Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Numerical model for analysis of offshore structures subjected to pool fires

Miguel R. Manco(1), Alexandre Landesmann(2), Murilo A. Vaz(1) and Julio C. Cyrino(1)

Received: xx xx/ Accepted: xx xx


Sociedade Brasileira de Engenharia Naval 2015

Abstract This paper introduces a numerical


model to assess the thermo-mechanical behavior of
offshore structures subjected to pool fires. The aim
of this model is to improve the representation of
fires in comparison to other existing models (e.g.
nominal fire curves or Zone Modeling). The
method proposed combines Two Zone Models
2ZM and a Modified Solid Flame Model MSFM
for calculating the convection and radiation
portions of the heat transfer process between the
fire and the structure. Thermal and mechanical
analyses are performed sequentially using the
commercial Finite Element Analysis (FEA)
software ABAQUS. To assess the structures
behavior, the transient temperature field resulting
from the preceding thermal analysis is used
together with pre-existing operating loads for the
mechanical analysis. A pool fire in a tank barge is
presented as a case study. Comparison of
temperature and strain fields obtained with the
method proposed to the results of other methods
for assessment of the casualty indicates that the
new model allows more realistic estimation of the
temperature fields produced by the fire with
relatively low computational costs.

Keywords
Pool Fire; Two Zone Model; Modified Solid Flame
Model.

Miguel R. Manco
mmanco1@oceanica.ufrj.br
Ocean Engineering Program(1), COPPE / Federal
University of Rio de Janeiro, Rio de Janeiro, Brazil
Civil Engineering Program(2), COPPE / Federal
University of Rio de Janeiro, Rio de Janeiro, Brazil

Introduction

The various floating systems employed by the


petroleum industry operate with equipment in high
temperatures and devices with active flames (e.g.
flares), as well as storing, transporting and/or distributing highly flammable materials in restricted
areas where people live and interact. Due to the
presence of such materials, the operators of these
facilities are required to follow strict fire protection
criteria, considered the most stringent among all
industrial facilities.
The eruption of a fire in these structures is one
of the worst scenarios possible, as the temperatures
produced by the fire degrade the mechanical
properties of steel, producing effects that
substantially alter the failure modes expected in the
original design (normal working temperature) [1].
For instance, such effects in offshore structures are
related to the emergence of plastic regions and
membrane behavior in deck plates, as indicated by
[2, 3, 4, 5]. One further aggravating factor for fires
in these structures is that external firefighting
support systems are usually very far away.
Currently, governments, classification societies
and numerous research centers have been
developing ever stricter criteria to avoid such
accidents. In this context, part 1.2 of Eurocode No.
1 [6] recommends several methods for analyzing
steel structures under fire. These rules mention
various methods for estimating compartment
temperature, such as nominal temperature vs. time
curves (e.g. for hydrocarbons - HC, ISO 834 and
external fire), Zone Models and advanced models
based on Computational Fluid Dynamics - CFD [7,
8, 9].
Zone models divide the compartment in control
volumes (zones) in which the temperature of gases
is supposed to be uniform. By applying the
principles of conservation of mass and energy and
the law of ideal gases, a set of ordinary differential
equations that describe the behavior of gas
temperatures and the interface between gases in
time may be derived [10, 11]. Nominal tempera-

ture vs. time curves may be understood as a


simplified single-zone case. CFD-based models are
the most elaborate, including models of turbulence, material burning rates, smoke propagation
and others, more reliably representing heat exchange processes [12, 13, 14], but require a large
amount of information and entail much higher
computation costs.
Furthermore, CFD-based models deal only with
the thermal part of the problem, so it is necessary
to feed its results to a Finite Element Analysis
(FEA) platform for the mechanical analysis, so yet
another difficulty is introduced when the grids of
the CFD and FEA models are not compatible [15].
Evidently, the better the representation of the thermal field, the more precise is the estimate of the
mechanical behavior of the structure due to the
fire.
The method proposed here aims to improve the
representation of pool fire phenomena, calculating
convection and radiation portions separately in a
single thermal analysis (ABAQUS [16]). The
convection portion considers gas temperature in
the compartment, previously estimated with the
Two Zone Model - 2ZM [17, 18], while the radiation portion makes use of flame temperature following the Modified Solid Flame Model (MSFM
[19]). The proposed model will be from now on
referred to as MSFM+2ZM. The models described
above are shown in Figure 1.
The numerical analyses are performed using
the free software OZone [20] and commercial code
ABAQUS [16], according to the finite element
method (FEM), taking into account the structural
and thermal effects resulting from the proposed
fire. Variation in thermal and mechanical properties of materials in case of high temperature condi-

