Sie sind auf Seite 1von 58

ERI/NRC 09-203

COMPARISON OF BLAST PRESSURES AND


EFFECTS METHODOLOGIES
WITH APPLICATION TO SOUTH TEXAS UNITS 3 & 4

Work Performed under the Auspices of:


U.S. Nuclear Regulatory Commission
Office of New Reactors (NRO)
Washington, D.C. 20555
Under Contract Number NRC 42-07-483

P. O. Box 2034
Rockville, Maryland 20847

February 2009
This page intentionally left blank.
ERI/NRC 09-203

COMPARISON OF BLAST PRESSURES AND


EFFECTS METHODOLOGIES
WITH APPLICATION TO SOUTH TEXAS UNITS 3 & 4

February 2009

Final Report

Edward A. Rodriguez, P.E. 1 and Wayne Schofield 2

Energy Research Inc.


6167 & 6189 Executive Blvd.
Rockville, Maryland 20852

Work performed under the auspices of


United States Nuclear Regulatory Commission
Washington, D.C.
Under Contract Number NRC 42-07-483 (Task Order 13)

1
Global Nuclear Network Analysis, LLC
2
Dade Moeller & Associates, Inc.
This page intentionally left blank
EXECUTIVE SUMMARY

The Nuclear Regulatory Commission queried the South Texas Project (STP) Units 3 & 4
engineering on the Final Safety Analysis Report (FSA) regarding Regulatory Guide 1.91
computations to determine safe distance from an explosion source. STP provided answers to the
NRC’s Request for Additional Information (RAI), which was further reviewed by subject matter
experts for Energy Research, Inc. The results of the technical independent review show that STP
engineering appropriately and correctly applied substance-specific thermodynamic data, under
specified conditions, to obtain the heats-of-combustion in calculating the TNT-equivalent
explosion energy for different fuels/oxidizer mixtures. As such, it should be understood that
specific thermodynamic parameters and atmospheric conditions applicable to the event of interest
must be well founded to yield conservative safe distances.

This report is intended to provide background information and comparison of engineering


methods, which are currently applied elsewhere in industry, to determine effects of combustions,
explosions, and detonations, and normalize these methods to utilize common TNT-equivalent
blast curves. Secondly, in reviewing Regulatory Guide 1.91, it has become evident that misuse
or misapplication of technical information within Regulatory Guide 1.91 may result in
inappropriate safe distance calculations for nearby explosions.

v ERI/NRC 09-203
TABLE OF CONTENTS

EXECUTIVE SUMMARY ..................................................................................................................................v

LIST OF TABLES........................................................................................................................................... viii

LIST OF FIGURES ........................................................................................................................................ viii

1. INTRODUCTION.....................................................................................................................................1

2. METHODOLOGIES USED BY OTHER UTILITIES................................................................................3


2.1 Bellefonte Nuclear Plant, Units 3 & 4 ........................................................................................... 3
2.2 Levy Nuclear Plant, Units 1 & 2 ................................................................................................... 3
2.3 Shearon Harris Nuclear Plant, Units 2 & 3................................................................................... 3
2.4 William States Lee III Nuclear Plant, Units 1 & 2 ......................................................................... 4
2.5 V. C. Summer Nuclear Plant, Units 2 & 3 .................................................................................... 4
2.6 South Texas Project Nuclear Plant, Units 3 & 4 .......................................................................... 4

3. BACKGROUND ......................................................................................................................................5

4. DOD HIGH EXPLOSIVES CURVES ......................................................................................................7

5. REGULATORY GUIDE 1.91 METHODOLOGY ...................................................................................11

6. THERMODYNAMICS FUNDAMENTALS .............................................................................................15


6.1 Thermophysics ........................................................................................................................... 15
6.2 Thermochemistry........................................................................................................................ 15
6.2.1 Heat-of-Reaction ........................................................................................................ 16
6.2.2 Heat-of-Combustion................................................................................................... 16
6.2.3 Heat-of-Detonation..................................................................................................... 17
6.3 Adiabatic Isentropic Gaseous Expansion................................................................................... 17

7. NUREG-1805 AND FM DATA SHEETS COMPARISON OF METHODOLOGIES ..............................21


7.1 NUREG-1805 (Confined Explosion-Leaking Flammable Gas) .................................................. 21
7.2 Factory Mutual Data Sheets....................................................................................................... 22
7.2.1 Gas Releases............................................................................................................. 22
7.2.2 Liquid Releases.......................................................................................................... 23
7.2.3 Calculation of Initial Flash Fraction ............................................................................ 23
7.2.4 Calculation of Liquid Pool Size .................................................................................. 24
7.2.5 Calculating of Evaporation from a Liquid Pool........................................................... 24
7.2.6 Calculation of Total amount of Vapor in a Cloud ....................................................... 25
7.2.7 Calculating TNT Equivalency..................................................................................... 26

8. BOUNDING METHODOLOGIES..........................................................................................................27
8.1 Solid Materials............................................................................................................................ 27
8.2 Liquids ........................................................................................................................................ 27
8.3 Vapor Clouds.............................................................................................................................. 27
8.3.1 Theory and Principle .................................................................................................. 28
8.3.2 Materials not Conducive to VCEs .............................................................................. 29
8.3.2.1 Liquefied Natural Gas (LNG) and Natural Gas (NG) (methane) ............... 29
8.3.2.2 Ammonia Gas ............................................................................................ 29
8.3.2.3 Gaseous Hydrogen.................................................................................... 29
8.3.2.4 Miscellaneous Flammable or Combustible Gases .................................... 30

vi ERI/NRC 09-203
8.3.2.5
Flammable Liquids or Gases processed above their Auto-ignition
Temperature .............................................................................................. 30
8.3.2.6 Flammable Liquids (High Viscosity) .......................................................... 30
8.3.2.7 Mixtures ..................................................................................................... 30
8.3.2.8 Hybrid Mixtures.......................................................................................... 30
8.3.3 VCE Methodologies ................................................................................................... 30
8.4 Additional Literature ................................................................................................................... 31

9. CONCLUSIONS....................................................................................................................................33

10. RECOMMENDATIONS.........................................................................................................................35

11. REFERENCES......................................................................................................................................37

APPENDIX: CONFIRMATORY CALCULATIONS.........................................................................................39

vii ERI/NRC 09-203


LIST OF TABLES

Table 1 TNT Equivalence ...................................................................................................... 6

Table 2 Combustion Parameters for Stoichiometric Gasoline and Oxygen ........................ 18

Table 3 Comparison of Calculated Safe Distances (ERI versus STP)................................ 34

LIST OF FIGURES

Figure 1 Blast parameters for HE charges [3] ........................................................................ 7

Figure 2 Peak “incident” and “reflected” pressures [3]............................................................ 8

Figure 3 Zoom view of incident and reflected pressures near 1-psi. ...................................... 9

Figure 4 Pressure-time history for free-air burst [3] ................................................................ 9

Figure 5 Exposure distance calculation [1] ........................................................................... 11

Figure 6 Radial distance to peak incident pressure of 1-psi [1] ............................................ 11

Figure 7 Radial distance from blast source for 1-psi incident overpressure......................... 13

viii ERI/NRC 09-203


1. INTRODUCTION

In accordance with US Nuclear Regulatory Commission (NRC) Regulatory Guide 1.91 [1],
postulated accident analyses on routes near nuclear power plants must be addressed for all
known combustible or detonable fuels.

However, Regulatory Guide 1.91 addresses only solid explosives and hydrocarbons liquefied
under pressure, and air blasts on highway, rail, and water routes. The regulatory guide does not
address liquids (pressurized or non-pressurized), cryogenically liquefied hydrocarbons (LNG,
CNG, propane), vapor clouds (confined or unconfined), fixed facilities, pipelines, etc.

Since Regulatory Guide 1.91 does not include methodologies that can be used to address these
additional scenarios it was important to review the different methodologies utilized by other
utilities as a basis for their FSAR Chapter 2, section 2.2.3, “Evaluation of Potential Accidents”. A
summary of some available evaluations follow in Section 2.

Finally, a note on engineering units utilized in this report. Wherever possible, Systems
International (SI) units are used, except in certain figures and equations, which were derived from
existing DOD documents. In some instances, both English and SI units will be included as
appropriate, however a thorough unified convention was used, where possible, herein.

