You are on page 1of 81

Service Behaviour of Reinforced Concrete Members

A thesis submitted in partial fulfillment of the requirements for the degree of Bachelor of Engineering in Civil Engineering (Structures)

John van Rooyen 309243947

Supervisor: Associate Professor Gianluca Ranzi

School of Civil Engineering University of Sydney, NSW 2006 Australia November 2012

Disclaimers

Student Disclaimer The work comprising this thesis is substantially my own, and to the extent that any part of this work is not my own I have indicated that it is not my own by acknowledging the source of that part or those parts of the work. I have read and understood the University of Sydney Student Plagiarism: Coursework Policy and Procedure. I understand that failure to comply with the University of Sydney Student Plagiarism: Coursework Policy and Procedure can lead to the University commencing proceedings against me for potential student misconduct under chapter 8 of the University of Sydney By-Law 1999 (as amended).

Departmental Disclaimer This thesis was prepared for the School of Civil Engineering at the University of Sydney, Australia, and describes the time dependent behaviour of reinforced concrete. The opinions, conclusions and recommendations presented herein are those of the author and do not necessarily reflect those of the University of Sydney or any of the sponsoring parties to this project.

ii

Table of contents

Table of contents ........................................................................................................................ iii Acknowledgements ...................................................................................................................... v Abstract ...................................................................................................................................... v Chapter summary ....................................................................................................................... vi List of tables and figures ........................................................................................................... vii Nomenclature ............................................................................................................................. xi Chapter 1 Introduction .............................................................................................................. 1 1.1. 1.2. General ......................................................................................................................... 1 Objectives ..................................................................................................................... 2

Chapter 2 Literature review ........................................................................................................ 3 2.1. 2.2. 2.3. 2.4. 2.5. 2.6. General ......................................................................................................................... 3 Shrinkage ...................................................................................................................... 3 Compressive creep ........................................................................................................ 7 Tensile creep ................................................................................................................. 9 Tensile strength .......................................................................................................... 10 Modelling time dependent behaviour .......................................................................... 10

Chapter 3 Time dependent behaviour in concrete ..................................................................... 12 3.1. 3.2. 3.3. Time dependent properties ......................................................................................... 12 Time dependent modelling step by step method ..................................................... 17 SSM assumptions ........................................................................................................ 20

Chapter 4 Cross sectional analysis ............................................................................................ 21 4.1. 4.2. 4.3. 4.4. 4.5. Background ................................................................................................................. 21 Uncracked formulation ............................................................................................... 21 Cracked formulation ................................................................................................... 22 Uncracked example layered approach ...................................................................... 24 Cracked example layered approach.......................................................................... 26 iii

4.6.

Comparison of results ................................................................................................. 27

Chapter 5 Finite element method.............................................................................................. 28 5.1. 5.2. 5.3. 5.4. 5.5. 5.6. 5.7. 5.8. 5.9. 5.10. 5.11. 5.12. 5.13. Assumptions and comments ....................................................................................... 28 Formulation ................................................................................................................ 29 Degrees of freedom and consistency ............................................................................ 29 Time dependency ........................................................................................................ 30 Transformation from local to global axes ................................................................... 31 Shrinkage .................................................................................................................... 31 Cracking ..................................................................................................................... 32 Gaussian quadrature ................................................................................................... 33 Programming .............................................................................................................. 34 Cracked example ..................................................................................................... 35 Uncracked validation .............................................................................................. 36 Cracked validation .................................................................................................. 39 AS3600-2009 comparison......................................................................................... 45

Chapter 6 Measurement of shrinkage profiles............................................................................ 47 6.1. 6.2. Previous techniques .................................................................................................... 47 Development of new sensors ....................................................................................... 48

Chapter 7 Conclusion ................................................................................................................ 50 Appendix A Comparison of cross sectional methods to analyse time dependent behaviour ...... 51 A.1. A.2. Constant Deformation ................................................................................................ 51 Constant load ............................................................................................................. 56

Appendix B Step by step cross sectional analysis formulation .................................................. 59 Appendix C Finite beam element formulation .......................................................................... 62 C.1. C.2. Displacement field ...................................................................................................... 62 Weak Formulation ...................................................................................................... 62

Appendix D Matlab finite element program.............................................................................. 65 References.................................................................................................................................. 67

iv

Acknowledgements
Thank you to my supervisor, Gianluca Ranzi, who was passionate, generous with his time and gave me the freedom to explore avenues of interest in this thesis. Thank you to my parents for always supporting me. Finally thank you to my wife, Carly, whose support and patience for my obsession allowed me this far.

Abstract
The main objective of this thesis is to predict the long term behaviour of reinforced concrete. To this end, a method of cross sectional analysis (based on the step by step method) is developed using a layered approach to model time dependent behaviour, including cracking, in beams under axial and/or bending loads. Calculated strains from this model are shown to agree with results from the literature. The cross sectional method is extended to a finite element framework, and a formulation for beam elements incorporating time dependent effects is presented. This formulation is implemented in Matlab, and calculated deflections are shown to agree with cracked and uncracked experiments on beams. The assumed shrinkage profile used to predict time dependent behaviour is explored, and nonuniform, curved shrinkage profiles are shown to significantly change calculated deflection by inducing cracked behaviour. As a result of the above finding, the measurement of shrinkage profiles is explored, and a humidity sensor developed. Finally, a comparison is made between results calculated by the refined FEM method, and the simplified method provided in AS3600-2009. It is suggested that improved accuracy in refined methods must be weighed up against complexity and additional time requirements.

Chapter summary
Chapter 1 provides a rationale behind the thesis and outlines its key objectives Chapter 2 describes current knowledge behind key time dependent properties including creep and shrinkage, compares conflicting research and identifies some gaps. It also describes some modelling considerations raised in the literature. Chapter 3 outlines the key time dependent properties in concrete, and introduces the step by step method for modelling time dependent behaviour. Chapter 4 shows the development of a cross sectional method of analysis based on the step by step method that considers axial loading and bending in cracked and uncracked sections. The method is validated against results in the literature. Chapter 5 presents a finite element formulation based on the step by step method outlined in chapter 4. Implementation of the formulation in Matlab is described. Experimental results are compared to results from the model, using uniform and non-uniform shrinkage profiles. Chapter 6 describes the development of a sensor to measure humidity in concrete as a means to identify the shrinkage profile. Chapter 7 outlines the conclusions of this thesis.

vi

List of tables and figures


Figure 2.1: Relative magnitudes of drying and autogenous shrinkage Figure 2.2: Meniscus that forms as a result of evaporation of bleed water Figure 2.3: Forces acting on the meniscus Figure 2.4: Electron microscope images of the formation of plastic shrinkage cracking Figure 2.5: Comparison of creep in a sealed specimen (left) and creep in a drying specimen (right) Figure 3.1: Development of shrinkage with time Figure 3.2: Shrinkage without restraint Figure 3.3: Free shrinkage strains Figure 3.4: Concrete under creep with no shrinkage Figure 3.5: Creep coefficient vs time Figure 3.6: Generalised concrete compressive strength development over time for normal strength concrete relative to fc(28) Figure 3.7: Stress strain curve for concrete Figure 3.8: Stress strain curves for varying strengths of concrete Figure 3.9: Discreet stress intervals used in the SSM Figure 3.10: Strain in a beam under axial and bending loads where plane sections remain plane Figure 4.1: Cross section divided into layers Figure 4.2: Cross Section for example (all units in mm) Figure 5.1: Assumed axis system Figure 5.2: 7 degree of freedom beam element Figure 5.3: Relationship between local element axis and global axis Figure 5.4: Sampling points and weights for a cubic Figure 5.5: Results from FEM for first time step Figure 5.6 Cross sections for Slabs A and B used for long term deflection tests Figure 5.7: Support and loading conditions for long term deflection tests Figure 5.8: Shrinkage results from test cylinders 3 4 4 4 7 12 13 13 14 15 16 16 16 17 19 23 24 28 29 31 33 35 36 36 37

vii

Figure 5.9: Creep coefficients calculated from cylinder tests Figure 5.10: Comparison of mid-span deflection as measured by experiment and by FEM calculation Figure 5.11: Comparison of mid-span strains as measured by experiment and by FEM calculation for Slabs A and B

37 38

39

Figure 5.12: Slab cross section, and support and loading conditions for long term cracked tests (dimensions in mm) Figure 5.13: Beam cross section and support and loading conditions for long term cracked tests (dimensions in mm) Figure 5.14: Progression shrinkage strain profiles over time assumed for the validation test.

39

40

42

Figure 5.15: Comparison of free shrinkage stresses as a result of non-uniform shrinkage strains. The left hand side shows those measured by experiment and the right hand side, those calculated by the FEM program based on an assumed shrinkage profiles.

42

Figure 5.16: Progression of the relative humidity profile over the first 7 days of curing for a concrete prism Figure 5.17: Comparison of deflection s as calculated by the FEM model and as measured by experiment for the beam

43

43

Figure 5.18: Comparison of deflections as calculated by the FEM model and as measured by experiment for the slab Figure 5.19: Comparison of cracking for the one way slab as calculated and as recorded by experiment at 400 days Figure 6.1: Measuring RH by measuring electrical conductivity Figure 6.2: Measuring RH in sealed cavities by direct measurement Figure 6.3: Measuring RH in sealed cavities by direct measurement Invalid source specified. Figure 6.4: Circuit board design for the sensor Figure 6.5: Data logger used to connect to humidity sensors Invalid source specified. Figure 6.6: Finished humidity sensor

44

44 47 47 48 48 49 49 viii

Figure A.1: Concrete beam subject to shrinkage and creep Figure A.2: Concrete stresses over time as a result of immediate and sustained shrinkage Figure A.3: Concrete stresses over time calculated using EMM Figure A.4: Concrete stresses over time calculated using AEMM Figure A.5: Creep assumptions in the RCM Figure A.6: Concrete stress over time as a result of instant and constant shrinkage Figure A.7: Strain over time as a result of instant and constant shrinkage Figure A.8: concrete column subject to constant load and creep Figure A.9: Axial strain over time as calculated by each cross sectional method Figure A.10: Concrete stress over time as calculated by each cross sectional method Figure A.11: Steel stress over time as calculated by each cross sectional method Figure C.1: Admissible displacement field under the Euler-Bernoulli beam assumptions Figure C.2: Generalised beam loading Figure D.1: Main GUI input for FEM Figure D.2: Properties GUI input for FEM Table 4.1: Loading, elastic modulus, shrinkage parameters and times steps for example Table 4.2: Creep coefficients for example Table 4.3: Comparison of strains for cracked and uncracked methods Table 6.1: Gauss-Legendre sampling points and weights Table 5.2: Results from FEM Table 5.3: Cross sectional results from FEM, matching those of a cross sectional analysis. Table 5.4: Creep coefficients and shrinkage strain values Table 5.5: Tensile strength and modulus of elasticity values

51 52 53 54 54 55 55 56 57 57 57 62 63 65 66 24 24 27 33 35 36 40 40

ix

Table A.1: Creep and shrinkage values calculated as per AS 3600-2009 Table A.2: Elastic modulus and shrinkage values used in constant loading example

51 56

Nomenclature

, ,

Area, first moment of area, second moment of area, respectfully, calculated about the reference axis Area, first moment of area, second moment of area, respectfully, for concrete calculated about the reference axis Area, first moment of area, second moment of area, respectfully, for steel calculated about the reference axis Area of tension steel, area of compressive steel. Matrix of cross sectional (geometric) properties Distance from top of section to reference axis Elastic modulus of concrete at time Elastic modulus of steel Creep loading vector at time Shrinkage loading vector at time Creep factor at time for stresses applied at Characteristic compressive strength of concrete Characteric flexural tensile strength of concrete Cracked and uncracked second moments of area Effective second moment of area after cracking Creep function representing elastic and creep strain per unit of stress Long-term to short term deflection factor Element stiffness matrix Factor in AS3600-2009 adjusting creep for age at loading Length of span Moment Cracking moment Design service moment Internal axial force and bending moment resisted by the concrete External force and moment applied to the cross section Internal force and moment in the cross section Cross sectional rigidities at time Vector of external actions

, ,

, , , d_ref () , , ,, . , ( , ) , , , , , , , ,

xi

t Z

Vector of internal actions Time Vector of nodal displacements Deflection Uniformly distributed load Distance from the reference axis to the neutral axes Section modulus Concrete strain Creep strain Elastic strain Strain at the reference axis Shrinkage strain Final design shrinkage strain Curvature Tensile web reinforcement ratio and compressive web reinforcement ratio Stress in the concrete at time step j Shrinkage induced tensile stress in the concrete Creep coefficient at time t, for loads applied at time Ageing coefficient at time t, for loads applied at time

, , ( , ) ( , )

xii

Chapter 1 Introduction
1.1. General
Reinforced concrete is widely used in the construction of high rises, bridges, floor slabs, pipes and other structures. Compared to other methods of construction, it is low cost, durable, and widely available. In the design of reinforced concrete structures, self-weight can be a significant component of total load. Sustained long term loads, such as self-weight, lead to deformation in concrete which occurs gradually over time, in addition to that which occurs when the load is first applied. These time-based deformations are not insignificant. For example, it is not uncommon for deflection in a simply supported beam to double over a period of one year. Over 30 years, deflections can be 2.5 times those occurring instantaneously. In addition to loading and material considerations such as those above, deflection also depends on span and cross section. Trends in building design have required increased spans and thinner cross sections, a result of a combination of developers wanting to maximise building floor space and minimise storey heights, and architects pushing the limits of concrete design. Deflections are therefore often critical in concrete design. That is, the design will be governed by serviceability rather than strength. Rigorous methods to calculate deflection, however, are not well understood or widely used by practicing engineers (Ranzi & Gilbert 2011). The basis for any rigorous method to predict deflection is the interaction of creep and shrinkage, both time dependent properties of concrete, and the inclusion of cracking which considerably reduces the stiffness of a member and increases deflection in flexural members.

