You are on page 1of 16

Special Issue: Fire Safe Indoor and Built Environment

Indoor and Built


Environment
Indoor and Built Environment
2023, Vol. 32(1) 9–24

A review of thermal properties ! The Author(s) 2021


Article reuse guidelines:

of timber and char at elevated sagepub.com/journals-


permissions
DOI: 10.1177/
temperatures 1420326X211035557
journals.sagepub.com/home/ibe

Long Shi1 and Michael Y. L. Chew2

Abstract
Timber is one of the most frequently adopted combustible materials in the built environment. The
thermal properties are the determining factors for assessing the fire risk in a building. The main thermal
properties of timber and their char are reviewed, especially those temperature-dependent and moisture-
dependent properties, including kinetic properties, ignition properties, thermal conductivity, specific
heat capacity, effective heat of combustion and thermal diffusivity. The study has collected and sum-
marized various thermal properties data and empirical models of hardwood and softwood with different
mass percentages in cellulose, hemicellulose and lignin, as temperature increases. The average ignition
temperature and effective heat of combustion of softwood are about 12.9% and 9.5% higher than those
of hardwood, respectively. From most of the previous models, the thermal conductivity of timber char
increases as temperature rises. Cellulose with a high density shows a higher thermal conductivity, but
its impacts on the specific heat capacity are limited. Models to predict the main thermal properties of the
hardwood, softwood and char are recommended. The collected data, together with those empirical
models, can provide useful data resources and tools for the related fire risk assessments.

Keywords
Timber, Thermal properties, Temperature dependence, Moisture dependence, Fire safety, Elevated
temperatures
Accepted: 10 July 2021

Introduction
of internal temperature under effects of conduction,
As a green solution, the use of timber has gained pop- convection and radiation, is a dynamic one. Pyrolysis
ularity in the building sector during the past decade.1 processes trigger when the temperature arrives at a crit-
However, its massive adoption are accompanied by fire ical level, while gas volatiles are released, and a part of
risk. The concerns of timber fire or timber building fire the virgin timber gradually changes to char. The char
have been raised by the public2 and policymakers, espe- could provide a determining impact on the follow-up
cially for tall buildings.3 fire behaviours as it acts as a thermal barrier to the
Due to its combustible characteristics, its fire risk
should be seriously assessed according to the local con-
ditions before applying it in the buildings. Timber is 1
Civil and Infrastructure Engineering Discipline, School of
quite a unique combustible material that its many ther- Engineering, RMIT University, Melbourne, Australia
mal properties could affect the overall assessment and 2
Department of Building, National University of Singapore,
fire behaviours. These could include the ignition prop- Singapore
erties, thermal conductivity, specific heat capacity,
Corresponding author:
effective heat of combustion and thermal diffusivity. Michael Y. L. Chew, Department of Building, National
Under a building fire, the elevation of the surface tem- University of Singapore, Singapore 117566.
perature of a timber element, together with the change Email: bdgchewm@nus.edu.sg
10 Indoor and Built Environment 32(1)

external heat sources. Ignition happens when those gas Fire-resistance of timber presents a period of time
volatiles are ignited as it rises to a critical temperature.4 that the timber product fails in one or more ways under
Flame heat flux then adds a heat source to the timber fire conditions, such as loadbearing capacity, insula-
surface that could accelerate the burning processes. tion, and integrity.19 The fire resistance level is a func-
Thermal properties of timber at elevated tempera- tion of those thermal properties, which vary under
tures are critical inputs for accurate numerical model- elevated temperatures. Many studies have been under-
ling. Many pyrolysis models have been developed taken to assist the amendment of the related building
previously to predict the pyrolysis behaviours of codes and design standards on timber products and
timber slabs under external radiation, such as buildings.20,21
Gpyro,5 ThermaKin6 and the authors’ FiresCone.7 In Temperature-dependent thermal properties of
these models, thermal properties at elevated tempera- timber such as kinetic parameters, ignition tempera-
tures are critical to the modelling accuracy. The related ture, thermal conductivity, specific heat capacity, effec-
sensitivity analysis8 and optimization algorithm (e.g. tive heat of combustion and thermal diffusivity based
genetic algorithm9) are adopted to determine the on experimental tests are reviewed. This study aims to:
appropriate values for timbers. However, there is still
a lack of data that can provide a good data source for • Address the main thermal properties of various
pyrolysis and fire modelling, especially for various timber types under elevated temperatures;
timber types. • Collect and analyse the related experimental data of
Many factors could affect the thermal properties of various types of timber under elevated temperatures;
timber, such as moisture content and application types and
(e.g. cross-laminated timber10) The determination of • Recommend the temperature-dependent thermal
the thermal properties for timber is much more chal- properties of timber for potential future modelling
lenging than those isotropic construction materials, inputs and uses.
such as steel,11 concrete12 and polymers.13 For exam-
ple, as an anisotropic material, timber’s thermal con-
ductivity to the fibre orientation is generally twice of The classification of timber
that perpendicular to fibre.14 Some of its characteristics Timber usually contains the hardwood and softwood.
even complicate the situation, such as internal mois- Hardwood shows pores or vessel elements that exist
ture15 and pyrolysis behaviours under a high- among parenchyma cells and fibre. Softwood contains
temperature environment.7,16,17 Previous research overlapping tracheid, connected by bordered pit aper-
found that the ignition time of those wet timber tures, parenchyma cells and, in some cases, resin
could be 65.5% longer than those of dry wood under canals.22 Table 1 shows a typical classification of com-
the same level of radiation.15 monly used timber.
Research gaps on timber’s thermal properties still The main components of timber are cellulose, hemi-
exist, and the related collection and summary are celluloses and lignin. Table 2 shows the chemical com-
urgently needed for the fire safety area. Although tim- ponents of dry wood in mass percentage.23 In Table 2,
bers’ thermal properties are well-known, one of the limited difference can be seen for the mass fraction of
most critical aspects for accurate fire modelling is cellulose in softwood and hardwood. However, soft-
those temperature-dependent thermal properties. This wood shows a relatively lower mass fraction of hemi-
is because the temperature environment could greatly celluloses and a higher lignin fraction. The influences of
affect its thermal properties. Temperature could largely
affect the thermal properties of timber. They are Table 1. The classification of commonly used timbers.
temperature-dependent and change with elevated tem-
peratures. Specific experimental instruments are needed Type Items
to gain these temperature-dependent data, which Softwood Balsam fir, Cypress, Douglas fir, Eastern
makes the data acquirement more complex when com- white cedar, Hemlock, Jack pine,
pared to those thermal properties under ambient tem- Ponderosa pine, Red pine, Redwood,
perature. To tackle this, some simulations have Southern pine, Spruce, Sugar pine,
attempted to simply use temperature-independent ther- Western red cedar, Western larch,
mal properties or temperature-dependent thermal White fir, White pine, etc.
properties obtained from other species of timber as Hardwood Balsam poplar, Basswood, Beech, Birch,
input parameters. This has resulted in a dramatic Blackbutt, Cherry, Large-tooth aspen,
decrease in modelling accuracy.18 Hence, the gathering Mahogany, Maple, Oak, Sweet gum,
Trembling aspen, Walnut, White ash,
of as many viable temperature-dependent data as pos-
Willow, etc.
sible is necessary.
Shi and Chew 11

Table 2. Chemical components of dry wood in mass


percentage.23

Types Cellulose Hemicellulose Lignin

Softwood 40–44 20–32 25–35


Hardwood 40–44 23–40 18–25

different mass fractions of components on their ther-


mal properties are analysed in the following content.