1 ERI/NRC 09-203
This page intentionally left blank

2 ERI/NRC 09-203
2. METHODOLOGIES USED BY OTHER UTILITIES

Prior to evaluating the methodologies presented in Regulatory Guide 1.91 and other industry
guidance, it was important to determine what methodologies other utilities have used in their
submittal of Combined Operating License (COL) Application FSAR section 2.2.3, in order to
accommodate scenarios not encompassed by Regulatory Guide 1.91.

A search was performed in ADAMS on the NRC website to locate FSAR Chapter 2, section 2.2.3
for several utilities who have recently submitted their COL Applications for NRC review. These
included:

 Bellefonte Nuclear Plant, Units 3 & 4


 Levy Nuclear Plant, Units 1 & 2
 Shearon Harris Nuclear Plant, Units 2 & 3
 William States Lee III Nuclear Plant, Units 1 & 2
 V. C. Summer Nuclear Plant, Units 2 & 3
 South Texas Project Nuclear Plant, Units 3 & 4

These sections were reviewed and a summary discussion follows.

2.1 Bellefonte Nuclear Plant, Units 3 & 4

For solid fuels transported by truck, rail, or barge, the methodologies presented in Regulatory
Guide 1.91 were used to determine the safe standoff distance from a detonation/deflagration from
a mass equivalent of TNT.

For methodologies used as the basis for determining the effects and safe stand off distance from
Confined and unconfined VCEs at nearby facilities, only summary information was presented in
the FSAR and further review of the associated calculation package would be necessary to
determine what methodologies were used.

For flammable vapor clouds (delayed ignition), the ALOHA code [2] was used to evaluate the
dispersion and detonation/deflagration of the vapor clouds.

2.2 Levy Nuclear Plant, Units 1 & 2

For solid fuels transported by truck, the methodologies presented in Regulatory Guide 1.91 were
used to determine the safe standoff distance from a detonation from a mass equivalent of TNT.

It is stated in the FSAR section that unconfined VCEs from the rupture of natural gas transport
pipeline were not considered a credible event, as previously determined in NRC licensing actions,
and were not considered further in the document. Regulatory Guide 1.194 was used to calculate
concentrations via plume rise and dispersion of a vapor cloud.

For confined VCEs, it was determined that no credible mechanism existed to trap a natural gas
leak that would lead to a detonation/deflagration.

2.3 Shearon Harris Nuclear Plant, Units 2 & 3

For solid fuels transported by truck or rail, the methodologies presented in Regulatory Guide 1.91
were used to determine the safe standoff distance from a detonation/deflagration from a mass
equivalent of TNT.

3 ERI/NRC 09-203
For the rupture of a nearby LPG pipeline and analysis of the effects from a fire and explosion, the
scenario was previously analyzed in the HNP FSAR and the results extrapolated for the new
facilities.

2.4 William States Lee III Nuclear Plant, Units 1 & 2

For solid fuels transported by truck or rail, the methodologies presented in Regulatory Guide 1.91
were used to determine the safe standoff distance from a detonation/deflagration from a mass
equivalent of TNT.

Unconfined VCEs from the rupture of natural gas transport pipeline were not considered a
credible event, as previously determined in NRC licensing actions and not considered further in
the document.
For methodologies used as the basis for determining the effects and safe standoff distance from
Confined VCEs at nearby facilities, only summary information was presented in the FSAR and
further review of the associated calculation package would be necessary to determine what
methodologies were used.

For Flammable vapor clouds (delayed ignition), the ALOHA code [2] was used to evaluate the
dispersion and detonation/deflagration of the vapor clouds.

2.5 V. C. Summer Nuclear Plant, Units 2 & 3

For solid fuels transported by truck or rail, the methodologies presented in Regulatory Guide 1.91
were used to determine the safe standoff distance from a detonation/deflagration from a mass
equivalent of TNT.

NUREG 1805 was used as the basis for determining the volume of vapor at the upper
flammability limit capable of occupying the largest vessel considered available for combustion.
Regulatory Guide 1.91 was used to determine the safe standoff distance.

For Flammable vapor clouds (delayed ignition), the ALOHA code [2] was used to evaluate the
dispersion and detonation/deflagration of the vapor clouds.

2.6 South Texas Project Nuclear Plant, Units 3 & 4

For solid fuels transported by truck, rail, or barge, the methodologies presented in Regulatory
Guide 1.91 were used to determine the safe standoff distance from a detonation/deflagration from
a mass equivalent of TNT.

For atmospheric liquids and gases, the South Texas Project used guidance obtained from both
NUREG-1805 and the Factory Mutual Data sheets to account for the limited applicability of
Regulatory Guide 1.91.

For Flammable vapor clouds (delayed ignition), the ALOHA code [2] was used to evaluate the
dispersion and detonation/deflagration of the vapor clouds.

4 ERI/NRC 09-203
3. BACKGROUND

Basis for guidance provided in Regulatory Guide 1.91 [1] stems from the US Department of
Defense (DOD) laboratories, which conducted blast effects starting in the 1940’s, utilizing solid
state high-explosives (HE). The DOD developed these curves for all known solid explosives, yet
normalized the detonation parameters to TNT-equivalence. The equivalence method is quite
elementary, but effective, such that a multitude of different high explosives have been correlated
with great precision to TNT.

A US military tri-services manual titled, Structures to Resist the Effects of Accidental Explosions
[3], commonly referred to as “TM 5-1300,” has been used in blast engineering design by DOD
since the early 1960’s. The US Department of Energy (DOE) developed a similar, yet site-
specific manual titled, A Manual for the Prediction of Blast and Fragment Loading on Structures
[4], for design against accidental explosions primarily in HE facilities, with much the same
information and guidance as TM 5-1300.

TM5-1300 and Lawrence Livermore National Laboratory’s (LLNL) methodology for TNT-
equivalence provide similar results because they both use the heat-of-detonation  H 
o
Det in
scaling, but the former is based on scaling weight (or mass) while the latter is based on energy-
per-unit-density.

Further discussions on heats-of-detonation are expanded in Section 6.2. As such, the slight
difference in TNT-equivalence is due to the density ratio, and for all practical purposes, the two
methods are identical. Theoretical density of TNT is set at   1.654 g/cm3, which is typically
used with the LLNL [5] methodology.

The TM5-1300 Method [3]:

H Det
o
WE  WExp (1)
H TNT
o

 H o 
 Det

  o Exp  
  
 H  o
 TNT

  oTNT  
 

Or, the LLNL Method [5]:

 H o 
  0.2107  Det
 (2)
  o Exp  
 

Table 1 shows sample computations with both methods in determining the TNT-equivalence for
common high explosives utilized by DOD and DOE.

5 ERI/NRC 09-203
Table 1 TNT Equivalence
TNT Equivalence
High Explosive
TM5-1300 LLNL

TNT 1.00 1.00

HMX 1.26 1.27

PBX-9501 1.16 1.17

PBX-9502 0.84 0.85

6 ERI/NRC 09-203
4. DOD HIGH EXPLOSIVES CURVES

The two blast effects manuals from DOD [3] and DOE [4] previously mentioned, contain the
current technical philosophy for determining blast overpressures and analyzing blast effects on
structures. Figure 1 shows the TM 5-1300 blast curve(s) as a function of scaled distance, Z is
defined in Equation 1. As stated previously, these curves were developed by the US Department
of Defense (DOD) laboratories, which conducted blast effects starting in the 1940’s, utilizing solid-
state high explosives. The DOD developed these curves for all known solid explosives, yet
normalized the detonation parameters to TNT-equivalence.

R
Z
W 1/3 (3)
Where:
R  Distance from explosive to target, (ft)
W  Weight of Explosive, (lb)
The scaled distance equation, or the “one-third” scaling law, is commonly called the Hopkinson-
Crantz scaling law, as described by Baker [6].

100000
50000
Pr, psi
20000 Pso, psi
10000 Ir, psi-ms/lb^(1/3)
5000 Is, psi-ms/lb^(1/3)
2000 ta, ms/lb^(1/3)
1000 to, ms/lb^(1/3)
500 U, ft/ms
Lw, ft/lb^(1/3)
200
100
50
20
10
5
2
1
0.5
0.2
0.1
0.05
0.02
0.01
0.005
0.1 0.2 0.3 0.50.7 1 2 3 4 5 67 10 20 30 50 70 100
Scaled Distance Z = R/W^(1/3)
Figure 1 Blast parameters for HE charges [3]

Primarily, Figure 1 curves provide blast parameters such as peak incident pressure (or, side-on
pressure) Pso , reflected pressure Pr , incident impulse I so , reflected impulse I r , and time of arrival
ta as well as pulse period to . Additional parameters included in Figure 1 are the shock front
velocity, U , and the positive phase wavelength, Lw . As stated previously, these curves are
normalized to TNT-equivalent energy and therefore utilize the heat-of-detonation H Exp method
d

described in section 2.0 of this report. Figure 2 clearly shows the incident and reflected pressure

7 ERI/NRC 09-203
as a function of scaled distance, where at 1-psi incident positive overpressure, and the scaled
distance, Z , is approximately 36-ft/lb1/3.

