You are on page 1of 6

# 2007 IChemE

IChemE SYMPOSIUM SERIES NO. 153

DEFLAGRATION TO DETONATION TRANSITION (DDT) IN JET IGNITED HYDROGEN-AIR


MIXTURES: LARGE SCALE EXPERIMENTS AND FLACS CFD PREDICTIONS
Prankul Middha1, Olav R. Hansen1 and Helmut Schneider2
1
GexCon AS, P.O. Box 6015, Postterminalen, NO-5892 Bergen, Norway; e-mail: prankul@gexcon.com
2
Fraunhofer Institut Chemische Technology, Joseph von Fraunhoferstr. 7, 76327 Pfinztal, Germany
As a result of a long history of model development and experimental validation, FLACS is
established as a CFD-tool for simulating hydrocarbon gas deflagrations with reasonable precision.
FLACS is widely used in petrochemical industry and elsewhere for explosion predictions for
input to risk assessments and design load specifications. In recent years the focus on predicting
hydrogen explosions has increased. A dedicated project was carried out between 2001 and 2004
to improve the modelling and validation of hydrogen explosions wherein many small and largescale experiments were simulated [1]. With the latest release of FLACS, the validation status for
hydrogen explosions is therefore considered good. For hydrogen explosions, deflagration to detonation transition (DDT) can be a significant threat. Recently, FLACS has been extended to indicate
the possibility of DDT in realistic situations. As a part of the study, four practical scenarios were
simulated and the simulation results were found to compare well with experimental data [2].
The model has now been developed further and used to simulate the experimental investigations
performed by Fraunhofer Institute of Chemical Technology. These concerned the transition of a
deflagration into a detonation in jet ignited hydrogen air mixtures within a partial confinement
[3]. The background for this project was the investigation of the potential hazards for a nuclear
power plant, whose process heat is used for the operation of an adjacent chemical plant (e.g. for
the gasification of coal), which should be located close to the nuclear plant to minimize heat
losses. The test set up consisted of a rectangular container (3 m  1.5 m  1.5 m) with an
opening on its front side. The container was followed by 2 parallel walls at a distance of 3 m
with a length of 12 m and a height of 3 m. The whole volume was filled with a hydrogen air
mixture, enclosed within a very thin PE-foil. The mixture was ignited at the rear side of the
container. The experiments observed very high pressures and transition to detonation due to the
high turbulence generated by a jet flame shooting into a large, reactive gas cloud followed by reflections of the high speed combustion front from the ground and the walls. The experiments observed
DDT for 21% hydrogen concentration, but not for mixtures less sensitive than that [4].
The modeling results are able to capture the experimental observations, including pressure traces
and locations of DDT, reasonably well. The possibility of DDT is indicated in terms of a spatial
pressure gradient across the flame front. The effect of geometrical dimensions on the observation
of DDT is also discussed by comparison with the detonation cell size. The flame speeds of the
detonation front are somewhat lower than those observed in the experiments but the development
of a shock ignition model is ongoing which is expected to resolve this difference.

KEYWORDS: DDT, hydrogen, CFD, FLACS, large scale experiments

industry, and the ongoing development of hydrogen-fuelled


vehicles, which are deemed to be the symbol of a future
hydrogen economy. A dedicated project was carried out
between 2001 and 2004 to improve the modelling and
validation of hydrogen explosions in FLACS wherein
many small and large-scale experiments were carried out,
combined with simulations and model improvements
(Hansen, et al., 2005). Therefore, the validation status of
FLACS for hydrogen explosions is considered good.
For hydrogen explosions, deflagration to detonation
transition (DDT) can be a significant threat. Transition to
detonation can occur in a variety of situations, many of
which are commonly employed in industrial settings.
These include flame acceleration as a result of repeated
obstacles (e.g. Peraldi, et al., 1986) and jet ignition (e.g.

INTRODUCTION
As a result of a more than 25-year history of model development and validation on the basis of experimental results at
GexCon, FLACS is established as a CFD-tool for simulating
hydrocarbon gas deflagrations with reasonable precision. An
extensive knowledge database has been compiled using both
experimental and theoretical studies under the aegis of a
series of Gas Safety programs (GSPs) that started in 1980.
This information has been implemented in the CFD tool
FLACS, which was first released in 1986. Today, FLACS
is used widely in petrochemical industry and elsewhere
for explosion predictions for input to risk assessments and
design load specifications. In recent years, there has been
a lot of focus on predicting gas explosions involving hydrogen. This is driven by an increasing interest from the nuclear

