Você está na página 1de 13

International Journal of Hydrogen Energy 32 (2007) 2081 2093

www.elsevier.com/locate/ijhydene

Characterization of high-pressure, underexpanded hydrogen-jet flames


R.W. Schefer a, , W.G. Houf a , T.C. Williams a , B. Bourne b , J. Colton b
a Combustion Research Facility, Sandia National Laboratories, Livermore, CA 94551, USA
b SRI International, 333 Ravenwood Ave., Menlo Park, CA 94025, USA

Received 30 June 2006; accepted 2 August 2006


Available online 9 October 2006

Abstract
Measurements were performed to characterize the dimensional and radiative properties of large-scale, vertical hydrogen-jet flames. This data
is relevant to the safety scenario of a sudden leak in a high-pressure hydrogen containment vessel and will provide a technological basis for
determining hazardous length scales associated with unintended hydrogen releases at storage and distribution centers. Jet flames originating from
high-pressure sources up to 413 bar (6000 psi) were studied to verify the application of correlations and scaling laws based on lower-pressure
subsonic and choked-flow jet flames. These higher pressures are expected to be typical of the pressure ranges in future hydrogen storage
vessels. At these pressures the flows exiting the jet nozzle are categorized as underexpanded jets in which the flow is choked at the jet exit.
Additionally, the gas behavior departs from that of an ideal-gas and alternate formulations for non-ideal gas must be introduced. Visible flame
emission was recorded on video to evaluate flame length and structure. Radiometer measurements allowed determination of the radiant heat flux
characteristics. The flame length results show that lower-pressure engineering correlations, based on the Froude number and a non-dimensional
flame length, also apply to releases up to 413 bar (6000 psi). Similarly, radiative heat flux characteristics of these high-pressure jet flames obey
scaling laws developed for low-pressure, smaller-scale flames and a wide variety of fuels. The results verify that such correlations can be used to
a priori predict dimensional characteristics and radiative heat flux from a wide variety of hydrogen-jet flames resulting from accidental releases.
Published by Elsevier Ltd on behalf of the International Association for Hydrogen Energy.
Keywords: Hydrogen; Turbulent jet; Combustion; Hydrogen flames

1. Introduction
The development of an infrastructure for hydrogen utilization
will require new safety codes and standards that establish guidelines for building the components of this infrastructure. Based
on a recent workshop on unintended hydrogen releases, one of
the most common release scenarios involves leaks from pressurized hydrogen-handling equipment [1]. These leaks range
from small-diameter, slow-release leaks originating from holes
in delivery pipes to larger, high-volume releases resulting from
accidental breaks in high-pressure storage tanks. In all cases,
the resulting hydrogen fuel jet represents a potential fire hazard, and the buildup of a combustible cloud poses a hazard if
ignited downstream of the leak.

Corresponding author. Tel.: +1 925 294 2681; fax: +1 924 294 2595.

E-mail address: rwsche@sandia.gov (R.W. Schefer).

A scenario in which a high-pressure leak of hydrogen is


ignited at the source is best described as a classic turbulent-jet
flame [2]. While laboratory-scale, subsonic-jet flames burning hydrocarbon fuels have been studied extensively, data for
larger-scale, subsonic and, in particular, sonic (choked) jet
flames is less available. In a previous study by the present authors [3], measurements were performed in large-scale, vertical
flames to characterize the dimensional, thermal, and radiative properties of ignited hydrogen jets. Jet flames originating
from a high-pressure source up to about 172 bar (2500 psi)
were studied. The results showed that measured flame lengths,
for a wide range of operating conditions, collapsed onto the
same curve when plotted as a function of Froude number,
which measures the relative importance of jet momentum and
buoyancy. Furthermore, the good comparison with hydrocarbon jet flame lengths demonstrated that the non-dimensional
correlations are valid for a variety of fuel types. The radiative heat flux measurements for hydrogen flames also showed

0360-3199/$ - see front matter Published by Elsevier Ltd on behalf of the International Association for Hydrogen Energy.
doi:10.1016/j.ijhydene.2006.08.037

2082

R.W. Schefer et al. / International Journal of Hydrogen Energy 32 (2007) 2081 2093

good agreement with non-dimensional correlations and scaling laws developed for a range of fuels and flame conditions.
A well-behaved linear dependence of radiative fraction on the
log of the flame residence time were found, in agreement with
non-sooting hydrocarbon flame data, but the radiative fraction
for the H2 flames at a fixed residence time is nearly a factor of two lower. The results verified that such correlations
can be used to predict radiative heat flux from a wide variety of hydrogen flames and established a basis for predicting
a priori the characteristics of flames resulting from accidental
releases.
The objective of the present investigation is to extend these
previous measurements toward the higher-pressure ranges expected in future hydrogen storage vessels. Thus, measurements
were obtained at storage pressures up to 413 bar (6000 psi) in
this study. This pressure range is also of interest because departures from ideal-gas behavior become important. In the following section, the results of previous studies for relevant flame
characteristics will be summarized. In particular, correlations
for flame length and radiative heat flux from jet flames will be
presented, the concept of underexpanded jets will be briefly introduced, and equations based on a simplified model of these
flows will be described.

Froude number is defined as


3/2

Frf =

ue fs

(qe /q )1/4 [(DTf /T )gd j ]1/2

where ue is the jet exit velocity, fs is the mass fraction of fuel at


stoichiometric conditions, (qe /q ) is the ratio of jet gas density
to ambient gas density, dj is the jet exit diameter, and DTf is
the peak flame temperature rise due to combustion heat release.
Small values of Frf correspond to buoyancy-dominated flames
while large values of Frf correspond to momentum-dominated
flames. Further, a non-dimensional flame length, L , can be
defined as
L =