100000
50000

20000
10000
5000
Pr
2000 Pso
1000
Pressure, (psi)

500

200
100
50

20
10
5

2
1
0.5

0.2
0.1 0.2 0.3 0.50.7 1 2 3 4 5 67 10 20 30 50 70100
Scaled Distance Z = R/W^(1/3)
Figure 2 Peak “incident” and “reflected” pressures [3]

Figure 3 present a close-up view of the incident and reflected pressure scale near the low-
pressure regime. Here, one clearly sees the scaled distance of ~36 ft/lb1/3 for attaining a 1-psi
incident positive overpressure, while the reflected pressure is a factor of 2 higher, i.e., 2-psi.
Figure 4 provides a typical pressure-time history of a blast wave, showing both the incident and
reflected pressure waves.

The blast pressure moving away from the HE source and towards a target surface is termed the
incident pressure, Pso , and is also the pressure magnitude immediately before impacting on a
hard (reflecting) surface. Reflected pressure, Pr , is the magnitude of blast pressure that is
subsequently amplified after striking a hard surface. It is the reflected peak pressure and its
associated impulse that causes damage to structures.

8 ERI/NRC 09-203
10

7
6
5
4 Pr
Pso
Pressure, (psi) 3

0.7
0.6
0.5
0.4

0.3

0.2
10 20 30 40 50 60 70 80 100
Scaled Distance Z = R/W^(1/3)
Figure 3 Zoom view of incident and reflected pressures near 1-psi.

Figure 4 Pressure-time history for free-air burst [3]

9 ERI/NRC 09-203
This page intentionally left blank

10 ERI/NRC 09-203
5. REGULATORY GUIDE 1.91 METHODOLOGY

NRC Regulatory Guide 1.91 utilizes the methodology discussed in Section 4.0 for determining the
safe distance from an explosion source to a facility or critical location. Figure 5 shows the
Regulatory Guide exposure distance along a roadway, which potentially is accessed by vehicles
carrying hazardous and/or detonable materials. The radial distance, R , measured from a facility,
is based on the scaled distance, Z , of 45 ft/lb1/3. This difference from the TM 5-1300
methodology will be discussed in some detail later, showing there is a rationale for the increase.

Figure 5 Exposure distance calculation [1]

Figure 6 shows the Regulatory Guide 1.91 design curve for radial distance to TNT weight, based
on peak incident overpressure of 1-psi. It is interesting to note that the examples noted on the
graphical plot imply these are TNT weights for solid materials, yet there is no indication of how to
arrive at these weights for other substances or mixtures (i.e., gaseous or liquid).

Figure 6 Radial distance to peak incident pressure of 1-psi [1]

11 ERI/NRC 09-203
Herein, provided is a comparison of the different methods to evaluate their conservative or non-
conservative nature. The methodologies presented in Regulatory Guide 1.91 and TM 5-1300 are
illustrated in Equations 4 and 5 and are compared to show the differences between them.

English Units

RRG  45W 1/3 Regulatory Guide 1.91 (4)


RTM 5  36.225W 1/3
TM 5-1300 (5)

Where:
RRG  Radial distance, (ft)
W  Charge weight, (lb)

SI Units

RRG  17.85W 1/3 Regulatory Guide 1.91 (4a)


RTM 5  14.37W 1/3
TM 5-1300 (5a)

Where:
RRG  Radial distance, (m)
W  Charge mass, (kg)
For a given radial safe distance, the differential on HE mass to reach 1-psi incident (i.e., side-on)
positive overpressure is:

1/3
45WRG  36.225WTM
1/3
5 (6)
WTM 5  1.92WRG (7)

For a given HE mass, the differential on radial distance to reach 1-psi incident positive
overpressure is:

RRG R
 TM 5
45 36.225 (8)
RRG  1.24 RTM 5 (9)

Graphically, these differences represent a conservative factor of safety for applications of design
conditions under Regulatory Guide rules. Figure 7 shows the Regulatory Guide 1.91 and TM 5-
1300 design curves for 1-psi incident positive overpressure. Not having any knowledge of the
NRC’s original effort on developing Regulatory Guide 1.91, it appears from the plot that the TM 5-
1300 scaled distance factor was utilized and increased by 25%, if one uses whole numbers for
the scaled distance;

45
f   1.25
36 (10)

12 ERI/NRC 09-203
Comparison of Reg Guide and TM 5-1300
4
10

W = 92%
Distance "R" from Explosion (ft)

R = 24%
1000

Reg Guide 1.91


TM 5-1300

100
4 5 6 7
100 1000 10 10 10 10

TNT Equivalent W eight, "W" (lb)

Figure 7 Radial distance from blast source for 1-psi incident overpressure

It should be emphasized, and as previously described in Section 1.0, that Regulatory Guide 1.91
addresses only solid explosives, hydrocarbons liquefied under pressure, and air blasts on
highway, rail, and water routes. The regulatory guide does not address cryogenically liquefied
hydrocarbons (LNG), fixed facilities and pipelines. Proper treatment for other hazardous and
detonable mediums requires fundamental knowledge of the system thermodynamics, and
specifically thermo chemistry of mixtures.

13 ERI/NRC 09-203
This page intentionally left blank

14 ERI/NRC 09-203
6. THERMODYNAMICS FUNDAMENTALS

As has already been discussed, the Regulatory Guide does not address other types of detonable
fuels except for solid HE and hydrocarbons liquefied under pressure. However, for hydrocarbons
liquefied under pressure, the Regulatory Guide utilizes a factor of 240% increase in mass that
account for the differences with TNT. Although this might be acceptable for some hydrocarbons,
it is by no means a conservative assumption, as will be seen later. Furthermore, it would not be
prudent, nor technically correct, to assume that a given mass (or weight) of a fuel is energetically
equivalent to the same mass (or weight) of TNT. In order to determine heat energy from different
substances, it becomes necessary to approach the problem from fundamentals of
thermodynamics. Herein we will briefly discuss applications of thermodynamics to chemically
reacting mixtures, such as gases or other liquid fuels.

6.1 Thermophysics

Changes in physical states, such as temperature and pressure, involve changes in internal
energy U  and enthalpy H  of the system [7]. The heat capacity of a substance defines the
quantity of heat energy required that a given amount of a substance must absorb to raise its
internal temperature 1-degree. Therefore, the amount of heat required to raise the temperature
of a substance, such as gas or liquid, in a closed (i.e., constant) volume is:

dU
cv 
dT (11)

This implies that all the heat goes into increasing the gas, or liquid’s, internal energy, U  ,
where:

H  U  PV (12)

On the other hand, if the gas or liquid is allowed to expand during heating process, then some of
the heat will increase the internal energy U  and some utilized as “work” involved in expansion:

dH
cp 
dT (13)

Thus, the above implies that:

dH  c p dT , or d U  PV   c p dT

6.2 Thermochemistry

Having described the internal energy and enthalpy from the standpoint of heat capacities of a
substance, we will briefly describe the changes in the chemical state of a mixture. That is,
chemical changes of a substance, or changes in the composition of molecules, also involves
changes in internal energy and enthalpy of system [7]. The foundations of a chemically reacting
mixture is a balanced reaction, where mass is always conserved while maintaining the number of
atoms equal between the reactants and products.

