65. Neither values were directly employed in the Monte Carlo simulations. ............................................................ 101 Figure 4-17: Flexural strength reduction factors based on Ellingwood with β = 3.5 for Member A designed with 4000 psi concrete. Similar results were reported throughout the simulations. ............................................................ 103 Figure 4-18: Flexural strength reduction factors based on Ellingwood with β = 4.0 for Member A designed with 4000 psi concrete. As with the β = 3.5, similar results were recorded amongst all simulations. ............................... 104 Figure 4-19: Comparison of flexural strength reduction factors for all methods under consideration. The current values provided by ACI yield the most conservative design estimates for the moment capacity of GFRP reinforced concrete beams. ................................................................................................................................................................ 105

viii

LIST OF TABLES

Table 2-1: Life-cycle costs for traditional steel reinforced and FRP reinforced concrete bridge decks. The FRP premium represents the difference in cost estimates between the two bridges. Note that costs in parentheses indicate negative dollar amounts. ...................................................................................................................................................... 4 Table 2-2: Partial safety factors proposed for fiber-reinforced polymer reinforced concrete structures for usage in EUROCODE 8 ................................................................................................................................................................... 11 Table 2-3: Environmental reduction factor for various fibers and exposure conditions per ACI 440.1R-06 ............. 12 Table 2-4: Comparison of the existing code specified and experimentally determined environmental reduction factors for carbon, glass, and aramid fiber reinforced concrete structures. The highest, median, and lowest values are reported. ....................................................................................................................................................................... 13 Table 2-5: Reduction factors for GFRP bars suggested by the American Concrete Institute, Canadian Standards Association, and the Japanese Society of Civil Engineers. ............................................................................................ 14 Table 2-6: Tensile strength and modulus of elasticity for plain #3 GFRP bars. Values based on the nominal and measured cross-sectional areas of the bars are reported. ............................................................................................... 19 Table 2-7: Tensile strength and modulus of elasticity for incased but unconditioned bars #3 GFRP bars. Values based on the nominal and measured cross-sectional areas of the bars are reported.................................................... 20 Table 2-8: Tensile strength of #3 GFRP bars subject to 30, 45, and 57 °C temperatures for 10, 30, 60, 90, and 180 days. ..................................................................................................................................................................................... 22 Table 2-9: Average tensile strength and coefficient of variation of #3 GFRP bars subject to 30, 45, and 57°C temperatures for 10, 30, 60, 90, and 180 days. ................................................................................................................ 23 Table 2-10: Modulus of elasticity for #3 GFRP bars subject to 30, 45, and 57°C temperatures for 10, 30, 60, 90, and 180 days............................................................................................................................................................................... 24 Table 2-11: Average modulus of elasticity and coefficient of variation of #3 GFRP bars subject to 30, 45, and 57°C temperatures for 10, 30, 60, 90, and 180 days. ................................................................................................................ 25 Table 2-12: Moisture absorption data based on weight gain percentages for #3 GFRP specimens immersed in limewater solution. ............................................................................................................................................................. 25 Table 2-13: Maximum moisture and diffusion for #3 GFRP bars immersed in a limewater solution at 30, 45, 57°C............................................................................................................................................................................................... 26 Table 2-14: Arrhenius analysis results based on a Fickian curve fit with respect to immersion temperatures of 30, 45, and 57°C. ...................................................................................................................................................................... 26 Table 3-1: Chemical compositions of simulated concrete pore solutions. Solution 1 has a pH of 13.6, and Solution 2 has a pH value of 12.7, which closely resembles the pore environment found in typical high performance concretes. .............................................................................................................................................................................................. 29 Table 3-2: Short-term accelerated aging test matrix for GFRP bars immersed in concrete pore solutions. [20] ...... 29 Table 3-3: Coefficients of regression equations for GFRP bar tensile strength retention immersed in both concrete pore solutions 1 and 2. [20]............................................................................................................................................... 31 Table 3-4: Coefficient of Regression Equations for Arrhenius Plots immersed in both concrete pore solutions 1 and 2 [20] ................................................................................................................................................................................... 32 Table 3-5: Values for Acceleration Factors [20] ............................................................................................................. 32 Table 3-6: Temperature-based shift factors for GFRP tensile strength retention based on a reference temperature of 30°C..................................................................................................................................................................................... 39 Table 3-7: Predicted long-term tensile strength retention based on variable temperature data .................................. 43

ix

Table 3-8: GFRP tensile strength retention results for variable humidity ..................................................................... 44 Table 3-9: Relative humidity-based shift factors for GFRP tensile strength retention based on a reference relative humidity of 30%. ................................................................................................................................................................. 46 Table 3-10: Predicted long-term tensile strength retention based on variable relative humidity data ........................ 49 Table 3-11: Predicted long-term tensile strength retention for GFRP based on superimposed relative humidity and temperature data................................................................................................................................................................. 55 Table 3-12: Physical properties of #3 Aslan 100 GFRP Rebar by Hughes Brothers .................................................... 57 Table 3-13: Statistical and mechanical properties of GFRP bars ................................................................................. 58 Table 3-14: Statistical parameters of ordinary ready-mix concrete [26] ...................................................................... 59 Table 3-15: Statistical parameters of dimensional variables .......................................................................................... 60 Table 3-16: Statistical properties of variables involved in this study ............................................................................. 61 Table 3-17: Member dimensions ....................................................................................................................................... 62 Table 3-18: Taxonomy for Weibull Models. ..................................................................................................................... 70

1

CHAPTER 1 Introduction

1.1 Background

Steel bars are the predominant form of reinforcement for concrete structures. The natural

porosity of concrete and exposure to harsh environments causes steel to be highly susceptible to

corrosion. The limited long-term durability and high cost of maintenance of steel rebar has lead to

increased interest in other reinforcing materials, particularly fiber reinforced polymers (FRP).

First-generation FRP materials where characterized by poor stiffness and tensile properties.

Early FRP reinforced concrete design guidelines lead to conservative estimates of environmental and

strength reduction factors. These factors have limited empirical basis and are generally selected by

committee consensus. A greater understanding of durability mechanisms and improved

manufacturing techniques has lead to a significantly improved material. Unfortunately, design codes

have not been readily updated to reflect the enhanced mechanical properties of FRP.

The long-term tensile strength retention of FRP remains elusive. Current retention models

either consider degradation due to moisture diffusion or temperature. The models rely on the

collection of short-term accelerated aging testing data. However, there is little definitive research

verifying long-term strength based on short-term results. Furthermore, no tensile strength retention

models exist in which both thermal and moisture effects are considered simultaneously.

The application of FRP as a reinforcing material in bridge decks is relatively new. A large-

scale analysis of the reliability of FRP as reinforcing in concrete slabs has not been performed. This

study will utilize various simulation and modeling techniques to perform a parametric analysis on the

flexural strength reliability of glass FRP (GFRP) internally reinforced concrete slabs.

2

1.2 Objectives

1.2.1 Primary Goal

The primary goal of this study is to utilize the Monte Carlo simulation technique to

determine the flexural strength reliability of GFRP internal reinforced concrete slabs. Previously

obtained strength degradation and moisture data for No. 3 GFRP bars will be used to validate the

theoretical results. The effects of slab geometry, concrete strength, reinforcement ratios, and bar

strength on flexural strength will be considered.

1.2.2 Secondary Goals

A secondary goal of this study is to examine the effects of environmental exposure on the

strength of GFRP reinforcing models. Existing experimental data will be incorporated into models

of GFRP strength degradation to describe the results. Combined effects of moisture and

temperature will be considered in the models.

An additional objective of this study is to review the environmental and strength reduction

factors currently prescribed by the American Concrete Institute (ACI). The environmental factors

have been determined by committee consensus with little experimental or theoretical validation. The

strength factors are based on steel reinforcement for reinforced concrete theory with no

consideration to long-term strength loss. Some recommendations will be made on appropriate

environmental and strength reduction factors.

3

CHAPTER 2 Literature Review

2.1 General

The literature to date has shown that FRP can be a viable material for usage in civil

infrastructure applications. The literature review presented entails a brief discussion of some of

recent applications of FRP, provides a review of the life cycle cost and, assesses the durability,

reliability, and flexural strength of FRP reinforced concrete. Currently available literature regarding

the primary degradation mechanisms of alkali attack and moisture are also reviewed. A brief

overview of current design provisions, especially the application of environmental reduction factors,

is presented. Experimental strength degradation data of GFRP is reviewed and will be utilized in

future chapters.

2.2 Applications for civil infrastructure

Van Den Einde et al. provide an overview various civil engineering applications with extensive

usage of FRP material systems [1]. FRP has been successfully used to repair constructed facilities,

seismic column retrofitting, and structural replacement using new bridge systems. Structural wall

overlays can be accomplished by using shear, flexural, and slab strengthening. Both carbon fibers as

well as carbon/glass hybrid tube systems have been used in various civil engineering applications.

FRP materials were cited for promise as a lightweight structural material.

Uomoto et al. conducted a study utilizing FRP as a reinforcing material for concrete [2].

Applications of FRP rods in the construction of new reinforced concrete (RC) and pre-cast (PC)

structures are discussed. The design and construction methodologies of fiber-reinforced sheets are

mentioned as well. Some general applications of FRP rods are marine concrete structures, pre-cast

bridges, transport facilities, tunnel linings, building structures.

4

2.3 Life cycle costs of FRP bridge decks

Nystrom et al. conducted a cost feasibility study on using FRP instead of convention steel

reinforcing bars for bridge construction [3]. The United States has a quickly aging transportation

infrastructure; 29% of all bridges are classified as either “structurally deficient” or “functionally

obsolete.” Cost effective strategies must be implemented to maintain current transportation systems.

FRP provides a longer life cycle than reinforced concrete bridges (60 years versus 40 years). As

demonstrated in Table 2-1, longer life cycle and reduced disposal costs do not mitigate the cost

premium of an FRP bridge. FRP bridge decks do provide a cost effective alternative for standard

short span bridges.

Table 2-1: Life-cycle costs for traditional steel reinforced and FRP reinforced concrete

bridge decks. The FRP premium represents the difference in cost estimates between the

two bridges. Note that costs in parentheses indicate negative dollar amounts.*

Description RC bridges FRP bridge FRP Premium

Construction costs $430/m2 $740/m2 $310/m2

Impact of disposal $45/m2 $9/m2 ($36/m2)

Total cost including

disposal $475/m2 $749/m2 $274/m2

Impact of

replacements $119/m2 $67/m2 ($52/m2)

Total cost including

disposal and longer life $594/m2 $816/m2 $222/m2

Sahirman et al. assess the economic feasibility of FRP bridge decks [4]. Future costs of

construction are estimated. Life cycle costs analysis for several in-place FRP bridges is presented.

* Nystrom, Halvard E., et al., Financial Viability of Fiber-Reinforced Polymer (FRP) Bridges. Journal of Management in Engineering, 2003. 19(1): p. 2-8. Reprinted with permission from ASCE. This material may be downloaded for personal use only. Any other use requires prior permission of the American Society of Civil Engineers. This material may be found at http://cedb.asce.org/cgi/WWWdisplay.cgi?0300089

5

The competitiveness of FRP bridge decks versus steel reinforced concrete decks is addressed.

Overall cost comparisons between the two are incomplete due to insufficient cost data for FRP

bridge decks.

2.4 Durability of GFRP

Nkurunziza et al. conducted a comprehensive literature review regarding the durability of

GFRP bars [5]. Various methods for evaluating long-term performance of GFRP are reviewed.

Advancements in material durability due to improved manufacturing techniques were noted. Initially

poor mechanical properties lead to conservative strength reduction factors. It is recommended that

factors be revised to reflect the improved properties of FRP. Significant progress in understanding

the mechanically behavior has been made. However, there is still limited long-term in-situ data on

FRP. Accelerated aging techniques, which mostly rely on using elevated temperatures, have been

employed to fill this gap.

Mukherjee and Arwikar performed a two – part study of environmental effects on GFRP bars

[6]. The second portion of the study utilized various microscopy techniques to analyze the

microstructural properties of conditioned and unconditioned GFRP specimen. Reinforcing bars

were conditioned at 60°C for 3, 6, and 12-month intervals. The scanning electron microscopy (SEM)

showed air bubbles and microcracks on unconditioned bars. SEM images also revealed both

scattered and local damage zones in conditioned bars. The high pH, alkali environment naturally

present in concrete lead to excessive calcium and silica in the GFRP bars. The authors have

recommended that the environmental reduction factors be based on “severity of alkali attack” when

used in concrete

Almusallam and Al-Salloum compared the durability of GFRP in concrete subject to three

different environmental conditions while undergoing sustained loading [7]. Tap water and seawater

6

of both wet/dry and continuous submersion were considered. A 16-month exposure period of

sustained loading had a profound effect on GFRP bars. Tensile strength losses of 28.2 – 33.2% were

reported for the three exposure conditions. Sustained stresses should be considered when assessing

the durability of GFRP, especially in highly alkaline environments.

2.5 Degradation mechanisms

Karbhari et al. have identified seven different environmental conditions that have a

profound impact on durability for FRP: moisture/solution, alkali, thermal, creep and relaxation,

fatigue, ultraviolet, and fire [8]. Significant research has been conducted regarding the effects of

moisture and alkalinity for FRP and concrete composite applications. For the overall purposes of

this thesis, only moisture and alkali effects will be discussed.

2.5.1 Moisture degradat ion

Moisture diffusion occurs in any polymeric material, including FRP [8]. The primary

moisture degradation mechanism for GFRP is caused by the loss of ions in the fiber. Using materials

that have been fully cured prior to installation and determining an appropriate resin-rich region

thickness for the composite are seen as ways of decreasing the susceptibility to moisture diffusion.

Nishizaki and Meiarashi conducted a study comparing the effects of moisture due to

immersion and moisture from humidity [9]. The study included up to 557 days of immersion

exposure. The pH of the water was not altered to mimic a concrete environment. Temperature

ranges of 40 – 60°C were used for both conditions. Measurements of weight change, infrared

spectroscopy, and bending tests were performed on all specimens. High weight reduction rates and

poor bending strength are believed to be due to separation between the fiber and resin. Both water

7

and moisture levels are can contribute greatly to long-term deterioration of materials, especially

when exposed to water.

Pauchard et al. studied the occurrence of fiber failure for water-aged GFRP [10]. A stress

corrosion cracking model was used to examine damage accumulation in the bars. Higher testing

temperatures lead to the activation of the delayed fracture process. Thus, only lower temperature

results were subsequently used. Microscopic techniques were used to validate the stress corrosion

and macroscopic cracks. Fibers embedded in a matrix were subject to slower crack growth rates than

fibers exposed to the air directly.

Micelli and Nanni compared the performance of carbon and glass FRP bars [11]. Test

specimens were subject to freeze-thaw, high temperature, high relative humidity cycles as well as UV

radiation. Other specimens were used for an alkali-based accelerated aging study. Ultimate tensile

stress, modulus of elasticity, and ultimate strain were reported for all specimens. Some of the GFRP

bars exposed to the alkaline solution demonstrated 30 and 40% strength reductions after 21 and 42

exposure days, respectively. Environmental cycling of the specimens did not adversely affect their

tensile properties. Degradation due to alkali attack was validated by the SEM images.

Nkurunziza et al. describe the chemistry behind the moisture degradation mechanism[5]. A

pattern of moisture absorption for a composite material was developed, which is shown in Figure

2-1. Moisture can lead to a reduction in the glass transition temperature. Stress corrosion cracking

occurs at the fiber level.

8

Figure 2-1: Primary and secondary effects of moisture absorption on composite materials. *

2.5.2 Hygrothe rmal e f f e c ts

Vijay et al. studied the effect of different hygrothermal conditions on GFRP bars [12].

