Você está na página 1de 7

ACI MATERIALS JOURNAL

TECHNICAL PAPER

Title no. 96-M51

Fire Behavior of High-Performance Concrete Made with


Silica Fume at Various Moisture Contents
by Sammy Yin Nin Chan, Gai-Fei Peng, and Mike Anson
Behavior of high-performance silica fume concrete with various moisture contents in the ISO standard fire was investigated.
Spalling test results from 100-mm cube specimens revealed that
moisture content and strength are the two main factors governing explosive thermal spalling of concrete. Moisture content
has a dominant influence on spalling. The dependence of spalling on moisture content and strength confirmed the vapor
pressure mechanism of spalling. Spatial distribution of rebound
hammer test results on postfire concrete slabs correlated with
the temperature field within the slabs. Conventional methods
do not properly describe the residual mechanical properties of
concrete subjected to fire. A new method is needed that considers the chemical and physical changes inside concrete caused
by elevated temperature.
Keywords: fire resistance; high-performance concrete; moisture content;
silica fume; spalling; strength.

INTRODUCTION
With development and application of high-performance concrete (HPC), understanding of its behavior when subjected to fire
is needed to insure its safe application.1,2 HPC exhibits superior
performance in many aspects, e.g. possesses high strength, durability, and workability.3 However, HPCs inferior thermal behavior, with its dense microstructure, has not been fully
understood. Poor fire behavior is a threat to HPCs practical application in many types of engineering structures, including offshore platforms, tunnels, and high-rise buildings. Fire behavior
is the main obstacle to HPCs further development. In many
countries, research projects are either planned or in process on
this specific topic.4-7
HPC fire behavior results from its susceptibility to explosive
spalling when subjected to fire or high temperature conditions.
The explosive thermal spalling is a catastrophic failure of concrete that generally occurs above 300 C and is characterized by
the material explosively breaking into pieces, often without advance notice. Such explosive spalling needs in-depth investigations because it happens in an unpredictable manner. This
explosive spalling may cause loss of the concrete cover over reinforcing steel bars. Direct exposure of reinforcing bars to high
temperatures reduces the structural integrity of the reinforced
concrete structure.
The explosive thermal spalling mechanism remains unidentified.1,8 Among several viewpoints, two have become dominant:
the vapor pressure build-up mechanism9 and the thermal stress
mechanism.7 The vapor pressure build-up mechanism occurs
because the dense hardened cement paste prevents moisture
from escaping under high temperatures, thus causing a considerable pressure build-up resulting in spalling. The thermal stress
mechanism occurs because the exposure to fire produces a thermal gradient within the concrete, which causes internal stresses
and the initiation of spalling. A combination of these two mechanisms is also possible. 10 In fact, the material factors of strength

ACI Materials Journal/May-June 1999

grade, moisture content, and aggregate type; the high temperature conditions including rate of temperature rise and maximum
temperature reached; and other factors, such as specimen configuration and reinforcement distribution, may all simultaneously be relevant to the occurrence of explosive spalling.
With respect to their relevant quantitative effects, however,
there is a gap in our knowledge due to the lack of systematic experimental data and a subsequently established and verified analytical model.
An understanding of the explosive spalling of HPC requires
data obtained from testing HPC material and structural elements. Some experimental data on strength and behavior of
HPC in fires have been reported in the past decade.10-12 Similar
to normal strength concrete (NSC), HPC loses strength depending on its exposure to fire. The high temperature of fire causes
decomposition of hardened cement paste made from portland
cement, the main cementitious agent. Inner microstructural
cracking also occurs. Several researchers indicated that the occurrence of spalling is mainly dependent on the pore pressure
build-up.4,5,13 This has been confirmed by a report providing a
quantitative measurement of the factors influencing explosive
thermal spalling.9 More quantitative measurement data on spalling is needed, and so is experimental data on the effects of high
temperature on other properties of HPC.
Apart from explosive spalling, fire damage to HPC involves
strength reduction and crack development. In a severe fire, a
concrete element is subjected to a transient thermal gradient attributed to variable elevated temperatures. Thus, strength loss at
different points inside the element will vary due to the temperature gradient. Scanning electronic microscope (SEM), x-ray diffraction (XRD), and nondestructive evaluation (NDE)
techniques may be useful for the fire damage evaluation of an
element.
In this paper, 100-mm concrete cubes of five strengths, four of
which were HPC, were exposed to the ISO standard fire to explore the relationships between strength grade, moisture content,
and the frequency of explosive thermal spalling. Two slabs of reinforced HPC of different strengths, but the same as two from the
four HPC mixes, were used to determine their respective fire behavior. At various points on the two slabs, after exposure to fire,
rebound hammer tests were conducted to determine the distribution of residual compressive strength as a means of identifying the
fire damage distribution.