Figure 1. Fire analysis models

tions are taken into account in the analysis, in


accordance with the applicable standard
recommendations, such as is the case of part 1.2 of
EC3 [21]. The fundamentals of the applied analysis
model are described in item 2 below. A case study
is proposed and briefly described in item 3 hereto.
An oil barge steel compartment, is submitted to a
fire scenario, caused by the burning of
hydrocarbons type pool fire [19], in order to
evaluate the thermo-structural behavior for differrent instances of the fire. The results obtained with
the numerical model are presented and critically
assessed in item 4 being compared with other
existing methods. The mains conclusions extracted
from the analyses performed are mentioned in item
5, indicating that this methodology allows
assessing qualitatively and quantitatively the
behavior of the structure, and can be used in the
improvement of current regulations related to the
safety of structures in the event of fire.

Analysis Methodology

A finite element model has been developed, which


includes a direct and rigorous consideration of
nonlinear physical and geometric effects on the
numerical formulation, allowing the estimate of the
structural instability collapse mode [2, 22, 23, 24].
The proposed analysis procedure begins by
evaluating the compartment arrangement, as well
as the selection of the fire scenario. Then, the
thermal analysis is performed, which purpose is to
determine the variation of the temperature in the
elements exposed to fire. The main numerical
formulation aspects of this stage are discussed in
item 2.1. The final stage of the procedure aims at

determining the structural behavior as a function of


the elapsed time of fire, in other words, depending
on the thermal conditions of fire exposure and
applied
external
loads
(mechanical).
Computational characteristics adopted in this final
stage of the numerical simulations are briefly
described in item 2.2.

2.1

Thermal Analysis

This analysis starts with the description of fuel


characteristics, i.e. fuel volume ( V [m]), pool free
surface area ( Afs [m]), burning rate per unit area (

m" [kg/m-s]), fuel density ( [kg/m]), combustion specific heat ( H c ,eff [kJ/kg]) and an empirical
constant relative to the specific fuel in question (
k [m1 ] ). The following step is the determination
of the fires radiation (MSFM) and convection
(2ZM) portions.
Three values must be known to define the
MSFM: flame equivalent diameter (D [m]), flame
height H[m] and flame equivalent temperature
[K]. Even if the pools free surface is not
circular, the flames equivalent diameter may be
estimated according to Eq. (1) [19]:

estimating gas temperatures in the compartment


and the height of the interphase between them.
This curve is calculated according to NFSCs [27]
recommendations, and is composed of three
sections. The first section corresponds to the fires
onset and development, and shows a parabolic
behavior (the duration of which depends on the
fuel type); the following section shows constant
behavior; finally, the curve drops linearly when
70% of the fuel mass is consumed. The radiation
portion is ignored in this model (the structures
emissivity is considered negligible), thus estimating only the convection portion. This leads to an
overestimation of gas temperatures, since no heat
exchange between the structure and the outside
environment through radiation is allowed in the
model.