Conditioning occurred via tap water, salt water, or an alkaline solution. Temperature ranges included

room temperature, 150°F, and freeze-thaw. A total of 162 specimens were tested. Moisture

absorption and tensile tests were performed. Moisture data presented was based on 325 conditioning

days. Tensile strength gains of up to 8.7% occurred for salt water specimens. Conversely, tensile

strength losses of up to 15.6% and 16.9% for freeze-thaw and alkaline conditioning, respectively,

were reported. Stiffness losses were recorded for all conditioning environments. The losses were

greatest for bars subject to alkaline immersion in freeze-thaw temperatures. The alkaline

environment, regardless of temperature, was experimentally shown to be the most degrading to the

GFRP bars.

Mula et al. considered the impact of a thermal gradient on water absorption [13]. GFRP

composites were transferred between sub-zero and elevated temperatures to create a thermal

gradient. Freezing temperatures of -20°C and elevated temperatures of 60°C were used. Specimens

were tested for moisture absorption and subject to short beam shear tests. Longer exposure times

* Nkurunziza, Gilbert, et al., Durability of GFRP Bars: A Critical Review of the Literature. Progress in Structural Engineering and Materials, 2005. 7(4): p. 194 - 209. Copyright John Wiley & Sons Limited. Reproduced with permission.

9

lead to reductions in the inter-laminar shear stresses. However, an initial increase in the inter-laminar

shear stresses appear during the initial conditioning phase. The state of water, frozen or liquid, did

have some affect on inter-laminar shear stresses as well.

2.5.3 Alkal i attack

Bhise performed a comprehensive analysis of the strength degradation of GFRP bars [14].

The bars were set in cement mortar and subjected to an alkaline solution, which simulates the pH of

concrete. A maximum exposure time of 300 days and three different water temperatures were used.

Strength degradation predictions were made using the time-Temperature-Superposition and Fickian

models. The Fickian model estimated a 45% strength reduction over a 50-year service life. The

experimental data is used in future parts of this study.

2.6 Design Code Provisions

Structural codes vary from municipality to municipality. The American Concrete Institute

(ACI) Committee 440 oversees the guidelines for structural concrete reinforced with FRP bars in the

United States [15]. EUROCODE and FIB provide guidelines for the European Union. The

Japanese Society of Civil Engineers (JSCE) oversees Japanese guidelines. In subsequent sections,

design provisions regarding flexural strength and environmental factors in these various codes are

discussed.

2.6.1 Flexural s t reng th of GFRP

When designing in accordance to ACI 440, the structural designer must give consideration to

the desired failure mode of the FRP reinforced concrete structure. FRP reinforced concrete may fail

either in flexure or shear. The focus of this study is on flexural failure. The three failure modes for

10

FRP reinforced concrete members in flexure are FRP bar rupture, concrete crushing, and

simultaneous rupture and crushing.

Nanni studied the flexural strength of concrete reinforced with FRP [16]. When compared

to similarly designed steel reinforced beams, FRP beams were able to produce a higher maximum

moment capacity. However, FRP does not have the same flexural rigidity and ductility as its steel

counterparts. In a separate parametric study, structural designs utilizing the ultimate strength and

working stress methods are considered. The concrete strength and reinforcement ratios were varied

to obtain many theoretical values. Both aramid and glass FRP bars were considered. The theoretical

flexural strength of GFRP was less than AFRP due to the lower stiffness. Early recommendations

included using FRP for pre-stressed concrete elements or in combination with high-strength

concrete.

Pilakoutas et al. addressed some of the design issues prevalent when using FRP in a structure

[17]. All designs were based on EUROCODE 8. Several simulations utilizing all fiber types were

run. Partial safety factors, γFRP, or member safety factors are applied to obtain the desired failure

mode. A sample of proposed partial safety factors is shown in Table 2-2. The factors are similar to

the strength factors utilized by ACI, and do not implicitly incorporate environmental effects. The

preferred failure mode is concrete crushing to maximize the usage of the costlier FRP

reinforcement. Also, it was recommended that a minimum reinforcement level be set to reduce the

possibility of failure via bar rupture.

11

Table 2-2: Partial safety factors proposed for fiber-reinforced polymer reinforced concrete

structures for usage in EUROCODE 8 *

Parameter Material Partial safety factor, γFRP,

(short and long term)

Strength E-glass reinforced 3.6

Strength Aramid reinforced 2.2

Strength Carbon reinforced 1.8

Stiffness E-glass reinforced 1.8

Stiffness Aramid reinforced 1.1

Stiffness Carbon reinforced 1.1

2.6.2 Environmental re duc tion fac to rs

ACI provides design guidelines and has established values for the environmental reduction

factors FRP bars [15]. Table 2-3 displays these values for carbon, glass, and aramid fibers. The

environmental reduction factor only implicitly accounts for temperature and are conservative values.

The values have not been substantiated by research and were selected via committee consensus.

Regardless of the exposure condition, glass fibers are subject to an environmental reduction factor.

* Pilakoutas, Kypros, Kyriacos Neocleous, and Maurizio Guadagnini, Design Philosophy Issues of Fiber Reinforced Polymer Reinforced Concrete Structures. Journal of Composites for Construction, 2002. 6(3): p. 154 - 161. Reprinted with permission from ASCE. This material may be downloaded for personal use only. Any other use requires prior permission of the American Society of Civil Engineers. This material may be found at http://cedb.asce.org/cgi/WWWdisplay.cgi?0203656

12

Table 2-3: Environmental reduction factor for various fibers and exposure conditions per

ACI 440.1R-06 *

Exposure condition Fiber type Environmental reduction factor

CE

Carbon 1.0

Glass 0.8 Concrete not exposed to earth

and weather Aramid 0.9

Carbon 0.9

Glass 0.7 Concrete exposed to earth and

weather Aramid 0.8

Myers and Viswanath compared environmental reduction factors from around the world

[18]. The United States, Japan, Canada, Great Britain, Norway, and Europe were compared. Factors

are provided for glass (GFRP), aramid (AFRP), and carbon (CFRP) fibers. Although these countries’

values are not explicitly stated, highest, median, and lowest values are reported. A summary has been

reproduced in Table 2-4.

* American Concrete Institute Committee 440, ACI 440.1R-06: Guide for the Design and Construction of Structural Concrete Reinforced with FRP Bars. 2006, American Concrete Institute: Farmington Hills, MI. Reproduced with permission from the American Concrete Institute.

13

Table 2-4: Comparison of the existing code specified and experimentally determined

environmental reduction factors for carbon, glass, and aramid fiber reinforced concrete

structures. The highest, median, and lowest values are reported. *

Highest value used Median Lowest value used

Criteria

Type

of

fibers

used

Code

specified

Experi-

mentally

Determined

Code

specified

Experi-

mentally

Determined

Code

specified

Experi-

mentally

Determined

CFRP 1.00 1.00 0.88 0.90 0.60 0.67

GFRP 0.80 0.97 0.70 0.81 0.14 0.15

Reduction for

Environmental

Degradation AFRP 0.90 0.98 0.85 0.69 0.31 0.20

Nkurunziza et al. provided a literature review regarding the durability of GFRP [5].

American, Canadian, and Japanese guidelines all prescribe an environmental reduction factor of

some form. While the American and Japanese guidelines combine reductions taken for

environmental degradation and sustained loads, the Canadian system does not. Table 2-5 illustrates

the provisional differences.

* Myers, John J. and Thara Viswanath, A Worldwide Survey of Environmental Reduction Factors for Fiber-Reinforced Polymers (FRP). Structures 2006, 2006. Reprinted with permission from ASCE. This material may be downloaded for personal use only. Any other use requires prior permission of the American Society of Civil Engineers. This material may be found at http://cedb.asce.org/cgi/WWWdisplay.cgi?0203656

14

Table 2-5: Reduction factors for GFRP bars suggested by the American Concrete Institute,

Canadian Standards Association, and the Japanese Society of Civil Engineers. *

Code ACI 440.1R-03 CAN/CSA-S6-00 JSCE 1997

Reduction due to the

environmental

degradation

CE

‘Environmental

reduction coefficient’

0.70 – 0.80

φFRP

‘Strength factor’

0.75

1/γfm

‘Factor taking the

material into account’

0.77

Combined reduction

due to the

environment and

sustained load

0.70 – 0.80 0.60 – 0.75 0.77

2.6.3 Appl ica ti on of l oad and res is tance fac t or des ign (LRFD) to FRP

Ellingwood assessed the feasibility of implementing a LRFD approach to designing with

FRP composite materials [19]. Ultimate and serviceability limit states for load combinations are

reviewed. Structural members limit states in tension, compression, flexure and shear are also

discussed. The limit states can be derived from applying probability and reliability theories to

experimentally obtained material data. A greater implementation of LRFD and other performance-

based designs will not occur until structural design standards incorporate them.

2.7 FRP Tensile Strength Predictions

Long-term in-situ FRP data is not readily available. Short term accelerated aging has often

been extrapolated to make long-term mechanical property estimates. Accelerated aging has

previously utilized elevated temperatures in conjunction with alkaline solutions to mimic service life

* Nkurunziza, Gilbert, et al., Durability of GFRP Bars: A Critical Review of the Literature. Progress in Structural Engineering and Materials, 2005. 7(4): p. 194 - 209. Copyright John Wiley & Sons Limited. Reproduced with permission.

15

conditions. Other variations have also included applying a sustained stress on FRP bars while

subject to the aforementioned conditions [5]. Once the aging tests have concluded, models based

on either moisture absorption or tensile strength data have been generated to predict long-term

tensile strength. The following sections will highlight models that have incorporated either one of

these data sets.

2.7.1 Regress ion-based tens i l e s treng th pred ic t ion

Chen et al. compared the effect of alkaline solution concentration on GFRP bars during

accelerated aging [20]. Two different types of glass fibers were tested. Alkaline solutions were

developed to simulate normal strength and high performance concrete. Temperatures ranged from

40 – 60°C°, with room temperature being 20°C. Using regression analysis and developing an

Arrhenius relation, the tensile strength retention at a specified time and temperature is ascertained.

The model predicts a 50% strength reduction after a half-year of exposure time for bars at room

temperature in the simulated normal strength concrete. The tensile strength retention estimates

when using long-term time scales proved to be very poor.

2.7.2 Dif fus ion based tens i l e s treng th pred ic t ion

Katsuki and Uomoto examined the deterioration of FRP due to alkali attack [21].

Accelerated testing was conducted on GFRP, CFRP, and AFRP bars. The testing program lasted

120 days at a constant temperature of 40°C. The AFRP and CFRP bars were subject to a higher

alkaline concentration, 2 mol/L versus 1 mol/L for GFRP. Tensile tests were performed at 0, 7, 30,

90, and 120-day intervals. A diffusion-based model was then fit to the tensile strength data.

Although the model provides an excellent fit to the short-term data, it is insufficient for long-term

predictions. The carbon and aramid FRP bars performed significantly better than the glass bars.

16

Those types of fibers provided more resistance to the alkali attack than the glass fibers. It is

recommended that GFRP bars be constructed with a thicker layer of resin to provide more glass

fiber protection.

2.7.3 Litherland Method

Won et al. studied the effect of short term aging on GFRP dowels [22]. The dowels were

either round or elliptical in shape, and three different diameters were used. Testing conditions

included tap water immersion, an alkali environment, saline environment, snow-melting agent, and

cyclic freezing and thawing. The conditions were selected to provide laboratory simulations of year-

round Korean weather. Shear strength testing was conducted. The Litherland Method, as shown in

Eq. 2.1, relates temperature and diffusion to extrapolate natural time deterioration from accelerated

testing:

N

C= 0.098exp 0.558T[ ] 2.1

where N is the period of natural deterioration, C is the period of accelerated deterioration, and T is

the acceleration temperature (°C). The Litherland Method was used to derive time-temperature

relationships for the aging data. Shear strength losses were not nearly as great as tensile strength

losses due to the immersion conditions.

2.8 Experimental strength degradation data of GFRP bars

Bhise performed several short-term accelerated aging tests. Tensile strength and modulus of

elasticity data were obtained over a 180-day period. Moisture content data was available for an 80-

day period [14]. The moisture diffusion and strength reduction data are used to generate a strength

reduction model.

17

All experiments were conducted using Hughes Brothers No. 3 GFRP bars. Tensile test

specimens were 35 inches (889 mm) long. Moisture absorption specimens were 4 inches (102 mm)

long. The specimens were tested in accordance with the 2000 ACI 440 code. A Universal Testing

Machine (UTM) was used to perform the tensile tests. The entire program utilized 100 tensile test

and 66 moisture absorption specimens. The moisture content was determined in accordance with

ASTM D5229.

2.8.1 Experimen tal P roc edu re

2.8.1.1 Conditioning of tensile test specimens

Three tanks were used to condition the GFRP bars at 30, 45, and 57°C, respectively.

Temperature selection was in accordance with ACI 440. Water heaters of various wattages were used

to maintain water temperature. The middle 10 inches (254 mm) of each bar was encased in a cement

mortar paste. The paste was used to simulate the alkali environment of a concrete bridge deck. The

remaining segments of each bar were coated with epoxy paint. Figure 2-2 displays a schematic of the

test setup. The pH and temperature of the water were checked often.

18

Figure 2-2: Side and full view schematics of the conditioning tank used by Bhise for short-

term immersion testing of GFRP bars. *

Tank exposure times of 10, 30, 60, 90, 180, and 300 days were used. Five specimens were tested at

each interval. The 300-day tensile test results were not reported.

2.8.1.2 Tensile Testing Procedure

The UTM was used to perform the tensile tests. Prior to testing, an anchorage system was

placed on each bar. The system consisted of the conditioned GFRP bar with PVC anchors at each

end. The anchors were filled with epoxy resin, hardener and sand mixture. A 48-hour setting time

was used to cure the anchoring system prior to tensile testing. Figure 2-3 shows the test specimen

layout.

* Bhise, Vikrant, Strength Degradation of GFRP Bars, in Via Department of Civil and Environmental Engineering. 2002, Virginia Polytechnic Institute and State University: Blacksburg, VA, electronic thesis available at http://scholar.lib.vt.edu/theses/index.html

19

Figure 2-3: Layout of the specimen ready for tensile test. Anchors are used to provide better

gripping in the UTM and are applied after the specimen is removed from the solution.

The testing load rate range satisfied ACI code requirements. Failure loads were reported and,

ultimate tensile strengths were determined. Five unconditioned bars were also tested. The tensile

strength and modulus of elasticity of the plain bars are shown in Table 2-6. For comparison, the

tensile strength and modulus of encased but unconditioned bars are shown below in Table 2-7.

Table 2-6: Tensile strength and modulus of elasticity for plain #3 GFRP bars. Values based

on the nominal and measured cross-sectional areas of the bars are reported.

Tensile Strength (ksi) Modulus of Elasticity (ksi) Number

Nominal Areaa Measured Areab Nominal Areaa Measured Areab

1 100.2 84.3 6470 5450

2 N/A N/A N/A N/A

3 100.4 84.5 6550 5520

4 102 85.9 6530 5500

5 104.3 87.8 6680 5630

Average 101.8 85.6 6550 5520

Coefficient of Variation (%) 1.80 1.40 a – Strength calculated using the nominal area (0.11 in2) b – Strength calculated using the measured area (0.13 in2)

20

Table 2-7: Tensile strength and modulus of elasticity for incased but unconditioned bars #3

GFRP bars. Values based on the nominal and measured cross-sectional areas of the bars are

reported.

Tensile Strength (ksi) Modulus of Elasticity (ksi) Number

Nominal Areaa Measured Areab Nominal Areaa Measured Areab

1 81.6 68.7 6290 5300

2 72.9 61.4 6640 5590

3 93.3 78.5 6700 5640

4 81.3 68.5 6200 5220

5 104 87.5 N/A N/A

Average 86.6 72.9 6450 5430

Coefficient of Variation (%) 14.0 3.80 a – Strength calculated using the nominal area (0.11 in2) b – Strength calculated using the measured area (0.13 in2)

2.8.1.3 Moisture absorption experiments

Moisture absorption experiments were used to obtain weight gain measurements. The

specimens were 4 inches (102 mm). An epoxy coating was placed on the ends of each specimen.