ACI Materials Journal , V. 96, No. 3, May-June 1999.


Received July 20, 1998, and reviewed under Institute publication policies. Copyright 1999, American Concrete Institute. All rights reserved, including the making
of copies unless permission is obtained from the copyright proprietors. Pertinent discussion will be published in the March-April 2000 ACI Materials Journal if received
by December 1, 1999.

405

Sammy Yin Nin Chan is a visiting professor at Southeast University and Shenzhen
University. He received his Honors and doctorate from the University of Dundee,
Scotland.
Gai-Fei Peng is a PhD candidate in the Department of Civil and Structural Engineering at Hong Kong Polytechnic University. He received his BS and MS from Tongji
University in Shanghai, and Tsinghua University in Beijing, respectively. His research
interests include concrete technology and new materials development.
Mike Anson is Dean of the Faculty of Construction and Land Use at Hong Kong
Polytechnic University. He is a graduate of Oxford and London Universities. His

Table 1Mix proportions and compressive


strengths of concrete tested
Batch quantities,

Fig. 1Standard temperature-time curve recommended in ISO


834.

kg/m3

Concrete mix

NC-40

HPC-60

HPC-70

Ordinary portland
cement

HPC-110 HPC-120

333

369

431

518

565

Silica fume

40

44

52

62

68

Coarse aggregate,
20 mm

835

835

835

835

835

Coarse aggregate,
10 mm

420

420

420

420

420

Sand

440

440

440

440

440

Water

224

207

183

151

133

Water-cementitious
materials ratio

0.60

0.50

0.38

0.26

0.21

Compressive
strength at 28 days,
MPa

47

65

78

115

128

Fig. 2Reinforcement arrangement of fire test slabs.

RESEARCH SIGNIFICANCE
Fire behavior of HPC has attracted research because of its
susceptibility to explosive spalling in a fire. Understanding such
explosive thermal spalling requires experimental data on both
HPC material and HPC structural elements. This paper presents
an experimental investigation conducted with HPC material and
elements with different moisture contents, subjected to the ISO
standard fire.

EXPERIMENTAL DETAILS
Spalling test
The spalling test was conducted on 100-mm concrete cube
specimens of five strengths, designated NC-40, HPC-60, HPC70, HPC-110, and HPC-120. Mix proportions and compressive
strengths are given in Table 1. The coarse aggregate was
crushed granite. The five mixes all had a slump of approximately 200 mm. Superplasticizer dosages varied for the four HPC
mixes. No superplasticizer was used in the NC-40 mix.
The spalling test was conducted in an oil-burning furnace using
three cube specimens with the specified moisture m values. The
furnace time-temperature curve complied with the ISO 834 standard curve shown in Fig. 1. Explosive spalling was monitored in
the enclosed furnace by recording the loud bang during heating
and by inspecting the specimens after cooling.

Establishment of moisture content in concrete


Moisture content is defined as the ratio of evaporable water
within a specimen at the time of the test to its original evaporable water when initially taken from the curing tank. This definition can be expressed by Eq. (1)
m = ( W t W 0 ) 100 percent
where
m = moisture content of specimen, percentage;
W 0 = total evaporable water within the specimen when

406

(1)

initially removed from the curing water, i.e. the


evaporable water in a fully saturated specimen, in kg;
Wt = the evaporable water within the specimen at the time
of the test, in kg.
If evaporable water of the saturated specimen is r times the
weight of the saturated specimen, and r has been determined using representative specimens from the same concrete batch,
then m can be expressed by Eq. (2)
m = [ 1 G (G 0 r ) ] 100 percent

(2)

where
G = G0 - Gt, in kg;
G0 = the weight of the saturated specimen initially taken
from the curing water, in kg;
Gt = the actual weight of the specimen at the time of the
test, in kg.
After casting, the cube specimens were cured in water for 90
days. At 90 days, they were taken from the water and weighed
continually at an interval of a few hours until their weights were
compatible with the moisture contents shown in Table 2. Once
the desired weights were reached, the specimens were sealed
with polyethylene film. This sealing maintained the moisture
contents until the spalling test took place.

