Вы находитесь на странице: 1из 107

“HOT COMPRESSIVE BEHAVIOR OF CONCRETE:

TESTING METHODS”

Dissertation submitted in partial fulfilment of the requirements

of the degree of

MASTER OF TECHNOLOGY
IN

ENGINEERING STRUCTURES
By

AANCHAL JASUJA
(ROLL NO: 131501)

Under the esteemed guidance of


Dr. C.B. KAMESHWAR RAO

DEPARTMENT OF CIVIL ENGINEERING


NATIONAL INSTITUTE OF TECHNOLOGY,
WARANGAL- 506004
DEPARTMENT OF CIVIL ENGINEERING

NATIONAL INSTITUTE OF TECHNOLOGY,

WARANGAL

(DEEMED UNIVERSITY)

CERTIFICATE

This is to certify that the dissertation work entitled “HOT


COMPRESSIVE BEHAVIOR OF CONCRETE: TESTING
METHODS” is a bonafide work done by Aanchal Jasuja (Roll
No.131501) carried out under my guidance during the academic year 2014-
2015 in the partial fulfilment of the requirements for the award of the
degree of Master of Technology in Engineering Structures.

(Roberto Felicetti)
Professor,
Department of Civil
and Environmental
Engineering,
Politecnico Di Milano,
Milan, Italy - 20133
II
DEPERTMENT OF CIVIL AND ENVIRONMENTAL ENGINEERING

Milano, July 8, 2015

CERTIFICATE

I, the undersigned, certify that Aanchal Jasuja (Student ID: 833884), candidate for

the award of Master of Technology in specialization of ENGINEERING

STRUCTURES (Department of Civil Engineering) from National Institute of

Technology, Warangal, India has worked on “Hot Compressive Behavior Of Concrete:

Testing Methods” during the academic year 2014-2015.

During her thesis period of last 5 months she has really worked hard and was able

to meet the aspirations by us, well in time. And the work carried out is interesting and

very well documented.

(Signature) (Signature)
Prof. Roberto Felicetti Prof. Marco Di Prisco
Professor and Guide Professor and head
Department of Civil and Environmental Department of Civil and Environmental
Engineering Engineering
Politecnico di Milano, Italy Politecnico di Milano, Italy

III
Approval Sheet
This project work entitled “HOT COMPRESSIVE BEHAVIOR OF CONCRETE:

TESTING METHODS” by Ms. AANCHAL JASUJA (131501), is approved in

partial fulfillment of the requirements for the award of degree of Master of Technology

in Civil Engineering with specialization in Engineering Structures, from National

institute of Technology, Warangal.

Examiners

__________________________

__________________________

__________________________

Supervisor (s)

(Signature) (Signature)
Prof. Roberto Felicetti Dr. Francesco Lo Monte
Professor Lecturer
Civil and Environmental Engineering
Civil and Environmental Engineering
Department Department
Politecnico di Milano, Italy Politecnico di Milano, Italy

Chairman
___________________________
(Signature)

Dr. Deva Pratap

Professor and Head

Department of Civil Engineering

National Institute of Technology

Warangal

Date: _____________________

IV
Place: _____________________
Declaration
I declare that this written submission represents my ideas in my own words and where
others ideas or words have been included, I have adequately cited and referenced the
original sources. I also declare that I have adhered to all principles of academic honesty
and integrity and have not misrepresented or fabricated or falsified any idea/ data/ fact/
source in my submission. I understand that any violation of the above will be cause for
disciplinary action by the institute and can also evoke penal action from the sources
which have thus not been properly cited or from whom permission has not been taken
when needed.

________________________
(Signature)
AANCHAL JASUJA
Roll No: 131501
Date: __________
Place: NIT Warangal

V
ABSTRACT

The increasing use of High-Strength Concrete in many structures exposed to extreme


environmental conditions (tunnels, off-shore platforms, Liquefied Natural Gas
terminals, containment shells), and the need to repair/strengthen many existing
structures in order to meet the most recent code provisions or to increase their service
life (e.g. nuclear power plants), bring in new problems, which go beyond the excellent
knowledge we have on the behaviour of ordinary concrete in extreme conditions.
At high temperature, in fact, High-Strength Concrete shows to be more heat
sensitive than ordinary concrete in terms of mechanical behaviour, mainly due to its
denser matrix that leads to a more brittle response and to the development of higher
values of pore pressure during heating. It is very important to design the testing methods
for concrete induced to thermal stresses.
In the present research project, the mechanical and thermal behaviour of concrete
is studied under uniaxial compression at high temperature (T = 250°C). Results are
compared with the mechanical properties in virgin conditions. Numerical models are
built up aimed at focusing on the influence of heating rate and thermal stress, via the
finite element method. This investigation, however, will be part of an ongoing work.

VI
ABSTRACT …………………………………………………………………………VI
LIST OF FIGURES ................................................................................................. XIV
LIST OF TABLES ................................................................................................... XIV
Chapter 1 FIRE ENGINEERING ....................................................................... 15
1.1 Introduction to Fire Engineering ................................................................... 15
1.2 Sources and Mechanisms of Fire .................................................................. 17
1.3 Fire Scenario ................................................................................................. 18
1.3.1 Buildings .................................................................................................... 18
1.3.2 Offshore Structures and Petrochemical Plants .......................................... 19
1.3.3 Tunnels ...................................................................................................... 20
1.3.4 Bridges ....................................................................................................... 22
Chapter 2 LITERATURE REVIEW ...................................................................... 23
Chapter 3 CONCRETE BEHAVIOUR AT HIGH TEMPERATURE.................. 36
3.1 Thermal and Physical Properties of Concrete ............................................... 36
3.1.1 Porosity and Density.................................................................................. 36
3.1.2 Permeability ............................................................................................... 37
3.1.3 Thermal Expansion .................................................................................... 39
3.1.4 Thermal Conductivity ................................................................................ 40
3.1.5 Specific Heat.............................................................................................. 41
3.1.6 Thermal Diffusivity..................................................................................... 42
3.2 Mechanical Properties ................................................................................... 43
3.2.1 Compressive Strength ................................................................................ 44
3.2.2 Tensile Strength ......................................................................................... 46
3.2.3 Poisson Ratio ............................................................................................. 47
3.2.4 Elastic Modulus ......................................................................................... 48
3.2.5 Fracture Energy ......................................................................................... 49
3.3 Spalling at High Temperature........................................................................ 50
Chapter 4 THE THERMAL PROBLEM .......................................................... 52
4.1 Mechanisms of Heat Transport ..................................................................... 52
4.1.1 Conduction................................................................................................. 52
4.1.2 Convection ................................................................................................. 53
VII
4.1.3 Radiation .................................................................................................... 53
4.2 Thermal Problem ........................................................................................... 54
4.2.1 General Equations...................................................................................... 54
4.2.2 Initial and Boundary Conditions................................................................ 56
4.2.3 Initial and Boundary Conditions in Fire Design ........................................ 56
4.2.4 Calibration of LVDT Measurement ........................................................... 57
Chapter 5 THE HYGRAL PROBLEM .............................................................. 59
5.1 Liquid-Vapour Equilibrium ........................................................................... 61
5.2 Moisture Transport ........................................................................................ 62
5.2.1 Physical Mechanisms in Hygral Problems ................................................ 62
5.3 Hygro-Thermal Problem ............................................................................... 63
Chapter 6 NUMERICAL ANALYSIS .................................................................... 65
6.1 Introduction ................................................................................................... 65
6.2 Description of the Concrete Model ............................................................... 65
6.2.1 Concrete Model ......................................................................................... 65
6.2.2 Properties of Concrete ............................................................................... 68
6.2.2.1 Thermal Properties ............................................................................. 68
6.2.2.2 Mechanical Properties ........................................................................ 69
6.2.3 Properties of Other Materials .................................................................... 71
6.2.3.1 Thermal Properties of Steel Disc and Steel Actuator ......................... 71
6.2.3.2 Thermal Properties of Stone Disc ...................................................... 71
6.2.3.3 Mechanical Properties of Steel Disc, Steel Actuator and the Stone Disc
71
6.3 Thermal Analysis in Abaqus ......................................................................... 72
6.3.1 Steps........................................................................................................... 72
6.3.2 Boundary Conditions ................................................................................. 72
6.3.3 Predefined Fields ....................................................................................... 72
6.3.4 Interaction .................................................................................................. 72
6.3.5 Mesh .......................................................................................................... 73
6.4 Mechanical Analysis in Abaqus .................................................................... 73
6.4.1 Steps........................................................................................................... 73
6.4.2 Boundary Conditions ................................................................................. 74
6.4.3 Predefined Fields ....................................................................................... 74
VIII
6.4.4 Interaction .................................................................................................. 75
6.4.5 Mesh .......................................................................................................... 75
Chapter 7 RESULTS AND CONCLUSIONS ....................................................... 76
7.1 Temperature Monitoring and Numerical Modelling ............................ 76
7.2 Temperature profile obtained from Thermal Analysis ........................ 76
7.3 Results obtained from Mechanical Analysis ......................................... 81
Chapter 8 SUMMARY............................................................................................ 97
REFERENCES ........................................................................................................... 99
ACKNOWLEDGEMENTS ....................................................................................... 103

IX
LIST OF FIGURES

Figure 1-1 Time-temperature curve for full process of fire development (Buchanan
A.H., 2001). ................................................................................................................. 16
Figure 1-2 damaged columns due to spalling exposed to fire (Pentagon, Washington,
USA, 2001). ................................................................................................................. 17
Figure 1- 3 Standard Fire curve according to EN 1991-1-2 (2004). ............................ 19
Figure 1- 4 Hydrocarbon curve (Mobil Oil Company, 1970s). ................................... 20
Figure 1- 5 RWS Fire curve (Rijkswaterstaat and TNO, 1979). .................................. 20
Figure 1- 6 Fires in (a) Channel Tunnel (1996) and (b) the Mont Blanc Tunnel (1999).
...................................................................................................................................... 21
Figure 1- 7 RABT Fire curves (Germany). .................................................................. 21
Figure 1- 8 MacArthur Maze collapsed due to fire (Giuliani et al., 2012). ................. 22
Figure 2-1 Concrete microstructure (Aïtcin and Neville, 1993).Error! Bookmark not
defined.
Figure 2-2 Concrete thermometer (Khoury, 2000). ..... Error! Bookmark not defined.
Figure 2-3 Portland cement paste porosity as a function of temperature (Harmathy,
1970). ........................................................................... Error! Bookmark not defined.
Figure 2-4 Density of concrete with different coarse aggregates (Harmathy, 1970).
...................................................................................... Error! Bookmark not defined.
Figure 2-5 (a) Concrete intrinsic permeability (Kalifa and Tsimbrovska, 1998); (b)
concrete relative intrinsic permeability (Mindeguia et al., 2013).Error! Bookmark
not defined.
Figure 2-6 Scanning Electron Microscopy images of samples after heating at 400°C
(Mindeguia et al., 2013). .............................................. Error! Bookmark not defined.
Figure 2-7 Thermal strain of Portland cement paste (Bazant and Kaplan, 1996).Error!
Bookmark not defined.
Figure 2-8 Thermal strain for different aggregates (Bazant and Kaplan, 1996). . Error!
Bookmark not defined.
Figure 2-9 Thermal strain of Portland cement, mortar and concrete (Cruz and Gillen,

X
1980). ........................................................................... Error! Bookmark not defined.
Figure 2-10 Thermal conductivity of initially-saturated concrete (Blundell et al., 1976).
...................................................................................... Error! Bookmark not defined.
Figure 2-1l1 Thermal conductivity for different types of concrete (Harmathy and Allen,
1973). ........................................................................... Error! Bookmark not defined.
Figure 2-12 (a) Thermal conductivity as a function of density and moisture content; (b)
thermal conductivity as a function of temperature (Shin et al., 2002). ................ Error!
Bookmark not defined.
Figure 2-13 (a) Specific heat of various concrete (Bazant and Kaplan, 1996); (b)
specific heat of a Korean nuclear power plant concrete (Shin et al., 2002). ....... Error!
Bookmark not defined.
Figure 2-14 Thermal diffusivity of normal and lightweight concrete (Harmathy and
Allen, 1973). ................................................................ Error! Bookmark not defined.
Figure 2-15 Temperature effect on stress-strain curves of concrete.Error! Bookmark
not defined.
Figure 2-16 Residual compressive strength as a function of temperature according to
(a) EN1994-1-2 (2004) and (b) ACI 216-1.07 (2007). Error! Bookmark not defined.
Figure 2-17 Residual compressive strength of two HPC and two UHPC: (a) actual
value; (b) values normalized with respect to the virgin strength (Khoury, 2000).Error!
Bookmark not defined.
Figure 2-18 Decay of the normalized tensile strength in hot conditions (EN1992-1-2,
2004). ........................................................................... Error! Bookmark not defined.
Figure 2-19 Tensile strength of HSC and NSC in residual conditions (Noumowe et al.,
1996). ........................................................................... Error! Bookmark not defined.
Figure 2-20 (a) Normalized stress as a function of Poisson ratio (Ehm, 1985); (b)
Poisson ratio as a function of temperature (Marechal, 1972).Error! Bookmark not
defined.
Figure 2-21 Effect of aggregate type and concrete strength on Poisson ratio (Cruz,
1966). ........................................................................... Error! Bookmark not defined.
Figure 2-22 (a) Normalized concrete elastic modulus for different aggregate types
according to fib (2007); (b) for different sealing conditions according to Nanstad
(1976). .......................................................................... Error! Bookmark not defined.
Figure 2-23 Normalized plot of fracture energy Gf as a function of temperature
(Heinfling, 1998).......................................................... Error! Bookmark not defined.
XI
Figure 2-24 Explosive spalling: empirical envelope for Normal-Strength Concrete,
showing the influence of moisture content and applied stress (Khoury, 2000). .. Error!
Bookmark not defined.
Figure 3-1 Heat transfer within a continuous plate. ..... Error! Bookmark not defined.
Figure 3-2 Heat flux balance........................................ Error! Bookmark not defined.
Figure 4-1 Types of water existing in concrete (Feldman and Sereda, 1968). .... Error!
Bookmark not defined.
Figure 4-2 Schematic representation (a) of a capillary pore and (b) of a pore with liquid
water, vapour and dry air (after Meftah, 2005). ........... Error! Bookmark not defined.
Figure 4-3 Plot Mass transport mechanisms in an unsaturated porous material during
heating (after Meftah, 2005). ....................................... Error! Bookmark not defined.
Figure 5-1 Plots of the residual compressive strength, on cylinders fc and on cubes Rc.
...................................................................................... Error! Bookmark not defined.
Figure 5-2 Sketch of the insulating capsule and of the specimen: (a) ready to be
introduced into the furnace; (b) set up during the test; (c) picture of two capsules inside
the furnace; (d) picture of a capsule placed between the platens and the press. .. Error!
Bookmark not defined.
Figure 5-3 Normalized results of compressive strength compared with the hot curves
suggested in EC2 for NSC and HPC, and residual test.Error! Bookmark not defined.
Figure 5-4 Stress-strain curves for fibre-reinforced concrete at various temperatures.
...................................................................................... Error! Bookmark not defined.
Figure 5-5 Test set-up for high temperature compressive and tensile strength tests.
...................................................................................... Error! Bookmark not defined.
Figure 5-6 Thermal jacket and insulated steel bracket frame.Error! Bookmark not
defined.
Figure 5-7 Compressive strength of SCC and FRSCC as function of temperature.
...................................................................................... Error! Bookmark not defined.
Figure 5-8 (a) Variation of compressive strength with temperature; (b) variation of
elastic modulus with temperature. ............................... Error! Bookmark not defined.
Figure 5-9 Stress-strain relationships of (a) HSC made with fly ash; (b) NSC made with
ordinary Portland cement. ............................................ Error! Bookmark not defined.
Figure 5-10 (a) Compressive strength of 90-day concrete vs. temperature; (b) Elastic
modulus of 90-day concrete vs. temperature. .............. Error! Bookmark not defined.
Figure 5-11 (a) Compressive strength of 150-day concrete vs. temperature; (b) Elastic
XII
modulus of 150-day concrete vs. temperature. ............ Error! Bookmark not defined.
Figure 5-12 (a) Mass loss vs. Strength loss for 90-day concretes; (b) Mass loss vs.
Strength loss for 150-day concretes. ............................ Error! Bookmark not defined.
Figure 5-13 (a) Residual concrete strength vs. temperature; (b) Residual elastic modulus
vs. temperature. ............................................................ Error! Bookmark not defined.
Figure 5-14 Stress-strain curves for concrete after heating: (a) 95 MPa; (b) 72 MPa.
...................................................................................... Error! Bookmark not defined.
Figure 5-15 Thermal cycles for the reference temperature of 500°C and 750°C. Error!
Bookmark not defined.
Figure 5-16 Heating system: electric furnace. ............. Error! Bookmark not defined.
Figure 5-17 Insulated specimens ready for heating. .... Error! Bookmark not defined.
Figure 5-18 Measurement devices for displacement and strain: (a) LVDT, (b)
Controller, (c) DD1 in a trial test, and (d) Schematic diagram of the sensors’ positions.
...................................................................................... Error! Bookmark not defined.
Figure 5-19 Thermocouples plugged into the specimen: (a) specimen in furnace ready
for thermal check, and (b) position of thermocouples. Error! Bookmark not defined.
Figure 5-20 Hot compressive tests: (a) specimen in the furnace before heating, and (b)
specimen in the compression testing machine. ............ Error! Bookmark not defined.
Figure 5-21 Sketch of an eighth specimen modelled in ABAQUS.Error! Bookmark
not defined.
Figure 5-22 Temperature development within the specimen exposed to thermal cycle
up to 750°C. ................................................................. Error! Bookmark not defined.
Figure 5-23 Numerical modelling: thermal fields of an eighth specimen (a) before
extracted from the furnace, and after cooling for (b) 10 min, (c) 20 min, and (d) 30 min.
...................................................................................... Error! Bookmark not defined.
Figure 5-24 Mean stress-strain curves in uniaxial compression for five mixes: M45 S,
M95 S, M70 B, M70 C and M70 S. Influence of concrete grade and aggregate type.
...................................................................................... Error! Bookmark not defined.
Figure 5-25 Mean stress-strain curves in uniaxial compression for six mixes: M70 S
PM05, M70 S PM1, M70 S PM2, M70 S PF2, M70 S SF40 and M70 S SF60. Influence
of fibre type and content. ............................................. Error! Bookmark not defined.
Figure 5-26 Mean stress-strain curves in uniaxial compression for all mixes at different
temperatures. ................................................................ Error! Bookmark not defined.
Figure 5-27 Comparison between hot and residual conditions in terms of: (a)
XIII
normalized compressive strength, (b) normalized elastic modulus, (c) strain at the peak
stress, and comparison between (d) normalized hot compressive strength and curves
given by EC2 and ACI. ................................................ Error! Bookmark not defined.

