Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Fire Safety Journal 146 (2024) 104145

Contents lists available at ScienceDirect

Fire Safety Journal


journal homepage: www.elsevier.com/locate/firesaf

Design fire scenarios for hazard assessment of modern battery electric and
internal combustion engine passenger vehicles
Jonathan L. Hodges a ,∗, Urvin Salvi a , Anil Kapahi b
a Jensen Hughes, Inc., 2020 Kraft Drive Suite 3020, Blacksburg, 24060, VA, USA
b
Jensen Hughes, Inc., 3610 Commerce Drive Suite 817, Baltimore, 21227, MD, USA

ARTICLE INFO ABSTRACT

Keywords: Full-scale fire testing of battery electric (BEVs) and internal combustion engine (ICEVs) passenger vehicles has
Battery electric vehicles historically shown peak heat release rates (HRRs) ranging from 5–10 MW. However, the vehicles tested in
Performance based design these studies are smaller with less energy capacity than vehicles currently on the market. This paper presents
Heat release rate
a model to extrapolate data from full-scale vehicle fire tests to evaluate the fire hazard associated with larger
Fire safety
passenger vehicles. The model is validated on 20 full-scale vehicle experiments in the public literature and is
Passenger vehicles
NFPA 502
then used to develop conservatively bounding fire scenarios for vehicles representative of the current market
environment. No significant difference was observed in the HRRs between the similarly sized ICEVs and BEVs in
this work; however, the predicted HRRs for both vehicle types significantly exceeded existing design guidance.
While the model may be shown to be overly conservative in future testing, it is validated using the data that
currently exists, and points to an alarming trend in the fire protection industry, where the hazard posed by
single passenger vehicles may exceed existing design guidance.

1. Introduction turnaround facilities. Performance-based design (PBD) can be used to


assess and mitigate fire hazards when prescriptive guidance is not
Traditional guidance on fire safety design scenarios for passenger applicable.
vehicles, such as NFPA 502 [1], Standard for Road Tunnels, Bridges, and PBD is a framework for designers to customize safety measures to
Other Limited Access Highways, PIARC 05.05.BEN [2], Fire and Smoke specific circumstances and hazards which are relevant to a specific
Control in Road Tunnels, and the UK’s Design Manual for Roads and design. In this approach, specific performance metrics are established
Bridges [3], have recommended peak heat release rates (HRRs) ranging for an infrastructure design which are then compared with the set
from 5–10 MW based on data from real-scale vehicle fire tests [4– of potential hazards which may expose the infrastructure [9]. Design
6]. However, modern vehicles present a different hazard than historic basis scenarios represent the most severe credible hazard which a
vehicles due to the increase in the use of combustible plastics and the proposed trial design must effectively manage in order to meet the
presence of alternative fuel systems. Recent research has shown that established performance requirements [10]. These scenarios are created
test data from older vehicles are not representative of the hazard posed by evaluating the likely hazard sources in the area of interest and
by modern vehicles [7]. This topic was also mentioned in the recently
developing a representative progression of the hazard event based on
published PIARC summary report on the impact of new propulsion
available test data. The success of a PBD lies in the ability to charac-
technologies on road tunnel operations [8]. The authors noted that the
terize accurately the worst-case scenarios. Unfortunately, the limited
lack of data on modern vehicles makes it difficult to derive statistically
data on modern passenger vehicle fires and the rapid development
relevant information about the different hazards posed by these vehi-
of battery technologies makes it difficult to ensure that developed
cles. Ensuring robust fire safety measures for infrastructure exposed to
scenarios are conservatively bounding. There is no consensus in the fire
passenger vehicles is essential to protect occupants, first responders,
safety community on best practices in considering the degree of hazard
and infrastructure. As technology and consumer habits continue to
evolve, the regulatory landscape will likely struggle to keep pace with posed by these vehicles in design.
emerging risks, resulting in a lag in updating safety guidelines. While The primary combustible materials in internal combustion engine
passenger vehicle fires may not be the bounding design scenario for vehicles (ICEVs) consist of mixtures of plastics and polymers and liquid
tunnels, these scenarios may be bounding in parking garages and quick hydrocarbon fuels. Fully electric BEVs replace the liquid hydrocarbon

∗ Corresponding author.
E-mail address: jhodges@jensenhughes.com (J.L. Hodges).

https://doi.org/10.1016/j.firesaf.2024.104145
Received 10 October 2023; Received in revised form 27 March 2024; Accepted 29 March 2024
Available online 30 March 2024
0379-7112/© 2024 Elsevier Ltd. All rights reserved.
J.L. Hodges et al. Fire Safety Journal 146 (2024) 104145

