You are on page 1of 16

International Journal of Hydrogen Energy 32 (2007) 136 – 151

www.elsevier.com/locate/ijhydene

Predicting radiative heat fluxes and flammability envelopes from


unintended releases of hydrogen
William Houf ∗ , Robert Schefer
Sandia National Laboratories1 , P.O. Box 969, Livermore, CA 94551-0969, USA

Received 3 October 2005; received in revised form 1 April 2006; accepted 1 April 2006
Available online 17 July 2006

Abstract
The prediction of the radiative heat flux from a turbulent-jet flame issuing from a damaged, high-pressure hydrogen storage system is an
issue of importance for the safe use of hydrogen. Information about the variation of the thermal radiation exposure with distance from the
hydrogen jet flame as well as the length and duration of the flame is important in determining safe distances for the handling and storage of
hydrogen. An equally important issue is the determination of the concentration decay of an unignited hydrogen jet in the surrounding air, and
the envelope of locations where the concentration falls below the lower flammability limit for hydrogen.
This paper discusses and presents results from models for the radiative heat transfer from hydrogen jet flames and the concentration decay of
unignited hydrogen jets for unintended release events involving high-pressure gas storage systems and fuel dispensers. The models are based on
a combination of empirical correlations and analytical models, as well as a numerical model of the temporal blow-down of a hydrogen storage
tank. For hydrogen jet flames, predictions of the radiative heat flux are shown and compared to exposure limits documented in the literature.
For unignited hydrogen jets, concentration contours are presented and the distance to the lower flammability limit is computed. Comparisons
are made with natural gas leaks from tanks at the same pressure with the same effective leak area.
䉷 2006 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved.

Keywords: Hydrogen; Flame radiation; Flammability envelope

1. Introduction delivery pipes to larger, high-volume releases resulting from ac-


cidental breaks in the tubing from high-pressure storage tanks.
The development of an infrastructure for hydrogen utiliza- In all cases, the resulting hydrogen jet represents a potential fire
tion will require new safety codes and standards that establish hazard, and the buildup of a combustible cloud poses a hazard
guidelines for building the components of this infrastructure. if ignited downstream of the leak.
Based on a recent workshop on unintended hydrogen releases, A case in which a high-pressure leak of hydrogen is ignited
one release case of interest involves leaks from pressurized at the source is best described as a classic turbulent-jet flame,
hydrogen-handling equipment [1]. These leaks range from shown schematically in Fig. 1. The distances of importance are
small-diameter, slow-release leaks originating from holes in the radial distance from the geometrical flame centerline, r, and
the distance downstream of the jet exit, x. Other variables of
interest are the jet exit diameter, dj , and the jet exit velocity and
density, uj and j , respectively. A previous paper by Schefer
et al. [2] reported experimental measurements of large-scale
∗ Corresponding author. Tel.: +1 9252943184; fax: +1 9252943870.
hydrogen jet flames and verified that measurements of flame
E-mail address: will@sandia.gov (W. Houf).
1 Sandia is a multiprogram laboratory operated by Sandia Corporation, length, flame width, radiative heat flux, and radiant fraction are
a Lockheed Martin Company, for the United States Department of En- in agreement with non-dimensional flame correlations reported
ergy’s National Nuclear Security Administration under Contract DE-AC04- in the literature. This work verifies that such correlations can
94-AL85000. be used to predict the radiative heat flux from a wide variety of
0360-3199/$ - see front matter 䉷 2006 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijhydene.2006.04.009
W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151 137

Nomenclature
b coefficient for hydrogen in the Abel–Nobel r radial position, m
equation of state (7.691 × 10−3 m3 /kg) RH2 gas constant for hydrogen (4124.18 J/kg/K)
Btu British thermal unit Ru universal gas constant (8314.34 J/kmol/K)
C∗ non-dimensional radiant power Rmax the maximum radial position from the flame
CH4 methane centerline for the given heat flux level, m
C2 H 2 acetylene Srad the total emitted radiative power, W
C2 H 4 ethylene Tad adiabatic flame temperature of hydrogen in
C3 H 8 propane air (2390 K)
deff the effective diameter, m uj jet exit velocity, m/s
dj jet exit diameter, m ueff the effective velocity at the end of expansion,
d∗ jet momentum diameter, m m/s
Drad radiation distance, m Wf flame width, m
Frf Froude number (dimensionless parameter Wmix mean molecular weight of the products of
based on the ratio of momentum effects to stoichiometric combustion of hydrogen in air
buoyancy effects) (24.54 kg/kmol)
fs mass fraction of fuel at stoichiometric condi- x axial position, m
tions xo the virtual origin of the jet, m
g acceleration due to gravity (9.8 m/s2 ) X(Rmax ) the axial location at which the maximum heat
H2 molecular hydrogen flux level occurs, m
hr hour Xrad the radiant fraction or the fraction of the total
K the entrainment constant chemical heat release that is radiated to the
Kc the entrainment constant for a round jet surroundings
Lvis visible flame length, m Z the compressibility factor [Z = p/(RT)]
L∗ non-dimensional flame length Hc heat of combustion, J/kg
LFL lower flammability limit Tf peak flame temperature rise due to combus-
LFLDPF lower flammability limit for a downward- tion heat release, K
propagating flame  pi
LFLUPF lower flammability limit for an upward- f flame density, kg/m3
propagating flame gas the density of the exiting gas evaluated at
mfuel total fuel mass flow rate, kg/s ambient temperature and pressure, kg/m3
mfuel Hc total heat released due to chemical j jet exit density, kg/m3
reaction, W (j /∞ ) ratio of jet gas density to ambient gas density
pj the jet exit pressure, bar ∞ density of the ambient fluid, kg/m3
psupply the pressure in the supply, bar cl volume fraction (mole fraction) along the
ptank the pressure in the tank, bar centerline of the jet
p∞ the ambient pressure, bar f global flame residence time, s
qrad (x, r) the radiant heat flux measured at a particu-
lar axial location, x, and radial location, r,
W/m2

hydrogen flames. The present study builds upon this work by the hydrogen–air mixture is no longer ignitable is of inter-
incorporating the experimentally verified correlations into an est to hydrogen ignition studies. The present study develops
engineering model that predicts flame length, flame width, and an engineering model for the concentration decay of a high-
the radiative heat flux at an axial position, x, and radial distance, momentum turbulent-jet based on experimentally measured en-
r. The engineering model is then used to predict radiative heat trainment rates and similarity scaling laws for turbulent-jets.
fluxes for hydrogen flames. The model is then verified by comparing simulations for high-
For cases where the high-pressure leak of hydrogen is unig- pressure natural gas leaks with the experimental data of Birch
nited, a classic high-momentum turbulent-jet is formed that can [3] for the concentration decay of high-pressure natural gas jets.
be described using the same coordinate system shown in Fig. 1. The engineering model is then applied to hydrogen and used to
The hydrogen concentration within the jet varies with axial predict unignited jet mean (time-averaged over turbulent fluc-
and radial position due to entrainment and turbulent mixing tuations) concentration contours for high-pressure hydrogen
with the ambient air. The concentration contour beyond which leaks.
138 W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151

Fuel Srad (k
(kW)
Flame 1.2
Envelope C2H4 11.2
C2H4 20.2
CH4 12.5
1.0 CH4 6.4
C2H2 18.1
C2H2 56.5
Fit to data
0.80 Data From Large-Scale
Present H2 Tests
H2 data:
Listed Below:
d=1.905 mm
d=7.938 mm (5 sec)

C*
0.60 (10sec)
(20sec)

0.40

0.20

0.0
0.0 0.50 1.0 1.5 2.0 2.5 3.0
x/L vis

Fig. 2. Axial variation of normalized radiative heat flux [4,2].