1.2. Objectives
This thesis seeks to predict the behaviour of reinforced concrete members over time under service loading using numerical models. It is the intention this work will provide some insight into the effect the shrinkage profile has on long term behaviour. Specific objectives are as follows: 1. To develop a cross sectional method of analysis that incorporates time dependent behaviour including cracking. 2. To develop a finite element program that will evaluate deflections in cracked and uncracked beams. 3. To assess the impact of the assumed shrinkage profile on calculated deflection. 4. To compare the simplified method for calculating long term deflection given by AS36002009 with experimental and finite element results. 5. To develop a method to measure the humidity profile through a cross section which can be used as a proxy for the shrinkage profile.

Chapter 2 Literature review


2.1. General
Time dependent behaviour in concrete is a result of the interaction of creep, shrinkage, elasticity, and tensile strength. These properties change over time, and of particular interest is their development at early ages as this has significant bearing on cracking. This literature review will focus on shrinkage, compressive creep, tensile creep and early age behaviour, in that order.

2.2. Shrinkage
Shrinkage can be divided into four categories: plastic, drying, autogenous, carbonation and thermal shrinkage. At early ages, the concrete goes through three phases particulate suspension, skeleton formation and initial hardening (Nehdi & Soliman 2011). While the concrete is wet and acts as a fluid (with particles in suspension) it may be subject to plastic shrinkage, but as soon as the skeleton is formed drying, autogenous and thermal shrinkage occurs. At a glance, drying shrinkage is the result of a loss of water, autogenous shrinkage a result of chemical reactions taking place, and thermal shrinkage a consequence of temperature changes that come about from the exothermic reactions taking place. Relative magnitudes for normal strength concrete are shown in figure 2.1.

Shrinkag

Drying shrinkage Autegenous shrinkage

Age of concrete
Figure 2.1: Relative magnitudes of drying and autogenous shrinkage.

2.2.1. Plastic shrinkage


When the concrete is first cast, particles settle and excess water rises to the top forming a thin layer in a process known as bleeding. If the rate at which bleed water rises is less than the rate at which it evaporates the water level will eventually drop below that of the surface particles forming a meniscus as shown in figure 2.2. 3

Figure 2.2: Meniscus that forms as a result of evaporation of bleed water

Figure 2.3: Forces acting on the meniscus.

The surface tension in the meniscus of the water acts upwards as shown in figure 2.3. To achieve equilibrium the water pressure must decrease to balance the external air pressure. Because the concrete is wet and the particles are mobile, this pressure differential induces shrinkage (Slowik, Schmidt & Fritzsch 2008). The mechanism is known as the capillary effect. As evaporation continues, capillaries become smaller and the meniscus radii sharper, inducing a greater pressure difference. Eventually the forces required by the menisci are too big, and the pressure reaches what is known as the air entry value (Slowik, Schmidt & Fritzsch 2008), at which point air breaks through the meniscus. This creates high localised stresses, with particles subject to relatively large tensile forces by menisci on one side and negligible forces on the side where air is entrained. These localised stresses can lead to what is known as plastic shrinkage cracking (Slowik, Schmidt & Fritzsch 2008). As this shrinkage occurs while the concrete is wet, bonds have not yet formed between concrete and reinforcing steel. The result is two-fold. On one hand cracking is not restrained by the steel, and cracks may carry across the entire section (Slowik, Schmidt & Fritzsch 2008). On the other hand, there are no internal restraints creating tension in the concrete. Images of plastic shrinkage crack formation are shown in figure 2.4.

Figure 2.4: Electron microscope images of the formation of plastic shrinkage cracking (Slowik, Schmidt & Fritzsch 2008).

Plastic shrinkage is determined by the rate at which evaporation and bleeding occurs. It is also highly dependent on the rigidity of the concrete mix (Neville 1995). Plastic shrinkage increases for increasing cement content and decreasing water content (Neville 1995). As the concrete starts to set and a solid skeleton forms, the forces exerted by the capillary pressures have less effect and the importance of capillary action reduces dramatically (Wittmann 1976).

2.2.2. Drying shrinkage


Drying shrinking occurs once the concrete skeleton is formed and is the result of a loss of moisture. There are four mechanisms which are suggested to cause drying shrinkage: capillary action, disjoining pressure, surface free energy and loss of interlayer water (RILEM 1988). Capillary action is discussed in section 2.2.1 and also applies to drying shrinkage but with reduced effect compared to plastic shrinkage, due to the restraint provided by the rigid skeleton. Disjoining pressure is the pressure that separates two parallel surfaces attracted to each other by the Gibbs energy of the two surfaces (International Union of Pure and Applied Chemistry 2012). In concrete these surfaces are the cement particles. Surrounding each particle is a film of adsorbed water which separates it from a layered neighbouring particle. The particles are attracted to each other, but repelled by the film of adsorbed water resulting in a disjoining pressure in the film of water. The water films, or small gaps between the cement particles, are known as gel pores (Neville 1995) or micropores (RILEM 1988). There is also a network of larger pores that are longer but not completely continuous through the concrete, called capillary pores. If the relative humidity of the environment is lower than that of the capillary pores, water is drawn from the micropores to maintain equilibrium. The movement of water from the micropores reduces the thickness of the water films separating the particles and results in shrinkage (RILEM 1988). This process is reversible so that concrete may expand if subject to environments with higher relative humidities. Surface free energy is related to the surface tension of solids. Atoms at the surface of a solid are in a higher state of energy than those inside. This is because there is an imbalance of forces with greater attractive forces between atoms of the solid, than those with the external environment. This creates a net hydrostatic compressive force on the solid. Water adsorption, however, reduces this imbalance, and decreases the surface free energy of the solid cement particle. As a result the hydrostatic compression is also reduced. Thus increased water in micropores leads to decreased surface energy and therefore expansion, while decreased water 5

leads to shrinkage (RILEM 1988). The water in micropores is governed by the relative humidity of the external environment and therefore so is shrinkage through changes in surface free energy. Loss of Interlayer water is governed by the loss of water between sheets of Calcium Silicate Hydrates (CSH). CSH are one of the products from the cement reaction, also known as hydration. (The other is tricalcium aluminate hydrate. Together these are (and have been) referred to as the cement particles.) There is not a clear distinction between the layers of water between CSH particles and the micropores referred to previously, however it is suggested that a small amount of water lost in these regions can lead to a large bulk shrinkage strains (RILEM 1988). The effect is greatest below 11% relative humidity and the shrinkage induced is found to be partially reversible. Little information is available in the literature on the extent to which each of these mechanisms affects drying shrinkage and in which conditions. A final consideration regarding drying shrinkage is the development of the drying front, or shrinkage profile within a cross section. Of the limited research that has been done in this area, none relates to the effect on time dependent behaviour.

2.2.3. Autogenous shrinkage


Hydration of cement requires water which is drawn from capillary pores. This process is known as self-desiccation. The volume of the hydrated cement (the product) is less than the reacting constituents. This volume change is known as chemical shrinkage. It is also known as the internal volume change. Autogenous shrinkage is caused by chemical shrinkage but is measured as the external volume change (Nehdi & Soliman 2011). This means during the plastic stage when the cement is still wet and the particles are mobile, chemical shrinkage is identical to autogenous shrinkage (these effects are minor however, and not usually considered in plastic shrinkage because hydration is minimal in the first two hours (Wittmann 1976)). As the concrete sets, the skeleton provides some resistance to the chemical shrinkage, and autogenous shrinkage drops below chemical shrinkage with no external supply of water (if external water is available concrete can expand from continued hydration and water absorption (Neville 1995)). As hardening progresses, autogenous shrinkage becomes increasingly restrained (Nehdi & Soliman 2011). Autogenous shrinkage is typically minor for normal water to cement ratios, but can be as large as drying shrinkage for very low water to cement ratios as in high performance concretes (Faria, Azenha & Figueiras 2006). It is less affected by size and shape of the member and relative humidity than drying shrinkage (Ranzi & Gilbert 2011).

2.2.4. Thermal shrinkage


Thermal shrinkage (or expansion) is governed by the coefficient of thermal expansion (CTE). However, the behaviour of concrete subject to a change in temperature depends on the temperature gradient within the section, which can create internal stresses. This is determined by thermal conductivity. The CTE for concrete is a result of the mixed coefficients for aggregate and hydrated cement paste.

2.3. Compressive creep


Creep can be broadly defined as deformation as a result of constant load (Neville 1995) (excluding shrinkage deformations). It can be divided into recoverable and non-recoverable parts. Recoverable creep, termed delayed elastic strain, occurs immediately after loading, and constitutes approximately 10 20% of total creep (40-50% of elastic strain) (Ranzi & Gilbert 2011). It is determined by subtracting the instantaneous elastic recovery from the total strain recovered when a load is removed (Neville 1995). Bazant, however, points out that the separation of elastic and creep recovery strains is often ambiguous (RILEM 1988). Generally, it is found that creep recovery is independent of the factors that govern the magnitude of irrecoverable creep (Neville 1970). Creep may also be defined according to whether the concrete is subject to drying or not. Consider two identical specimens: one where drying is prevented and which is subject to sustained compression, and one which is unsealed and subject to drying conditions, but not loaded. Intuition would suggest the combined response of these two specimens would be the same as a specimen subject to drying and loading. However, it is found that deformation exceeds the sum of the strains of the two independent tests. The difference is known as drying creep. Creep which occurs in the absence of shrinkage is known as basic creep. Figure 2.5 expresses this graphically.
Drying Basic creep Basic creep Strain Elastic Time Time Strain Shrinkage Elastic

Figure 2.5: Comparison of creep in a sealed specimen (left) and creep in a drying specimen (right)

2.3.1. Factors influencing creep


There are a number of factors that influence the rate of creep in general: It is heavily influenced by relative humidity, with lower relative humidities causing increased creep strains (Neville 1995). As the age at which loading occurs increases, creep decreases. It is found to be non-linearly related to the volumetric contents of cement paste and aggregate (Neville 1995). It is also influenced by the stiffness of the aggregate which provides restraint to the potential creep of the paste alone (Neville 1995). The porosity of the aggregate appears to influence creep rates (Neville 1995). Up to a stress to strength ratio of 0.5 fc, the relationship between stress and creep is linear, but becomes non-linear after this as micro cracking occurs (Neville 1995). It is found as temperature increases or decreases an increase the transient rate of creep occurs (Bazant, Cusatis & Cedolin 2004). There is a size effect where an increase in surface area to volume ratio leads to a decrease in creep.