D 2 Afs /

(1)
Flame height H[m] is calculated as shown in Eq.
(2) [25]:

m"
H

42
gD
D
a

0.61

(2)

where a is air density [kg/m] and g is the


acceleration of gravity [m/s].
Flame equivalent temperature
is obtained
using Eq. (3) [26]:

fl . . Eav .
fl .
4

fl ,eq 4

(3)
Figure 2. Stages in the thermal analysis

where
represents flame emissivity, is StefanBoltzmanns constant [
],
is environmental temperature [K],
[kW/m]
is the flames average emission power and is
atmospheric transmissibility.
The flames average emission power is calculated
following Eq. (4) [26]:

E
4 4 1
E
(4)
av
lum fl fl
lum soot

where
is the percentage of visible flame,
is the flames radiation temperature [K] and
[kW/m] is the emission power of soot.
The necessary input data for the software
OZone [20], which works with the 2ZM, includes the heat release rate (HRR) curve, used for

In the proposed FE model, developed in


ABAQUS [16], it is calculated the thermal field
in each time step due to radiation and
convection processes [28, 29, 30, 31]. The effect of
radiation is calculated by considering the
characteristics of the MSFM through the CAVITY
command. This option allows considering the
interaction between flame and structure through
the relative emissivity calculating the respective
shape factors. The heat transfer due to convection
is analyzed using the subroutine FILM, which
considers the distribution of the convection heat
transfer coefficient and gas temperature in contact
with the structure. A fixed convection heat transfer
coefficient is assumed and the gas temperature
distribution is obtained by the OZone program.

The flames equivalent temperature (


) is
modeled according to the behavior of the HRR
curve. Figure 2 shows a schematic representation
of the thermal analysis process.
The temperature dependent thermal properties
of the steel considered in the FE model follow the
recommendations given in part 1.2 to EC3 [21] and
are shown in Figure 3.

Apart from the thermal deformation imposed


on the structural model, variations in the mechanical properties of steel as a result of temperature,
as shown in Figure 4(a), are considered reduction:
of effective yield strength (
), slope of the linear
elastic range (
) and proportional limit (
),
obtained based on recommendations of [21].
The values of
,
and
at temperature
, are obtained by multiplying their values at
environment temperature (
,
and
, respectively) by the respective reduction
factors (
,
and
), shown in Figure 4(b). In this
figure it is also shown the behavior of the
coefficient of linear thermal expansion of the steel
as a function of temperature ( ) according to
[21].

3
Figure 3. Specific heat and thermal conductivity of
carbon steel as a function of the temp. [21].

2.2

Structural Analysis

Since the variation of the temperature field was


established in the previous analysis stage, the finite
element mesh used, i.e., the nodal coordinates, the
elements connectivity and the results for heat
fluxes are used in the simulation of structural
behavior under the postulated fire conditions. The
procedure is initialized by the application of
external loads, including the structural weight,
fluid action and other operational loads. At this
stage, deformations and their respective stresses,
corresponding to normal operating conditions of
the structure, can be seen. The variation of the
temperature field determined in the thermal
analysis is imposed to the structural model along
with other external loads applied.

Figure 4.

Case Study

This case study analyses a compartment 9.5 m


long, 20 m wide and 4m high (with a 2m x 0.8 m
entrance) simulating the engine room of a barge
tank, as in [3]. A fire is started at the center of this
compartment in a pool with a diameter of 1.5m
holding 40 gallons of fuel oil. To avoid the lateral
support conditions affecting the behavior of the
stiffener and to reduce computational costs, only
the part of the compartment situated between two
half frame spacings is modeled (Figure 5). Heat
transfer between the burning compartment and the
contiguous environments is done considering
constant room temperature of 20 C and a heat
exchange coefficient (including radiation and
convection) of
following the
recommendations
in
[6].
A
schematic
representation of the model considered is shown in
Figure 5.

(a) Stress-strain curve and (b) Reduction factor (RF) in relation to 20 C values and thermal expansion
coefficient for carbon steel at elevated temperatures [21].

Figure 5.

Model geometry and boundary conditions

The computational mesh used employs


elements type DS4 and S4R in thermal and
mechanical analyses, respectively. These elements
are composed of 4 nodes, with capacity for
developing nonlinear physical and geometrical
analyses. The solid flame represented only in the
thermal model is made up of elements type
DC3D8. The complete Newton-Raphson solution
process is adopted to update the matrices and the
linear solution of equations. The von Mises
criterion is adopted for determining the element
plastification. Figure 5 shows the finite element
model as well as the boundary conditions
employed in the analysis. The union between the
deck and transverse bulkhead has a denser mesh in
order to capture the plastics effects.
As boundary conditions the left cross-section in
Figure 5 is considered clamped, while on the right
cross-section the symmetry condition was
considered to reduce the computational cost.
Finally, in the side edges the boundary conditions
is considered that represent the continuity of the

compartment, as shown in Figure 5. The


structures own-weight and side pressure of 0.025
and 0.015 MPa in deck and bottom, respectively.
The case study, besides being assessed with the
proposed methodology, is studied using the
hydrocarbon curve - HC and the 2ZM [6].