XIV
LIST OF TABLES

Table 2-1 Concrete Mix Design (Bamonte and Gambarova , 2009)Error! Bookmark
not defined.
Table 2-2 Concrete Mix Design (Lie and Kodur, 1995)Error! Bookmark not
defined.
Table 2-3 Concrete Mix Design (Khaliq and Kodur, 2011)Error! Bookmark not
defined.
Table 2-4 Concrete Mix Design (Diederichs et al., 1988)Error! Bookmark not
defined.
Table 2-5 Concrete Mix Design (Hammer, 1995) ....... Error! Bookmark not defined.
Table 2-6 Mix designs and mechanical properties of concrete mixes ................. Error!
Bookmark not defined.
Table 2-7 Experimental plan: experimental campaign and amount of specimens for
testing ........................................................................... Error! Bookmark not defined.
Table 2-8 Position of thermocouples ........................... Error! Bookmark not defined.
Table 2-9 Physico-mechanical properties of concrete mixes in virgin conditionsError!
Bookmark not defined.
Table 6-1 Thermal properties of concrete ................. Error! Bookmark not defined.7
Table 6-2 Mechanical properties of concrete ............ Error! Bookmark not defined.8
Table 6-3 Thermal properties of steel ..................... Error! Bookmark not defined.70
Table 7-1 Temperature profile at different time intervals for the top surface AC Error!
Bookmark not defined.77
Table 7-2 Temperature profile at different time intervals for the inner vertical surface
AC ........................................................................... Error! Bookmark not defined.78
Table 7-3 Temperature profile at different time intervals for the outer vertical surface
BD ........................................................................... Error! Bookmark not defined.79
XVI
Chapter 1 FIRE ENGINEERING

1.1 Introduction to Fire Engineering

Fire, used to generate heat and light, brought a dramatic change in the habits of early humans.
Nevertheless fire is seen a symbol of progress, it may become a destructive force for the structures,
causing heavy consequences for people, great economic losses and severe damage to the structural
bearing capacity. Structural fire safety is now one of the key points in the design of buildings and
it is attracting worldwide attention. According to the Institute of Fire Engineering in UK, Structural
Fire Engineering is:
The application of scientific and engineering principles, rules (codes), and expert
judgement, based on an understanding of the phenomena and effects of fire and of
the reaction and behaviour of people to fire, to protect people, property and the
environment from the destructive effects of fire. (IFE, 2014)
There are two strategies of designing safe structures for a given severity: a) to prevent ignition
or fire spreading outside the compartment to other parts of the building, by applying active or
passive fire protection systems and b) to ensure the fire resistance of the structure in order to
provide enough time to evacuate people, before the occurrence of collapse. Hence, time is the
critical parameter for the strategies.
The two strategies are represented in Figure 1-1 for a typical fire development in a
compartment, with the assumption of no suppression devices or fire fighters to prevent fire
spreading. Compartment fire development can be divided into four stages: incipient, growth,
burning and decay.
Flashover is the sudden transition from the growth stage (pre-flashover) to the burning stage
(post-flashover, also known as fully developed stage). The phenomenon involves a rapid transition
to the phase in which all the directly exposed combustible material within the compartment is
burning. In the post-flashover stage, the rate of heat release and the temperature in the compartment
increase rapidly, resulting in high temperature. The performance of a structural member exposed
to fire depends on the thermal and mechanical properties of the materials of which the member is
made.

17
Figure 1-1 Time-temperature curve for full process of fire development (Buchanan A.H., 2001).

In the present work of thesis, the performance of concrete structures is investigated by focusing
on concrete compressive behaviour subjected to heating, since it is a major aspect in structural fire
performance.
Within this context, great attention will be devoted to High-Performance Concrete. The
increasing use of this cementitious material in structures exposed to extreme environmental
conditions (tunnels, marine platforms, Liquefied Natural Gas – LNG terminals, containment shells
for industrial facilities), and the need to repair/strengthen existing structures in order to adhere to
more recent codes, in fact, brings in new problems, which go beyond the good knowledge of the
behaviour of ordinary concrete in extreme conditions. High-Performance Concrete, in fact, tends
to be more sensitive to heating, and in particular to the phenomenon of spalling.
Progressive spalling involves gradual fall out of small pieces of concrete from the heated
surface or sudden failure with large release of energy and slough off of large pieces, and it imperils
the load-bearing capacity of the structure, or even worse, may lead to complete failure (examples
of spalling see Figure 1-2). Therefore, spalling should be considered to evaluate the fire
performance of a given R/C structure.
Unfortunately, it is quite difficult to fully understand the phenomenon of spalling since it is a
function of several different factors, such as temperature, concrete type, moisture content, vapour
pressure, cements paste, etc., often leading to unpredictable behaviour.
The best economical method to prevent spalling is the addition of polypropylene fibres to the
concrete mix (0.15-0.3%), since they melt (at about 160-170°C) when exposed to fire, so
increasing the porosity by leaving cavities through which the water vapour can escape. Therefore,
the possibility of spalling is reduced (Buchanan A. H., 2001). This is the reason why, the design of
building with these new materials requires a new approach. Technical solutions should concentrate
on improving the structural design with regards to reinforcement type, external protection, element
dimensions, material microstructure, aggregate type, cement paste, etc.

18
Figure 1-2 damaged columns due to spalling exposed to fire (Pentagon, Washington, USA, 2001).

1.2 Sources and Mechanisms of Fire

Ignition occurs when a gaseous combustible mixture is heated to the temperature that can trigger
the exothermic oxidation reaction of combustion. Generally, the heat required for ignition is
provided by external sources. A few cases that spontaneous combustion occurs inside solid
materials caused by self-heating are studied in a special subject. Multiple possible heat sources can
lead to fire ignition:
• sources with flame (matches, candles, gas heaters, open fires);
• smouldering sources (cigarettes);
• electrical sources (arching, overheating);
• Radiant sources (sunlight, hot items, heaters, fires).
There are some less common sources related to mechanical factors (friction), hot surfaces,
natural phenomena (lighting), transportation (car, bus and train accidents), crime (arson) and socio-
political issues (terrorism, war). With these sources fire could occur in the building and also spread
from the first burning object to another by flame contact or by radiant heat, if the second object is
very close or further away, respectively. The time to ignite a second object depends on the intensity
of radiation from the flame and also the distance between the objects.
The ignition temperature and the amount of heat needed depend on the material properties of
the fuel, the dimensions and shape of the ignited object, and the time of exposure. An efficient
ignition source is the one has competent heat and temperature to cause ignition in the expected
time of exposure. The ignition time of a material depends on its thermo-physical properties, such
as thermal conductivity, specific heat and density, whose outcome is defined thermal inertia. When
exposed to the same heat source, the surface of material with lower thermal inertia will be heated

19
faster than the surface of material with higher material inertia, resulting in more rapid ignition. The
comprehension of the development of a fire requires the knowledge of the phenomena and the
laws governing the transmission of heat, which occurs by conduction, convection and radiation.
Conduction is a significant parameter for the ignition of solid surfaces and fire resistance of
barriers and structural members.
Convection is the heat transfer by the movement of molecules within fluids (e.g., liquids,
gases). It’s also a crucial aspect in flame spreading and upward movement of smoke and hot gases
to the ceiling or through the ventilations during a room fire. Usually, convective heat transfer
calculations consider heat transfer between the surface of a solid and a surrounding fluid that heats
or cools the solid material. The rate of heating or cooling is related to several factors, such as the
velocity of the fluid at the surface.
Radiation is the transfer of energy by electromagnetic waves. It is the main mechanism for heat
transfer from flames to fuel surfaces, from hot smoke to building objects and from a burning
building to the adjacent one.

1.3 Fire Scenario

During the service life of a structural system, fire can be one of the most severe conditions it
undergoes. The main categories of structures faced to fires are:
• residential buildings, office buildings and warehouses;
• offshore structures and petrochemical plants;
• tunnels;
• bridges and viaducts;
• Nuclear power plants.
Generally, the fire load applied to a structure is based on standard temperature-time curves of
the hot gasses, suggested by Code provisions. A number of parameters are taken into account when
the temperature-time curves are defined, and, among all the parameters, the source of combustion
plays a major role.

1.3.1 Buildings

The standard temperature-time curve (also known as Standard Fire curve) represents the typical
development of the temperature of a building for a cellulosic fire, whose fuel source is usually
wood, paper fabric, etc. The curve refers to a single exposure condition of a fully developed fire
without considering cooling. In the first 30 minutes, the temperature rises rapidly to 842°C. The
fire profile keeps increasing up to 1000°C over a period of 120 min, and it will decrease once most
of the combustible material is consumed.
In real cases, the growth phase of fires can be slower or longer, and the maximum temperature
can be higher. However, as subjected to pronounced fluctuations, the fires are rarely sustained.

20
More serious situations can happen with faster fire development, leading to high vapour pressure
in the pores and high thermal gradients.
In EN 1991-1-2 (2004), the expression of Standard Fire is defined as expressed in Equation
[1.1] and plotted in Error! Reference source not found. (initial temperature T0 = 20°C):
T = 345 log (8t + 1) + T0 [1.1]

1200

1000

800
T [°C]

600
Standard Fire (ISO 834)
400
T = 345 log (8t + 1) + T0
200

0
0 30 60 90 120
t [min]
Figure 1- 3 Standard Fire curve according to EN 1991-1-2 (2004).

1.3.2 Offshore Structures and Petrochemical Plants

For offshore structures and petrochemical plants, a new and more severe fire model is required
since the temperature development in hydrocarbon fires is much faster and can reach higher level
than that of cellulosic fires. Thanks to the Mobil Oil Company, in the early 1970s the hydrocarbon
test curve was developed, with a rapid temperature increment up to 900°C in the first 4 min and a
maximum temperature of 1100°C (as shown in Error! Reference source not found., where T0 =
20°C). The curve is adopted by a number of organizations and used as a common method to assess
fire-protecting materials for offshore structures and petrochemical plants.

21
1200

1000

800

T [°C]
600
Hydrocarbon Fire
400
T = 1080 (1 - 0.325e -0.167t - 0.675e -2.5t ) + T0
200

0
0 30 60 90 120
t [min]
Figure 1- 4 Hydrocarbon curve (Mobil Oil Company, 1970s).

1.3.3 Tunnels

Tunnels represent a significant mean connecting people and regions since the latter half of 19th
century. In Europe, Italy (representing 12% of the population and 6% of the surface of the
Continent) owns 30% of the railway tunnels and 37% of the roadway tunnels. Italy, in fact, has
more than 1200 km of railway tunnels and more than 600 km of roadway tunnels, and over 75% of
them were built before 1940, when the fire problems were not taken into consideration in the
structural design. All the data indicate that the assessment of the fire resistance of existing tunnels
is big issue.
1400

1200

1000
T [°C]

800

600 RWS Fire


400

200

0
0 30 60 90 120
t [min]
Figure 1- 5 RWS Fire curve (Rijkswaterstaat and TNO, 1979).
To design the fire resistance of tunnels, several temperature-time curves have been developed.
In the Netherlands, the Ministry of Public Works, namely Rijkswaterstaat (RWS), and the TNO
Centre for Fire Research have established a fire curve for the evaluation of passive protecting
materials in tunnels, based on test results. The fire load modelled by such RWS Fire curve (Error!
Reference source not found.) represents a hydrocarbon fire with high severity, rapidly increasing
up to 1200°C in 10 min and peaking at 1350°C (melting temperature of concrete) after 60 min, and

22
gradually dropping to 1200°C at 120 min. The RWS Fire curve is based on the assumption that in
the worst case, a fire with 300MW fire load was caused by a petrol tanker, burning for 120 min.
Compared to the Hydrocarbon Fire curve, the higher peak of the RWS Fire curve is due to the
relatively enclosed area in which the fire occurs with little heating dissipation into the ambient
atmosphere.
Several accidents of the past century represent very well-known examples of fire in tunnels. In
1996, a train caught fire in the Channel Tunnel linking UK to France (Error! Reference source
not found.a) and the maximum temperature was 1100°C (much lower than the peak of the RWS
curve). The similar conclusion was made according to the fire occurred in the road tunnel of Mont
Blanc (Error! Reference source not found.b) with a maximum temperature of 1000°C.

Figure 1- 6 Fires in (a) Channel Tunnel (1996) and (b) the Mont Blanc Tunnel (1999).

Another fire curve, the RABT curve (Error! Reference source not found.), was developed
considering a less severe fire scenario compared to the RWS curve. The temperature rises rapidly
up to 1200°C (melting point of some aggregates) within 5 min, keeps the peak for 30 min (for car
fires) or 60 min (for train fires), and afterwards drops to 15°C in 110 min. The initial temperature
is considered as 15°C.

23
1400

1200 Car Fires


1000 Train Fires

T [°C]
800

600

400 RABT Fire


200

0
0 30 60 90 120 150 180
t [min]
Figure 1- 7 RABT Fire curves (Germany).

1.3.4 Bridges

On April 29, 2007, the MacArthur Maze, a large freeway interchange bridge near Oakland in the
USA, collapsed because of a severe fire. The fire was triggered by the overturning of a tank truck
carrying 32600 litres of unleaded gasoline, and the peak temperature was 1650°C. The collapse
happened just 20 min after the fire broke out, indicating a very low fire resistance of the bridge.
The weakened steel structure leaded to a significant deflection of the road section and the failure
of the overstressed connections, resulting in the collapse of the bridge.

Figure 1- 8 MacArthur Maze collapsed due to fire (Giuliani et al., 2012).

24
Chapter 2 LITERATURE REVIEW

The aim of this part is to give a brief overview of the works made by other authors in order to
understand the main parameters controlling hot compressive strength of High-Strength Concrete
(HSC) and to plan an experimental campaign aimed at giving answer to problems still open. In the
literature review, attention is paid to concrete mix design, heating cycles and test and their effects
on the compressive strength of HSC. The reference articles are discussed below:
• Bamonte P., Gambarova P. G., Thermal and mechanical properties at high temperature of a very
high-strength durable concrete [J]. Journal of Materials in Civil Engineering, 2009, 22(6): 545-
555.
• Lie T. T., Kodur V. K. R., Thermal and mechanical properties of steel- fibre-reinforced concrete
at elevated temperatures [J]. Canadian Journal of Civil Engineering, 1996, 23(2): 511-517.
• Khaliq W., Kodur V., Thermal and mechanical properties of fibre reinforced high performance
self-consolidating concrete at elevated temperatures [J]. Cement and Concrete Research, 2011,
41(11): 1112-1122.
• Diederichs U., Jumppanen U-M., Schneider U., High temperature properties and spalling
behaviour of high strength concrete. Proceedings of the Fourth Weimar Workshop on High
Performance Concrete: Material Properties and Design, held at Hochschule fur Architektur und
Bauwesen (HAB), Weimar, Germany, October 4th and 5th, 1995, pp. 219-236.

25
• Hammer T. A., “HIGH-STRENGTH CONCRETE PHASE 3, Compressive strength and E-
modulus at elevated temperatures,” SP6 Fire Resistance, Report 6.1, SINTEF Structures and
Concrete, STF70 A95023, February, 1995.
• Felicetti R., Gambarova P. G., Rosati G. P., Corsi F., Giannuzzi G., Residual mechanical
properties of high-strength concretes subjected to high-temperature cycles. Proceedings of 4th
International Symposium on Utilization of High-Strength/High-Performance Concrete, Paris,
France, 1996, pp. 579-588.

Bamonte P., Gambarova P. G., 2009: The aim of the work is to evaluate the thermal diffusivity
and the mechanical decay as a function of temperature of the Ultra High-Strength Concrete (fc =
125 MPa) containing polymeric fibres and quartzitic aggregates.