Kinetic properties of timbers


The pyrolysis process of timber has its own character-
istics. The thermal degradation of timber above 100 C Figure 1. Two-steps reaction scheme for thermal
can be divided into four phases:24 decomposition of timber.

• Within 100–200 C, timber becomes dehydrated, the order of reaction; E is the activation energy, J/mol;
which then generates water vapour and non- and R is the universal gas constant, J/mol  K.
combustible gases and liquids, including CO2, However, practical situations at elevated tempera-
formic acid, acetic acid and H2O. Due to prolonged tures are much complicated than the one-step reaction.
exposures at higher temperatures, timber can For example, some parts of virgin timber change into
become charred. tar initially, and then the tar changes into char and gas
• Within 200–300 C, significant pyrolysis then starts volatiles when the temperature increases. This two-
for some timber components; and besides those steps reaction for timber’s thermal decomposition is
gases and liquids, large amounts of CO are released shown in Figure 1.
during this phase from the pyrolysis layer, together This two-steps reaction can also be expressed by the
with the other gas products25 such as CO2, CH4 and Arrhenius law. The virgin timber changes into gas vol-
lower amounts of H2 and C2 hydrocarbons; atiles, tar and char simultaneously during pyrolysis
• The third phase happens between 300 and 450 C process. So, the mass-loss rate of timber is dependent
accompanied by the vigorous production of flamma- on these three processes, which can be given by
ble volatiles. Significant depolymerization of cellu- equation (2)
lose takes place within the range 300–350 C. Also,
aliphatic side chains are split off from the aromatic @ms
¼ km0 ¼ ðkg1 þ kt þ kc1 Þm0 (2)
rings in the lignin at around 300 C. Finally, the @t
carbon–carbon linkage between lignin structural
units is cleaved within 370–400 C; and where kg1, kt and kc1 are the production reaction rate of
• When the temperature goes over 450 C, the left gas, tar and char, respectively, 1/s.
timber residue is char, which could undergo further For all five processes, the related mass loss can be
degradation when the temperature keeps rising. described by equation (3)

Virgin timber changes to char after water evapora- @mi


¼ ki mi (3)
tion and the burning of fuel inside. This process is usu- @t
ally considered a one-step reaction. The virgin timber
changes to char directly during drying and combustion where i represents g1, t, c1, g2, c2 for the five processes,
processes. An nth order Arrhenius law, equation (1), respectively.
can describe the mass reduction in this process The rate of reactions, k, is governed by the
  Arrhenius Law in equation (4)
 
@ms ms  mf n E  
¼ Am0 exp  (1)  n
@t m0 RT ms  mf Ei
ki ¼ Ai exp  (4)
m0 RT
where ms, m0, and mf are the dynamic mass, initial
mass, and final mass of timber, respectively, kg; t is To obtain A and E in the above Arrhenius equa-
the time, s; A is the pre-exponential constant, 1/s; n is tions, some thermal analysis methods were used,
12 Indoor and Built Environment 32(1)

namely thermogravimetric analysis (TGA) and differ- Table 4 shows a summary of timber components’
ential scanning calorimetry (DSC).26 At the early stage experimental kinetic data, including cellulose, hemicel-
of thermal analysis, the isothermal method has usually lulose and lignin. Hemicellulose shows the highest
been adopted.27 However, accurate isothermal condi- activation energy, and it is followed by cellulose and
tions are difficult to be maintained and controlled lignin, respectively. The activation energy of hemicellu-
because of the limited experimental conditions. lose is higher than 180 kJ/mol, cellulose is between
Furthermore, one non-isothermal curve contains the 71 and 213 kJ/mol, and lignin is between 25 and
same amount of information as several isothermal 145 kJ/mol. It means the reactions of hemicellulose
curves. Thus, the non-isothermal method28 replaces and cellulose are more complex than those of lignin.
the isothermal method later under the rapid develop- Table 4 shows that lignin’s pre-exponential
ment of experimental conditions. Table 3 shows a sum- frequency factor is less than 1 when the temperature
mary of experimental kinetic data for timber. is higher than 617 K.

Table 3. A summary of experimental kinetic data for timber.

Experimental
References Timber condition Temp. range Kinetic expression

Samolada and vasalos29 Firwood Isothermal 673–773 K kg1 þ kt ¼ 2.40  104exp(94/RT)


Wagenaar et al.30 Pine TGA 553–673 K kc1 ¼ 3.05  107exp(125/RT)
773–873 K kt ¼ 9.28  109exp(149/RT)
kg1 ¼ 1.11  1011exp(177/RT)
k ¼ 1.4  1010exp(150/RT)
Chan et al.31 Pine – – kc1 ¼ 1.08  107exp(121/RT)
kt ¼ 2  108exp(133/RT)
kg1 ¼ 1.3  108exp(140/RT)
Liu and Fan32 Nanmu TGA 435–568 K k ¼ 5.53  108exp(116.57/RT)
568–623 K k ¼ 1.99  1024exp(290.53/RT)
623–728 K k ¼ 5.91  105exp(109.37/RT)
728–787 K k ¼ 2.30  1021exp(320.37/RT)
Liu and Fan32 Paulownia TGA 455–571 K k ¼ 7.64  1011exp(149.0/RT)
571–631 K k ¼ 1.44  1018exp(215.21/RT)
631–671 K k ¼ 3.90  1020exp(287.32/RT)
671–775 K k ¼ 7.42  1048exp(645.17/RT)
Thurner and Mann33 Oak sawdust Isothermal 573–673 K kg1 ¼ 8.607  105exp(88.6/RT)
kt ¼ 2.475  108exp(112.7/RT)
kc1 ¼ 4.426  107exp(106.5/RT)
kg1 þ kt ¼ 1.039  108exp(106.5/RT)
k ¼ 1.481  108exp(106.5/RT)
Di Blasi and Beech – 573–708 K kg1 þ kt ¼ 1.5  1010exp(149/RT)
Branca34 kc1 ¼ 3.3  106exp(112/RT)
kg1 ¼ 4.4  109exp(153/RT)
kt ¼ 1.1  1010exp(148/RT)
Gorton and Knight34 Hardwood Isothermal 677–822 K k ¼ 1.483  106exp(89.52/RT)
Ward and Brashlaw34 Wild cherry Isothermal 538–593 K k ¼ 1.19  1012exp(173.7/RT)
Nunn et al.35 Sweet gum Non-iso 600–1400 K kg1 þ kt ¼ 3.338  105exp(69/RT)
Font et al.34 Almond shell – 733–878 K kc1 ¼ 2.89  103exp(73/RT)
kt ¼ 5.85  106exp(119/RT)
kg1 ¼ 1.52  107exp(139/RT)
k ¼ 1.885  106exp(92.1/RT)
Swann et al.36 Maple plywood TGA 353–823 K k ¼ (2.0  0.33)1013exp[(170  23)/RT]
DSC k ¼ (1.33  0.83)1011exp[(146  3)/RT]
Liu and Fan32 Willow TGA 446–595 K k ¼ 2.54  108exp(118.73/RT)
595–658 K k ¼ 2.53  1024exp(296.93/RT)
658–699 K k ¼ 8.13  1014exp(226.56/RT)
699–768 K k ¼ 4.41  1051exp(711.36/RT)
Note: This table shows only the first-order pyrolysis reaction.
Shi and Chew 13