fuel with lithium-ion batteries (LIBs) which present a different hazard vehicle and the other a larger. Four tests on single passenger sedans
profile than ICEVs. The prevailing industry guidance is that the HRR collected from Okamoto et al. [20] and data from three minivans
of burning passenger vehicles is linked primarily to the vehicle class collected from Okamoto et al. [21] were also utilized. The more recent
and size, rather than the propulsion system [7]. However, the tests on studies have primarily focused on BEVs and alternative fuel vehicles.
which these conclusions are based are limited to vehicles with small Each BEV fire experiment used a LIB of energy capacity ranging from
energy capacities relative to newer vehicles on the market. Recent tests 16–80 kWh. Relevant insights provided by the authors on their results
conducted by Sturm et al. indicate that the HRR may be higher for BEVs are summarized in the following paragraphs.
with larger energy capacities than for similarly sized ICEVs, with the A study conducted by Lecocq [22] tested two small passenger BEVs
tested vehicles having up to 80 kWh battery capacities [11]. However, with 16.5 kWh and 23.5 kWh batteries and two similarly sized ICEVs.
the 200 BEV models on the market with the largest batteries in the The tests showed that the peak HRR and the total heat released (THR)
United Stated (e.g., extended range) have capacities greater than 80 remained relatively similar between the ICEV and BEV from similar
kWh, with the largest battery over 3 times larger at 246 kWh [12]. In manufacturers. The authors noted that the ICEVs used a full tank of
addition to the increased HRR potential from the battery, the total mass gas but did not indicate the size of the tanks. Watanabe et al. [23]
of these vehicles is larger to accommodate the large batteries which conducted a similar full-scale experimental study involving a Nissan
further increases the combustible mass. Leaf electric vehicle which was compared to a gasoline-powered car.
Several researchers have proposed design fires for passenger vehi- The 24 kWh LIB of the BEV was compared to a 10 L gasoline car
cles in recent years. Boehmer et al. [7] reviewed the existing vehicle which resulted in higher maximum values for HRR in the BEV than
fire data as of 2020 and found that the majority of test data showed the gasoline vehicle. The authors advocated that BEVs could present a
HRRs in the range of 5–10 MW which is in alignment with guidance more severe fire hazard than ICEVs based on their results. It is unclear
in NFPA 502 [1]. However, the authors note that this is a wide range how full the gasoline tank was or how involved it was in the fire due
which gives the designer significant discretion in the severity of the fire to the low peak of 2 MW observed during testing. The authors used
hazard to consider. Willstrand et al. [13] developed a design scenario a heat of combustion of 22 MJ/kg in calculation of the HRR, but this
based on the full-scale test data collected in their study. The scenario value was assumed and not measured.
had a peak HRR of 5 MW which was reached approximately 22 min Lam et al. [24] tested seven full-scale vehicle fires including three
after ignition. The authors assumed a medium-slow growth rate of BEVs, two PHEVs, and two ICEVs. All three BEVs were small passenger
900 s after ignition to reach the peak. Brzezinska et al. [14] proposed vehicles with a battery capacity of 20 kWh. Two of the BEVs were the
a peak HRR of 7 MW based on an average heat release rate per unit same make and model, the only difference being the state of charge
area (HRRPUA) released from the battery and the exposed surface (one at 100% and the other at 85%). The third BEV was a different
area of the battery. Wang et al. [15] reviewed the existing data on make and model but similar in size to the other two. The primary
BEV fires and found a similar range of HRRs. While the scenarios difference in this BEV was the LIB consisted of two separate packs in
evaluated in each of these studies are relevant to the vehicles tested, different locations rather than a single pack used in the other two BEVs.
none proposed a systematic approach to extrapolate the experimental The state of charge in the third BEV was 100%. The authors observed a
results to develop conservatively bounding scenarios for design based slightly higher growth rate in the HRR of the first two BEVs as well as
on the hazard posed by larger vehicles with a larger energy capacity. the presence of a second peak which did not occur in the related ICEV;
This paper addresses this gap in fire safety design by presenting a however, the peak HRR was similar between the BEVs and the ICEV.
new approach to develop design basis scenarios for BEV, ICEV, and The third BEV had a lower peak HRR compared to the similar ICEV
Plug-in Hybrid Electric Vehicle (PHEV) fires based on the curb weight, in this study. The authors did not provide information on the energy
liquid fuel tank size, and battery energy capacity. The method is based capacity of the PHEVs or the size of the fuel tanks in the PHEVs or
on a synthesis of publicly available data on ICEV, BEV, PHEV, and ICEVs.
LIB fire experiments. The conservatism of the model is evaluated by Willstrand et al. [13] tested an ICEV full-size van, a BEV full-size
comparing it with data available in the literature. Section 2 describes van with an 40 kWh battery, and a BEV family car with a 24 kWh
the available data in the literature and the approaches used to develop battery. While the ICEV and BEV at 40 kWh had similar vehicle sizes,
the design basis scenarios. Section 3 compares design basis scenarios the BEV exhibited a higher peak heat release rate (7 MW for BEV
developed using the proposed approach to existing test data for various and around 6 MW for ICEV). The report also highlighted that the
vehicles currently available on the market. Section 4 discusses these effective heat of combustion was nearly the same for both full-size
results and the broader implications for fire safety design. vans, but considerably smaller for the BEV family car. The authors
noted these differences in burning characteristics primarily stemmed
2. Methodology from variations in fire development within the vehicles. The heat of
combustion for the ICEV measured by the authors is low compared
2.1. Full-scale test data to that presented elsewhere in the literature and may be indicative
that the gas tank was either not full or did not have a significant
There are limited full-scale fire tests conducted on passenger vehi- rupture during testing. The authors presented empirical relationships
cles due to the significant cost and the desire to protect proprietary for the peak HRR and total energy of LIB batteries related to the
information. The test data used to develop and validate the model in electrical energy capacity. These relationships are used to estimate the
this work included three older studies documented in the SFPE Hand- contribution of the battery in the design fire calculations.
book [16] as well as more recent studies. The HRR was characterized in Sturm et al. [11,27] conducted five full-scale fire tests involving
each experiment, either through oxygen consumption calorimetry [17] three BEVs in a tunnel environment. Each of these tests provides a
or fuel consumption calorimetry [18]. Table 1 summarizes key data value data point; however, it is difficult to compare with a model
collected in several recently conducted passenger vehicle fire studies. due to the uncertainties. The sedan BEV had manual intervention after
Additional details on the batteries used in the BEV experiments are 5 min with the application of a fire blanket. The utility van BEV had
provided in Table 2. an unfortunate loss of data shortly after the measured peak, leading
There were a total of 14 experiments included from the SFPE to some uncertainty if the peak HRR was achieved in the test. The
Handbook. This included seven experiments conducted by [19]. Five SUV ICEV had an unknown amount of fuel within the gas tank. The
of these tests (test numbers 2, 3, 4, 7, and 8) utilized single passenger total energy released in the utility van ICEV test (1,540 MJ) was
vehicles of varying sizes. Two of these tests (test numbers 9 and 10) less than the total energy capacity of the reported fuel tank size (50
utilized two passenger vehicles in each test, with one being a smaller L × 0.83 kg/L × 40.2 MJ/kg = 1,670 MJ). The SUV BEV had ignition