The specimens were then submerged an alkaline solution at 30, 45, and 57°C. The testing program

lasted 80 days with samples being drawn 1, 2, 4 hours after initial immersion and 1,2, 5, 10, 30, 65,

and 80 days. Two specimens, a total of 66, were each weighed directly after the immersion and an

oven treatment. The percent weight gain for each specimen was reported using Eq. , in which Wx is

the wet weight at time x and, Wd is the oven dry weight.

%M =W

x!W

d

Wd

"100 2.2

21

2.8.2 Experimen tal Resu l ts

2.8.2.1 Tensile test results

The complete tensile experimental program consisted of 5 plain, 10 incased and

unconditioned and; 90 incased and conditioned specimens. The plain and encased but

unconditioned bar tensile and modulus results were previously reported in Table 2-6 and Table 2-7.

Tensile strength and modulus of elasticity have been determined for each specimen. For the time-

dependent samples, all specimen results were averaged and a coefficient of variation (COV) was

calculated. Table 2-8 and Table 2-9 display all of the reported tensile results. Table 2-10 and Table

2-11 show all of the reported modulus of elasticity values. All reported results are based on the

measured bar area of 0.13 in2 (83.9 mm2). When a failure has been noted, it occurred in an

unconfined area of the bar.

22

Table 2-8: Tensile strength of #3 GFRP bars subject to 30, 45, and 57 °C temperatures for 10, 30, 60, 90, and 180 days.

Tensile strength (ksi) Days

Specimen

Number 30°C 45°C 57°C

1 94.6 88.6 81.5

2 95.3 78.5 90.3

3 91.2 91.1 90.7

4 85.6 81.8 76.5

10

5 93.6 79.5 81.2

1 72.3 68.5 62.7

2 86.5 78.1 66.1

3 72.8 65.5 68.1

4 90.1 67.8 61.5

30

5 69.8 83.3 59.6

1 70.3 69.1 52.1

2 71.1 59.8 65.1

3 68.6 63.1 74.7

4 86.1 62.8 51.2

60

5 70 71.7 66.4

1 66.1 59.7 62.1

2 60.7 61.7 60.4

3 63.9 66.9 60.4

4 73 Anchor Slip 59.8

90

5 67.4 69.8 60.2

1 Failure Anchor Slip 40.0

2 Anchor Slip 61.2 39.9

3 59.1 Failure 48.8

4 Anchor Slip Anchor Slip 52.4

180

5 Anchor Slip Anchor Slip 46.8

23

Table 2-9: Average tensile strength and coefficient of variation of #3 GFRP bars subject to

30, 45, and 57°C temperatures for 10, 30, 60, 90, and 180 days.

Tensile strength (ksi)

30°C 45°C 57°C Days

Average COV (%) Average COV (%) Average COV (%)

10 92.1 4.2 83.9 6.7 84.1 7.4

30 78.3 11.8 72.6 10.5 63.2 5.4

60 73.2 9.9 65.3 7.5 62 16.2

90 66.2 6.9 64.5 10.9 60.5 1.4

180 59.1 N/A 61.2 N/A 43.9 12.6

24

Table 2-10: Modulus of elasticity for #3 GFRP bars subject to 30, 45, and 57°C temperatures

for 10, 30, 60, 90, and 180 days.

Modulus of Elasticity (ksi) Days

Specimen

Number 30°C 45°C 57°C

1 5270 5160 5080

2 N/A 5410 5000

3 5340 5230 5260

4 5400 5210 5360

10

5 5270 5380 5230

1 5270 5290 5400

2 5540 N/A 5260

3 5670 5140 5240

4 4640 5170 5500

30

5 5360 5130 4680

1 4470 5380 5420

2 4130 5380 5420

3 5310 5370 5330

4 5420 5320 5230

60

5 5170 5190 5480

1 5270 5300 5330

2 5200 5090 5350

3 5190 5430 5250

4 5390 5330 5440

90

5 5290 5300 5200

1 5860 4970 5200

2 5160 5070 4650

3 5660 4830 5030

4 4550 4720 5070

180

5 5140 4460 5170

25

Table 2-11: Average modulus of elasticity and coefficient of variation of #3 GFRP bars

subject to 30, 45, and 57°C temperatures for 10, 30, 60, 90, and 180 days.

Modulus of elasticity (ksi)

30°C 45°C 57°C Days

Average COV (%) Average COV (%) Average COV (%)

10 5320 5.8 5270 2.52 5180 1.36

30 5290 4.33 5180 1.68 5210 5.54

60 4900 2.3 5380 0.97 5340 2.6

90 5260 2.32 5290 2.6 5310 2.01

180 5070 6.73 4810 7.2 5020 7.17

2.8.2.2 Moisture absorption results

Two specimens were weighed at a given time. Table 2-12 exhibits the moisture absorption

data for the specimens. The reported results are the average of the two specimens.

Table 2-12: Moisture absorption data based on weight gain percentages for #3 GFRP

specimens immersed in limewater solution.

Time (Days) 30°C 45°C 57°C

0.042 0.12 0.16 0.17

0.084 0.16 0.155 0.19

0.167 0.19 0.19 0.22

1 0.26 0.25 0.43

2 0.32 0.39 0.57

5 0.39 0.61 0.74

10 0.51 0.89 1.33

15 0.7 0.95 1.29

30 0.88 1.41 1.81

65 1.06 1.29 1.87

80 1.02 1.57 1.92

26

Utilizing the data in Table 2-12, the maximum moisture and diffusion coefficient for the different

temperatures were determined. The values have been reported in Table 2-13. An Arrhenius analysis

was conducted. The results of the analysis have been summarized in Table 2-14.

Table 2-13: Maximum moisture and diffusion for #3 GFRP bars immersed in a limewater

solution at 30, 45, 57°C.

Temperature (°C) Maximum Moisture (M∞) Diffusion Coefficient (D)

30 1.076 0.168

45 1.484 0.205

57 1.93 0.236

Table 2-14: Arrhenius analysis results based on a Fickian curve fit with respect to immersion

temperatures of 30, 45, and 57°C.

Type Temp

(°C)

EM∞

(kcal/mol)

M∞o

(%MC) R2

ED

(kcal/mol)

Do

(mm2/s) R2

#3 GFRP 30 – 57 4.28 13.394 0.9991 2.49 1.22E-04 0.9999

2.9 Conclusion

A review of literature on the factors affecting durability of FRP was presented. The moisture

and alkali attack degradation mechanisms were discussed. Comparisons and contrasts between

international design code governing FRP reinforced concrete structures were provided. Amongst the

codes compared, all were seen as providing overly conservative guidelines for FRP strength and the

flexural capacity of the structure. Several tensile strength retention models were also discussed. A

comprehensive model incorporating both temperature and moisture effects on GFRP tensile

strength retention is sorely lacking. Existing strength degradation for GFRP bars, which will be used

27

in later chapters, is presented. A cohesive strength retention model is needed, which can be used to

assess the long-term flexural strength reliability of FRP reinforced concrete members.

28

CHAPTER 3 Computational Procedure

3.1 Tensile Strength Retention Models

Tensile strength tests may be performed on a material to determine its’ mechanical properties.

This strength is often used as a parameter when calculating flexural capacity and/or strength of

reinforced concrete structures. Tensile tests are often performed on unconditioned specimens.

Researchers have used accelerated aging techniques to condition specimens and garner long-term

information on a material system.

3.1.1 Existing Tens i l e Streng th Mode l s

Mathematical models have been created to describe the long-term tensile strength retention

of GFRP. Testing specimens often incur damage via elevated temperatures and moisture ingress.

GFRP has extensive usage in civil infrastructure applications. Solutions are generally created to

mimic the alkaline and porous environment of concrete, which is a common infrastructure material.

The proceeding models have been proposed by researchers to determine long-term tensile strength

using limited laboratory conditioning of specimens.

3.1.1.1 Durability prediction via short-term accelerated aging

Chen et al. conducted short-term experiments on GFRP bars to predict long-term tensile

strength via accelerated aging. Two types of GFRP bars were used. All bars were No. 3 in size and

differed in the type of e-glass fibers used [20].

The bars were aged by placing them in a simulated concrete pore solution at elevated

temperatures. The concrete pore solutions’ composition is shown in Table 3-1. Solution 1 has a pH

value of 13.6, and is similar to that of normal concrete. Solution 2 has a pH of 12.7, and is similar to

29

high performance concrete. Elevated temperatures of 40 and 60°C were used with 20°C as a

reference temperature. Table 3-2 shows the experimental test matrix included exposure times [20].

Table 3-1: Chemical compositions of simulated concrete pore solutions. Solution 1 has a pH

of 13.6, and Solution 2 has a pH value of 12.7, which closely resembles the pore environment

found in typical high performance concretes. *

Quantities (g/L)

Solution type NaOH KOH Ca(OH)2

1 2.4 19.6 2

2 0.6 1.4 0.037

Table 3-2: Short-term accelerated aging test matrix for GFRP bars immersed in concrete

pore solutions. [20]

Exposure time

Bar Type

Solution

type

Temperature

(°C) (days) (days) (days) (days)

60 60 90 120 240

40 60 90 120 240

GFRP1 1

20 60 90 120 240

60 60 70 90 120

40 60 70 90 120

GFRP2 2

20 60 70 90 120

The tensile strength retention percentages were determined for both bars at each exposure

interval. Figure 3-1 displays the results for GFRP1 bars, which were exposed to Solution 1. The

* Chen, Yi, Julio F. Devalos, and Indrajit Ray, Durability Prediction for GFRP Reinforcing Bars Using Short-Term Data of Accelerated Aging Tests. Journal of Composites for Construction, 2006. 10(4): p. 279 - 286. Reprinted with permission from ASCE. This material may be downloaded for personal use only. Any other use requires prior permission of the American Society of Civil Engineers. This material may be found at http://cedb.asce.org/cgi/WWWdisplay.cgi?0608541

30

three exposure temperatures have been displayed. For the purposes of this thesis, only the normal

strength concrete experimental data will be considered.

Ten

sile

str

engt

h re

tent

ion

[%]

Exposure time [days]

Figure 3-1: Tensile strength retention of GFRP1 bars exposed to Solution 1 at 20, 40, and

60°C [20]

According to Chen et al, tensile strength retention, Y, can be determined as a function of

exposure time, t, and the inverse of the degradation rate, τ, which is shown in Eq. 3.1 [20].

Y = 100exp !t

"#$%

&'(

3.1

The degradation rate is based on a constant for a material system, A; the activation energy,

Ea; universal gas constant, R; and Kelvin temperature, K. Eq. 3.2 expresses the relationship between

τ and the aforementioned factors.

31

! =1

k=1

Aexp

Ea

RT

"#$

%&'

3.2

these factors are unknown, � , τ can be determined via regression analysis of the tensile

strength retention versus exposure time data. Values of τ and the correlation coefficient have been

reported for both sets of bars and are listed in Table 3-3 [20].

Table 3-3: Coefficients of regression equations for GFRP bar tensile strength retention

immersed in both concrete pore solutions 1 and 2. [20]

GFRP1 Bars in Solution 1 GFRP2 Bars in Solution 2

Temperature

(°C) τ r τ r

60 143 0.93 222 0.99

40 200 0.98 714 0.96

20 256 0.96 1,667 0.94

A master curve can be generated based on the data presented in Table 3-1 and Figure 3-1.

Acceleration factors for the exposure time can be determined using Eq. 3.3. Two temperatures and

the ratio of activation energy to the universal gas constant, Ea/R are needed. The reference

temperature, T0, is 20 °C. The ratio is determined by curve fitting Eq. 3.2 to the data presented in

Table 3-3. The results of the curve fitting are shown in Table 3-4. Table 3-5 displays the acceleration

factors for both GFRP bars. The master curve for GFRP1 bars has been generated and is shown in

Figure 3-2. A single value for τ was determined via regression analysis of the master curve to be 256

days for GFRP1 bars in Solution 1. The corresponding correlation coefficient was 0.92 [20].

32

AF =t0

t1

=

ck0

ck1

=k1

k0

=

Aexp !Ea

RT1

"#$

%&'

Aexp !Ea

RT2

"#$

%&'

= exp !Ea

RT

1

T0

!1

T1

"

#$%

&'(

)*

+

,- 3.3

Table 3-4: Coefficient of Regression Equations for Arrhenius Plots immersed in both

concrete pore solutions 1 and 2 [20]

GFRP1 bars in Solution 1 GFRP2 bars in Solution 2

Tensile strength

retention (%) Ea/R r Ea/R r

50 1,420 0.99 4,891 0.99

60 1,423 0.99 4,892 0.99

70 1,420 0.99 4,891 0.99

80 1,420 0.99 4,892 0.99

Table 3-5: Values for Acceleration Factors [20]

Temperature

(°C) GFRP1 in Solution 1 GFRP2 in Solution 2

60 1.80 7.50

40 1.28 2.33

20 1.00 1.00

33

Ten

sile

str

engt

h re

tent

ion

[%]

Exposure time [days]

Figure 3-2: Tensile strength retention versus exposure time for GFRP1 bars [20]

To determine the long-term validity of the model developed by Chen et al, tensile strength

retention values were obtained for a range of 1 – 100,000 days, which corresponds to a long-term

value of approximately 274 years. The results have been plotted and are shown in Figure 3-3. The

exposure range covers a period of 0.0027 – 274 years. As mentioned in Chapter 2, the service life for

GFRP bridge deck structures is 60 years (21,900 days). The model proposed by Chen et al would

predict a tensile strength retention value of 2.65 !10-17% and 7.05 ! x10-36 %, for 30 and 60

service life years, respectively, which is approximately zero. Thus, at the mid and end points of

service life, the reinforcing bars are projected to have no tensile strength. Therefore, the model is

34

inadequate at projecting long-term tensile strength retention of GFRP bars over a typical GFRP

reinforced bridge deck’s life span.

Ten

sile

str

engt

h re

tent

ion

[%]

Exposure time [days]

Figure 3-3: Tensile strength retention versus long-term exposure time for GFRP1 bars

3.1.1.2 Deterioration prediction from alkali attack

Katsuki and Uomoto performed accelerated testing experiments on AFRP, CFRP, and

GFRP rods. The rods were accelerated by placing them in an alkali solution, which can be used to

simulate a concrete’s pore environment. For the purposes of this research, only the GFRP rods are

35

discussed. The GFRP rods were subject to a 40 °C testing environment and a 1.0 mol/l aqueous

NaOH solution. The bars were exposed for 7, 30, 90, and 120 days. Tensile tests were performed on

all bars. Figure 3-4 contains the tensile test results for the GFRP bars [21].

Ten

sile

str

engt

h [M

Pa]

Ten

sile

str

engt

h [k

si]

Curing time [days]

Figure 3-4: GFRP tensile strength results for specimens subject to accelerated aging at 40 C

with a 1.0 mol/l aqueous NaOH solution. [21]

Failure strength can be determined as a function of the depth of alkali penetration and time.

Although alkali penetration is non-uniform, the residual cross-sectional area can be represented as a

circle. The depth of penetration, x, is obtained via Eq. 3.4.

x = 2 ! k !C ! t 3.4

36

where k is the diffusion coefficient (mm2/he); C is the alkaline concentration (mol/L); and t is the

curing time (hrs). A diffusion coefficient of

2.8!10-6 (cm2/hr) was reported [21].

A relationship between the tensile strength before and after alkali exposure is derived.