2.3.2. Mechanisms behind creep


Many theories have been proposed to explain creep behaviour including the factors that influence it. The most relevant developments are outlined below: 1. Basic creep is explained by the cement paste being a visco-elastic material: part elastic and part viscous (Neville 1995). Under load, the viscous phase of the cement paste flows. The flow is a result of bonds breaking and reforming. Alone, however, this theory does not explain drying creep, ageing or any of the influencing factors outlined above. Bazant suggested an improvement, known as Solidification theory which still assumes a viscoelastic material, but explains the ageing of concrete by an increase in the volume fraction of load bearing concrete. This increase is brought about by continuing cement hydration over time. Thus, the load bearing volume fraction of the concrete increases, along with stiffness. This explains short term ageing of concrete and importantly from a modelling point of view also allows visco-elastic parameters to remain time independent. It was found however, long term ageing could not be explained by solidification because volume growth of hydrated cement is too short lived (Bazant et al. 1997).

2. A mechanism behind drying creep was proposed by Wittman, who suggested that tensile stresses induced by shrinkage, caused microcracks in unloaded specimens, reducing measured shrinkage. Axially loaded specimens however, are not subject to any cracking, and shrinkage deformations are therefore greater. However, experiments with symmetrical members under pure bending, which shrinkage does not affect, still display drying creep, showing microcracks do not explain all of drying creep (Bazant & Xi 1994). Bazant proposed that stress induced shrinkage may explain additional drying creep. It is based on the notion that micro-diffusion between the micro-pores (which occurs as a result of drying) increases the ability of bonds to break and reform and therefore increases creep (Bazant & Chern 1985). No physical explanation behind this behaviour could be found however. 3. Bazant solved the issues in 1 and 2, with the development of microprestress theory. Microprestress is proposed to develop in the micropores as a result of differences in the energy of the water vapour and adsorbed water. These energy differences can be brought about by volume or temperature changes in the micropores. Because microprestress is transmitted through the bonds that exist between the opposing walls of micropores, this increases the breakage of these bonds, and promotes shear slip (viscous flow). Bazant shows this not only resolves issues in 1 and 2, explaining ageing and drying creep, but also explains the temperature effects on creep. Together, Bazant suggests the micrprestress and solidification theories explain almost all creep behaviour and together form a grand unified theory.

2.4. Tensile creep


Research described for creep so far is based on compressive creep. No consensus was found in the literature as to the magnitude of tensile creep compared to compression creep with conflicting research suggesting it was bigger, the same, and smaller (Neville 1970). Bissonnette found that Tensile creep was subject to drying creep as in compressive creep (Bissonnette 2007). However, research done by Illston suggests drying has no influence on the magnitude of creep in tension, while studies done by Davis et al. on plain concrete beams showed that drying creep on the compression face was three times greater than drying creep on the tension side (Neville 1970). Microcracking appears to play a minor role in tensile creep according to Bissonnette, who explains that micro cracking reduces the modulus of elasticity and is found not to be significantly changed after loading. As for compression creep, tensile creep is found to be proportional to applied stress, up to a limit of 50 67% of short term ultimate tensile strength (Bissonnette 2007) (Neville 1970). 9

2.5. Tensile strength


Tensile strength in concrete is a function of the propensity for concrete to fracture. If the concrete is assumed to be homogenous and flawless, theoretical tensile strengths are calculated to be 2000 times actual. The discrepancy is explained by the presence of flaws, which attract high stress concentrations despite low average stresses in the medium. Bigger flaws attract bigger stress concentrations. This can lead to microscopic failures but not necessarily entire failure. The propensity of the entire medium failing depends on the behaviour and state of the material surrounding the local failure. The number and size of flaws is stochastic, and means that strength is governed by probability. Therefore larger specimens are more likely to have a great number of bigger flaws leading to reduced tensile strength. All of this may explain why tensile strengths based on flexural tests are greater than those based on uniaxial stress (such as the Brazilian test). There is a size effect (less material is subject to tensile stress in a flexure test) and also a difference in the state of the material surrounding a potential flaw. In flexure, stresses reduce as distance from the extreme tensile fibre increases reducing the likelihood of cracks propagating (Neville 1995).

2.6. Modelling time dependent behaviour


Modelling time dependent behaviour requires spatial and time discretisation. Spatial discretisation refers to breaking down a structure into constituent elements, elements into cross sections, and cross section into layers. For each time step, the discretised structure must be solved by iteration, where deformations are adjusted progressively until equilibrium is reached (Kawando & Warner 1996). Time dependent behaviour is the result of two effects those that are stress dependent such as creep, and those that are stress independent such as shrinkage. To separate these two effects, stress dependent behaviour is measured as the difference in deformation between a loaded specimen and an identically sized and aged specimen that has undergone the same environmental conditions but unloaded (Bazant 1975). Stress-dependent behaviour over time may be modelled in essentially two ways. Firstly by an integral-type model, and secondly by a rate-type creep model. The integral-type model is based on the assumption that the relationship between stress and strain is linear. This is roughly true under serviceability conditions, where stresses are less than 40% concrete strength (Neville 1995). As a result of this linearity the stress dependent deformation may be expressed by the compliance function , , which is defined as the strain

10

at time t, caused by a unit application of constant stress applied at time . It incorporates both creep and elastic strains. Strain is then calculated as the sum of the stress changes over time by their respective compliance functions. A shortfall of this approach is that it does not directly model some extrinsic state variables that affect the rate of creep. Extrinsic state variables are factors that can change creep after casting, and are properties within the material. They include things such as temperature, degree of hydration and pore humidity (Bazant 1988). To account for this, behaviours such as drying creep, are often incorporated into the compliance function (as is done by AS3600-2009), rather than being modelled directly. Another disadvantage of the integral-type model is that stress increments for each time period for each discretised element or layer must be stored, decreasing computational efficiency and increasing memory requirements (Kawano & Warner 1996). It has been found the integral type formulation does not model creep recovery accurately, and should not be used in scenarios where unloading occurs (stress reduction as a result of redistribution is not problematic) (Bazant 1988). Under the rate-type method, concrete is modelled as a viscoelastic material. That is, it undergoes a time dependent shearing strain under shearing stress as would a liquid (albeit highly viscous), and also undergoes a non-time dependent elastic strain as a result of an applied stress. It is essentially represented by dampers (dashpots) and springs combined in series and parallel as required to produce the appropriate response. This is the same method used to model polymers, however unlike polymers, concrete is also subject to ageing. This means time dependent behaviour in concrete is not only a function of time lag, but of time lag and the time of loading. A result of this is that solutions must be solved numerically, not analytically (Bazant 1975). Bazant maintains the rate-type approach is most realistic, as it is based on the physical processes behind the solidification-microprestress theory (see section 2.3.2) and can incorporate the effects of ageing, varying pore humidity and temperature (Bazant 1997). The rate-type method is particularly suited to finite element applications because creep calculations are not dependent on stress histories and therefore do not need to be stored improving computationally efficiency. Warner, however, shows that this method can be unstable if time discretisation is not fine enough. Warner shows that the two methods, integral-type and rate-type, produce similarly accurate results for given stress histories, as long as the integral-type method is not used for unloading scenarios or where stresses in the concrete reach more than 0.4fc (Kawando & Warner 1996). For ease of application to experimental data, and because computational efficiency is not critical, the integral-type approach is used in this thesis. 11

Chapter 3 Time dependent behaviour in concrete


3.1. Time dependent properties
Deformations in concrete can be classified as either instantaneous or time dependent. When subject to load, concrete will effectively deform instantly. The extent of this deformation will depend on the stiffness of the concrete at the time, and the magnitude of the load. After loading and instantaneous deformation, the concrete will continue to deform over time. This is a result of three phenomena: shrinkage, creep and ageing.

3.1.1. Shrinkage
After concrete is poured and begins to set, it will shrink as water is lost and chemical reactions take place. This process occurs gradually, with shrinkage approaching an asymptotic upper limit as shown in figure 3.1.

Shrinkage

sh

Time
Figure 3.1: Development of shrinkage with time

As shrinkage depends on a range of factors as outlined in chapter 2, such as aggregate type, mix, and drying conditions, shrinkage strains can vary, but are typically in the range of
200 10 to 1100 10 (Wight & Macgregor 2012). The majority of this strain is reached

within 100 days and can be attributed to drying. The exception to this is high performance concretes with very low water to cement ratios which undergo significant autogenous shrinkage, making up as much as 50% of total shrinkage strain (Yang, Sato & Kawai 2005). In AS3600-2009 shrinkage is given as the sum of drying and endogenous shrinkage (autogenous not made between shrinkage types, and shrinkage is given simply as . and thermal shrinkage), so that = + . For the purposes of this thesis, distinction is

12

There are some important considerations regarding the effects of restraint and shrinkage on behaviour worth elaborating at this point. Consider a beam subject to shrinkage as shown in figure 3.2. Without restraint, the concrete will deform without stress. If restraint is applied, the concrete will want to shrink, but because it is restrained from doing so, will be drawn in tension.

Free shrinkage no induced stresses Restraint pulls the concrete specimen into tension from its free shrinkage state.
Figure 3.2: Shrinkage without restraint

Restraint can be in the form of end restraints as shown in figure 3.2 or as internal restraint in the form of reinforcing. Shrinkage profiles need not be, and in most cases are not, linear across a section. Shrinkage occurs more quickly the closer regions are to drying surfaces, and slower the further away they are. Consider a plain concrete slab drying from top and bottom only as shown in figure 3.3a. The outer surface will shrink more than the core, as shown by the free shrinkage strains in figure 3.3b . This induces stresses that produce strains acting in the opposite direction resulting in a uniform strain profile as shown in figure 3.3c. For design purposes it is common to assume a uniform shrinkage profile as these effects are not usually considered to affect calculated deflections significantly.

(a) Slab subject to free shrinkage

(b) Shrinkage strain profile

(c) Resulting strain = elastic strains + shrinkage strains

Figure 3.3: Free shrinkage strains

13

3.1.2. Creep
Creep describes the deformation of concrete under load over time. It is mostly irrecoverable deformation, so that once the load is removed the concrete does not go back to its original shape, but remains deformed. A small portion is recoverable, however the distinction between this and instantaneous elastic deformation is not easily made. Consider a specimen under constant load, disregarding shrinkage for the moment, as shown in figure 3.4. Load is applied over a certain time period, and then removed. The strains that occur as a result are shown in the strain vs time diagram, and shown schematically in the specimens above the graph.

Load

Load

Load Load

Elastic recovery

Creep recovery

Load removed Load applied Strain

Elastic Creep strain Elastic or instantaneous strain 0 t0 t Time Permanent deformation T recovery Creep recovery

Figure 3.4: Concrete under creep with no shrinkage

The magnitude of creep depends on the strength of the concrete, the age of the concrete when loaded, the composition of the concrete, dimensions of the specimen and humidity (Wight & Macgregor 2012). If the specimen is unsealed and allowed to dry, creep will increase, through a process termed drying creep, discussed in chapter 2. Typical values for creep are of the order of 2.5 times instantaneous deformation. 14

From a material modelling perspective, creep strain can be expressed as a proportion of initial elastic strain;
, = ,

(3.1)

Equation 3.1 may be also expressed as a function of stress:


, = ,

(3.2)

deformation at time , and , is known as the creep coefficient. The creep coefficient can be measured or calculated. AS 3600-2009 Concrete Structures provides a method to calculate the creep coefficient based on empirical studies, allowing for concrete strength, humidity, exposed concrete and concrete maturity. Accuracy of the resulting coefficient is in the order of 30% (Standards Australia 2009). A typical curve showing the creep coefficient versus time is shown in figure 3.5.

In equations 3.1 and 3.2, , is the creep strain at some time t past the initial elastic

( , )

(, )

(, )

Time

Figure 3.5: Creep coefficient vs time (Ranzi & Gilbert 2011)

3.1.3. Ageing
Over time, concrete strength (compressive and tensile) and stiffness gradually increase due to the continued hydration of the cement paste and other reasons outlined in chapter 2. Creep deformations also depend on age. For a given load, creep strains are smaller the later the load is applied. Collectively, these effects are termed ageing. Each of these will be discussed briefly. Compressive strength in concrete is usually specified as the lower characteristic cylinder strength at 28 days, denoted . Standards dictate 95% of cylinder tests of the same concrete must exceed this strength. Though this thesis is not concerned with ultimate strength, the

15

development of compressive strength with time, shown in figure 3.6, is important as it reveals an aging process also associated with tensile strength and stiffness.
1.4

Ratio fc(T)/fc(28)

1.0

0.6

0.2 1 3 7 28 90 365

Time (days) Figure 3.6: Generalised concrete compressive strength development over time for normal strength concrete relative to fc(28) (Wight & Macgregor

Compared to compressive strength, tensile strength develops over at a slower rate. As a result the relationship between the two is not linear. AS3600-2009 gives this relationship as
. = 0.6 for tensile strength in flexure, and = 0.36 for uniaxial tensile strength.