2
J.L. Hodges et al. Fire Safety Journal 146 (2024) 104145

Table 1
Full-scale vehicle fire experiments.
Study Vehicle Battery Gas Initial Peak Time to THR 𝛥Hc Lost
capacity tank mass HRR peak mass
(kWh) (L) (kg) (MW) (min) (GJ) (MJ/kg) (%)
CTICM [19] ICEV Car – 38a 951 1.2 27 1.8 9.5 19.5%
ICEV Car – 29a 757 3.5 10 2.1 15.2 18.2%
ICEV Car – 38a 955 2.2 30 3.1 21.2 15.2%
ICEV Car – 38a 1303 8.3 25 6.7 24.0 21.3%
ICEV Car – 29a 830 4.1 22 4.1 22.2 22.2%
2 ICEV Cars – 11 2096 7.5 13 8.9 30.0 14.1%
2 ICEV Cars – 11 2096 8.2 – 8.4 24.6 16.3%
Okamoto et al. [20] ICEV Car – 10 1360 3.5 26 5.0 22.0 16.5%
ICEV Car – 10 1360 3.0 25 4.9 22.0 16.3%
ICEV Car – 20 1360 1.9 15 4.9 22.0 16.5%
ICEV Car – 10 1360 2.4 56 5.0 22.0 16.8%
Okamoto et al. [21] ICEV Minivan – 10 1440 3.6 55 5.4 22.0 17.0%
ICEV Minivan – 10 1440 3.1 55 5 22.0 15.8%
ICEV Minivan – 10 1440 4.1 20 5.2 22.0 16.3%
Lecocq et al. [22] ICEV Car – 45a 1128 4.8 15 6.3 35.9 17.0%
ICEV Car – 45a 1501 6.1 17 10.0 36.4 19.6%
BEV Car 17 – 1122 4.2 25 6.3 29.8 19.0%
BEV Car 24 – 1502 4.7 30 8.5 30.7 18.6%
Watanabe et al. [23] ICEV Car – 10 1275 2.0 37 4.3 22.0b 15.3%
BEV Car 24 – 1520 6.3 40 6.4 22.0b 29.0%
Lam et al. [24] ICEV Car – 45a 1096 7.1 6 3.3 12.0 25.0%
BEV Car 20 – 1448 6.0 5 N/A N/A 23.0%
BEV Car 20 – 1475 5.9 5 4.9 17.0 20.0%
ICEV Car – 57a 1344 10.8 8.0 4.7 13.0 22.0%
BEV Car 20 – 1650 6.9 10 4.7 13.0 22.0%
PHEV Car 8a 20a 1467 6.0 7.5 4.7 15.0 21.0%
PHEV Car 8a 20a 1712 7.9 8.3 5.9 13.0 26.0%
Willstrand et al. [13] ICEV Van – 44 1326 5.9 7 5.9 23.0 19.0%
BEV Van 40 – 1587 7.0 20 5.2 21.0 15.6%
BEV Car 24 – 1540 5.0 17 6.7 17.0 25.6%
Sturm et al. [11] BEV Car 80 – 1865 7.0 5 2.9 N/A 34.0%
BEV Van 24 – 1564 6.1 5.3 4.6 N/A 17.0%
BEV SUV 80 – 1951 4.9/8.6 10/15 4.5 N/A 35.0%
Kang et al. [25] ICEV Car – 45a 1320 7.7 15 8.1 27.4 22.3%
BEV Car 39 – 1540 6.5 15 8.5 29.8 18.4%
BEV Car 64 – 1685 7.3 11 9.0 30.5 17.6%
Hynynen et al. [26] BEV SUV, no bat. 50 – 2110a 2.8 45 5.0 20.8 21.0%
BEV SUV, sprink. 50 – 2110a 3.0 107 5.7 17.3 21.0%
a
Estimated based on typical vehicle sizes/available information.
b
Assumed by the authors but not measured.

Table 2 the battery was initiated by injecting a saline solution into the battery
Battery characteristics in BEV fire experiments.
casing which resulted in a rapid growth in HRR to 8.6 MW. The authors
Study Battery SOC (%) Chemistry Cell form
noted that the HRR of BEVs is higher than conventional gasoline cars;
capacity (kWh) factor
however, the amount of additional heat released hinges on the level of
Lecocq et al. [22] 17 100 * *
24 100 * *
the battery’s involvement in the fire. Due to the high uncertainties in
the quantity of liquid fuel included in the test, the ICEV tests are not
Watanabe et al. [23] 24 100 LMO/LNO Pouch
included in benchmarking.
Lam et al. [24] 20 100 * *
20 85 * *
Kang et al. [25] compared the HRR between ICEVs, BEVs, and
20 100 * * Hydrogen Fuel Cell Electric Vehicles (FCEV). The BEVs tested by the
Willstrand et al. [13] 40 80 NMC Pouch authors included a 39 kWh and 64 kWh capacity. The authors also
24 80 NMC Prismatic tested subassemblies to apportion the HRR into its constitutive com-
Sturm et al. [11] 80 100 NMC Pouch ponents. These tests included tests of individual LIB pack fires as well
24 * LMO Pouch as fires involving the entire BEV body with the LIB pack removed prior
80 100 NMC Pouch
to testing. The apportioned HRR between the LIB packs and the BEV
Kang et al. [25] 39 100 NMC Pouch provide a basis for calculating the HRR of BEVs in this study.
64 100 NMC Pouch
Hynynen et al. [26] investigated the effectiveness of sprinklers at
Hynynen et al. [26] 50 90 NMC Prismatic
controlling the HRR of small SUV BEV fires. Data were presented from
*Unknown/not provided by authors. two BEV tests, one with sprinklers and one without sprinklers. The
non-sprinklered BEV scenario had the LIB removed prior to testing
to evaluate the burning of the non-battery combustibles. The authors
located in the rear seats and the battery was not involved in the fire found that the peak HRR and THR were affected by the fire scenario
for the first 800 s, resulting in a lower peak at 4.9 MW. However, and size of the vehicle but not significantly different between ICEVs
after the fire had caused some initial degradation, thermal runaway in and BEVs.

3
J.L. Hodges et al. Fire Safety Journal 146 (2024) 104145

Table 3
Representative vehicles on the U.S. market.
Class BEV ICEV
Curb weight Battery energy Curb weight Gas
(kg) capacity (kWh) (kg) tank (L)
Compact car 1600 40 1300 34
Sports car 2450 150 1800 61
Large car 2600 115 1760 57
8 passenger van 3000 100 2500 68
Pickup truck 3250 135 2600 87
Large SUV 4300 250 2900 121

2.2. Representative vehicles

As discussed in the introduction, modern vehicles can be signifi-


cantly larger with higher energy capacities than what has been tested to
date. Table 3 shows a selection of representative BEV and ICEV vehicles
which was created to analyze the potential fire hazard of these vehicles
coming to market using the approach presented in this work. These Fig. 1. Comparison of consumed mass with measured peak HRR.
vehicles were selected by manual observation of vehicles online [12].
Representative curb weights and battery capacities were developed by
examining high range vehicles in each vehicle category available in the
aligning with observations on the experimental data. The impact of this
U.S. BEV market. A similar approach was used to identify comparably
quantity on the HRR in the model is further discussed in Section 4.
sized ICEV vehicles for comparison. Note that the pickup truck class
There are similar difficulties in estimating the peak HRR of the non-
vehicle corresponds to a light-duty personal vehicle common in the
propulsion related vehicle components, 𝑄̇ 𝑣𝑒ℎ . The majority of passenger
U.S. market, and the large SUV class corresponds to heavy-duty off-
vehicle test data in the literature includes some quantity of liquid
road personal vehicles fairly unique to the U.S. The SUVs tested by
fuel in the gas tanks which increases the uncertainty in scaling these
Sturm et al. [11] and Hynynen et al. [26] correspond to typical SUV
components. Fig. 1 compares the consumed mass from ICEV vehicle fire
sizes in many markets; however, the large SUV class in the US can be
tests in the literature with the peak HRR reported by the authors. These
significantly larger.
cases include the 21 ICEV cases from Table 1. The linear fit corresponds
to the equation
2.3. Design fire HRR calculation
𝑄̇ 𝑣𝑒ℎ,𝑟𝑒𝑓 𝑒𝑟𝑒𝑛𝑐𝑒
𝑄̇ 𝑣𝑒ℎ,𝑠𝑐𝑎𝑙𝑒𝑑 = 𝑚𝑐,𝑣𝑒ℎ,𝑠𝑐𝑎𝑙𝑒𝑑 = 0.02 × 𝑚𝑐,𝑣𝑒ℎ,𝑠𝑐𝑎𝑙𝑒𝑑 (4)
The empirical correlations presented by Willstrand et al. [13] and 𝑚𝑐,𝑣𝑒ℎ,𝑟𝑒𝑓 𝑒𝑟𝑒𝑛𝑐𝑒
the apportioned HRR results presented by Kang et al. [25] were used
where the subscript reference indicates the data collected by the authors
as the primary basis to scale the design fire HRR and THR proposed in
and scaled indicates the design vehicle, and the subscript c indicates the
this study. The data provided by the authors for the full vehicle as well
consumed mass during testing. The slope in Eq. (4) was calculated using
as independently for the battery and vehicle are used to scale the HRR
a linear regression. Note that the 11 MW peak in Fig. 1 corresponds to
and THR based on their individual heats of combustion.
a relatively short-lived peak HRR from one of Lam et al.’s tests from
significant liquid fuel involvement. Outside of the short lived peak, the
2.3.1. Peak heat release rate
sustained peak was around 6 MW which is also in line with the linear
The peak HRR of a vehicle fire in this analysis consists of the linear
fit. 𝑄̇ 𝑣𝑒ℎ,𝑠𝑐𝑎𝑙𝑒𝑑 was scaled using Eq. (4) in this work.
combination of individual contributions of the LIB, liquid gasoline (for
The mass of the other vehicle components (i.e., not battery and not
ICEVs and PHEVs), and other vehicle components,
liquid fuel) is related to the curb weight of the vehicle, battery mass,
𝑄̇ 𝑚𝑎𝑥 = 𝑄̇ 𝑏𝑎𝑡 + 𝑄̇ 𝑣𝑒ℎ + 𝑄̇ 𝑔𝑎𝑠 (1) and the size of the gas tank using the equation