Results

Structural failure is considered to occur in this


study when: (a) the material fractures (assumed to
happen when the materials ultimate deformation is
reached) or (b) the structure becomes unstable, as a
result of drastic increases in vertical deflections or
local buckling of any structural element [32]. As
stated by [33], more than one of these phenomena
may occur simultaneously in some cases. In this
type of casualty, it is also common practice to
present the results as a function of the plastic
deformation fields, since strains are related to
temperature fields which may be non-uniform.

Figure 6. (a) Considered temperatures in each fire model (HG: Hot Gases; CG: Cold Gases), (b) HRR curve and height of
separation of zones.

4.1

Temperature Domain

Figure 6 shows: (a) temperatures used in each


model, (b) the HRR curve considered and the
interphase height between zones in the 2ZM.
Figure 7 shows the temperature fields in each part
of the structure for the models studied in different
times after onset of the fire. As can be seen,
different thermal fields are obtained in each model
as a result of their different assumptions. HC fire is
the most conservative model, considering strictly
increasing temperatures reaching over 400C in
only 5 minutes, and consequently generates the
most severe results. The 2ZM, in turn, shows lower
temperatures due to the low thermal load considered. Finally, the results of the MSFM+2ZM model
show the highest temperature to be concentrated
above the flame, providing a more realistic
representation of the temperature field of the
localized fire.
In the Zone Models (HC and 2ZM), linearization of the radiation portion forces the adoption
of a single heat flow value in the whole control
volume for each time step. This condition implies
that elements in the control volume have almost

Figure 7.

constant temperature profiles despite their location.


Figure 8a shows temperature profiles in the deck
plate with the HC and 2ZM models in various time
instants, while temperature profiles in the stiffener
web are shown in Figure 8c for the HC model and
Figure 8d for the 2ZM model. It may be seen that
the temperature profile of the stiffener web shows
higher values, as a result of its greater massivity
factor (area exposed to the fire/ element volume),
producing conduction heat flows towards the plate
and the stiffener flange.
This heat flow is responsible for the higher
temperatures in the connections of the stiffener
web with the plate (
in Figure 8a) and
the stiffener flange (
in Figures 8c and
8d). The temperature profile in the stiffener flange
is not shown, as it keeps essentially uniform
temperatures along its width, with the same value
seen in its connection with the stiffener web
(
). Figure 10a shows that the deck
reaches thermal equilibrium after approximately 45
minutes in the HC model, afterwards keeping a
constant temperature profile. This equilibrium
occurs due to the model not considering the fires
extinction phase, as the hot gas temperature is

Temperature fields obtained for each considered fire model.

Figure 8. Temperature profile to plate deck (a)HC/2ZM, (b) MSFM+2ZM S1/S2 and stiffener web (c) HC, (d) 2ZM, (e)
MSFM+2ZM/S1, (f) MSFM+2ZM/S2.

maintained after 40 minutes (Figure 6a). The


plates temperature is lower because it releases
heat to the environment through its side not
exposed to the fire. Since the 2ZM considers the
extinction phase, a temperature reduction occurs in
all parts of the structure after 45 minutes (Figures
8a and 8d). As it considers the shape factor of
radiation, the MSFM + 2ZM shows non-uniform
heat flows, and therefore presents different
temperature profiles in each transversal section of
the compartment (Figure 7). Figures 8b,e,f show
the temperature profile for two transversal sections
of the deck, one directly above the flame (S1) and
the other halfway between the flame and the
transversal bulkhead (S2). Due to having the

highest shape factor (lowest distance from the


flame), the stiffener flange in section S1 shows the
highest temperatures, reaching values over 500C.
In section S2, however, the temperature does not
reach even half this value. This model also allows
representation of the so-called shading effect, in
which an element does not receive the radiation
emitted by the flame due to being covered by
another element. In this case, for example, the
connection of the plate with the stiffener is covered
by the stiffener flange (Figures 8b,e,f).
It needs to be emphasized that, the HC model
predicts very high temperatures, resulting
significant degradation of steel properties. The
2ZM produces moderate temperatures which do

not substantially affect the structures stiffness.