Concrete mix design, Specimen geometry and Thermal cycles


Most of the tests were performed on small cylinders and small prisms and cubes (minimum size =
36-40mm), larger notched prisms were used for measuring fracture energy. In this study, two types
of polymeric fibres were used, in order to minimize plastic and hydraulic shrinkage and to prevent
spalling during the heating process. A limited number of cylinders and prisms were cast without
polymeric fibres, in order to compare the sensitivity of plain and fibre-reinforced concretes to
spalling, and to check the effectiveness of the fibres against spalling.
The concrete mix design is indicated in Table 2- 1.
The heating and cooling modalities were as follow:
• Heating (cooling) rate: vh = ΔT/Δt = 30°C/h (-15°C/h)
• Reference temperatures: T = 20, 150, 300, 450, 600 and 750°C;
• Rest at the reference temperatures: Δt = 2 h
The hot tests were performed immediately after the rest period at each reference temperature.
As for the residual tests, the specimens were stored in ambient conditions (T =20°C, RH = 70%)
for 1 week prior to testing.

Table 2- 1 Concrete Mix Design (Bamonte and Gambarova , 2009)

Constituents Contents and bulk properties

Cement R52.5, Type 1 c = 635 kg/m3 (29% by solid mass)


Quartzitic aggregates (da = 4 mm) c = 1480 kg/m3 (68% by solid mass)
Polypropylene fibres:

• Macrofibres (L = 20 mm; Ø = 200-250 μm) c = 5.2 kg/m3 (0.24% by solid mass)

• Microfibres (L = 6 mm; Ø=40-50μm) c = 1.1 kg/m3 (0.05% by solid mass)


Water (w/c = 0.31) 200 kg/m3
Acrylic superplasticizer (0.6% c) 3.8 kg/m3 (dry fraction)

26
Solid mass per unit volume 2170 kg/m3
Mass per unit volume 2370kg/m3

Residual Compressive Strength


Three nominally identical cylinders (Ø = 36 mm, h = 110 mm, Error! Reference source not
found.) and six nominally identical cubes (a = 40 mm, Error! Reference source not found.) were
tested for each of the six reference temperatures (including ambient temperature); the cubes
resulted from the breaking off of the prisms tested in three-point bending. All tests were force-
controlled.
The results exhibit a limited scatter, since the standard deviation is well below 10% of the mean
value for the cylinders and below 5% for the cubes. As for the ratio between the cylindrical and
cubic compressive strengths (fc / Rc), the value is close to 0.90 up to 150°C and then starts
decreasing almost linearly, because of the increasing role of friction effects (between the press
platens and the specimen), that are definitely lower in the case of cylindrical specimens.

Figure 2- 1 Plots of the residual compressive strength, on cylinders fc and on cubes Rc

Hot Compressive Strength


The hot compressive strength at different temperatures was measured on cylinders, with the same
heating rate and reference temperatures adopted in the residual tests. The specimens are installed
inside an insulating cylindrical shield made of high performance fibre-reinforced concrete (fc = 80
MPa; see Figure 2- 2 Sketch of the insulating capsule and of the specimen: (a) ready to be introduced into the
furnace; (b) set up during the test; (c) picture of two capsules inside the furnace; (d) picture of a capsule placed between
the platens and the press.
).
After the reference temperature has been reached, there is the usual rest of 2 hours. After being
taken out of the furnace, the whole assembly is installed in 3-4 min between the platens of the
press, and the test is performed in accordance with a force-controlled procedure. The average time
requested by the tests was less than 15 min.

27
Figure 2- 2 Sketch of the insulating capsule and of the specimen: (a) ready to be introduced into the furnace; (b) set
up during the test; (c) picture of two capsules inside the furnace; (d) picture of a capsule placed between the platens
and the press.

The test results are summarized in Error! Reference source not found. in a normalized for,
in order to make comparisons (1) with ordinary concrete (EC2-siliceous concrete); (2) with High-
Performance Concrete (EC2-HPC, Class 3, fc20 ≥ 90MPa); and (3) with the residual strength.

Figure 2- 3 Normalized results of compressive strength compared with the hot curves suggested in EC2 for NSC and
HPC, and residual test.
Between 100 and 250°C the hot strength is lower than the residual strength, since at high
temperature there is vapour pressure inside the pores during the test, while the vapour pressure is
totally released prior to testing after cooling. As a consequence, in hot conditions the interaction
between the mostly isotropic tensile stresses due to vapour pressure and the compressive load
brings in a decrease of the compressive strength.

28
Above 250-300°C, there are no sizeable differences between residual and hot compressive
cylindrical strengths. Furthermore, with reference to EC2 provisions (EC2-EN1992-1-2 2004) and
to temperatures higher than 300°C, both plots are (1) very close to the curves valid for ordinary
siliceous concretes and (2) definitely above the curve corresponding to High-Performance
Concrete.

Lie and Kodur, 1995: The aim of this work is to evaluate the thermal and mechanical properties
of steel fibre-reinforced concrete at elevated temperature. These properties included the thermal
conductivity, specific heat, thermal expansion and mass loss, as well as the strength and
deformation properties of steel fibre-reinforced siliceous- and carbonate-aggregate concretes.

Concrete Mix
For determining the mechanical properties, the following specimens were made from each batch
of concrete:
• 18 cylinders with a diameter of 80 mm and a length of 300 mm;
• 6 cylinders with a diameter of 150 mm and a length of 300 mm;
• 6 prisms with a cross section of 100 mm × 100 mm and a length of 400 mm.
Eighteen cylinders of 80 mm diameter and six prisms of each concrete type were used for
determining the mechanical properties of concrete at elevated temperatures. Two thermocouples
were installed in each of the 80 mm cylinders and in each prism.

Mechanical Properties
The mechanical properties that were measured were the strength and deformation properties of the
concretes at elevated temperatures. The tests were conducted in a special testing machine,
consisting of an electrical furnace, in which the temperatures and the heating rate could be
controlled, and a loading machine capable of performing tests in both displacement- and load-
control.

Table 2- 2 Concrete Mix Design (Lie and Kodur, 1995)


Batch (specimen type)
Property
1 (NRC1) 2 (NRC 2) 3 (NRC3)
3
Cement content (kg/m ) 380 439 439
Fine aggregate (kg/m3) 673 621 621
Coarse aggregate (kg/m3)
19 mm 678 788 788
9.5 mm 438 340 340
Total 1116 1128 1128
Aggregate type Siliceous Carbonate Carbonate
3
Water (kg/m ) 167 161 161
Water-cement ratio 0.44 0.37 0.37

29
Retarding admixture (mL/m3) 745 - -
Superplasticizer (mL/m3) 2500 300 1200
Steel fibre (kg/m3) 42 - 42
28-day compressive strength (MPa) 39.9 32.6 43.2
Compressive strength at test date (MPa) 40.9 37.1 43.3

Figure 2- 4 Stress-strain curves for fibre-reinforced concrete at various temperatures.

The stress-strain tests were carried out at a number of selected temperatures in the range
between room temperature and 750°C. The creep tests were carried out under selected loads that
varied from 15% to 60% of the compressive strength of the concretes at 20°C, in the temperature
range between room temperature and 750°C.
The results of the measurements of the mechanical properties of the concretes showed that the
compressive strength at elevated temperatures of fibre-reinforced concrete is higher than in plain
concrete. The presence of steel fibres increases the ultimate strain and improves the ductility.
However, the effect of aggregate type on the compressive strength is not significant.
Based on the results, stress-strain curves for the concretes at elevated temperatures were
derived as shown in Error! Reference source not found..

Khaliq and Kodur, 2011: The aim of this work is to evaluate the thermal and mechanical properties
of Self-Consolidating Concrete (SCC) and Fibre-Reinforced SCC (FRSCC). As regards the
thermal properties, the specific heat, thermal conductivity, and thermal expansion were measured,
whereas for mechanical properties compressive strength, tensile strength and elastic modulus were
measured in the temperature range T = 20-800°C.

Concrete Mixes and Specimens


Four types of Self-Consolidating Concrete (SCC) were designed for the test: SCC, SCC reinforced

30
with steel (SCC-S), polypropylene (SCC-P) and hybrid fibres (SCC-H).

Thermal properties
The SCC and FRSCC specimens were exposed to high temperature in a furnace connected to Hot
Disk apparatus. Specimen and furnace temperatures were controlled by Hot Disk software. In each
test the furnace temperature was increased to the target temperature and maintained at that level
till the entire test specimen reached equilibrium conditions. At this stage the thermal conductivity
and specific heat were recorded by the data acquisition system. Then the temperature in the furnace
was increased to following target temperature and this procedure was continued till 800°C.

Hot Compressive Strength


In this study, compressive and splitting tensile strength tests were carried out at various
temperatures using unstressed test method. During heating, three Type K thermocouples (TC) were
used to record temperature in the furnace, at the surface and mid-depth of the cylinder. The test
specimens were exposed to target temperatures of 100, 200, 400, 600, and 800°C in an electric
furnace at a heating rate of 2°C/min as proposed by RILEM test procedure. When the target
temperature was attained in the furnace, the cylinder continued to be maintained at this temperature
for 2 hours till desired steady state conditions were reached throughout the specimen.

Table 2- 3 Concrete Mix Design (Khaliq and Kodur, 2011)

Components SCC SCC-S SCC-P SCC-H

Cement type-1, kg/m3 327 327 327 327


3
Fine aggregate, kg/m 735 735 735 735
Coarse aggregate, kg/m3 904 904 904 904
Fly ash (Class C), kg/m3 101 101 101 101
3
Slag St Lawrence (Grade 120), kg/m 76 76 76 76
3
Water, kg/m 143 143 143 143
Water cement ratio, (w/c) 0.44 0.44 0.44 0.44
Water to cementitious ratio (w/cm) 0.28 0.28 0.28 .028
3
Air entraining admixture, kg/m 3 3 3 3
3
High range water reducer and plasticizer, kg/m 81 81 81 81
Slump flow/spread, mm 440 410 420 410
VSI index 0.5 0.5 0.5 0.5
Humidity at casting, % 45 45 45 45
Ambient temperature at casting, °C 23 23 23 23
Concrete mix temperature at casting, °C 20 20 20 20

31
Steel fibres, kg/m3 - 42 - 42
3
Polypropylene fibres, kg/m - - 1 1
Compressive strength, MPa
7 days 46 48 45 46
28 days 61 57 56 57
90 days 72 70 68 72

After achieving steady state conditions in cylinders inside the furnace, the hot cylinders were
taken out from the furnace, were covered by thermal jacket for compressive strength and in steel
bracket frame for splitting tensile strength testing apparatus, as shown in Error! Reference source
not found..

Figure 2- 5 Test set-up for high temperature compressive and tensile strength tests.

Figure 2- 6 Thermal jacket and insulated steel bracket frame.

32
Figure 2- 7 Compressive strength of SCC and FRSCC as function of temperature.

The tests results are shown in Error! Reference source not found.. In all four concretes,
physical and chemical changes that occur at higher temperatures lead to significant reduction in
compressive strength. For SCC, there is a reduction in compressive strength at 100°C initially and
then strength regain at 200°C. The strength loss till 100°C can be attributed to initial moisture loss
in SCC. The increase in strength in 100-200°C temperature range can be attributed to rehydration
and moisture migration. SCC showed higher compressive strength as compared to FRSCC.
The relative strength loss in all concretes follows similar trend to its absolute strength. There
is noticeable strength retention in all SCC and FRSCC even at 800°C and this ranges from 30 to
44%. This level of strength retention is much higher than conventional HSC, which has less than
20% relative compressive strength at 800°C.

Diederichs et al., 1988: In this study, unstressed (strain rate controlled) tests, transient creep tests,
and transient relaxation tests on specimens made of three different HSC concretes were performed.

Concrete Mixes
Among all the three types of concrete, one was a blast furnace slag cement concrete (series Tr),
one was a Portland cement with silica fume concrete (series Si), and one was a Portland cement
with class F fly ash concrete (series Lt). In addition, normal strength concrete with ordinary
Portland cement (series OPC) was also tested for the sake of comparison.

Test procedures and results


Besides compressive strength, specimen shape (cubes and cylinders) and heating rate (2°C/min
and 32°C/min) were also examined. After reaching the desired temperature, the specimens were
then kept for 2 hours to reach a steady state condition, and tested in compression at a constant
strain rate of about 0.05% per minute. Concrete cubes (L = 100 mm) and concrete cylinder (Ø = 80
mm, h = 300 mm) were exposed to temperatures up to 850°C.
Table 2- 4 Concrete Mix Design (Diederichs et al., 1988)

Concrete series Si Lt Tr OPC

Density (kg/m3) 2648 2594 2654 2390

33
Cube strength (MPa)
-28 days 114.4 87.3 91.4 48.0
-90 days 100.8 106.9 111.9 36.0
Cylinder strength (MPa)
-90 days 106.6 91.8 84.5 32.9

Error! Reference source not found. shows the variations of compressive strengths and
modulus of elasticity of all concretes with respect to temperature.

Figure 2- 9 Stress-strain relationships of (a) HSC made with fly ash; (b) NSC made with ordinary Portland
cement.

shows typical stress-strain relationships for HSC and NSC.

Figure 2- 8(a) Variation of compressive strength with temperature; (b) variation of elastic modulus with temperature.

34
Figure 2- 9 Stress-strain relationships of (a) HSC made with fly ash; (b) NSC made with ordinary Portland cement.

The following observations were made:


• HSC specimens failed in a more brittle manner than Normal-Strength Concrete specimen.
• Destructive spalling did not occur in any of the specimens heated at rate of 2°C/min.
• Slight spalling occurred to some cylinders used in the transient creep and transient relaxation
tests (preloaded prior to heating), which were heated at a high rate of 32°C/min.
• Spalling occurred in all pre-loaded cubes heated at the rate of 32°C/min.
• The so-called “dry-hardening” which causes the increase of strength in Normal-Strength
Concrete between 150 to 350°C was not observed for HSC, instead the strength of HSC
decreased in this temperature range.
• Besides heating rate and paste density, the dimensions and shapes of specimens are also
important with respect to explosive spalling.

Hammer, 1995: Unstressed (stress rate controlled) tests were performed in order to study the
effects of the following variables on the fire performance of HSC:
• Concrete Compressive strength: varied between 69 to 118 MPa. The variation in compressive
strength was achieved by changing the water/cement ratio from 0.27 to 0.50.
• Aggregate type: lightweight aggregate and normal weight crushed gravels.
• Temperature: varied from 20, 100, 200, 300, 450, and 600 °C.

Specimen Geometry and Test Procedure


Five different concrete mixes were used as shown in Error! Reference source not found..

Table 2- 5 Concrete Mix Design (Hammer, 1995)

Concrete Mixtures w/c ratio Silica Fume

ND65 0.50 5% (by mass of cement)


ND95 0.36 5%
ND95-0 0.36 0%

35
ND115 0.27 5%
LWA75 0.36 5%

ND: Normal density concrete


LWA: Lightweight aggregate concrete
95-0: Concrete compressive strength is 95 MPa and without silica fume

The specimens were 100 mm × 310 mm cylinders. Two specimens were tested at each
temperature level, one at 90 day of age and one at 150 day of age. The cylinders were stripped, the
ends were cut, then cured at room temperature (20°C) two days before testing.
Five cylinders, to be tested on each at the target temperatures of 100, 200, 300, 450 and 600°C,
were weighted and heated together at a rate of 2°C/min. Each cylinder was removed from the oven
at the target temperature and weighted again for mass loss (moisture evaporation). The cylinder
was reheated using a steel tube covered by a heating element and loaded. The time between
removal from the oven and to the start of testing was typically 8 minutes. The loading plates were
also heated to minimize heat loss. The loading rate was 0.80 MPa/s.

Test Results
The test results, in terms of variations of compressive strengths and modulus elasticity with respect
to temperatures are shown in Error! Reference source not found. to Error! Reference source
not found.. Also plotted in Error! Reference source not found. are the relationships between the
measured mass loss and the strength loss.

Figure 2- 10(a) Compressive strength of 90-day concrete vs. temperature; (b) Elastic modulus of 90-day concrete vs.
temperature

36
Figure 2- 11(a) Compressive strength of 150-day concrete vs. temperature; (b) Elastic modulus of 150-day concrete
vs. temperature.

The following conclusions were drawn:


• The difference in terms of strength loss with increasing temperature for the different concretes
examined in this study was not relevant, except for ND95-0 (no silica fume concrete), which
showed a gain in strength between 200 to 300°C.
• A typical “breakpoint” in the strength-temperature curves was observed at 300°C. The study
reported that at this temperature an explosive failure followed by the release of steam was
observed. No such explosive failure or steam release was observed at the other temperatures.
The author speculated that the reason for the reduced strength even at low temperatures
(between 100 to 300°C), is the high internal pressure due to the reduced ability of moisture
to escape from HSC.
• An increase in compressive strength was observed for all concretes, except for ND95-0,
between 300-450°C. For mix ND95-0, the strength increase occurred earlier, between 200-
300°C.
• Modulus of elasticity of all concretes decreased at a higher rate at temperature above 300°C.
• Concrete without silica fume (ND95-0) show slightly better performance in terms of lower
strength loss.
• The replacement of normal weight coarse aggregate with light-weight aggregate does not
seem to affect the temperature dependent strength loss.

37
Figure 2- 12(a) Mass loss vs. Strength loss for 90-day concretes; (b) Mass loss vs. Strength loss for 150-day
concretes.

Felicetti et al., 1996: Unstressed residual strength (strain rate controlled) tests on silica fume based
HSC specimens with specified strengths of 72 and 95 MPa were performed.