x
0.3
CO/H2 (Turns & Myhr, 1991)
CH4 "
0.25 CH4 (d=1.91 mm)
H2 (d=1.91 mm)

Radiant Fraction
H2 (d=7.94 mm)
Jet Exit r H2 (d=7.94 mm)
0.2 H2 (d=5.08 mm)

0.15
Hydrogen
Flow 0.1

Fig. 1. Coordinate system for turbulent-jet flame. 0.05 Fit to H2 Data

0
1 10 100 1000
2. Description of engineering models Flame Residence Time (ms)

2.1. Flame radiation heat flux and flame length model Fig. 3. Radiant fraction as a function of flame residence time (lab H2 flame
data for diameters of 1.905 and 3.75 mm, large-scale H2 flame test data at
Gaseous flame radiation is the primary heat transfer mech- diameter of 7.94 mm).
anism from hydrogen flames. The flame radiation heat flux
model follows the approach of Sivathanu and Gore [4] where
ized axial distance. Fig. 2 shows typical profiles of C ∗ mea-
the flame properties of importance are the visible flame length,
sured in six different turbulent-jet flames using CH4 , C2 H2 and
Lvis , total radiative power emitted from the flame, Srad , and
C2 H4 as the fuel [4] as well as the measurements of Schefer
total heat released due to chemical reaction, mfuel Hc where
et al. [2] for large-scale H2 jet flames.
mfuel and Hc are the total fuel mass flow rate and the heat of
The use of Eq. (2) to calculate flame radiation heat flux
combustion, respectively. The radiant fraction, Xrad , is defined
levels requires knowledge of the flame radiant fraction. Turns
as the fraction of the total chemical heat release that is radiated
and Myhr [5] measured the radiant fraction from turbulent-
to the surroundings and is given by an expression of the form
jet flames using four hydrocarbon fuels with a wide variety
Xrad = Srad /mfuel Hc . (1) of sooting tendencies. These fuels included methane, ethylene,
propane, and a 57% CO/43% H2 mixture. A plot of the radiant
For turbulent-jet flames, the radiative heat flux at an axial po- fraction data from Turns and Myhr [5] along with the radiant
sition x and radial position r can be expressed in terms of the fraction data for large-scale H2 flames is shown in Fig. 3. The
non-dimensional radiant power, C ∗ , and, Srad , the total emitted radiant fraction data, Xrad , is plotted versus the global flame
radiative power. The radiative heat flux is given by an expres- residence time where the residence time is given by an expres-
sion of the form [4] sion of the form
qrad (x, r) = C ∗ Srad /4r 2 , (2) f = (f Wf2 Lvis fs )/(3j dj2 uj ), (3)
where qrad (x, r) is the radiant heat flux measured at a partic- where f , Wf , and Lvis are the flame density, width, and length,
ular axial location, x, and radial location, r. Experimental data and fs is the mass fraction of hydrogen in a stoichiomet-
further show that C ∗ may be expressed in non-dimensionalized ric mixture of hydrogen and air. For turbulent-jet flames, the
form as a function of burner diameter, flow rate and fuel type flame width, Wf , is approximately equal to 0.17Lvis [2]. This
and, for turbulent-jet flames, is dependent only on the normal- definition of residence time takes into account the actual flame
W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151 139

density and models the flame as a cone. The flame density, f , 102
is calculated from the expression f =p∞ Wmix /(Ru Tad ), where
p∞ is the ambient pressure, Wmix is the mean molecular weight
of the stoichiometric products of hydrogen combustion in air,
L*=13.5Fr2/5/(1+0.07Fr 2)1/5 L*=23
Ru is the universal gas constant, and Tad is the adiabatic flame
temperature for hydrogen. The figure suggests that for flames
with a lower sooting tendency, there is a well-defined relation-
ship between radiant fraction and global flame residence time.

L*
10
Both methane and the CO/H2 mixture show a well-behaved de-
pendence on residence time and nearly collapse onto the same
curve over the range of conditions studied. Values for the large- H2@172 bar, choked (d=7.94 mm)
H2@172 bar, unchoked (d=7.94 mm)
scale hydrogen jet flames are approximately a factor of two CH4 (d=1.91 mm)
H2 (d=1.91 mm)
lower than the hydrocarbon flames for the same flame residence CH4 Kalghatgi
time. C3H8 Kalghatgi
H2 Kalghatgi
The visible flame length, Lvis , is required for computing H2@413 bar (d=5.08 mm)

the global flame residence time, f , to determine the flame 1.0


0.1 1.0 10.0 100.0
radiant fraction. Based on an analysis of the transition from
Fr
momentum-controlled to buoyancy-controlled turbulent-jet
flame dynamics, Delichatsios [6] developed a useful corre- Fig. 4. Variation of dimensionless visible flame length with flame Froude
lation for turbulent flame lengths. The correlation is based number. Listed pressures refer to initial tank pressures for the large-scale
on a non-dimensional Froude number that measures the ratio hydrogen jet flame tests.
of buoyancy to momentum forces in jet flames. Using the
nomenclature of Turns [7] the Froude number is defined as
3/2
to reduce the hydrogen flame length measurements for plotting
uj fs in terms of L∗ in Fig. 4. The simulation also uses this same
Frf = , (4)
(j /∞ )1/4 [(Tf /T∞ )gdj ]1/2 effective diameter approach to recover the visible flame length,
Lvis , from the values of L∗ computed from Eqs. (6a) and (6b).
where uj is the jet exit velocity, fs is the mass fraction of fuel at If the jet exit velocity and density of a hydrogen flame are
stoichiometric conditions, (j /∞ ) is the ratio of jet gas density known, then Eq. (4) can be used to calculate the flame Froude
to ambient gas density, dj is the jet exit diameter, and Tf is number and Eqs. (5) and (6) can then be used to compute the
the peak flame temperature rise due to combustion heat release. visible length of the flame, Lvis . The flame width, Wf , can be
Small values of Frf correspond to buoyancy-dominated flames computed from the expression Wf =0.17Lvis and used in Eq. (3)
while large values of Frf correspond to momentum-dominated to compute the global flame residence time, f . Knowing the
flames. Note that the parameters known to control turbulent flame residence time, a curve-fit to the hydrogen radiant fraction
flame length such as jet diameter, flow rate, stoichiometry, and data in Fig. 3 can be used to determine the radiant fraction of
(j /∞ ) are included in Frf . Further, a non-dimensional flame the hydrogen flame. Knowing the radiant fraction and using a
length, L∗ , can be defined as curve-fit to the C ∗ curve shown in Fig. 2, Eq. (2) can be used
Lvis fs Lvis fs to compute the radiant heat flux from the hydrogen flame at
L∗ = = , (5) any axial position, x, and radial position r.
dj (j /∞ ) 1/2 d∗

where Lvis is the visible flame length and d ∗ is the jet momen- 2.2. Unignited jet concentration decay model
tum diameter. Fig. 4 shows the resulting correlation of flame
length data for a range of fuels (H2 , C3 H8 and CH4 ) and in- For cases where the high-pressure leak of hydrogen is unig-
let flow conditions. In the buoyancy-dominated regime, L∗ is nited, a classic high-momentum turbulent-jet is formed that can
correlated by the expression be described using the same coordinate system shown in Fig. 1.
2/5 The hydrogen concentration within the jet varies with axial po-
13.5F r f
L∗ = for Frf < 5 (6a) sition, x, and radial position, r, due to entrainment and turbulent
(1 + 0.07F r 2f )1/5 mixing with the ambient air.
The nature of the concentration field of subsonic, momentum-
and in the momentum-dominated regime by the expression
dominated incompressible turbulent free jets is well docu-
L∗ = 23 for Frf > 5. (6b) mented in the literature [8]. The decay of the mean volume
fraction, ¯ cl , (or mean mole fraction) along the centerline of
The flame length data of Schefer et al. [2] for large-scale hy- the jet is given by an expression of the form
drogen flames is shown on the plot and is found to be in good
 1/2
agreement with the L∗ correlations given by Eqs. (6a) and (6b). Kd j ∞
For choked flow conditions the concept of a notional expan- ¯ cl (x) = , (7)
x + xo gas
sion and effective source diameter (see next section) was used
140 W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151