Table 4. A summary of experimental kinetic data of timber components.

References Component Experimental conditions Temp. range Kinetic constants

Moghtaderi 37
Cellulose – 600–850 K k ¼ 1.2  106exp(100.5/RT)
– 450–700 K k ¼ 6.79  103exp(71/RT)
– 520–1270 K k ¼ 6.79  109exp(139.6/RT)
– 520–1270 K k ¼ 3.9  1011exp(166.4/RT)
Dollimore 38 Cellulose Isothermal 481–513 K k ¼ 1.26  1014exp(150/RT)
Ozturk and Merklin 39
Cellulose Non-isothermal – k ¼ 7.4  107exp(130.524.7/RT)
Varhegyi et al. 40 Hemicellulose DTG (10 K/min)a – k ¼ 1.0  1017exp(187/RT)
DTG (80 K/min)a – k ¼ 1.0  1013exp(148/RT)
Cellulose DTG (10 K/min)a – k ¼ 2.0  1015exp(213/RT)
DTG (80 K/min)a – k ¼ 5.0  1013exp(195/RT)
Chan and Krieger 41
Lignin Non-isothermal 433–953 K k ¼ 7.83exp(25.12/RT)
Non-isothermal 683–2163 K k ¼ 1.5  101exp(30.56/RT)
Tang 41
Lignin – 553–617 K k ¼ 1.65  104exp(87.9/RT)
– 617–708 K k ¼ 9.3  102exp(37.7/RT)
Hirata 41
Lignin – 553–573 K k ¼ 7.17  1010exp(145.7/RT)
a
Those values in brackets are the heating rates during experimental tests.

Thermodynamic properties Effective heat of combustion


Effective heat of combustion is the released energy
Ignition properties
when a material goes through burning processes.
Critical heat flux (qcr) represents the lowest heat flux Although the effective heat of combustion may fluctu-
per unit area capable of initiating a combustion reac- ate as temperature rises, researchers would prefer to
tion on a material. It reflects the difficulty of igniting a consider that they are temperature independent. In
material. Ignition temperature (Tig) is the surface tem- 1917, Thornton46 found that for many organic liquids
perature of a material when it is ignited. In fire model- and gases, under complete combustion, a more or less
ling, the material is considered to be ignited when its fixed net amount of heat is released per unit mass of
surface reaches the ignition temperature. consumed oxygen. Later, Huggett47 confirmed this and
Table 5 shows the ignition temperature and critical obtained a constant for this, namely 13.1 kJ/g. This is
heat flux of typical timber.15,23,42–44 The ignition tem- the basic theory for measuring the heat release rate
peratures of hardwoods (e.g. red oak, Victoria ash, under standard conditions by ISO 9705.48,49
blackbutt and maple) are within 423–613 K, where Table 7 shows the effective heat of combustion of
the ignition temperatures of those softwoods (e.g. typical timber.45,50,51 From Table 7, the values for soft-
Chinese scholartree, redwood, western red cedar, pine woods are between 11.3 and 14.2 kJ/g, and for the
and Douglas fir) are between 531 and 755 K. The aver- hardwood they are between 10.6 and 13.2 kJ/g. The
age ignition temperature of those softwoods is about average effective heat of combustion of softwoods is
12.9% higher than those of hardwoods. In the studies 9.5% higher than those of hardwoods.
undertaken by Wang et al.,42 different altitudes repre-
sent the environment with various oxygen concentra-
Thermal conductivity
tions and pressure, so the study addressed the impacts
of oxygen concentration and pressure on the ignition The thermal conductivity (k) of material reflects its
properties. ability of heat conduction. Figure 2 shows the thermal
The different ignition behaviours of softwood and conductivity of timber dependent on temperature. In
the hardwood may be due to their chemical compo- 1970, Wenzl52 addressed a relationship between ther-
nents. Table 6 shows the temperature ranges of the mal conductivity and temperature. The empirical
three timber components decompose and release gas model gained by him shows an increasing trend of ther-
volatiles.45 They show quite different thermal degrada- mal conductivity along with the temperature. Later in
tion characteristics. Among these three components, 1986, Parker53 indicated that the thermal conductivity
lignin has a higher temperature range for degradation. of timber is related to temperature also mass retention
Softwood has a higher fraction of lignin, which fraction.
explains why those softwoods’ ignition temperatures For a better comparison between these two models,
are higher than those of hardwoods. a mass retention fraction was assumed in Parker’s
14 Indoor and Built Environment 32(1)

Table 5. Ignition temperature (Tig) and critical heat flux (qcr) of typical timber.