Fire resistance test on slabs


The fire resistance test was conducted on two slabs prepared
from the same batch of concrete as the spalling test Specimen
HPC-60 and HPC-110. They were named Slab HPC-60 and
HPC-110, respectively. The reinforcement of the slabs is shown
in Fig. 2, and the fire test configuration is given in Fig. 3.
During the fire test, the time-temperature curve of the furnace
complied with the ISO standard curve. Midspan deflections and
temperatures at locations inside the slabs were monitored continuously throughout the test. The furnace temperature was
measured using a thermocouple placed below the slabs, as
shown in Fig. 3. Fig. 4 shows the thermocouple locations within
a 1/4 segment of each slab.
After demolding, the slabs were subjected to moist curing for
28 days. After 28 days, they were exposed to air, allowing the
inner moisture to escape freely until the fire test at 90 days. The

ACI Materials Journal/May-June 1999

research interests include concrete technology and construction productivity.

407

ACI Materials Journal/May-June 1999

Fig. 3Fire test configuration.


Fig. 5Relationship between spalling frequency, strength, and
moisture content.

Fig. 4Thermocouple locations with quarter segment of slab.


moisture contents for Slab HPC-60 and HPC-110 were 60 and
62 percent, respectively, at the time of the fire test.

Post-fire evaluation of slabs


When the slabs cooled to room temperature after the fire test,
crack patterns were measured, and residual mechanical properties were determined at various slab surface locations using the
rebound hammer test.

RESULTS AND DISCUSSION


Spalling dependence on moisture content and
strength
The relationship between spalling frequency, strength, and
moisture content is shown in Fig. 5. The frequency was calculated as the average from the three cubes for each set. These results
show that spalling depended on both the strength and moisture
content of concrete. For concrete strengths less than 60 MPa,
there was no spalling even with the concrete fully saturated. For
concrete strengths greater than 60 MPa, the higher the moisture
content, the greater the spalling frequency when the moisture
content was above a threshold value mt. The value of mt for HPC70, HPC-110, and HPC-120 was found to be 88, 63, and 63 percent, respectively, as shown in Fig. 5.
For concretes HPC-70, HPC-110, and HPC-120, a difference
in the moisture content caused a difference in spalling frequency,
while all the other specimen conditions remained unchanged.
This revealed that moisture content had a dominant influence on
spalling frequency for concrete strengths greater than 60 MPa.
Spalling generally occurred within the temperature range of
480 to 500 C, which differs from previously reported values of
320 to 360 C 12 and 715 C.8 This discrepancy may be due to the
difference in pore structure of hardened cement paste and other
specimen configurations if the time-temperature curves are the
same. The reason spalling of concretes of different strengths in
this experiment was concentrated in the narrow temperature
range from 480 to 500 C is of interest and needs more research.

Fire resistance test


During the fire test of the slabs, no explosive thermal spalling
took place. This is reasonable because the 100-mm cube spal-

ACI Materials Journal/May-June 1999

Fig. 6Furnace and Slab HPC-110 temperatures (refer to Fig.


4 for locations).