where 𝑄̇ is the peak HRR and the subscripts indicate the total, battery, 𝑚𝑖,𝑣𝑒ℎ = 𝑚𝑖,𝑡𝑜𝑡 − 𝑚𝑖,𝑏𝑎𝑡 − 𝑚𝑖,𝑔𝑎𝑠 (5)
vehicle, and liquid fuel contributions, respectively.
where 𝑚𝑖,𝑣𝑒ℎ is the mass of other vehicle components, 𝑚𝑖,𝑡𝑜𝑡 is the total
𝑄̇ 𝑏𝑎𝑡 is scaled based on the empirical relationship presented by
curb weight of the vehicle, and 𝑚𝑖,𝑔𝑎𝑠 is the mass of the liquid fuel in
Willstrand et al. [13],
the tank in the design scenario. The mass of an arbitrary LIB battery
( 0.67 )
𝑄̇ 𝑏𝑎𝑡 = 0.16 × 𝐸𝑒𝑙𝑒𝑐 (2) pack is estimated based on the energy capacity and mass measured by
Kang,
where 𝑄̇ 𝑏𝑎𝑡 is in MW and 𝐸𝑒𝑙𝑒𝑐 is the electrical energy capacity of the 𝑚𝑖,𝑏𝑎𝑡,𝑟𝑒𝑓 𝑒𝑟𝑒𝑛𝑐𝑒
battery in kWh. Note the units in Eq. (2) have been converted from the 𝑚𝑖,𝑏𝑎𝑡,𝑠𝑐𝑎𝑙𝑒𝑑 = 𝐸 (6)
𝐸𝑒𝑙𝑒𝑐,𝑟𝑒𝑓 𝑒𝑟𝑒𝑛𝑐𝑒 𝑒𝑙𝑒𝑐,𝑠𝑐𝑎𝑙𝑒𝑑
original presentation to be MW and kWh.
It is difficult to estimate 𝑄̇ 𝑔𝑎𝑠 in a vehicle fire due to the high vari- where m is the mass in kg, the subscript i indicates the initial value be-
ability of events which can occur. For example, the tank may rupture fore the fire, and the subscript bat indicates the battery. The reference
violently resulting in rapid vaporization and subsequent combustion of battery had a capacity of 64 kWh and an initial mass of 446 kg. The
the fuel, or the gasoline may leak and accumulate as a pool on the consumed mass of the other vehicle components, 𝑚𝑐,𝑣𝑒ℎ , is calculated
ground. In this work, 𝑄̇ 𝑔𝑎𝑠 is instead related to a fixed time required to based on the consumed mass fraction observed during testing,
consume all of the gasoline, according to the equation 𝑚𝑐,𝑣𝑒ℎ,𝑟𝑒𝑓 𝑒𝑟𝑒𝑛𝑐𝑒
𝑚𝑐,𝑣𝑒ℎ = 𝑚𝑖,𝑣𝑒ℎ × (7)
𝑚𝑔𝑎𝑠 𝛥𝐻𝑐,𝑔𝑎𝑠 𝑚𝑖,𝑣𝑒ℎ,𝑟𝑒𝑓 𝑒𝑟𝑒𝑛𝑐𝑒
𝑄̇ 𝑔𝑎𝑠 = (3)
𝛥𝑡 As shown in Table 1, the consumed mass fraction varies widely
where 𝛥𝑡 is the fixed time interval of the gasoline burning. A fixed value across tests and vehicle types. Fig. 2 compares the curb weight of tested
of 𝛥𝑡 of 600 s (10 min) is used in the statistical analysis in this work vehicles with the fraction of vehicle mass that was consumed during

4
J.L. Hodges et al. Fire Safety Journal 146 (2024) 104145

Fig. 2. Comparison of consumed mass with curb weight. Fig. 3. Comparison of heat of combustion with consumed mass fraction.