Katsuki and Uomoto relied on two primary assumptions. First, regions in which alkali penetration

has occurred have no tensile strength. Secondly, non-alkali penetrated regions have the same tensile

strength as the region prior to immersion in the NaOH solution. Initial strength, σ0, can then be

related to the failure load, Pt, and the sectional area, St, at a certain age as follows in Eq. 3.5

!0=P0

S0

=Pt

St

"Pt= S

t#!

0

3.5

The strength at a certain age, σt , an also be related to initial strength and cross-sectional area

as demonstrated by Eq. 3.6.

!t=Pt

S0

= !0"St

S0

= !0"# R

0$ x( )

2

# " R0

2 3.6

where R0 is the initial bar radius. The GFRP bars had a 3 mm radius. The equations for depth of

penetration and strength at a certain age can be combined in to a single function, the results of

which are presented in Eq. 3.7. Figure 3-5 displays the experimental and calculated tensile strength

results for GFRP bars. The calculated strength values are an overestimate of the experimental

values. The error may be due to the circular approximation of the penetrated area [21].

!t= 1"

2 # k #C # tR0

$

%&'

()

2

#!0

3.7

37

Ten

sile

str

engt

h [M

Pa]

Ten

sile

str

engt

h [k

si]

Time [days]

Figure 3-5: Experimental and projected tensile strength results for GFRP bars

An attempt was made to model long-term strength using Katsuki and Uomoto’s model.

When using long-term values, the projected strength retention increased over time. This result is

implausible due to the primary concept of strength degradation due to alkali penetration. Over time,

it is expected that the tensile strength retained would be reduced from alkali attack. Hence, this

model lacks long-term validity. A shorter and more appropriate time range has been determined.

The results have been plotted and are shown in Figure 3-6.

38

T

ensi

le s

tren

gth

[MP

a]

Ten

sile

str

engt

h [k

si]

Time [days]

Figure 3-6: Predicted tensile strength retention for GFRP based on a diffusion model for

strength retention.

3.1.2 Model ing exis t ing tens i l e da ta

As seen in section 3.1.1 , there are inadequate long-term tensile strength retention models.

This study seeks to develop a long-term tensile strength retention model. The superposition

principle will be applied to create time-temperature superposition as well as time-relative humidity

superposition. Separate tensile strength studies will be integrated using these principles. The

assumptive power-law model for long-term decay will be applied to create an integrated temperature

39

and relative humidity model for tensile strength. Tensile strength retention predictions will be made

and utilized in the subsequent Monte Carlo simulations.

3.1.2.1 Tensile strength with variable conditioning temperatures

Bhise conducted short-term tensile tests on GFRP specimens subject to variable

temperatures. The specimens were tested for a total of 300 days for temperatures of 30, 45, and 57

°C. The specimens were manufactured by Hughes Brothers and contained a minimum fiber volume

fraction of 70% [23].

3.1.2.1.1 Temperature -based t ime shi f t factors

Temperature-based time shift factors were determined. A plot of the temperature versus the

shift factor, aT, is generated. A temperature of 30 °C is used for the reference temperature. A linear

trend line is fit to the data such that the shift factors generate a line with a R2 value closest to one.

Table 3-6 displays the temperature-based shift factors. Figure 3-7 illustrates the linearity of the shift

factors as well as contains the equation for the best-fit line. The temperature-based shift factors are

used to horizontally shift the tensile retention data.

Table 3-6: Temperature-based shift factors for GFRP tensile strength retention based on a

reference temperature of 30°C

Temperature (°C)

30 45 57

aT 1 2.8 4.86

40

Tem

pera

ture

-bas

ed ti

me

shift

fact

or

Temperate [°C]

Figure 3-7: Temperature-based time shift factors for GFRP versus exposure temperature.

Note that the reference temperature is 30 °C.

Once determined, the temperature-based shift factors can now be applied to the exposure

time. Each shift factor is applied to its respective time range. The results of the time shifting process

are shown below in Figure 3-8.

41

T

ensi

le s

tren

gth

rete

ntio

n [%

]

Shifted time [days]

Figure 3-8: Tensile strength retention of GFRP bars subject to variable temperatures after

the temperature-based shift factors have been applied.

3.1.2.1.2 Appli cat ion o f the power law

The power-law model can be generated using the temperature-based shift factors. A least-

squares linear regression analysis is performed on the tensile strength retention data versus the

logarithm of shifted time. The slope was determined to be -0.127, and the intercept is 2.16. Raising

10 to the power of the intercept value yields an approximate value of 145. The tensile strength

retention can be modeled using the power law as shown in Eq. 3.9. The model has been overlaid

with the data presented in Figure 3-8 , and is show in Figure 3-9.

42

Y = 145 ! a

T! t( )

"0.127

3.8

Ten

sile

str

engt

h re

tent

ion

[%]

Shifted time [days]

Figure 3-9: Tensile strength retention versus shifted time for GFRP bars subject to variable

temperatures. The power-law model for GFRP tensile strength retention is overlaid.

3.1.2.1.3 Predi c t i on o f long-t e rm t ens i le s t rength

The power-law model as indicated by Eq. 3.9 can be used to predict long-term tensile strength

retention values. 1, 3, 10, 30, and 60-year tensile strength retention values can be ascertained. They

are present in Table 3-7. Based on the information presented, it is projected that GFRP would have

a retain 41% of their tensile strength at the end of the service life for GFRP reinforced concrete

bridge deck.

43

Table 3-7: Predicted long-term tensile strength retention based on variable temperature data

Time Modeled tensile strength

retention

(years) (days) (%)

1 365 69

3 1095 60

10 3650 52

30 10950 45

60 21900 41

3.1.2.2 Tensile strength with variable relative humidity conditions

Momose et al. conducted tensile tests on GFRP specimens subject to atmospheric humidity.

The specimens were tested for a total of 120 days with a relative humidity range from 30 to 95% and

a constant temperature of 71 °C. The specimens were manufactured by Tolay, Ltd and contained a

fiber volume fraction of 58.7%. Tensile strength retention values were reported and are displayed

below in Table 3-8. Figure 3-10 displays the effects of relative humidity on tensile strength retention

[24].

44

Table 3-8: GFRP tensile strength retention results for variable humidity

Relative Humidity (%)

Days 30 40 60 80 >95

10 106 103 98 87 72

30 99 101 85 72 61

50 105 94 72 65 65

120 104 93 80 62 60

45

T

ensi

le s

tren

gth

rete

ntio

n [%

]

Time [days]

Figure 3-10: Effect of a variable moisture absorbing condition on the tensile strength of

GFRP plates as a function of exposure time. *

3.1.2.2.1 Relat ive humidi ty -based t ime shi f t fac tors

A plot of the relative humidity versus the shift factor, aRH, is generated. A relative humidity

of 30% was used as the reference humidity. A linear trend line is fit to the data such that the shift

factors generate a line with a R2 value closest to one.

* Momose, Yutaka, et al. Effects of moisture on glass fiber reinforced composites. in Proceedings of the 1999 44th International SAMPE Symposium and Exhibition 'Envolving and Revolutionary Technologies for the New Millennium', SAMPE '99. 1999. Long Beach, CA. Copyright by Taylor & Franics. Reproduced with permission.

46

Table 3-9 displays the relative humidity shift factors. Figure 3-11 illustrates the linearity of

the shift factors as well as contains the equation for the best-fit line. The relative humidity-based

time shift factors are used to horizontally shift the tensile retention data.

Table 3-9: Relative humidity-based shift factors for GFRP tensile strength retention based

on a reference relative humidity of 30%.

Relativity Humidity (%)

30 40 60 80 >95

aRH 1 2.8 34 309 1083

Rel

ativ

e hu

mid

ity-

base

d ti

me

shift

fact

or

Relative humidity [%]

Figure 3-11: Relative humidity based time shift factors for GFRP versus exposure relative

humidity. Note that the reference relative humidity is 30%.

47

The relative humidity-based time shift factors can be applied to the exposure time. Each

shift factor is applied to its respective time range. The results of the time shifting process are shown

in Figure 3-12.

Ten

sile

str

engt

h re

tent

ion

[%]

Shifted time [days]

Figure 3-12: Tensile strength retention of GFRP plates subject to variable relative humidity

after the relative humidity-based shift factors have been applied.

3.1.2.2.2 Appli cat ion o f the power law

Using a similar process as applied to Bhise’s data, a power-law model can also be generated

based on the relative humidity-based time shift factors. A least-squares linear regression analysis is

performed on the tensile strength retention data versus the logarithm of shifted time. The slope was

48

determined to be -0.068, and the intercept is 2.13. Raising 10 to the power of the intercept values

yields an approximate value 134. The tensile strength retention can be modeled using the power law

as shown in Eq. 3.9. The model has been overlaid with the data presented in Figure 3-12, and is

show in Figure 3-13.

Y = 134 ! a

RH! t( )

"0.068

3.9

Ten

sile

str

engt

h re

tent

ion

[%]

Shifted time [days]

Figure 3-13: Tensile strength retention versus shifted time for GFRP plates subject to

variable relative humidity. The power-law model for the GFRP tensile strength retention is

overlaid.

49

3.1.2.2.3 Predi c t i on o f long-t e rm t ens i le s t rength

The power-law model as indicated by Eq. 3.9 can be used to predict long-term tensile strength

retention values. 1, 3, 10, 30, and 60-year tensile strength retention values can be ascertained. The

projected tensile strength retention percentages are presented in Table 3-10. The relative humidity-

based tensile strength retention model projects that 68% of the tensile strength will remain in the

GFRP bars at the end of a 60-year service life. It should be noted that this value is over 50% higher

than the retention value from the temperature-based model. The significant difference can be

possibly attributed to the difference in diffusion rates present in each of the experimental conditions.

Table 3-10: Predicted long-term tensile strength retention based on variable relative

humidity data

Time Modeled tensile strength

retention

(years) (days) (%)

1 365 89

3 1095 83

10 3650 76

30 10950 71

60 21900 68

3.1.2.3 Tensile strength with combined relative humidity and temperature effects

Damage may occur to a composite material in many ways. The data presented by Bhise and

Momose et al. provide isolated short-term tensile strengths. A more accurate tensile strength model

must incorporate the effects of both relative humidity (moisture) and temperature. The combined

model will be used to predict long-term tensile strength, and will be employed in the Monte Carlo

simulations.

50

3.1.2.3.1 Combined t ime shi f t fac tors

The time shift factors based on temperature and relative humidity were independently

obtained. Using a reference temperature of 30° and a reference relative humidity of 30%, general

mathematical expressions can be obtained describing these time shift factors.

The generic form of the temperature-based shift factor is shown in Eq. 3.10

aT = Aexp B1

Tref!1

T

"

#$

%

&'

(

)**

+

,--

3.10

where A and B are constants that will need to be fit the variable temperature data.

The linear form of Eq. 3.10 can be obtained by taking the natural logarithm of both sides.

The result of which is shown in Eq. 3.11. A least-square linear regression analysis is performed by

taking the natural logarithm of the temperature-based shift factors versus the inverse the

temperature differences. Figure 3-14 illustrates the linearity of Eq. 3.11.

ln aT( ) = ln A( ) + B !1

Tref + 273.15"

1

T + 273.15

#

$%

&

'( 3.11

51

ln

(th

ime

shift

fact

or)

1

T

!"#

$%&'

1

Tref

!

"#

$

%&

1

K

()*

+,-

Figure 3-14: Natural logarithm of temperature-based shift factors versus the inverse of

temperature. Note that the reference temperature is 30 °C.

The constant A is computed as the exponential of the intercept. It is taken to be 0.958. The

constant B is identical to the slope, and is determined to be 5804.

A similar process may be performed on the relative humidity based shift factors. However,

the generic form of these factors varies slightly from the temperature-based factors. The relative

humidity-based shift factors can be obtained via Eq. 3.12

52

aRH = C exp D RH ! RHref( )"#

$% 3.12

The linear form of Eq. 3.12 is expressed by Eq. 3.13

ln aRH( ) = ln C( ) + D RH ! RHref( ) 3.13

where C and D are constants that will need to be fit to the variable relative humidity data. A least-

squares linear regression analysis is performed by taking the natural logarithm of the relative

humidity-based shift factors versus the absolute relative humidity. Figure 3-15 illustrates the linearity

of Eq. 3.13.

ln (

rela

tive

hum

idit

y-ba

sed

tim

e sh

ift fa

ctor

)

RH ! RHref

Figure 3-15: Natural logarithm of relative humidity-based shift factors versus difference in

relative humidity. Note that the reference relative humidity is 30%.

53

The constant C is computed as the exponential of the intercept. It is taken to be 1.05. The

constant D is identical to the slope, and is determined to be 11.0.

The shift factor equations can be combined through multiplication, as shown by Eq. 3.14.

Eqs. 3.10 and 3.12 can be substituted in to Eq. 3.14 to yield Eq. 3.15.

aT ,RH

= aT!a

RH 3.14

aT ,RH = A !C exp B !1

Tref + 273.15"

1

T + 273.15

#

$%

&

'( + D ! RH " RHref( )

)

*++

,

-..

3.15

3.1.2.3.2 Long-t erm t ens i le s t rength predi c t i on

The data generated by Bhise and Momose et al. will be combined for form one complete

data set. The sample specimens used by Bhise were submerged. Thus, a relative humidity of 100%

was assumed. A least-squares linear regression analysis is performed on the logarithms of both

tensile strength retention and shift time. The slope is reported to be -0.056, and the intercept is

1.981. Raising 10 to the power of the intercept values yields a value of 95.8. The tensile strength

retention can be modeled using the power law as shown in Eq. 3.16. The model and combined

tensile strength retention data are shown in Figure 3-16. Bhise’s early experimental values do not

closely match the power-law model. The discrepancy is potentially due to the initial post-cure effects

while subject to the higher temperatures and 100% relative humidity. The GFRP tensile strength

retention percentages that will be incorporated in to the Monte Carlo simulations are presented in

Table 3-11.

Y = 95.8 ! t"0.056 3.16

54

T

ensi

le s

tren

gth

rete

ntio

n [%

]

Shifted time [days]

Figure 3-16: Tensile strength retention versus shifted time for GFRP composites with

superimposed temperature and relative humidity effects. The power-law model for the

GFRP tensile strength retention is overlaid.

55

Table 3-11: Predicted long-term tensile strength retention for GFRP based on superimposed

relative humidity and temperature data.

Time Modeled tensile strength

retention

(years) (days) (%)

1 365 69

3 1095 65

10 3650 61

30 10950 57

60 21900 55

It should be noted that the glass fiber reinforced composite materials used by Bhise and

Momose et al. are not the same. The fiber volume fraction differed significantly between the two

sample types. The bars created by Hughes Brother contained a minimum of 70% glass fiber content

by weight [23]. The constant temperature of 71°C used in Momose et al. experiments is a higher

temperature than what any of Bhise’s bars were subject to in a water environment. The tensile

strengths between samples have been normalized by using tensile retention percentages to develop

the model, which negates the intrinsic sample material differences. Bhise’s samples were also subject

to an exposure time of 300 days of 100% relative humidity versus 120 days of variable humidity for

Momose et al. The combined model adequately reflects the combined effects of temperature and

moisture on the projected long-term tensile strength retention of GFRP bars.

3.2 Monte Carlo Simulations

The Monte Carlo technique was used to generate random variables. The concrete beam

geometry, concrete strength, as well as the GFRP bar tensile strength and modulus of elasticity was

56

randomized. Ten thousand random sample points were generated for each parameterized elements.

The sample points were used to determine the flexural capacity of a FRP reinforced concrete beam.

3.2.1 Random number g enera tion

The random numbers were generated using the built-in Random Number Generator in

Microsoft Excel. A uniform distribution of numbers between 0 and 1 is used. All randomly

generated numbers that are 0 and 1 are discarded, and new numbers are generated. The inverse of

the normal distribution is computed based on the mean and standard deviation of the parameterized

element as well as the probability of the event. The probability is equivalent to the randomly

generated number.