Lf fs
dj (qe /q )1/2

Lf fs
,
d

2.1. Flame length


Studies in the literature have defined a flame length based
on the visible flame length, which is determined using either
visual observation [4] or photographs of the visible flame emission [5]. Alternative definitions of flame length can be based
on infrared flame emission or ultraviolet flame emission [3],
which give flame lengths that are 12% longer and 22% shorter,
respectively, than the visible flame length. Since most correlations presented in the literature are based on visible flame
length, we will confine the present results to a flame length
based on visible emission.
Becker and Liang [4] obtained extensive experimental data
comparing flame lengths of hydrogen, carbon monoxide,
methane, ethane, ethylene and propane. These data, combined
with results in the literature, were found to correlate well over
a range of operating conditions extending from the forcedto natural-convection limits when using the appropriate nondimensional parameters. Kalghatgi [5] studied jet flames of
hydrogen, methane, propane and ethylene. In these studies,
the flame length data was extended from subsonic jet flows
to underexpanded sonic jets. Using the non-dimensional parameters of Becker and Liang, the data were found to collapse
onto a single curve. However, in the forced convective limit,
fitting constants determined by Kalghatgi differed from those
of Becker and Liang.
A correlation for flame length developed by Delichatsios [6]
is based on a non-dimensional Froude number that measures
the ratio of buoyancy-to-momentum forces in jet flames. The

(2)

where Lf is the visible flame length and d is the jet momentum diameter (=dj (qe /q )1/2 ). In the buoyancy-dominated
regime, L is correlated by the expression
2/5

L =

13.5Frf

(1 + 0.07Fr2f )1/5

for Frf < 5

(3a)

and in the momentum-dominated regime by the expression


L = 23 for Fr > 5.

2. Theory

(1)

(3b)

It has been found that turbulent flame lengths are wellcorrelated over a large range of flow conditions using these
non-dimensional parameters. Recent results by Schefer et al.
[3] showed that this correlation works well for vertical turbulent hydrogen jets, both subsonic and choked, originating from
sources at pressures up to 172 bar (2500 psi).
2.2. Flame radiation
The characterization of radiative heat flux is integral to the
development of safety codes and standards. In fuel-rich hydrocarbon flames where significant amounts of sooty particles are
formed, radiation from soot dominates the radiative heat flux.
In hydrogen flames, gaseous emission accounts for nearly all
the radiative heat flux, with excited-state H2 O* molecules being the only significant source of radiative emission. A useful
quantity to characterize the radiative heat flux from flames is
the radiant fraction, Xrad , which is defined as the fraction of the
total chemical heat release that is radiated to the surroundings:
Xrad =

Srad
,
mfuel DHc

(4)

where Srad is the total radiative power emitted from the flame,
mfuel is the total fuel mass flow rate, DHc is the heat of combustion and mfuel DHc is the total heat released due to chemical reaction. For turbulent-jet flames, the radiative power can
be expressed in terms of a non-dimensional radiant power, C ,
given by the general expression
C (x/L, r/L) =

4pR 2 qrad (x/Lf , r/Lf )


,
Srad

(5)

R.W. Schefer et al. / International Journal of Hydrogen Energy 32 (2007) 2081 2093

where R is the radial distance from the flame centerline to the


location at which the radiant flux is measured and qrad (x, r)
is the radiant heat flux measured at a particular axial location,
x, and radial location, r. Experimental data further show that
C may be expressed in non-dimensional form as a function
of burner diameter, flow rate and fuel type and, for turbulentjet flames, is dependent only on the normalized axial distance.
Under these conditions Eq. (5) reduces to
C (x/L) =

4pR 2 q

4pR 2 q

rad (x/Lf )
rad (x/Lf )
=
.
Srad
Xrad mfuel DHc

(6)

The form of Eq. (6) was verified by Sivathanu and Gore [7] in
laboratory-scale turbulent-jet flames over a range of conditions.
Previous measurements by Schefer et al. [3] verified the validity of these heat flux scaling relations in larger-scale choked
hydrogen-jet flames at pressures up to 172 bar (2500 psi).
Turns and Myhr [8] showed that the flame radiative fraction
correlated well with the flame residence time for a wide variety
of fuels. These fuels included methane, ethylene, propane and a
57% CO/43% H2 mixture. The flame residence time as defined
by Turns and Myhr is given by the expression
sf =

qf Wf2 Lf fs
3qf df2 uf

(7)

where qf , Wf , and Lf are the flame density, width and length,


respectively. This definition of residence time takes into account
the actual flame density and models the flame as a cone. For
turbulent-jet flames, Wf is approximately equal to 0.17 Lf in
hydrocarbon [8] and hydrogen-jet flames [3]. Verification that
the high-pressure, high-velocity jet flames studied here obey
these scaling laws would enable us to a priori quantify the heatflux characteristics of large-scale jet flames.
2.3. Underexpanded jets
For subsonic jets issuing into ambient air, the exit velocity
corresponds to a Mach number less than unity and the pressure
at the exit plane is just equal to atmospheric. For jets involving
a high-pressure storage device, the pressure drop across the jet
is often sufficient to result in an underexpanded jet in which the
flow is choked at the jet exit and the exit pressure is considerably
greater than atmospheric. The flow will then rapidly expand to
atmospheric pressure through a series of expansion shocks. A
convenient simplification of this expansion process involves the
concept of a notional, or fictional, nozzle. Kalghatgi [5]
used the notional nozzle concept to extend flame-length correlations developed for subsonic jet flames to underexpanded jet
flames. Birch et al. [9,10] further developed this concept into
a pseudo-diameter, or an effective source diameter, that was
used to explain the concentration and velocity field decays in
non-reacting, high-pressure jet flows.
The concept of a notional nozzle as it relates to a highpressure release through a jet is illustrated in Fig. 1. Level 1
corresponds to the high-pressure source, Level 2 is the jet exit
and Level 3 corresponds to the exit conditions after expansion
through the notional nozzle. This situation is equivalent to re-

2083

V3
deff
Level 3
P3, T3
V2

Notional Nozzle
Expansion Region

P2, T2

Level 2
d2
P1, T1
Level 1

High-pressure
Reservoir

Fig. 1. Notional nozzle concept. (Birch et al. [10].)

placing the expansion process that occurs as the gas, which is


choked at the jet exit, expands to ambient conditions. The diameter of the notional nozzle, deff , is larger than the jet exit
diameter, d2 (=dj ). Assuming no viscous forces and a uniform
velocity profile across the notional nozzle cross section, the
conservation of mass and momentum are written as follows:
q2 A2 u2 = q3 A3 u3