The MSFM+2ZM presents very localized high
temperatures. The connection bracket between
deck reinforcements and transversal bulkhead is
the most severely stressed element (in the
mechanical analysis), and does not reach
temperatures higher than 250 C in the 2ZM and
MSFM + 2ZM models.

4.2

Strength Domain

The state of stresses and strains is determined


based on the information regarding variation in the
temperature field. Thermal load is applied after
taking into account the structures own weight and
lateral pressures of 0.025 MPa in the deck and
0.015 MPa in the compartment bottom. Buckling
of the support bracket happens in all cases, the
differences between models being found only in
the time when it occurs.
In the HC fire model, degradation of the
mechanical properties of steel (due to high temperatures) combined with thermal strains produce

Figure 9.

considerable displacements and deformations in


the structure. Figure 9a displays the plastic
deformation field (PEEQ) in the structure in the
instant preceding its collapse, and shows that
ultimate deformation is localized.
Figure 10a shows displacements of three points
in the structure, found at: (A) the connection
between plate and reinforcement in the extremity
where symmetry is considered, (B) the union
bracket between deck and transversal bulkhead
reinforcements, and (C) halfway on the transversal
bulkhead, respectively.
Vertical displacement of point (A) evidences
the start of membrane behavior in the upper deck.
Transversal displacement of point (B), in turn,
evidences failure of the bracket due to local
buckling of this element. Finally, longitudinal
displacement of point (C) is caused by instability
of the transversal bulkhead, which induces global
collapse of the structure only approximately 23
minutes after the onset of the fire.
Figure 9b shows plastic deformations for the

Deformation field preceding instability in different models (a) HC pre-collapse, (b) HC Collapse, (c)
2ZM and (d) MSFM + 2ZM.

Figure 10. Displacement in points A, B and C for the (a) HC, (b) 2ZM and MSFM+2ZM

structures final configuration. It can be seen that


the ultimate deformation is exceeded in the deck
bracket and halfway through the vertical
reinforcement of the transversal bulkhead. Material
fracture is not considered here, as its determination
needs the establishment of criteria outside the
scope of this work. [24] proposes a model which
considers weld rupture due to fire. Figures 9c,d
show the structures plastic deformation field for
the 2ZM and MSFM+2ZM models, respectively. It
is evident from these images that ultimate
deformation is not exceeded in both cases. The
displacement of points (A) and (B) for 2ZM and
MSFM+2ZM is shown in Figure 10b. As
mentioned in the thermal analysis, the 2ZM
produces temperature profiles of low magnitude
and almost uniform throughout the deck,
generating slight, but generalized weakening in the
whole structure. In this case there is no global
instability of the structure, but the bracket buckles
locally (continuous black line).
Figures 9c,d show the structures plastic
deformation field for the 2ZM and MSFM+2ZM
models, respectively. It is evident from these
images that ultimate deformation is not exceeded
in both cases. The displacement of points (A) and
(B) for 2ZM and MSFM+2ZM is shown in Figure
10b. As mentioned in the thermal analysis, the
2ZM produces temperature profiles of low
magnitude and almost uniform throughout the
deck, generating slight, but generalized weakening
in the whole structure. In this case there is no
global instability of the structure, but the bracket
buckles locally (continuous black line).
Finally, with the MSFM+2ZM the thermal field
generates temperatures up to 580C in the stiffener
flange at the section above the flame, weakening
the deck in this region and thus overloading the
support bracket, which presents lower temperature
than in the previous case, but is subjected to
greater stress, as evidenced by the greater vertical
and transversal displacement of the deck and
bracket shown in Figure 10b (dashed lines).