where K is the entrainment constant, ∞ is the density of the At each axial position, x, the radial variation of the concentra-
ambient fluid, gas is the density of the exiting gas evaluated at tion is computed from the expression
ambient temperature and pressure, and xo is the virtual origin
¯ (x, r) = cl (x)e−Kc (r/(x+xo )) ,
2
of the jet [8]. (11)
For high-pressure leaks of hydrogen, the exit flow chokes
at the sonic velocity if the pressure ratio across the leak is where the value of Kc = 57 for a round jet [8]. Eqs. (8)–(11)
greater than the critical pressure ratio (approximately 1.9 for can be used to compute the concentration field from a high-
hydrogen). At pressure ratios higher than the critical value, the momentum turbulent-jet resulting from the supercritical release
exit velocity remains locally sonic. For these supercritical re- of hydrogen. For the studies performed in this paper, a value of
leases, the flow leaves the exit to form an underexpanded jet the entrainment coefficient equal to K = 5.40 [9] was used for
that quickly expands to ambient pressure through a complex the simulations. The value of the virtual origin, xo , is typically
flow structure involving one or more shocks. As a result, the a small multiple (less than 5) of the jet exit diameter and was
concentration field behaves as if it were produced by a larger set to zero for these studies in accordance with the work of
source than the actual exit diameter and the diameter of this Birch [9].
effective source is referred to as the effective diameter, deff .
The work of Birch et al. [3,9] for natural gas jets indicates that 3. Comparison of models with experimental data
the classical laws for concentration decay for turbulent-jets in
pressure equilibrium (i.e. Eq. (7)) can be applied to underex- 3.1. Flame radiation heat flux and flame length model
panded jets resulting from supercritical releases provided that
the jet exit diameter, dj , is replaced by the effective diameter The hydrogen flame radiation and flame length models were
deff . The reports of Britter [10,11] discuss various approaches compared against the large-scale hydrogen jet flame experi-
for computing effective diameter source models for underex- ments of Schefer et al. [2]. In these experiments, hydrogen
panded jets. gas was released from a “six-pack” of high-pressure cylin-
The effective source diameter model used in this work is for- ders, each connected to a central manifold with a common
mulated by considering a notional expansion [9] that conserves outlet. Typical pressure in the full cylinders was 137.9 bar
both mass and momentum while retaining the assumption that (2000 psia)–172.3 bar (2500 psia).
the pressure is reduced to ambient pressure at the end of the To obtain jet exit conditions, a network flow model of the
expansion. Based on the work of Birch et al. [9], the equation piping and high-pressure cylinders used in the experiment was
for the effective source diameter is developed using the Sandia developed Topaz code [12]. The
 1/2 network flow model considers the non-ideal gas behavior of
j uj hydrogen through an Abel–Nobel equation of state [13] of the
deff = dj , (8) form
gas ueff
RH2 T
p= , (12)
where j is the jet exit density, uj is the jet exit velocity, gas is (1 − b)
the density of the exiting gas evaluated at ambient pressure and
temperature, dj is the jet exit diameter, and ueff is the velocity where the values of RH2 = 4124.18 J/kg K and b = 7.691 ×
at the end of the expansion. The effective velocity at the end 10−3 m3 /kg were used for hydrogen. The model can also be
of the expansion is given by an expression of the form used with an ideal-gas equation of state by setting the value b
equal to zero.
ueff = uj + (pj − p∞ )/(j uj ), (9) The tank blow-down and network flow model was used to
predict the flow and pressure drop through the piping leading
where pj is the jet exit pressure and p∞ is the ambient pressure. to the jet exit. These jet exit conditions were then used with
Eqs. (8) and (9) can be used to compute the effective source the flame length and radiant fraction correlations described in
diameter for supercritical releases and are valid for real gas the previous section to predict the hydrogen jet flame charac-
as well as ideal gas models as long as the jet exit conditions teristics. Comparisons of the measured and predicted pressure
are computed properly. For hydrogen at 200 bar and 300 K, the history curves in the high-pressure cylinders were used to vali-
compressibility factor Z (where Z =p/(RT)) is approximately date the tank blow-down network flow model [2]. Simulations
1.12; at a pressure of 800 bar and the same temperature the with the network flow model indicated that significant pressure
compressibility factor is approximately 1.51. For an ideal gas, drop occurred in the piping of the experiment with the total
Z is equal to unity. pressure at the jet exit being approximately 16.4 bar (226 psig)
For supercritical releases the effective source diameter re- or a static pressure of approximately 13.6 bar (182 psig) at 0.1 s
places the jet diameter in Eq. (7) and centerline concentration into the blow-down.
decay equation becomes Fig. 5 shows a comparison of the flame length predictions
from the model with the large-scale hydrogen jet flame length
 1/2 data. Because an approximate ±10% scatter occurs in the data
Kd eff ∞
¯ cl (x) = . (10) around the L∗ correlation (see Fig. 4) used in the model, an
x + xo gas uncertainty analysis was performed where the L∗ correlation
W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151 141

6.0 7.0
Time = 10 sec
Simulation
6.0 (C*+10%,L*+10%, Xrad+10%)
5.0 Data
Data
Simulation (L*: +10%)
5.0
Simulation
Flame Length (m)

4.0

(kW/m2)
Simulation (Nominal) (Nominal)
4.0
Simulation (L*: -10%) Simulation
3.0 (C*-10%,L*-10%, Xrad-10%)

RAD
3.0

q
2.0 2.0

1.0 1.0

0.0
0.0 0 0.5 1 1.5 2 2.5 3
0 10 20 30 40 50 60
x/Lvis
Time (sec)

Fig. 5. Comparison of simulation of hydrogen visible flame length with the Fig. 7. Comparison of simulation of radiative heat flux from a hydrogen flame
hydrogen jet flame data of Schefer et al. [2]. at a radial position of r = 1.82 m with the data at 10 s into the blow-down.

10 calculations were performed with the nominal values of these


Time= 5 sec correlations, and an increase of 10% to each of the three cor-
Simulation relations (upper bound on radiative heat flux), and a decrease
8 (C*+10%,L*+10%,X
rad
+10%)
of −10% to each of the correlations (lower bound on radiative
Data heat flux). The results of these calculations are shown in Fig. 6.
An additional comparison with data using the same approach
qRAD (kW/m2)

6 Simulation is shown in Fig. 7 at a time of 10 s into the blow-down. The


(Nominal)
range of the calculations with either an increase of 10% or de-
Simulation crease of 10% in each of the correlations for L∗ , C ∗ , and Xrad
4 (C*-10%,L*-10%,Xrad-10%) are able to bound the range of experimental data adequately at
both times.