3.2.2 Parameter ized e l ements

The Monte Carlo simulation method relies on the application of probabilistic distribution

information to determine the parameters’ uncertainties. Beam geometry, concrete strength, and

tensile and modulus of elasticity of the GFRP bars are the critical parameterized elements in this

study. These variables can be used to determine flexural strength and failure strain criteria. The area

of the bars is not parameterized due to its deterministic nature.

3.2.2.1 GFRP tensile strength

The guaranteed tensile strength, as determined by manufacturers, can be used to determine

the bias, λ, and the coefficient of variation, V, for a set of GFRP bars. Bhise utilized #3 Hughes

Brothers bars [14]. Table 3-12 displays physical properties of the #3 bar as determined by Hughes

Brothers, which have been reproduced here with permission [23].

57

Table 3-12: Physical properties of #3 Aslan 100 GFRP Rebar by Hughes Brothers

Bar Size

Cross Sectional

Area* Nominal Dia.

Guaranteed

Tensile Strength

Tensile Modulus

of Elasticity

(mm) (inches) (mm2) (in2) (mm) (in) (MPa) (ksi) (GPa) psi 106

9 #3 84.32 0.1307 9.53 0.375 760 110 40.8 5.92

The bias of the tensile bar strength,

! f fu, can be determined via Eq. 3.17

! f fu=µ f fu

f fu 3.17

where

µ f fuis the mean experimental bar strength, and

f fu is the guaranteed tensile strength as based

on the retention percentages that were reported Table 3-12. The original tensile strength to be

applied to the retention percentages is approximately 85.6 ksi. The coefficient of variation, V, can be

computed using Eq. 3.18

Vffu=! f fu

µ f fu

3.18

where

! f fuis the standard deviation of the tensile strength.

A Weibull distribution can be fit to the experimental bar strength data [25]. The two main

parameters of the Weibull distribution are the shape parameter, m, and the scale parameter,

!o. The

parameters can be determined using either linear estimators or the process of maximum likelihood

estimation, MLE, both of which will be discussed in detail in Section 3.3.1.2 . Maximum likelihood

estimation will be used to determine all Weibull distribution parameters.

The Weibull parameters can be used to determine the mean bar strength and coefficient of

variation [25]. The mean bar strength is computed via Eq. 3.19

58

µ f fu= !

0"1+ m

m

#

$%&

'() !

0 3.19

where Γ the gamma function. The coefficient of variation, V, can then determined by Eq. 3.20.

Vffu=

!2 + m

m

"

#$%

&'

!2 1+ m

m

"

#$%

&'

(1 )1.2

m 3.20

A Weibull distribution has been fit to Bhise’s unconditioned plain bar specimens. The mean,

bias, and coefficient of variation have been determined. Okeil and Kulkarni conducted a similar

analysis on #2 through #10 FRP bars [25].

Table 3-13 provides a summary of the results for a #3 GFRP bar. There are differences

between the values reported by Bhise and Okeil and Kulkarni. These differences can be attributed to

dissimilar sample sizes and bar manufacturers.

Table 3-13: Statistical and mechanical properties of GFRP bars

Standard

Deviation COV

Weibull Distribution

Parameters Nominal

Tensile Strength Bias

! f fu

Vf fu

!o

Reference (MPa) (ksi)

! f fu (MPa) (ksi) (%) m (MPa) (ksi)

Bhise 590 85.6 0.778 11.14 1.615 2.07 61.8 595.3 86.3

Okeil and

Kulkarni

[25]

760 110 1.18 105.1 15.24 11.8 10.2 938.4 136.1

3.2.2.2 Concrete Strength

Concrete strength is used when determining the depth to the neutral axis. A quadratic

equation must be solved to obtain this parameter for the cases in which the GFRP bar does not

59

rupture. For this study, concrete strengths of 4000 (27,560), 5000 (34,450), and 6000 psi (41,340

kPa) will be considered.

Nowak and Szerszen conducted a statistical study to calibrate the steel reinforced concrete

design code. Strength of concrete, reinforcing steel bars, and pre-stressing strands were analyzed.

The bias and coefficient of variation values for ready-mix concrete under consideration are reported

in Table 3-14 [26].

Table 3-14: Statistical parameters of ordinary ready-mix concrete

! f c Mean

! f c

(kPa) (psi)

Number of

Samples (kPa) (psi) λ V

27,560 4000 116 34037 4937 1.235 0.145

34,450 5000 30 39,480 5726 1.15 0.058

41,340 6000 30 46,163 6695 1.12 0.042

3.2.2.3 Concrete beam and slab dimensions

Three dimensions are used to describe a concrete beam or slab. Figure 3-17 illustrates the

dimensions

Figure 3-17: Concrete beam with dimensions

b is the width of rectangular cross section; d is the distance from extreme compression fiber to the

neutral axis; and h is the overall height of the member. The dimensions are often reported in inches

or millimeters.

60

As part of their calibration study, Nowak and Szerszen also reviewed fabrication factors

when designing steel reinforced concrete structures. The fabrication factor refers to the natural

variations in dimensions and geometry during construction Fabrication factors for cast-in-place,

plant-cast, pre-stressed, and post-tensioned have been provided [26]. Only factors for cast-in-place

concrete were taken into consideration for this study.

Okeil et al. examined the reliability of bridge girders that had been strengthened by carbon

fiber-reinforced polymer laminates. Various statistical properties, namely the dimensions, were

reviewed prior to performing Monte Carlo simulations. Statistical property values were compared to

previously reported values [27]. Table 3-15 shows the results of the parameters provided by Nowak

and Szerszen as well as Okeil et al. The values for Okeil et al. will be used for the overall height of

the member, h.

Table 3-15: Statistical parameters of dimensional variables

Researcher Variable λ V

b 1.01 0.04 Nowak and Szerszen

[26] d 0.99 0.04

Okeil et al. [27] b,h,d 1.00 0.03

3.2.2.4 Summary of parameterized elements

Several variables were taken in account for the Monte Carlo simulations. Statistical

parameters of these variables have been obtained. Table 3-16 provides a summary of the variables as

well as the bias and coefficients of variation that will be incorporated into the simulations.

61

Table 3-16: Statistical properties of variables involved in this study

Researcher Variable λ V

b 1.01 0.04 Nowak and Szerszen

[26] d 0.99 0.04

Okeil et al. [27] h 1.00 0.03

4000 psi

(27,560 kPa) 1.235 0.145

5000 psi

(34,450 kPa) 1.15 0.058

Nowak and Szerszen

[26]

! f c

6000 psi

(41,340 kPa) 1.12 0.042

Bhise [14]

f f fu 0.78 0.0207

3.2.3 Flexural des ign of FRP re in fo rc ed concre te beams

The flexural strength of GFRP reinforced concrete beams was simulated using the Monte

Carlo technique. Each simulation consisted of the 10,000 data points.

3.2.3.1 Input values

The height and width of the concrete slab was input. Three geometric configurations werew

considered. The distance from extreme compression fiber to the centroid of tension reinforcement,

d, is computed by subtracting the clear cover distance from the overall height, h. A standard clear

cover value of 2.625 in. (66.675 mm) has been assumed for all of the configurations and is

completely arbitrary. Table 3-17 displays the dimensions and member name for each configuration.

62

Table 3-17: Member dimensions

b h d Member

Name (in) (mm) (in) (mm) (in) (mm)

A 12 304.8 24 609.6 21.375 542.925

B 9 228.6 16 406.4 13.375 339.725

C 18 457.2 48 1,219.2 45.375 1,152.5

The deterministic values presented in Table 3-17 were used as “mean values” to obtain

randomized dimensional elements. The standard deviation can be acquired by using the variance

information presented in Table 3-16 and Eq. 3.18.

Concrete compression strength values of 4000 (27,560), 5000 (34,450), and 6000 psi (41,340

kPa) will be considered. A similar procedure can be conducted to determine the “mean” and

standard deviation values for each of the concrete strengths.

Reinforced concrete design also utilizes a concrete calibration factor, β1, which was derived

by Whitney. β1 is 0.85 for concrete strengths of up to 4000 psi (28 MPa). A reduction of 0.05 is

taken for every 1,000 psi (7 MPa) increase in concrete strength. Eq. 3.21 will be used to

automatically calculate β1 for each of the concrete strengths. It should be noted that β1 can not be

less than 0.65 [15].

!1= 0.85 + 0.05 4 "

#fc1000

$%&

'()

3.21

3.2.3.2 GFRP reinforcement

The ACI 440 committee has established minimum levels of FRP reinforcement ratios. The

minimum ratios are of importance when designing reinforced structures with an expected failure

mode of FRP bar rupture. The minimum reinforcement area, Af, is similar in prescription to that of

63

steel reinforced concrete structures. Eq. 3.22 will be used to establish the minimum reinforcement

area.

Af ,min =4.9 !fc

f f fu

bwd "330

f f fu

bwd 3.22

where bw is the width of the web in inches. It should be noted that there is a separate equation for

metric units. All configurations under consideration have rectangular cross-sectional areas.

Therefore, the width of the web is taken to be the same as the width of the beam. Thus, Eq. 3.22

can be re-written as follows in Eq. 3.23.

Af ,min =4.9 !fc

f f fu

bd "330

f f fu

bd 3.23

The minimum reinforcement area does not apply to sections designed to fail via concrete

crushing, as the conditions of Eq. 3.22 are satisfied. The minimum reinforcement ratio, ρmin, can be

determined once Eq. 3.22 has been calculated via Eq. 3.24.

! f =Af

bd 3.24

Reinforcement ratios ranging from 0.1 to 1.4 times the balanced reinforcement ratio, ρfb,

were considered. The level of reinforcement is indicates whether over or under-reinforcement of the

FRP bars has been allocated by the designer. Under-reinforced sections are more susceptible to

failure via FRP bar rupture. Conversely, over-reinforced sections are designed to have concrete

crushing occur. It should be noted that a balanced reinforcement ratio is subject simultaneous FRP

bar rupture and concrete rushing. The balanced reinforcement ratio takes in to account the

mechanical properties for the FRP bars as well as the concrete strength. This ratio can be computed

by Eq. 3.25.

64

! fb = 0.85"1#fc

f f fu

E f$cu

E f$cu + f fu 3.25

where εcu is the ultimate strain in the concrete, and Ef is the design or guaranteed modulus of

elasticity of the FRP. The ultimate strain in concrete is always taken to be 0.003. The balanced area

of reinforcement is then back-calculated using both Eqs. 3.24 and 3.25.

The total number of reinforcement bars can be determined once the minimum balanced

reinforcement area is known. The reinforcement ratio associated with the total reinforcing bar area

is then computed. Fractions of the reinforcement ratio are then applied. The fraction of balanced

reinforcing ratio is increased by one-tenth for each round of simulations until 1.4 times the balanced

reinforcement ratio is achieved. The reduced reinforcement ratio and area are then determined. The

reduced values are compared to the minimums as established by Eqs. 3.23 and 3.24.

Using the power law model developed in Section 3.1.2.3 , the predicted age of the GFRP

bars can be applied to the model. The standard deviation is based on the unconditioned specimens,

which is 1615 psi (11 MPa). The unconditioned values for bar modulus of rupture and its standard

deviation are also used.

As per ACI 440 code, environmental reduction factors must be applied to the guaranteed

tensile strength of the FRP bars. The design tensile strength, ffu, is determined by Eq. 3.26 as follows

f fu = CE f fu* 3.26

where CE is the environmental reduction factor, and f fu* is the guaranteed tensile strength as defined

by Eq. 3.27. Environmental reduction factors of 0.7, 0.8, and 1.0 will be considered. The guaranteed

tensile strength is taken as the mean bar strength, fu,ave, minus three times the standard deviation [15].

f fu*= fu ,ave ! 3" 3.27

65

3.2.3.3 Strain-compatibility analysis

The flexural strength of under-reinforced FRP reinforced concrete members cannot be

calculated using the Whitney rectangular stress block. The maximum concrete strain of 0.003 may be

violated or not achieved, and is determined by the mode of failure. The depth to neutral axis, c, must

be determined. A quadratic equation for concrete design can be solved to determine the parameters

of c by using an approximation of the Whitney stress block. The three constants for the concrete

design quadratic are a1, b1, and c1. They can be calculated using Eqs. 3.28 - 3.30.

a1= 0.85 ! "

1!b ! #fc 3.28

b1= Af !Ef ! "cu 3.29

c1= Af !Ef ! "cu !d 3.30

Once the constants have been obtained, the quadratic formula may be applied determine the

depth to neutral axis. As shown below in Eq. 3.31, only the positive value of the quadratic is used.

c =b1+ b

1

2! 4 "a

1" c1

2 "a1

3.31

The FRP strain at rupture and at the ultimate condition of concrete strain may now be

ascertained. The ultimate strain, εcu, and FRP strain at rupture, εrupt, obtained via Eqs. 3.32 and 3.33,

respectively.

! fu =!cu

cd " c( ) 3.32

!rupt =f f

E f

3.33

A ratio of strains is also computed. A strain-based failure mode is selected. FRP bar rupture

is presumed to have occurred when the FRP strain rupture is greater than the ultimate strain.

Concrete crushing is the failure mode when ultimate strain is achieved before the bars have

66

ruptured. The strain-based failure mode then confirms the governing nominal flexural strength, as

will be discussed in Section 3.2.3.4 .

3.2.3.4 Nominal flexural strength

Although strain-compatibility analysis provides insight in to the failure mode, the flexural

strength of the FRP reinforced member is still unknown. Concrete crushing and FRP bar rupture

have separate processes to determine the appropriate flexural capacity.

The bar strength must be determined when designing for failure due to concrete crushing.

The bar strength is taken as the lesser of design tensile strength, ffu, and the tensile stress in the FRP

reinforcement, ff. The design tensile strength has been specified as an input value. The tensile stress

may be computed via. Eq. 3.34 [15].

f f =Ef ! "cu( )

2

4+0.85 ! #

1! $fc

% f

E f ! "cu & 0.5 !Ef ! "cu

'

(

))

*

+

,,- f fu 3.34

The flexural strength due to concrete crushing can be derived using Eqs. 3.35 and 3.36 [15].

Mn = Af f f d !a

2

"#$

%&'

3.35

where a is the depth equivalent of the rectangular stress block.

a =Af f f

0.85 ! "fc !b 3.36

The determination of a member’s flexural strength due to FRP bar rupture is more complex.

The quadratic equation for concrete stress-strain behavior must be solved to determine the flexural

capacity. The bar strength is based on the design tensile strength. Also, the depth to the neutral axis

is used instead the depth to the neutral axis as calculated by Eq. 3.31. Once obtained, the flexural

strength can be computed as follows in Eq. 3.37 [15].

67

Mn = Af f fu d !"1c

2

#$%

&'(

3.37

The flexural strength for each simulation is taken as the lesser of the flexural strength due to

concrete crushing or FRP bar rupture. The minimum values should coincide with the failure mode

predicted by the strain-compatibility analysis.

Information regarding the nominal slab design and Monte Carlo results are reported for each

simulation series. The nominal design moment is taken the flexural strength of the member based

on the deterministic input values. The design flexural strength is computed by applying a strength

reduction factor for flexure to nominal design value. The strength reduction factor is dependent

upon the balanced and calculated values for the reinforcement ratio. Eq. 3.38 provides the guidelines

for selecting the correct strength reduction factor.

! =

0.55 for " f # " fb

0.3+ 0.25" f

" fb

for " fb

0.65 for " f $ 1.4" fb

%

&

''

(

''

< " f < 1.4" fb 3.38

The average flexural strength for the Monte Carlo simulations is computed. The standard

deviation and coefficient of variation are also reported for each set. All flexural strengths are

obtained in both English and SI units. The number of simulations below the nominal and design

flexural strengths are tabulated. The number of rupture failures is also counted.