The results presented evidence that HC fire is


the most severe model, predicting collapse of the
structure, while the zone model imposes less stress
on the structure, but does not reliably represent the
localized fire.

Conclusions

The
numerical-computer
methodology for
analyzing the behavior of steel structures under
pool fire conditions presented in this paper was
applied to evaluate the behavior of an oil barge
compartment submitted to a fire scenario. Despite
the idealized load conditions the thermomechanical behavior could be observed at different
times of the postulated fire, but it should be
mentioned that the results obtained are valid only
for the load and boundary conditions considered.
In a real situation, when the structure suffers the
action of waves and other loading conditions, a
different behavior can occur.
The results presented make evident that the
different models employed to obtain the transient
temperature field produce different results due to
their different assumptions. In this context, the
refinement of thermal models produce increasingly
realistic simulations, and the mechanical behavior
of the structure can thus be estimated more
reliably. On the other hand, more refined models
imply greater computational costs and increased
difficulty in defining input data, restricting their
applicability during the structures design stage.
The proposed thermal model achieved the goal
of improving representation of the thermal field for
localized pool fires with relatively low computational costs and implementation complexity.
Thermal field variation produced not only
quantitative but also qualitative changes in the
structures behavior, as shown in the results of the
case study.
This paper attempts to show the structures
global behavior and the advantages of the thermo-

mechanical model proposed in comparison to the


simpler models recommended by classification
societies.
The conclusions obtained from the numerical
simulations performed indicate that the method
presented here may be successfully applied to
assess the performance of offshore structures under
pool fire conditions.

Acknowledgment
The authors wish to express their gratitude to the
National Petroleum Agency of Brazil (ANP) and
National Council of Scientific and Technological
Development of Brazil (CNPq) for their support
for the development of this work.

References
[1] Yang, X.J., Gao R. (2004), Factors Affecting
the behavior of Steel Structures in Fire,
Proceedings of NASCC 2004, California,
EUA.
[2] Skallerud, B. and Amdahl J. (2002),
Nonlinear Analysis of Offshore Structures,
Research Studies Press Ltd., Baldock,
Herforshire, England.
[3] Manco, M.R., Vaz, M.A., Cyrino, J.C.,
Landesmann, A. (2013), Behavior of
stiffened panels exposed to fire, Proceedings
of IV MARSTRUCT, Espoo, Finland, pp. 101
- 108.
[4] Manco, M.R., Vaz, M.A., Cyrino, J.C.,
Landesmann, A. (2014), Analysis of oil
tanker deck under hydrocarbon fire,
International Journal of Modeling and
Simulation for the Petroleum Industry, Vol. 8,
No 2, pp. 17-24.
[5] Landesmann, A., Mendes, J.R., Ellwanger, G.
(2010), Numerical Model for the Analysis of
Offshore Structural Elements under Fire
Conditions, Proceedings of XXXIV Jornadas
Sudamericanas de Ingeniera Estructural, San
Juan, Argentina.
[6] European Committee for Standardization
(2004), EUROCODE No. 1: Actions on
Structures, Part 1-2: Actions on Structures
exposed to Fire, ENV 1991-1-2, British
Standards Institution, London, UK.
[7] ABS (2013), Accidental Load Analysis and
Design for Offshore Structures, American
Bureau of Shipping.
[8] ABS Guidance Notes on Alternative Design
and Arrangements for Safety 2004
[9] International Maritime Organization 2001
GUIDELINES ON ALTERNATIVE DESIGN
AND ARRANGEMENTS FOR FIRE
SAFETY