Type Timber Conditions Tig (K)d qcr (kW/m2)c

Softwood Chinese scholar tree42 a


20.8 kW/m2, altitude 3650 mb, in house apparatus 755e –
26.2 kW/m2, altitude 3650 m 704e –
31.6 kW/m2, altitude 3650 m 694e –
37 kW/m2, altitude 3650 m 645e –
Chinese scholar tree42 20.8 kW/m2, altitude 50 m, in house apparatus No ignitione –
26.2 kW/m2, altitude 50 m No ignitione –
31.6 kW/m2, altitude 50 m 683e –
37 kW/m2, altitude 50 m 584e –
Redwood43 21 kW/m2, cone calorimeter 723f –
30 kW/m2 663f –
42 kW/m2 653f –
54 kW/m2 703f –
Western red cedar23 Lateral ignition and flame spread test (LIFT) 613–649f 11.9–15.7
Redwood23 LIFT 623–646f 12.6–15.5
Pine23 LIFT 608–639f 11.6–15.0
Douglas fir23 LIFT 609–643f 11.6–15.2
Redwood43,44 Along grain, cone calorimeter 648f 15.5
Across grain 577f 5.9
Douglas fir23,44 Along grain, cone calorimeter 657f 16.0
Across grain 531f 8.4
Pine15 Dry, along grain, cone calorimeter 587–752e –
Wet, along grain 580–831e –
Hardwood Red oak43 27 kW/m2, cone calorimeter 568f –
52 kW/m2 623f –
76 kW/m2 553f –
Victoria ash23 LIFT 570–613f 9.3–13.1
Blackbutt23 LIFT 559–610f 8.7–12.8
Red oak44 Along grain, cone calorimeter 677f 10.8
Across grain 548f 9.2
Maple44 Along grain, cone calorimeter 627f 13.9
Across grain 423 3.8
Beech15 Dry, along grain, cone calorimeter 540–714e –
Wet, along grain 544–777e –
Cherry15 Dry, along grain, cone calorimeter 562–799e –
Wet, along grain 608–817e –
Oak15 Dry, along grain, cone calorimeter 586–729e –
Wet, along grain 627–785e –
Maple15 Dry, along grain, cone calorimeter 588–738e –
Wet, along grain 606–764e –
Ash15 Dry, along grain, cone calorimeter 555–712e –
Wet, along grain 538–761e –
a
It represents the external heat flux.
b
Altitude means the place where experiments took place.
c
It represents the critical heat flux under the piloted ignition conditions.
d
The surface temperature of the sample measured at ignition was considered as the ignition temperature.
e
The ignition temperatures were obtained under autoignition conditions.
f
The ignition temperatures were obtained under piloted ignition.15

model to obtain the same value with Wenzl’s prediction Table 6. Timber components release volatiles over different
temperature ranges. 45
at 250 K. As shown in this Figure 2, the increasing rate
of thermal conductivity from Wenzl’s model is bigger Components Temperature ranges (K)
than that of Parker’s model. The data in Eurocode 554
seem to be non-linear. These values are between Parker’s Cellulose 513–623
Hemicellulose 473–533
and Wenzl’s predictions except those values between
Lignin 553–773
523 and 873 K. Those values from Gupta et al.55
Shi and Chew 15

Table 7. The effective heat of combustion of common timbers.

Type Timber h (kJ/g) Type Timber h (kJ/g)

Softwood Japanese red pine50 13.1 Hardwood Japanese walnut50 10.6


Douglas fir51 12.4 Zelkova50 11.0
Sugar pine51 11.5 Japanese oak50 10.6
Western red cedar45 13.1 Red oak51 13.2
Redwood45 12.6 Victorian ash45 11.7
Radiate pine45 11.9 Blackbutt45 10.6
Douglas fir45 12.0 Poulownia50 12.3
Japan cedar50 13.0 Japanese beech50 11.3
Hiba abor-vitae50 14.2 50

Japanese larch50 11.3


Poulownia50 12.3

Figure 2. Thermal conductivity of timber dependent on temperature.52–56 Softwood in Gupta et al. 55 is composed of balsam
fir, white spruce, and black spruce.

are smaller than those from Parker’s model. From et al.’s59 models are very close. However, for pines
Suuberg et al.,56 three lines in the figure reflect the from Alves and Figueiredo’s model,59 thermal conduc-
thermal conductivity of cellulose with different initial tivity is higher than the other two types of timber, when
densities. Cellulose with a high density shows a larger they have the same moisture content. Figure 3 suggests
thermal conductivity. The detailed empirical models that the thermal conductivity of softwood may increase
mentioned above are also shown in Table 8. faster than those of the hardwood when moisture con-
Figure 3 shows the thermal conductivity of timber tent rises.
dependent on moisture content. Similarly, for a better Figure 4 shows the temperature-dependent thermal
comparison, the mass retention fraction in Parker’s conductivity of various types of char. As can be seen,
model57 and the density of Dry wood in Tenwolde the thermal conductivity of timber char increases as
et al.’s model59 are assumed to gain the same thermal temperature rises, observed from most of the models’
conductivity under zero moisture content. As can be predictions. Differently in Koufopanos et al.’s model,60
seen, values gained by Park’s57 and Tenwolde the thermal conductivity decreases when the
16 Indoor and Built Environment 32(1)

Table 8. A summary of empirical models to predict the thermal conductivity of timber and timber char.

References Material Empirical model (W/mK) Effective region

Wenzl52 Timber k ¼ 0.13 þ 3  10-4 (T-273) –


Parker53 Douglas fir k ¼ 0.124(0.35 þ 0.65Z)[1 þ 6.8(T-T0) 104] –
Suuberg et al.56 Cellulose k¼0.0222 þ 3.13  104T a
q ¼ 0.928 g/cm3, T ¼ [459, 562] K
k¼0.0162 þ 2.29  104T a
q ¼ 0.678 g/cm3, T ¼ [459, 562] K
k¼0.011 þ 1.55  104T a
q ¼ 0.458 g/cm3, T ¼ [459, 562] K
Parker57 Douglas fir k ¼ 0.0273 þ 0.2S(1 þ 0.02M) S ¼ [0.3, 0.8], M ¼ [0, 40]%
Alves and Figueiredo58 Pine k ¼ 0.166 þ 0.00396M –
Tenwolde et al.59 Timber k ¼ 0.001qD(0.1941 þ 0.00406M)þ0.01864 M ¼ [0, 40]%
Koufopanos et al.60 Timber char k ¼ 0.08-(T-273) 104 –
Alves and Figueiredo58 Pine char k ¼ 0.091 þ 8.2  105T –
Suuberg et al.56 Cellulose char k¼0.0993 þ 4.29  104T-1.89  107T2 a
q ¼ 0.41 g/cm3, T ¼ [447, 533] K
k¼0.117 þ 5.03  104T-2.65  107T2 a
q ¼ 0.52 g/cm3, T ¼ [422, 530] K
k¼0.0612 þ 2.67  104T-1.80  107T2 a
q ¼ 0.18 g/cm3, T ¼ [423, 522] K
k¼0.235 þ 1.02  103T-7.77  107T2 a
q ¼ 0.33 g/cm3, T ¼ [423, 548] K
a
The density is obtained under ambient temperature. T is the temperature, and all the temperatures in this table are described by Kelvin; Z is
the mass retention fraction of timber, T0 is ambient temperature, K; S is the specific gravity of timber; and M is the moisture content of timber.

Figure 3. Thermal conductivity of timber dependent on moisture.57–59 Timber in Tenwolde et al.’s59 represents plywood,
fiberboard, and several types of particleboard.