The elastic modulus is measured as the slope of the secant for the linear portion of the stress strain curve as shown in figure 3.7. It is a measure of material stiffness, and is a result of the combined stiffness of the cement and aggregate.

fc
Strength (MPa)

80

60

Stress

40

Secant modulus (elastic modulus) Strain


Figure 3.7: Stress strain curve for concrete

20

10

20 Strain x10

30
-6

40

Figure 3.8: Stress strain curves for varying strengths of concrete (Neville 1995)

16

From figure 3.8 it can be seen that the greater the concrete strength, the greater the elastic modulus. It follows from this the elastic modulus must increase with time if compressive strength does. This development of elastic modulus with time is reflected in various codes including AS3600 and Eurocode.

3.2. Time dependent modelling step by step method


To model time dependency, creep and shrinkage components must be included in the expression for strain in the concrete, thus:
= + + ()

Where ( )is the elastic strain, ( )creep strain, and ( )shrinkage strain. step by step method (SSM) is the most accurate and general. For brevity, explanation and comparison of the other methods is relegated to appendix A.

(3.3)

There are many methods that can be used as a basis for calculating creep strain of which the

The SSM is based on a stepwise approach, where gradual changes in stress are broken down into discreet intervals as shown in figure 3.9.

( )

()

Stress

Time

Figure 3.9: Discreet stress intervals used in the SSM

For any given stress change in the concrete there will be both an elastic strain and creep strain component, which using equation 3.2, can be given by:
+ = + ( , )

This can be expressed more compactly as:

17

+ = , ( )

(3.4)

Where , is known as the creep or compliance function. It represents the combined elastic and is given by:
, =

and creep strain for the time period ( ) resulting from the application of one unit of stress,
1 + ,

(3.5)

Total elastic and creep strain in the concrete can then be calculated by summing the elastic and creep strains for each of the changes in stress. From equation 3.3
= + + ()

= , + , + ()

(3.6)

Equation 3.6 can be approximated by:


= , + , + ()

(3.7)

This can be re-written in short hand as follows:


, , = , , + , ,

(3.8)

where j represents the current time step t = tj. Equation 3.8 can be re-arranged as follows:
, , = , , + , (, , ) + , ( , , ) , , = , , + , , , , + , , , , , , = , , + , , , , + , , , ,

, , = , , + , , , , + , , + , , , , , , , , = , , , + , , + , , , + (, , ) , , , = , , + (, , ),

(3.9)

Equation 3.9 can now be solved for the stress in the concrete at time tj

18

, , , , = + , , ,

, = , (, , ) + , , ,

(3.10)

where , , =

, , ,

(3.10a)

And from equation 3.5, , = , since , = 0 (there is no creep because no time has
,

elapsed). Stress in the steel is given as


, = ,

(3.11)

To maintain compatibility, the strain in the concrete must match the strain in the steel at a given position in the cross section. Thus;
, = , =

(3.12)

represents the strain at any point in the cross section as shown by figure 3.10 and is given by:
= +
d_ref

(3.13)

Cross section A-A

Strain

Figure 3.10: Strain in a beam under axial and bending loads where plane sections remain plane

In non-time dependent analyses, the x-axis is normally set to the position of the neutral axis of the cross section, where the first moment of area about the x-axis is zero. However, in an analysis involving cracking over time, the position of the neutral axis changes, making it more practical to refer to the x-axis by an arbitrary reference distance, d_ref, as shown in figure 3.10.

19

3.3. SSM assumptions


There are a number of assumptions behind the SSM. Firstly it assumes a linear relationship between stress and strain. This holds true for stresses up to 0.4 fc, and under service loading this is a valid assumption (Bazant 1988). Secondly, the SSM assumes the principal of superposition for creep. Creep strain (at a given point in time) is calculated as the sum of creep strains from loads, regardless of when the loads were placed. This has been found to agree with experimental observations when the stress history is increasing. However, when the stress history is decreasing, creep strains are found to be overestimated by super position (Kawando & Warner 1996). Finally, all methods, including the SSM, are limited by the accuracy of available inputs. Tensile creep, as mentioned in chapter 2, is not well researched compared to compressive creep. That which has been done shows conflicting results. For lack of a better alternative, this model will assume tensile creep is the same as compressive creep.

20

Chapter 4 Cross sectional analysis

4.1. Background
A cross sectional analysis may be used to calculate strain and curvature at a cross section based on the moment and axial force at that point. Using the SSM as a basis, formulations are developed for cracked and cracked sections and examples given.

4.2. Uncracked formulation


At a cross section, at any time, the internal forces will be equal to the external applied forces. The external forces and moments at a cross section refer to the external moments and axial forces that would need to be applied to maintain the position of the beam if the beam was cut at that point. Thus:
= =

(4.1) (4.2)

Ne and Me are the externally applied axial loads and moments, and Ni and Mi are the equal and opposite internal resisting forces, a portion of which comes from the steel, and a portion of which comes from the concrete. Thus;
= + = +

(4.3) (4.4)

The forces and moments in the concrete are given by equations 4.5 and 4.6 respectively.
= ,

(4.5) (4.6)

= ,

Combining equations 3.10, 3.13 and 4.5 and 4.6 yields:


= , (, + , ) + ,, ,

(4.7)

= , (, + , ) + ,, ,

(4.8)

21

After some manipulation (full derivation can be found in Appendix B), the resulting equilibrium equation is found to be:
, = + , ,

(4.9)

Where
, , , = , = , , , , ,, , , = , , = ,, , , = , , ,, ,

(4.10a-e)

And ,, = , + , ,, = , + and ,, = , + follows:


= , , + ,

Where subscript c denotes concrete, and s steel. Equation 4.9 is then solved for strain as

(4.11)

step will then be passed into equations 4.7 and 4.8. Values calculated for N0 and M0 will then

The first time step will have no creep history so that , = . The solution to the first time

time steps as necessary.

be used in the calculation of , for the next time step. The process is repeated for subsequent

4.3. Cracked formulation


The previous section described the solution for an uncracked section, where geometrical properties do not change with time. When cracked sections are taken into account, the concrete cross section must be analysed in layers. The greater the number of layers in the cross section the greater the accuracy of the solution. The appropriate number of layers can be determined by ensuring the second moment of area, as calculated by equation 4.12c, is within 1% of the analytical value . To begin the analysis, it is assumed the layers are uncracked, and the cross section properties Ac, Bc and Ic are calculated assuming each layer is a rectangle with width calculated at the centre of the layer, and height as shown in figure 4.1. For clarity, the number of layers in the diagram has been limited. Typical calculations could involve 500 layers.

22

d_ref

yl
h w
Layer l

Figure 4.1: Cross section divided into layers

For the layer shown in figure 4.1, = , = , = .

Ac, Bc and Ic are then given as:


, ,

(4.12a-c)

Where m is the number of layers in the cross section. Equation 4.11 is then called, and strains calculated. These strains are used to calculate stress in each layer with equation 3.10. Any layer with a stress greater than the tensile strength of the concrete are ignored for the recalculation of Ac, Bc and Ic using equations 4.12a-c. Equation 4.11 is again called, and strains, and concrete stresses recalculated and compared with tensile strength. New concrete geometric properties are calculated. This process is repeated until the value for strain converges to an acceptable limit. Once this limit is reached, and strains for the first (instantaneous) time period have been determined, it is necessary to include the values for Nc,0 and Mc,0 in the next time step. For this purpose, the integrals in equations 4.5 and 4.6 are approximated by the stresses calculated in the previous time step, so that:
,,

(4.13)

,,

(4.14)

Where , , is the stress in concrete layer h, at time tj.

23

The converging process outlined in the first time period is called again to solve for strains in the next time period. The process continues for as many time periods as required. It may be noted from this formulation that stresses in each layer must be stored for each time period so that solutions for subsequent time periods can be found. There are additional rules regarding stresses and stress histories which should be mentioned. Firstly, the stress history in a given layer is completely removed when that layer is cracked (when stress is greater than tensile strength). Secondly, once a layer is cracked it may only take compressive stresses from that point in time onwards. If it does take subsequent compressive stresses, these should become part of the layers new stored stress history. The limiting tensile strength may also be set to 0, rather than the tensile strength, if a conservative answer is required. It may also be set so that the concrete cannot crack, so the resulting solution will closely match that of the uncracked solution, enabling a check of the layered procedure.

4.4. Uncracked example layered approach


Strains for the cross section in figure 4.2 are to be calculated with the parameters in tables 4.1 and 4.2, assuming a constant load and moment of -30 kN and 50 kNm respectively, dividing the section into 500 layers.
Table 4.1: Loading, elastic modulus, shrinkage parameters and times steps for example j (days) 28 100 30,000 Ec,j (MPa) 25,000 28,000 30,000 0 -300E-06 -600E-06 sh(t) Table 4.2: Creep coefficients for example

(j,i) 28 100 30,000

28 0.0 1.5 2.5

100

30,000

0.0 2.0

0.0

b = 300 dst(1) = 50 Ast(1) = 620 dref = 200 dst(2) = 550 ys(1) = -150 ys(2) = +350 Ast(2) = 1800 Figure 4.2: Cross Section for example (all units in mm)

D = 600

24

4.4.1. Analysis at = days


The concrete section is divided into 500 layers. Each layer is of width 300mm, and height 600/500 = 1.2mm.
, , , and , are calculated with E= 25,000 MPa, and assuming no cracking. , = 4.50 10 + 484.0 10 = 4.98 10 , = 450.0 10 + 107.4 10 = 557.4 10

, = 180.0 10 + 46.9 10 = 226.9 10

Shrinkage is zero during the first time period, as is creep, so equation 4.11 reduces to:
= ,

, = 4.98 10 557.4 10

557.4 10 226.9 10

10 30 10 = 42 50 10 324 10

4.4.2. Analysis at = days


, , , and , are calculated in a similar fashion to the previous time step, however = 28,000 MPa is now used, giving the following values: , = 5.524 10 , , = 611.4 10 , , = 248.5 10

In order to calculate , Nc,0 and Mc,0 are calculated for each layer using the strains calculated
= 44.4 10 1.8 = 79.9 10 39.3 10 70.8 10

and must now be calculated in order to determine strains.

from = 28 days by equations 4.13 and 4.14. The results are summed giving:

In order to calculate , , ,and ,, are calculated using = 28,000 giving:


= 5.04 10 300 10 = 1.51 10 504.0 10 151.2 10 , = 4.98 10 557.4 10

Thus strains may be calculated using equation 4.11 as:


557.4 10 226.9 10
79.9 10 + 1.51 10 30 10 50 10 70.8 10 151.2 10

, 10 = 385 825 10

25

4.4.3. Analysis at = , days


, = 5.88 10 , , = 647.4 10 , , = 262.9 10
, = 3.24 10 = 106.1 10 155.4 10 324 10

, 10 = 669 1.20 10

4.5. Cracked example layered approach


The same cross section (and parameters) is analysed assuming the concrete can take no tensile stress, so that fct.f = 0 MPa.

4.5.1. Analysis at = days


The first calculation for the instantaneous analysis is the same as for the uncracked example with resulting strains:
, 10 = 42 324 10

These strains are used to calculate stresses in each of the layers, those with tensile stresses are excluded and , , , and , are recalculated. Strains are then calculated again, and this iterative procedure is continued until there is negligible difference between strains of successive iterations. Final properties in the section are found to be:
, = 4.98 10 , , = 557 10 , , = 227 10

With strains:
, 10 = 42 324 10

4.5.2. Analysis at = days


, = 5.52 10 , , = 611 10 , , = 248 10

Nc,0 and Mc,0 for , are calculated using only the active layers excluding those layers that are subject to tensile stresses. and are found to be:
80 10 , = 1.51 10 71 10 151 10

, 10 = 385 825 10

26

4.5.3. Analysis at = , days


, = 1.97 10 , , = 67 10 , , = 70.8 10 = , 10 = 526 2.16 10 260 10 , = 894 10 33.5 10 105 10

4.6. Comparison of results


Results for the example given in section 4.5 are compared with results from the analytical (uncracked) method and shown in table 4.3. Results from the analytical method are sourced from Ranzi and Gilbert (Ranzi & Gilbert 2011).
Table 4.3: Comparison of strains for cracked and uncracked methods

Analytical uncracked

Layered uncracked

Layered cracked

r,j

42.7 10 386 10 670 10 331 10 841 10 1,220 10

42.3 10 385 10 669 10 324 10 825 10 1,197 10

42.2 10 385 10 526 10 324 10 825 10 2,160 10

Differences in strains between the analytical method, and the layered uncracked method arise because the layered uncracked method calculates concrete properties based on the gross area of concrete, over stating the stiffness of the beam slightly. The analytical uncracked method calculates properties based on concrete net area, not including concrete where the steel exists. Results from the cracked section indicate cracking only occurs in the third time step, as strain and curvature at this time step are greater than for the uncracked sections.