3.3 Reliability assessment

The reliability assessment relies on determining an appropriate cumulative density function to

model the flexural strength data. The Monte Carlo simulations are based on normally distributed

random variables. The Weibull distribution will be used to model the flexural strength. The specific

parameters of the distribution must be obtained for each simulation. Confidence intervals will be

68

determined for both the Weibull parameters as well as the reliability function for the flexural

strength.

3.3.1 Weibul l dis tr ibut ion mode l s

Scientists and engineers often use the Weibull distribution. The distribution is readily

applicable to many different types of data sets. The distribution is flexible, and its parameters are

easily obtained through a variety of numerical methods.

Extreme value theory can be used to identify the third asymptotic distribution. Fisher and

Tippet originally derived this distribution. Waloddi Weibull used the distribution to review breaking

strengths, and now bears it namesake [28]. The Weibull distribution will be used to model the

flexural strength for each configuration.

3.3.1.1 Background on the Weibull distribution

A Weibull model can be based on either two or three parameters. The three parameters are

shape, scale, and location. The notation describing the distribution varies. For this study, Dodson’s

notation will be used. The Weibull probability density function is described below in Eq. 3.39

f x( ) =! x " #( )

!"1

$! exp "x " #$

%&'

()*!+

,--

.

/00, x 1 # 3.39

where β is the shape parameters, θ is the scale parameter, and δ is the location parameter [29]. The

shape parameter may be also referred to as the Weibull modulus.

The Weibull cumulative density function is described by Eq. 3.40 as

FXx( ) = 1! exp !

x ! "#

$%&

'()*+

,--

.

/00

3.40

69

The Weibull reliability function will be used more often than the Weibull cumulative density

function. The function is often computed as the one minus the cumulative density function. A more

formal definition of the Weibull reliability function is shown below in Eq. 3.41.

R x( ) = exp !x ! "#

$%&

'()*+

,--

.

/00

3.41

The shape parameter, β, alters the shape of the probability density function, as seen in Figure

3-18 [29]. The shape parameter is characteristic property of the modeled data set. For example, the 0

and 60 year Monte Carlo simulation configurations will have the same shape parameter but different

scale parameters.

Figure 3-18: The Weibull probability density function [29]

Murthy et al. have proposed that there are seven different Weibull models. Table 3-18 shows

the types and descriptions of Weibull models [30].

70

Table 3-18: Taxonomy for Weibull Models.

Type Description

I Transformation of Weibull Variable

II Transformation of Weibull Distribution

III Univariate Models involving Multiple Distributions

IV Varying Parameters

V Discrete Models

VI Multivariate Models

VII Stochastic Models

A goodness-of-fit test such as Chi-square or Kolmogorov-Smirnov can be used to determine

the appropriate type of model. The standard Weibull distribution will be used in this study, which is

a Type II model. Although the two-parameter distribution results in lower nominal strengths, there

is only a 1% decrease in values [31]. The location parameter has been assumed to be equal to zero.

Therefore, a two-parameter Weibull distribution will be used.

3.3.1.2 Methods for estimating parameters

Graphical and statistical methods may be used to determine the parameters of the Weibull

distribution. The statistical methods include moment estimator, percentile estimator, maximum-

likelihood estimator, Bayesian estimator, and interval estimator. The graphical method often relies

on plotting the data and fitting it to a line [30]. The linear and maximum likelihood estimators are

the only methods of interest in this study.

The easiest way to obtain the Weibull parameters is by using linear estimators. The slope of

the line is the Weibull modulus. The scale parameter can be taken as the exponential of the negative

ratio of between slope and intercept. The linear estimators provide a good estimate. However, the

method of maximum likelihood provides the most accurate results for the Weibull parameters [29].

71

The method of maximum likelihood provides a more robust estimate for the Weibull

parameters. The parameters may be obtained by simultaneously solving Eqs. 3.42 - 3.44 [31]. The

Solver feature of Microsoft Excel was used to solve the simultaneous equations.

xi! "( )

#ln x

i! "( )

i=1

n

$

xi! "( )

i=1

n

$!1

#%&'

()*!

ln xi! "( )ni=1

n

$ = 0

3.42

! =

xi" #( )

$

i=1

n

%n

&

'

((((

)

*

++++

1

$

3.43

!

"!xi# $( )

i=1

n

%!#1

# ! #1( ) xi# $( )

i=1

n

%#1

= 0 3.44

3.3.1.3 Development of confidence intervals

Given the large sampling size for each simulation, nearly exact confidence intervals may be

obtained. All data points will be considered uncensored. Ninety-fifth percentile confidence intervals

for the parameters and overall distributions can be determined after the Weibull scale and shape

parameters are known. Dodson outlines the process for ascertaining both sets of confidence

intervals [29]. The confidence intervals for the parameters will be determined prior to the

distribution’s intervals.

72

3.3.1.3.1 Confidence int ervals for Weibu l l parameters

Confidence intervals were obtained for the scale and shape parameters. The location

parameter has been assumed to be equal to 0. Thus, no confidence intervals can be determined. The

flexural strength simulation points must be sorted in to an ascending order [29].

First, the confidence intervals for the shape parameter were calculated. The second

derivative of the of the log-likelihood function with respect to the scale parameter is determined for

each uncensored data point using Eq. 3.45.

!2Lu

!" 2=

#" 2

$xi

"%&'

()*# #

" 2%&'

()*# +1( ) 3.45

Next, The second derivative of the of the log-likelihood function with respect to the shape

parameter is determined for each uncensored data point using Eq. 3.46.

!2Lu

!" 2= #

1

" 2#

xi

$%&'

()*"

lnxi

$%&'

()*

+

,-

.

/0

2

3.46

Third, the second derivate of the log-likelihood function with respect to both the scale and

shape parameters is calculated for each uncensored data point using Eq. 3.47.

!2Lu

!"!#= $

1

#+

xi

#%&'

()*"1

#" ln

xi

#%&'

()*+1

+

,-

.

/0 3.47

A similar process is conducted for each of the censored points. The second derivatives for

the censored points are equal to zero, as there are no censored points. The equations have been

included for future reference though. The second derivative of the log-likelihood function with

respect to the scale parameter is obtained via Eq. 3.48.

!2Lc

!" 2= #

xi

"$%&

'()* *

" 2$%&

'()* +1( ) 3.48

73

The second derivative of the log-likelihood function with respect to the shape parameter is

determined via Eq. 3.49.

!2Lc

!" 2= #

xi

$%&'

()*"

lnxi

$%&'

()*

+

,-

.

/0

2

3.49

The second derivate of the log-likelihood function with respect to both the scale and shape

parameters is calculated for each censored data point using Eq.

!2Lc

!"!#=

xi

#$%&

'()"1

#$%&

'()

" lnxi

#$%&

'()+1

*

+,

-

./ 3.50

The second derivatives for both the uncensored and censored data points with can be used

to determine the local information matrix. This matrix will be used when obtaining confidence

intervals for both the scale and shape parameters. The local information matrix is obtained using Eq.

3.51.

F =

!"L

T

2

"# 2!

"LT

2

"$"#

!"L

T

2

"$"#!"L

T

2

"$ 2

%

&

''''

(

)

****

3.51

where

!2LT

!"2=

!2Lui

!"2

i=1

r

# +!2Lci

!"2

i= r+1

n

# 3.52

!2Lt

!" 2=

!2Lui

!" 2+

!2Lci

!" 2

i= r+1

n

#i=1

r

# 3.53

!2LT

!"!#=

!2Lui

!"!#+

!2Lci

!"!#i= r+1

n

$i=1

r

$ 3.54

Additionally, r is the number of failures, while n is the total number of data points. Both are equal to

10,000 for the data sets used because the system is assumed to have failed. Furthermore, the number

74

of failures is the identical to the number of uncensored data points. It should be noted that all

summations involving censored points are equal to zero.

The local information matrix is then inverted. The parameters are often referred to the

following as noted in Eq. 3.55. The variance of the scale parameter is equal to F(1,1)

!1 . The variance of

the shape parameter is equal to F(2,2)

!1 . The covariance of the scale and shape parameters is equal to

F(1,2)

!1 .

F!1=

F(1,1)

!1F(1,2)

!1

F(2,1)

!1F(2,2)

!1

"

#$%

&' 3.55

Ninety-five percent confidence limit states were used. The lower and upper confidence limits

for the shape parameter are determined by Eqs. 3.56 and 3.57, respectively.

!L=

!

expK F(2,2)

"1

!

#

$%%

&

'((

3.56

!U= ! exp

K F(2,2)"1

!

#

$%%

&

'((

3.57

where K is the standard normal percentile for the percent limit as determined by Eq. 3.58.

K = 100 1!"2

#$%

&'(

3.58

The percent limit is based on Eq. 3.59

100 1!"( ) 3.59

and α is the percentile of interest. The 95th percentile has an α value of 0.05 and is equal to a 97.5th

standard normal percentile. Therefore, K can be computed as the inverse of the standard normal

distribution, and is approximately equal to 1.96.

75

The lower and upper confidence limits for scale parameter can be readily calculated via Eqs.

3.60 and 3.61, respectively.

!L=

!

expK F(1,1)

"1

!

#

$%%

&

'((

3.60

!U= ! exp

K F(1,1)"1

!

#

$%%

&

'((

3.61

3.3.1.3.2 Confidence int ervals for Weibu l l di s t r ibut ions

Confidence intervals for Weibull distributions examine the reliability of the data set. The

method of maximum likelihood will be used to determine the 95th percentile confidence intervals.

In recalling Eq. 3.55, variances of the estimates are automatically calculated. The variance of the

scale parameter is equal to F(1,1)

!1 . The variance of the shape parameter is equal to F(2,2)

!1 . The

covariance of the scale and shape parameters is equal to F(1,2)

!1 .

RLx( ) = e

! exp u+K var u( )( ) 3.62

RUx( ) = e

! exp u!K var u( )( ) 3.63

where

u = ! ln x( ) " ln #( )$% &' 3.64

var u( ) = ! 2 var "( )

" 2#$%

&'(+

u2var !( )

! 4

#$%

&'()2u cov !,"( )

! 2"#$%

&'(

*

+,

-

./ 3.65

Eq. 3.65 can be re-written in terms of the inverted local information matrix values, and is

shown in Eq. 3.66.

76

var u( ) = ! 2F(2,2)

"1

# 2$

%&'

()+

u2 *F

(1,1)

"1

! 4

$

%&'

()"2u *F

(1,2)

"1

! 2#

$

%&'

()+

,--

.

/00

3.66

3.3.2 Determinat ion of env ironmental re duc tion fac tors

The American Concrete Institute requires the application of an environmental reduction

factor to the guaranteed tensile strength of FRP. Values are based on the exposure conditions that

the concrete is subject to, and were shown in Table 2-3. The goal is to determine a reduction factor

such that the initial FRP tensile strength is reduced by a sufficient amount, and thereby

approximates the projected long-term tensile strength of the material. A reliability-based approach

was used to determine appropriate environmental reduction factors.

3.3.2.1 Monte Carlo simulations for bar strength

This portion of the study seeks to determine more appropriate factors. According to Eq.

3.26, the design tensile strength is directly related to the environmental reduction factor. Therefore,

it is the only randomized variable for these Monte Carlo simulations. Age-defined tensile strength

values were based on the Power Law model described in Sec. 3.1.2.3 , and were used as the mean

bar strength. The standard deviation is based on the unconditioned GFRP bar data provided by

Bhise.

The concrete members were presumed to be subject to flexure. They were designed in

accordance with the procedure outlined in Sec. 3.2.3 . Ten thousand simulations for GFRP tensile

strength were performed for 0, 1, 3, 10, 30, and 60 years. Unique values for the environmental

reduction will be determined for each of the years.

77

3.3.2.2 Computational procedure

The development of the design tensile strength of FRP takes in to account the mean minus

three standard deviations value, which is the guaranteed tensile strength. For a normal distribution,

the 3 standard deviations pertain to 6 standard deviations over the data range. Therefore, 99.73% of

the data will lie under the area of the normal distribution curve. The lower tail of -3σ represents the

bottom 0.135% of the normal cumulative distribution function (CDF). Conversely, the upper tail of

+3σ refers to 99.865% of the normal CDF.

The Weibull distribution is generally not symmetric like the normal distribution. The upper

and lower tails can be obtained manually or via approximation. The approximation of tails selects

data values at 99.9% and 0.01%, which is very similar to having 99.73% of the data lying under the

normal CDF. The alternative approach is to obtain the mean and standard deviation of the Weibull-

modeled data. The mean plus or minus three standard deviations can be determined. The manual

method of determining the tails were used. Once the lower tail of the Weibull distribution has been

determined, its corresponding reliability value may be obtained via Eq. 3.41.

Okeil and Kulkarni have shown that FRP bar strength can be modeled using a Weibull

distribution [25]. The Weibull distribution for the bar strength can be transformed in tot a

distribution for flexural strength. Ninety-five percent upper and lower reliability limits will be

determined based the flexural strength. The mean, standard deviation, and coefficient of variation

will be computed.

The tail analysis was performed on the initial data to determine the corresponding reliability

level. In the following derivation, the initial conditions and time dependent conditions are denoted

as t1 and t2, respectively. The upper reliability tail of minus three standard deviations is computed by

Eq. 3.67.

78

Mn( )

t1

= µt1! 3"

t1 3.67

The flexural strength can then be substituted in to Eq. 3.41. The re-written equation is

shown below in Eq. 3.68.

R = exp !M

n( )t1

"t1

#

$%

&

'(

)*

+

,,

-

.

//= exp !

µt1! 30

t1

"t1

#

$%

&

'(

)*

+

,,

-

.

//

3.68

The reliability will be applied to obtain the age specific flexural strength. The flexural

strength can be written in terms of either the reliability and Weibull distribution parameters or the

deterministic flexural strength equation. Both options are displayed in Eq. 3.69.

Mn( )t2= !t2 ln

1

R

"#$

%&'

1

()

*

++

,

-

.

.= Af CE fu ,ave / 30( ))* ,- d /

(1c

2

)*+

,-.

123

456t2

3.69

The environmental reduction factor can be see as the fraction of strength retained under

harsh exposure conditions. Thus, the environmental reduction factor can be approximated as the

ratio of the reliability-based aged tensile strength and the unconditioned guaranteed tensile strength.

The overall mathematical expression for the environmental reduction factor is presented in Eq. 3.70.

CE =

!t2 ln1

R

"#$

%&'

1

()

*

++

,

-

.

.

Af d /(1c

2

)*+

,-.

0

1

222

3

222

4

5

222

6

222t2

fu ,ave / 37( ){ }t1

3.70

The process was repeated for each of the simulation years. Concrete strength and beam

geometry will also be varied. The environmental reduction factor at both time intervals and the 60-

year mark will also be incorporated in to the Monte Carlo simulations. Comparisons between the

recommended environmental reduction factor and those mandated by the ACI code will be made.

79

3.3.3 Determinat ion of s t reng th reduc ti on fac t or s

The current ACI values for the strength reduction factor are based on the desired failure

mode and level of reinforcement. Factors can be selected using Eq. 3.38. The basis for the values is

derived from the ACI code for steel reinforced concrete structures. There is a need to determine

strength reduction values based on the behavior FRP material itself.

Three different approaches were used to determine strength reduction factors. The

corresponding design flexural strength will be resolved for each Monte Carlo simulation using Eq.

3.71

Mn Design( )= ! Mn Nomin al( )( ) 3.71

where φ is the strength reduction factor; and MnNomin al( )

is the nominal flexural strength. The nominal

flexural strength is taken as the minimum flexural strength value per the limit state flexural strength

function for either concrete crushing or FRP bar rupture.

The first factor was selected in accordance with Eq. 3.38. Based on the fractions of

reinforcement ratios under consideration, most design values will incorporate a strength reduction

factor of 0.55.