Specific heat capacity


temperature rises. The trend for the cellulose char is not
the same as those of cellulose.56 The 0.701 g/cm3 cellu- Figure 5 shows the temperature-dependent specific
lose char shows the highest thermal conductivity when heat capacity of timber. From Figure 5, values from
they are at the same temperature. Then it goes to 0.934, Wenzl,52 Koch,61 Volberhr,62 Gupta et al.,55 Eurocode
0.944 and 0.682 g/cm3, respectively. Thermal conduc- 554 and Dunlap65 are very close. These experimental
tivities from 0.934 and 0.944 g/cm3 cellulose char are tests took within different temperature regions, as
very close, which overlap in this figure. The detailed shown in Table 9. Gronli and Melaaen’ model63
empirical models for timber char are listed in Table 8 shows a relatively lower increasing rate than those of
as well. the other models. However, it shows a relatively higher
Shi and Chew 17

Figure 4. Thermal conductivity of timber char dependent on temperature.55,56,58,60 The densities from Suuberg et al.56 are the
initial cellulose char densities under the ambient environment. The two lines with initial cellulose char densities of 0.934 and
0.944 g/cm3 overlap in this figure, with different applied temperature ranges. Softwood in Gupta et al.55 is composed of balsam
fir, white spruce and black spruce.

Figure 5. Temperature-dependent specific heat capacity of timber.52,54,55,61–64 Softwood in Gupta et al.55 is composed of
balsam fir, white spruce, and black spruce. Dry wood in Gronli and Melaaen63 includes Norwegian birch, pine and spruce.
18 Indoor and Built Environment 32(1)

Table 9. A summary of empirical models to predict the specific heat capacity of timber and timber char.

References Material Empirical model (kJ/kgK) Effective region

Wenzl52 Timber Cp ¼ 1.112 þ 0.00485(T-273) –


Koch61 Spruce pine Cp ¼ 0.0366 þ 0.0042T T ¼ [333, 413] K
Gupta et al.55 Softwood Cp ¼ 0.52477 þ 0.00546T T ¼ [313, 413] K
Dunlap65 Dry wood Cp ¼ 0.212 þ 0.00485T T ¼ [293, 379] K
Volbehr62 Timber Cp ¼ 0.297 þ 0.00506T M ¼ [0, 27]%
Gronli and Melaaen63 Dry wood Cp ¼ 1.5 þ 103T –
Perry and Green66 Timber char Cp ¼ 1.0032 þ 0.00209(T-273) –
Parker57 Douglas fir Cp ¼ 1.43 þ 3.55  104T-7.32  104/T2 –
Stull67 Timber char Cp ¼ 1.39 þ 0.00036T T ¼ [700, 2000] K
Gronli and Melaaen63 Timber char Cp ¼ 4.2  10-4þ2.09  106T þ 6.85  106T2 –
Gupta et al.55 Softwood char Cp ¼ 0.79528 þ 5.98  103T-3.8  106T2 –
Suuberg et al.56 Cellulose char Cp ¼ 0.960 þ 2.02  103T-33505/T2 a
q ¼ 0.41 g/cm3
T ¼ [423, 823] K
Cp ¼ 0.922 þ 2.04  103T-40036/T2 a
q ¼ 0.52 g/cm3
T ¼ [398, 548] K
Cp ¼ 1.19 þ 1.49  103T-49626/T2 a
q ¼ 0.18 g/cm3
T ¼ [423, 823] K
Cp ¼ 2.08  1.11  104T-77136/T2 a
q ¼ 0.33 g/cm3
T ¼ [423, 548] K
a
The density is obtained under ambient temperature. T is the temperature, and all the temperatures in this table are in Kelvin.

value when the temperature is less than 450 K. Ragland where CD p is the specific heat capacity of dry wood,
and Aerts64 obtained four groups of data with moisture KJ/kg.K; Cwp is the specific heat capacity of water,
contents of 0, 5%, 12% and 20%. The higher moisture 4.186 kJ/kgK; and M is the moisture content of
content of timber shows the biggest specific heat capac- timber, %.
ity. Also, the data with zero moisture content are close
to the data from Wenzl,52 Koch,61 Volberhr,62 Gupta
et al.,55 Eurocode 5 54 and Dunlap.65 Thermal diffusivity
Figure 6 shows the specific heat capacity of timber
char dependent on temperature. Data from Parker57 Thermal diffusivity indicates the heat flow rate through
and Stull67 are very close. Stull67 addressed a relationship a material. It is usually used together with the other
between the specific heat capacity and temperature within thermal properties to quantify the cooling time of an
700–2000 K. Effective region was not mentioned in object. A bigger thermal diffusivity of a material indi-
Parker’s study.57 For Gronli and Melaaen’s model,63 cates a lower temperature gradient inside the material
the specific heat capacity increases drastically as temper- under the same heat conditions. The thermal diffusivity
ature rises. Also, Suubery et al.56 indicated that the influ- is usually defined in equation (6)
ence of density of cellulose on the specific heat capacity is
limited. The related empirical models for timber char are k
D¼ (6)
listed in Table 9. Perry and Green’s model66 has a similar qCp
increasing rate with Suubery et al.’s model56 and Gupta’s
model.55 Those predicted specific heat capacities from where D is the thermal diffusivity of a material, m2/s;
Parker’s model57 seem to be much stable than others. and q is the density of a material, kg/m3.
Moisture content also has an influence on the spe- Figure 7 shows the thermal diffusivity of cellulose
cific heat capacity. Tenwolde et al.59 took a large
dependent on temperature. Suuberg et al.56 found that
number of experimental tests on various species of
the cellulose’s thermal diffusivity keeps constant within
timber, and developed a model for the specific heat
389–562 K. The thermal diffusivity of cellulose char
capacity with different temperatures and moisture con-
increases as temperature rises. The thermal diffusivity
tents, as shown in equation (5)
is the highest when the density of the cellulose char
CDp þ 0:01MCp
W
Cp ¼ is 0.33 g/cm3. Then it is followed by 0.41, 0.18 and
1 þ 0:01M 0.52 g/cm3 cellulose char, respectively. A summary of
þ ð0:0002355T  0:0001326M  0:06191ÞM
empirical models for cellulose and its char are listed in
(5) Table 10.56
Shi and Chew 19

Figure 6. Specific heat capacity of timber char dependent on temperature.55–57,63,66,67 Softwood in Gupta et al.55 is composed
of balsam fir, white spruce, and black spruce. Dry wood in Gronli and Melaaen63 includes Norwegian birch, pine, and spruce.

Figure 7. Thermal diffusivity of timber dependent on temperature.


20 Indoor and Built Environment 32(1)

Table 10. A summary of empirical models to predict the thermal diffusivity of cellulose and its char.