27

Chapter 5 Finite element method


5.1. Assumptions and comments
In this section, a finite element formulation is derived for a beam governed by Euler-Bernoulli beam theory and extended to incorporate time-dependent effects using the SSM. Assumptions that follow are: Plane sections remain plane, and shear deformations are ignored. This widely used assumption has been shown to hold true for long slender beams. A perfect bond between concrete and steel. This assumption is not completely valid, however variability in bond slip experiments suggest inclusion of such behaviour cannot be relied on (Kotsovos & Pavlovic 1995). Service stresses in concrete remain in the elastic region. At service loads this assumption holds true (Bazant 1988) The cross section is symmetric about the y axis as shown in figure 5.1 so that no torsional or out of plane bending effects are considered. An axis system is used where the z axis is at an arbitrary height as shown in figure 5.1. z x x y y

Figure 5.1: Assumed axis system

For design, where creep and shrinkage parameters are estimated with great variability, this formulation would not be warranted and simpler methods preferred. However, this FEM model will be used in conjunction with experimental data that is far more accurate.

28

5.2. Formulation
Using the principal of virtual work, the weak formulation for the Euler-Bernoulli beam is given as:
= . .

where = , the internal axial force and moment

(5.1)

= , the reference axis strain and curvature = = , the axial and transverse distributed loads = , the virtual displacements =

a differential operator

and represent displacements in the longitudinal (z axis) and transverse (y axis) directions respectively. Equation 5.1 equates internal strain energy (LHS) with external work (RHS). A derivation for this can be found in Appendix C.

5.3. Degrees of freedom and consistency


The deformed shape of the beam is assumed to take the form of a polynomial. The order of the polynomial chosen determines the number of nodal points required in the element, so that the number of unknowns is equal to the number of equations. A 7 node (7 degrees of freedom) beam element as shown in figure 5.2 is chosen for this study, with a cubic describing deformation in the y direction along the beam, and a quadratic describing the deformation in the x-direction along the beam. These order of polynomial avoid a locking problem where elements become overly stiff when the centroid moves away from the reference z axis - an issue where cracking is involved. Locking can arise due to uneven contribution of strain in the x and y directions. Using cubic and quadratic polynomials ensures that strain becomes a linear function of position in both directions (Ranzi & Gilbert 2011).

Figure 5.2: 7 degree of freedom beam element

29

The polynomial used to describe the deformed shape may be expressed in terms of the coefficients of each order, or in terms of the displacements at particular points along the deformed shape. This latter form of the polynomial is known as the shape function, and is expressed as:

] and, Where = , = [ 1 =

(5.2)

0 1

0 +

5.4. Time dependency


To incorporate the time dependent effects of creep and shrinkage, equation 4.9 and 5.1 are combined, where ri = re, giving:
= . + .

(5.3)

Substituting equation 5.2 into 5.1 gives:


= . + .

(5.4)

which can be rearranged to form:


= + . .

(5.5)

expressed as:

using the matrix property . = . , equation 5.2 and equation C.8, equation 5.5 can be
( ) = ( + )

(5.6)

This is equivalent to = , where is the element loading vector given by


( ) . =

+ , and is the element stiffness given by =

This system of equations is then solved taking the inverse of ke and using matrix partitioning.

30

5.5. Transformation from local to global axes


The formulation given in 6.8 applies to a single beam element. A finite element model consists of multiple elements, and this system must be solved by creating a global stiffness matrix and global displacement and loading vectors. To facilitate this, local axes must be converted to a consistent global axis by the following transformation:
= and =

(5.7)

Where d and q are the local displacement and loading vectors respectively, and D and Q are the global displacement and loading vectors respectively. T is the transformation matrix and for a 7 degree of freedom beam element is given as:
0 = 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1

(5.8)

where is the angle between the local and global x axis, taken counter clockwise from the global x axis as shown in figure 5.3.

Figure 5.3: Relationship between local element axis and global axis

Equation 5.7 is combined with equation 5.6 to give:


=

So that the stiffness for the element in global terms becomes =

Global Y Global X (5.9)

5.6. Shrinkage
A shrinkage profile in a concrete member is often assumed to be uniform across its crosssection. This is not what is found according to relative humidity profiles (discussed in chapter 7). To account for other possible shrinkage profiles, the FEM model is adjusted so that the shrinkage profile is approximated by a polynomial. To achieve this, a shape function similar to that described by equation 5.2, is used. 31

The approximating polynomial is given as:


= + + +

(5.10)

Values for shrinkage down the cross section are then assumed to be known by experiment or otherwise. So that at = = , = = , = = , and at
= = . This can be expressed in matrix form as follows: 1 1 1 = 1

or =
=

(5.11)

Solving for a, gives: Where is the vector of known shrinkage values, at points , , . Shrinkage strain at
= 1

(5.12)

any point along the y axis can now be given by:

(5.13)

Equation 5.13 can be used to calculate shrinkage in each layer. These values are then multiplied by the appropriate geometric values for each layer as per equations 4.12 a) and b), and summed giving the shrinkage vector , = , . , ,
,,

5.7. Cracking
A layered model is used, similar to that outlined in section 4.3, to account for cracking. For the first iteration, it is assumed the section is uncracked and the calculated values for the matrix D are based on the summed values in each of the layers. Recall from equation 4.10b, = ,
, , ,

After strains are solved for, values for ,, , ,, and ,, are recalculated where layers with stress greater than the concrete tensile strength are not included. This is repeated until resulting strains converge. In the next time step, it is necessary to calculate the creep

, , component , where , = , , . , and , must be calculated using

stored stress values at a cross section in each of the layers, and integrated along the element.

32

5.8. Gaussian quadrature


Integration along an element where stresses and active layers change, mean that closed form solutions are problematic. To get around this, integration is done using Gaussian quadrature, specifically Gauss-Legendre quadrature. The basis for quadrature is that the definite integral of any polynomial of a particular order may be given exactly by the sum of weighted values at certain points along the curve. The weighting applied to each of these values, and the points at which these values are evaluated by the polynomial depend on the order of the polynomial. If the polynomial is evaluated at n points (known as sampling locations), the integral is exactly given for any polynomial up to order 2n-1. To generalise the rule, the domain is normalised to [-1,1]. Weightings and sampling locations for the first 3 gauss points are shown in table 6.1
Table 6.1: Gauss-Legendre sampling points and weights

Points 1 2 3

Sampling location (xi)


0 3/5, 0, 3/5 1/3, 1/3

Weights (Wi) 2
1,1 8/9, 5/9, 5/9

The sampling locations in table 6.1 are based on the roots for Legendre functions of order n, while the weightings are given by =

(Abramowitz & Stegun 1964).

As an example, a graphical representation for a cubic function is shown in figure 5.4. The definite integral is given by the weighted sum of the function evaluated at the sampling points

and

, giving 1.14+1.53 = 2.67 matching the analytical solution.


3.0 2.0 f(x) 1.14 1.0 0.0 -1.0 -0.5 0.0 0.5 1.0 x 1.53
= + +

Figure 5.4: Sampling points and weights for a cubic

To convert from the normalised domain to a domain of a-b, the following operation is used:
=

+ + 2 2 2

(5.14) 33

5.9. Programming
Matlab was chosen as the programming language to implement the FEM program. It was chosen because of its ability to work with matrices efficiently and easily, despite difficulties with the implementation of a user interface. The FEM program was written with the following calculation steps: For each time step Calculate shrinkage While the difference in displacement compared to the prior loop is greater than 1e-7 For each element For each gauss point If first loop, calculate geometric properties assuming layers are uncracked If time period is greater than 1, calculate creep vector based on geometric properties from previous loop, and stored stress history Calculate shrinkage vector Assemble element stiffness matrix using geometric properties and gauss weighting Loop Loop For each element, transform element stiffness matrix, and loading vector from local to global coordinates Assemble the global stiffness matrix Solve the system of equations by partitioning the matrices For each element For each gauss point Calculate global reactions, global displacements and local displacements Determine strains at the gauss sampling points, and identify uncracked layers Calculate and store stresses in each layer using equation 3.10 for current time period Calculate geometric properties based on uncracked layers Loop Loop Loop Remove stresses from previous time periods for layers that are now cracked Loop

34

5.10. Cracked example


The same cross section and loading analysed in section 4.4 is analysed using the FEM model, with two gauss points. The tensile strength of concrete is set to 3MPa, and the beam is subject to a constant bending moment over its length of 5m. Input into the Matlab model is via a graphical user interface shown in appendix D. Output for the first time step is shown in figure 5.5.

Rx = -30 kN Ry = 1.09e-14kN Dzz = 0.00811

Dx = -0.211 mm Ry = -1.09e-14kN Dzz = -0.00811

Figure 5.5: Results from FEM for first time step

Results for the three time periods are shown in table 5.2. These results can be compared to the cross sectional analysis because the loads and moments are constant across the beam.
Table 5.2: Results from FEM

Time (Days) N1x (mm) N1y (mm) N1zz (rad) N2x (mm) N3x (mm) N3y (mm) N3zz (rad)

28
0 0 811 10

100
0 0 2100 10

30000
0 0 5400 10 1300 10 2600 10 0

106 10 211 10 0 811 10

962 10 1900 10 0 2100 10

5400 10

To compare strains and curvature at a cross section, the displacement results at any node (in this case the right hand side has been selected) are converted to strain and curvature using equations 5.2 and C.8 so that
= =

(5.15)

This gives results shown in table 5.3, which are the same as for the cracked cross sectional analysis in section 4.5.

35

Table 5.3: Cross sectional results from FEM, matching those of a cross sectional analysis.

Time (days) e K (mm-1)

28
42.3 10 324 10

100
385 10 825 10

30000
526 10 2,160 10

5.11. Uncracked validation


Tests carried out by Al-Deen and Ranzi (Al-Deen, Ranzi & Uy 2012) described below, were used to validate the FEM model for uncracked sections. Two slabs (one way) with cross sections shown in figure 5.6, were subject to their self-weight with strains and deflections measured over a period of 119 days. Specimens were cast at the same time, and propped and moist cured for 15 days. After 15 days, props were removed leaving the beam simply supported as shown in figure 5.7. Mid-span deflections were measured using linear variable differential transformers (LVDTs). Mid span strains were measured with strain gauges on the top and bottom surfaces of the concrete and internally.

900

900

Slab A
162 180

Slab B
62 165 180

5 x N16 reinforcing bars

10 x N16 reinforcing bars

Figure 5.6 Cross sections for Slabs A and B used for long term deflection tests

180 Bearing Roller support 150 3,000 150

Figure 5.7: Support and loading conditions for long term deflection tests

Shrinkage strains were measured from two concrete cylinders allowed to deform freely. These cylinders were poured from the same batch as for the beams and subject to the same curing and drying conditions. Measured strains are shown in figure 5.8.