2 3.2. Unignited jet concentration decay model

There appears to be a lack of data in the literature for the


0 concentration decay of momentum-dominated, choked flow,
0 0.5 1 1.5 2 2.5 3
x/L vis
unignited turbulent hydrogen jets resulting from supercritical
releases. Hence the unignited jet model was compared with the
Fig. 6. Comparison of simulation of radiative heat flux from a hydrogen flame jet concentration decay data of Birch et al. [3] for supercritical
at a radial position of r = 1.82 m with the data at 5 s into the blow-down. releases of natural gas. Birch measured the concentration de-
cay of natural gas into air for a 2.7 mm diameter round nozzle
connected to a regulated high-pressure natural gas supply. The
was increased and then decreased by 10% from its nominal method of concentration measurement in the experiment inte-
value. Calculations are shown in Fig. 5 for the nominal L∗ grated the turbulent concentration fluctuations in the flow and
correlation, and an increase in L∗ of 10% and a decrease in L∗ resulted in a time-averaged concentration measurement at each
of −10%. Predictions from the model are found to be in good axial location. Measurements of the mean concentration level at
agreement with the measured hydrogen flame lengths. different axial positions along the jet centerline were made for
Fig. 6 shows a comparison of simulations and measured ra- supply pressures ranging from 3.5 to 71 bar. Birch found if the
diation heat flux data along the axis of a hydrogen jet flame mean concentration decay along centerline was plotted in terms
at a radial distance of 1.82 m (6 ft) from the flame centerline of the non-dimensional coordinate x/(dj (psupply /p∞ )0.5 ), then
at a time 5 s into the blow-down of the high-pressure hydro- the data collapsed onto a single curve.
gen cylinders. An approximate ±10% scatter occurs in the data Calculations with the unignited jet model discussed in the
around the L∗ correlation (see Fig. 4), the C ∗ correlation (see previous section were performed using natural gas properties
Fig. 2), and the radiant fraction correlation (see Fig. 3), Xrad . and generating jet exit conditions for a large high-pressure sup-
Hence, an uncertainty analysis was performed where model ply attached to a short round nozzle. The Topaz network flow
142 W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151

Data of Birch (1984)


Simulation (ptank = 18.25 Bar, dia = 0.794 mm)
30 Simulation (ptank = 18.25 Bar, dia = 1.158 mm)
Simulation (ptank = 207.85 Bar, dia = 0.794 mm)
Simulation (ptank = 207.85 Bar, dia = 1.158 mm)
Simulation (ptank = 207.85 Bar, dia = 1.158 mm) K+10%
25 Simulation (ptank = 207.85 Bar, dia = 1.158 mm) K-10%

Simulation (K+10%)
1/ Mole Fraction

20 Simulation (Nominal)
Data

15

10

Simulation (K-10%) Fig. 9. Simulation of radiation heat flux from a hydrogen jet flame with a
5
leak diameter of 3.175 mm and a tank pressure of 207.85 bar (3000 psig).

0
0 20 40 60 80
x/(dj(Psupply/P∞)1/2)
1035.21 bar (15,000 psig) and leak diameters in the range from
9.525 mm (3/8 inch) to 0.25 mm were suggested for analysis.
Fig. 8. Comparison of simulation of centerline concentration decay for natural
For the simulations reported in this section a storage tank
gas unignited jets with the data of Birch [3]. volume of 29.7 m3 was used based on the recommendation
of the expert panel. Calculations are reported for pressures
of 18.25 bar (250 psig), 207.85 bar (3000 psig), 518.11 bar
code with an ideal gas equation of state for natural gas was (7500 psig), and 1035.21 bar (15,000 psig) and leak diameters
used to generate jet exit conditions for this geometry. Cal- ranging between 1.587 mm (1/16 in.) and 6.35 mm (1/4 in.).
culations were performed at pressures of 18.25 bar (250 psig) Jet exit conditions were computed using the Topaz network
and 207.85 bar (3000 psig) for jet exit diameters of 0.794 and flow code with an Abel–Noble equation of state for hydro-
1.158 mm. The axial variation of the reciprocal of the mean gen to simulate a large tank of hydrogen connected to a short
concentration (1/¯cl ) on jet centerline was plotted in terms of length of tubing (3.175 mm) with a diameter equal to the di-
the non-dimensional axial coordinate, x/(dj (psupply /p∞ )0.5 ), ameter of the leak under consideration. The tank temperature
where dj is the jet exit diameter, psupply is the pressure in high- was assumed to be initially at ambient temperature (294 K)
pressure supply, and p∞ is the ambient pressure. Comparison with the end of the tubing exiting to the ambient environment
of the calculations from the model with the data of Birch [3] (1.0133 bar, 294 K). Calculations were performed for hydro-
using the nominal value of the turbulent entrainment constant gen jet flames and unignited jets with the results for radiative
(K = 5.40) is shown in Fig. 8. Based on data reported by Birch heat flux and concentration decay being reported at 1 s into the
[3,9] there appears to be approximately ±10% variation in the tank blow-down for each case. At 1 s, the tank pressure has not
value of the turbulent entrainment constant, K. Hence, in ad- changed significantly from its initial value and the radiative
dition to using the nominal value of K, calculations were per- and concentration length scales are at their largest values.
formed for the 207.85 bar 1.158 diameter nozzle by varying For the hydrogen jet flames, radiative heat flux contours were
K ±10% from the nominal value. Results of the calculations us- recorded for heat flux levels of 1577 W/m2 (500 Btu/hr ft 2 ),
ing the nominal value of K are in excellent agreement with the 4732 W/m2 (1500 Btu/hr ft 2 ), and 25,237 W/m2 (8000 Btu/
data of Birch. Moreover, the calculations at 207.8 bar, which are hr ft2 ). These heat flux levels corresponding to values listed in
well beyond the maximum pressure of 71 bar used in Birch’s the 2003 International Fire Code [16] for exposure at prop-
experiments, are found to be in excellent agreement with the erty line, exposure for employees for a maximum of 3 min,
collapsed data curve plotted in terms of x/(dj (psupply /p∞ )0.5 ). and exposure for non-combustible equipment, respectively.
The work of Ruffin et al. [14] also appears to confirm the no- Fig. 9 shows results for the radiative heat flux from a hydro-
tional expansion concentration decay model of Birch for super- gen jet flame with a tank pressure of 207.85 bar (3000 psig)
critical jets of methane and hydrogen at a pressure of 40 bar. and a leak diameter of 3.175 mm (1/8 in.). Important safety
related information recorded from the simulations includes
4. Simulation of unintended releases the maximum radial position from the flame centerline for the
given heat flux level, Rmax , the axial location at which the
4.1. Hydrogen jet flame radiation and unignited jet maximum occurs, X(Rmax ), the combination of these two dis-
concentration decay tances, Drad = (Rmax + X(Rmax )), and the visible flame length,
Lvis . Fig. 10 shows a plot of Drad and the visible flame length
Simulations for unintended releases of hydrogen were per- for various leak diameters for a tank pressure of 207.85 bar.
formed by considering a break in the tubing directly connected Also included on the plot are the upper and lower bounds for
to a large hydrogen storage container. Based on a survey of a Drad and Lvis assuming an uncertainty of ±10% in each of
panel of experts [15] familiar with current and intended uses of the values of C ∗ , L∗ , and Xrad . Fig. 11 shows a plot of Rmax
hydrogen, pressures in the range from 18.25 bar (250 psig) to and X(Rmax ) for various leak diameters for a tank pressure of
W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151 143

25
Ptank = 207.85 bar (3000 psig)
Heat Flux Level = 1577 W/m2
20 Drad
Distance (m)

15

10

5
Flame Length

Fig. 12. Simulations of concentration decay of an unignited hydrogen jet with


0 a diameter of 3.175 mm (1/8 in.) and a tank pressure of 207.85 bar (3000 psig).
1 2 3 4 5 6 7
Contour lines correspond to mole fraction levels shown in the color legend.
Diameter (mm)

Fig. 10. Simulations of hydrogen jet flame radiation from a tank at pressure
207.85 bar (3000 psig) for various diameter leaks. Results showing radiation 60
distance, Drad = (X(Rmax ) + Rmax ), for a heat flux level of 1577 W/m2 and P = 207.85 bar (3000 psig)
tank
the visible flame length. Solid lines show distances using nominal values of
C ∗ , L∗ , and Xrad . Dashed lines show upper and lower bounds for Drad and 50 2% m.f.