# 2007 IChemE

IChemE SYMPOSIUM SERIES NO. 153

Knystaus, et al., 1979). There has been a strong debate on


the mechanisms underlying the transition to detonation
and it is still an active research area. Detailed description
of all processes following ignition that may lead to DDT
is extremely challenging. This is due to a complicated interaction of compressible flow, chemical reaction, and turbulence that needs to be described at very high spatial and
temporal resolution. Much theoretical effort has been
focused on development of criteria for DDT (Breitung,
et al., 2000) but these criteria and scaling arguments are difficult to apply in a process facility.
Although the validation of the current version of
FLACS in simulating flame acceleration and high-speed
deflagrations is good, a detonation model is lacking.
We are currently involved in an activity sponsored by
Norwegian Research Council that aims at predicting the
extent to which DDT may be expected using FLACS. As
a part of this work, FLACS has been extended to indicate
the possibility of DDT in realistic situations. Previously,
many practical scenarios have been simulated and the
simulation results have been found to compare well with
experimental data (Middha, et al., 2006). The model has
now been developed further and used to simulate the large
scale experiments carried out by Fraunhofer Institut
Chemische Technology (Fh-ICT) which concerned the transition of a deflagration into a detonation in jet ignited
hydrogen air mixtures within a partial confinement (Pfortner
and Schneider, 1984; Schneider, 2005). Comparisons
have been performed in terms of pressure traces, location
and time of DDT, and the possibility of propagation of a
detonation front.

and corrected for curvature at scales equal to and smaller


than the reaction zone. All flame wrinkling at scales less
than the grid size is represented by sub-grid models,
which is important for flame interaction with objects
smaller than the grid size. FLACS uses a standard k-1
model for turbulence. However, some modifications are
implemented, the most important being a model for generation of turbulence behind sub-grid objects and a model
for flame folding around them (Arntzen, 1998). With the
close coupling between sub-grid modelling and turbulence
model, it is not believed that using a more advanced turbulence model with more equations and constants will give
much added value for the typical simulations carried out
with FLACS.
The representation of geometry using a distributed
porosity concept is one of the key advantages of FLACS
compared to several other CFD tools. The geometry is represented with area and volume porosities, as well as wake
generating sub-grid object areas in all flow directions.
FLACS can therefore be used to simulate all kinds of complicated geometries using a Cartesian grid. Large objects
and walls will be represented on-grid; smaller objects will
be represented sub-grid. Sub-grid objects will contribute
to flow resistance, turbulence generation, and flame
folding (for explosions). More details on FLACS can be
found elsewhere (Hansen, et al., 2005).

DESCRIPTION OF EXPERIMENTS
This section provides a short description of the experiments
that were conducted at Fh-ICT in 1984 (Pfortner and
Schneider, 1984). The background for this project was the
investigation of the potential hazards for a nuclear
power plant, whose process heat is used for the operation
of an adjacent chemical plant (e.g. for the gasification
of coal), which should be located close to the nuclear
plant to minimize heat losses. The test set up consisted of
a driver section that was a rectangular container
(3 m  1.5 m  1.5 m). In the front side of the driver
section there is a square spaced opening with blocking
ratio 0.1 (tests IA1, IA2, IA3) and 0.3 (tests IA4 and IA5).
The container was followed by a lane which consisted
of 2 parallel walls at a distance of 3 m with a length of
12 m and a height of 3 m (see Figure 2 for details). The
whole volume was filled with H2-air mixture, enclosed
within a very thin PE-foil. The mixture was ignited at the

BRIEF DESCRIPTION OF FLACS


FLACS is a computational fluid dynamics (CFD) code that
solves the compressible Navier-Stokes equations on a 3-D
Cartesian grid. The basic equations used in the FLACS
model as well as the explosion experiments to develop
and validate FLACS initially have been published
(Hjertager, 1985; Hjertager, et al., 1988). A model for development of the flame that describes how the local reactivity
changes with parameters like concentration, temperature,
pressure, turbulence, etc. is implemented. A good description of geometry and the coupling of geometry to the
flow, turbulence, and flame is one of the key elements in
the modelling. The real flame area is described properly

Figure 1. For explosion and dispersion studies representation of the detailed geometry is important for the quality of the predictions.
In FLACS this is handled with a porosity concept