Conservation of Mass

(8)

and
q2 A2 u22 q3 A3 u23 = A2 (P3 P2 )
Conservation of Momentum,

(9)

where A is the cross-sectional area. Solving Eq. (8) for q3 A3 ,


substituting into Eq. (9) and rearranging gives
u3 = u 2

(P3 P2 )
.
q2 u2

(10)

Also, from Eq. (8), the notional nozzle diameter is given by


r
r
q2 u 2
q2 u 2
= dj
.
(11)
deff = d2
q3 u 3
q3 u 3
This approach is entirely analogous to that of Birch et al. [10]
in which both the conservation of mass and momentum were
used to develop an expression for the notional nozzle diameter.
Conditions at Level 1, the high-pressure source, are typically
known quantities (in the present experiments they are measured). Following the approach of Birch and coworkers, conditions at the jet exit, where the flow is sonic, can be determined
from Level 1 conditions using isentropic flow relations [11].
Since expansion to ambient conditions occurs through the notional nozzle, the pressure P3 is equal to the ambient pressure.
The velocity u3 can then be calculated from Eq. (10) and the
notional nozzle diameter, deff , is given by Eq. (11). The notional nozzle diameter and flow properties at the notional nozzle exit are then used in the correlations for the flame length
(Eqs. (1)(3)) and the flame residence time (Eq. (7)).
The remainder of this paper highlights the experimental system, measurement techniques, and presents results for flame

2084

R.W. Schefer et al. / International Journal of Hydrogen Energy 32 (2007) 2081 2093

tests using a 5.08-mm-diameter hydrogen (H2 ) jet at pressures


up to 413 bar (6000 psi). Comparisons are made to measurements obtained in laboratory-scale hydrogen flames and to jet
flames for a variety of fuels found in the literature. These comparisons will highlight the applicability of jet scaling laws and
similarity variables to the present H2 jets and identify any differences between hydrocarbon and hydrogen-fuel jets.

3. Experimental methods
3.1. Flow system
The tests were carried out at the SRI International Corral
Hollow test facility in Tracy, CA. A schematic of the flow delivery system is shown in Fig. 2. The hydrogen was provided
by a storage tube trailer supplied by Air Products. The trailer
consisted of eight high-pressure hydrogen storage tubes. The
volume of each tube was 617 l (21.8 ft 3 ) at a nominal pressure of 431 bar (6258 psig) (actual initial pressures for the three
tests conducted were 431 (6253 psi), 425 (6173 psi) and 438 bar
(6349 psi)). These pressures are considerably higher than the
previous release tests where the initial storage pressure was
about 172 bar (2500 psi) [3]. In previous tests, the hydrogen
was provided by a six-pack of conventional gas cylinders,
each with a volume of 49 l. Typically only two cylinders were
used during each blowdown test, which limited the test duration to about 100 s. The higher-pressure storage tubes in the
current tests are more representative of expected future hydrogen storage pressures. Two storage tubes were used for each
of three flame tests, with the remaining six tubes closed to the
manifold. During a typical tank blowdown test, the pressure
decreased from its initial value to near atmospheric pressure at
an exponential decay rate over a period of approximately 500 s.

It should be noted that the conventional hardware in the


commercially available tube trailer was modified to remove
several flow restrictions that would have limited the flow
throughput. Thus, the piping transporting the hydrogen out of
the trailer was maintained at a minimum inner diameter (ID) of
5.08 mm. The hydrogen was delivered to a stagnation chamber
by two lengths of 1.9 and 1.75 cm ID stainless steel tubing.
The stagnation chamber, located just prior to the jet exit, was
29.2 cm in length by 15.2 cm inside diameter and was sized
to maintain a low flow Mach number (less than 1 103 ). At
this Mach number, the measured pressure and temperature in
the stagnation chamber were in excellent agreement with the
true stagnation conditions. Both the stagnation chamber pressure and temperature histories were measured for the duration
of each test. The temperature was measured with a type-K
thermocouple, while the pressure was measured using a
piezoresistive pressure transducer. Both voltage outputs were
digitized at a 500 Hz rate and stored on a Nicklet-digitized
storage scope for post-processing. The jet exit conditions
could then be calculated assuming isentropic expansion (see
Section 3.4) between the stagnation chamber and the vertically orientated, 5.08-mm diameter jet exit. The stagnation
chamber was designed into the present experimental setup to
provide a more controlled, well-defined flow into the jet. In
the previous 172 bar (2500 psi) tests, no stagnation chamber
was used and the jet exit conditions were calculated using
the Sandia Topaz network flow model [1214] to predict the
flow through the piping leading up to the jet exit. Also note
that the jet diameter in the present experiments was 5.08 mm,
which is smaller than the 7.94-mm jet diameter of the previous
tests. This smaller diameter was selected so that the 5.08-mm
diameter pipe leading out of the trailer could supply sufficient
hydrogen flow to maintain a high pressure in the stagnation
chamber.

413 bar (6000 PSI) Test Setup

29.2 cm

Jet

0.5 cm ID
Stagnagtion
Chamber
0.79 cm ID
1.9 cm ID
1.75 cm ID
x 3 m length x 8.2 m length
H2 storage tubes: 2 Banks of 4 tubes
connected to 0.5 cm ID tubing .
Volume of each tube is 21.8 cubic feet.

15.2 cm ID
dj= 0.508 cm ID
4.44 cm
0.508 cm
0.79 cm ID
Jet Nozzle
Details

Fig. 2. Schematic of experimental flow delivery system for 413 bar (6000 psi) tests.