15 ERI/NRC 09-203
6.2.1 Heat-of-Reaction

The heat-of-reaction H r  is the change in enthalpy between the starting and ending chemical
o

states of a substance or mixture. Heats-of-formation  H  are merely the heats-of-reaction, or


o
f

changes in enthalpies, involved in forming a particular compound from its elements, where these
elements and final compound are at a standard-state, i.e., standard temperature and pressure
(STP). The standard-state is chosen arbitrarily as 1-atm pressure and 298.15K temperature,
however, it is also for consistency and because it is generally easier to determine chemical states
at room temperature and pressure. These states are known as heats-of-formation. For example,
if the starting state for hydrogen  H 2  and oxygen O2  is a gaseous mixture at constant
pressure, and we burn the mixture, the chemical balance becomes:

2 H 2  O2  2 H 2O (14)

Implying that upon burning 2 molecules of  H2  with one molecule of  O2  , we obtain water

vapor  H 2O  . If the heat-of-reaction occurs at constant pressure, the energy balance can be
written between enthalpies of formation of gaseous reaction products and reactants (i.e., fuels
and oxidizers):

H ro   H of  products    H 0f  reactants 
(15)

Thus, the substance’s heat-of-reaction from using heat-of-formation data for the reactants and
reaction products of the H2-O2 mixture is,
H ro   H of  2 H 2O    H 0f  2H 2 O2 
(16)

No matter what terminology is used, heat-of-combustion, heat-of-detonation, or heat-of-explosion,


these terms all imply the same phenomena. That is, it implies the enthalpy difference between
reaction products and reactants at STP.

6.2.2 Heat-of-Combustion

Heat-of-combustion  H  of a substance, or mixture, is the heat-of-reaction for a complete


o
c

combustion (or burning) with molecular oxygen, to its most oxidized state. For the majority of
substances in gaseous, liquid, or solid form, these tests have been conducted and reference
values of heats-of-combustion may be found in most standard reference books [8 - 9]. Where
heats-of-combustion for unusual substances cannot be found, then it becomes necessary to
calculate heats-of-formation for both reactants and reaction products.

Heats-of-formation for many substances are found in the CRC Handbook [10], Cooper [7],
Glassman [11], and the most comprehensive are contained in the tri-services sponsored
compendium titled, JANAF Thermochemical Tables [12].

H co   H of  products    H 0f  reactants 
(17)

16 ERI/NRC 09-203
6.2.3 Heat-of-Detonation

Heat-of-detonation is yet another special case of heats-of-reaction associated primarily with


propellants or solid explosives that release heat through the process of detonation. Again, the
same process is applied in determining the heat-of-detonation as was accomplished for the heat-
of-reaction or heat-of-combustion. However, in this case, heats-of-formation for the actual
explosive compound and reaction products must be obtained utilizing a chemical balance.

H do   H of  detonation products    H 0f  explosive 


(18)

Section 7 provides examples of these calculations for typical fuels. Alternatively, there are a
number of thermochemical kinetics, or thermal equilibrium, codes available to solve the
thermodynamic parameters, as well shock and detonation conditions. STANJAN [13], CHEMKIN
[14], CEA [15], and CHEETAH [16] are a few codes capable of evaluating the chemical species
and heats-of-formation.

6.3 Adiabatic Isentropic Gaseous Expansion

As an initial approximation to the explosive fuels, assume that the energy of explosion is an
adiabatic isentropic expansion of detonation gases [17] to atmospheric conditions, per Equation
19. That is, in this particular case, the assumption is that the contents of the reactions products in
the vessel will expand from the Chapman-Jouguet (C-J) pressure (i.e., ideal peak detonation
pressure) to ambient conditions through an adiabatic isentropic process. The energy, or work,
accomplished by the expansion of a gas at an initial volume, Vo , and pressure Po , into an
external atmosphere at pressure, Pe ;

  P  P  dV
e
E e (19)
o

PoVo
E k (20)
 1

 1
   1

     
1   e  
k  1        1 e
 Pe P P
  Po  
(21)
  Po   Po
   

Where:
E = Energy (kJ or MJ)
Pe = External (atmospheric) pressure
Po = Detonation peak pressure
Vo = Initial volume
cp
 = Isentropic expansion coefficient 
cv
k = Proportion of available gas energy converted to kinetic energy

17 ERI/NRC 09-203
Once the energy of explosion is determined, it becomes necessary to convert this energy to a
TNT equivalent energy for utilizing the scaled distance curves of TM 5-1300 [3] or Regulatory
Guide 1.91 [1]. For simplicity, the energy of explosion per unit volume (kJ/m3) will be considered,
such that:

E P
 o k (22)
Vo   1
  1
  1

E Po    Pe    Pe  P 
1  e  

or   1        1   Po   
(23)
Vo   1   Po   Po
     
In his seminal text, Baker [6] utilizes a slightly modified form for the blast energy assuming
isentropic gas expansion from a closed initial volume, Vo , to external atmospheric pressure, Pe :
 1

PV   P   P  

Eexp  e o
     
o e
(24)
  1  Pe   Po  
 
In comparison to Baum [17], Baker’s [6] results are higher for a given gas-mixture and associated
isentropic expansion coefficient. The expansion energy for a given volume of gas, can equated
to a TNT-equivalent weight by:

WTNT 
E 
exp Gas

 H o
d TNT

Where:
E 
exp Gas = Energy of expansion (kJ or MJ)

 H  o
d TNT = Heat-of-detonation (kJ/kg or MJ/kg)

Table 2 shows thermodynamic parameters, as derived from the NASA CEA code [15], for typical
hydrocarbon, i.e., gasoline (C8H18), combustion in air. Therefore, 1-mole of gasoline plus 12.5-
moles of oxygen makes this a balanced stoichiometric reaction.

C8 H18  12.5O2  8CO2  9 H 2O (25)

Table 2 Combustion Parameters for Stoichiometric Gasoline and Oxygen

Gas Composition Thermodynamic Parameters

%C8H18 %O2 MW
 
cp co H 
0
f CH
8 18
 H 
o
f CO  H O
2 2
(moles) (moles) (g/mol) cv (m/s) (kJ/mole) (kJ/kg)

1 12.5 114 1.111 1103 -213 -5722

Note: Initial pressure @ 1.0132 bar (1-atm).

18 ERI/NRC 09-203
Calculating the heat-of-combustion from the difference of heat-of-formation of products and
reactants, as shown in Eq. 17:
H co  5722  ( 213) (26)

H co  5509 kJ / mol (27)


Where:
co  Sonic velocity of unburned gas
H co  Heat-of-combustion
H of  Heat-of-formation
MW  Molecular weight (g/mol)

The total heat-of-combustion is:


5509 kJ / mol
H co 
114 g / mol (28)
H  48.3 kJ / g
o
c

Comparing this to the heat-of-detonation for TNT: H d  4.65 kJ


o
/ g , it becomes clear that
the ratio of heats-of-reaction is about a factor of 10. The TNT-equivalent weight would be:

H co
WTNT  Eq  WExp
H TNT
o
(29)
 48.3 
WTNT  Eq    WExp  10.4WExp
 4.65 
With actual heat-of-combustion for gasoline as calculated above, the safe distance to reach 1-psi
incident positive overpressure may be determined using Regulatory Guide curve as,
RRG  45W 1/3 , provided that the yield fraction, , associated with the vapor-phase is
determined. It is important to note that hydrocarbons stored at atmospheric conditions, or below
their boiling point, will combust only that portion within the vapor phase between the LFL and
UFL. Thus, the fraction available for immediate combustion is generally about 1/20th of the total
mass. Also, according to NPFA 325, the atmospheric boiling point for gasoline lies between 100 -
400°F (311-478K) depending on the grade of gasoline. Thus, it is incumbent upon the analyst to
ensure that flash point limits, boiling points, and other relevant thermodynamic data are taken into
consideration in the calculation of safe distances. For boiling points below atmospheric
conditions, flashing vapor quantity must be considered in the analysis. See Section 7.2 for
additional guidance in utilizing these parameters based on the FM Global Datasheets, and
calculations provided in the attached Appendix for gasoline combustion.

19 ERI/NRC 09-203
This page intentionally left blank

20 ERI/NRC 09-203
7. NUREG-1805 AND FM DATA SHEETS COMPARISON OF METHODOLOGIES

Section 6 briefly described and detailed the fundamentals of calculating the enthalpies of
reactants and reaction products, to achieve the heat-of-combustion of a mixture. NUREG-1805
[18] and Factory Mutual Data sheets [19] utilize this philosophy throughout in calculating the TNT-
equivalent energy of combustion/detonation. Calculating the mass of a flammable gas or the
mass of evaporative vapors from a liquid spill is not encompassed in the guidance provided by
Regulatory Guide 1.91. Several utilities have used the methodologies presented in NUREG-1805
and/or that which is presented in the Factory Mutual Data Sheets to calculate the mass of vapor
in a cloud. Presented below is a comparison of the methodologies:

7.1 NUREG-1805 (Confined Explosion-Leaking Flammable Gas)

One typical explosion in an enclosure is caused by flammable gas leaking, which mixes with air in
the enclosure and subsequently ignites to cause an explosion. The energy released by
expansion of compressed gas upon rupture of a pressurized enclosure may be estimated using
the following equation.