Table 2Moisture content of concrete specimens


Specimen configuration
Moisture content, percent
Specimen no. for each moisture content value

100-mm cube
100, 88, 75, 63, 50, 25, 0
3

ling tests showed HPC-60 would not spall under the fire regime
adopted, and HPC-110 would spall only when the moisture content was higher than 63 percent. The moisture content of Slab
HPC-110 was 62 percent when tested.
The furnace and slab temperatures are shown in Fig. 6. The fire
resistance rating14 is the time when either of the following two
conditions is reached: 1) the temperature on the unexposed slab
surface reaches 220 C; or 2) the midspan slab deflection reaches
1/20 of the clear span. The midspan slab deflections during the
fire test are shown in Fig. 7. The fire resistance ratings for Slab
HPC-60 and HPC-110 were 110 and 123 min, respectively. At
these times, the temperatures on the unexposed slab surfaces were
104 and 135 C, respectively. Therefore, Slab HPC-110 had a
greater fire resistance rating than Slab HPC-60.

Crack pattern of slabs


After the fire tests, slab crack patterns were measured and recorded. Cracking of the two slabs appeared similar. Fig. 8 indicates a dense crack network in which three main cracks
propagated across the entire width of the slabs. Other crack
lengths were generally 30 to 50 mm. The three major crack
widths of the two slabs ranged from 0.7 to 1.5 mm. No cracks
were observed on the slab top surfaces. The nearly identical
crack patterns of both slabs suggest that surface crack patterns
cannot be used as an indicator of the degree of fire damage to
concrete.

408

Fig. 9Schematic diagram of temperature field in slabs.

Fig. 7Slab midspan deflections during fire test.

Fig. 10Residual compressive strength (from rebound hammer


tests) and schematic diagram of its distribution at different
depths on sides of slabs.
has a spatial distribution correlating with the temperature field.
Conventional methods, such as the compressive strength test
and the fire resistance test, seem inappropriate for the evaluation
of residual mechanical properties of these structural elements. A
new method, which takes into consideration both the chemical
and physical changes caused by the temperature field inside
concrete, is needed to evaluate the fire damage in terms of spalling, strength decrease, and cracking.

CONCLUSIONS
Fig. 8Schematic diagram of slab crack pattern (after fire

Rebound hammer tests


The rebound hammer measurements (assumed to be fairly
uniform before fire exposure) would be expected to decrease
relative to the thermal history at the different points. The measured temperatures for the points at different depths showed that
the slabs experienced temperatures as described schematically
in Fig. 9. As the maximum temperature on the slab top surfaces
was 250 C about 1 hr after they reached their respective fire resistance rating, it may be assumed that the mechanical property
variations on the top surface were small.11 Thus, the research
was focused on properties at different depths.
The rebound hammer results on the slab sides at different
depths within three zones have been converted into equivalent
compressive strengths, as shown in Fig. 10. The conversion is
according to the calibration relationship in Chinese Standard
JGJ/T 23-92.15 Measurement of residual compressive strengths

409

Based on the experimental investigation with the ISO standard fire, the following conclusions are drawn.
1. Moisture content and strength are the two main factors governing thermally induced explosive spalling of concrete. If the
strength of concrete is below a certain value, generally no spalling will occur, even at a high moisture content level. In this investigation, this strength value was 60 MPa. When the concrete
strength exceeds 60 MPa, the higher the moisture content, the
greater the spalling probability, as long as the moisture content
is greater than a threshold value mt . For HPC-70, HPC-110 and
HPC-120, mt values were found to be 88, 63, and 63 percent, respectively. Thus, moisture content has a dominant influence on
spalling.
2. The effects of moisture content and strength on spalling
confirmed the vapor pressure build-up hypothesis as the mechanism for spalling.
3. That no spalling occurred during the fire test of Slab HPC60 and HPC-110 is consistent with the test results obtained with
100-mm cube specimens.

ACI Materials Journal/May-June 1999

4. Slab HPC-110 had a higher fire resistance rating than Slab


HPC-60, but the crack patterns on both slab surfaces are almost
the same. Thus, the surface crack patterns may not be a good indicator for assessing fire damage to concrete.
5. The rebound hammer test on fire-tested slabs revealed that
measured residual mechanical properties possessed a spatial
distribution within the tested slabs that correlated with its temperature distribution. Routine strength determination and fire
resistance test methods are too simple to describe the residual
mechanical properties of concrete subjected to fire. A new
method involving the chemical and physical analyses at elevated temperatures is needed in evaluating fire damage of concrete
elements.
6. The reason for the narrow temperature range of 480 to 500
C within which most spallings occurred is not yet determined.
Further research in this specific subject is desirable.