R.W. Schefer et al. / International Journal of Hydrogen Energy 32 (2007) 2081 2093

2085

Table 1
Flame conditions
Flame

dJ (mm)

Lab flame: H2

1.91
1.91
1.91
1.91
1.91

Lab flame: CH4

1.91
1.91
1.91
1.91

SRI flame: H2
t = 10 s
t = 50 s
t = 100 s
t = 200 s
t = 300 s
t = 400 s

5.08
5.08
5.08
5.08
5.08
5.08
5.08

a Dh

Q (slm)

m (g/s)

mDh1c a (kW)

87.7
116.3
174.5
261.6
349.0

15
20
30
45
60

0.021
0.028
0.042
0.062
0.083

2.64
3.34
5.01
7.52
10.0

58.2
87.3
98.9
116.6

10
15
17
20

0.110
0.166
0.188
0.222

5.54
8.32
9.42
11.08

2.57 105
1.63 105
1.02 105
4.59 104
2.02 104
8.02 103

359.3
228.6
142.6
64.2
28.2
11.2

43 105
27 419
17 112
7705
3385
1344

uJ (m/s)

1140
1079
1056
1052
1059
1067

= 118 830 kJ/kg for hydrogen heat of combustion; 50 016 kJ/kg for methane.

Results of the present 5.08-mm-diameter jet flame tests were


compared with measurements obtained in a laboratory-scale,
1.91-mm-diameter H2 jet flame in which measurements were
obtained as part of the present study. In addition to a H2 jet,
methane was also used as a fuel in the laboratory flame to verify the measured difference in radiative heat flux from CH4
and H2 flames. These measurements removed the effect of
any differences in jet geometry and radiometer calibration on
the measurements and facilitated a more direct comparison of
the radiative properties of the two fuels. Shown in Table 1 are
the flow conditions associated with each of these flames. The
times shown in the table for the 5.08-mm-diameter flame correspond to different times in the cylinder blowdown tests.
3.2. Flame length measurements
Digital video images of the flame were obtained to characterize the flame structure and length. The visible flame images
were recorded using two Sony Model DCR TRV27 video cameras. The cameras here were located approximately 14 m from
the jet centerline and located vertically so that the two cameras covered a field of view extending from the jet exit (x = 0)
to 15 m in the vertical direction. The fields of view were overlapped to provide a continuous view of the entire flame length.
The images were stored at a standard 30 fps video frame rate,
which is not of sufficient temporal resolution to follow the
flame movement. In addition, the individual frame exposure
time of 33 ms is insufficient to capture the instantaneous flame
structure, which is averaged over this exposure time. Multiple images were averaged together at selected times to provide
information on the time-averaged flame properties and provide quantitative data on the relevant flame length scales. Because of the relatively weak flame luminescence, all tests were
run at night to eliminate background light and improve flame
visibility.

3.3. Radiative flux measurements


Similar to previous tests [3] heat flux measurements were
obtained using Medtherm Model 64P-1-22 SchmidtBoelter
thermopile detectors with a 150 view angle. A zinc selenide
(ZnSe) window on the face of the radiometer has 70% transmission between 0.7 and 17 lm. Three separate tests were performed to maximize the amount of data obtained. In the first
two tests our procedure followed the method of Sivathanu and
Gore [7], in which a thermopile detector was placed at a radial
distance, R, of half the visible flame length, Lf /2, at a selected
time during the test. Using a previously developed model for
flame length variation during the tank blowdown [15], the flame
length was estimated to be 9.1 m at a time of about 20 s after
the initiation of a test. Thus, radiometers were placed at a radial
distance, R, of 4.57 m from the jet centerline and at axial increments of 1.52 m along the flame length. The radiometer locations for tests 1 and 2 are indicated to the left side of the flame
shown schematically in Fig. 3. These measurements enabled us
to verify whether the scaling laws for radiative heat flux profiles along the flame length at the higher tank pressures of the
present investigation matched results measured at lower pressures. Four additional radiometers were also placed in the jet
exit plane at 0.3 m increments in the radial direction and aimed
in the direction of the jet flow to record the radial component
of the heat flux. Two tests were performed with this radiometer
orientation to verify the reproducibility of the data sets.
Once these scaling relationships were verified for the current
tests and the flame length history was established, a third test
was performed in which the six radiometers were placed along
an approximately 45 line from the jet centerline and originating from the jet exit. These radiometer placements are shown
on the right side of the flame in Fig. 3. Since the flame length
varied with time during each tank blowdown, these radiometer locations resulted in six times during blowdown where the

2086

R.W. Schefer et al. / International Journal of Hydrogen Energy 32 (2007) 2081 2093
Table 2
Compressibility factors for hydrogen at 300 K

Radiometer
Locations for
Tests 1 & 2

x=9.14 m
Flame
Tip

Luminous
flame zone

x=7.62 m

Radiometer
Locations for
Test 3

x=6.09 m

x=r=4.57 m

x=4.57 m

,
x=3.04 m

x=1.52 m

x=r=3.04 m

Ambient
Still Air

x=r=1.52 m
x

x=r=0.76 m

r=4.57 m
r

P (bar)

P (psi)

68
172
344
689
1034

1000
2500
5000
10 000
15 000

1.04
1.10
1.23
1.43
1.65

ditions at Level 1 with those at the jet exit, Level 2. In the


underexpanded jet case, the isentropic relations are again used
to relate the conditions at Level 2 to Level 1. In addition, Eqs.
(10) and (11) for underexpanded jets are used to calculate conditions at the notional nozzle exit.
The application of isentropic flow relations to tank blowdown
is straightforward at lower pressures where ideal-gas behavior
is valid. In fact, at pressures up to about 172 bar (2500 psi), the
gas can be treated to a good approximation as ideal. However, at
higher pressures the gas behavior increasingly departs from that
of an ideal-gas and the relations must be modified to account
for gas compressibility. Equations of state such as the twoconstant van der Waals equation and the BeattieBridgeman
equation with five constants have been used to describe real-gas
behavior [16]. It has also been shown that real-gas behavior of
hydrogen in the present context of a sudden gas expansion from
a high-pressure reservoir can be adequately described through
an AbelNobel equation of state of the form
P=

r=0.3 m
H2

Fig. 3. Experimental setup and coordinate system for turbulent-jet flame.

radial and axial location of one radiometer are both equal to half
the visible flame length. Thus, the measured value of the radiative heat flux at those locations, qrad (x/Lf = 0.5, r/Lf = 0.5),
and the known value of C (x/Lf = 0.5, r/Lf = 0.5) could be
used to determine the total radiative power emitted from the
flame, Srad , from Eq. (5), The radiant fraction, Xrad , could then
be determined from the known value of the heat release due
to chemical reaction, mfuel DHc using Eq. (4). This approach is
similar to that used by Turns and Myhr [8].