E  H co M f (30)
Where:
E = Explosive energy released (kJ)
 = Yield (i.e., the fraction of available combustion energy participating in blast wave
generation)
H co = theoretical net heat-of-combustion (kJ/kg)
Mf = mass of flammable vapor release (kg)

The yield,  , is typically in the range of 1 percent (0.01) for unconfined vapor releases, to 100
percent (1.0) for confined vapor releases.

One of the most common methods used to estimate the effects of an explosion is to relate the
exploding fuel to trinitrotoluene (TNT). This method converts the energy contained in the
flammable cloud into an equivalent mass of TNT, primarily because blast effects of TNT have
been extensively studied as a function of TNT weight and distance from the source. Hence, the
blast effects of an explosion can be inferred by relating an explosion to an “equivalent” explosion
of TNT. To do so, we relate a given fuel type and quantity to an equivalent TNT charge weight, as
follows:

E
WTNT  (31)
4500
Where:
WTNT = Weight of TNT (kg)
E = Explosive energy released (kJ)

Blast effects can also be related to the equivalent weight of TNT using by the relationship
between the distance from the source, the charge weight, and the overpressure caused by the
blast wave, including the reflected shock wave. Scaled distance is the distance at which the
overpressure is calculated divided by the cube root of the TNT charge weight.

21 ERI/NRC 09-203
D
Dsc  1/ 3
WTNT
(32)

Where:
Dsc = Scaled distance [m/(kg)1/3]
D = Distance at which the overpressure is calculated (m)
WTNT = Weight of TNT (kg)

7.2 Factory Mutual Data Sheets

The Factory Mutual Data Sheets (formerly Data Sheet 7-0S) [19] provides technical guidance to
evaluate the physical property damage consequences to buildings and process structures from
explosion overpressures and ensuing fires, caused by the outdoor release and delayed ignition of
a flammable vapor cloud.

The FMDS uses a TNT equivalency methodology to evaluate approximate effects of a worst
credible case VCE. Other refined VCE prediction or estimation methods, while briefly discussed,
are beyond the scope of the data sheets.

The techniques and procedures described in the guidelines are a simplified approach to a
complex problem. Application is not appropriate for non-chemical plants that may have storage or
use of hazardous materials, or for smaller chemical plants that lack the congestion or confined
layout needed to produce outdoor VCEs.

Loss history supports that VCEs have occurred primarily in large petrochemical or refinery
facilities and/or as a result of transportation incidents and that, they have occurred on a very low
frequency as compared to other major events.

7.2.1 Gas Releases

If the material exists in the system as a gas, the following equation can be used to estimate the
mass of gas released, Wg , from a break.

Wg  KCd Ar t 2  l Pd
(33)

Where:
Wg = Mass of gas released (kg)
Ar = Area of release opening (m2)
Cd = Discharge coefficient (use 1.0)
P1 = Process or reservoir pressure (Pa)
Pa = Ambient pressure (Pa) (absolute) (@ sea level Pa =1.014×105 pa)
l = Vapor density at process conditions (kg/m3)
t = Discharge duration (sec) (use 600)
K = Gas Constant. K usually falls in the range 0.63 to 0.73. An average value of
K  0.68 may be used with little error when calculating vapor or gas flow rates for most
materials.

22 ERI/NRC 09-203
Pd = Use actual process or reservoir pressure  P1  when P1 is in excess of 20 psi (135
kPa) (sonic flow). For pressures less than this subsonic flow, Pd  P1  Pa (Pressure- Pa
absolute).

7.2.2 Liquid Releases

If the material exists in a system as a liquid, the mass of the liquid released Wl  , assuming
gravity and/or vessel pressure as the driving force with no vaporization (two-phase flow) in the
orifice, can calculated by Equation 32:

2  P1  Pa 
Wl  Cd  l Ar t  2 gh (34)
l
Where:
Wl = Mass of liquid released (kg)
Ar = Area of release opening (sq m)
Cd = Discharge coefficient (use 0.62)
g = Gravitational constant (9.81 m/s2)
h = Height of liquid in tank above discharge point (m)
Pa = Ambient pressure (Pa absolute)
P1 = Process or reservoir pressure (Pa absolute)
l = Density of Liquid (kg/m3) @ process temperature  T1 
T = Discharge Duration (sec) (Use 600)

7.2.3 Calculation of Initial Flash Fraction

The pressurized liquid will flash once it has escaped and is at atmospheric pressure. The heat
required for vaporization is taken from the liquid itself so that any liquid which is left will have
been cooled to its atmospheric boiling point. The initial flash fraction Fvap of liquid is given by:

CP1  T1  Tb 
Fvap  (35)
H vapb
Where:
Fvap = Fraction of liquid flashed to vapor
CP1 = Specific heat of liquid at constant pressure (J/kg° K) averaged between process
temperature T1 and atmospheric BP. If range figures not available, use CP1 at boiling
point to be conservative.
H vapb = Heat of vaporization (J/kg) at boiling point
Tb = Atmospheric Boiling point (°K or °C)
T1 = Temperature of liquid in vessel (°K or °C)

23 ERI/NRC 09-203
As suggested in the FM Global Guidelines, Fvap is doubled to account for aerosol mist in the
cloud. Therefore if the initial flash fraction Fvap is 50% or greater, the entire liquid contents can be
assumed to vaporize, and Wl equals Wv and then Equation 37 can be used in conjunction with
Wv to solve for the mass of equivalent TNT.

If Fvap is less than 50%, then that portion of Wl that is not vaporized will rainout into a ground
pool. The next step, Equation 34, can be used to calculate the amount of this ground pool that is
subsequently vaporized due to thermal and atmospheric effects. For example, if the initial flash
fraction, Fvap , is calculated to 15% and is doubled to 30% to account for aerosol mist, then 70%
of the initial liquid release, Wl , will rain out into the liquid pool.

For a more accurate prediction of initial flash fraction, a source term computer model can be
used.

7.2.4 Calculation of Liquid Pool Size

The size of the liquid pool will be a factor in the evaporation rate. The extent of an unconfined spill
on a relatively non-porous surface can be calculated as follows:

A  t gVo (36)

Where:
A = Spill area (m2)
g = Gravitational constant (use 9.81 m/s2)
t = Time (Use 600 even if discharge duration is less) (Sec)
Vo = Volume spilled (m3)

For unconfined spills on essentially flat surfaces, the area of the spill should be limited so that the
pool depth is not less than 6 mm (1⁄4 in.). For a confined spill, such as inside a diked or curbed
area, use the ground surface as well as vertical surfaces of walls up to the maximum level of
liquid. When a dike or process area has drainage, the pool calculation becomes difficult. While
drainage is favorable and required for controlling pool fires, it may be of little help in limiting vapor
formation. The hot liquid can continue to vaporize within the drainage system, often more rapidly
due to standing water and vapor can be released to the process area environment through the
drainage system openings. The presence of drainage within a dike or curbed area should
generally be ignored and the full amount spilled should be used for vaporization calculations.
However, for very large capacity drainage systems designed specifically to both rapidly remove
spilled material to a safe remote location and to retard vapor formation to the atmosphere, credit
can be given after using considerable engineering judgment.

7.2.5 Calculating of Evaporation from a Liquid Pool

Any material discharged as a liquid, if not initially flash vaporized or carried into the cloud as
aerosol droplets, will form a pool on the ground and vaporization of the boiling pool may occur. If
the atmospheric boiling temperature of the discharged liquid is below the ambient temperature,
the liquid will continue boiling after it is spilled on the ground. The heat necessary for boiling is
supplied by conduction heat transfer from the ground.

24 ERI/NRC 09-203
The amount of liquid evaporated Wboil  by ground heat transfer is calculated using Equation 35.

1/2
2B  t 
Wboil    Ta  Tb  A (37)
H vapb   

Where:
Wboil = Amount of liquid (kg) evaporated, not to exceed the amount of liquid spilled
B = Thermal property of spill surface (Table 3, Section 7-42, Page 21 of the FMDS)
t = Time (Sec) (use 600 even if discharge duration is less)
A = Area of the spill (m2) in contact with solid surface.
Ta = Ambient temperature (°K or °C) (Use actual summer high for ambient temperature)
Tb = Atmospheric Boiling point of liquid (°K or °C)
H vapb = Heat of vaporization (J/Kg) at atmospheric boiling point

If the liquid boiling point is above ambient, this equation is no longer applicable. According to
NFPA 325, gasoline has a boiling point between 100-400°F (311-478K) * , depending on the grade
of gasoline. For example; assume gasoline, whose boiling point is at the lower-limit of 100°F, is
stored in a hot summer day with air temperatures at, or below, 100°F, no amount of liquid
evaporation will take place, and Equation (37) is invalid. Ambient atmospheric temperatures
would need to exceed the boiling point in order to achieve evaporation. In this situation,
diffusional evaporation is the controlling factor. Equations to determine this rate are highly subject
to atmospheric conditions, particularly wind speed. Since wind speed is not known, this factor is
excluded from this simplified approach. A computer model must be used to predict the release
under this type of conditions.