REFERENCES
1. Lin, W.-M.; Lin, T. D.; and Powers-Couche, L. J., Microstructures of
Fire-Damaged Concrete, ACI Materials Journal, V. 93, No. 3, May-June
1996, pp. 199-205.
2. Jahren, P. A., Fire Resistance of High-Strength/Dense Concrete with
Particular Reference to the Use of Condensed Silica FumeA Review,
Proceedings of the Third International Conference on Fly Ash, Silica
Fume, Slag, and Natural Pozzolans in Concrete, SP-114, American Concrete Institute, Farmington Hills, Mich., 1989, pp. 1013-1049.
3. Mehta, P. K., and Aitcin, P.-C., Principles Underlying Production of
High-Performance Concrete, Cement, Concrete, and Aggregates, V. 12,
Winter 1990, pp. 70-78.
4. Jensen, J. J.; Hammer, T. A.; and Hansen, P. A., Fire Resistance and
Residual Strength of HPC Exposed to Hydrocarbon Fire, Proceedings of

410

International Workshop on Fire Performance of High-Strength Concrete,


NIST SP 919, NIST, Gaithersburg, Feb. 13-14, 1997, pp. 59-68.
5. Anderberg, V., Spalling Phenomena of HPC and OC, Proceedings
of International Workshop on Fire Performance of High-Strength Concrete,
NIST SP 919, NIST, Gaithersburg, Feb. 13-14, 1997, pp. 69-74.
6. Kodur, V. K. R., Studies on the Fire Resistance of High-Strength
Concrete at the National Research Council of Canada, Proceedings of
International Workshop on Fire Performance of High-Strength Concrete,
NIST SP 919, NIST, Gaithersburg, Feb. 13-14, 1997, pp. 75-86.
7. Ahmed, G. N., and Hurst, J. P., An Analytical Approach for Investigating the Causes of Spalling of High-Strength Concrete at High Temperatures, Proceedings of International Workshop on Fire Performance of
High-Strength Concrete, NIST SP 919, NIST, Gaithersburg, Feb. 13-14,
1997, pp. 95-108.
8. Sanjavan, G., and Stocks, L. J., Spalling of High-Strength Silica
Fume Concrete in Fire, ACI Materials Journal, V. 90, No. 2, Mar.-Apr.
1993, pp. 170-173.
9. Hisaka, M., Physical Properties of High-Strength and High-Quality
Concrete Using High-Range Water-Reducing AgentsPart 2, Journal of
Cement and Concrete (Tokyo), No. 549, Nov. 1992, pp. 9-18. (in Japanese)
10. Bazant, Z. P., and Kaplan, M. F., Concrete at High Temperatures:
Material Properties and Mathematical Models, Longman Group Ltd.,
England, 1996, pp. 5-196.
11. Chan, Y. N.; Peng, G. F.; and Chan, K. W., Comparison between
High-Strength Concrete and Normal Strength Concrete Subjected to High
Temperature, RILEM Materials and Structures (Paris), V. 29, Dec. 1996,
pp. 616-619.
12. Castillo, C., and Durrani, A. J., Effect of Transient High Temperature on High-Strength Concrete, ACI Materials Journal , V. 87, No. 1,
Jan.-Feb. 1990, pp. 47-53.
13. Consolazio, G. R.; McVay, M. C.; and Rish, J. W., Measurement
and Prediction of Pore Pressure in Cement Mortar Subjected to Elevated
Temperature, Proceedings of International Workshop on Fire Performance of High-Strength Concrete, NIST SP 919, NIST, Gaithersburg, Feb.
13-14, 1997, pp. 125-148.

ACI Materials Journal/January-February 1999

14. Chinese Standard GBJ 45-82, Code of High-Rise Civil Building


Design for Fire Protection, Beijing, June 1, 1983, p. 47. (in Chinese)
15. Chinese Standard JGJ/T 23-92, Technical Specification for Inspection of
Concrete Compressive Strength Method, Beijing, Oct. 1, 1993, p. 32. (in Chinese)

ACI Materials Journal/January-February 1999

411

Você também pode gostar