Material Empirical models (m2/s) Effective regiona

Cellulose D ¼ 0.806  0.174 q ¼ 0.928 g/cm3, T ¼ [389, 562] K


D ¼ 0.806  0.174 q ¼ 0.678 g/cm3, T ¼ [389, 562] K
D ¼ 0.806  0.174 q ¼ 0.458 g/cm3, T ¼ [389, 562] K
Cellulose char D ¼ 1.57 þ 7.95  103T-5.97  106T2 q ¼ 0.41 g/cm3, T ¼ [447, 533] K
D ¼ 1.51 þ 7.50  103T-5.99  106T2 q ¼ 0.52 g/cm3, T ¼ [422, 530] K
D ¼ 2.27 þ 1.09  102T-9.21  106T2 q ¼ 0.18 g/cm3, T ¼ [410, 522] K
D ¼ 4.47 þ 2.02  102T-1.68  105T2 q ¼ 0.33 g/cm3, T ¼ [403, 580] K
a
The density is obtained under ambient temperature.

Table 11. Temperature independent thermal conductivity of timber.

Type Timber k (W/mK) Type Timber k (W/mK)

Softwood Norway spruce 0.10 Hardwood Birch 0.350


Spruce 0.109 Red oak 0.44
Douglas fir 0.11 Maple 0.35
Redwood 0.19 Birch 0.214–0.25
Douglas fir 0.23
Fir 0.138

Recommendation and thermal conductivities of softwoods, in which the mass


implementation retention fraction can be considered as 1, namely virgin
timber. For hardwoods, thermal conductivity predic-
There are thousands of species of timber available tions are recommended to be based on Wenzl’s
worldwide. It is impossible to place each one of them model.52
in experiments to measure their temperature-dependent For timber char’s thermal conductivity, the differ-
thermal properties. The above sections list the main ence between softwood and hardwood is limited. In
thermal properties of typical timber in our daily lives. Figure 4, except Koufopanos et al.’s model,60 the ther-
However, until now, there is still a huge gap in timber’s mal conductivity of char rises under a higher tempera-
thermal properties. It is significant to summarize these ture. Those data from Gupta et al.55 are close to those
data to guide their appropriate uses in the practical fire from Alves and Figueiredo’s model.59 So Alves and
modelling or relevant implementation. Figueiredo’s model59 is recommended for predicting
From Table 7, the range of the effective heat of the thermal conductivity of timber char under elevated
combustion for softwood is 11.3–14.2 kJ/g, and for temperatures.
hardwood is 10.6–13.2 kJ/g. The average value for soft- Figure 5 shows the specific heat capacity of timber.
wood in Table 7 is 12.49 kJ/g, and for hardwood is From Figure 5, the data from Wenzl,52 Kouch,61
11.41 kJ/g. Therefore, if the timber species is not avail- Gupta et al.,55 Dunlap65 and Volbehr62 are close to
able in Table 7, the effective heat of combustion for each other. Experimental data for dry timber from
softwood and hardwood is recommended as 12.49 and Ragland and Aerts64 are close to the other models’
11.41 kJ/g, respectively. data. So, this group of models could be adopted to
Table 11 shows a summary of timber’s thermal con- predict the specific heat capacity of dry wood. In addi-
ductivity, which are temperature independent.44,56 The tion, moisture content is known to have a great impact
range of softwood is 0.10–0.23 W/mK, and for hard- on the specific heat capacity. Therefore, after combin-
wood it is 0.214–0.44 W/mK. The thermal conductiv- ing Wenzl’s model52 (in Table 9) and Tenwolde et al.’s
ities of hardwoods are averagely 134.7% higher than model59 (equation (5)), equation (7) is gained
those of softwoods. Most of the obtained thermal con-
ductivities from the literature are between the predic- 0:00485T þ 0:04186M0:21205
Cp ¼
tions of Parker’s model53 and Wenzl’s model.52 The 1 þ 0:01M (7)
increasing rate from Wenzl’s model52 is higher than þ ð0:0002355T  0:0001326M  0:0691ÞM
that of Parker’s model.53 As softwoods’ thermal con-
ductivities are normally lower than those of the hard- In equation (7), the specific heat capacity of water is
woods, Parker’s model53 is recommended to predict the considered as temperature independent. Figure 8 shows
Shi and Chew 21

Figure 8. A comparison of specific heat capacity between equation (7) and experimental data for timber.

Table 12. Recommended thermal properties of softwood, hardwood and char.

Thermal property Timber Items

Kinetic properties Softwood See Table 3


Hardwood See Table 3
Cellulose See Table 4
Hemicellulose See Table 4
Lignin See Table 4
Ignition properties Softwood See Table 5
Hardwood See Table 5
Effective heat of combustion(kJ/g) Softwood If not seen in Table 7, recommend 12.75.
Hardwood If not seen in Table 7, recommend 11.9.
Thermal conductivity (W/mK) Hardwood k ¼ 0.13 þ 3  104 (T-273)
Softwood k ¼ 0.124 þ 0.8432  104(T-293)
Char k ¼ 0.091 þ 8.2  105T
0:00485T þ 0:04186M0:21205
Specific heat capacity (kJ/kgK) Timber Cp ¼
1 þ 0:01M
þ ð0:0002355T  0:0001326M  0:0691ÞM
Char Cp ¼1.0032 þ 0.00209(T-273)
Thermal diffusivity (m2/s) Timber D ¼ qCk p
Char
Note: T is the temperature, and all the temperatures in this table are in Kelvin; and the above suggestions are for the conditions when the
specific timber species are not found in the above tables.

that equation (7) and experimental data from Ragland Similarly, for char’s specific heat capacity, it is
and Aerts64 are in good agreement. So, equation (7) is assumed that there is no difference between softwood
recommended for predicting the specific heat capacity and hardwood. Figure 6 shows that the data from
of timber dependent on both temperature and moisture Gronli’s model68 increase drastically as temperature
content. rises. Data from Suubery et al.56 and Gupta et al.55
22 Indoor and Built Environment 32(1)

increase smoothly along with the temperature. Data Authors’ contribution


from Perry and Green’s model66 are between them. Long Shi: Writing – original draft, methodology, formal
So, Perry and Green’s model66 is recommended for analysis; Michael Yit Lin Chew: Conceptualization, writing
predicting timber char’s specific heat capacity for – review and editing, supervision, project administration.
both the softwood and hardwood.
Few studies are found to measure the thermal diffu- Declaration of conflicting interests
sivity for timber and char directly. Maybe it is because The author(s) declared no potential conflicts of interest with
the thermal diffusivity can be calculated by equation respect to the research, authorship, and/or publication of this
(6). So, equation (6) is recommended for predicting the article.
thermal diffusivity of timber and timber char. Those
above recommended thermal properties of softwood, Funding
hardwood and char are summarized in Table 12. The author(s) received no financial support for the research,
authorship, and/or publication of this article.