7.2.6 Calculation of Total amount of Vapor in a Cloud

The amount of material used to solve the TNT equivalency equation (Equation 37), should be
either Wg (contents initially gas) or Wv (contents initially liquid or mixed liquid vapor). Wv can be
calculated by adding the amount of released liquid Wl (Equation 32) to that which flashes
(Equation 33) and the amount evaporated from the liquid pool Wboil (Equation 35).

Wv  2 FvapWl  Wboil (38)

Note that the initial flash fraction Fvap is doubled to account for aerosolization. Doubling the Fvap
cannot exceed unity (1.0) and Wv cannot exceed the total mass of material initially released from
the vessel.

*
Note: For gasoline types, whose boiling point ranges vary between 70-400°F, the amount of evaporation
would be significant on a hot summer day. For conservatism, the highest temperature difference Ta  Tb  ,
i.e., the lowest boiling point of chemical Tb  and highest ambient temperature Ta  may be used.

25 ERI/NRC 09-203
7.2.7 Calculating TNT Equivalency

The energy released in an explosion of a vapor cloud is expressed as a TNT equivalent. This
methodology is often called the Ideal Blast Wave method. Based on an approximate energy of
decomposition for TNT of 2000 Btu/lb, the following may be used to calculate a TNT equivalent
for a vapor cloud containing a known weight of flammable gas Wg or vapor Wv in SI Units:

H co f
We  W (39)
1.11x106
Where:
We = Mass of Equivalent TNT Energy Yield (t)
W = Mass of vapor in cloud (kg) of gas (Wg) or vapor (Wv)
H co = Heat of combustion of material (Kcal/kg)
f = Explosive yield (efficiency) factor (Section 3.4.3, Section 7-42, Page 22 of the FMDS)

The regulatory Guide 1.91 guidance is dated and requires revision to encompass the differing
scenarios that need to be addressed by the utilities. The FMDS appear to be the best resource
for those scenarios not captured in Regulatory Guide 1.91 and the guidance needs to be revised
to reflect the additional augmenting methodologies.

26 ERI/NRC 09-203
8. BOUNDING METHODOLOGIES

The NRC has posed the following question:

“What methodologies could be used by utilities, for detonable solids, liquids


and gases, that would provide bounding safe distance values for those
materials that are being transported or stored in close proximity to a Nuclear
Power station such that the associated peak positive incident overpressure
from the detonation never exceeds 1 psi at the critical facility being analyzed.”

Provided in this section is a discussion regarding the difficulties in calculating universal bounding
values for every scenario and type of detonable material analyzed.

8.1 Solid Materials

For solid materials and hydrocarbons liquefied under pressure, the methodologies and maximum
probable quantity of hazardous cargo transported, as presented in Regulatory Guide 1.91, would
provide bounding safe standoff values. The maximum probable quantity of hazardous cargo
transported or stored is dependent on the transportation mode, the transportation vehicle used, or
the storage method. As suggested in Regulatory Guide 1.91, the maximum probable hazardous
solid cargo for a single highway truck is 50,000 pounds (23,000 kg). Similarly, the maximum
explosive cargo in a single railroad boxcar is approximately 132,000 pounds (60,000 kg). The
largest probable quantity of explosive material transported by ship is approximately 10,000,000
pounds (4,500,000 kg). When shipments are made in connected vehicles such as railroad cars or
barge trains, an investigation of the possibility of explosion of the contents of more than one
vehicle is necessary. However, it must be understood that heat-of-reaction for each of these
different hazardous material quantities are correlated to TNT-equivalence.

Regulatory Guide 1.91 does provide some rudimentary bounding information regarding Vapor
Clouds and as stated in the guide, “A reasonable upper bound to the blast energy potentially
available based on experimental detonations of confined vapor clouds is a mass equivalence of
240 percent.”

8.2 Liquids

If the material in the vessel exists as a liquid, the amount discharged into a confined or
unconfined area, and the fraction that will flash vaporize and evaporate can be estimated using
the methodologies illustrated in section 7.2. Calculating one bounding scenario that would
comfortably encompass most is very difficult based on the all the variables discussed in section
7.2. For example, the results of the calculations in section 7.2 depend heavily on the
assumptions used which include; amount released, discharge coefficient, release duration,
density of the discharged liquid, fraction of the liquid that flashes to vapor, spill area, thermal
property of the spill surface, ambient temperature, heat of vaporization, heat of combustion, and
explosive yield factor.

8.3 Vapor Clouds

A vapor cloud explosion is defined as an explosion occurring outdoors which produces damaging
overpressure. It is initiated by the unplanned release of a large quantity of flammable vaporizing
liquid or high-pressure gas from a storage tank or system, process vessel, pipeline, or
transportation vessel.

27 ERI/NRC 09-203
8.3.1 Theory and Principle

Generally speaking, for a VCE with damaging overpressure development to occur, several factors
must be present.

First, the material released must be flammable and processed or held under suitable conditions of
pressure or temperature. Examples of such materials are liquefied gases under pressure (e.g.,
propane, butane,); ordinary flammable liquids at high temperatures and/or pressures (e.g.,
cyclohexane, naptha) and non-liquefied reactive flammable gases (e.g., ethylene, acetylene).

Second, a cloud of substantial size and concentration also must form prior to ignition. With most
common flammable materials, should ignition occur instantly with release of the material, a large
vapor cloud fire may occur, causing extensive localized heat radiation damage; however,
significant blast pressures causing widespread damage will likely not occur. (Exception: some
highly reactive materials, such as ethylene oxide under some conditions, might produce
overpressures even with immediate ignition.) Should the cloud be allowed to form over a period of
time within a confined process area and subsequently ignite, blast overpressures away from the
cloud center can equal or even exceed those developed from detonation of high explosives and
result in extensive damage over a wide area. Ignition delays of from one to five minutes are
considered most probable, although major incidents with ignition delays as low as a few seconds
and higher than 60 minutes have occurred.

Third, a sufficient amount of the cloud must be within the flammable range of the material to
cause extensive overpressure. The percent of the vapor cloud in each region varies, depending
upon many factors including type and amount of the material released, pressure at release, size
of release opening, direction of release, degree of outdoor confinement of the cloud, and wind
speed, atmospheric stability and other environmental effects. The cloud will move over time,
changing the flammable regions. For example, a continuous release over a long period of time
will generally have a rich region near the source, a lean region at the cloud leading edge, and a
flammable region in between. A puff release (essentially instantaneous release) will usually have
a rich region at the leading edge with flammable regions following.

Important factors that must be present for an ignited vapor cloud to produce overpressure are
outdoor confinement and turbulence generation. Research testing, incident investigation, and
computer modeling have demonstrated that the greater the horizontal and vertical confinement
and the more turbulence in the gas cloud, the greater the potential for overpressure development.
Turbulence can be caused by two primary mechanisms. First, repeated obstacles in the center of
a cloud can accelerate gas mixing (due to eddy and shear layer effects), which in turn can
increase flame speeds within the cloud. This highly influences pressure development due to
flame instabilities. Second, turbulence can be directly initiated from a high-pressure release.

Because obstacles cause turbulence, obstacles between the material and the plant play a
principal role in the ability of a released vapor cloud to burn as a flash fire with only radiant heat
effects or transit to an explosion with overpressure effects as well as radiant heat effects. The
amount of congestion, confinement, the horizontal and vertical spatial arrangement of obstacles
in the flow path of a cloud, are all important in determining if an ignited cloud of vapor will transit
from a fire to an explosion.

Wide-open spaces between the released material and the critical structure do not easily promote
VCE events unless the area presents unusual conditions of confinement (such as long, narrow
ravines) or repeated obstacles (such as dense forests or large railroad staging yards). However,
a cloud released in an open area may be of sufficient size and winds may be of suitable velocity
and direction to disperse the cloud into a congested process area at great distances from the
actual release. While the ignition of the cloud could occur anywhere in the cloud (even in the

28 ERI/NRC 09-203
open space, for example, by a vehicle), the apparent explosion epicenter will be the area where
the cloud is confined and where obstacles exist that can cause transition from a cloud fire to a
cloud explosion. Remaining portions of the cloud outside the congested area will not contribute to
blast effects, although radiant heat effects will occur.