visible flame length with ±10% uncertainty in each of the values of C ∗ , L∗ ,


4% m.f.
and Xrad .
40 6% m.f.
Distance (m)

25 8% m.f.
30
Ptank = 207.85 bar (3000 psig)

Heat Flux Level = 1577 W/m2


20
20
Distance (m)

Rmax
15
10

10 0
1 2 3 4 5 6 7
Diameter (mm)
5
X(R max) Fig. 13. Simulations of concentration decay for a turbulent high-momentum
supercritical unignited hydrogen jet from a tank at pressure 207.85 bar
0 (3000 psig) for various diameter leaks. Results showing axial distance from
1 2 3 4 5 6 7 jet origin to the point where jet concentration reaches 2.0%, 4.0%, 6.0%,
Diameter (mm) and 8.0% mole fraction on jet centerline. Solid lines show distances using the
nominal value of the turbulent-jet entrainment constant, K = 5.40. Dashed
Fig. 11. Simulations of hydrogen jet flame radiation from a tank at pressure lines show upper and lower bounds for distances with ±10% uncertainty in
207.85 bar (3000 psig) for various diameter leaks. Results showing maximum the value of K.
radial distance from the flame centerline, Rmax , for a heat flux level of
1577 W/m2 and the axial location on centerline, X(Rmax ), where the maxi-
mum occurs. Solid lines show distances using nominal values of C ∗ , L∗ , and 3.175 mm (1/8 in.). Important safety information recorded from
Xrad . Dashed lines show upper and lower bounds for Rmax and X(Rmax )
with ±10% uncertainty in each of the values of C ∗ , L∗ , and Xrad .
the simulations is the distance from the jet exit to where the
mean concentration decays to a given concentration level on
the jet centerline. Although the generally accepted value for
207.85 bar, including the upper and lower bounds for Rmax and the upward-propagating lower flammability limit of hydrogen
X(Rmax ) assuming ±10% uncertainty in each of the values in air is 0.04 mole fraction, experimental data in the literature
of C ∗ , L∗ , and Xrad . At this pressure the value of Drad can indicate that the limit may be as high as 0.072 mole fraction
be computed to approximately ±14% to ±18% depending on for horizontal-propagating flames and 0.095 mole fraction for
the jet diameter, while the flame length can be computed to downward-propagating flames [17,18]. For the unignited hy-
approximately ±10%. drogen jet simulations, distances from the origin to jet cen-
Fig. 12 shows mole fraction contours for the simulation of terline concentration levels of 0.08, 0.06, 0.04, and 0.02 mole
the concentration decay of an unignited jet of hydrogen for a fraction were recorded, and these distances are referred to as
tank pressure of 207.85 bar (3000 psig) and a leak diameter of X8%, X6%, X6%, and X6%, respectively. Fig. 13 shows a
144 W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151

Table 1
Hydrogen jet flame radiation distances for selected leak diameters and tank pressures

Ptank (bar) dj (mm) X(Rmax ) (m) Rmax (m) Drad (m) Lvis (m) Heat flux (W/m2 )

18.25 1.00 0.35 0.10 0.45 0.55 1577


18.25 1.00 0.35 0.059 0.41 0.55 4732
18.25 1.00 0.35 0.026 0.38 0.55 25,237
18.25 2.3810 0.84 0.52 1.36 1.32 1577
18.25 2.3810 0.84 0.30 1.14 1.32 4732
18.25 2.3810 0.84 0.13 0.97 1.32 25,237
18.25 4.2333 1.49 1.59 3.09 2.35 1577
18.25 4.2333 1.49 0.92 2.41 2.35 4732
18.25 4.2333 1.49 0.39 1.89 2.35 25,237
18.25 6.35 2.24 2.90 5.14 3.52 1577
18.25 6.35 2.24 1.67 3.91 3.52 4732
18.25 6.35 2.24 0.72 2.96 3.52 25,237
207.85 1.00 1.13 0.96 2.08 1.77 1577
207.85 1.00 1.13 0.55 1.68 1.77 4732
207.85 1.00 1.13 0.24 1.36 1.77 25,237
207.85 2.3810 2.68 3.75 6.43 4.22 1577
207.85 2.3810 2.68 2.16 4.84 4.22 4732
207.85 2.3810 2.68 0.93 3.61 4.22 25,237
207.85 4.2333 4.76 7.94 12.71 7.50 1577
207.85 4.2333 4.76 4.58 9.35 7.50 4732
207.85 4.2333 4.76 1.98 6.75 7.50 25,237
207.85 6.35 7.14 13.09 20.23 11.25 1577
207.85 6.35 7.14 7.55 14.70 11.25 4732
207.85 6.35 7.14 3.27 10.42 11.25 25,237
518.11 1.00 1.68 1.91 3.60 2.65 1577
518.11 1.00 1.68 1.10 2.79 2.65 4732
518.11 1.00 1.68 0.48 2.16 2.65 25,237
518.11 2.3810 4.01 6.46 10.47 6.31 1577
518.11 2.3810 4.01 3.73 7.74 6.31 4732
518.11 2.3810 4.01 1.61 5.62 6.31 25,237
518.11 4.2333 7.13 13.27 20.40 11.23 1577
518.11 4.2333 7.13 7.66 14.79 11.23 4732
518.11 4.2333 7.13 3.31 10.45 11.23 25,237
518.11 6.35 10.69 21.58 32.28 16.84 1577
518.11 6.35 10.69 12.46 23.16 16.84 4732
518.11 6.35 10.69 5.39 16.09 16.84 25,237
1035.21 1.00 2.21 2.89 5.10 3.48 1577
1035.21 1.00 2.21 1.67 3.88 3.48 4732
1035.21 1.00 2.21 0.72 2.93 3.48 25,237
1035.21 2.3810 5.26 9.30 14.56 8.29 1577
1035.21 2.3810 5.26 5.37 10.63 8.29 4732
1035.21 2.3810 5.26 2.32 7.59 8.29 25,237
1035.21 4.2333 9.35 18.84 28.20 14.73 1577
1035.21 4.2333 9.35 10.88 20.23 14.73 4732
1035.21 4.2333 9.35 4.71 14.07 14.73 25,237
1035.21 6.35 14.03 30.46 44.49 22.08 1577
1035.21 6.35 14.03 17.58 31.61 22.08 4732
1035.21 6.35 14.03 7.61 21.64 22.08 25,237

Note: Assuming worst case of no pressure loss in tubing.

plot of unignited jet concentration decay distances for a tank shows a summary of concentration decay distances for unig-
pressure of 207.85 bar (3000 psig) for various leak diameters. nited hydrogen jets for the same tank pressures and selected
Upper and lower bounds for the concentration decay distances leak diameters using the nominal value of the entrainment
are also shown on the plot assuming a ±10% uncertainty in the constant K. Detailed plots of radiation and concentration decay
turbulent-jet entrainment constant K. distances for the range of parameters considered are presented
Table 1 shows a summary of radiation distances recorded in the Appendix.
from hydrogen jet flame simulations for tank pressures Fig. 14 shows a comparison of hydrogen jet flame radiation
of 18.25 bar (250 psig), 207.85 bar (3000 psig), 518.11 bar hazard distances with unignited jet concentration decay dis-
(7500 psig), and 1035.21 bar (15,000 psig) for selected leak di- tances for a range of tank pressures and leak diameters. Re-
ameters using the nominal values of C ∗ , L∗ , and Xrad . Table 2 sults are shown for the visible flame length and the radiation
W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151 145