ORCID iD
Long Shi https://orcid.org/0000-0003-0763-7293
Conclusions
Thermal properties (especially those temperature- References
dependent or moisture-dependent) of softwood, hard- 1. Szasz A, Hlavicka V, Lubloy E and Biro A. Numerical
wood and char are reviewed and summarized in this modelling of the fire resistance of double sheared steel-to-
study, including kinetic properties, ignition properties, timber connections. J Build Eng 2021; 37: 102150.
2. Barber D. Tall timber buildings: what’s next in fire
thermal conductivity, specific heat capacity, effective
safety? Fire Technol 2015; 51: 1279–1284.
heat of combustion and thermal diffusivity. The related 3. Australasian Fire and Emergency Service Authorities
data and empirical models are also collected. The fol- Council Limited. Fire safety principles for massive
lowing main conclusions can be obtained: timber building systems. Melbourne: Australasian Fire
and Emergency Service Authorities Council, 2018.
• Timber usually includes hardwood and softwood. 4. Shi L and Chew MYL. Experimental study of woods
Both show a similar mass percentage in cellulose, under external heat flux by autoignition: ignition time
while softwood shows a relatively higher and lower and mass loss rate. J Therm Anal Calorim 2013; 111:
mass percentage, respectively, in lignin and hemicel- 1399–1407.
5. Fawaz M, Lautenberger C and Bond TC. Prediction of
lulose. The pyrolysis behaviours of these chemical
organic aerosol precursor emission from the pyrolysis of
components then result in different pyrolysis behav- thermally thick wood. Fuel 2020; 269: 117333.
iours at elevated temperatures; 6. Leventon IT, Korver KT and Stoliarov SI. A generalized
• Among these three components, lignin has a higher model of flame to surface heat feedback for laminar wall
temperature range for degradation. Softwood shows flames. Combust Flame 2017; 179: 338–353.
a higher fraction of lignin, which explains why the 7. Shi L, Chew MYL, Novozhilov V and Joseph P.
ignition temperatures of those softwoods are higher Modelling the pyrolysis and combustion behaviours of
than those of hardwoods. The average effective heat non-charring and intumescent-protected polymers using
‘FiresCone. Polymers 2015; 7: 1979–1997.
of combustion of softwoods is 9.5% higher than
8. Shi L, Chew MYL, Liu XY, Cheng XD, Wang B and
those of hardwoods;
Zhang GM. An experimental and numerical study on fire
• The thermal conductivities of softwoods increase behaviors of charring materials frequently used in build-
faster than those of hardwood when moisture con- ings. Energy Build 2017; 138: 140–153.
tent rises. The thermal conductivity of timber char 9. Lautenberger C and Fernandex-Pello AC. Optimization
increases as temperature rises, observed from most algorithms for material pyrolysis property estimation.
of the previous models. Cellulose with a high density Fire Saf Sci 2011; 10: 751–764.
shows larger thermal conductivity, but its influences 10. Shen YL, Schneider J, Tesfamariam S, Stiemer SF and
on the specific heat capacity are limited; and Chen ZW. Cyclic behavior of bracket connections for
• Recommended models to predict the temperature- cross-laminated timber (CLT): assessment and compari-
son of experimental and numerical models studies.
dependent thermal properties of the hardwood, soft-
J Build Eng 2021; 39: 102197.
wood, and char are summarized through this study. 11. Langer AM and Morse RG. The world trade center
A model was also recommended to predict the ther- catastrophe: was the type of spray fire proofing a factor
mal diffusivity of timber and timber char, including in the collapse of the twin towers? Indoor Built Environ
those at elevated temperatures. 2010; 10: 350–360.
Shi and Chew 23