(12)

where Z = 1/(1 bq) is the compressibility factor, b is the


co-volume constant, and RH2 is the gas constant for (RH2 =
4.12 N m/gm K). The value of Z for an ideal-gas is unity (b = 0
for an ideal-gas) and values different from unity indicate departures from ideal-gas behavior. Shown in Table 2 are values of Z
for hydrogen at a temperature of 300 K and various pressures.
The pressure range shown is indicative of expected future hydrogen storage pressures. It can be seen that departures from
ideal-gas behavior do indeed become significant at pressures
above 172 bar (2500 psi).
Rearranging the AbelNobel equation of state, Eq. (12), the
gas density at the stagnation temperature and pressure, T0 and
P0 , is given by
q0 =

3.4. Jet exit conditions


Stagnation chamber temperature and pressure were measured
directly during tank blowdown. The non-dimensional flame
length, L , and Froude number, Fr, used to correlate the flame
length (Eqs. (1)(3)) and the flame residence time (Eq. (7)) require conditions at the jet exit (Level 2 in Fig. 1) if the flow is
subsonic, and at the notional nozzle exit (Level 3) if the flow
is choked and the jet is underexpanded. In the case of subsonic
flow, isentropic relations can be used to relate the known con-

qRH2 T
= ZqRH2 T ,
(1 bq)

P0
,
P0 b + RH2 T0

(13)

where b is 7.691 103 m3 /kg for hydrogen. It can further


be shown for the isentropic flow of an AbelNoble gas from a
high-pressure stagnation state to the jet exit that gas density in
the stagnation reservoir and at the jet exit are related by


c
q0
=
1 bq0

qj
1 bqj

!c "

1+

c1
2(1 bqj )2

Mj2

#c/(c1)

(14)

R.W. Schefer et al. / International Journal of Hydrogen Energy 32 (2007) 2081 2093

where M is the Mach number, c is the specific heat ratio


and the subscripts 0 and j refer to stagnation conditions and
conditions at the jet exit, respectively. Assuming choked (sonic)
flow at the jet exit, then Mj = 1 and Eq. (14) reduces to


!c "
!#c/(c1)
c
qj
c1
q0
1+
.
=
1bq0
1bqj
2(1bqj )2

(15)

The left-hand side is known from the measured stagnation tank


conditions through Eq. (13) so that Eq. (15) can be solved to
obtain the jet exit density, qj . Further, for isentropic flow of an
AbelNoble gas the ratio of the temperature in the stagnation
tank to the jet exit temperature is given by
!
T0
c1
Mj2
=1+
(16)
Tj
2(1 bqj )2
which for choked flow at the jet exit becomes
Tj =

T0
.
1 + (c 1)/2(1 bqj )2
%

(17)

Thus, knowing the measured stagnation tank temperature and


the jet exit density from Eq. (15), Eq. (17) can be used to
calculate the jet exit temperature. Now, from the AbleNoble
equation of state the pressure at the jet exit can be obtained
from
Pj =

qj RT j
1 bqj

(18)

Finally, the sonic velocity, a, of an AbleNoble gas at the jet


exit can be determined from the known jet exit temperature,
the jet exit density, and the equation
uj = a =

p
1
cRT j .
(1 bqj )

(19)

4. Experimental results
4.1. Tank blowdown history
Fig. 4 shows the measured stagnation chamber pressure and
temperature during the tank blowdown. From Fig. 4a it can be
seen that the initial storage tube pressure of 431 bar (6258 psi)
decreases to near atmospheric pressure at the end of the test.
The double exponential equation given in the figure provides a
best fit to the data. The total blowdown time for each test was
approximately 500600 s (for a two-tube blowdown). The stagnation temperature decreased rapidly to a minimum of about
45 C at about 140 s into the test before increasing due to heat
transfer from the surroundings. A best fit to the temperature
data is also given in the figure.
An interesting aspect of the stagnation chamber data is seen
during early times in the blowdown. Shown in Fig. 4b is an
expanded view of the temperature and pressure profiles corresponding to the first 10 s after the test was initiated. The stagnation pressure undergoes a rapid decrease over the first 0.5 s,

2087

followed by a more gradual rate of decrease for times greater


than about 1 s. The stagnation temperature shows a similar initially rapid decrease during the first 0.5 s, passes through a
minimum between 0.5 and 1 s into the release, followed by a
broader maximum at about 3 s. A gradual decrease then occurs
until 140 s into the release (see Fig. 4) where a second minimum is located.
The cause for the more complex behavior at short release
times is unclear, but it can be speculated that the initially rapid
pressure decrease might be due to the longer time constant
associated with flow entering the stagnation chamber than that
associated with flow exiting the chamber through the relatively
short pipe leading to the jet exit. This scenario is based on the
sequence of operations that occurs before actual testing. Prior
to the initiation of a test, solenoid valves located at the exit
of the tube trailer and immediately upstream of the jet exit
are closed and residual gas in all lines is evacuated to assure
that the test is started with pure hydrogen in the system. Next
the solenoid valve at the exit of the tube trailer is remotely
opened to allow hydrogen to enter the system. Since the valve
between the stagnation chamber and the jet exit is closed, the
pressure equalizes with pure hydrogen throughout the system.
Thus, a constant initial pressure is attained in the hydrogen
storage tubes, all piping and in the stagnation chamber. To
initiate a test, the solenoid valve downstream of the stagnation
chamber is remotely opened and the pressurized hydrogen is
released through the jet exit. Due to the small volume of piping
after the stagnation chamber, it is speculated that the initial
flow of hydrogen from the stagnation chamber occurs rapidly,
while the large volume of piping between the storage tubes
and the stagnation chamber requires a longer time to provide
an equal amount of hydrogen to the stagnation chamber. This
imbalance results in an initially rapid decrease in stagnation
chamber pressure before the hydrogen can be replenished at a
rate equal to that being released.
Shown in Fig. 5 is a comparison of measured stagnation
chamber pressure and temperature and the predicted behavior using the Topaz network flow model. Several heat transfer boundary conditions were tried in the model. In the initial
calculations, an adiabatic system was assumed in which zero
heat transfer to the high-pressure reservoir (H2 storage tubes),
the stagnation chamber and the piping occurred. The results
(curve labeled adiabatic) showed a more rapid decrease in stagnation chamber pressure than measured and a stagnation chamber temperature significantly below the experimental value.
Subsequent calculations assuming heat transfer to a constant
temperature wall showed that heat transfer to the reservoir dominated the predicted temperature history and that the calculation
was insensitive to whether adiabatic or non-zero heat transfer
was imposed on the stagnation chamber and piping. Thus, predicted histories are shown in Fig. 5 assuming full heat transfer
to the reservoir, stagnation chamber and piping at a constant
wall temperatures of 289 and 267 K. The predicted pressure
and temperature histories show good agreement with measured
results for both values of wall temperature. The agreement
is not exact because the actual wall temperature likely varies
in time and is dependent on the thermal mass of the walls.