[10] Quintiere, J.G. (2006), Fundamentals of Fire


Phenomena, 1st Ed., John Wiley and Sons,
England.
[11] Chow WK, Meng L. Analysis of Key
Equations in a Two-layer Zone Model and
Application with Symbolic Mathematics in
Fire Safety Engineering, J. of Fire Sciences,
vol 22 2004 97-124.
[12] McGrattan K, Hostikka S, McDermott R,
Floyd J, Weinschenk C, Overholt K. NIST
Special Publication 1018-1 Fire Dynamics
Simulator (Version 6.2) Technical Reference
Guide, Volume 1: Mathematical Model, NIST
2015.
[13] McGrattan K, Hostikka S, McDermott, R,
Floyd J, Weinschenk C, Overholt K. NIST
Special Publication 1018-3, Fire Dynamics
Simulator (Version 6.2) Technical Reference
Guide, Volume 3: Validation. NIST 2015.
[14] McGrattan K, Hostikka S, McDermott R,
Floyd J, Weinschenk C, Overholt K. NIST
Special Publication 1019 Fire Dynamics
Simulator (Version 6.2) User's Guide. NIST
2015.
[15] Silva, J. (2013), Tridimensional interface
model to fire-thermomechanical analysis of
structures under fire conditions (portuguese),
D.Sc. thesis COPPE/UFRJ, Brazil.
[16] ABAQUS (2011), Hibbitt, Karlsson e
Sorensen, Version 6.11-3.
[17] Han, S.S. (2004), Review on key equations in
the Two-Layer Zone Model CFAST,
International
Journal
on
Engineering
Performance-Based Fire Codes, Vol. 6, No 4,
p.277-283.
[18] Zhang, C., Li, G. (2013), Modified One Zone
Model for fire resistance design of steel
structures, Advanced Steel Construction, Vol.
9, No. 4, pp. 282-297.
[19] SFPE. (2002). SFPE Handbook of Fire
Protection Engineering, 3th edition, Society
of Fire Protection Engineers.
[20] Ozone (2001), Universit of Lige, Version
2.2.
[21] European Committee for Standardization
(2004), Eurocode No. 3: Design of steel
structures, Part 1.2: Structural fire design,
ENV 1993-1-2, British Standards Institution,
London, UK.
[22] Sun, R., Huang Z., Burgess, I.W., (2012),
Progressive collapse analysis of steel
structures under fire conditions, Engineering
Structures Journal, Vol. 34, pp. 400-413.
[23] Hosseini, S. A., Zeinoddini, M., Saedi Daryan,
A., and Rahbari, M. (2014), Model fire tests
on a beam-to-leg connection in an offshore
platform topside, Fire and Materials Journal,
Vol. 38, pp. 529-549.

[24] Ufuah, E. (2012), Fundamental Behaviour of


Offshore Platform steel deck under running
poll
fires,
Ph.D.
thesis
University
Manchester.
[25] Thomas, P.H. (1962), The Size of Flames
from Natural Fires, Ninth Symposium
(International) on Combustion, Combustion
Institute, Pittsburgh, EUA, pp 844 859.
[26] Silva, F. (2014), Thermal performance-based
analysis of fuel storage tanks exposed to fire
(portuguese),
M.Sc.
dissertation
COPPE/UFRJ, Brazil.
[27] NFSC (2013), Natural Fire Safety Concept:
Full-scale tests, implementation in the
Eurocodes and development of a user-friendly
design tool ECSC Research, European
Commission, Directorate General for
Research and Innovation, Draft Final Report.
[28] Ozisik, M.N. (1994), Finite Difference
Methods in Heat Transfer, 1st Ed., CRC
Press, USA.

[29] Cook, R.D., Malkus, D.S., Plesha, M.E. and


Witt, R.J. (2002), Concepts and Applications
of Finite Element Analysis, 4th Ed., John
Wiley and Sons, New York.
[30] Lewis R.W., Nithiarasu, P. and Seetharamu,
K.N. (2004), Fundamentals of the Finite
Element Method for Heat and Fluid Flow,
John Wiley and Sons, England.
[31] Reddy, J.N. (2004), An Introduction to
Nonlinear Finite Element Analysis, Oxford
University Press, New York.
[32] Zhang, C., Li, G., Usmani, A. (2013),
Simulating the behavior of restrained steel
beams to flame impingement from localizedfires, Journal of Constructional Steel
Research, Vol. 83, pp. 156-165.
[33]
Pay-Zaforteza, I., Garlock, M.E. (2012),
A numerical investigation on the fire response
of a steel girder bridge, Journal of
Constructional Steel Research, Vol. 75, pp. 93103.

You might also like