36

Shrinkage 400E-6

Shrinkage vs time
Creep coefficient (15,t)

1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0

Creep coefficient vs time

300E-6 200E-6 100E-6 0 0 50 100 Time (days) 150

50 Time (days)

100

150

Figure 5.8: Shrinkage results from test cylinders

Figure 5.9: Creep coefficients calculated from cylinder tests

A polynomial curve was fitted to the data (shown by the dotted line in figure 5.8), and used to determine shrinkage strains for the FEM model. This was done to ensure a smooth shrinkage profile, rather than using actual values which may include outliers. Creep strains were measured from the average response of three cylinders subject to a constant stress of 5.75 MPa. Creep coeffients were calculated based on equations 4.1 and 4.3, so that:
15, = =

(5.16)

where represents the total concrete strain. Creep coefficients are shown in figure 5.9. As for shrinkage, a curve was fitted to the creep data as shown by the dotted line in figure 5.9, to provide a more representative value of creep for modelling purposes. As the step by step method is used, a family of creep curves must be generated based on the creep curve in figure 5.9. This is required because of the ageing effect. The creep curve from figure 5.9 is used as a baseline, and the remaining creep curves adjusted for the time step using AS3600-2009. The creep coefficient, according to AS3600-2009, is given by a basic creep value multiplied by factors to account for humidity, slab thickness, time after loading and age of concrete at loading (Standards Australia 2009). All factors, except the age of concrete at loading, are accounted for in the experimental creep coefficient values in figure 5.9. This factor in AS3600, known as k3, adjusts for the age of concrete at loading and is given by the following equation (Ranzi & Gilbert 2011):
= 2.7/(1 + log () )

(5.17) 37

Thus for any other time periods where the age at loading is not 15 days, the creep coefficient is adjusted and is given by:
2.7 1 + log , = (15, ) 2.7 1 + log15

(5.18)

Where is the time lag ( ). The density of concrete is taken to be 2,400 kg.m-3, and steel 8,000 kg.m-3, giving distributed loads of 3.87 kN.m-1, and 3.92 kN.m-1 for slabs A and B respectively. Youngs modulus was found to be 19,000 MPa at 15 days. The increase in Youngs modulus over time was estimated using a modified relationship from Eurocode 90, with s = 0.38 for normal strength concrete (Comit Euro-International du Bton 1993):
=

(15)

(5.19)

This estimate produces a slightly higher value for E than would occur if equation 5.19 was based on (28) but the effect on the compliance function (, ) is small.

Calculated deflections mostly follow actual deflections closely and are shown in figure 5.10. Strains are compared in figure 5.11. Curvatures seem to deviate more than reference strain. Causes for this most likely relate to variability in the estimated parameters of (, ) and .

It is also possible material parameters in the slabs are not reflected in the cylinder tests but this is unlikely.
2.0 1.8 Mid span deflection (mm) 1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0 20 40 60 Time (days)
Figure 5.10: Comparison of mid-span deflection as measured by experiment and by FEM calculation

Calculated vs Actual Mid-Span Deflection

Slab A - actual Slab A - FEM calculated Slab B - actual Slab B - FEM calculated 80 100 120

38

Long term strain profiles for Slab A


Distance from bottom of cross section (mm) 119 days 60 days 45 days 30 days

Long term strain profiles for Slab B


Distance from bottom of cross section (mm) 119 days 60 days 45 days 30 days 200

200 180 160 140 120 100 80 60

180 160 140 120 100 80 Theoretical Actual 60 40 20 0 -600E-06 -400E-06 Strain -200E-06 0

Theoretical Actual

40 20 0 -200E-06 0

-600E-06

-400E-06

Strain

Figure 5.11: Comparison of mid-span strains as measured by experiment and by FEM calculation for Slabs A and B

5.12. Cracked validation


Tests by Gilbert and Nejadi were used to validate the FEM model for cracked sections (Gilbert & Nejadi 2004). Deflections for a one way slab and beam were measured over a period of 400 days under sustained load. The cross section and loading conditions are shown in figures 5.12 and 5.13.

2.9 kN.m-1 (plus dead load) 130 3 x N12 reinforcing 161 161 Bearing Roller 400 3,500

Figure 5.12: Slab cross section, and support and loading conditions for long term cracked tests (dimensions in mm)

39

18.6 kN

18.6 kN

300

348 Bearing Roller 3,500

348

2 x N16

Figure 5.13: Beam cross section and support and loading conditions for long term cracked tests (dimensions in mm)

Shrinkage, creep, flexural tensile strength and modulus of elasticity values were separately measured using cylinders and prisms according to AS 1012-2000 Methods of testing concrete, with results shown in tables 5.4 and 5.5.
Table 5.4: Creep coefficients and shrinkage strain values

Age (days) cc sh

14 0 0

16 0.14 -14

21 0.36 -109

27 0.48 -179

53 0.92 -403

96 1.15 -591

136 1.29 -731

200 1.4 -772

242 1.5 -784

332 1.64 -816

394 1.71 -825

Table 5.5: Tensile strength and modulus of elasticity values

Age (days) f'ct (MPa) E (MPa)

7 3.0 21,090

14 3.7 22,820

21 4.3 23,990

28 5.6 24,950

As for the uncracked validation, creep coefficients and shrinkage values are fitted to a curve for use in the model, and the family of creep coeffficients calibrated against AS3600-2009. The values for the modulus of elasticity beyond 28 days are calculated using equation 5.19. The tensile strength of concrete is critical to the calculation of deflection for cracked sections, because it determines the stress at which the concrete will crack in the model as described in chapter 4. AS3600-2009 provides a relation between concrete compressive strength and flexural tensile strength where
= 0.6

(5.20)

Equation 5.20 is not ideal because the relationship between the two variables is weak (Wight & Macgregor 2012). The flexural tensile strength shown in table 5.5 is the most useful measure of tensile strength for the model. It should be noted, it is still not ideal because in measuring the flexural tensile strength, it is difficult to remove the effects of shrinkage. That is, the measured 40

flexural tensile strength does not take into account the additional tensile stresses near the surface caused by non-uniform shrinkage. Thus, tensile stresses reported may be understated. In order to calculate tensile strength beyond 28 days for use in the model, increase in tensile strength over time is based on the increase in compressive strength. Since tensile strength is given at 28 days, strengths beyond this point in time are calibrated by using figure 3.6 and equation 5.20. For example, the tensile stress at 90 days is calculated as follows; From figure 3.6, fc at 90 days is given as: 90 = 1.2 (28) The ratio of flexural tensile strength at 90 days to that at 28 days is then given by:
90 0.61.2 = = 1.1 28 0.6

So that 90 = 1.1 28 = 1.1 5.6 = 6.1 MPa.

Deflections are calculated using both uniform shrinkage and an assumed shrinkage profile as shown in figure 5.14. The assumed shrinkage profile is based on a measured relative humidity profile from the literature (Cement Concrete and Aggregates Australia 2007), and for any given time period is shaped such that the average of the shrinkage strains is equal to the uniform shrinkage value for the same time period. To validate the shrinkage profiles, the resulting free shrinkage stresses generated at 14 days in the concrete without reinforcing steel are compared with those measured by Grasley and Lange on a 76 mm thick mortar prism at five days allowed to dry on two opposing sides only, and shown in figure 5.15. The maximum stresses in the samples from Grasley are limited by the tensile strength of the concrete at approximately 2 MPa, while the limit for the FEM calculation is 3.7MPa. This aside, the similarity between the two graphs confirms the shrinkage profiles shown in figure 5.14 are reasonable estimates. The average of each of the non-uniform profiles in figure 5.14 is the same as the uniform shrinkage values.

41

Assumed shrinkage profile over time


400,332, 242,200 days

136 days

96 days

53 days

28 days Cross section height (mm) 160 140 120 100 80 60 40 20 0 0

-900E-6 -800E-6 -700E-6 -600E-6 -500E-6 -400E-6 -300E-6 -200E-6 -100E-6 Free shrinkage strain

Figure 5.14: Progression shrinkage strain profiles over time assumed for the validation test.

Free shrinkage stresses - measured

Free shrinkage stresses - calculated

80 70 60 50 40 30 20 10 0 -1 0 1 2 3
Stress (MPa)

160 Height in cross section (mm) 140 120 100 80 60 40 20 0 -1 0 1 2 3 Stress (MPa)

Figure 5.15: Comparison of free shrinkage stresses as a result of non-uniform shrinkage strains. The left hand side shows those measured by experiment and the right hand side, those calculated by the FEM program based on an assumed shrinkage profiles.

The development of the assumed shrinkage profiles over time in figure 5.14 is based on a similar progression found experimentally over the first 7 days of curing and shown in figure 5.16.

Height in cross section (mm)

42

Progression of relative humidity profile 80 Height in cross section (mm) 70 60 50 40 30 20 10 0 0 10 20 30 40 50


% change in relative humidity since pouring Figure 5.16: Progression of the relative humidity profile over the first 7 days of curing for a concrete prism (Grasley & Lange 2004)

3 days 5 days 7 days

Calculated deflections for the slab and beam are shown in figures 5.16 and 5.18 respectively. These include deflections calculated using uniform shrinkage at each time period, and those calculated assuming the shrinkage profiles in figure 5.14.

14.0 Mid span deflection (mm) 12.0 10.0 8.0 6.0 4.0 2.0 0.0 0

Calculated vs Actual Mid-Span Deflection

Actual FEM model - non-uniform shrinkage FEM model - uniform shrinkage

100

200 Time (days)

300

400

Figure 5.17: Comparison of deflection s as calculated by the FEM model and as measured by experiment for the beam

From figure 5.17, FEM calculated deflections for the beam are similar to actual deflections regardless of shrinkage profile. In contrast to this, results for the slab in figure 5.18 show significant variance between deflections calculated with and without a uniform shrinkage profile. 43

30 Mid span deflection (mm) 25 20

Calculated vs Actual Mid-Span Deflection

Actual 15 10 5 0 0 50 100 150 200 Time (days) 250 300 350 400 FEM model - Non-uniform shrinkage FEM model - Uniform shrinkage

Figure 5.18: Comparison of deflections as calculated by the FEM model and as measured by experiment for the slab

Calculated deflections with uniform shrinkage for the slab are much lower than actual, because the slab does not crack under these conditions. This is confirmed by comparison of the cracking moment to the midspan moment. The cracking moment is calculated by transformed section and found to be 6.85 kN.m, while the mid span moment is given by /8, and found to be 6.79 kN.m. When non-uniform shrinkage strains are introduced the slab cracks due to additional shrinkage induced tensile stresses at and near the surfaces of the slab. The resulting cracked beam profile and stresses are compared to actual beam cracks at 400 days in figure 5.19, and show good agreement. Calculated cracking and stresses

Actual beam cracks


160 mm

3500 mm Figure 5.19: Comparison of cracking for the one way slab as calculated and as recorded by experiment at 400 days

44

It is noted, however, that autogenous shrinkage is not measured, nor included in the model, so that at = 14 days, = 0. This may introduce some error. However, in light of the variability of the other inputs such as tensile strength and elastic modulus, a more refined approach may not be any more meaningful. The results in this section show the model predicts the cracked behaviour of beams and one way slabs well. They also highlight the sensitive nature of deflection to cracking and the critical role shrinkage can play in the onset of cracking.

5.13. AS3600-2009 comparison


AS3600-2009 provides a simplified approach to calculating long-term deflection in section 8.5.3.2, by multiplying short term deflection by a multiplier kcs, given as follows:
= 2 1.2 / 0.8

Where is the area of steel in the compressive zone (if any) and is the area of steel in the tensile zone, both taken at midspan in a simply supported beam. For the uncracked beams analysed in section 5.11, short term (immediate) deflection is given by:
= 384

(5.21)

(5.22)

For these beams, the transformed section properties (all concrete) are; = 484.4 10 mm4 and for both. This closely matches the short term values calculated by the FEM of 0.43mm. The
= 498.0 10 mm4 for slabs A and B respectively, giving deflections of 0.44 mm at 28 days factor is calculated to be 2 for slab A and 0.8 for slab B. For slab A, this gives a long term

deflection of 0.89 mm compared to actual long term deflection of 1.97 mm, and for slab B, 0.35

mm compared to an actual of 1.34 mm. FEM on the other hand provided closer results of 1.74 mm (vs 1.97 mm actual) and 1.21 mm (vs 1.34 mm actual). For the cracked beam analysed in section 5.12, deflection can be calculated using equation 5.22 however I is now calculated according to clause 8.5.3.1 to incorporate cracking:
= + + 0.6

(5.23)

where:

is the cracked transformed section found to be 28.7 10 is the cracking moment given by:

is the uncracked transformed section found to be 144.3 10

is the service design moment found to be 6.79 kNm based on a UDL of 4.43 kN.m-1.

45

= . +

where:

+ 0

(5.24)

. is calculated to be 3 MPa using equation 7.XXX based on = 25 MPa. is the maximum shrinkage-induced tensile stress given by: = 2.5 0.8 1 + 50

P is a prestressing/post-tensioning force not applicable to this example.