2088

R.W. Schefer et al. / International Journal of Hydrogen Energy 32 (2007) 2081 2093

6000

20

6500

20
P=-45.2173+3141.95*exp(-0.0076595*t)
+1635.99*exp(-0.0379651*t)
Pstag(psig)

10

P_fit (psig)

5000

15

6000

10

T=-51.375+0.035753*t+34.713*exp(-0.022375*t)
+23.426*exp(-0.023006*t)

Pstag (psig)

3000

Tstag (C)

Pstag(psig)

-10
-20

Tstag(C)
T_fit (C)

2000

5
5000
0
4500
-5

-30
1000

4000

-40

0
0

100

200

300

400

-10
-15

3500

-50
500

Time (s)

(a)

Tstag (C)

5500

4000

10

Time (sec)

(b)

Fig. 4. (a) Stagnation chamber pressure and temperature tank blowdown history for initial tube pressure of 431 bar (6258 psig). (b) Stagnation chamber pressure
and temperature tank blowdown history during initial 10 s of blowdown. Initial tube pressure of 431 bar (6258 psig).

Measured
Full Heat Transfer, Tw=289 K
Full Heat Transfer, Tw=267 K
Adiabatic

6000

300
Twall = 289 K
250

5000

200

Twall = 267 K

3000

150

Adiabatic

2000

100

1000

50

Tstag (K)

Pstag (psig)

4000

0
0

100

200

300

400

500

Time (s)
Fig. 5. Comparison of experimentally measured and Topaz predicted stagnation chamber pressure and temperature during tank blowdown for initial tube
pressure of 431 bar (6258 psig).

R.W. Schefer et al. / International Journal of Hydrogen Energy 32 (2007) 2081 2093

7000
Pstag
Pjet
Pjet
Pstag
Pstore
Pjet

6000

Pressure (psi)

5000
Topaz Storage Tube
4000

Topaz Stagnation Chamber


Measured Stagnation Chamber

3000

Isentropic Ideal GasJet Exit


2000

Isentropic Abel-Noble Jet Exit


Topaz Jet Exit

1000

0
0

50

100

150

200

250

300

Time (sec)
Fig. 6. Comparison of experimentally measured pressure history and Topaz
predictions during tank blowdown. Initial tank pressure 431 bar (6258 psig).

2089

gas pressure is highest. For times greater than about 40 s the


two assumptions show good agreement and indicate ideal-gas
behavior is a good approximation. The predicted jet exit
pressure using the Topaz model is generally lower than the
isentropic expansion curves, reflecting the lower stagnation
chamber pressures predicted by the model relative to the experimentally measured values. However, Topaz predictions
assuming an non-isentropic expansion between the stagnation
chamber and jet exit show good agreement with isentropic expansion results, indicating the isentropic expansion assumption
used to reduce the data is a good approximation.
Fig. 7 shows a comparison of calculated jet exit conditions
based on isentropic expansion laws assuming an ideal-gas and
a gas that obeys the AbelNoble equation of state. As expected,
the maximum differences occur at the shortest blowdown times
where the pressures are highest. For example, at a time of
2 s into blowdown the stagnation chamber pressure is 312 bar
(4534 psi). The corresponding calculated jet exit properties for
an ideal gas and an AbelNoble gas are given in Table 3 at a
time of 2 s. The differences at greater times become increasingly
small as the pressure drops.
4.2. Flame length

In contrast, the Topaz network flow calculation assumes a constant wall temperature during the entire tank blowdown.
Fig. 6 shows the pressure-history curves in the hydrogen
storage tube, the stagnation chamber and at the jet exit. The
measured stagnation pressure history is shown, while Topaz
predicted curves for the storage tube, stagnation chamber and
jet exit are shown for comparison. The curves labeled isentropic
ideal gas jet exit and AbelNoble jet exit were calculated using
the measured values of stagnation chamber pressure and the
isentropic expansion of either an ideal-gas or a non-ideal gas
that obeys the AbelNoble equation of state. Note that the Topaz
model accounts for frictional losses and heat transfer in the
piping during expansion and thus does not assume an isentropic
expansion.
A comparison of the measured and predicted stagnation
chamber pressures shows that the measured pressure initially
decreases more rapidly than indicated by the model. Thus,
the initially rapid measured pressure drop over the first 0.5 s
seen in Fig. 4b is underpredicted by the model, which does
not appear to predict the initial transient blowdown behavior.
However, for times greater than about 35 s the model predicts
only a slightly lower stagnation pressure.
Also shown in Fig. 6 is the predicted storage tube pressure.
It can be seen that the pressure in the tube is generally about
5% higher than the predicted stagnation chamber pressure at
the same time. This difference is due to losses in the pipes
connecting the storage tubes to the stagnation chamber, which
were minimized so that the highest stagnation chamber pressure
could be realized.
A comparison of jet exit pressures assuming isentropic flow
between the stagnation chamber and the jet exit shows that
the ideal-gas assumption predicts up to 8% higher pressures
than an AbelNoble gas early in the tank blowdown where the