Specimen Geometry and Test Procedure


Both concretes used siliceous aggregates, consisting mostly of crushed flint particles, with
maximum size of 25 mm. The specimens included cylinders of two different sizes: 100 mm × 300
mm and 100 mm × 150 mm; and deep beams with dimensions of 80 mm × 275 mm × 500 mm. The
100 mm × 300 mm cylinders (37 specimens) were tested in uniaxial compression. The 100 mm ×
150 mm cylinders (20 specimens) were notched at mid-height and tested in direct tension. For
uniaxial compression tests, batches of 2 to 4 cylinders were exposed to temperatures of 20, 105,
250, 400 and 500°C.
All specimens were cured under water for one week and stored in air for three weeks at 20°C
and 92% R.H. and one month at 20°C and 65% R.H. Specimens heated to 105°C were kept at this
temperature for 7 days in an oven, and then allowed to cool to room temperature in the closed
oven. For specimens exposed to higher temperatures, a heating rate of 12°C/min was used to heat
the specimens up to the target temperatures, which were maintained for 12 hours. A cooling rate
of 12°C/min was used to cool the specimens to room temperature.

Experimental Results
The experimental results, in terms of variations of residual compressive strengths and modulus
elasticity with respect to temperatures are shown in Error! Reference source not found.. Also
plotted in Error! Reference source not found. are the stress-strain relationships for HSC.

38
Figure 2- 13(a) Residual concrete strength vs. temperature; (b) Residual elastic modulus vs. temperature.

The study revealed similar trends of reduction in compressive strength and modulus of
elasticity with increasing temperatures as was observed in other experimental campaigns. An
exposure temperature of 250°C appears to be the level which marks the higher rate of strength and
modulus reduction. At 400 to 500°C, most of the flint aggregates appeared cracked or split, with a
different colour compared with aggregates not exposed to heat.

Figure 2- 14 Stress-strain curves for concrete after heating: (a) 95 MPa; (b) 72 MPa.

39
Chapter 3 CONCRETE BEHAVIOUR AT HIGH
TEMPERATURE

3.1 Thermal and Physical Properties of Concrete

In order to investigate the temperature distribution inside concrete element thus to determine the
thermal field, the thermal properties of concrete material should be known.
With reference to concrete, the aggregate type and the initial moisture content are important
factors on which the thermal parameters depend. The type of coarse aggregate used in concrete
mix is one of the main factors affecting the change in concrete properties with the temperature. In
the following, 3 aggregate types will be considered: carbonate, siliceous and lightweight.
Carbonate aggregates involve limestone and dolomite, siliceous aggregates involve silica, granite
and sandstone, while lightweight aggregates are usually produced by heating shale, slate of clay.

3.1.1 Porosity and Density

The loss of evaporable and absorbed water, and the dehydration create additional pore space to
effectively increase the porosity as the temperature increases. At 850°C, porosity is about 40% by
volume greater than at 105°C, as noted by Harmathy (1970, see Figure 3- 1 Portland cement paste
porosity as a function of temperature (Harmathy, 1970).

According to Tsimbrovska (1997), the increase in porosity is caused by two main mechanisms:
• 20°C ≤ T ≤ 250-300°C: loss of adsorbed and interlayer water;
• T > 250-300°C: micro-cracking (due to dehydration and thermal incompatibility between
aggregate and cement paste).

40
Figure 3- 1 Portland cement paste porosity as a function of temperature (Harmathy, 1970).

As a result of the pore volume increase, the bulk mass density of the cement paste decreases
with temperature. Changes in mass density result also from thermal expansion and drying
shrinkage, diffusion of water or released gases, and dehydration, melting, or sintering.
Concrete density depends on the density of the aggregates and on its moisture content in the
temperature range from 20°C to 150°C. Figure 3- 2Error! Reference source not found.
presents the effect of temperature on the density of concrete as per EC 2(T = 20-1000°C).
ρ
2600
2500
Density kg/m3

2400
2300
2200
2100
2000
0 250 500 750 1000 1250
Temperature

Figure 3- 2 Density of concrete (EC 2).

3.1.2 Permeability

Permeability is one of the most important parameters in understanding concrete behaviour at high
temperature, in particular as regards explosive spalling. The intrinsic permeability K is a material
property (not related to any specific fluid), from which the conventional permeability K’ can be
evaluated as shown in Equation [2.1]:
kr
K' = -K [2.1]
μf

Where,

41
kr is the relative permeability of the fluid, depending on the fluid and on the saturation degree;
μf is dynamic viscosity of the fluid.
Kalifa and Tsimbrovska (1998) studied the influence of the temperature on the intrinsic
permeability. Figure 3- 3(a) shows that at 105°C, the permeability of High-Performance Concrete
is 10 times smaller than that of Normal-Strength Concrete; between 105°C and 400°C the
permeability of High-Performance Concrete increases much faster than that of Normal-Strength
Concrete. Above 300°C, micro-cracks play the major role in increasing the permeability.
Mindeguia et al. (2013) studied the influence of temperature and fibres on concrete intrinsic
permeability (see Figure 3- 3(b)). The relative intrinsic permeability (ratio between permeability
at high temperature and in virgin conditions) increases significantly with the temperature (from 2
to 4 orders of magnitude). For concretes without polypropylene fibres, the permeability is little
affected by heating up to 250°C while at higher temperatures, it strongly increases. For concrete
containing polypropylene fibres, the permeability increases significantly in the range 120-250°C,
since the temperatures are around the melting point (160-170°C) of polypropylene fibres.

Figure 3- 3(a) Concrete intrinsic permeability (Kalifa and Tsimbrovska, 1998); (b) concrete relative intrinsic
permeability (Mindeguia et al., 2013).
The effect of polypropylene fibres becomes particularly evident between 120-250°C due to
two main reasons: (a) melting of fibres increases the pore volume, and (b) fibres behave like micro-
defects causing a concentration of the thermal stresses, and favouring micro-cracking, to the
advantage of pore connectivity (Pistol et al., 2011). Moreover, according to several authors (Kalifa
et al., 2001; Khoury, 2008; Pistol et al., 2011), polypropylene fibres first expand while melting (at
160-170°C) and then shrink.
The permeability increase is favoured, also, by the thermal incompatibility between cement
paste and aggregates. In fact, while the cement paste shrinks due to drying, the aggregates dilate
due to thermal expansion. This mismatch brings in tensile stresses in the matrix and, hence, micro-
cracking, which can be seen as percolation paths for fluids, explaining the high measured values
of intrinsic permeability (see
Figure 3- 4 Scanning Electron Microscopy images of samples after heating at 400°C (Mindeguia et al., 2013 )).
Furthermore, the mechanical loads influence the evolution of intrinsic permeability due to
load-induced macro-cracking (Gérard et al., 1996; Wang et al., 1997; Torrenti et al., 1999), as also
shown in an experimental study carried out by Picandet et al., (2001). Compared to undamaged

42
specimens, a uniaxial compressive load at 90% of the ultimate strength can increase concrete
permeability by about one order of magnitude. This increase is related to the maximum applied
strain and is caused by the formation of a connected network of micro-cracks. It can be stated that,
during heating, porosity changes from a system of more or less closed and isolated pores to a rather
open interconnected network.

Figure 3- 4 Scanning Electron Microscopy images of samples after heating at 400°C (Mindeguia et al., 2013).

3.1.3 Thermal Expansion

The resultant coefficient of thermal expansion for concrete is dominated by the aggregate type.
The coefficient of thermal expansion of concrete depends also on the age and type of curing, being
influenced by the moisture content of the cement paste, with minimum values in dry and saturated
conditions. The thermal dilation increases with the temperature due to aggregate expansion
dominating over cement paste shrinking.
Lightweight-aggregate concretes (e.g. pumice and expanded shale aggregates) may shrink at
temperatures higher than 300°C. Tests on sealed concrete indicate that under thermal cycling (up
to 150°C): (a) permanent expansive strains occur after the first cycle, and (b) the permanent
expansion increases with increasing number of cycles, but at a decreasing rate. Also the thermal
coefficient decreases as the number of cycles increases; furthermore, specimens allowed to dry
after the initial heating show less expansion during any subsequent thermal cycles. Results indicate
an almost monotonic increase in the thermal expansion coefficient for limestone concrete, until
decarbonation (CaCO3 → CaO + CO2) leads to a reduction of the coefficient.

The coefficient of thermal expansion represents the volume change of a material due to varying
temperatures, and is expressed as a variation in length per degree of temperature change. The
coefficient is important as a measure of the structural displacements and thermal stresses resulting
from a temperature change that can lead to cracking and/or spalling.
Concrete thermal expansion is a complex phenomenon because of the interaction of its two main
components, cement paste and aggregates, which have their own thermal behaviours. The intrinsic
differences in the thermal expansion between cement paste and aggregates cause tensile stresses
to arise in the former, followed by micro-cracking. Any observation of the thermal expansion in

43
concrete is rather difficult due to various extraneous effects that accompany the temperature
change (additional volume changes caused by variations in moisture content, chemical reactions
leading to dehydration, creep and micro-cracking resulting from non-uniform thermal stresses).
Figure 3- 5 represents the thermal strain of Normal Strength concrete as per Eurocode 2 and Figure
3- 6 represents examples of linear thermal expansion of various rocks with temperature.

0.016
Thermal dilation as per EC2
0.014
Siliceous
0.012 aggregate
Thermal Strain ϵth

0.01 s
0.008
0.006 Calcareou
s
0.004
aggregate
0.002 s
0
-300 200 700 1200
Temperature

Figure 3- 5 Thermal strain of Concrete (EC 2)

coefficient of thermal expansion α


0.000018
0.000016
Thermal Expansion

0.000014
0.000012
0.00001
0.000008
0.000006
0.000004
0.000002
0
0 200 400 600 800 1000 1200 1400
Temperature
Figure 3- 6 Coefficient of thermal Expansion of concrete (EC 2).

3.1.4 Thermal Conductivity

Thermal conductivity is the heat flux transmitted through a unit area of a material under a unit
temperature gradient (i.e. ability to conduct heat). At normal temperatures the thermal conductivity
of concrete depends primarily on the thermal conductivity of the aggregates and on the moisture
content at the time of heating (the higher the aggregate-to-cement ratio and the lower the water-
to-cement ratio, the higher the concrete conductivity).
Other important factors influencing concrete thermal conductivity are the hardened cement

44
paste, the pore volume and distribution, and the water content, while age does not appear to have
effects at high temperatures. Relatively high values for thermal conductivity can be shown at low
temperatures, while at higher temperatures (T ≥ 100°C) thermal conductivity decreases. Figure 3-
7 presents the effect of temperature on thermal conductivity of concrete (EC 2).

2.5 Conductivity λ
2 lower bound

Conductivity W/m K
1.5 upper bound

0.5

0
0 200 400 Temperature
600 800 1000 1200

Figure 3- 7 Thermal conductivity of concrete (EC 2).

The major factors influencing concrete thermal conductivity are moisture content, aggregate
type, cement paste, and pore volume and distribution. Concrete conductivity varies linearly with
moisture content. As aggregate conductivity increases, concrete thermal conductivity increases as
well. Cement paste has a lower conductivity than the aggregates, thus concretes with higher cement
paste content tend to have a lower conductivity than lean mixes.

3.1.5 Specific Heat

The heat capacity of a material is the amount of heat per unit mass required to change the
temperature of the material by one degree, and can be expressed through Equation [2.2].
∂H
Cp = ( ) [2.2]
∂Tp
Where H is enthalpy, T is the temperature, and p is the pressure.
The specific heat of ordinary concrete at room temperature ranges from 0.50 to 1.13 kJ/(kg°C),
while the specific heat of cement paste ranges from 0.63 to 1.72 kJ/(kg°C). At room temperature,
aggregate type, mix proportions, and age do not have a great effect on the specific heat of concrete;
however, as the moisture content increases, the specific heat capacity increases at low temperature.
The specific heat of aggregates can be calculated from their mineralogical content and the specific
heat of concrete can be determined based on its relative proportions.
Figure 3- 8 represents the effects of temperature on experimentally determined specific heats
of concrete with different moisture content. At elevated temperature, specific heat is sensitive to
the various transformations that take place in concrete. This includes the vaporization of free water
at about 100-250°C, the dissociation of Ca(OH)2 at about 450-500°C and the α-β quartz
transformation in some aggregates. Heating of initially-saturated concrete causes a rapid but

45
temporary rise in the specific heat at about 90°C due to the rapid release of latent heat of
vaporization (Bazant and Kaplan, 1996).
C
2500
0% moisture
2000

Speicific Heat J/ kgK


1.5% moisture
3% moisture
1500

1000

500

0
0 250 500 750 1000 1250
Temperature
Figure 3- 8 Specific heat of concrete (EC 2)

3.1.6 Thermal Diffusivity

Thermal diffusivity is a measure of the rate at which heat will diffuse through a material in all
directions due to a temperature variation. It is the ratio between the heat transmitted and the heat
stored by the unit mass of the material, therefore it indicates the ability of a material in transmitting
heat due to temperature gradients (the smaller the thermal diffusivity, the higher the insulation
ability). Thermal diffusivity is defined according to Equation [2.3].
λ
D= [2.3]
ρc
Where,
D is the thermal diffusivity [m2/s] (generally ranging between 0.3 and 0.8 mm2/s);
λ is the thermal conductivity [W/m°C];
c is the specific heat [J/KgK];
ρ is the density [Kg/m3].
Thermal diffusivity is important, for instance, in nuclear power plant structures such as pre-
stressed concrete pressure vessels. Concrete thermal diffusivity can be determined by the thermal
properties of its constituents. At normal temperatures the diffusivity of concrete is mainly governed
by the diffusivity of the aggregate. Aggregates with increasing values of thermal diffusivity include
basalt, rhyolite, granite, limestone, dolerite and quartzite. Factors that affect thermal conductivity
generally have the same influence on thermal diffusivity.
Figure 3- 9 represents thermal diffusivity for normal-weight concrete as per EC 2. As noted in
the figure, at about 600°C the thermal diffusivity of all concretes investigated becomes
approximately 0.3 mm2/s.

46
Diffusivity
0.7
0.6

Diffusivity mm2sec
0.5
0.4
0.3
0.2
0.1
0
0 200 400 600 800 1000 1200
Temperature
Figure 3- 9 Thermal diffusivity of normal concrete (EC 2)

3.2 Mechanical Properties

The chemical reactions involving concrete constituents during heating together with the mutual
interaction between aggregates and cement paste affect concrete mechanical properties, inducing
a remarkable deterioration of its mechanical response. Moreover, durability and serviceability are
remarkably influenced by high temperature due to the physical deterioration processes which may
lead to severe cracking. The mechanical decay of concrete during a fire may be attributed to three
main material factors:
• Physico-chemical changes in the cement paste;
• Physico-chemical changes in the aggregates;
• Thermal incompatibility between aggregates and cement paste.
Cement-based materials, however, lose dramatically their bearing capacity only above 550-
600°C. At lower temperatures (i.e. those reached, for instance, in the core of R/C members during
fire), the decay of the mechanical properties (compressive strength and elastic modulus) can be
reduced by choosing an appropriate mix design, whereby thermally stable aggregates of low
thermal expansion are employed, and cement blends are selected (which produces a low CaO/SiO2
ratio).
Concrete deformation is usually described by assuming that the total strain ε consists of four
components as described in Equation [2.4] (Buchanan, 2001; En 1992-1-2, 2004).
ε = εth (T) + εσ (σ, T) + εcr (σ, T, t)+ εtr (σ, T) [2.4]
where
εth (T) is the thermal strain, which is solely a function of the temperature;
εσ (σ, T) is the instantaneous stress-related strain, which is a function of both the applied stress σ
and the temperature;
εcr (σ, T, t) is the creep strain, which is a function of the stress, the temperature and the time;

47
εtr (σ, T) is the transient strain, which is a function of both the applied stress and the temperature.

3.2.1 Compressive Strength

Concrete compressive strength is at the basis of the load-carrying capacity of a structure. The hot
behaviour of concrete in compression is generally well described by means of a temperature-
dependent stress-strain diagram (see Figure 3- 10), which depends on three parameters: (a)
compressive strength fcT; (b) strain at the peak stress εc1T and (c) elastic modulus EcT, considered
dependent on the temperature.

1.2
σ/fc,20 Vs ϵ
1

0.8 T 400
Stress MPa

0.6 T 300
T 200
0.4
T 100
0.2
T 20
0
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035
Strain
Figure 3- 10 Temperature effect on stress-strain curves of concrete.

Hot Behaviour
Above 100°C the physically combined water is released from concrete. During this process, the
strength does not change, but above 100°C the elastic modulus decreases by about 10-20%. Above
300°C the silicate hydrates decompose and above 500°C the portlandite is dehydrated, so the
compressive strength is limited reduced below 450-500°C and rapidly above 500°C. Temperatures
higher than 500°C can reduce concrete compressive strength to only a small fraction of its original
value, especially in highly siliceous concrete (Felicetti and Gambarova, 1998), and such concretes
are unlikely to possess any useful structural strength as they may contain a considerable amount
of calcium oxide (CaO), which expands under moisture attack.
At temperatures above 600°C some aggregates begin to convert or to decompose. In any case,
if spalling does not occur, the external layers are no longer able to carry any load, but still prevent
the inner core from being overheated, something positive for the fire resistance of a structural
member.

Residual Behaviour
The effect of temperature on the residual compressive strength of Normal-Strength Concrete is
shown in
Figure 3- 11 Residual compressive strength as a function of temperature according to (a) EN1994-

48
1-2 (2004) and (b) ACI 216-1.07 (2007).