# 2007 IChemE

IChemE SYMPOSIUM SERIES NO. 153

Figure 2. Test facility, with details of sensors and cameras. The tubes were not installed in test IA1

rear side of the container with 5 pyrotechnic igniters,


distributed over the area. In tests IA2 IA5, 2 vertical
tubes (diameter 14 cm) were installed, respectively, with a
distance of 5 cm from the wall each in the middle and at
the end of the lane. It was ensured that the mixture is homogeneous by mixing and sampling. We simulated all experiments, except IA3, and the relevant scenario parameters
are presented in Table 1. More information, including all
results is presented in Pfortner and Schneider (1984).
The experiments observed very high pressures and
transition to detonation due to the high turbulence generated
by a jet flame shooting into a large, reactive gas cloud followed by reflections of the high speed combustion front
from the ground and the walls. The experiments observed
DDT for 21% hydrogen concentration, but not for less
sensitive mixtures, with the exception of test IA1 where
no tubes were installed. Also, DDT was observed to occur
near the tube for test IA2 but near the ground for test IA4.
No detonation was seen for test IA5.

simulations. The simulations were carried out on a


LINUX PC with 12 processors and 34 GB RAM, and
took 2 days to complete. The possibility of DDT is indicated
in terms of a spatial pressure gradient across the flame front
(DPDX) as it is hypothesized that this parameter is able to
visualize when the flame front captures the pressure front,
which is the case in situations when fast deflagrations
transition to detonation (Middha, et al., 2006). The effect
of geometrical dimensions on the observation of DDT is
also discussed by comparison with the detonation cell size.
Figure 3 presents the comparison of simulated
pressure traces for selected sensors at different distances
inside the geometry with experimental observations for
test IA2. It is seen that the simulations agree reasonably
well with measurements, and similar agreement was
observed for other sensors. The arrival times of the pressure
peaks were consistent with those seen in the experiments,
while some discrepancies were seen in peak pressures.
Similar comparisons were seen for other tests. Detailed
results were not presented for all tests due to lack of
space. The maximum simulated pressure in test IA2 was
10.2 barg at sensor 12, compared to 9.2 barg in the experiments also at sensor 12. The flame arrival times were
calculated to be 50 and 90 ms at photo transistors F2 and
F3, compared to the observed values of 50 and 80 ms. For
test IA4, the maximum simulated was 9 barg at sensor 8,
compared to around 12 barg in the experiments also at
sensor 8. In this case, the flame arrival times at F2 and F3
were calculated to be 35 and 52 ms, compared to observed
values of 38 and 49 ms. The maximum pressure in the
driver section was also found to compare very well with
experimental results for all tests. 2D snapshots of the
pressure field at the ground, along with the flame and
DPDX for test IA2 at different times with a detailed
description of various different stages during the simulation
are shown in Fig. 4. The pressure is seen to rise very quickly
as the hot flame jet shoots out into the lane, creating a lot of
turbulence and mixing of hot products with the unburnt
mixture. The hot products act as ignition sources, and a

RESULTS
This section presents some of the key results of the
simulations, and comparisons with experimental data. The
simulation domain was resolved by around 3.25 million
grid cells with a grid resolution of 5 cm. The grid
was according to the FLACS guidelines for explosion
Table 1. Relevant scenario parameters for the four tests
considered
Parameters
Ambient Temp. (K)
Ambient pressure (bar)
H2 concentration in
driver unit (%)
H2 concentration in lane (%)

IA1

IA2

IA4

IA5

279.2
0.988
21.9

281.5
0.991
20.8

293
0.993
22.3

293.4
0.996
19.9

21.0

21.1

22.5

20.0

# 2007 IChemE

IChemE SYMPOSIUM SERIES NO. 153

Figure 3. Comparison of simulated pressure traces (left) for selected sensors with experimental data (right) for test IA2. Similar
agreement was seen for other monitor points

large amount of the unburnt mixture is simultaneously


ignited. In the shear layer near the wall, still higher turbulence levels and reflections are seen, with very high pressures and maximum likelihood of DDT. The shock wave is
seen to sustain a pressure of 20 barg before it decays as
seen in the last picture in Fig. 4. A more pertinent parameter
for this work is the parameter DPDX, shown in the bottom
part of each snapshot (a value  10 indicates a strong
likelihood of DDT if dimension of high DPDX region is
significant compared to detonation cell size).
For test IA2, the maximum value of DPDX
(maximum likelihood of DDT) was seen at the first set
of pipes. DDT could also occur before, but with lesser

probability. This agrees with the experimental observation


of DDT at the left pipe. For test IA4, DDT was predicted
to occur at or near the ground next to the sidewalls, and
no special likelihood was seen at the pipes. This was also
consistent with the experimental result. The simulations
also indicated a similar, but somewhat lower possibility of
DDT for test IA1 compared to test IA4, as the H2 concentrations in this test were actually higher than those in test
IA2. The values and dimensions of the simulated highly
tumultuous region for test IA5 were much smaller, and
thus indicated a small chance of DDT.
A comparison of the geometrical dimensions with the
detonation cell size was also carried out, but that was found