Fig. 8 shows typical single-frame video images of the visible flame at various times into the blowdown test. The fieldof-view in the images, which have been cropped in Fig. 8, is
about 11.3 m in the vertical direction, and about 5.0 m in the
horizontal direction. In each image, the flame is vertically oriented with jet flow from bottom to top, and the jet exit is located near the bottom center of the image. Fig. 8a corresponds
to a time very close to the initial flame ignition. Also seen in
the frame are two glowing wires extending horizontally across
the flame. These are the electrically heated nichrome wires
that are used as the ignition sources. They are located at distances of x = 1.9 and 2.6 m downstream of the jet exit and pass
through the jet centerline. Typically ignition occurs at the second downstream wire where the flow velocity is sufficiently low
and the hydrogen/air mixture is within the flammability limits.
Figs. 8b and c, which are images taken at slightly later times,
show that from the initial ignition point the flame propagates
both upstream toward the jet exit where it stabilizes as a lifted
flame (Fig. 8b) and downstream into the flammable mixture
created by the hydrogen and ambient air entrained into the jet.
The final image, Fig. 8d, corresponds to a time of 5 s where
the flame has reached steady state. From this time until the end
of the test, the flame length slowly decreases over a period of
about 500 s as the pressure in the hydrogen storage tube decreases. With the exception of changes in the total flame length,
the flame appearance remains constant during the test.
Flame lengths based on the visible flame video images were
used to determine the time-average flame length. The average
flame length, indicated by the data points in Fig. 9, was determined from the flame length averaged over five successive
frames around the indicated time for each point. The flame
length decreases with time due to the decrease in mass flow
rate as tank pressure is reduced. Results are shown for all three

2090

R.W. Schefer et al. / International Journal of Hydrogen Energy 32 (2007) 2081 2093

Fig. 7. Comparison of calculated jet exit conditions using ideal-gas assumption and AbelNoble equation of state.
Table 3
Jet exit properties at a blowdown time of 2 s
Exit property

Ideal gas

AbelNoble gas

% Difference

Pe (bar)
Te (K)
qe (kg/m3 )
ue (m/s)
m (kg/s)

1.65
228
17.6
1151
0.41

1.53
223
14.8
1284
0.385

7.3
1.7
16.0
1.2
6.1

blowdown tests to quantify the reproducibility of the flame


length data. Based on the three tests, the estimated uncertainty
in the time-averaged flame length is +/ 5%.

Shown in Fig. 10 is the non-dimensional flame length, L , as


a function of Froude number, Frf . The non-dimensional flame
length and the Froude number are given by Eqs. (1) and (2),
respectively. Included in Fig. 10 are flame length data from
Kalghatgi [5] for a range of fuels (H2 , C3 H8 and CH4 ) and
inlet flow conditions. Also shown is the flame length data of
Schefer et al. [3] for momentum-dominated jets at a lower
initial storage tube pressure of 2500 psi, which was found to
be in good agreement with the momentum-dominated value of
L = 23 for Fr > 5. It can be seen that turbulent flame lengths
are well-correlated over a large range of flow conditions using
these non-dimensional parameters.

R.W. Schefer et al. / International Journal of Hydrogen Energy 32 (2007) 2081 2093

2091

12

10

Lvis (m)

100

200

300

400

500

600

Time (sec)
Fig. 9. Flame length history using visible flame emission.

102

L*

L*=13.5Fr2/5/(1+0.07Fr2)1/5

L*=23

10

H2 choked (d=7.94 mm)


H2 unchoked (d=7.94 mm)
CH4 (d=1.91 mm)
H2 (d=1.91 mm)
CH4 (Kalghatgi)
C3H8 (Kalghatgi)
H2 (Kalghatgi)
Present data (d=5.08 mm)

1.0
0.1

1.0

10.0

100.0

Fr
Fig. 8. Photographs of visible flame luminosity at different times during
blowdown. Ignition wires are visible in first three times. Jet diameter is
5.08 mm.

Values of Fr and L for the present data were calculated


using the time-averaged flame lengths in Fig. 9. Since the
present flow is choked at the jet exit, the concept of a notional nozzle expansion and the effective source diameter
(as described in Section 2.3) was used in Eqs. (1) and (2)
to reduce the hydrogen flame data and calculate the Froude
number. The present data are shown as solid red symbols in

Fig. 10. Variation of visible flame length with Froude number. Data is for
vertical jet orientation. Solid lines indicate correlations for buoyancy- and
momentum-dominated regimes as described by Eqs. (2) and (3).

Fig. 10 over the range of 10 < Frf < 20. The data collapse well
onto the momentum-dominated correlation given by Eq. (3b),
again verifying that these correlations, developed from lowerpressure jet studies, are valid for higher-pressure hydrogen-jet
releases.

2092

R.W. Schefer et al. / International Journal of Hydrogen Energy 32 (2007) 2081 2093

0.3

Fuel S (kW)
1.2

C2H4 11.2
C2H4 20.2
CH4 12.5
CH4 6.4
C2H2 18.1
C2H2 56.5
Fit to data of Ref. [7]
H2 data:

C*

0.80

Radiant Fraction

1.0

d=7.94mm / 2500psi
t=5 sec
t=10 sec
t=20 sec
d=5.08mm / 6000psi
t=20 sec

0.60
0.40

CO/H2 (Turns&Myhr, 1991)


CH4 "
C3H8 "
C2H4 "
CH4(d=1.91 mm)
H2 (d=1.91 mm)
H2 (d=7.94 mm)
"
H2
Present data

0.25
0.2
0.15
0.1
0.05
0

"

0.20

10

100

1000

Flame residence Time (ms)


Fig. 12. Radiant fraction as a function of flame residence time.

0.0
0.0

0.50

1.0

1.5

2.0

2.5

3.0

x/Lvis
Fig. 11. Profiles of normalized radiative heat flux in a turbulent, hydrogen-jet
flame. Jet diameter is 5.08 mm. Jet orientation is vertical.