Table 2
Unignited hydrogen jet concentration decay distances on jet centerline for selected leak diameters, tank pressures, and mole fractions (i.e. X2% indicates the
distance from jet origin to the point where the centerline concentration has decayed to a mean concentration of 2% mole fraction)

Ptank (bar) dj (mm) X2% (m) X4% (m) X6% (m) X8% (m)

18.25 0.25 0.67 0.33 0.22 0.16


18.25 0.50 1.34 0.67 0.44 0.33
18.25 1.00 2.67 1.34 0.89 0.67
18.25 2.3810 6.36 3.18 2.12 1.59
18.25 4.2333 11.31 5.65 3.77 2.82
18.25 6.35 16.97 8.48 5.65 4.24
207.85 0.25 2.13 1.07 0.71 0.53
207.85 0.50 4.26 2.13 1.42 1.07
207.85 1.00 8.53 4.26 2.84 2.13
207.85 2.3810 20.30 10.15 6.76 5.07
207.85 4.2333 36.10 18.05 12.03 9.02
207.85 6.35 54.13 27.06 18.04 13.53
518.11 0.25 3.19 1.59 1.06 0.80
518.11 0.50 6.38 3.19 2.13 1.60
518.11 1.00 12.77 6.38 4.25 3.19
518.11 2.3810 30.39 15.19 10.13 7.598
518.11 4.2333 54.03 27.01 18.01 13.50
518.11 6.35 81.03 40.51 27.01 20.25
1035.21 0.25 4.18 2.09 1.39 1.05
1035.21 0.50 8.37 4.18 2.79 2.09
1035.21 1.00 16.74 8.37 5.58 4.18
1035.21 2.3810 39.86 19.93 13.29 9.96
1035.21 4.2333 70.85 35.42 23.62 17.71
1035.21 6.35 106.24 53.12 35.41 26.56

10 30
Ptank = 18.25 bar (250 psig) Ptank = 207.85 bar (3000 psig)

Drad (1577 W/m2) 25 Drad (1577 W/m2)


8
Drad (4732 W/m2) Drad (4732 W/m2)
Distance (m)

Distance (m)

20
6 4% m.f. 4% m.f.

6% m.f. 15 6% m.f.

4 8% m.f. 8% m.f.
10

2
Flame Length 5 Flame Length

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

50 70
Ptank = 518.11 bar (7500 psig) Ptank = 1035.21 bar (15000 psig)

Drad (1577 W/m2)


60 Drad (1577 W/m2)
40
2
Drad (4732 W/m ) 50 Drad (4732 W/m2)
Distance (m)

Distance (m)

30 4% m.f. 4% m.f.
40
6% m.f. 6% m.f.

20 8% m.f. 30 8% m.f.

20
10
Flame Length 10 Flame Length

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

Fig. 14. Comparison of simulations of hydrogen jet flame radiation hazard distances with unignited hydrogen jet centerline concentration decay distances for
various tank pressures (18.25 bar (250 psig), 207.85 bar (3000 psig), 518.11 bar (7500 psig) and 1035.21 bar (15,000 psig)) and leak diameters. Dashed lines
show the radiation hazard distance, Drad = (X(Rmax ) + Rmax ), for radiation heat flux levels of 1577 and 4732 W/m2 and the visible flame length. Solid lines
show unignited jet concentration decay distances along jet centerline for concentration levels of 4%, 6%, and 8% mole fraction.
146 W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151

Table 3 hazard distance, Drad , for heat flux levels of 1577 and
Comparison of unignited hydrogen and natural gas jet concentration decay 4732 W/m2 . These radiation hazard distances are compared
distances on jet centerline for selected leak diameters and tank pressures
with unignited jet concentration decay distances from ori-
Ptank (bar) dj (mm) Axial distance to Axial distance to gin to jet centerline mean concentration levels of 0.08, 0.06,
5% mole fraction for 4% mole fraction for and 0.04 mole fraction. For the range of pressures stud-
natural gas (m) hydrogen (m) ied, the unignited jet concentration decay distance to the
18.25 3.175 1.19 4.24 generally accepted lower flammability limit of hydrogen
18.25 1.587 0.59 2.12 in air (0.04 mole fraction) is greater than the radiation jet
207.85 3.175 3.92 13.54 flame hazard distance (Drad ) for exposure at property line
207.85 1.587 1.96 6.77
(1577 W/m2 ).

0.2 0.2
Ptank = 18.25 bar (250 psig) Ptank = 207.85 bar (3000 psig)

UFL:Natural Gas
0.15 0.15 Hydrogen
Mole Fraction

Mole Fraction
Natural Gas
LFLDPF:Hydrogen
LFLDPF:Hydrogen
0.1 0.1
LFL:Natural Gas
LFL : Natural Gas
LFLUPF:Hydrogen LFL :Hydrogen
UPF
0.05 Hydrogen 0.05
Natural Gas

0 0
0 5 10 15 20 0 5 10 15 20
X (m) X (m)

Fig. 15. Predicted centerline concentration decay for unignited jets of natural gas and hydrogen for a 3.175 mm diameter leak at tank pressures of 18.25 bar
(250 psig) and 207.85 bar (3000 psig). Distances at which the natural gas concentration falls below the lower flammability limit (LFL) for natural gas and
the hydrogen concentration falls below the upward-propagating flame (LFLUPF ) and downward-propagating flame (LFLDPF ) lower flammability limits for
hydrogen are indicated.

20 60
P = 18.25 bar (250 psig) Ptank = 207.85 bar (3000 psig)
tank
50 2% m.f.
2% m.f.
15 4% m.f.
4% m.f.
Distance (m)

Distance (m)

40 6% m.f.
6% m.f.
8% m.f.
10 8% m.f. 30

20
5
10

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

100 120
P = 518.11 bar (7500 psig) P = 1035.21 bar (15000 psig)
tank tank

80
2% m.f. 100
2% m.f.
4% m.f.
Distance (m)

Distance (m)

80 4% m.f.
60 6% m.f.
6% m.f.
60
8% m.f. 8% m.f.
40
40
20 20

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

Fig. 16. Simulations of concentration decay for turbulent high-momentum supercritical unignited hydrogen jets from tanks at pressures from 18.25 bar (250 psig)
to 1035.21 bar (15,000 psig) for various diameter leaks. Results showing axial distance from jet origin to the point where jet concentration reaches 2.0%, 4.0%,
6.0%, and 8.0% mole fraction on jet centerline. Solid lines show distances using the nominal value of the turbulent-jet entrainment constant, K = 5.40. Dashed
lines show upper and lower bounds for distances with ±10% uncertainty in the value of K.
W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151 147

7 7
Ptank = 18.25 bar (250 psig) Ptank = 18.25 bar (250 psig)
6 Heat Flux Level = 1577 W/m2
6 Heat Flux Level = 1577 W/m2

5 Drad 5

Distance (m)

Distance (m)
4 4 Rmax

3 3

2 2

1 Flame Length 1 X(Rmax)

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

7 7
Ptank = 18.25 bar (250 psig) Ptank = 18.25 bar (250 psig)
6 Heat Flux Level = 4732 W/m2
6 Heat Flux Level = 4732 W/m
2

5 5
Distance (m)

Distance (m)
Drad
4 4
X(R )
max
3 3

2 2
Flame Length
1 1
Rmax
0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

7 7
Ptank = 18.25 bar (250 psig) Ptank = 18.25 bar (250 psig)
6 6
Heat Flux Level = 25237 W/m2 Heat Flux Level = 25237 W/m2

5 5
Distance (m)

Distance (m)

Flame Length
4 4
X(R )
max
3 3

2 2
Drad Rmax
1 1

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

Fig. 17. Simulations of hydrogen jet flame radiation from a tank at pressure 18.25 bar (250 psig) for various diameter leaks. Results showing maximum radial
distance from the flame centerline, Rmax , for a heat flux levels of 1577, 4732, and 25,237 W/m2 and the axial location on centerline, X(Rmax ), where the
maximum occurs. Solid lines show distances using nominal values of C ∗ , L∗ , and Xrad . Dashed lines show upper and lower bounds for Rmax and X(Rmax )
with ±10% uncertainty in each of the values of C ∗ , L∗ , and Xrad .