According to EN1994-1-2 (2004) and ACI 216-1.07 (2007), respectively.

Figure 3- 11 Residual compressive strength as a function of temperature according to (a) EN1994-1-2 (2004)
and (b) ACI 216-1.07 (2007).

Recent research has extended the testing of compressive strength from Normal-Strength
Concrete to High-Performance Concrete (HPC, fc = 60-130 MPa) and even Ultra-High
Performance Concrete (UHPC containing steel fibres, fc = 130-200 MPa). The results plotted in
Error! Reference source not found. shows that the residual strength of the HPC declines less
with the temperature than that of the UHPC, but that the reverse is true for the residual strength
measured after cooling (possibly because of the steel fibres).

Figure 3- 12 Residual compressive strength of two HPC and two UHPC: (a) actual value; (b) values normalized with
respect to the virgin strength (Khoury, 2000).

The relatively poorer hot performance of the UHPC could be due to its very dense structure,
from which moisture escapes less readily. This has two effects: (a) a physical effect due to reduced
Van der Waals forces as water expands upon heating, and (b) a chemical effect whereby detrimental
transformations can take place under hydrothermal conditions.

49
3.2.2 Tensile Strength

The tensile strength of concrete is important because it determines the ability of concrete to face
cracking. Direct measurement of concrete tensile strength is only seldom made because of
difficulties in gripping the specimen faces and in maintaining a uniform tensile stress across the
section. An indication of concrete tensile strength can be obtained by split and/or flexural tests.
The split test is an indirect method for evaluating tensile strength; the experimental procedure
consists in a horizontal concrete cylinder loaded in compression through bearing strips placed
along two axial lines that are diametrically opposite. Tensile strength reduction with the
temperature in hot conditions takes place at rather low temperature (T ≥ 100°C), as shown in Figure
3- 11 (EN1992-1-2, 2004).

Figure 3- 13 Decay of the normalized tensile strength in hot conditions (EN1992-1-2, 2004).

Figure 3- 14 Tensile strength of HSC and NSC in residual conditions (Noumowe et al., 1996).

Figure 3- 14 shows that the residual tensile strengths for both Normal-Strength Concrete and
High-Strength Concrete decrease similarly and almost linearly with increasing temperature.
Tensile strength of the High-Strength Concrete remained 10-15% higher than that of the Normal-
Strength Concrete over the entire temperature range. As expected, at any temperature the tensile
strength by splitting (and the same has been proved for the flexural tests) is larger than that in
direct tension, thanks to the beneficial effect of stress redistribution in the specimens characterized
by plane stresses.

50
Figure 3- 14 Tensile strength of HSC and NSC in residual conditions (Noumowe et al., 1996).

High-Performance Concrete containing silica fume (fc = 60-130 MPa) is more sensitive to high
temperature (T = 100-800°C) than ordinary concrete, owing to its denser cementitious matrix, with
lower diffusivity, and to the more important role of the aggregates, whose thermal decay is much
more harmful.
From the limited results on the tensile strength available in the literatures, aggregate type and
mix proportions have a significant effect on the tensile strength, whose decrease in calcareous-
aggregate concrete is twice as high as that of siliceous-aggregate concrete at 500°C; lean mixes
(with a rather low cement content) are less heat-sensitive than rich mixes (with a rather high cement
content). Finally the residual tensile strength is somewhat lower than the tensile strength measured
at elevated temperature.

3.2.3 Poisson Ratio

The knowledge of the Poisson ratio is needed for structural analyses of flat slabs, arch dams,
tunnels, tanks, and other statically indeterminate members. At normal ambient conditions, the
Poisson ratio of concrete ranges from 0.11 to 0.32, but is generally comprised between 0.15 and
0.20. Under uniaxial compression loading, the beginning of matrix micro-cracking corresponds
to an apparent increase in the Poisson ratio value. At ambient temperature there is no consistent
relationship between the Poisson ratio and concrete characteristics (water-cement ratio, curing
age, and aggregate size and type), however it is generally lower in High-Performance Concrete
and higher in saturated concrete and in dynamically-loaded concrete.
At 20˚C the Poisson ratio is constant before the load reaches about 70% of the ultimate load,
while as the temperature increases the deviation of Poisson ratio increases Figure 3- 15(a).
According to Figure 3- 15(b), the Poisson ratio decreases with temperature, from about 0.2 at
room temperature to about 0.1 at 400°C. This drop is attributed to a change in state due to
heating that resulted in water desorption (e.g. weakening of the microstructure caused by
breakage of bonds due to heating and by micro-cracking). When the specimen is permitted to

51
cool after a given change in state, the variation of the Poisson ratio with temperature is slight,
and nil when the evaporable water is removed.

Figure 3- 15(a) Normalized stress as a function of Poisson ratio (Ehm, 1985); (b) Poisson ratio as a function of
temperature (Marechal, 1972).

Figure 3- 16 shows the Poisson ratio as a function of the temperature for different aggregate
types and two nominal concrete compressive strengths. The values generally varied from 0.11 to
0.25. Even though at higher temperature the results were unstable and there was no general trend,
the results did show lower Poisson ratio values at room temperature for higher strength concretes.
Some results by Bamonte and Gambarova (2010) clearly indicate that, the residual Poisson
ratio increases up to 0.30-0.35 for T = 250-350°C in a siliceous High-Strength Concrete (fc = 125
MPa), and then tends to decrease. At 600°C, the value of the Poisson ratio is close to 0.15, which
is value of the virgin material.

Figure 3- 16 Effect of aggregate type and concrete strength on Poisson ratio (Cruz, 1966).

3.2.4 Elastic Modulus

Concrete elastic modulus is a measure of its stiffness or resistance to deformation, which is used
extensively in the analyses of reinforced concrete structures to determine stresses and
displacements developed in R/C members or structures. Several potential factors affecting the
elastic modulus due to heating have been investigated (e.g., mix proportions, aggregate type,
cement type, concrete strength, sealed or unsealed conditions, presence of sustained stress during

52
heating, and duration of thermal exposure). Besides these parameters, however, only the aggregate
type and – to a lesser extent – the loading conditions during heating have a sizeable effect on the
elastic modulus. Elastic modulus is more temperature-sensitive than the compressive strength both
at high temperature and after cooling, as concrete stiffness is directly affected by micro-cracking.
The significant influence of aggregate type on the elastic modulus is presented in Error!
Reference source not found.. For siliceous aggregates, such as basaltic and quartzitic aggregates,
the normalized elastic modulus EcT/ Ec20 is 0.4-0.5 at 500°C, while the reduction of elastic modulus
in light-weight aggregates is close to, or even less than that of calcareous aggregates (at 500 °C,
EcT/ Ec20 is up to 0.70-0.75 for expanded clay).

Elastic Modulus
1.2
1
0.8
ET/E20

0.6
0.4
0.2
0
0 200 400 600 800 1000 1200
Temperature
Figure 3- 17 Normalized concrete elastic modulus for (EC 2)

Figure 3- 18 Normalized plot of fracture energy Gf as a function of temperature (Heinfling, 1998).

3.2.5 Fracture Energy

Fracture energy Gf is defined as the total amount of energy dissipated over a unit crack area
(Heinfling, 1998) and is presented in Figure 2-23 as a function of temperature. According to Zhang
and Bicanic (2001), the decrease at 105°C is mainly caused by the rise of pore pressure due to
water vaporization. So, the higher the moisture content, the higher the decrease of fracture energy.

53
For higher values of the temperature, the effect of pore pressure becomes negligible and fracture
energy increases. Up to 300°C further cement hydration helps this recovery. Above 300°C micro-
cracks, dehydration and decomposition cause a reduction of fracture energy. In any case, the
experimental results show a sizeable scattering also because of the lack of uniformity in the
definition of fracture energy among the different authors. Generally, fracture energy is much less
temperature-sensitive than the other mechanical parameters (such as tensile and compressive
strength, and elastic modulus), as shown in Error! Reference source not found. (see also Felicetti
and Gambarova, 1998).

3.3 Spalling at High Temperature

Spalling is the violent or non-violent breaking off of layers or pieces of concrete from the surface
of a structural member when exposed to high and more or less rapidly increasing temperatures, as
experienced in fires with heating rates typically of 10-30°C/min.
Spalling can be grouped into four categories: (a) aggregate spalling; (b) explosive spalling;
(c) surface spalling; (d) corner/sloughing-off spalling. The first three categories occur during the
first 20-30 min of the fire and are influenced by the heating rate, while the last one occurs after
30-60 min and is influenced by the maximum temperature (Khoury, 2000).

Figure 3- 19 Explosive spalling: empirical envelope for Normal-Strength Concrete, showing the influence of
moisture content and applied stress (Khoury, 2000).

The material is breaking-off in different ways, as stated by the Institution of Structural


Engineers-ISE (1975):
• Explosive spalling: is a violent form of spalling, which occurs at an early stage of the heating
process and may result in extensive damage or complete destruction of the concrete member;
• Local spalling: consists of sometimes violent surface spalling, aggregate splitting or corner
separation. Aggregate spalling is caused by aggregate failure close to the surface and involves

54
small pieces expelled from the surface. This type of spalling does not adversely affect the
structural performance, as only superficial damage is produced;
• Sloughing off: is a progressive form of breakdown, involving the partial separation of
material from the concrete member. Corner spalling occurs later, often in the decay stages of
the fire. It is characterized by larger corner pieces breaking-off of the concrete due to tensile
cracks developing close to the corners and edges. Due to the late occurrence of this type of
spalling, concrete and reinforcing bars are already weakened. Thus, corner spalling may not
influence the performance of structures significantly when subjected to fire.
In current design codes such as Eurocode 2, there is not much information about spalling. It is
stated that no checks for spalling are necessary if the moisture content of the member is less than
3%, as shown is Figure 3- 19, and that the tabulated data can be used to prescribe generic fire
ratings for concrete members.
For concrete grades in the range C80/95-C90/105, spalling can occur in any situation when
there is a direct exposure to the fire and four methods are recommended to be used, regarding
reinforcement mesh, concrete type, protective layers and inclusion in the concrete mix of more
than 2 kg/m3 of monofilament polypropylene fibres.

55
Chapter 4 THE THERMAL PROBLEM

The physical and mechanical properties of concrete at high temperature are dependent on the
thermal field inside the material during the heating procedure, as described in the previous chapter.
For a given point, the local thermal degradation is linked to the maximum temperature reached.
However, most experimental measurements of temperatures are not able to provide continuous
thermal field inside a specimen or a structural member, which makes the numerical simulation
essential for the investigation and the design of structural fire performance.
In numerical simulations of heat transfer, the thermal properties of materials, boundary
conditions and fire scenarios are taken into account to evaluate the thermal field. This chapter will
briefly describe the theoretical basis of heat transfer in order to set up a finite element code for
simulations.

4.1 Mechanisms of Heat Transport

Heat transfer is defined as a process of energy exchange, which is not static. Heat transfer occurs
between physical systems with different temperatures by dissipating heat. The fundamental
mechanisms of heat transfer are: conduction, convection and radiation.

4.1.1 Conduction

Conduction is the transfer of internal energy by microscopic diffusion and collisions of particles
within a body, because of the thermal gradient. The kinetic energy among the molecules is
transmitted without macroscopic mass transport.
For a homogeneous and isotropic plate (see Figure 4-20) with two parallel surfaces Γ1 and Γ2
at a distance of L, the corresponding temperatures are T1 and T2 (T1 > T2). During a time instance,
assuming that the temperatures of the surfaces are constant, the isothermal planes within this plate
are parallel to the surfaces and the heat flux is perpendicular to the isothermal planes.

56
Figure 4-20 Heat transfer within a continuous plate.

Heat transfer always takes places from a region of high temperature to another region of lower
temperature, which means for the plate at issue, that the heat moves from Γ1 to Γ2. The total amount
of heat Q moving though the area A in a time instance Δt is calculated using Equation [4.1] (Fourier
Law):
λA
Q= ( T1 - T2 ) ∆t [4.1]
L
According to the Fourier Law, thermal conductivity λ of a material is defined in Equation [4.2],
expressed in terms of specific heat flux:
Q λ
q̇ = = ( T1 - T2 ) [4.2]
A ∆t L

4.1.2 Convection

Convection is the heat transfer mechanism occurring between two bodies in relative motion to
each other, and is usually describing the transmission through the macroscopic movement of fluids.
Convection mainly characterizes the heat transfer between a solid body and its surrounding fluids.
Considering a solid body with external surface temperature of T1 and its surrounding fluids at
a temperature of T2 (T1 > T2), the specific heat flux can be evaluated by the Equation [4.3] (Newton
Law):
q̇ = αc ( T1 - T2 ) [4.3]
The thermal convection coefficient αc depends on both the physical properties of the fluid and
the physical situation in which convection occurs, indicating that the heat loss rate of this solid
body is proportional to the temperature difference between its surrounding fluids and itself.

57
4.1.3 Radiation

Radiation is a mechanism of heat transfer relating to electromagnetic radiation generated through


the thermal motion of the charged particles. For a body with a temperature higher than absolute
zero, the collisions of molecules will lead to the change of their kinetic energy, resulting in the
emission of thermal radiation.
The Stefan-Boltzmann Law provides the algorithm of thermal flux radiated per unit surface
area of a black body per unit time, proportional to the fourth power of the body’s thermodynamic
temperature T, as expressed in Equation [4.4]:
q̇ = σ T 4 [4.4]
The Stefan-Boltzmann constant σ = 5.67∙10-8 [W/m2K4].
For a member with emissivity εm at a thermodynamic temperature of Tm exposed to a fire with
emissivity εf (generally equals to 1) at a thermodynamic temperature of Tf, the radiative heat flux
per unit surface area is determined according to EN 1991-1-2 by:
q̇ = σ ∙ εm ∙ εf ∙ Φ ∙ ( Tf4 - Tm4 ) [4.5]
The reference emissivity value of the member εm and the reference configuration factor
accounting for the visibility of the flames Φ are given in EN 1991-1-2 (2004), equal to 0.8 and 1.0
respectively if no specific data given.
Some reference emissivity values are listed as following (EN 1991-1-2, EN 1992 to EN 1996
and EN 1999):
• generic: 0.8 • concrete: 0.7
• carbon steel: 0.7 • stainless steel: 0.4

4.2 Thermal Problem

4.2.1 General Equations

As we discussed before, the main goal of the simulation is to obtain the thermal field inside a
specimen or a structural member, namely the temperatures at every node. Generally speaking, the
temperature in a continuum can be expressed as a function of spatial coordinates and time, as
shown in Equation [4.6]:
T = T ( x, y, z, t ) [4.6]
In Civil Engineering, conduction is the main mechanism of heat transfer, while convection and
radiation are usually considered as boundary conditions defining the thermal loads acting on the
surfaces of the domain. For a homogeneous, isotropic and continuous body, the specific heat flux
of two surfaces at a distance dx can be evaluated according to the Fourier Law, as in Figure 4-21.

58
Figure 4-21 Heat flux balance.

∂ ∂T
The balance of fluxes can be written as: ∂x ( λ ∂x ) = ∇ ∙ ( λ ∇T )
Let us define the concepts of stored energy and internal heat source, and enforce the
conservation of energy at any time for the unit volume. The stored energy can be expressed by
specific heat (per unit volume) ρc and the partial derivative of the temperature T with respect to
the time t, and is the sum of ingoing and outgoing fluxes and the internal energy (per unit volume)
H. The Energy Conservation Equation is expressed by Equation [4.7]:
∂T
ρc = ∇ ∙ ( λ ∇T ) + H [4.7]
∂t
To simplify the problem, two assumptions are introduced in our case. The first assumption is
that the conductivity of material is space invariant, with a constant value λ. So Equation [3.7] can
be rewritten as Equation [4.8]:
∂T λ 2 𝐻
= ∇ T+ [4.8]
∂t ρc ρc
The other assumption is that there is no internal heat sources, which means H = 0. Therefore,
Equation [3.8] is simplified as Equation [4.9], where Dq = λ/ρc is the thermal diffusivity as defined
in Chapter 2.
∂T
= Dq ∇2 T [4.9]
∂t
Specially, in an axisymmetric cylindrical coordinate system (r, θ and z), the general Energy
Conservation Equation and its particular form, once taking the two assumptions into account, are
expressed in Equation [4.10] and [4.11], respectively:
∂T ∂ ∂T
ρc∙ r = (λ∙r )+H∙r [4.10]
∂t ∂r ∂r

59
∂T Dq ∂T ∂2 T
= ( ) + Dq 2 [4.11]
∂t r ∂r ∂𝑟

4.2.2 Initial and Boundary Conditions

To obtain the thermal field within a domain, meaning to work out the special solution of the
differential Equation [4.9], the imposition of initial and boundary conditions is required.
Considering a domain Ω with the boundary Γ and the normal vector n, the most common initial
condition is expressed in Equation [3.12]:
T ( x, y, z, 0 ) = f ( x, y, z ) for ( x, y, z ) ∈ Ω [4.12]
Three different ways are given to define the required boundary conditions (Reinhard
Hinkelmann, 2005):
1) Specified temperature (Dirichlet-boundary condition)
T ( x, y, z, t ) = h ( x, y, z, t ) for ( x, y, z ) ∈ Γ [4.13]
This condition is simple but important as it determines the temperature history of boundary.
However practically relevant data cannot be obtained easily.
2) Specified heat flux (Neumann-boundary condition)
∂T ( x, y, z, t )
= q ( x, y, z, t ) for ( x, y, z ) ∈ Γ [4.14]
∂n
The Neumann-boundary condition defines a fixed heat flux for the surface exposed to fire
load. For an adiabatic surface, the thermal flux should be zero.
3) Convection condition (Cauchy-boundary condition)
∂T ( x, y, z, t )
= b [ p ( x, y, z, t ) - T ( x, y, z, t )] for ( x, y, z ) ∈ Γ [4.15]
∂n
The Cauchy-boundary condition is a linear combination of Dirichlet- and Neumann-
boundary condition, providing a thermal load driven by convection. The heat flux through
the boundary is proportional to the temperature difference between the surface and the
ambient fluid. A more common expression of Equation [4.15] is given in Equation [4.16]
for the ith point on the boundary with thermal convection coefficient αi and conductivity λi.
Temperatures of the point and the ambient fluid are, respectively Ti and Tf.
∂T ( x, y, z, t ) 𝛼i
= [ 𝑇f ( x, y, z, t ) - 𝑇i ( 𝑥i , 𝑦i , 𝑧i , t )] for ( x, y, z ) ∈ Γ [4.16]
∂n λi
Radiation is not taken into account in the previous discuss. If it is, the heat transfer problem is
no more linear and the computation is more complex.