12. Siddique R and Noumowe AN. An overview of the prop- 30. Wagenaar BM, Prins W and Van Swaaij WPM. Flash
erties of high-strength concrete subjected to elevated tem- pyrolysis kinetics of pine wood. Fuel Process Technol
peratures. Indoor Built Environ 2010; 19: 612–622. 1993; 36: 291–298.
13. Huang DM, Wang CY, Shen YM, Lin P and Shi L. Fire 31. Chan WCR, Kelbon M and Krieger BB. Modelling and
behaviors of vertical and horizontal polymethylmetha- experimental verification of physical and chemical pro-
crylate slabs under autoignition conditions. Proc Safety cesses during pyrolysis of a large biomass particle. Fuel
Prog 2019; 39: e12109. 1985; 64: 1505–1513.
14. Klippel M. Fire safety of bonded structural timber ele- 32. Liu NA and Fan WC. Modelling the thermal decompo-
ments. Zürich: ETH Zurich, 2014. sitions of wood and leaves under a nitrogen atmosphere.
15. Shi L and Chew MYL. Influence of moisture on auto- Fire Mater 1998; 22: 103–108.
ignition of woods in cone calorimeter. J Fire Sci 2012; 30: 33. Thurner F and Mann U. Kinetic investigation of wood
158–169. pyrolysis. Ind Eng Chem Proc Des Dev 1981; 20: 482–488.
16. Gong JH, Stoliarov SI, Shi L, Li J, Zhu SB, Zhou Y and 34. Di Blasi C. Modeling chemical and physical processes of
Wang ZR. Analytical prediction of pyrolysis and ignition wood and biomass pyrolysis. Progr Energy Combust Sci
time of translucent fuel considering both time-dependent 2008; 34: 47–90.
heat flux and in-depth absorption. Fuel 2019; 235: 913–922. 35. Nunn TR, Howard JB, Longwell JP and Peters WA.
17. Shi L and Chew MYL. A review of fire processes model- Product compositions and kinetics in the rapid pyrolysis
ing of combustible materials under external heat flux. of sweet gum hardwood. Ind Eng Chem Proc Des Dev
Fuel 2013; 106: 30–50. 1985; 24: 836–844.
18. Gruyter D. A novel approach to determine charring of 36. Swann J, Hartman J and Beyler C. Study of radiant smol-
wood in natural fire implemented in a coupled heat- dering ignition of plywood subjected to prolonged heat-
mass-pyrolysis model. Holzforschung 2020; 75: 148–158. ing using the cone calorimeter, TGA, and DSC. In: The
19. Structural Timber Association. Fire safety in timber build- 9th international symposium on fire safety science, Baden-
ings. Alloa: Structural Timber Association, 2015. Württemberg, Germany, September 21–26 2008: 98–102.
20. Gales J, Chorlton B and Jeanneret C. The historical nar- Karlsruhe: University of Karlsruhe.
rative of the standard temperature – time heating curve 37. Moghtaderi B. The state-of-the-art in pyrolysis modelling
for structures. Fire Technol 2021; 57: 529–558.
of lignocellulosic solid fuels. Fire Mater 2006; 30: 1–34.
21. Verma N and Salem O. Comparitive study on the flexural
38. Dollimore D. The kinetics of the oxidative degradation of
behaviour of glulam built-up beams based on ambient
cellulose and cellulose doped with chlorides. Thermochim
and standard fire tests. Eng Struct 2021; 233: 111759.
Acta 1987; 121: 273–282.
22. Popescu CM, Singurel G, Popescu MC, Vasile C,
39. Ozturk Z and Merklin JF. Fast pyrolysis of cellulose in a
Argyropoulos DS and Willfor S. Vibrational spectrosco-
single pulse shock tube. Biomass Bioenergy 1993; 5:
py and X-ray diffraction methods to establish the differ-
ences between hardwood and softwood. Carbohyd Polym 437–444.
2009; 77: 851–857. 40. Varhegyi G, Antal MJ, Szekely T and Szabo P. Kinetics
23. Janssens M. Fundamental thermophysical characteristics of the thermal decomposition of cellulose, hemicellulose,
of wood and their role in enclosure fire growth. PhD and sugarcane bagasse. Energy Fuels 1989; 3: 329–335.
Thesis, University of Gent, Ghent, 1991. 41. Chan RWC and Krieger BB. Kinetics of dielectric-loss
24. White RH and Dietenberger MA. Wood products ther- microwave degradation of polymers: lignin. J Appl Polym
mal degradation and fire. In: Buschow KHJ, Cahn RW, Sci 1981; 26: 1533–1553.
Flemings MC, Ilschner B, Kramer EJ, Mahajan S, 42. Spearpoint MJ and Quintiere JG. Predicting the burning
Veyssiere P (eds) Encyclopedia of materials: science and of wood using an integral model. Combust Flame 2000;
technology. Amsterdam: Elsevier, 2001, pp.9712–9716. 123: 308–324.
25. Di Blasi C. Combustion and gasification rates of lignocel- 43. Wang YF, Yang LZ, Zhou XD, Dai JK, Zhou YP and
lulosic chars. Progr Energy Combust Sci 2009; 35: 121–140. Deng ZH. Experiment study of the altitude effects on
26. Shi L, Chew MYL, Xie QY, Zhang RF, Li LM and Xu spontaneous ignition characteristics of wood. Fuel 2010;
CM. Experimental study on fire characteristics of PC 89: 1029–1034.
monitors – part I: combustion properties and pyrolysis 44. Hopkins DJ. Predicting the ignition time and burning rate
characteristics. J Appl Fire Sci 2010; 19: 23–39. of thermoplastics in the cone calorimeter. Gaithersburg:
27. Basan S and Guven O. A comparison of various isother- National Institute of Standards and Technology, 1995.
mal thermogravimetric methods applied to the degrada- 45. Janssens M and Douglas B. Wood and wood products.
tion of PVC. Thermochimica Acta 1986; 106: 169–178. In: Harper CA (ed) Handbook of building materials for
28. Dollimore D, Evans TA, Lee YF and Wilburn FW. fire protection, chapter 7. New York: McGraw-Hill, 2004,
Calculation of activation energy and pre-exponential fac- pp.7.1–7.58.
tors from rising temperature data and the generation of 46. Thornton W. The relation of oxygen to the heat of com-
TG and DTG curves from A and E values. bustion of organic compounds. Philosoph Mag J Sci
Thermochimica Acta 1991; 188: 77–85. 1917; 33: 196–203.
29. Samolada MC and Vasalos IA. A kinetic approach to the 47. Huggett C. Estimation of the rate of heat release by
flash pyrolysis of biomass in a fluidized bed reactor. Fuel means of oxygen consumption measurement. Fire
1991; 70: 883–889. Mater 1980; 4: 61–65.
24 Indoor and Built Environment 32(1)

48. ISO 9705:2016. Reaction-to-fire tests - full-scale room 58. Alves SS and Figueiredo JL. A model for pyrolysis of wet
tests for surface products - Part 2: technical background wood. Chem Eng Sci 1989; 44: 2861–2869.
and guidance. Geneva: International Organization for 59. TenWolde A, McNatt JD and Krahn L. Thermal proper-
Standardization, 2016. ties of wood and wood panel products for use in buildings.
49. Shi L and Chew MYL. Fire behaviors of polymers under Madison: Oak Ridge National Laboratory, 1988.
autoignition conditions in a cone calorimeter. Fire Safety 60. Koufopanos CA, Papayannakos N, Maschio G and
J 2013; 61: 243–253. Lucchesi A. Modelling of the pyrolysis of biomass par-
50. Harada T. Time to ignition, heat release rate and fire ticles: studies on kinetics, thermal and heat transfer
endurance time of wood in cone calorimeter test. Fire effects. Can J Chem Eng 1991; 69: 907–915.
Mater 2001; 25: 161–167. 61. Koch P. Specific heat of ovendry spruce pine wood and
51. Heskestad G. Heat of combution in spreading wood crib bark. Wood Sci 1968; 1: 203–214.
fires with application to ceiling jets. Fire Safety J 2006; 62. Volbehr B. Swelling of wood fibers. PhD Thesis,
41: 343–348. University of Kiel, Kiel, Schleswig-Holstein,
52. Wenzl H. The chemical technology of wood. New York: Germany, 1896.
Academic Press, 1970. 63. Gronli MG and Melaaen MC. Mathematical model for
53. Parker WJ. Prediction of the heat release rate of wood. wood pyrolysis – comparison of experimental measure-
Fire Saf Sci. 1986; 207–216. ments with model predictions. Energy Fuels 2000; 14:
54. BS EN 1995-1-2:2004. Eurocode 5. Design of timber struc- 791–800.
tures – part 1-2: general – structural fire design. London: 64. Ragland KW and Aerts DJ. Properties of wood for com-
British Standards Institution, 2004. bustion analysis. Biores Technol 1991; 37: 161–168.
55. Gupta M, Yang J and Roy C. Specific heat and thermal 65. Dunlap F. The specific heat of wood. Washington D.C. U.
conductivity of softwood bark and softwood char par- S. Department of Agriculture, 1912.
ticles. Fuel 2003; 82: 919–927. 66. Perry RH and Green D. Perry’s chemical engineers’ hand-
56. Suuberg EM, Milosavijevic I and Lilly WD. Behavior of book. 6th ed. New York: McGraw-Hill, 1984.
charring materials in simulated fire environments. 67. Stull DR. JANAF thermochemical tables. Washington, D.
Gaithersburg: National Institute of Standards and C.: US Government Printing Office, 1971.
Technology (NIST), 1994. 68. Gronli MG. A theoretical and experimental study of the
57. Parker WJ. Prediction of the heat release rate of Douglas thermal degradation of biomass. PhD Thesis, The
fir. In: The second international symposium on fire safety Norwegian University of Science and Technology,
science, Tokyo, June 13–17 1988: 337–346. Trondheim, Ålesund, Gjøvik, Norway, 1996.

You might also like