Ellingwood has outlined the critical need to apply a load and factor resistance (LFRD)

approach in determining the strength reduction factors. LFRD has been successfully applied to

wood and steel structure design. Composite materials, such as FRP, can also be designed using

LFRD. An approximate value of a resistance factor can be obtained using Eq. 3.72

! =µR

Rn

"

#$%

&'exp ()

R* + *V

R[ ] 3.72

where µR is the mean resistance level; Rn is the nominal strength; αR is the sensitivity coefficient; β is

the safety index; and VR is the variability.

80

The mean resistance level is the average flexural strength of all simulations. The variability

will be taken as the coefficient of variation of the flexural strength. The sensitivity coefficient has

been assumed to be 0.6 [19]. ACI has a minimum factor of safety, β, of 3.5, with an allowable range

of 3.5 – 4 [15]. For the purposes of this study, safety factors of 3.5 and 4 are considered using Eq.

3.72. Numerical substitutions can be made, and Eq. 3.72 can then be re-written as Eqs. 3.73 and

3.74 for the minimum and maximum safety factors, respectively. Ellingwood’s approach will

constitute the second strength reduction factor.

! =µM

n

Mn Nominal( )

"

#$$

%

&''exp (2.1 )V

R[ ] 3.73

! =µM

n

Mn Nominal( )

"

#$$

%

&''exp (2.4 )V

R[ ] 3.74

The third strength reduction factor was obtained using the brute force method. A maximum

value for Eq. 3.71 was obtained by solving for a strength reduction factor. An additional constraint

is imposed on Microsoft Excel Solver. The design value is considered by design practice as the

minimum flexural strength value. Therefore, the strength reduction factor must be achieved such

that the number of simulations falling below the design value is identically zero.

The ACI strength reduction factors serve as the control factors. Comparisons between ACI

and the second and third approaches are made. The number of simulations falling below the revised

design flexural strength is tabulated.

81

CHAPTER 4 Results

4.1 Introduction

Monte Carlo simulations were performed to model the flexural strength and durability of

GFRP reinforced concrete beams. Simulations were based on varying the parameters involved in

calculating the flexural capacity. The modeled long-term tensile strength of GFRP bars was used to

assess fitting environmental reduction factors. Flexural strength reduction factors were computed

and compared.

4.2 Environmental Reduction Factors

Prior to beginning the full-scale Monte Carlo simulation procedure, appropriate

environmental reduction factors were ascertained. Environmental reduction factors were expected

to act as “knock down” factors when applied to initial GFRP bar tensile strengths. The reduction

was expected to be significant enough to such that the newly computed bar tensile strength values

and the projected long-term tensile strength values would be in close proximity to each other. The

mean minus three standard deviations for flexural strength of a GFRP reinforced concrete member

corresponds to a reliability of approximately 0.9890 for all simulations. Figure 4-1 is a representative

illustration, which highlights the initial difference in Weibull-based moment capacities when the

projected 0 and 60 year GFRP tensile strength values are applied.

82

Rel

iabi

lity

Flexural strength [k-in]

Figure 4-1: Comparison of Weibull-based moment capacities for Member A using projected

0 and 60 year GFRP tensile strength values

Eq. 3.70 was used to calculate the environmental reduction factor to convert the initial

strength distribution. A factor of 0.5 was obtained. Figure 4-2, which is a representative illustration,

displays Weibull-based flexural strengths for Member A based on initial, revised, 60-year, and ACI

based GFRP tensile strengths. The revised strength was determined by applying the calculated

environmental reduction factor of 0.5 to the initial GFRP tensile strength. The ACI based

distribution utilizes an environmental reduction factor of 0.7, which is in accordance with current

design code provisions.

All four of the Weibull-based flexural strength distributions have been included on the figure

for comparative purposes. The ACI-based flexural strength distribution does not closely match the

83

projected strength values at 60 years, and provides an overestimate of the strength retention. The

application of the ACI exposure factor could potentially lead to unsatisfactory flexural strength for

the reinforced member over the anticipated service life.

Rel

iabi

lity

Flexural strength [k-in]

Figure 4-2: Comparison of Weibull-based moment capacities for Member A using both 0

year and 60-year projected GFRP tensile strength values. The 0 year revised and ACI

moment capacities have incorporated environmental reduction factors of 0.5 and 0.7,

respectively, to the 0-year GFRP tensile strength.

All of the time-dependent environmental reduction factors were collected and compared to

the ACI values. A representative plot is shown in Figure 4-3. The currently prescribed ACI values

are shown as constant horizontal lines over the service life of the structure. The environmental

reduction factors shown were obtained for all three members at the specific service age level.

84

Concrete strength and beam geometry were determined to have no bearing on the environmental

reduction factor using the outlined approach. Given a life-cycle of 60 years of exposure, a more

appropriate environmental reduction factor would be 0.5. This factor is also taken in to

consideration for the remaining Monte Carlo simulations. The trend of decreasing environmental

reduction factors with time is expected due to the loss of GFRP strength over time.

Aside from the ACI prescribed values, the environmental reduction factors presented in

Figure 4-3 represent GFRP reinforced concrete members that are exposed to earth and weather. It

has been assumed that these reinforced concrete members are subject to nearly 100% relative

humidity. However, it is possible that alternative environmental reduction values exist for situations

in which the FRP reinforced concrete members are designed for other relative humidity conditions.

Such instances would indicate an indoor exposure level, which could produce an environmental

value similar to what ACI currently proposes. This study has not accounted for these instances, and

entirely deals with GFRP reinforced concrete members that are exposed to earth and weather.

85

Env

iron

men

tal r

educ

tion

fact

or [

CE]

Time [years]

Figure 4-3: Environmental reduction factors for Member A designed with 4000 psi concrete

as a function of time.

4.3 Monte Carlo simulation results

Many Monte Carlo simulations were performed during this study. The environmental

reduction factors proposed in Sec. 4.2 were utilized to determine the appropriate projected long-

term GFRP bar tensile strength. The effects of beam geometry, concrete strength, and fractions of

the balanced reinforcement ratio will be considered. Nominal flexural strength, percentage of

simulation failures due to GFRP bar rupture, percentage of simulations with a flexural strength

below the nominal value are used to draw comparisons and contrasts between each of the effects

under consideration.

86

4.3.1.1 Effect of environmental reduction factors

The environmental reduction factors, which were established in Sec. 4.2 , were used in the

Monte Carlo simulations. The factors were used in Eq. 3.26 to determine the design tensile strength.

The flexural strength was then computed for either concrete crushing or FRP bar rupture. Figure

4-4 shows the Weibull mean for nominal flexural strength for Member A with 4000 psi. The error

bars shown are based on the Weibull standard deviation for the modeled flexural strength.

The balanced reinforcement condition is shown. The flexural strength increases with

increasing environmental reduction factors. The trend, as demonstrated by the representative

illustration, is expected due to the direct relationship between the environmental reduction factor

and design tensile strength. The ACI values of 0.7 and 0.8, for exposed and unexposed concrete,

produced greater flexural strength values than those calculated in Sec. 4.2 . The ACI prescribed

environmental reduction values inflate the retained tensile strength, and thus, produces higher than

achievable long-term nominal flexural strength.

87

Wei

bull-

base

d no

min

al fl

exur

al s

tren

gth

[k-i

n]

Environmental reduction factor [CE]

Figure 4-4: Weibull-based nominal flexural strength as a function of environmental

reduction factor for Member A designed with 4000 psi concrete at balanced reinforcement

ratio conditions.

Figure 4-5 shows the percentage of simulations occurring below the nominal flexural

strength as a function of environmental reduction factor. The percentages range from 25 – 50% for

all simulations conducted. The lowest percentage for all simulations occurred for the configurations

designed with a concrete strength 6000 psi and a reinforcement ratio fraction of 1.2.

In general, lower environmental reduction factors lead to greater susceptibility to fail via

FRP bar rupture. The trend can be tied to the underestimation of long-term strength via the

application of the environmental reduction factor. As corroborated by Figure 4-4, lower

environmental reduction factors lead to lower nominal flexural strengths for the GFRP reinforced

concrete members. When an environmental reduction factor is not considered, as is the case when it

88

is equal to 1, the number of simulations falling below the nominal flexural strength slightly increases.

The balanced condition indicates the potential to have simultaneous concrete crushing and FRP bar

rupture. Therefore, the increase may be attributed to the randomness of the simulations, and the

propensity to fail via FRP bar rupture. The result for the configuration is in relative proximity to the

other values.

Per

cent

age

of s

imul

atio

ns b

elow

nom

inal

flex

ural

str

engt

h [%

]

Environmental reduction factor [CE]

Figure 4-5: Percentage of simulations occurring below the nominal flexural strength for

GFRP reinforced concrete members as a function of the environmental reduction factor for

Member A designed with 4000 psi concrete and balanced reinforcement ratio conditions.

The only two failure modes under consideration are FRP bar rupture and concrete crushing.

The third area of analysis pertains to the number of simulations that failed due to FRP bar rupture.

In a mathematical sense, one of the failure modes will always precede the other mode. However,

89

physically, there are many instances, especially around the region of balanced reinforcement ratio, in

which both failure modes due successfully occur simultaneously. Figure 4-6 is shown as a

representative illustration. The projected environmental reduction values produced yielded all

simulations failing due to FRP rupture. The design tensile strength is nearly halved by the

application of the most of the computed factors. It is expected that the reduction would adversely

impact the strength of the member and lead to more bar ruptures. The lowest percentages achieved

in this area were for the simulations in which the environmental reduction factor was not included.

If failure due to FRP bar rupture is desired, the proposed factors succeed in guaranteeing the

preferred failure mode.

Per

cent

age

of F

RP

bar

rup

ture

failu

res

[%]

Environmental reduction factor [CE]

Figure 4-6: Percentage of failures out of 10,00 simulations due to FRP bar rupture for

Member A designed with 4000 psi concrete and balanced reinforcement ratio conditions.

90

4.3.1.2 Effect of beam geometry

Three different member sizes were considered for the simulations. Members A and C are

analyzed for comparative purposes. The reinforcing area and fractions of reinforcement ratio will be

compared.

Figure 4-7 and Figure 4-8 show the reinforcing bar area as a function of the fraction of

balanced reinforcement ratio for Members A and C, respectively. The initial constant values for

reinforcing area is due to the minimum area requirements imposed by Eq. 3.23. The minimum area

is invoked for fractions of reinforcement ratio ranging from 0.1 – 0.7. The total reinforcing bar area

is dependent upon the member width, b, and the distance from extreme compression fiber to the

centroid of tension reinforcement, d. It is expected that the larger geometry of Member C will

require more FRP bars for reinforcing. As shown by the figures, the effect of geometry on the area

is minimized by an increase of the concrete strength. The required reinforcing linearly increases with

fractions of reinforcement ratio for all other values.

91

GF

RP

rei

nfor

cing

bar

are

a [i

n2 ]

Fraction of balanced reinforcement ratio

Figure 4-7: Total GFRP reinforcing bar area for Member A with respect to varying concrete

strength and fractions of balanced reinforcement ratio.

92

GF

RP

rei

nfor

cing

bar

are

a [i

n2 ]

Fraction of balanced reinforcement ratio

Figure 4-8: Total GFRP reinforcing bar area for Member C with respect to varying concrete

strength and fraction of balanced reinforcement ratio.

Figure 4-9 compares the nominal flexural strength as a function of reinforcing ratio for

members A and C. The balanced conditions with a concrete strength of 4000 psi were used the

illustration shown. There is a kink in the data for both of the members. The kink coincides with the

transition from relying on the minimum reinforcing to utilizing reinforcing areas based on the

applied fraction of the balanced reinforcement ratio. There is a tremendous gap in the nominal

strength achieved by the two members. The gap can be attributed to the significantly larger moment

arm produced by the dimensions of Member C. Namely, the depth of Member C is twice as large as

the depth of Member A.

93

Nom

inal

flex

ural

str

engt

h [k

-in]

Total GFRP reinforcing area [in2]

Figure 4-9: Nominal flexural strength for GFRP reinforced concrete members with respect

to total GFRP reinforcing bar area for Members A and C. The illustrations shown are used

to draw contrast to different beam geometries of the different members.

The flexural strength of a FRP reinforced member is both directly and indirectly dependent

upon the beam geometry. As seen by Eqs. 3.35 and 3.36, for concrete crushing failure, flexural

strength will increase with larger distances from extreme compression fiber to the centroid of

tension reinforcement. Larger d values can be compensated by utilizing a larger beam width. The

beam geometry is also an integral component of in determining the constants of the concrete design

quadratic. It is expected that larger d and b values will produce bigger flexural strength values.

Figure 4-10 and Figure 4-11 show the nominal flexural strength for Members A and C as a

function of the fraction of balanced reinforcement ratio. All environmental reduction factors under

94

consideration have been plotted. The initial constant values for the flexural strength are present due

to the aforementioned consequence of the minimum FRP reinforcement. The flexural strength for

member A ranges from approximately 800 – 3000 k-in. Conversely, the flexural strength range for

Member C ranges from 5000 – 19000 k-in. The maximum value achieved by Member A is barely

60% of the minimum value for Member C.

Nom

inal

flex

ural

str

engt

h [k

-in]

Fraction of balanced reinforcement ratio

Figure 4-10: Nominal flexural strength for GFRP reinforced concrete Member A with

respect to environmental reduction factor and fraction of balanced reinforcement ratio. The

member was designed with 4000 psi concrete.

95

Nom

inal

flex

ural

str

engt

h [k

-in]

Fraction of balanced reinforcement ratio

Figure 4-11: Nominal flexural strength for GFRP reinforced concrete member C with respect

to environmental reduction factor and fraction of balanced reinforcement ratio. The member

was designed with 4000 psi concrete.

Figure 4-12 and Figure 4-13 display the percentage of failures due to FRP bar rupture for

both members A and C. For the information shown, the members were designed with 4000 psi. All

fractions of reinforcement ratio as well as environmental reduction factors are demonstrated. One

hundred percent of all simulations for both cases up to 0.8 of the reinforcement ratio were

determined to fail due to FRP bar rupture. The declines in the percentages have been attributed to

the environmental factors, and not the increase in geometry. Therefore, and although Member C is

significantly larger than Member A, it is equally prone to the FRP bar rupture failure mechanism.

96

Per

cent

age

of s

imul

atio

n fa

ilure

s du

e to

FR

P b

ar r

uptu

re [

%]

Fraction of balanced reinforcement ratio

Figure 4-12: Percentage of simulation failures due to FRP bar rupture for Member A

designed with 4000 psi concrete with respect to fractions of balanced reinforcement ration

and environmental reduction factor. All environmental reduction factors were subject to

100% FRP bar rupture failure except for factors of 0.8 and 1.0.

97

Per

cent

age

of s

imul

atio

n fa

ilure

s du

e to

FR

P b

ar r

uptu

re [

%]

Fraction of balanced reinforcement ratio

Figure 4-13: Percentage of simulation failures due to FRP bar rupture for Member C design

with 4000 psi concrete and with respect to fractions of balanced reinforcement ration and

environmental reduction factor. All environmental reduction factors were subject to 100%

FRP bar rupture failure except for the factors of 0.8 and 1.0.

4.3.1.3 Effect of concrete strength

The minimum reinforcing area is dependent on the beam geometry and concrete strength, as

per Eq. 3.23. The effects of concrete strength on the nominal flexural strength and the percentage of

failures due to FRP bar rupture will be examined in this section. For illustrative purposes, the figures

shown have been based on data from Member A with no environmental reduction factor applied.

Figure 4-14 displays the nominal flexural strength versus the fraction of balanced

reinforcement ratio. The results for 4000, 5000, and 6000 psi concrete are shown. The initial

expectation was that higher concrete strengths would produce larger nominal flexural strength

98

values. The trend is exemplified by the figure. Greater concrete strengths lead to the minimum

reinforcement being considered over a smaller range due to its indirect relationship. It has been

corroborated Figure 4-10 and Figure 4-10. The concrete strength also impacts the β1 value, which is

based on the concrete strength and is an important factor in reinforced concrete designed for

flexure.