8.3.2 Materials not Conducive to VCEs

Throughout the industry, it is generally accepted that the following materials do not present a
significant or credible outdoor VCE exposure. These exclusions are based on many factors such
as heats of combustion, fundamental burning velocities, research testing, ease of dispersal, other
experts’ opinions and loss history.

8.3.2.1 Liquefied Natural Gas (LNG) and Natural Gas (NG) (methane)

LNG or NG, when the ethane component is less than 15% by volume. If LNG or NG has an
ethane component in excess of 15%, it may be susceptible to VCE with overpressure
development. However, with 85% or greater methane concentration, an explosion outdoors with
damaging overpressures is not considered likely. Normal NG or LNG for fuel use is composed of
92–94% methane, 3–4% ethane and other hydrocarbons, and 3% nitrogen. While this varies by
region, supplier and time of year, rarely would ethane content exceed 5% due to cost and value
of ethane for other uses. The only condition where LNG or NG might have high ethane content
would be where it is used as a pure feedstock in chemical processing. For all practical purposes,
the ethane content of NG and LNG in pipelines and storage systems will never exceed 5%, and
will normally be much lower. The validity of excluding LNG or NG is supported by testing by the
Institution of Gas Engineers (Great Britain) in small- and full-scale obstacle arrays simulating
natural gas processing units. While it was concluded that NG with high methane content could
produce moderate overpressures if released into process arrays of extremely close packed
obstacles, it was further stated that the tests “demonstrated that both the probability of a vapor
cloud explosion occurring and its consequences will be lower for natural gas than with other
common hydrocarbons.” Further, there have been no reported VCEs involving LNG or NG with
high methane content, even though numerous ignited vapor releases resulting in flash fires
without overpressures have occurred worldwide. Coupling testing and loss history with the low
reactivity, the lightness of the vapor, and the relatively low flame speeds of LNG or NG, a VCE
involving these materials is considered beyond the scope of a worst credible case scenario.

8.3.2.2 Ammonia Gas

Ammonia Gas can and has exploded when substantially confined inside equipment or buildings.
There is no indication, based on loss history, testing or combustion features, that ammonia will
produce overpressures if released and ignited outdoors.

8.3.2.3 Gaseous Hydrogen

Hydrogen gas such as in tube trailers, in a pipeline, or in a process, regardless of system


pressure, should not be considered to present an outdoor VCE potential. Only one outdoor
explosion incident involving hydrogen is widely reported. This involved a small release of high-
pressure gaseous hydrogen during an acoustical test at the US Department of Energy Nevada
Test Site. The extremely turbulent aerial cloud ignited, probably due to static energy caused by
the turbulence of release. The overpressures were low and caused minimal damage to test
equipment and structures. According to large users of gaseous hydrogen, aerial clouds are
occasionally released from hydrogenation units and other processes with slightly delayed ignition.
However, these so called ‘‘aerial detonations’’ have not caused significant far-field overpressures
and have resulted in localized damage. For this reason, gaseous hydrogen is not deemed an
unusually severe or credible VCE exposure.

29 ERI/NRC 09-203
8.3.2.4 Miscellaneous Flammable or Combustible Gases

Such as ammonia synthesis gas (a hydrogen/carbon monoxide mix), coal and blast furnace
gases, methylene chloride, and trichloroethylene are excluded from this document, because of
low flame speeds and heats of combustion and lack of loss history.

8.3.2.5 Flammable Liquids or Gases processed above their Auto-ignition Temperature

Flammable liquids or gases processed above their auto-ignition temperature will immediately
ignite on contact with air. A severe flash fire may result, but delayed cloud ignition, necessary for
development of significant overpressure, will not occur with these materials.

8.3.2.6 Flammable Liquids (High Viscosity)

Flammable liquids having a high viscosity (greater than 1×105 centipoises) will likely not present
normal vapor formation and will form pools of non-vaporizing liquid rapidly.

8.3.2.7 Mixtures

Mixtures of two or more liquid or gaseous materials in a process system may occur, creating
difficulty in choosing a material class, calculating release, etc. There is no easy solution for
calculating mixtures, even with advanced computer models. Considerable judgment must be
used. Generally, the more hazardous material should be selected as if it were the entire volume.
If equal amounts exist, several calculations may be necessary to determine a worst credible
event. If the actual heat of combustion of the mixture is known, it may be used in energy
equations.

8.3.2.8 Hybrid Mixtures

Hybrid Mixtures are a mixture of dust in a flammable gas medium. VCEs have occurred in hybrid
systems, notably at polyolefin manufacturing facilities. Some VCE researchers feel suspended
polymer dust in a gas cloud might contribute to overpressure development. While possible, there
is no calculation procedure or model known to accurately factor in hybrid mixtures.

8.3.3 VCE Methodologies

The conditions necessary to produce a VCE are fairly well understood, and a number of
calculation methods are available to convert the VCE scenario into an assessment of damage
effects. Most calculation techniques include methods for determining amount of material
released, cloud size and energy release upon ignition.

The energy of the material released is often converted into TNT equivalency (following principles
of the so-called Ideal Blast Wave methodology), by assigning an explosion efficiency number.
Also referred to as explosive yield, explosion efficiency is an estimation of the explosive effect of
the mass in the cloud relative to an equivalent mass of TNT. Once a TNT equivalency is
determined, published test data is used to calculate blast overpressures. Other methods have
been developed and published to evaluate VCE overpressure effects. Section 7.2 abstracts
several of the more well-known methods, including the TNT equivalency method.

Calculating one bounding scenario that would comfortably encompass most is very difficult based
on all the variables discussed in this section. Calculation of vapor cloud releases and
consequences can either be performed in a spread sheet or be estimated by using one of many
computer models available (for example ALOHA [2]). The primary feature of the more
sophisticated modeling codes is their ability to more accurately calculate source term release,

30 ERI/NRC 09-203
cloud dispersion and drift, primarily because they factor in wind speed and atmospheric stability
conditions.

Many, however, apply a TNT Equivalency method for determining energy release, which requires
the practitioner to determine a credible release scenario and assign variables such as explosion
efficiency and blast epicenter.

8.4 Additional Literature

The Center for Chemical Process Safety (CCPS) of the American Institute of Chemical Engineers
(AIChE) has devoted a tremendous amount of effort into development of guidance to assess
vapor-cloud explosions, flash fires, and BLEVES. General engineering solutions to these
phenomena are provided along with detailed analyses of specific hazardous release conditions.
The CCPS has two volumes that are widely used in the process industry to assess these
phenomena through much of the same philosophy as described in the body of this report:

(1) “Guidelines for Evaluating the Characteristics of vapor Cloud Explosions, Flash Fires,
and BLEVES,” [20] and

(2) “Guidelines for Use of Vapor Cloud Dispersion Models” [21].

These two volumes were a collaborative effort between the AIChE and TNO Prins Maurits
Laboratory in the Netherlands.

31 ERI/NRC 09-203
This page intentionally left blank

32 ERI/NRC 09-203
9. CONCLUSIONS

As stated in the introduction section of this document, according to the Regulatory Guide 1.91,
postulated accident analyses on routes near nuclear power plants must be addressed for all
known combustible or detonable fuels utilizing the methodologies presented in the guide.

However, Regulatory Guide 1.91 addresses only solid detonable materials, hydrocarbons
liquefied under pressure, and air blasts on highway, rail, and water routes. The regulatory guide
does not address liquids (pressurized or non-pressurized), cryogenically liquefied hydrocarbons
(LNG, CNG, propane), vapor clouds (confined or unconfined), fixed facilities, pipelines and other
scenarios.

Figure 6 shows the Regulatory Guide 1.91 design curve for radial distance to TNT weight, based
on peak incident overpressure of 1-psi. It is interesting to note that the examples noted on the
graphical plot imply these are TNT weights for solid materials, yet there is no indication of how to
arrive at these weights for other substances or mixtures (i.e., gaseous or liquid). Proper
treatment for other hazardous and detonable mediums requires fundamental knowledge of the
system thermodynamics, and specifically thermochemistry of mixtures.