4.3. Radiative flux


The radiative heat flux measurements were integrated to
obtain the total heat radiated, Srad , which allowed the nondimensional radiant power, C , to be calculated from Eq. (6).
The distribution of C for the present 5.08-mm-diameter H2 jet flame is shown in Fig. 11 for a tank blowdown time of 20 s.
Data are shown for tests 1 and 2 in which the radiometers are
located along the length of the flame at a constant radial distance R (see configuration on left of Fig. 3). The blowdown time
of 20 s corresponds to a visible flame length that is twice the
radial distance R, in accord with the method of Sivathanu and
Gore [7]. It can be seen that the present data collapse onto a single curve and show excellent agreement with the previous data
for lower-pressure hydrocarbon and hydrogen-jet flames. Thus,
the functional dependence for radiative heat flux expressed by
Eq. (6) applies to the present 413 bar (6000 psi) hydrogen-jet
flames.
The radiative heat flux data was also used to calculate the
radiant fraction, Xrad , using Eq. (4). In this case the data from
tests 1 and 2, as well as the data for the maximum radiative
heat flux obtained from the radiometers placed along a 45 line
to the flame centerline were used (configuration on right of
Fig. 3). The variation of radiant fraction with flame residence
time, given by Eq. (7), is shown in Fig. 12. Also shown is the
data of Turns and Myhr [8] for turbulent-jet flames using four
fuels with a wide variety of sooting tendencies. These fuels
included methane, ethylene, propane and a 57% CO/43% H2
mixture. Also shown are previous results obtained in 1.91 mm,
subsonic hydrogen and methane jet flames at various flow rates
and in the 7.94-mm diameter hydrogen-jet flame during the
blowdown of a 172 bar (2500 psi) source [3]. Note that the
data for non-sooting flames (CO/H2 and hydrocarbon flames
at shorter residence times) show a well-behaved, nearly linear
dependence on log of the residence time. In contrast, at longer
residence times, the radiant fraction for fuels that have a high

sooting tendency (propane and ethylene) becomes strongly dependent on residence time and behaves in a non-linear fashion.
The present data show excellent agreement with the previous
hydrogen-jet data and again indicate that, for a fixed flame residence time, the radiant fraction of hydrogen jets is about a
factor of two lower than non-sooting hydrocarbon flames.
5. Summary and conclusions
Measurements were performed to characterize the dimensional and radiative properties of large-scale, vertical hydrogenjet flames. High-pressure jets up to 413 bar (6000 psi) were
studied to verify the application of correlations and scaling laws
based on lower-pressure subsonic and choked-flow jet flames.
At these pressures, the flows exiting the jet nozzle are categorized as underexpanded and the flow is choked at the jet
exit. Additionally, the gas behavior departs from that of an
ideal-gas and alternate formulations for non-ideal gas must be
introduced. The flame length results show that lower-pressure
engineering correlations based on the Froude number and a nondimensional flame length also apply to releases from storage
vessels at pressures up to 413 bar (6000 psi). Similarly, radiative
heat flux characteristics of these high-pressure jet flames obey
scaling laws developed for low-pressure, smaller-scale flames
and a wide variety of fuels. The results verify that such correlations can be used to a priori predict dimensional characteristics and radiative heat flux from a wide variety of hydrogen-jet
flames.
Acknowledgments
This research was supported by the US Department of
Energy, Office of Energy Efficiency and Renewable Energy,
Hydrogen, Fuel Cells and Infrastructure Technologies Program. The large-scale hydrogen-jet experiments were conducted at the SRI International Corral Hollow Experimental
Site. Laboratory-scale flame experiments were conducted at
the Sandia Combustion Research Facility in laboratories supported by the US Department of Energy, Office of Basic
Energy Sciences, Chemical Sciences.

R.W. Schefer et al. / International Journal of Hydrogen Energy 32 (2007) 2081 2093

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.
[2] Turns SR. An introduction to combustion, second ed., New York:
McGraw-Hill; 2000.
[3] Schefer RW, Houf WG, Bourne B, Colton J. Spatial and radiative
properties of an open-flame hydrogen plume. Int J Hydrogen Energy
2006; 31: 13311340.
[4] Becker HA, Liang D. Combust Flame 1978;32:11537.
[5] Kalghatgi GT. Combust Sci Technol 1984;41:1729.
[6] Delichatsios MA. Combust Flame 1993;92:34964.
[7] Sivathanu YR, Gore JP. Combust Flame 1993;94:26570.
[8] Turns SR, Myhr FH. Combust Flame 1991;87:31935.
[9] Birch AD, Brown DR, Dodson MG, Swaffield F. Combust Sci Technol
1984;36:24961.
[10] Birch AD, Huges DJ, Swaffield F. Combust Sci Technol 1987;52:
16171.

2093

[11] John JEA. Gas dynamics. Boston: Allyn and Bacon, Inc.; 1969.
[12] Winters WS. TOPAZa computer code for modeling heat transfer and
fluid flow in arbitrary networks of pipes, flow branches, and vessels.
SAND83-8253, Sandia National Laboratories, Livermore, CA; January
1984.
[13] Winters WS. TOPAZthe transient one-dimensional pipe flow analyzer:
users manual. SAND85-8215, Sandia National Laboratories, Livermore,
CA; July 1985.
[14] Winters WS. TOPAZthe transient one-dimensional pipe flow analyzer:
an update on code improvements and increased capabilities. SAND878225, Sandia National Laboratories, Livermore, CA; September 1987.
[15] Houf W, Schefer RW. Predicting radiative heat fluxes and flammability
envelopes from unintended releases of hydrogen. In: 16th Annual
hydrogen conference and hydrogen expo USA, March 29April 1,
Washington, DC, 2005.
[16] Chenoweth DR. Gas-transfer analysis section H-real gas results via the
van der Waals equation of state and virial expansion extensions of its
limiting AbelNoble Form, Sandia Report SAND83-8229; June, 1983.

Você também pode gostar