Nom

inal

flex

ural

str

engt

h [k

-in]

Fraction of balanced reinforcement ratio

Figure 4-14: Nominal flexural strength for GFRP reinforced concrete Member A with

respect to fractions of balanced reinforcement ratio and concrete strength.

Figure 4-15 compares the percentage of failures attributed to FRP bar rupture for the three

concrete strengths. All three concrete strengths behaved similarly for fractions of reinforcement

ratio up to approximately 0.8. Over this range, all simulations failed due to FRP bar rupture. The

99

results are consistent with effects on rupture from beam geometry and reinforcement area.

However, 4000 psi concrete produced fewer FRP bar rupture simulation failures at a lower fraction

of balanced reinforcement ratio. For the over-reinforced sections, which should be more prone to

concrete crushing by definition, the 5000 and 6000 psi behaved identically and ultimately failed to

reduce the vulnerability to FRP bar rupture failure.

Per

cent

age

of s

imul

atio

n fa

ilure

s du

e to

FR

P b

ar r

uptu

re [

%]

Fraction of balanced reinforcement ratio

Figure 4-15: Percentage of simulation failures due to bar rupture with respect to fraction of

balanced reinforcement ratio for Member A designed with 4000, 5000, and 6000 psi concrete.

Values of 100% indicate all 10,000 simulations experienced this failure mode.

100

4.4 Strength reduction factors

Four methods were used to determine flexural strength reduction factors. The first method

is based current ACI code provisions. The second and third methods are based on an LRFD derived

equation by Ellingwood for two different factors of safety. The final method is based on using brute

force calculations. The figures shown are representative of the trends demonstrated by the three

concrete members for all three concrete strengths. For illustrative purposes, the results of member A

with a concrete strength of 4000 psi are displayed.

The ACI flexural strength reduction factors were determined for each simulation using Eq.

3.38. Figure 4-16 displays the relationship between fractions of the reinforcing ratio and the ACI

strength reduction values. The strength reduction factor has a constant value for fractions of the

balanced reinforcement ratio less than one. For other fractions of the balanced reinforcement ratio,

the flexural strength reduction increases as expected, which is in accordance with the ACI guidelines.

Figure 4-16 also extends the fraction of balanced reinforcement ratio beyond what has been

considered for the Monte Carlo simulations. The extension was used to verify the constant strength

reduction factor of 0.65 for reinforcement conditions greater than 1.4 times the balanced

reinforcement ratio. The FRP bar rupture governed simulations are subject to more strength

reduction than members who fail via concrete crushing. There were no simulations achieving a

flexural strength less than the ACI-based design value.

101

Fle

xura

l str

engt

h re

duct

ion

fact

or

Fraction of balanced reinforcement ratio

Figure 4-16: ACI-based flexural strength reduction factors with respect to fractions of the

balanced reinforcement ratio. 1.5 and 1.6 times the balanced reinforcement ratio are also

presented to verify the strength reduction factor of 0.65. Neither values were directly

employed in the Monte Carlo simulations.

Figure 4-17 and Figure 4-18 show the flexural strength reduction factors for the safety

factors of 3.5 and 4.0, respectively. The factors were computed using Eqs. 3.73 and 3.74. Due to

similar performance among the simulations for reinforcement ratios of less than 0.6 times the

balanced reinforcement ratio, it is expected that the flexural strength factors would be constant over

that reinforcement ratio range. The Weibull-based mean, standard deviation, and coefficient of

variation for those simulations are identical for a given environmental reduction factor. Hence, the

flexural strength reduction factors are equivalent for that range of reinforcement ratios. The factors

102

also slightly increase as the fraction of balanced reinforcement ratio increases, which is expected due

to modest increases of the mean, standard deviation, and coefficient of variation values.

The strength reduction factors often create further conservative designs when using FRP as

a reinforcing material for concrete. These factors decrease with increasing environmental reduction

factors. The environmental reduction factors reduce FRP tensile strength, and, subsequently, the

overall flexural strength of the structure. Therefore, larger tensile strength values reduce the

susceptibility of FRP bar rupture failure, and promote the concrete crushing failure mode. As the

reinforcement ratio is increased such that concrete crushing should occur, the flexural strength

reduction factors also increase due to greater levels of FRP bar reinforcement in concrete member.

103

F

lexu

ral s

tren

gth

redu

ctio

n fa

ctor

Fraction of balanced reinforcement ratio

Figure 4-17: Flexural strength reduction factors based on Ellingwood with β = 3.5 for

Member A designed with 4000 psi concrete. Similar results were reported throughout the

simulations.

104

Fle

xura

l str

engt

h re

duct

ion

fact

or

Fraction of balanced reinforcement ratio

Figure 4-18: Flexural strength reduction factors based on Ellingwood with β = 4.0 for

Member A designed with 4000 psi concrete. As with the β = 3.5, similar results were

recorded amongst all simulations.

Figure 4-19 provides a representative comparison between the four methods used to

determine flexural strength reduction factors. The factors determined by ACI and via the brute force

method yield zero simulations that with flexural strength values less than the design value.

Furthermore, using Ellingwood’s approach with a factor of safety of 4.0 generates only 1 simulation

(0.01%) falling below the design flexural strength value. Thus, The ACI-based values produce a

conservative estimate of the design strength when compared to the other methods. The brute force

method values are the maximum values in which the number of simulations with a flexural strength

105

less than the design value occurs. A more appropriate value would be closer to 0.80, which is

approximately the strength reduction value obtained via the brute force method. F

lexu

ral s

tren

gth

redu

ctio

n fa

ctor

Fraction of balanced reinforcement ratio

Figure 4-19: Comparison of flexural strength reduction factors for all methods under

consideration. The current values provided by ACI yield the most conservative design

estimates for the moment capacity of GFRP reinforced concrete beams.

106

CHAPTER 5 Conclusions

5.1 Introduction

The chief objective of this study was to assess the long-term reliability of GFRP reinforced

concrete members in flexure. The secondary goal was to determine appropriate environmental and

strength reduction factors for moment capacity. The study incorporates the application of the

power-law model and the Monte Carlo simulation technique to GFRP bars.

5.2 Conclusions

5.2.1 Environmental re duc tion fac to rs

The current ACI 440 code’s environmental reduction factors may not be appropriate for

concrete structures exposed to the environment. Based on the 60-year GFRP tensile strength

retention model generated, and through reliability analysis, it is recommended that an environmental

reduction factor 0.5 be used for concrete exposed to earth and weather. The revised factor takes in

to account the loss of GFRP tensile strength over time and provides a conservative, yet appropriate,

estimate for flexural design. The currently prescribed ACI values for environmental exposure alone

do not provide a significant reduction to account for long-term reinforcing bar tensile strength

reduction.

5.2.2 Monte Carlo s imula ti on parame ters

Flexural strength of FRP reinforced concrete beams is comprised of the tensile strength of

the FRP reinforcing bars and the moment arm generated by the distance between the neutral axis

the reinforcing bars. The tensile strength is a function of the total reinforcing in the concrete

107

member and the reinforcement ratio. Greater values of either component lead to increases in the

flexural strength. However, the application of the environmental reduction factors can lead to a loss

in strength for which other components of the beam geometry may compensate. The moment arm

calculation for concrete structures primarily relies on the depth of the member. Greater depths also

lead to tremendous increases in the flexural capacity of the members. Enlarged member dimensions

were found to have no positive effect on reducing the number of simulations that contributed to

failure of the GFRP bars due to rupture.

5.2.3 Streng th reduc ti on fac to rs

Strength reduction factors are intrinsically based on the reinforcement ratio. The current code

provisions produce a conservative estimate for the design flexural strength, and the factors range

from 0.55 – 0.65. It has been shown that an increase in the strength reduction value could be made

without reducing the number of reinforced concrete members subject to premature flexural failure.

A more appropriate strength reduction factor is approximately 0.80, which does indirectly

incorporate the environmental effects.

5.3 Recommendations for future work

Although many simulations were performed for this study, there is a still a tremendous need

for future work. The following are recommendations for additional research:

1. The strength reduction factors generated from Ellingwood’s equation could benefit

from the application of the Hasofer-Lind safety index and its iterative process. More

accurate values for the safety index and sensitivity for the Monte Carlo simulation

should be developed using this process.

108

2. Long-term data must be obtained combining temperature and relative humidity

effects into a single study. If possible, weight gain measurements in addition to

tensile strength tests should be performed on all sample specimens. The assumption

of the power-law model for long-term testing extrapolation needs to be validated.

3. Additional simulations should be performed assuming a 100-year service life for a

bridge deck member. Environmental and strength reduction factors for this time

range.

4. Further simulations should be performed on carbon and aramid fiber reinforced

structures to assess the validity of the ACI environmental reduction factor guidelines.

Comparisons should be made regarding its performance under the use of the

strength reduction factor.

109

REFERENCES

[1] Einde, L. Van Den, L. Zhao, and F. Seible, Use of FRP Composites in Civil Structural

Applications. Construction and Building Materials, 2003. 17: p. 389 - 403.

[2] Uomoto, Taketo, et al., Use of Fiber Reinforced Polymer Composites as Reinforcing Material for Concrete. Journal of Materials in Civil Engineering, 2002. 14(3): p. 191 - 209.

[3] Nystrom, Halvard E., et al., Financial Viability of Fiber-Reinforced Polymer (FRP) Bridges. Journal of Management in Engineering, 2003. 19(1): p. 2-8.

[4] Sahirman, Sidharta, Robert C. Creese, and Bina R. Setyawati, Evaulation of the Economic Feasibility of Fiber-Reinforced Polymer (FRP) Bridge Decks, in ISPA/SCEA International Joint Conference. 2003: Orlando, Florida.

[5] Nkurunziza, Gilbert, et al., Durability of GFRP Bars: A Critical Review of the Literature. Progress in Structural Engineering and Materials, 2005. 7(4): p. 194 - 209.

[6] Mukherjee, Abhijit and S.J. Arwikar, Performance of Glass Fiber-Reinforced Polymer Reinforcing Bars in Tropical Environments - Part II: Microstructural Tests. ACI Structural Journal, 2005. 102(6): p. 816 - 822.

[7] Almusallam, Tarek H. and Youself A. Al-Salloum, Durability of GFRP Rebars in Concrete Beams under Sustained Loads at Severe Environments. Journal of Composite Materials, 2006. 40(7): p. 623 - 637.

[8] Karbhari, V.M., et al., Durability Gap Analysis for Fiber-Reinforced Polymer Composites in Civil Infrastructure. Journal of Composites for Construction, 2003. 7(3): p. 238 - 247.

[9] Nishizaki, Itaru and Seishi Meiarashi, Long-Term Deterioration of GFRP in Water and Moist Environment. Journal of Composites for Construction, 2002. 6(1): p. 21 - 27.

[10] Pauchard, V., et al., In-situ Analysis of Delayed Fibre Failure within Water-aged GFRP under Static Fatigue Conditions. International Journal of Fatigue, 2002. 24: p. 447 - 454.

[11] Micelli, Francesco and Antonio Nanni, Durability of FRP Rods for Concrete Structures. Construction and Building Materials, 2004. 18(7): p. 491 - 503.

[12] Vijay, P.V., H.V.S. GangaRao, and Rajesh Kalluri. Hygrothermal Response of GFRP Bars Under Different Conditioning Schemes. in Durability of Fibre Reinforced Polymer Composites for Construction. 1998. Sherbrooke, Canada.

[13] Mula, S., et al., Effects of Hygrothermal Aging on Mechanical Behavior of Sub-zero Weathered GFRP Composites. Journal of Reinforced Plastics and Composites, 2006. 25(6): p. 673 - 680.

[14] Bhise, Vikrant, Strength Degradation of GFRP Bars, in Via Department of Civil and Environmental Engineering. 2002, Virginia Polytechnic Institute and State University: Blacksburg, VA.

110

[15] 440, American Concrete Institute Committee, ACI 440.1R-06: Guide for the Design and Construction of Structural Concrete Reinforced with FRP Bars. 2006, American Concrete Institute: Farmington Hills, MI.

[16] Nanni, Antonio, Flexural Behavior and Design of RC Members Using FRP Reinforcement. Journal of Structural Engineering, 1993. 119(11): p. 3344 - 3359.

[17] Pilakoutas, Kypros, Kyriacos Neocleous, and Maurizio Guadagnini, Design Philosophy Issues of Fiber Reinforced Polymer Reinforced Concrete Structures. Journal of Composites for Construction, 2002. 6(3): p. 154 - 161.

[18] Myers, John J. and Thara Viswanath, A Worldwide Survey of Environmental Reduction Factors for Fiber-Reinforced Polymers (FRP). Structures 2006, 2006.

[19] Ellingwood, Bruce R., Load and Resistance Factor Design (LFRD) for Structures Using Fiber-Reinforced Polymer (FRP) Composites, U.S.D.o. Commerce, Editor. 2000, National Institutes of Standards and Technology. p. 52.

[20] Chen, Yi, Julio F. Devalos, and Indrajit Ray, Durability Prediction for GFRP Reinforcing Bars Using Short-Term Data of Accelerated Aging Tests. Journal of Composites for Construction, 2006. 10(4): p. 279 - 286.

[21] Katsuki, F. and T. Uomoto, Prediction of Deterioration of FRP Rods Due to Alkali Attack, in Non-metallic (FRP) Reinforcement for Concrete Structures, L. Taerwe, Editor. 1995, E & FN Spon: London. p. 82 - 89.

[22] Won, Jong Pil, Yong Chin Cho, and Chang Il Jang, The Durability of Glass Fibre-Reinforced Polymer Dowel after Accelerated Environmental Exposure. Polymers and Polymer Composites, 2006. 14(7): p. 719 - 730.

[23] Hughes, Brothers, Glass Fiber Reinforced Polymer (GFRP) Rebar Aslan 100. 2002, Hughes Brothers, Inc.: Seward, Nebraska. p. 20.

[24] Momose, Yutaka, et al. Effects of moisture on glass fiber reinforced composites. in Proceedings of the 1999 44th International SAMPE Symposium and Exhibition 'Envolving and Revolutionary Technologies for the New Millennium', SAMPE '99. 1999. Long Beach, CA.

[25] Okeil, Ayman M. and Sujata Kulkarni, Flexural Resistance Models for Concrete Decks Reinforced with Fiber-Reinforced Polymer Bars. Transportation Research Record, 2006. 1976: p. 190 - 196.

[26] Nowak, Andrzej S. and Maria M. Szerszen, Calibration of Design Code for Buildings (ACI 318): Part 1 - Statistical Models for Resistance. ACI Structural Journal, 2003. 100(3): p. 377 - 382.

[27] Okeil, Ayman M., Sherif El-Tawil, and Mohsen Shahawy, Flexural Reliability of Reinforced Concrete Bridge Girders Strengthened with Carbon Fiber-Reinforced Polymer Laminates. Journal of Bridge Engineering, 2002. 7(5): p. 290- 299.

[28] Bain, Lee J. and Max Engelhardt, Statistical analysis of reliability and life-testing models: theory and methods. Second ed. 1991, New York, NY: Marcel Dekker, Inc. 496.

111

[29] Dodson, Bryan, The Weibull Analysis Handbook. Second ed. 2006, Milwaukee, WI: ASQ Quality Press. 167.

[30] Murthy, D.N. Prabhakar, Min Xie, and Renyan Jiang, Weibull Models. 2004, Hoboken, NJ: John Wiley & Sons, Inc. 383.

[31] Alqam, Maha, Richard M. Bennett, and Abdul-Hamid Zureick, Three-parameter vs. two-parameter Weibull distribution for pultruded composite material properties. Composite Structures, 2002. 58(4): p. 497 - 503.

112