# 2007 IChemE

IChemE SYMPOSIUM SERIES NO. 153

Figure 4. 2D snapshots of simulation results (test IA2): P (top), flame (middle), DDT indication parameter DPDX (bottom)

to be much larger than the minimum required for possible


propagation of detonation waves for all tests. The
maximum flame speed was seen to be 1208 m/s in test
IA2 and 1374 m/s in test IA 4 (compared to 1651 m/s
and 1740 m/s in the experiments). These are somewhat
lower than those observed in the experiments but the development of a shock ignition model is ongoing which is
expected to resolve this difference. However, our calculations also indicated the possibility of DDT in test IA1,
which was not seen in the experiments. But as mentioned
above, DDT is a very complex phenomenon, and is extremely difficult to predict accurately. Also, the absence of a
detonation model leads to the decay of the shock front,
even after DDT is expected to happen. In general, the
modeling results are able to capture the experimental observations, including pressure traces and locations of DDT,

reasonably well. We hope that the current model, when


coupled with the additional features currently under development, can be used by the process industry to get a fair
idea of the danger of DDT.

CONCLUSIONS
Large-scale experiments carried out at Fh-ICT have been
simulated using the CFD tool FLACS. In general, the
modeling results are able to capture the experimental observations, including maximum pressures, arrival times, and
locations of DDT, reasonably well. However, some
discrepancies are seen, which may be attributed to experimental uncertainties and the very difficult nature of the
simulations. The flame speeds of the detonation front are
somewhat lower than those observed in the experiments

# 2007 IChemE

IChemE SYMPOSIUM SERIES NO. 153

5. Hjertager, B., Fuhre, K., Bjorkhaug, M., 1988, Gas


explosion experiments in 1:33 and 1:5 scale offshore
separator and compressor modules using stoichiometric
homogeneous fuel/air clouds, J. Loss Prevention Proc.
Ind., 1: 197 205.
6. Knystautas, R., Lee, J. H., and Wagner, H. G., 1979, Direct
initiation of spherical detonation by a hot turbulent gas jet.
Proc. Comb. Inst., 17: 1235 1245.
7. Middha, P, Hansen, O. R., and Storvik, I. E., 2006. Prediction of deflagration to detonation transition in hydrogen
explosions. Proceedings of the AIChE Spring National
Meeting and 40th Annual Loss Prevention Symposium,
Orlando, FL, April 23 27, 2006.
8. Peraldi, O., Knystautas, R., and Lee, J. H., 1986, Criteria
for transition to detonation in tubes. Proc. Comb. Inst.,
21: 1629 1637.
9. Pfortner, H., Schneider, H., Tests with Jet Ignition of
Partially Confined Hydrogen Air Mixtures in View of
the Scaling of the Transition from Deflagration to
Detonation, Final Report for Interatom GmbH, Bergisch
Gladbach, Germany, Oct. 1984, Fraunhofer ICT Internal
Report.
10. Schneider, H., 2005, Deflagration and deflagration to detonation transition within a partial confinement similar to a
lane Proceedings of International Conference on Hydrogen
Safety, Pisa, Italy, September 2005.

but the development of a shock ignition model is ongoing


which is expected to resolve this difference. Also, the
absence of a detonation model leads to the decay of the
shock front, even after DDT is expected to happen. The
support of Norwegian Research Council for this work is
acknowledged.

REFERENCES
1. Arntzen, B.A., 1998. Modeling of turbulence and combustion for simulation of gas explosions in complex geometries, PhD Thesis, NTNU, Trondheim, Norway.
2. Breitung, W., et al., 2000. Flame Acceleration and
Deflagration to Detonation Transition in Nuclear Safety.
State-of-the-Art Report, OECD Nuclear Energy Agency,
Ref. NEA/CSNI/R/2000/7.
3. Hansen, O.R., Renoult, J., Sherman, M.P., and Tieszen,
S.R., 2005, Validation of FLACS-Hydrogen CFD Consequence Prediction Model Against Large Scale H2
Explosion Experiments in the FLAME Facility, Proceedings of International Conference on Hydrogen Safety,
Pisa, Italy, September 2005.
4. Hjertager, B.H., 1985, Computer simulation of turbulent
reactive gas dynamics. J. Model. Identification Control,
5: 211 236.

You might also like