(5.25)

where:

is the ratio of tensile web reinforcement and found to be 0.0053 is the ratio of compressive web reinforcement which is 0.
is the final design shrinkage strain of concrete. From table 3.1.7.2 in AS3600-2009, for

25 MPa concrete, and interior environments this is found to be 789 10 (interpolated value).

This gives calculated values of = 1.65 , = 2.48 and = 34.3 10 . Short term deflection is calculated to be 11.3 mm, and long term deflection 22.6mm. This compares to an actual value of 25.2mm, and 24.1 mm as calculated by FEM. These results show that in design, the increased accuracy that comes about from a refined approach comes at the cost of time and increased complexity, and this trade off must be evaluated in light of the design problem. In addition, consideration must be given to the variability of the properties that must be estimated for the calculation of time dependent behaviour, and whether a refined and accurate model would improve results significantly.

46

Chapter 6 Measurement of shrinkage profiles

6.1. Previous techniques


As previously outlined, most shrinkage can be attributed to drying shrinkage. Drying shrinkage, as discussed in chapter 2, occurs as a result of a loss of moisture by evaporation at the surface. Thus the shrinkage profile will closely follow that of the moisture profile. Previous experimental methods used to measure moisture content include overall weight change, destructive sectioning and internal humidity/water content measurements (Weiss 1999). Weight change and destructive sectioning methods either do not provide adequate resolution or are not practical. Humidity measurements have previously been taken by measuring electrical conductivity of a porous medium, such as porous siltstone, inside the concrete (Rajabipour, Sant & Weiss 2007). An example of this device is shown in figure 6.1.

Figure 6.1: Measuring RH by measuring electrical conductivity (Rajabipour, Sant & Weiss 2007) Figure 6.2: Measuring RH in sealed cavities by direct measurement (Grasley, Lange & D'Ambrosia 2006)

One disadvantage is that the porous medium itself absorbs some of the water thereby reducing the relative humidity and electrical conductivity. This is known as hysteresis. Hysteresis can be accounted for and the effect removed, but this requires calibration measurements. Accuracy of this method is not confirmed. Another method used to measure relative humidity is by measuring relative humidity directly in sealed cavities in the concrete as shown in figure 6.2.

47

The drawback of this procedure is that the cavities run the depth of the sensor, and the air in this large cavity must come into equilibrium with the concrete which may affect readings.

6.2. Development of new sensors


The approach taken in this thesis is to measure relative humidity by burying small humidity sensors in the concrete, and sealing them with a breathable and waterproof cap. It is expected impact on concrete performance would be minimised. Their small size would also suit real time measurement if an application arises for measuring humidity in concrete in real buildings. Sensors used are Sensirion SHT25s, with an accuracy of 1.8% RH, an operating range of 0-100% RH, and size 3 x 3 x 1.1 mm. An image of the sensor is shown in figure 6.3.

3 mm

3 mm
Figure 6.3: Measuring RH in sealed cavities by direct measurement Invalid source specified.

These sensors must be integrated into an electronic system. They cannot be used off the shelf, and a circuit board is required to measured data. Design of a printed circuit board was sourced from a local design company in Sydney. Size was minimised by printing on both sides of the board. Final designs are shown in figure 6.4.

(a)

(b)

Figure 6.4: Circuit board design for the sensor

48

The white headerboard (4 pin connector) shown in figure 6.4a was not used, and wires directly soldered to the PCB to reduce size. The grey cap shown in figure 6.4b is a PTFE membrane, similar to Gore-tex commonly used in rain jackets, that is waterproof but also breathable. The remainder of the PCB was sealed in an epoxy resin to ensure moisture would not interfere with components. A 3m flat RJ11 (telephone) cable was used to connect to the PCB (by soldering). The other end was connected to a male RJ45 connector, to connect to a data logger, as shown in figure 6.5.

Figure 6.5: Data logger used to connect to humidity sensors Invalid source specified.

Manufacture and assembly of the circuit boards and sensors was completed by a company in South Australia. The finished product is shown in figure 6.6 and has been tested successfully in wet environments.

Figure 6.6: Finished humidity sensor

49

Chapter 7 Conclusion

The objective of this thesis was to predict the behaviour of reinforced concrete members over time and under service conditions. In this vein, a cross sectional method of time dependent analysis was formulated for cracked sections using a layered approach, and based on the step by step method. This was validated against examples in the literature. A finite element program was developed in Matlab, extending application of the step by step method to frames and non-uniform loading. A comparison of experimental results from the literature with the program showed agreement in both cracked and uncracked beams. In the analysis of cracked beams, the incorporation of a curved shrinkage profile in the finite element model was found to closely predict measured deflections, while a uniform shrinkage profile predicted only half those measured. This highlights a critical role the shrinkage profile can play in the calculation of cracking and deflection. A comparison between the simplified method for calculating deflection provided by AS3600-2009, and a more refined approach such as the finite element program, showed improved accuracy for the refined model. It also highlighted the need to assess the trade-off between accuracy and complexity, and whether a refined model would significantly improve predictions in light of the variability in estimating time dependent properties.

50

Appendix A Comparison of cross sectional methods to analyse time dependent behaviour

Four methods to model time dependent behaviour are compared.: The effective modulus method (EMM), the age-adjusted effective modulus method (AEMM), the rate of creep method (RCM) and the step by step method (SSM).

A.1. Constant Deformation


To compare methods used to analyse time dependent behaviour, consider a symmetrically reinforced concrete beam unrestrained and subject to shrinkage of -600 that is constant over time (while not realistic, this will facilitate explanation of the methods), as shown in figure A.1.
100mm

Where; fc = 40 MPa As = 200 mm2

Figure A.1: Concrete beam subject to shrinkage and creep

Values for the creep coefficient and the elastic modulus are given in table A.1 (calculated using AS 3600-2009 Concrete Structures). Shrinkage values are not realistic.
Table A.1: Creep and shrinkage values calculated as per AS 3600-2009 Time (days) sh(t) E()

100mm

14 0.00 -600E-6 26.8E+3

50 2.22 -600E-6 29.3E+3

100 2.62 -600E-6 30.1E+3

200 2.85 -600E-6 30.8E+3

400 2.98 -600E-6 31.2E+3

1000 3.08 -600E-6 31.6E+3

2000 3.12 -600E-6 31.8E+3

5000 3.15 -600E-6 32.0E+3

10000 3.16 -600E-6 32.1E+3

From an intuitive perspective, one may reasonably expect the unrestrained specimen to contract immediately as a result of the shrinkage, with compressive stresses in the restraining steel and tensile stresses in the concrete. Creep in concrete will reduce these stresses over time (effectively allowing it to stretch), as shown in figure A.2.

51

Intial tensile stress as a result of shrinkage

Stress

stepwise reductions in tensile stress as a result of creep from the initial tensile stress

Time
Figure A.2: Concrete stresses over time as a result of immediate and sustained shrinkage

The gradual reduction in tensile stress over time may be modelled as step wise drops in tensile stress. Each drop in tensile stress is effectively the same as an application of compressive stress. Each compressive stress is also subject to creep, as the original tensile stress is.

A.1.1. Effective modulus method (EMM)


The EMM adjusts the modulus of elasticity to include the effect of creep. Instead of using a value of E based on instantaneous stiffness, it is reduced to take into consideration the creep that has occurred over time. Consider equation 3.3, now in terms of stress;
= = =

+ , + () 1 + , + () 1 + , + ( ) 1 + , +

Where (, ) = )( (1 + (, )) and is known as the effective modulus. The EMM assumes the concrete stress applied at the end of the time history, is applied constantly from the beginning to the end of the time history. In the example, this means creep will be based on the final tensile concrete stress as shown in figure A.3, which is less than the

+ ,

(A.1)

52

tensile stress at the beginning of the time history, thus creep will be underestimated, stresses over estimated, and total contraction of the beam over-estimated as shown in figure A.3.
2.5

Concrete stress vs time


This region of stress not accounted for under EMM in calculation of creep and final strain

2.0

Concrete Stress (Mpa)

1.5

1.0

Creep based on final stress value

0.5

0.0 10 100 1000 Time (days) Figure A.3: Concrete stresses over time calculated using EMM 10000 100000

A.1.2. Age-adjusted effective modulus method (AEMM)


The AEMM is similar to the EMM, however creep is calculated from two components. The first is based on the initial stress at t=0 with the effective modulus as per the EMM. The second is based on the total change in stress over the time period with an adjusted effective modulus. The adjustment is required because the change in stress is gradual and the resulting creep is reduced. The adjusted effective modulus is given as:
, = 1 + , (, )

(A.2)

Where , is the ageing coefficient and is less than one. Determination of the ageing coefficient requires knowledge of the step by step method and is not readily available. However, for load durations greater than 100 days, the following values provide reasonably accurate results (Ranzi & Gilbert 2011); Constant load , = 0.65

Constant deformation , = 0.80 The results in figure A.4 agree with those of the step by step method closely.

53

Creep based partly on initial stress

2.5

Concrete stress vs time

value using effective modulus as defined in EMM

Concrete Stress (MPa)

2.0
Creep also based on the total change in stress using the adjusted effective modulus

1.5

1.0

0.5

0.0 10 100 1000 Time (days)


Figure A.4: Concrete stresses over time calculated using AEMM

10000

100000

A.1.3. Rate of creep method (RCM)


The RCM assumes that the rate of creep that occurs at a given point in time is the same regardless of when the load is first applied. This means that the creep curves are parallel as shown in figure A.5.
(, ) ( , ) (, )

, (Actual) (, ) = (, ) , (RCM Assumed)

Time

Figure A.5: Creep assumptions in the RCM

54

Because the assumed rate of creep for loads applied after the first is greater than the rate of creep that is actually occurring (refer to curve for , in figure A.5), creep strains are over estimated for increasing load histories, and under estimated for decreasing load histories. In the current example, the compressive stresses in the concrete are increasing, so that compressive creep strains are over estimated. This results in lower stresses, and under estimation of contraction in the beam.

A.1.4. Step by step method (SSM)


This method is outlined in section 3.2

A.1.5. Stress and strain results


Resulting concrete stresses over time and total beam contraction (strain) are shown in figure A.6 and A.7 respectively.
2.2 2.1 Concrete Stress (MPa) 2.0 1.9 1.8 1.7 1.6 1.5 1.4 1.3 10 100 1000 Time (days) Figure A.6: Concrete stress over time as a result of instant and constant shrinkage 10000 100000 EMM AEMM SSM RCM

Concrete stress vs time

Total strain vs time


10 -300E-6 -350E-6 -400E-6 -450E-6 -500E-6 EMM -550E-6 Figure A.7: Strain over time as a result of instant and constant shrinkage AEMM SSM RCM 100 1000 Time (Days) 10000 100000

Strain

55

A.2.Constant load
Methods for analysing time dependent properties may also be compared by looking at resulting stresses and strains in a short column under constant load, as shown in figure A.8. In this example, the column is free to deform, and shrinkage is ignored. 50 kN 100mm 100mm Where: fc = 40 MPa As = 200 mm2

50 kN
Figure A.8: concrete column subject to constant load and creep

The time steps, elastic modulus, and shrinkage values used in the calculations are shown in table A.2
Table A.2: Elastic modulus and shrinkage values used in constant loading example Time (days) sh(t) E() 14 0.00 0 26.8E+3 50 2.22 0 29.3E+3 100 2.62 0 30.1E+3 200 2.85 0 30.8E+3 400 2.98 0 31.2E+3 1000 3.08 0 31.6E+3 2000 3.12 0 31.8E+3 5000 3.15 0 32.0E+3 10000 3.16 0 32.1E+3

Resulting total strain and concrete and steel stresses calculated by the four methods are shown in figure A.9 to A.11.