testing for each of the studies discussed in Section 2.1. The data shows peak HRR. The same values were used in calculation of 𝐸𝑡𝑜𝑡 in the
there is not a strong correlation between curb weight and the mass two configurations; however, because 𝛥𝐻𝑐 scales with the consumed
loss fraction; however, the variability does tend to increase with larger mass fraction, 𝐸𝑡𝑜𝑡 is fairly independent of the consumed mass fraction.
vehicles. The two data points from Sturm et al. in particular have a Instead, conservatism is added by directly applying a safety factor to
much higher mass loss fraction than the other data points. Two values the total energy predicted by the model,
for mass loss percent were used in this work, 17% and 25%, which
were determined statistically to be the best fit to the data and meet the 𝐸𝑡𝑜𝑡,𝑐𝑜𝑛𝑠 = 𝑆𝐹 × 𝐸𝑡𝑜𝑡 (10)
conservatively bounding objective, respectively.
where 𝐸𝑡𝑜𝑡,𝑐𝑜𝑛𝑠 is the conservatively bounding 𝐸𝑡𝑜𝑡 and 𝑆𝐹 is the safety
factor. The best fit results shown in Section 3 use the 17% consumed
2.3.2. Total heat released
mass fraction with 𝑆𝐹 = 1.0, and the conservatively bounding results
Similarly, the total energy associated with a vehicle fire in this
use the 25% consumed mass fraction with 𝑆𝐹 = 1.35.
analysis consists of the linear combination of individual contributions
of the LIB, other vehicle components, and liquid gasoline (for ICEVs
and PHEVs), 2.3.3. Characterizing the time-resolved HRR
Design fire heat release rates are often characterized as t-squared
𝐸𝑡𝑜𝑡 = 𝐸𝑏𝑎𝑡 + 𝑚𝑐,𝑣𝑒ℎ 𝛥𝐻𝑐,𝑣𝑒ℎ + 𝑚𝑐,𝑔𝑎𝑠 𝛥𝐻𝑐,𝑔𝑎𝑠 (8) fires according to the equation
where 𝐸𝑡𝑜𝑡 is the THR of the full vehicle including the battery, other 𝑄 = 𝛼𝑡2 (11)
components, and liquid fuel in MJ, 𝛥𝐻𝑐 is the heat of combustion,
the subscript veh indicates the value corresponding to the non-battery where t is the time in minutes and 𝛼 is the fire growth rate in MW∕min2 .
vehicle components only, and the subscript gas indicates the gasoline. NFPA 72 defines growth rates based on the time required for a fire
The total heat released by the battery is calculated using the empir- to grow to 1.055 MW [29]. A medium growth rate is characterized
ical relationship from Willstrand et al. by a growth time of 5 min, and a fast growth rate by a time of
2.5 min. Kang et al. identified that the growth rate of the two BEVs
𝐸𝑏𝑎𝑡 = 48.5 × 𝐸𝑒𝑙𝑒𝑐 (9) tested was generally described by a medium to medium-fast growth
where 𝐸𝑏𝑎𝑡 is the total heat released by the battery in MJ. Note the rate [25]. In this work, the time to peak HRR was calculated based on
units in Eq. (9) have been converted from the original presentation to the time required to reach the peak based on a medium-fast growth
be MJ and kWh. The contribution of liquid fuels to the total energy is rate (i.e., characteristic time of 3.75 min corresponding to 𝛼 = 0.075
calculated based on the size of the fuel tank assuming a liquid density MW∕min2 ).
of 720 kg∕m3 [28]. The heat of combustion of the gasoline, 𝛥𝐻𝑐,𝑔𝑎𝑠 , is The time-resolved profile of a design fire is difficult to quantify
considered to be 44.1 MJ/kg [28]. when deviating from directly using experimental measurements. One
An estimate of the heat of combustion of the other vehicle compo- common approach used is to consider the t-squared growth until the
nents (i.e., non-battery and non-liquid fuel mass), 𝛥𝐻𝑐,𝑣𝑒ℎ , is needed fire reaches its peak and then freeze the HRR at that level until the
to scale the total energy content of these components. The heat of fuel is consumed. While this may be conservative in some scenarios,
combustion of other vehicle components measured by Kang et al. was this is not representative of the typical burning behavior of a vehicle.
28.8 MJ/kg; however, this value is high compared to many of the Instead, the exponential curve approach originally proposed by Numa-
other studies presented in Table 1. There are many potential sources jiri et al. [30] and later adapted by Ingasson [31] for design fires in the
for variation in the heat of combustion across tests, including the transportation section is used in this work. This model represents the
ignition methodology, calorimetry methods, laboratory conditions, and time resolved HRR according to the following equation
localized flame extinction resulting in variable quantities of unburned
hydrocarbons being exhausted. Fig. 3 compares the measured mass loss 𝑄(𝑡) = 𝑄𝑚𝑎𝑥 𝑛𝑟(1 − 𝑒−𝑘𝑡 )𝑛−1 𝑒−𝑘𝑡 (12)
fraction with the 𝛥𝐻𝑐 across the available full-scale test data. Note that where 𝑄𝑚𝑎𝑥 is the peak HRR in kW, n, r, and k are shape factors related
the data where the curb weights (Hynynen et al.) or 𝛥𝐻𝑐 (Sturm et al. to the time to peak, THR, and 𝑄𝑚𝑎𝑥 . The shape factor parameters are
and Watanabe et al.) were not reported are omitted from the figure. calculated using the empirical fits,
As discussed in Section 2.3.1, two values for consumed mass fraction
were used to provide a best estimate and a conservatively bounding 𝑛 = 0.74294𝑒2.9𝑄𝑚𝑎𝑥 𝑡𝑚𝑎𝑥 ∕𝐸𝑡𝑜𝑡 (13)

5
J.L. Hodges et al. Fire Safety Journal 146 (2024) 104145

Table 4
Calculated fire hazard of representative vehicles on the market.
𝑟 = (1 − 1∕𝑛)1−𝑛 (14) Class BEV ICEV
𝑡𝑚𝑎𝑥 𝑄𝑚𝑎𝑥 𝐸𝑡𝑜𝑡 𝑡𝑚𝑎𝑥 𝑄𝑚𝑎𝑥 𝐸𝑡𝑜𝑡
(min) (MW) (GJ) (min) (MW) (GJ)
𝑘 = 𝑟𝑄𝑚𝑎𝑥 ∕𝐸𝑡𝑜𝑡 (15) Compact car 10.6 8.5 9.8 10.5 8.3 8.4
Sports car 12.4 11.6 17.4 12.8 12.2 12.2
where 𝑡𝑚𝑎𝑥 is the time to peak in seconds, and 𝐸𝑡𝑜𝑡 is in kJ. Large car 13.1 12.8 17.2 12.5 11.8 11.8
8 passenger van 14.1 15.0 19.0 14.6 16.1 16.3
Pickup truck 14.5 15.8 21.3 15.3 17.6 17.6
2.4. Conservatism characterization Large SUV 16.0 19.3 30.2 16.7 20.9 20.6