60
4.2.3 Initial and Boundary Conditions in Fire Design

The definition of the appropriate initial and boundary conditions is essential to solve the thermal
problems in the most appropriate way, as previously mentioned. In real cases, generally the initial
condition is the initial temperatures within the member or structure, simplified as a single value
for all the points (usually the ambient temperature).
The boundaries of the domain can be considered as two types: hot and cold surface. For the
hot surface, thermal load is usually described by a fire curve for convection and radiation heat
transfer, with corresponding coefficients taking a reference EN 1991-1-2 (2004). The cold surface
can be dealt imposing the ambient temperature, an adiabatic boundary (Tan and Yao, 2003) or a
heat sink, which is close to the real situation. Generally, the thermal gradient at the cold surface is
low, leading to a small value of heat fluxes. Imposing the adiabaticity will lead to higher
temperatures, which brings the benefit for safety consideration.

4.2.4 Calibration of LVDT Measurement

The evaluation of the elastic modulus is usually based on the measurement performed by means
of DD1, which are very accurate since they directly measure the displacements in the specimen.
In hot compressive tests, however, DD1s were not available due to the high temperature at the
specimen surface, and only LVDTs measuring the platen-to-platen distance could be used. Since
LVDTs were influenced by the deformability of the steel disks (placed between the specimen and
the testing machine) a calibration is needed.

Figure 4-3 (a) Specifications of the specimen and positions of LVDT and DD1 (b) Top view

61
σ F
dsd = Hsd = H
Es Es . As sd
σ F
dDD1 = H𝐷𝐷1 = H
Es Es . As DD1
The average displacement measured by LVDT and the displacement measured by DD1 are
defined as dLVDT and dDD1, respectively. The two measurements are linked by Equation [4.1].
dLVDT = dDD1 + d𝑠𝑡𝑜𝑛𝑒 + d𝐻𝑂𝐿𝐷𝐼𝑁𝐺 𝐹𝑅𝐴𝑀𝐸 [4.1]
F
dLVDT = H + d𝑠𝑡𝑜𝑛𝑒 + d𝐻𝑂𝐿𝐷𝐼𝑁𝐺 𝐹𝑅𝐴𝑀𝐸 [4.2]
Es . As DD1
d𝑠𝑡𝑜𝑛𝑒 + d𝐻𝑂𝐿𝐷𝐼𝑁𝐺 𝐹𝑅𝐴𝑀𝐸 = d𝐶𝑂𝑅𝑅𝐸𝐶𝑇𝐼𝑂𝑁 = d𝐶 [4.3]
F
d𝐶 =dLVDT - H
Es . As DD1
F HDD1
d𝐶 =dLVDT - . H
Es . As HDD1 SP
HSP
d𝐶 =dLVDT - dDD1 . [4.4]
HDD1
d𝐶 =f (F) [4.5]
Hence, d𝐶 can be expressed as a polynomial function of the force applied F.
dLVDT CORRECTED = dLVDT - dC [4.6]
The strain of concrete εc is defined in Equation [5.4].
dLVDT CORRECTED
εc = [4.7]
H𝑆𝑃
The elastic modulus of specimen is expressed as:
σ
Ec =
εc
This procedure was applied to all the tests performed in virgin conditions, with a steel specimen
stone discs at the end, in which both LVDT and DD1 were used. That is because for all the tests,
the properties of stone disks were same for all the specimens, but the initial settlements between
specimen, testing machine and instrumentation was specific for every test.

62
Chapter 5 THE HYGRAL PROBLEM

In structural analysis of R/C members, aimed at understanding the global mechanical response,
concrete is usually considered as a homogeneous material, with well-defined thermo-mechanical
properties; this approach is not far from the reality when the local behaviour is not of primary
concern. This simplification is useful and convenient in numerical modelling. On the other hand,
when explosive spalling can be relevant and requires to be considered, concrete microstructure
local heat/water transfer phenomena should be taken into account. In such cases, concrete
heterogeneity needs to be considered for a better evaluation of the behaviour of the material and
even of the whole structure at high temperature.
As a composite material, concrete is basically composed of aggregates (sand, gravel or crushed
stone), water and cement paste, with dry air and vapour inside the voids. The voids (i.e. the
porosity) can influence the mechanical behaviours, such as stiffness and strength, although they
are believed to depend mainly on the components (aggregates, cement paste and fibres).
Concrete porosity is directly related to the transport properties, e.g. the permeability. The water
transport has a significant impact on the concrete behaviour, including shrinkage, creep and
relaxation, heat transport, chemical diffusion etc. Actually the service life of R/C structures has a
strong dependency on concrete transport properties.
In order to have a better understanding of the transport properties, two main parameters should
be defined:
• the total porosity ϕ: the fraction between the volume of voids Vp and the total volume of the
material V:
Vp
ϕ= [5.1]
V
• the degree of saturation Si: the fraction between the volume of the evaporable water Vl and
the volume of voids Vp :

63
Vl
Si = [5.2]
Vp
In the hydrated cement paste, voids are classified into three types on the basis of their size:
– gel pores (diameter, d < 10 nm), namely interlayer hydration space;
– capillary pores (10 nm < d < 100 μm);
– macro pores (d > 100 μm).
Voids are occupied by the bulk water (evaporable water) and gaseous phase (dry air and
vapour). The bulk water is going to evaporate at T = 105°C. Macro pores have little influence on
concrete shrinkage, while capillary pores and gel pores have great impact, especially for those of
size between 2.5 and 50 nm (Guo Y. C. et al., 2013).
In the hydrated cement paste, water can exist in many forms, according to the strength of the
bond with the cementitious matrix (see also Error! Reference source not found.):
– capillary water, existing as free water in large capillaries, while in small capillaries it can
cause shrinkage when lost; it can be easily expulsed during drying;
– adsorbed water: is physically adsorbed, and causes shrinkage when lost; it can be expulsed
during dying for about 30% relative humidity;
– interlayer water, existing as monomolecular layers between two C-S-H layers, causing
substantial shrinkage if lost; it can be expulsed during drying for about 11% relative
humidity;
– chemically combined water: is part of hydration products, and can be expulsed only during
dehydration due to the decomposition during heating.
Since the cement paste and aggregates are different constituents in concrete, their mutual
interaction results in the separation of water from the cement particles and there will be a narrow
region around aggregate particles with fewer cement particles and more water, namely the
Interfacial Transition Zone (ITZ). This zone of discontinuity is not uniform and is characterized
by larger porosity, playing a major role as weakest link in micro-cracking (even before loading).

Figure 5- 1 Types of water existing in concrete (Feldman and Sereda, 1968).

The behaviour of concrete is affected by many microscopic properties and their interaction,

64
and it can be conjectured that the macroscopic properties of concrete will be more complex at high
temperature considering the complex physical and chemical reactions involved.
A rough summary of the main processes occurring during heating is listed in the following:
• water evaporation (capillary water, adsorbed water and interlayer water) in the temperature
range of 100-250°C;
• dehydration (C-S-H and carbon hydroxide);
• expansion of aggregates;
• shrinkage of the cement paste;
• micro-cracking because of thermal kinematic incompatibility between aggregates and cement
paste (T ≥ 250-300°C).

5.1 Liquid-Vapour Equilibrium

In High-Performance Concrete, voids are extremely fine and contain water in different forms.
When concrete is exposed to fire or other severe thermal conditions, large amount of water inside
concrete will be removed via release and evaporation, and temperature gradients will be induced.
Water evaporation leads to pore-pressure build up and to concrete shrinkage, while thermal
gradients can induce thermal stresses interacting with the external loads. Once the internal stresses
exceed the maximum allowable tensile stresses, the thermal cracks and spalling occur (Fu and Li,
2010).
In water evaporation, an important relationship is the liquid-vapour equilibrium. In the
capillary pores, there is an interfacial tension of water that causes the generation of concave
meniscus between liquid and gas, as shown in Figure 5- 2 Schematic representation (a) of a
capillary pore and (b) of a pore with liquid water, vapour and dry air (after Meftah, 2005).

a. Thus, a discontinuity of fluid pressure, i.e. capillary pressure, is formed; it represents the
pressure difference between liquid water and gas (Equation [4.3]). The capillary pressure is a
function of liquid water saturation Sl.
pc (Sl ) = pv + pa - pl [5.3]
where,
pc is the capillary pressure;
pv and pa are the pressures of vapour and dry air, respectively (their sum is the gas pressure,
according to Principle of Dalton, as shown in Figure 5- 2 Schematic representation (a) of a
capillary pore and (b) of a pore with liquid water, vapour and dry air (after Meftah, 2005).

b);
pl is the liquid water pressure.

65
(a) (b)
Figure 5- 2 Schematic representation (a) of a capillary pore and (b) of a pore with liquid water, vapour and dry air
(after Meftah, 2005).

According to the Young-Laplace Equation, the capillary pressure pc is proportional to the


interfacial tension σ, and inversely proportional to the pore radius r, also depends on the wetting
angle of the liquid on the surface of the capillary. Assuming the wetting angle equal to zero, the
capillary pressure can be expressed as:
2σ (T)
pc (Sl ) = = 2σ (T)χ [5.4]
r
where χ is the meniscus curvature.
Liquid pressure is defined according to the generalized Clapeyron relation, Equation [5.5].
ρl (T) RT p
pl = ln [ v ] + pvs (T) + pa [5.5]
Ml pvs (T)

5.2 Moisture Transport

5.2.1 Physical Mechanisms in Hygral Problems

The hygro-thermal problem is the basic issue for the evaluation of pore pressure at high
temperature. Such problem is strictly related to the physical mechanisms of mass transport
phenomena. Water in concrete exists in different forms and in mass transport both liquid and gas
phases need to be considered.
The physical mechanisms include advection and diffusion. In the former, liquid and gas
transport is governed by pressure gradients and permeability K’, while in the latter, it is governed
by concentration gradients and diffusivity Ddiff. The expressions of these two mechanisms (Darcy’s
Law and Fick’s law) are shown as Equation [4.18] and [4.19], respectively (Schrefler and
Pesavento, 2004; and Daïan, 2001):

66
v = -K'∙∇(p) [5.6]
∂w
= Ddiff ∙∇2 (w) [5.7]
∂t
The parameter K’ is the conventional permeability, defined as:
kri
K' = -K [5.8]
μf

where K is the intrinsic permeability, only related to the material itself; kri is the relative
permeability of a particular fluid, dependent on the degree of saturation of the liquid phase; μf is
the dynamic viscosity of the fluid.
Generally speaking, advection is the main mechanism for liquid water and dry air, while for
vapour both mechanisms exist. The abovementioned mechanisms are plotted in Error! Reference
source not found..

Figure 5- 3 Plot Mass transport mechanisms in an unsaturated porous material during heating (after Meftah, 2005).

5.3 Hygro-Thermal Problem

Thermal problem in concrete is usually addressed considering the heat transfer due to three
mechanisms only: conduction, convection and radiation, thus ignoring internal heat sources, as
already discussed in Chapter 3. Conduction is regarded as the only mechanism of heat transfer
inside concrete, while convection and radiation play a role as boundary conditions, controlling the
energy ingoing to the external surface. In this way, however, some physical phenomena are
neglected:

67
• the energy transferred by the fluids in relative motion with respect to the solid phase;
• the energy absorbed by the endothermic reactions of vaporization and dehydration. .
Considering these phenomena, the Enthalpy Conservation Equation becomes:
∂T ∂ρvap ∂ρdehydr
ρc + Hvap + Hdehydr = ∇ ∙ ( λ∇T ) [5.9]
∂t ∂t ∂t
where,
c and ρ are specific heat and density of concrete, respectively;
Hvap and Hdehydr are the enthalpies of water vaporization and cement dehydration, respectively;
ρvap is the concentration of the vapour;
ρdehydr is the density of the dehydrated cement.
The Equation [4.21] makes it possible to evaluate the thermal fields inside the concrete with
considering the hygral aspect. However, there are two more unknowns added, ρvap and ρdehydr,
which means more equations are needed to solve the hygro-thermal problem. One possible way is
imposing the mass conservation.
In the present Chapter, the hygral framework has been discussed. In the following, transport
phenomena and water evaporation are considered. This approach has been adopted as concrete
behaviour will be studied in the present work of thesis for lower temperatures (250°C), when
evaporation of pore water is at its peak. Physical processes related to water, mainly influence
concrete behaviour in the range T = 100-300°C.

68
Chapter 6 NUMERICAL ANALYSIS

6.1 Introduction

When exposed to fire, concrete members experience high thermal gradients due to the low thermal
diffusivity of the material, and pore pressure builds up because of water evaporation in the voids.
Pressure gradients drive the mass transfer of gas and water both towards the exposed face and the
inner (and colder) region. In the latter case, vapour condensates, so locally increasing the saturation
degree of the pores with a subsequent reduction of the gas permeability (Kalifa et al., 2000). This
quasi-saturated layer is often referred as moisture clog and can favour the significant increase of
the pore pressure. Due to its denser matrix and low permeability, High-Performance Concrete
proved to be more prone to such phenomenon and, in general, more heat-sensitive with respect to
Normal-Strength Concrete.
As already mentioned, High-Performance Concrete becomes more widely used, especially for
structures subjected to extreme environments. This cementitious material, on the one hand have
denser matrices, which ensure better durability and mechanical properties (at the cost of greater
brittleness and sensitivity to high temperature due to the low porosity), and on the other hand
justify the optimization of the mix design in terms of aggregates, fibres and other constituents. In
the design of structures, to guarantee the desired level of safety during fire, the concrete mix should
be defined in order to minimize the decay of mechanical properties and the risk of spalling.
Within this context, an experimental campaign has been launched in 2015 at the Politecnico di

69
Milano. The main purpose of the project is to investigate the behaviour during heating and after
cooling of concrete and concrete structures. This goal is pursued by studying the compressive
behaviour and the spalling sensitivity of eleven different concrete mixes. In the present work of
thesis, the characterization of the compressive behaviour of the concretes have been investigated
during heating (hot conditions), at the reference temperature of 250°C.

6.2 Description of the Concrete Model

6.2.1 Concrete Model

The main target of investigation is that Designing the compression test for concrete in hot
conditions using numerical simulations: reference temperatures T = 20°C and 250°C. The
mechanical properties fc, Ec, and εc1 in hot conditions are evaluated by means of displacement-
controlled compressive tests on cylinders. For all the tests in compression, cylindrical specimens
(Ø = 100 mm and h = 200 mm) have been used.
For all the tests in hot condition, the heating rate was 1°C/min with an initial temperature 20°C,
and after target temperature (250°C) was reached, it was maintained for 1 hour. The specimen was
then subjected to compression loading.
As discussed in the following, the specimens were thermally insulated before the introduction
in the electric furnace, and this lead, inside the specimen, more pronounced thermal gradients
during the heating process, with respect to bare specimens. Thus, although the heating rate was
quite low (1°C/min), the thermal gradients were not negligible during heating; however, after two
hours at the target temperature, the thermal field inside the specimen was practically uniform.
Before starting the programme of heating, the crucial problem that needs to be deal with is to
check if the thermal field within the concrete coincides with the designed one. Thus, the
experimental temperature monitoring process was performed on a plain concrete cylinder with six
thermocouples plugged inside. The models as per Abaqus are described below.

Input Temp Profile


300

250

200
Temperature

150

100

50

0
0 100 200 300
Time (min)

70
Figure 6-1 Thermal cycles for the reference temperature of 250°C.

Figure 6-2 Whole 2D Model

Figure 6-3 Quarter of the 2D Model as in the analysis

71
Figure 6-4 3D and 2D schematic descriptions of the model with dimensions

6.2.2 Properties of Concrete

6.2.2.1 Thermal Properties

As mentioned in Chapter 3, all the properties of concrete used for the analysis are as per EN 1992-
1-2:2004. The following table shows the values of the thermal properties of concrete which are
applied in ABAQUS.