56

600E-6 500E-6 400E-6 Strain 300E-6 200E-6 100E-6 0 10 100 EMM

Total strain vs time

AEMM

SSM

RCM Time (Days)

1000

10000

100000

Figure A.9: Axial strain over time as calculated by each cross sectional method

4.4 4.2 Concrete Stress (MPa) 4.0 3.8 3.6 3.4 3.2 3.0 2.8 2.6 10 100

Concrete stress vs time

EMM

AEMM

SSM

RCM

Time (days) 1000 10000 100000

Figure A.10: Concrete stress over time as calculated by each cross sectional method

120.0 100.0 Concrete Stress (MPa) 80.0 60.0 40.0 EMM 20.0

Steel stress vs time

AEMM

SSM

RCM

Time (days) 0.0 10 100 1000 10000 100000

Figure A.11: Steel stress over time as calculated by each cross sectional method

57

For constant load, the EMM calculates creep at the final time step based on the final concrete stress. Due to creep, the column has been squashed. This deformation increases stress in the steel, and reduces the stress in the concrete. Because the creep is based on this final value of concrete stress, it is lower than it should be, and total strain is therefore understated. Conversely the RCM over estimates the rate of creep, and as a result the total creep and total strain are overstated. This leads to greater stresses in the steel and reduced stresses in the concrete than given by the more accurate SSM. AEMM gives similar results to the SSM as expected, as the rate of creep is based both on the initial concrete stress, and the gradual reduction in compressive stress. The reason for the slight difference between the two methods is the value for , which has been set at 0.8. A more methods under scenarios such as this, where loads do not change over time. accurate value for , (based on the SSM) would produce an exact match for the two

58

Appendix B Step by step cross sectional analysis formulation


At a cross section, at any time, the internal forces will be equal to the external applied forces at a cross section. The external forces and moments at a cross section refer to the external moments and axial forces that would need to be applied to maintain the position of the beam if the beam was cut at that point. Thus:
= =

(4.1) (4.2)

Ne and Me are the externally applied axial loads and moments, and Ni and Mi are the equal and opposite internal resisting forces, a portion of which comes from the steel, and a portion of which comes from the concrete. Thus;
= + = +

(4.3) (4.4)

The forces and moments in the concrete are given by equations 5.11 and 5.12 respectively.
= ,

(4.5) (4.6)

= ,

Combining equations 5.3, 5.5, and 5.11 and 5.12 yields:


= , (, + , ) + ,, ,

(4.7)

= , (, + , ) + ,, ,

(4.8)

After some manipulation, the resulting equilibrium equation is found to be: Equation 5.13 can be rearranged as follows;
= , , + , , , + , , ,

(B.1)

Where = , and , = ,

Equation 5.14 can also be rearranged in a similar fashion; 59

= , , + , , , + ,, ,

(B.2)

Where = and , = ,

Equations 5.15 and 5.16 can be more conveniently stated as:


= , , (, , ) + ,, + ,, , = ,, (, , ) + , , + ,, ,

(B.3) (B.4)

Where ,, = , , ,, = , and ,, = , As with concrete, the forces and moments in steel may be expressed as;
= ,

(B.5) (B.6)

= ,

Equations 5.4, 5.5 and 5.19 and 5.20 can be combined to give:
= (, + )

(B.7) (B.8)

= (, + )

Or in more compact form:


= , , + ,

= , , + ,

(B.9) (B.10)

Where , = , , = = and , = = Here is not strictly the second moment of area of the steel. The parallel axis theorem gives the second moment of area about an axis other than its own as = + . The calculation for in equation 5.23 ignores , however this is not significant as the area of the steel is small relative to the concrete (usually about 2% (Standards Australia 2009)). Combining equations 5.16, 5.17, 5.22 and 5.23 gives:
= , , + , ,, , + ,, ,

(B.11)

60

= , , + , ,, , + , , ,

(B.12)

Where , = , + ,, , , = , + ,, and , = , + , , Using equations 5.7, 5.8, 5.24 and 5.25 it is possible to express equilibrium in matrix form as follows:
, = + , ,

(4.9)

Where
, , , = , = , , , , , , , , = , , = , , , , = , , , , ,

(4.10a-e)

Strains are then solved for, as shown in equation 5.27.


= , , + ,

(4.11)

61

Appendix C Finite beam element formulation


C.1.Displacement field
For any particular point P in the beam, the generalised displacement model is shown in figure C.1, which shows the admissible movement of that point in the orthogonal axis system y,z as the beam deflects.

z y yr uk P Where: yr is distance from the centroid to the reference z axis. y is the vertical distance to the point P P = v For small
( ) ( ) =

vj

y-yr (y-yr)cos

(y-yr)sin

Figure C.1: Admissible displacement field under the Euler-Bernoulli beam assumptions

From figure C.1, the displacement field is given as:


, = + ( ) =

(C.1)

Where the only non-vanishing strain is given as:


. = =

(C.2)

C.2. Weak Formulation


The principle of virtual work is used to derive the weak formulation (the weak formulation is also referred to as the global balance condition because it expresses the balance between stiffness and displacement and loading for the element as a whole). Consider the generalised beam loading shown in figure C.2, where subscript L refers to point loads on the left of the beam and subscript R to point loads on the right. n and m are the distributed axial loads and moments respectively. 62

p(z) ML NL SL z n(z) SR L ML NR

Figure C.2: Generalised beam loading

For a virtual displacement of the beam, internal work is equated to external work in equation C.3. From C.2 the only nonvanishing strain was found to be . The hat sign represents a virtual displacement.
+ = + + + + + +

(C.3)

Combining equations C.1 and C.3 gives:


+ [ ] = + + + + + +

(C.4)

Recalling that = and = gives:


= + + + + +

(C.5)

Equation C.5 can be expressed in matrix form as:


. = . + . + .

(C.6)

Since the nodal forces on the right hand side can be incorporated in the assembly of the loading vector when solving the system, it is ignored for now. This gives:
= . .

(C.7)

or more simply:
= . .

(C.8)

= , = , and = = . where = ,

Equation C.8 is a statement of global equilibrium. The strain vector may be expressed as a function of displacement as:

63

= 0

Equation C.9 can be combined with C.8 to give


= . .

0 =

(C.9)

where: =

(5.1)

This is the weak formulation for a beam element.

64

Appendix D Matlab finite element program

Figure D.1: Main GUI input for FEM

65

Figure D.2: Properties GUI input for FEM

66

References
Unsupported source type (InternetSite) for source Ste12. Abramowitz, M & Stegun, I 1964, Handbook of Mathematical Functions, 10th edn, NY Wiley, New York, US. Al-Deen, S, Ranzi, G & Uy, B 2012, ' Non-uniform shrinkage in simply-supported composite steel-concrete slabs', Steel & Composite Structures, vol under review. Altoubat, SA 2000, Early age stresses and creep-shrinkage interaction of restrained concrete, University of Illinois at Urbana-Champaign, Illinois, United States of America. Bazant, Z 1975, 'Theory of Creep and Shrinkage in Concrete Structures: A Precis of Recent Developments', Mechanics Today, vol 2, pp. 1-93. Bazant, Z 1988, 'State of the art in mathematical modeling of creep and shrinkage of concrete', RILEM Committee TC-69, John Wiley and Sons, Chichester and New York. Bazant, Z 1997, 'Prediction of concrete creep and shrinkage: past, present and future', Joint WANO/OECD-NEA Workshop on Prestress Losses in NPP (Nuclear Power Plant) Containments, OECD Nuclear Energy Agency, Civaux, France. Bazant, Z & Chern, J 1985, 'Concrete creeep at variable humidity: consitutive law and mechanism', RILEM: Materials and Structures, vol 18, pp. 1-20. Bazant, Z, Cusatis, G & Cedolin, L 2004, 'Temperature effect on concrete creep modeled by microprestress-solidification theory', Journal of Engineering Mechanics, vol 130, no. 6, pp. 691699. Bazant, Z, Hauggaard, A, Baweja, S & Ulm, F 1997, 'Microprestress-Solidification Theory for concrete creep. I: Aging and drying effects', Journal of Engineering Mechanics, vol 123, no. 11, pp. 1188-1194. Bazant, Z & Xi, Y 1994, 'Drying creep of concrete: consitutive model and new experiments separating its mechanisms', Materials and Structures, vol 27, pp. 3-14. Bissonnette, B 2007, 'Tensile creep of concrete: study of its sensitivity to basic parameters', ACI Materials Journal, vol 104, no. 4, pp. 360-368. Bissonnette, B & Pigeon, M 1995, 'Tensile creep at early ages of ordinary, silica fume and fiber reinforced concretes', Cement and Concrete Research, vol 25, no. 5, pp. 1075-1085. 67

Cement Concrete and Aggregates Australia 2007, 'Moisture in Concrete and Moisture-sensitive Finishes and Coatings', Datasheet, Sydney. Comit Euro-International du Bton 1993, CEB-FIB Model Code 1990: Design Code., Thomas Telford, London. Faria, R, Azenha, M & Figueiras, J 2006, 'Modelling of concrete at early ages: Application to an externally restrained slab', Cement & Concrete Composites, vol 28, pp. 572-585. Gilbert, I & Nejadi, S 2004, 'An experimental study of flexural cracking in reinforced concrete members under sustained loads', in Studies from School of Civil and Environmental Engineering, University of New South Wales, School of Civil and Environmental Engineering, Sydney, Australia. Grasley, Z & Lange, D 2004, 'Modeling Drying Shrinkage Stress Gradients in Concrete', Cement, Concrete and Aggregates, vol 6, no. 2. Grasley, Z, Lange, D & D'Ambrosia, M 2006, 'Internal relative humidity and drying stress gradients in concrete', Materials and Structures, vol 39, pp. 901-909. International Union of Pure and Applied Chemistry 2012, 'Gold Book'. Kawando, A & Warner, R 1996, 'Model Formulations for Numerical Creep Calculations for Concrete', Journal of Structural Engineering, vol 122, no. 3, pp. 284-290. Kawano, A & Warner, F 1996, 'Model Formulations for Numerical Creep Calculations for Concrete', Journal of Structural Engineering, vol 122, no. 3, pp. 284-290. Kotsovos, M & Pavlovic, M 1995, Structural Concrete: Finite-element Analysis for Limit-state Design, Telford, New York. Nawy, E 2008, Concrete Construction Engineering Handbook, 2nd edn, CRC Press, Florida, United States. Nehdi, M & Soliman, A 2011, 'Early-age properties of concrete: overview of fundamental concepts and state-of-the-art research', Construction Materials, vol 164, no. CM2, pp. 57-77. Neville, AM 1970, Creep of Concrete: Plain, Reinforced and Prestressed, North-Holland Publishing Company, Amsterdam, Holland. Neville, AM 1995, Properties of Concrete, 4th edn, Pearson Education Limited, Essex, UK.

68

stergaard, L, Lange, D, Altoubat, S & Stang, H 2001, 'Tensile basic creep of early-age concrete under constant load', Cement and Concrete Research, vol 31, pp. 1895-1899. Rajabipour, F, Sant, G & Weiss, J 2007, 'Development of Electrical Conductivity-Based Sensors for Health Monitoring of Concrete Materials', Transportation Research Board, Washington, DC. Ranzi, G & Gilbert, RI 2011, Time-dependent behaviour of concrete structures, Spon Press, Oxford, UK. RILEM 1988, Mathematical Modeling of Creep and Shrinkage of Concrete, John Wiley & Sons Ltd, Essex, UK. Slowik, V & Ju, JW 2011, 'Discrete modeling of plastic cement paste subjected to drying', Cement & Concrete Composites, vol 33, pp. 925-935. Slowik, V, Schmidt, M & Fritzsch, R 2008, 'Capillary pressure in fresh cement-based materials and identi cation of the air entry value', Cement & Concrete Composites, vol 30, pp. 557-565. Standards Australia 2009, 'AS3600 Concrete Structures', Sydney. Sutherland, J, Humm, D & Chrimes, M 2001, Historic concrete background to appraisal, Thomas Telford Publishing, Longon. Vandamme, M & Ulm, F-J 2009, 'Nanogranular origin of concrete creep', PNAS, vol 106, no. 26, pp. 10552-10557. Weiss, J 1999, 'Prediction of Early-Age Shrinkage Cracking in Concrete Elements', PhD Thesis, Northwestern University, Illinois. Wight, J & Macgregor, J 2012, Reinforced concrete mechanics and design, 6th edn, Pearson Education, New Jersey, US. Wittmann , FH 1976, 'On the action of capillary pressure in fresh concrete', Cement and Concrete Research, vol 6, pp. 49-56. Wittmann, FH & Roelfstra, PE 1980, 'Total deformation of loaded drying concrete', Cement and Concrete Research, vol 10, pp. 601-610. Yang, Y, Sato, R & Kawai, K 2005, 'Autogenous shrinkage of high-strength concrete containing silica fume under drying at early ages', Cement and Concrete Research, vol 35, pp. 449 - 456. 69

You might also like