The objective of a typical modeling approach is to have a model


prediction without any systematic bias. However, this approach may
3. Results
not be desirable in a performance-based design because the model
predictions may prove to be non-conservative compared with the ex-
Scatter plots demonstrating the performance of the model are pro-
periments due to the uncertainties (i.e., half of model predictions vided in Fig. 4 and Fig. 5 for 𝑄̇ 𝑚𝑎𝑥 and 𝐸𝑡𝑜𝑡 , respectively. The left plot
would be under-predicting the HRR compared with experiments). The in each figure shows the statistical best fit of the model, whereas the
objective of this analysis is instead to have as little variability about right plot shows the proposed conservatively bounding fit. In each plot
the mean as possible, while purposely having systematic bias in the the 𝑥-axis corresponds to the experimental measurement provided by
conservative direction to account for the variable uncertainty. A model the authors and the 𝑦-axis corresponds to the model prediction based
in which the bound in the non-conservative direction is above the mean on the curb weight, battery capacity, and liquid fuel capacity. The
experimental value is considered to be conservatively bounding in this solid black line shows the 1:1 line with the experiments, and the two
work. dashed black lines show 95% confidence intervals on the experimental
The conservatism of the design fires generated in this work is measurements. The solid red line shows the bias-adjusted centerline of
characterized by evaluating the model uncertainty using the approach the design basis predictions, and the two dashed red lines show the
presented by McGrattan et al. [32]. The relative bias describes the 95% confidence interval on these predictions. The overall uncertainty
systematic tendency of the model to over-predict (> 1) or under-predict is documented in the bottom-right of each figure. Note that the optimal
(< 1) the measured values. The model standard deviation describes behavior of the best fit model would be for the red lines to overlap
the scatter of the predicted value about the experimental measurement with the solid lines; however, the optimal behavior for the conservative
after correcting for any systematic bias present in the predictions. In the design fire model is for the lower red dashed line to be in line with the
context of a conservatively bounding design scenario, the desired bias solid black line (conservatively bounding) and the spread between the
is sufficiently larger than 1 such that the standard deviation in error two red dashed lines be as small as possible (not overly conservative).
will result in conservative predictions at the lower end (i.e., higher than The time-resolved design fire for each vehicle calculated using the
experimental measurements). However, the bias should not be higher methodology proposed in this work is compared with the experimental
than needed to accomplish this objective as higher values will be overly data in Fig. 6. Each figure compares the experimentally measured
conservative. The model uncertainty in this approach is calculated for HRR profile (dashed line) with the representative design fire using the
a discrete set of comparisons using the equations proposed methodology with the conservatively bounding configuration
(solid line). Each study used different ignition and initiation strategies.
2
[ ]
𝜎𝑀 = 𝑚𝑎𝑥 𝑉 𝑎𝑟 (𝑙𝑛 (𝑀∕𝐸)) − 𝜎𝐸2 , 𝜎𝐸2 (16) As a result, several of the cases had long incipient periods prior to the
ignition and rapid fire growth associated with the LIB. The test timeline
( ) was offset in these cases such that a time of zero corresponds closer to
2
𝛿 = 𝑒𝑥𝑝 𝑙𝑛 (𝑀∕𝐸) + 0.5𝜎𝑀 − 0.5𝜎𝐸2 (17)
the onset of rapid growth in HRR. Note the study by Hynynen et al. is
where M is the set of model predictions on the set of experimental not benchmarked due to the presence of sprinklers in the testing.
observations, E, 𝜎𝐸 is the experimental deviation, Var indicates the Similar design fire curves for a selection of representative BEVs and
variance assuming a normal distribution, and the overbar indicates ICEVs in Table 3 are shown in Fig. 7. The calculated peak HRR, THR,
the mean. The uncertainty in heat release rate measurements varies and time to peak for each of the representative vehicles in Table 3 are
significantly based on measurement technique and operating condi- provided in Table 4. The peak 𝑄̇ 𝑚𝑎𝑥 is similar between the similarly
sized BEVs and ICEVs, with the exception of the ICEV pickup truck
tions. An experimental uncertainty of 7.5% is assumed in this work
which has a 10% higher HRR than the BEV counterpart. However,
and is used in calculation of the model uncertainty. This value is at
across all vehicle classes the BEVs have a higher calculated 𝐸𝑡𝑜𝑡 than
the upper limit of theoretical uncertainties in oxygen consumption
the ICEV counterparts. Qualitatively, this leads to sharper peak in the
calorimetry (2.5–7.0%) [33] and in the middle of the reproducibility
predicted HRR profile of the ICEVs than is predicted for the BEVs. This
of peak and time-averaged peak HRRs measured in round robin stud-
trend is particularly evident in the large SUV case shown in Fig. 7.
ies [34]. While this estimate is reasonable for laboratory-scale testing,
These results are discussed in more detail in Section 4.
the true experimental uncertainty in full-scale vehicle fire testing is
likely higher. 4. Discussion
One of the key uncertainties in the test data and the proposed
model is the consumed mass percent. Experimentally, this percent has The comparisons in Fig. 6 demonstrate qualitatively that the pro-
been observed to vary from 15%–35%. This parameter was used to posed design fire calculation methodology generates conservatively
tune the model performance to achieve the performance objective. The bounding descriptions of each test. In this context, we consider conser-
value of 17% was found to provide the best fit agreement with the vatively bounding to mean that the HRR of the design basis scenario
experimental measurements (i.e., 𝛿 = 1), and the value of 25% was is higher than the experiment for the duration of testing. This criterion
found to be representative as conservatively bounding based on the is generally true except in the cases tested by Lecocq et al. [22]. In
experiments. Model performance results are provided for both of these these cases, the model over-predicted the peak HRR, particularly in
cases in Section 3. Recommended passenger vehicle design fire HRRs the 24 kWh experiment, which resulted in the decay phase of the
are provided using the conservatively bounding 25% in Section 4. fire occurring more quickly than observed experimentally. Typically, in

6
J.L. Hodges et al. Fire Safety Journal 146 (2024) 104145

Fig. 4. Statistical performance of model on 𝑄̇ 𝑚𝑎𝑥 .

Fig. 5. Statistical performance of model on 𝐸𝑡𝑜𝑡 .

PBD scenarios the earlier parts of the fire are of particular interest due and an associated variation of 14%. This was deemed to be overly
to the need to consider the impact on occupant tenability during egress conservative, leading to the recommended 10 min duration in this
and the fast response time of local fire service. In these applications the work. While it is possible that the variability could be decreased below
two scenarios would be conservative compared with Lecocq’s experi- 14%, the lower limit in the statistical assessment approach is to match
mental measurements. The model can be readily adjusted to remove the experimental uncertainty of 7.5%.
this non-conservatism by changing the assumed consumption percent The same trend is observed in the THR statistics as well, with the
of the vehicle. However, the trade-off to this adjustment would be lower end of model predictions consistently above the experimental
significantly worse predictions across the other cases. measurements. However, the bias factor and associated variation is
The quantitative results shown in Fig. 4(b) and Fig. 5(b) demon- higher in this metric at 1.71 and 36%, respectively. Physically, the
strate that the model achieves the overall objectives posed in Sec- over-estimation of the THR results in a longer decay phase than is
tion 2.4. The systematic bias of 1.34 in the peak HRR and the asso- typically seen experimentally. As discussed previously, the earlier parts
ciated variation of 14% shows that, after accounting for uncertainty, of the fire are of particular interest in typical PBD applications. Over-
the model predictions are consistently above the experimental mea- estimating the THR and the resulting duration of the exposure can be
surements. The one exception to the model providing conservatively beneficial in design to ensure that sufficient fire protection features and
bounding peak HRRs is the ICEV case where a peak HRR of 10 MW suppression resources are available for extended fireground operations.
was measured. The initiating fire in this case was a 2 MW burner The variability in the model is also within the same range of
located near the fuel tank. This resulted in the liquid fuel burning variability observed in testing. Consider each of the small vehicles with
significantly faster than the 10 min burning duration assumed in this a battery capacity in the 20–24 kWh range and a curb weight ranging
work. While a more conservative burning duration could be assumed from 1440–1540 kg in Table 1. The experimental range in peak HRRs
to better align with this test, this would increase the predictions of the observed in these 5 cases was 4.7–6.3 MW, corresponding to a 29%
model for all ICEV and PHEV cases. For example, assuming a 7 min difference across experiments. Similarly, the range of THRs observed in
burning duration for the gasoline aligns the model predicted peak HRR these 5 cases was 4.9–8.5 GJ which corresponds to a 54% difference.
with the measurement. However, this results in a bias factor of 1.40 There are many reasons for the spread in experiments, some examples