4.2. Comparison of unignited jet concentration decay to puted for the natural gas unignited jets. For the hydrogen unig-
lower flammability limit for hydrogen and natural gas nited jet, a value of 0.04 mole fraction was used as the lower
flammability limit [17] and the concentration decay distance
A limited number of simulations were performed to com- to the 0.04 mole fraction level was computed. Table 3 shows a
pare the unignited jet lengths for hydrogen and natural gas comparison of the concentration decay distances to the lower
assuming the same tank pressure and leak diameter. For the flammability limit for hydrogen and natural gas for tank pres-
natural gas simulations, natural gas properties were used in sures of 18.25 bar (250 psig) and 207.85 bar (3000 psig) for
the models and the Topaz network flow simulator assuming an leak diameters of 3.175 mm (1/8 in.) and 1.1587 mm (1/16 in.).
ideal equation of state. The generally accepted lower flamma- The jet concentration decay distance to the lower flammabil-
bility limit for natural gas in air is 0.05 mole fraction [19] ity limit for hydrogen is approximately a factor of 3.5 times
and hence the centerline concentration decay distance from jet greater than that for natural gas at the same pressure and leak
origin to the 0.05 mole fraction concentration level was com- diameter.
148 W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151

25 25
P = 207.85 bar (3000 psig) P = 207.85 bar (3000 psig)
tank tank
2 2
Heat Flux Level = 1577 W/m Heat Flux Level = 1577 W/m
20 20
Drad
Distance (m)

Distance (m)
15 15 R
max

10 10

5 5
Flame Length X(Rmax)

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

25 25
Ptank = 207.85 bar (3000 psig) Ptank = 207.85 bar (3000 psig)
2 2
20 Heat Flux Level = 4732 W/m 20 Heat Flux Level = 4732 W/m
Distance (m)

Distance (m)
D
rad
15 15
Rmax
10 10

5 5 X(R )
Flame Length max

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

25 25
P = 207.85 bar (3000 psig) P = 207.85 bar (3000 psig)
tank tank
2 2
20 Heat Flux Level = 25237 W/m Heat Flux Level = 25237 W/m
20
Distance (m)

Distance (m)

15 D
rad 15

10 10 X(R )
max
Rmax

5 Flame Length 5

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

Fig. 18. Simulations of hydrogen jet flame radiation from a tank at pressure 207.85 bar (3000 psig) for various diameter leaks. Results showing maximum radial
distance from the flame centerline, Rmax , for a heat flux levels of 1577, 4732, and 25,237 W/m2 and the axial location on centerline, X(Rmax ), where the
maximum occurs. Solid lines show distances using nominal values of C ∗ , L∗ , and Xrad . Dashed lines show upper and lower bounds for Rmax and X(Rmax )
with ±10% uncertainty in each of the values of C ∗ , L∗ , and Xrad .

An interesting observation by Swain [20] is that it is difficult studies were found to be very consistent once differences in
to ignite horizontal issuing turbulent hydrogen jets at any loca- test apparatus were accounted for. Based on this literature re-
tion where the mean (time-averaged) hydrogen concentration is view, it was concluded that the flammability limits of hydrogen
less than approximately 0.08 mole fraction. At first appearance are well established and do not need further research. How-
this result is somewhat surprising, since the generally accepted ever, a unique aspect of hydrogen is that the lower flammabil-
value for the lower flammability limit of hydrogen is 0.04 mole ity limit is significantly different for upward-, downward-, and
fraction. Several possibilities were explored to explain this dis- sideward-propagating flames. This is a buoyancy effect due to
crepancy, and an initial thought was that the flammability limits the low density of hydrogen relative to air. In contrast, the lower
for hydrogen were not sufficiently well established. To resolve flammability limits for conventional hydrocarbon fuels such as
this issue a literature search was completed on flammability methane are independent of flame propagation direction. Al-
limits of mixtures of hydrogen and air. Nearly 80 investigations though the generally accepted value for the upward-propagating
of the hydrogen flammability limits were identified between lower flammability limit for hydrogen in air is 0.04 mole frac-
1920 and 1960 and the flammability limits measured in these tion, experimental data in the literature indicate that the limit
W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151 149

40 40
P = 518.11 bar (7500 psig) P = 518.11 bar (7500 psig)
tank tank
35 2
35
Heat Flux Level = 1577 W/m Heat Flux Level = 1577 W/m2
30 D
30

Distance (m)
Distance (m)
rad
25 25
R
max
20 20
15 15
10 10
5 Flame Length 5 X(R
max
)

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

40 40
Ptank = 518.11 bar (7500 psig) Ptank = 518.11 bar (7500 psig)
35 2
35 2
Heat Flux Level = 4732 W/m Heat Flux Level = 4732 W/m
30 30

Distance (m)
Distance (m)

D
25 rad 25
20 20
Rmax
15 15
10 10
5 Flame Length 5 X(Rmax)

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

40 40
Ptank = 518.11 bar (7500 psig) Ptank = 518.11 bar (7500 psig)
35 35
Heat Flux Level = 25237 W/m2 Heat Flux Level = 25237 W/m2
30 30
Distance (m)
Distance (m)

25 Flame Length
25
20 20
X(Rmax)
15 15
Rmax
10 10
Drad
5 5
0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

Fig. 19. Simulations of hydrogen jet flame radiation from a tank at pressure 518.11 bar (7500 psig) for various diameter leaks. Results showing maximum radial
distance from the flame centerline, Rmax , for a heat flux levels of 1577, 4732, and 25,237 W/m2 and the axial location on centerline, X(Rmax ), where the
maximum occurs. Solid lines show distances using nominal values of C ∗ , L∗ , and Xrad . Dashed lines show upper and lower bounds for Rmax and X(Rmax )
with ±10% uncertainty in each of the values of C ∗ , L∗ , and Xrad .

may be as high as 0.072 for horizontal-propagating flames, and provided in Fig. 15 where the predicted centerline concentra-
between 0.085 and 0.095 for downward and spherically prop- tion decay profiles are plotted for both fuels for a 3.175 mm
agating flames [18]. It is noteworthy that this range of values diameter leak for tank pressures of 18.25 bar (250 psig) and
agrees well with the range of values for hydrogen ignition in 207.85 bar (3000 psig). Also indicated are the lower (LFL)
turbulent hydrogen jet flows observed experimentally by Swain and upper flammability limits (UFL) of natural gas (0.05
[20]. Considering the fact that flame propagation from a single and 0.15 mole fraction, respectively) and the lower flamma-
ignition point in a turbulent-jet is likely to be in a random direc- bility limits of hydrogen for upward (LFLUPF ) propagating
tion, and not exclusively in an upward direction, the range of flames (0.04 mole fraction) and downward (LFLDPF ) propa-
ignitable hydrogen concentrations observed by Swain is consis- gating flames (0.085–0.095 mole fraction). Fig. 15 shows that
tent with the range of lean hydrogen flammability limits found while the decay distance to the lower flammability limit for
in the literature. upward-propagating hydrogen flames (0.04 mole fraction) is
Considering the above discussion, a more detailed com- approximately 3.5 times greater than that for natural gas, the
parison of unignited hydrogen and natural gas jet lengths is hydrogen decay distance to the lower flammability limit for
150 W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151