Table 6- 1 Thermal properties of concrete

Temperature Conductivity Density Specific Diffusivity


T k ρ Heat c D
°C W/m K kg/m3 J/kg K mm2/s
0 1.36 2400 900 0.63

72
100 1.2297 2400 900 0.57
115 1.211138 2400 900 0.56
200 1.1108 2352 1000 0.47
300 1.0033 2316 1050 0.41
400 0.9072 2280 1100 0.36
500 0.8225 2259 1100 0.33
600 0.7492 2238 1100 0.30
700 0.6873 2217 1100 0.28
800 0.6368 2196 1100 0.26
900 0.5977 2175 1100 0.25
1000 0.57 2154 1100 0.24
1100 0.5537 2133 1100 0.24
1200 0.5488 2112 1100 0.24

6.2.2.2 Mechanical Properties

Table 6- 2 Mechanical properties of concrete

Temperature °C Modulus of Poisson’ Thermal Coefficient


Elasticity s ratio Strain of Thermal
MPa Expansion
T E µ ϵth α
20 36000 0.2 1.48E-07 7.4E-09
100 22500 0.2 0.000619 6.19E-06
200 15709.09 0.2 0.001498 7.49E-06
300 11314.29 0.2 0.0026 8.67E-06
400 7200 0.2 0.004034 1.01E-05
500 4020 0.2 0.005913 1.18E-05
600 1890 0.2 0.008346 1.39E-05
700 1314 0.2 0.011441 1.63E-05
800 756 0.2 0.013 1.63E-05
900 414 0.2 0.013 1.44E-05
1000 180 0.2 0.013 0.000013
1100 54 0.2 0.013 1.18E-05
1200 0 0.2 0.013 1.08E-05

73
Concrete Damage Plasticity
CDP data

Dilation Eccentricity fb0/fc0 K Viscosity


Angle Parameter
30 0.1 1.16 0.67 0
Compression Hardening

σ Vs ϵ
70
60
50
Stress MPa

T 400
40
T 300
30
T 200
20
T 100
10
T 20
0
0 0.01 0.02 0.03 0.04
Strain

Tension Hardening

σ Vs ϵ
5
4.5
4
3.5
Stress MPa

T 400
3
2.5 T 300
2
T 200
1.5
1 T 100
0.5 T 20
0
0 0.01 0.02 0.03 0.04
Strain

74
6.2.3 Properties of Other Materials

6.2.3.1 Thermal Properties of Steel Disc and Steel Actuator

As mentioned in Chapter 3, all the properties of concrete used for the analysis are as per EN 1992-
1-2:2004. The following table shows the values of the thermal properties of concrete which are
applied in ABAQUS.

Table 6- 3 Thermal properties of Steel

Temperature Conductivity Density Specific Heat


T k ρ c
°C W/m K kg/m3 J/kg K
20 53.334 7850 439.8018
100 50.67 7850 487.62
200 47.34 7850 529.76
300 44.01 7850 564.74
400 40.68 7850 605.88
500 37.35 7850 666.5
600 34.02 7850 759.92
601 34.02 7850 760.9051
700 30.69 7850 1008.158
735 30.69 7850 5000
736 30.69 7850 4109
800 27.3 7850 803.2609
900 27.3 7850 650.4438
901 27.3 7850 650
1000 27.3 7850 650
1100 27.3 7850 650

6.2.3.2 Thermal Properties of Stone Disc

The thermal properties for the stone disc (aggregate) are very similar to that of concrete, hence in
the analysis, the same thermal properties are used for the stone disc as that of concrete (see 1.2.2.1).

6.2.3.3 Mechanical Properties of Steel Disc, Steel Actuator and the Stone Disc

During the mechanical analysis, we want to evaluate the stresses (thermal and compression) only
in the concrete model, hence, the properties of rest of the materials is kept negligible, as mentioned

75
below.

Elastic Modulus Poisson’s ratio


E = 0.001 MPa µ = 0.2

6.3 Thermal Analysis in Abaqus

The various steps involved in the thermal analysis are described below:

6.3.1 Steps

For the thermal analysis, one step is created, i.e. Heat Transfer (transient), with the
following values:

-duration of 17400 seconds (230 minutes of heating, and 60 minutes of loading)


-Maximum number of increments-5000
Initial Increment size-60
Minimum Increment size-5
Maximum Increment size-60
Maximum allowable temperature change per increment-200
Maximum allowable emissivity change per increment-1

6.3.2 Boundary Conditions

We provide a constant boundary condition at the bottom of the actuator, for a temperature of 20°C,
throughout the analysis.

6.3.3 Predefined Fields

We apply a predefined field for the whole model, creating the initial temperature equal to 20°C,
which is constant throughout the region.

6.3.4 Interaction

As discussed in previous chapters, heat transfers can be through convection, conduction or


radiation. Among these conduction occurs within the material subjected to heat. So subject our
model to the other two interactions, which are convection and radiation. The values of coefficients
for convection and radiation are referred from the Eurocode 2, mentioned below. These

76
interactions are applied to the surface subjected to heat which includes the concrete model, the
steel disc, and 0.015m of the stone disc, which means 0.12m in length for the quarter of the model.
For the rest of the surface, we have applied a convection with an instantaneous amplitude.
Coefficient of convection-4
Coefficient of radiation-0.7

6.3.5 Mesh

A uniform quadratic mesh of size 0.01 is chosen for the whole model, as we have limited number
of nodes (being the student edition). The element type of the mesh chosen is Heat Transfer.

6.4 Mechanical Analysis in Abaqus

The various steps involved in the thermal analysis are described below:

6.4.1 Steps

For the mechanical analysis, two steps are created, Heating (general static), and Loading (general
static) with the following values:

Heating step
-duration of 17400 seconds (230 minutes of heating, and 60 minutes of loading)
-Maximum number of increments-5000
Initial Increment size-60
Minimum Increment size-5
Maximum Increment size-60

Loading step

77
-duration of 300 seconds
-Maximum number of increments-5000
Initial Increment size-5
Minimum Increment size-0.0001
Maximum Increment size-5

6.4.2 Boundary Conditions

We provide three different boundary conditions for the concrete model, for the bottom surface, the
top surface (being the mid surface the actual model), and the inner vertical surface, described
below.
-For the Top horizontal surface
-Created at Step 1 (the heating step)
-Type of Boundary Condition- Symmetry/Antisymmetry
-Direction- Y-SYMMETRIC (U2=UR1=UR3=0)
-For the Inner vertical surface
-Created at Step 1 (the heating step)
-Type of Boundary Condition- Symmetry/Antisymmetry
-Direction- Y-ANTI-SYMMETRIC (UR2=U1=U3=0)
-For the Bottom surface
-Created at Step 2 (the loading step)
-Type of Boundary Condition- Displacement/Rotation
-Direction- U2
-Amplitude
Time Frequency
0 0
300 0.001

6.4.3 Predefined Fields

We apply a predefined field for the whole model, using the results of the thermal analysis, which
was hence analysed separately, as shown below.

78
6.4.4 Interaction

For the heating phase, we use the output database of the thermal analysis in the form of predefined
fields, as described earlier. So we do not subject our model to any other interaction surface. But
we provide a constraint (tie) between the concrete model and the steel disc, as it is subjected to a
compression load.

6.4.5 Mesh

A uniform quadratic mesh of size 0.01 is chosen, same as that for thermal analysis, as mesh should
be the same in both the analysis. The element type of the mesh chosen is Axis symmetric stress.

79
Chapter 7 RESULTS AND CONCLUSIONS

7.1 Temperature Monitoring and Numerical Modelling

As introduced previously, six thermocouples were put inside one specimen exposed to a thermal
cycle up to 250°C, including a heating phase at 1°C/min, a steady phase at 250°C for 1 hour. The
temperature of the six thermocouples and temperature inside the furnace were recorded as shown
in Figure 7- 1
The mechanical behaviour of the specimen during the compressive tests was numerically
simulated as well, in order to understand if the thermal gradients developed inside concrete
occurring while the compressive test is performed may have some influence. Numerical
simulations showed that the abovementioned difference in terms of temperature inside the
specimen have negligible influence on the failure stress and on the elastic modulus evaluated by
means of the compressive test.

7.2 Temperature profile obtained from Thermal Analysis

(a) (b)

80
(c) (d)
Figure 7- 1 Numerical modelling: thermal fields of the specimen (a) at the end of thermal analysis i.e. 290 min (b) At
the mid of heating, 150 min, (c) At the end of heating, 230 min, and (d) At the mid of constant one hour, 260 min .

Figure 7- 2 Description of the Abaqus model

Table 7- 1 Temperature profile at different time intervals for the top surface AB
x T-0 T-30 T-60 T-90 T-120 T-180 T-240 T-310
0 20 22.1039 30.6569 45.2208 65.0588 117.451 179.929 236.937
0.008333 20 22.1855 30.8449 45.5133 65.4555 118.045 180.699 237.371
0.016667 20 22.4355 31.4145 46.3965 66.6516 119.835 183.015 238.657
0.025 20 22.8694 32.3812 47.8867 68.6642 122.841 186.897 240.753

81
0.033333 20 23.5131 33.7711 50.0112 71.5223 127.095 192.375 243.589
0.041667 20 24.4026 35.6209 52.8086 75.2664 132.645 199.496 247.066
0.05 20 25.5847 37.978 56.3288 79.9495 139.552 208.315 251.06

AB- Top surface


300

250 T-0

200 T-30
Temperature

T-60
150
T-90
100 T-120

50 T-180
T-240
0
0 0.01 0.02 0.03 0.04 0.05 T-310

Distance

Figure 7- 3 Temperature Time curve for the top surface

Table 7- 2 Temperature profile at different time intervals for the vertical surface AC
x T-0 T-30 T-60 T-90 T-120 T-180 T-240 T-310
0 20 22.1039 30.6569 45.2208 65.0588 117.451 179.929 236.937
0.007692 20 22.1034 30.6486 45.193 65.0005 117.313 179.714 236.646
0.015385 20 22.1017 30.6233 45.1079 64.8229 116.894 179.061 235.761
0.023077 20 22.0986 30.5788 44.9613 64.5186 116.18 177.944 234.243
0.030769 20 22.0938 30.5122 44.7459 64.0755 115.145 176.32 232.03
0.038462 20 22.0869 30.4193 44.4524 63.4778 113.755 174.134 229.037
0.046154 20 22.0777 30.296 44.0702 62.7074 111.972 171.318 225.158
0.053846 20 22.067 30.1393 43.5904 61.7478 109.758 167.8 220.276
0.061539 20 22.0575 29.9499 43.0096 60.5898 107.083 163.525 214.279
0.069231 20 22.0557 29.7367 42.337 59.242 103.949 158.474 207.087
0.076923 20 22.0752 29.5234 41.6053 57.7461 100.409 152.709 198.685
0.084615 20 22.1402 29.3566 40.8834 56.194 96.6018 146.413 189.176
0.092308 20 22.2898 29.3124 40.2859 54.7422 92.777 139.926 178.804
0.1 20 22.5783 29.4961 39.9728 53.6098 89.2864 133.742 167.944
0.105 20 22.5751 29.481 39.9381 53.549 89.1567 133.53 167.659
0.1125 20 21.9101 27.4905 36.0755 47.3025 76.7132 113.322 142.585
0.12 20 21.3892 25.7555 32.5616 41.4947 64.9094 94.0003 117.877

82
0.1275 20 20.9701 24.2045 29.2985 36.0013 53.565 75.3292 93.4842
0.135 20 20.6207 22.783 26.2136 30.7337 42.5644 57.1739 69.4663
0.1425 20 20.3164 21.4486 23.2519 25.6282 31.8347 39.4662 45.9012
0.15 20 20.0365 20.1637 20.3652 20.6298 21.3184 22.1619 22.8624

Figure 7- 4 Temperature time curve for the left surface

Table 7- 3 Temperature profile at different time intervals for the outer vertical surface BD
x T-0 T-30 T-60 T-90 T-120 T-180 T-240 T-310
0 20 25.5847 37.978 56.3288 79.9495 139.552 208.315 251.06
0.007692 20 25.5841 37.9703 56.3041 79.8995 139.444 208.166 250.883
0.015385 20 25.5819 37.9463 56.2283 79.7468 139.114 207.712 250.343
0.023077 20 25.5777 37.9034 56.096 79.4829 138.546 206.928 249.409
0.030769 20 25.5705 37.837 55.8979 79.093 137.714 205.773 248.03
0.038462 20 25.5583 37.7398 55.6198 78.5548 136.574 204.185 246.128
0.046154 20 25.5383 37.6011 55.2411 77.8366 135.069 202.074 243.594
0.053846 20 25.5056 37.4051 54.7324 76.8935 133.115 199.312 240.274
0.061539 20 25.452 37.1279 54.0514 75.6618 130.595 195.717 235.954
0.069231 20 25.3629 36.7316 53.1327 74.0443 127.332 191.013 230.323
0.076923 20 25.2142 36.1574 51.8779 71.8975 123.067 184.791 222.933
0.084615 20 24.9436 35.2677 50.0646 68.9017 117.235 176.16 212.877
0.092308 20 24.4722 33.8904 47.4168 64.6627 109.141 164.023 199.038
0.1 20 23.0378 30.4569 41.4743 55.7089 92.7992 139.034 172.694
0.105 20 23.03 30.4322 41.4252 55.6292 92.644 138.795 172.405

83
0.1125 20 23.5236 30.8693 41.3997 54.8299 89.6946 133.609 162.022
0.12 20 22.5427 28.1603 36.315 46.7548 73.8672 107.94 131.052
0.1275 20 21.4129 25.0939 30.6165 37.7602 56.3336 79.4323 96.6931
0.135 20 20.8361 23.1935 26.7827 31.4444 43.5604 58.5516 70.1293
0.1425 20 20.3968 21.5807 23.4007 25.7696 31.9183 39.4861 45.4398
0.15 20 20.0215 20.0921 20.2028 20.3477 20.724 21.1852 21.5604

BD-right surface
300

250 T-0
T-30
Temperature

200
T-60
150
T-90
100
T-120
50 T-180

0 T-240
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 T-310
Distance

Figure 7- 5 Temperature time curve for the left surface

84
7.3 Results obtained from Mechanical Analysis

7.3.1 Temp Vs Time at diff locations

Figure 7- 6 Positions of the six thermocouples inserted

300
T_ove
Temperature vs Time-Abaqus n
250
T_5

200 T_15
Temperature

T_30
150
T_50
100
T_100

50 T_100-
ext
0
0 50 100 150 200 250 300 350
Time (min)
Figure 7- 7 Temperature time curves for the six thermocouples

85
Temp vs Time
300

250

200

Temp
150
Exp-T oven
100
T_oven
50

0
0 100 200 300 400 500
Time min

These graphs are obtained by comparing the results of the experiment and from Abaqus plots. It is
observed that the graphs don’t coincide exactly, due to differences in the thermal properties of
concrete used and the actual ones. For abaqus data, we have used the properties assuming the
following conditions;
1. 0% moisture content in concrete- for variations in specific heat
2. Lower bound - for thermal conductivity
In order to coincide the graphs, we tried changing the assumptions, but the difference was
increased on increasing the conductivity, and increasing the moisture content. But according to
Eurocode 2, these are the limits for the properties. So the following might be the possible reasons:
(a) The concrete might be totally dry (0% moisture).
(b) The conductivity of the concrete model is even lesser than the lower bound used. This
might be because according to the code, after heating the concrete stays in open
atmosphere, but during the test we keep the concrete insulated in the same heating press
for one more hour after heating, so the heat couldn’t pass.

Temp vs Time
180
160
140
120
100
Temp

80 Exp-T_5
60 T_5
40
20
0
0 100 200 300 400 500
Time min

86
Temp vs Time
200
180
160
140
Temp 120
100
Exp-T_15
80
60 T_15
40
20
0
0 100 200 300 400 500
Time min

Temp vs Time
250

200

150
Temp

Exp-T_30
100
T_30
50

0
0 100 200 300 400 500
Time min

87
Temp vs Time
250

200

Temp 150

Exp-T_50
100
T_50
50

0
0 100 200 300 400 500
Time min

Temp vs Time
300

250

200
Temp

150
Exp-T_100
100 T_100

50

0
0 100 200 300 400 500
Time min

88
Temp vs Time
300

250

200
Temp
150
Exp-T_100 ext
100 T_100-ext

50

0
0 100 200 300 400 500
Time min

89
7.3.2 Stress curves

s11 x radial stress


s22 y vertical stress
circumferential
s33 z stress
s12 x-y shear stress

 Heating Phase- Top surface

The top surface here corresponds to the mid surface of the concrete model in the experiment.
During the heating phase, there is compression at the ends and tension in the mid region, as seen
in fig. Both the circumferential and the vertical stress follow this trend, but radial stress are tensile
along this region, due to linear heating, which reduce to half when the concrete is kept at rest after
230 minutes of heating at the rate 1°c/min. The magnitude of these stresses is very less as compared
to the strength of concrete. And the magnitude of shear stress observed are negligible.

Figure 7- 8 Stresses for the Top surface- Heating phase

90
Circumferential stress at the mid surface
1.5
1
5 min
0.5
10 min
0
Stress MPa -0.5 0 0.02 0.04 0.06 0.08 0.1 30 min
-1 60 min
-1.5 120 min
-2
230 min
-2.5
-3 290 min
Distance (m)

 Loading Phase- Top surface

During the loading phase, the magnitude of both the circumferential and the radial stress reduce to
approximately 10% of the previous values, but follow the same trend as during the heating phase,
but vertical stresses are observed to increase to a magnitude of 53.7 MPa, which is obviously due
to the compression loading applied, in order to obtain a deformation of 0.001m..The magnitude of
shear stress are again negligible. So the stress graphs for the mid surface are justified.