7
J.L. Hodges et al. Fire Safety Journal 146 (2024) 104145

Fig. 6. Comparison of design fire HRRs with experimental measurements. Solid lines are the design fire fits and dashed lines are the experimental data.

include differences in cell composition, arrangement, other vehicle vehicle components are the largest individual contributor to the overall
components, and test methodologies. Unless the model is adapted to peak HRR. While the overall curb weight of BEVs is higher than that of
be more specific to these parameters, it is unlikely that the overall the similarly sized ICEV, the contribution of other vehicle components
statistical performance will be significantly improved. This may be to the total HRR was less than in the ICEV for the majority of cases. The
possible in the future as more vehicles are tested and data shared with reason this occurs in the model is due to the battery representing a large
the community; however, the current set of available data is insufficient fraction of the curb weight. After removing the estimated battery mass
to further adapt the model to include these other parameters. from the curb weight, the total combustible mass of the other vehicle
Fig. 8 breaks down the HRR of each representative vehicle from Ta- components is less in most cases. The contribution of the fuel source to
ble 3 by the contribution from the individual fuel sources. These results the peak HRR is generally similar between the BEV and ICEV vehicles;
indicate that the peak HRR for comparable vehicle classes of BEVs and however, in the majority of cases, the gasoline contributes a slightly
ICEVs is similar; however, the peak HRR was generally predicted to be higher HRR than the battery in the proposed model. This is due to the
higher for the ICEV vehicles than for the BEV vehicles. In each case the conservative assumption that all of the gasoline in the tank is consumed

8
J.L. Hodges et al. Fire Safety Journal 146 (2024) 104145

Fig. 7. Representative single passenger vehicle design fire HRRs.

Fig. 8. Representative single passenger vehicle peak HRRs by fuel.

within a fixed ten minute interval. Adjusting the gasoline burning time present unique toxic and explosion hazards. These additional hazards
interval to a longer duration representative of more of the test data become more prominent in enclosed spaces and will be addressed in
would yield more similar HRRs. There is additional conservatism built our future work.
into both the BEVs and ICEVs analysis in the 100% state of charge The recommended design fires presented here may be found to
assumed for the battery and the full gas tank assumed in the ICEV. be overly conservative in the future based on additional testing of
This study focused primarily on the fire behavior of a single pas- larger vehicles with higher fuel capacities. However, the model has
senger vehicle. This provides a starting point for the growing hazard been validated on the range of data that currently exists and points
associated with larger vehicles; however, it is often necessary to con- to a potential problem with existing and future designs. As consumers
sider the spread of the fire to adjacent vehicles in design. Existing PBD adopt larger vehicles with higher loads of combustible plastics, fuels,
guidance has recommended 5–8 MW as a representative peak HRR for a and battery energy capacities, the potential fire hazard posed by one of
vehicle for use in these analyses. NFPA 502 recommends a peak value of these vehicles may exceed prior design guidance for multiple passenger
8 MW for a single passenger car, and a peak HRR for multiple passenger cars.
cars of 15 MW [1], and ISO guidance on car park design provides
similar recommendations in the absence of utility vehicles [35]. While 5. Conclusion
Fig. 7 does not show a significant difference between the HRRs of
BEVs and ICEVs of similar vehicle class, it is possible that the potential This paper presented a new approach to establish design fire scenar-
for flame spread between vehicles could differ between these vehicle ios for passenger vehicles based on the data available in the literature.
classes. Funk et al. noted a similar concern due to the directionality of The model was designed to provide conservatively bounding predic-
fire due to the jetting of flammable gases from battery cells, which is tions of the time resolved HRR profile of BEVs, ICEVs, and PHEVs based
unique to BEVs, could increase the likelihood of flame spread to com- on the energy capacity of the batteries, the curb weight of the vehicle,
bustibles or vehicles in the vicinity [36]. However, it is unclear whether and the capacity of the gas tank. Model predictions were validated with
this additional hazard is more severe than the unique hazard posed by experimental data from 20 full-scale fire tests available from 6 studies
a pool fire developing in an ICEV which could also expose adjacent in the publicly available literature. Overall, the model predictions were
vehicles. In addition, though this paper focuses on the fire hazard, BEVs found to be consistently conservative in both HRR and THR with bias