60 60
Ptank = 1035.21 bar (15000 psig) Ptank = 1035.21 bar (15000 psig)
50 Heat Flux Level = 1577 W/m2 50 Heat Flux Level = 1577 W/m
2

Drad

Distance (m)
Distance (m) 40 40
Rmax
30 30

20 20

10 10
Flame Length X(Rmax)

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

60 60
Ptank = 1035.21 bar (15000 psig) Ptank = 1035.21 bar (15000 psig)
50 Heat Flux Level = 4732 W/m2 50 Heat Flux Level = 4732 W/m2

Distance (m)
Distance (m)

40 40
Drad

30 30
Rmax

20 20

10 10
Flame Length X(R )
max
0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

60 60
Ptank = 1035.21 bar (15000 psig) Ptank = 1035.21 bar (15000 psig)
50 Heat Flux Level = 25237 W/m2 50 Heat Flux Level = 25237 W/m2
Distance (m)

Distance (m)

40 40
Drad
30 30

20 20 Rmax
X(R )
max

10 10
Flame Length

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Diameter (mm) Diameter (mm)

Fig. 20. Simulations of hydrogen jet flame radiation from a tank at pressure 1035.21 bar (15,000 psig) for various diameter leaks. Results showing maximum
radial distance from the flame centerline, Rmax , for a heat flux levels of 1577, 4732, and 25,237 W/m2 and the axial location on centerline, X(Rmax ), where
the maximum occurs. Solid lines show distances using nominal values of C ∗ , L∗ , and Xrad . Dashed lines show upper and lower bounds for Rmax and
X(Rmax ) with ±10% uncertainty in each of the values of C ∗ , L∗ , and Xrad .

downward-propagating flames (0.085–0.095 mole fraction) is An uncertainty analysis of the hydrogen jet flame radiation
only a factor of 1.5 greater. model (207.85 bar case) using an uncertainty of ±10% in each
of the three experimentally measured correlations (C ∗ , L∗ ,
5. Summary and conclusions Xrad ), indicates that the radiation distance, Drad , can be com-
puted to approximately ±14% to ±18% for the jet diameters
The previous sections presented methods by which the radi- studied. The flame length can be computed to approximately
ant heat flux from hydrogen jet flames and the concentration ±10%. Assuming a ±10% uncertainty in the experimentally
decay of supercritical high-momentum unignited hydrogen jets measured turbulent-jet entrainment constant, K, an uncertainty
may be computed. If the jet exit conditions can be computed analysis of the unignited jet concentration decay model indi-
at the leak [21], then these methods can be used to compute cates that concentration decay distances can be computed to
hydrogen jet flame radiation and unignited jet concentration ±10%. Figs. 16–20 in the appendix of this paper give de-
decay based on the models. tailed plots of the hydrogen jet flame radiation hazard distances
W. Houf, R. Schefer / International Journal of Hydrogen Energy 32 (2007) 136 – 151 151

(Drad , X(Rmax ), Rmax ) and unignited jet concentration decay [2] Schefer R, Houf W, Bourne B, Colton J. Experimental measurements
distances (including upper and lower bounds) for tank pres- to characterize the thermal and radiation properties of an open-flame
sures of 18.25 bar (250 psig), 207.85 bar (3000 psig), 518.11 bar hydrogen plume. 15th annual hydrogen conference and hydrogen expo
USA, April 26–30, Los Angeles, CA; 2004.
(7500 psig), and 1035.21 bar (15,000 psig) over a range of leak [3] Birch AD, Brown DR, Dodson MG, Swaffield F. The structure and
diameters from 0.25 to 6.35 mm. concentration decay of high pressure jets of natural gas. Combust Sci
Finally, an analysis comparing the concentration decay of Technol 1984;36:249–61.
supercritical high-momentum unignited hydrogen and natural [4] Sivathanu YR, Gore JP. Combust Flame 1993;94:265–70.
gas jets indicates that the decay distance to the lower flamma- [5] Turns SR, Myhr FH. Combust Flame 1991;87:319–35.
[6] Delichatsios MA. Combust Flame 1993;92:349–64.
bility limit for the hydrogen jet (with the same tank pressure [7] Turns SR. An introduction to combustion. second ed., New York:
and effective leak diameter) is about 3.5 times greater than the McGraw-Hill; 2000.
natural gas jet when the lower flammability limit for hydrogen [8] Chen C, Rodi W. Vertical turbulent buoyant jets—a review of
based on upward-propagating flames (0.04 mol fraction) is used experimental data. Oxford: Pergamon Press; 1980.
in the analysis. If, however, the lower flammability limit for [9] Birch AD, Hughes DJ, Waffield F. Velocity decay of high pressure jets.
Combust Sci Technol 1987;52:161–71.
downward-propagating hydrogen flames is used, then the de- [10] Britter RE. Dispersion of two phase flashing releases—FLADIS field
cay distance for the hydrogen jet is only about 1.5 times greater experiment, the modelling of a pseudo-source for complex releases.
than the natural gas jet. Report FM89/2, Cambridge Environmental Research Consultants Ltd.;
There appears to be a lack of data in the literature for the December 1994.
concentration decay of momentum-dominated, choked flow, [11] Britter RE. Dispersion of two phase flashing releases—FLADIS
field experiment, a further note on modelling flashing releases.
unignited turbulent hydrogen jets resulting from supercritical Report FM89/3, Cambridge Environmental Research Consultants Ltd.;
releases. Data of this type would be useful for validating the use November 1995.
of the unignited jet concentration decay model for hydrogen. [12] Winters WS. TOPAZ—a computer code for modeling heat transfer and
fluid flow in arbitrary networks of pipes, flow branches, and vessels.
Acknowledgements SAND83-8253, Sandia National Laboratories, Livermore, CA; January
1984.
[13] Chenoweth DR. Gas-transfer analysis section H—real gas results via the
This work was supported by the US Department of Energy, van der Waals equation of state and virial expansion extensions of its
Office of Energy Efficiency and Renewable Energy, Hydrogen, limiting Abel–Nobel form. Sandia Report SAND83-8229; June 1983.
Fuel Cells and Infrastructure Technologies Program under the [14] Ruffin E, Mouilleau Y, Chaineaux J. Large scale characterization of the
Codes and Standards subprogram element managed by Patrick concentration field of supercritical jets of hydrogen and methane. J Loss
Prev Process Ind 1996;9(4):279–84.
Davis. [15] International Code Council (ICC) Ad-Hoc committee for hydrogen gas;
October 2003.
Appendix [16] 2003 International Fire Code. International Code Council, Inc.; 2002.
[17] Zebetakis MG. US Bureau of Mines. Bulletin 627; 1965.
Fig. 16 shows simulation results of unignited hydrogen jet [18] Coward HF, Jones GW. Limits of flammability of gases and vapors.
Bureau of Mines Bulletin 503; 1952.
concentration decay distances and their uncertainty for the [19] Questar gas, Material safety data sheet, natural gas; April 1998.
range of leak diameters and tank pressures studied. Figs. 17–20 [20] Swain M. Hydrogen properties testing and verification. Presented at fuel
show simulation results of hydrogen jet flame radiation dis- cell summit meeting, Coral Gables, Florida; June 17, 2004.
tances and their uncertainty for the range of leak diameters and [21] Chernicoff W, Engblom W, Schefer W, Houf W, San Marchi C.
tank pressures studied. Characterization of leaks from compressed hydrogen dispensing systems
and related components. 16th annual hydrogen conference and hydrogen
expo USA, March 28–April 1, Washington, DC; 2005.
References

[1] Schefer RW, Houf WG, Moen CD, Chan JP, Maness MA, Keller JO,
et al. Hydrogen codes and standards unintended release workshop:
workshop analysis. Workshop held December 12, 2003 at Sandia
National Laboratories, Livermore CA; 2004.

You might also like