Figure 7- 9 Stresses at the top surface- Loading Phase

91
 Heating surface- Bottom surface

At the bottom surface, all the stresses are negligible during the heating phase, as there is an
interface of the concrete surface with the steel disc.

Figure 7- 10 Stresses at the bottom surface- Heating phase

92
 Loading surface- Bottom surface

During the loading phase, a compression load is applied at the bottom surface equivalent to obtain
a displacement of 0.001m. So High vertical stresses are observed, around 54MPa (compressive).
Also, tensile radial stresses and compressive circumferential stresses occur in this region but the
magnitude is very less as compared to the vertical stress.

Figure 7- 11 Stresses at the bottom surface- Loading phase

93
 Heating surface- Inner vertical surface

The trend of compression and tension in the inner vertical surface is similar to the horizontal mid
surface, i.e. compression at the mid region and tension at the ends. This trend is followed by both
radial and circumferential stresses during the heating phase. And the vertical stresses are always
tensile, because during heating, the outer surface tends to expand, and so the inner surface contracts
in order to prevent this expansion, resulting in tensile thermal stresses vertically.

Figure 7- 12 Stresses at the inner vertical surface- Heating phase

94
 Loading surface- Inner vertical surface

During the loading phase, the radial and circumferential stresses drop to about 5 to 10% of the
previous values. And compressive vertical stresses develop as a result of loading.

Figure 7- 13 Stresses at the inner vertical surface- Loading phase

95
7.3.3 Force - Displacement Curve

This graph shows the restraining force at the bottom surface at equidistant nodes during the loading
phase, and they are negligible during the heating phase.

Restraining Force vs time-Loading Phase


50

40
Node 1
30
Node 2
Force kN

20 Node 3
Node 4
10
Node5

0 Node 6
289 290 291 292 293 294 295 296
-10
Time (mintue)

Figure 7- 14 Force Displacement curve during the loading phase


The following is the graph of vertical displacement in the concrete during the heating phase. Due
to tensile thermal stresses, the displacement has negative magnitude. Maximum observed
displacement is about 0.16mm.

Vertical Displacement vs time Heating Phase

0
-0.00002 0 100 200 300 400
1
-0.00004 2
-0.00006
3
Disp (m)

-0.00008
4
-0.0001
5
-0.00012
-0.00014 6

-0.00016
-0.00018
Time min
Figure 7- 15 Displacement vs Time curve during heating phase

The following graph of displacement during the loading phase is clearly linear as this is the input

96
displacement curve, and our concrete is subject to a compressive load in accordance with this
curve.

Displacement vs time loading phase


0.0012

0.001

0.0008
1
0.0006 2
Displ (m)

0.0004 3

0.0002 4
5
0
289 290 291 292 293 294 295 296 6
-0.0002

-0.0004
Time min

Figure 7- 16 Displacement vs Time curve during loading phase


This is the total Force-Displacement curve during the whole cycle, used to obtain the stress-strain
curve.

160 F-U
140

120

100
Force kN

80

60

40

20

0
-0.0002 0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012
Displacement (m)
Figure 7- 17 Force Displacement curve

97
7.3.4 LVDT correction

The LVDT calibration is required before the tests as described in previous chapter. We have
installed 3 DD1s and 3 LVDTs to the steel specimen for measuring the displacement. For DD1,
we have obtained an average stress-strain graph, as the results of the 3 LVDTs are very similar.
But for LVDT, we have three result stress-strain graphs and one average graph, as shown below.

80.0
DD1_average
70.0

60.0
Stress [MPa]

50.0

40.0

30.0

20.0

10.0
strain_DD1 [-]
0.0
0.000E+00 1.000E-04 2.000E-04 3.000E-04 4.000E-04 5.000E-04

Figure 7- 18 Stress- Strain curve from DD1

80.0

70.0
LVDT
60.0
Stress [MPa]

50.0
LVDT1
40.0
LVDT2
30.0

20.0 LVDT3

10.0
strain_LVDT
0.0
0.000E+00 1.000E-03 2.000E-03 3.000E-03 4.000E-03 5.000E-03 6.000E-03

Figure 7- 19 Stress- Strain curve from LVDT


Then, the correction for the LVDT is calculated in the terms of dLVDT - dDD1 as derived in the
equation [5.4]

98
HSP
d𝐶 =dLVDT - dDD1 . [5.4]
HDD1
Then the displacement load curve is obtained with load in the x-axis in order to obtain the
displacement as a polynomial function of load.

1.000E+00
9.000E-01
LVDT1 Correction on LVDT
8.000E-01 LVDT2
LVDT Correction [mm]

7.000E-01
LVDT3
6.000E-01
Series4
5.000E-01
4.000E-01 Series5
3.000E-01 Series6
2.000E-01
1.000E-01
Load F [MN]
0.000E+00
0.000 0.050 0.100 0.150 0.200 0.250 0.300 0.350 0.400 0.450 0.500 0.550 0.600 0.650

Figure 7- 20 LVDT Correction curve with polynomial functions

Now as we have obtained the polynomial function, we perform the test with the concrete specimen
at 20°C, and obtain the LVDT values, and apply the correction.
Hence, we obtain the final stress-strain curve with DD1 values, LVDT values, and the corrected
LVDT values as shown below.

40.0 Stress-Strain
35.0

30.0

25.0
Stress MPa

LVDT corrected
20.0
DD1m
15.0
LVDT
10.0

5.0

0.0
-0.001 0 0.001 0.002 0.003 0.004 0.005 0.006
Strain

Figure 7- 21 Stress strain curves obtained from DD1, LVDT, and corrected curve

99
7.3.5 Stress strain curve obtained

The stress-strain curve obtained from the respective force-displacement curve is shown below. It
is compared to the stress-strain curve for concrete at 250°C, as per the euro code. It is observed
that the compressive strength of concrete achieved is comparable to the expected strength. But the
curve breaks at a certain point whereas the reference graphs continues to achieve zero stress. The
reason might be that the overall stiffness of the concrete is reduced due to the application of high
temperature.

σ Vs ϵ
60

50

40
Stress

30

20 Temp 250
output
10

0
0 0.005 0.01 0.015 0.02 0.025 0.03
-10
Strain

Figure 7- 22 Stress strain curves obtained from Eurocode and Abaqus

Comparison of stress strain curve


Input Peak Stress 53.9 MPa
Output Peak Stress 54.5 MPa

100
Chapter 8 SUMMARY

Concrete has low thermal diffusivity and good mechanical response at high temperature as
compared to the other construction materials, hence it shows an acceptable behaviour. However,
several well-known fire accidents (English Channel, 1996; Stucky Mill fire, Venice, 2003;
Shanghai, High Rise Apartment-Fire, 2010), were reported through time and serious damage
occurred, concerning both human safety and bearing capacity in terms of stability and integrity.
These examples indicate that concrete exposed to fire represents an important problem, which
needs further research to be fully understood in order to provide design tools, able to predict the
mechanical response of different concrete mixes and R/C structures
In the last decades, many studies have been carried out on the abovementioned topics stressing
the role of some issues, but understanding all the intricacies of heat-exposed concrete needs further
efforts. In this research work, the main objective was to study the variations in thermal and
mechanical properties of concrete, and to design a test set-up for Hot-Compressive test in order to
attain required temperature and displacement.
The main conclusions that can be drawn are summarized as follows:
• High-Performance Concrete is significantly more sensitive to high temperatures than
Normal-Strength Concrete, in terms of normalized decay of the mechanical properties.
• From the temperature-time plots obtained from abaqus and experiments, it can be seen that
the concrete specimen had very less moisture content and low conductivity.
• During heating, Compressive stresses (circumferential and radial) develop at the ends of
concrete specimen due to the thermal gradients, and similarly, tensile stresses develop at the
mid-region of the specimen (at both horizontal middle face and vertical face); no stresses
develop at the bottom face;

101
 During the loading face, high vertical stresses develop specimen due to the compressive
load applied.
 A negative displacement of about 0.16 mm is observed during heating due to the thermal
expansion. A linear displacement curve is obtained with a maximum displacement of 0.1
mm which is actually applied externally in the numerical simulations.
 The stress-strain curve obtained from the respective force-displacement curve is compared
to the stress-strain curve for concrete at 250°C, as per the euro code. The compressive
strength of concrete achieved is comparable to the expected strength. But the overall
stiffness of the concrete is reduced due to the high thermal gradients developed.

102
REFERENCES

Bamonte P. and Gambarova P. G., (2010), “Thermal and Mechanical Properties at High
Temperature of a Very High-Strength Durable Concrete”, Journal of Materials in Civil
Engineering - ASCE, pp. 545-555.

Bazant Z. P. and Kaplan M. F., (1996), Concrete at High Temperature: Material Properties and
Mathematical Models, Longman, London (United Kingdom).

Buchanan A. H., (2001), Structural design for fire safety, John Wiley & Sons, Chichester, New
York (USA), 448 pp.

Cruz C. R., (1966), “Elastic Properties of Concrete at High Temperatures”, Department Bulletin
191, Journal of the Portland Cement Association Research and Development Laboratories, 8
(1), pp. 37-45.

Daïan J. F., (2001), “Evaluation des proprieties de transfert dans les materiaux cimentaires”, Reveu
Francaise de Genie Civil, Union, pp. 179-202.

Diederichs U., Jumppanen U-M., Schneider U., (1995), “High temperature properties and spalling
behaviour of high strength concrete”, Proceedings of the Fourth Weimar Workshop on High
Performance Concrete: Material Properties and Design, held at Hochschule fur Architektur und
Bauwesen (HAB), Weimar, Germany, October 4th and 5th, 1995, pp. 219-236.

Ehm C., (1985), “Experimental Investigations of the Biaxial Strength and Deformation of Concrete
at High Temperatures”, Thesis, Technical University of Braunschweig (Germany).

EN 12390-2:2009, Testing fresh concrete – Slump test, European Committee For Standardization,
Brussels (Belgium).

EN 12390-3:2009, Testing hardened concrete – Compressive strength of test specimen, European


Committee For Standardization, Brussels (Belgium).

103
EN 12390-6:2009, Testing hardened concrete, Part 6: Tensile splitting strength of test specimens,
European Committee For Standardization, Brussels (Belgium).

EN 1991-1-2:2004, Eurocode 1: Actions on structures Part 1-2: General actions - Actions on


structures exposed to fire, European Committee for Standardization (CEN), Brussels, 2004.

EN 1992-1-2:2004. Eurocode 2: Design of concrete structures Part 1-2: General rules - Structural
fire design, European Committee for Standardization (CEN), Brussels (Belgium).

Feldman R. F. and Sereda P. J. (1968), “A Model for Hydrated Portland Cement Paste as Deduced
from sorption-length Change and Mechanical Properties”, Materiaux et Construction, 1 (6),
pp. 509-520.

Felicetti R. and Gambarova P. G., (1998), “On the Residual Tensile Properties of High
Performance Siliceous Concrete Exposed to High Temperature”, Special Volume in honour of
Z. P. Bazant’s 60th Anniversary, Prague, March 27-28, Ed. Hermes (Paris), pp. 167-186.

Felicetti R., Gambarova P. G., Rosati G. P., et al., (1996), “Residual mechanical properties of high-
strength concretes subjected to high-temperature cycles”, Proceedings, 4th International
Symposium on Utilization of High-Strength/High-Performance Concrete, Paris, France, 1996,
pp. 579-588.

fib Bulletin 38 (2007), Fire Design of Concrete Structures - Materials, Structures and Modelling,
International Federation for Structural Concrete (fib), Lausanne (Switzerland), April 2007.

Fu Y. F. and Li L. C., (2011), “Study on mechanism of thermal spalling in concrete exposed to


elevated temperatures”, Materials and Structures, 44, pp. 361-376.

Giuliani L, Crosti C and Gentili F. (2012), “Vulnerability of bridges to fire”, proc. of the 6th
International Conference on Bridge Maintenance, Safety And Management – IABMAS, Stresa
(Italy), p. 313.

Guo Y. C., Qian J. S. and Wang X., (2013), “Pore Structure and Influence of Recycled Aggregate”,
Mathematical Problems in Engineering, Research Article, pp. 1-7.

Hammer T. A., “HIGH-STRENGTH CONCRETE PHASE 3, Compressive strength and E-


modulus at elevated temperatures,” SP6 Fire Resistance, Report 6.1, SINTEF Structures and
Concrete, STF70 A95023, February, 1995.

104
Harmathy T. Z., (1970), “Thermal properties of concrete at elevated temperatures”, Journal of
Materials, 5 (1), pp. 47-74.

Heinfling G., Stabler J., Baker G. and Reynouard J.M., (1998), “ Effects of high temperatures on
the fracture energy of concrete”, Proceedings of the 3rd International Conference on Fracture
Mechanics of Concrete Structures – FRAMCOS 3.

Khaliq W., Kodur V., (2011), “Thermal and mechanical properties of fibre reinforced high
performance self-consolidating concrete at elevated temperatures”, Cement and Concrete
Research, 41(11), pp. 1112-1122.

Khoury G. A. (2000) “Effect of fire on concrete and concrete structures”, Progress in Structural
Engineering and Materials, 2, pp. 429-447.

Khoury G. A., (2008), “Polypropylene Fibres in Heated Concrete. Part 2: Pressure Relief
Mechanisms and Modelling Criteria”, Magazine of Concrete Research, 60 (3), pp. 189-204.

Lie T. T., Kodur V. K. R., (1996), “Thermal and mechanical properties of steel- fibre-reinforced
concrete at elevated temperatures”[J], Canadian Journal of Civil Engineering, 23(2), pp. 511-
517.

Lo Monte F., (2014), Reinforce Concrete in Fire: from Materials Behaviour to Spalling Sensitivity
and Structural Modelling, (Supervisors: P. G. Gambarova and R. Felicetti).

Marechal J. C., (1972), “Variations in the Modulus of Elasticity and Poisson’s Ratio with
Temperature”, Paper SP 34-27 in Concrete for Nuclear Reactors, American Concrete Institute
- ACI, Farmington Hills (Michigan), pp. 495-503.

Meftah F., (2005), Comportement des bétons en température, Master course of the Ecole Nationale
des Ponts et Chaussées, Champs sur Marne (France), 68 pp.

Mindeguia J. C., Pimienta P., Carrè H., and Borderie C. L., (2013), “Experimental Analysis of
Concrete Spalling due to Fire Exposure”, European Journal of Environmental and Civil
Engineering, 17 (6), pp. 453-466.

Noumowe A. N., Clastres P., Debicki G. and Costaz J. L., (1996), “Thermal Stresses and Water
Vapor Pressure of High Performance Concrete at High Temperature”, Proceedings of the 4th
International Symposium on Utilization of High-Strength/High-Performance Concrete, Paris

105
(France).

Picandet V., Khelidj A. and Bastian G., (2001), “Effect of axial compressive damage on gas
permeability of ordinary and high-performance concrete”, Cement and Concrete Research, 31,
pp. 1525-1532.

Pistol K., Weise F., Meng B. and Schneider U., (2011), “The Mode of Action of Polypropylene
Fibres in High Performance Concrete at High Temperatures”, Proceedings of the 2nd
International Workshop Concrete Spalling due to Fire Exposure, E.A.B. Koenderss and F.
Dehn (Eds), RILEM, Delft, October 5-7, 2011, pp. 289-296.

Reinhard Hinkelmann, (2004), Efficient Numerical Methods and Information –Processing


Techniques for Modeling Hydro – and Environmental Systems, Springer, Berlin (Germany),
320 pp.

Schrefler B. A. and Pesavento F. (2004), “Multiphase flow in deforming porous material”,


Computers and Geotechnics, 31, pp. 237-250.

Shin K. Y., Kim S. B., Kim J.H., Chung M. and Jung P. S., (2002), “Thermo-Physical Properties
and Transient Heat Transfer of Concrete at Elevated Temperatures”, Nuclear Engineering and
Design 212, pp. 233-241.

Zhang B. and Bicanic N., (2001), “Fracture energy of high performance concrete at temperatures
up to 450°C”, Proceedings of the 4th International Conference on Fracture Mechanics of
Concrete and Concrete Structures - FRAMCOS 4, 28th May-1st June, 2001, Cachan (France).

106
ACKNOWLEDGEMENTS

I would like to express my deepest appreciation to all those who provided me the possibility to
complete this report. A special gratitude I want to give to Prof. Roberto Felicetti, and the co-
supervisor Dr. Francesco Lo Monte. Their contribution in stimulating suggestions and
encouragement, helped me to coordinate my project especially in writing this report.
Furthermore I would also like to acknowledge with much appreciation the crucial role of Dr. C.
B. K. Rao, Professor, Engineering Structures department, without whose guidance it would not
have been possible for me to perform this thesis.
It was a great opportunity for me to pursue my thesis at Politecnico Di Milano, and I would like to
give the major credit to Dr. P. Ratheesh, Professor and coordinator of Engineering Structures
Division, Department of Civil Engineering, for his immense help in my selection for this
scholarship program. I take this opportunity to express my sincere thanks to Dr. DEVA PRATAP,
Professor and Head of Civil Engineering Department for his valuable, inspiring guidance,
encouragement, and excellent motivation.
I would also like to thank my colleague Xuezing wang, for helping me through this thesis in Italy
and K. Vignesh Kumar who is solely responsible for my successful application and selection for
ERASMUS MUNDUS Program at NIT Warangal. Last but not the least, I would like thank my
Parents for their unconditional love and support.

AANCHAL JASUJA
ROLL NO. 131501

107

Вам также может понравиться