9
J.L. Hodges et al. Fire Safety Journal 146 (2024) 104145

factors of 1.34 and 1.71, respectively. The model was used to evaluate [11] P. Sturm, P. Fößleitner, D. Fruhwirt, R. Galler, R. Wenighofer, S.F. Heindl, S.
the fire hazard posed by several representative vehicles currently on Krausbar, O. Heger, Fire tests with lithium-ion battery electric vehicles in road
tunnels, Fire Saf. J. 134 (September) (2022).
the market or soon to be on the market. The total energy released
[12] Electric cars battery gross capacity comparison chart, 2023, https://www.
was consistently higher in the BEVs than in the equivalent ICEVs, myevreview.com/comparison-chart/battery-capacity-kwh. (Accessed 21 Septem-
particularly in the larger vehicles. While no significant difference was ber 2023).
observed between the HRRs predicted for ICEVs and BEVs, the HRRs [13] O. Willstrand, R. Bisschop, P. Blomqvist, A. Temple, J. Anderson, Toxic Gases
predicted for larger vehicles of both propulsion types were found to from Electric Vehicle Fires, tech. rep., RISE Research Institutes of Sweden, 2020.
[14] D. Brzezinska, P. Bryant, Performance-based analysis in evaluation of safety in
significantly exceed existing design guidance on passenger vehicles. car parks under electric vehicle fire conditions, Energies 15 (2) (2022).
As consumers continue to adopt larger vehicles with higher energy [15] X. Wang, M. Wang, R. Jiang, J. Xu, B. Li, X. Wang, M. Lei, P. Su, C. Liu, Q. Yang,
capacities, it will be important for the fire protection community to J. Yu, Impact of battery electric vehicles on ventilation design for road tunnels:
ensure that sufficiently conservative design fires are used to represent A review, Tunnell. Undergr. Space Technol. 134 (February) (2023) 105013.
[16] V. Babrauskas, Heat release rates, in: SFPE Handbook of Fire Protection
these increasing hazards. While the approach presented in this paper
Engineering, fifth ed., 2016, pp. 799–904.
is focused primarily on passenger vehicle fires, the general concepts on [17] ASTM E2067, Standard Practice for Full-Scale Oxygen Consumption Calorimetry
creating conservative design fires and the statistical analysis approach Fire Tests, tech. rep., ASTM International, 2016, pp. 1–24.
can be readily adapted to other fire hazard analyses. [18] R.A. Bryant, M.F. Bundy, The NIST 20 MW Calorimetry Measurement System for
Large-Fire Research, NIST Technical Note 2077, 2019.
[19] E. Cresson, Development of Design Rules for Steel Structures Subjected to Natural
CRediT authorship contribution statement Fires in Closed Car Parks, tech. rep., CTICM, 1997, p. 160.
[20] K. Okamoto, N. Watanabe, Y. Hagimoto, T. Chigira, R. Masano, H. Miura, S.
Jonathan L. Hodges: Writing – review & editing, Writing – original Ochiai, H. Satoh, Y. Tamura, K. Hayano, Y. Maeda, J. Suzuki, Burning behavior
draft, Visualization, Validation, Supervision, Software, Project admin- of sedan passenger cars, Fire Saf. J. 44 (3) (2009) 301–310.
[21] K. Okamoto, T. Otake, H. Miyamoto, M. Honma, N. Watanabe, Burning behavior
istration, Methodology, Investigation, Formal analysis, Conceptualiza-
of minivan passenger cars, Fire Saf. J. 62 (PART C) (2013) 272–280.
tion. Urvin Salvi: Writing – original draft, Visualization, Methodology, [22] A. Lecocq, M. Bertana, B. Truchot, G. Marlair, Comparison of the fire conse-
Data curation. Anil Kapahi: Writing – review & editing, Visualization, quences of an electric vehicle and an internal combustion engine vehicle, in:
Methodology. International Conference on Fires in Vehicles-FIVE, Vol. 2, 2012, pp. 183–194.
[23] N. Watanabe, O. Sugawa, T. Suwa, Y. Ogawa, M. Hiramatsu, H. Tomonori, H.
Miyanoto, K. Okamoto, M. Honma, Comparison of fire behaviors of an electric-
Data availability
battery-powered vehicle and gasoline-powered vehicle in a real-scale fire test, in:
Proceedings from 2nd International Conference on Fires in Vehicles-FIVE, 2012,
The data used in this paper is all available in the public literature. pp. 195–206.
[24] C. Lam, D. MacNeil, R. Kroeker, G. Lougheed, G. Lalime, Full-scale fire testing of
electric and internal combustion engine vehicles, in: 4th International Conference
Acknowledgment
on Fire in Vehicle, Baltimore, 2016, pp. 95–106.
[25] S. Kang, M. Kwon, J. Yoon Choi, S. Choi, Full-scale fire testing of battery electric
All authors approved the version of the manuscript to be published. vehicles, Appl. Energy 332 (November 2022) (2023) 120497.
[26] J. Hynynen, O. Willstrand, P. Blomqvist, P. Andersson, Analysis of combustion
Declaration of competing interest gases from large-scale electric vehicle fire tests, Fire Saf. J. 139 (March) (2023)
103829.
[27] P.-J. Sturm, P. Fößleitner, D. Fruhwirt, S.F. Heindl, B. Kohl, O. Heger, R.
The authors declare that they have no known competing finan- Galler, R. Wenighofer, S. Krausbar, Brandauswirkungen Von Fahrzeugen Mit
cial interests or personal relationships that could have appeared to Alternativen Antriebssystemen: Ergebnisbericht, Technische Universität Graz,
influence the work reported in this paper. Institut für Verbrennungskraftmaschinen und . . . , 2021.
[28] SFPE, Appendix 3: Fuel properties and combustion data, in: SFPE Handbook of
Fire Protection Engineering, fifth ed., Society of Fire Protection Engineers, 2016.
References [29] NFPA, NFPA 72: National Fire Alarm and Signaling Code, National Fire
Protection Association, 2019, p. 5406.
[1] NFPA, NFPA 502 Standard for Road Tunnels, Bridges, and Other Limited Access [30] F. Numajiri, K. Furukawa, Mathematical expression of heat release rate curve
Highways, National Fire Protection Association, 2020. and proposal of ’Burning Index’, Fire Mater. 22 (1) (1998) 39–42.
[2] F. PIARC, Smoke Control in Road Tunnels, PIARC Committee on Road Tunnels, [31] H. Ingason, Design fire curves for tunnels, Fire Saf. J. 44 (2) (2009) 259–265.
C5, 05.05, Vol. 32, B, Paris, France, 2002. [32] K. McGrattan, B. Toman, Quantifying the predictive uncertainty of complex
[3] S.H. Company, Design Manual for Roads and Bridges, UK, 2020. numerical models, Metrologia 48 (3) (2011) 173–180.
[4] D. Brzezinska, R. Ollesz, P. Bryant, Design car fire size based on fire statistics [33] M.L. Janssens, Variability in Oxygen consumption calorimetry tests, in: Thermal
and experimental data, Fire Mater. 44 (8) (2020) 1099–1107. Measurements: The Foundation of Fire Standards, ASTM STP. Vol. 1427, ASTM
[5] M.K. Cheong, M.J. Spearpoint, C.M. Fleischmann, Design fires for vehicles in road International, 2002, pp. 147–162.
tunnels, in: Proc. 7th International Conference on Performance-Based Codes and [34] ISO, ISO 5660-1:2015 Reaction-To-Fire Tests - Heat Release, Smoke Production
Fire Safety Design Methods, 2008, pp. 229–240. and Mass Loss Rate - Part 1: Heat Release Rate (Cone Calorimeter Method)
[6] C. Mayfield, D. Hopkin, Design Fires for Fire Safety Engineering, VTT Technical and Smoke Production Rate (Dynamic Measurement), tech. rep., International
Research Centre of Finland, 2011. Organization for Standardization, 2015, 2015.
[7] B. Haavard, M. Klassen, S. Olenick, Modern vehicle hazards in parking structures [35] ISO, ISO/TR 24679-3: Fire Safety Engineering - Performance of Structure in Fire
and vehicle carriers, Fire Protect. Res. Found. (July) (2020) 60. - Part 3: Example of an Open Car Park, tech. rep., International Organization
[8] P. Sturm, L. Mellert, D. Sprakel, H. Ingason, A. Mos, C. Willmann, J. Thompson, for Standardization, 2013.
G. Clark, Impact of New Propulsion Technologies on Road Tunnel Operations [36] E. Funk, K.W. Flecknoe-Brown, T. Wijesekere, B.P. Husted, B. Andres, Fire
and Safety, PIARC, 2022. extinguishment tests of electric vehicles in an open sided enclosure, Fire Saf.
[9] E.R. Rosenbaum, SFPE Engineering Guide to Performance-Based Fire Protection, J. 141 (August) (2023) 103920.
second ed., SFPE and NFPA, Quincy, MA, 2007.
[10] SFPE, Code Official’s Guide to Performance-Based Design Review, Society of Fire
Protection Engineers, 2004.

10

You might also like