The Regulatory Guide 1.91 guidance is dated and requires revision as it is applicable to only solid
explosives and hydrocarbons liquefied under pressure. Since Regulatory Guide 1.91 does not
address other scenarios utilities are forced to utilize the methodologies inherent in the regulatory
guide for the various bounding scenarios or seek out other sanctioned industry guidance to
address these additional scenarios.

The South Texas Project, lacking sufficient guidance in Regulatory Guide 1.91, sought out other
sanctioned industry guidance to address these additional scenarios in order to provide a
bounding analysis to the NRC in their FSAR COLA submittal. Their submittal resulted in an RAI,
which is summarized in the following text:

“The minimum safe distance values shown in Table 2.2S-9 are said to be
based on TNT equivalency method using Regulatory Guide (RG) 1.91
methodologies. But they seem smaller than generally expected. Please
explain the methodology in detail.”

The South Texas Project, in response to the RAI, provided a detailed description of the methods
utilized for different sources of explosion phenomena such as, atmospheric liquids, liquefied
gases, and gases. The South Texas Project described the methodologies and approach to
ascribing minimum safe-distances from explosion sources, utilizing methods in NUREG-1805 and
the Factory Mutual Data Sheets.

ERI utilized the methodologies presented in this document to evaluate a few of the more
detonable materials selected by STP in their analysis. The attached Appendix contains a
comparison of ERI calculated safe distances versus STP calculated safe distances for those
detonable materials. The results are presented in Table 3 below.

It is concluded, based on this technical independent review, that the South Texas Project
engineering staff appropriately and correctly applied substance-specific thermodynamic data to
obtain the heats-of-combustion in calculating the TNT-equivalent explosion energy for different
fuels/oxidizer mixtures.

33 ERI/NRC 09-203
Table 3 Comparison of Calculated Safe Distances (ERI versus STP)
ERI Calculated Value ERI Calculated Value STP Calculated
Detonable Material
(m) (ft) Value (ft)

Gasoline 72 235 266

Hydrogen 326 1070 1047

Hydrazine 13 42 86

Ethylene 2293 7537 7575

Acetic Acid 230 754 814

34 ERI/NRC 09-203
10. RECOMMENDATIONS

It is recommended that the Regulatory Guide 1.91 guidance be updated as it only addresses
solids or hydrocarbons liquefied under pressure. Furthermore, it is recommended that Regulatory
Guide 1.91 be revised to either provide a tangible link to those sanctioned industry methodologies
that encompass scenarios not included in the guidance provided by Regulatory Guide 1.91, or
incorporate the guidance provided in NUREG-1805, Factory Mutual Data Sheets, or other
acceptable methodologies approved by the NRC (e.g. modeling codes, AIChE recommendations,
international standards/methodologies, etc.).

35 ERI/NRC 09-203
This page intentionally left blank

36 ERI/NRC 09-203
11. REFERENCES

1. “Evaluation of Explosions Postulated to Occur on Transportation Routes Near Nuclear Power


Plants,” Nuclear Regulatory Commission, Regulatory Guide 1.91, Washington, D.C.,
February 1978.

2. "ALOHA User's Manual: Aerial Locations and Hazardous Atmospheres," Version 5.4.1, US
Environmental Protection Agency (EPA), Office of Environmental Management, Washington,
D.C. and U.S. National Oceanic and Atmospheric Administration (NOAA), Emergency
Response Division, Seattle, Washington, 2007.

3. “Structures to Resist the Effects of Accidental Explosions,” U.S. Army, Navy, and Air Force,
U.S. Department of Defense, TM 5-1300, Revision 1, Washington, D.C., 1990.

4. “A Manual for the Prediction of Blast and Fragment Loadings on Structures,” U.S.
Department of Energy, DOE/TIC-11268, Albuquerque Operations Office, Amarillo Area
Office, Pantex Plant, Albuquerque, New Mexico, 1980.

5. B. Dobratz, “LLNL Explosives Handbook: Properties of Chemical Explosives and Explosive


Simulants,” Lawrence Livermore National Laboratory, University of California, UCRL-52997,
Livermore, California, 1981.

6. E. E. Baker, Explosions in Air, Wilfred Baker Engineering, San Antonio, Texas, 1973.

7. P. W. Cooper, Explosives Engineering, VCH Publishers, Inc., New York, 1996.

8. SFPE Handbook of Fire Protection Engineering, Society of Fire Protection Engineers,


National Fire Protection Association, Quincy, Massachusetts, 2002.

9. V. Brabauskas, Ignition Handbook, Society of Fire Protection Engineers, Fire Science


Publishers, Issaquah, 2003.
th
10. Robert C. Weast (Ed.), Handbook of Chemistry and Physics, 55 Edition, CRC Press,
Cleveland, Ohio, 1975.

11. I. Glassman, Combustion, 3rd Edition, Academic Press, New York, New York 1996.

12. David Lide. s.l (Ed), JANAF Thermochemical Tables, American Chemical Society and
American Institute of Physics for the National Bureau of Standards, Journal of Physical and
Chemical Reference Data, 1985..

13. W. C. Reynolds, “STANJAN Thermochemical Code,” Department of Mechanical Engineering,


Stanford University. Stanford, California, 1987.

14. “CHEMKIN,” Reaction Design, Inc., San Diego, CA, 2008.

15. S. Gordon, B. McBride, “Computer Program for Calculation of Complex Chemical Equilibrium
Compositions and Applications,” National Aeronautic and Space Administration (NASA),
NASA Reference Publication RP-1311, May 2004.

16. L. Fried, L., C. Souers, M. Howard, “CHEETAH: A Next Generation Thermochemical Code,”
Ver. 2.0. Lawrence Livermore National Laboratory University of California, UCRL-ID-117240,
Livermore, California, 1994.

37 ERI/NRC 09-203
17. M. R. Baum, S. J. Brown [Ed.], “The Velocity of Missiles Generated by the Disintegration of
Gas Pressurized Vessels and Pipes,” American Society of Mechanical Engineers, 1984
Pressure Vessels and Piping Conference and Exhibition, PVP-Vol. 82, pp. 67-83, San
Antonio, Texas, 1984.

18. “Fire Dynamics Tools (FDT's): Quantitative Fire Hazard Analysis Methods for the US Nuclear
regulatory Commission Fire Inspection Protection Program,” Nuclear Regulatory
Commission, NUREG-1805, Washington, D.C., 2004.

19. “Guidelines for Evaluating the Effects of vapor Cloud Explosions Using a TNT Equivalency
Method,” Factory Mutual Global, Property Loss Prevention Data Sheet 7-42, Boston,
Massachusetts, 2006.

20. “Guidelines for Evaluating the Characteristics of Vapor Cloud Explosions, Flash Fires, and
BLEVES,” American Institute of Chemical Engineers, Center for Chemical Process Safety,
Park Ave., New York, New York.

21. “Guidelines for Use of Vapor Cloud Dispersion Models,” American Institute of Chemical
Engineers, Center for Chemical Process Safety, Park Ave., New York, New York.

22. “Guide to Fire Hazard Properties of Flammable Liquids, Gases, and Volatile Solids,” National
Fire Protection Association, NFPA 325, Quincy, Massachusetts, 2001.

38 ERI/NRC 09-203
APPENDIX: CONFIRMATORY CALCULATIONS

39 ERI/NRC 09-203
GASOLINE (Liquid)

ΔHc (Kcal/kg) - Gasoline 10,437.60 From the internet


Explosion efficiencies listed in paragraph 3.4.3 of the FMDS
(page 22) are based on historical evidence and on literature
classification of materials by several different sources, including
f - Explosive Yield Factor
Dave J. Lewis of ICI (England), Unconfined Vapor Cloud
(Class I Materials FMDS)
Explosion—Historical Perspective and Predictive method Based
on Incident Records, (1980) Prog. Energy Comb. Sci., VC, pp
0.05 151-165.
Vapor density of gasoline
(lbs/ft3) 0.2516 Density of air (0.074 lb/ft3) x Specific Gravity of Gasoline (3.4)

Mass in
Vapor Cloud
Material at Risk Amount of Liquid Released Units (kg) Assumptions
9,000 gallons of gasoline leaks from confinement and spills to
surrounding asphalt. The liquid is assumed to flash once it has
Gasoline 9000 Gal 137.3 escaped confinement and is vaporized and concentrations in the
vapor are between the LFL and the UFL. The vapor remains
confined in a cloud and is not dispersed and detonates.

 H co f
Equation 7 from FMDS We  W
1.11x10 6
Where:
We Mass of Equivalent TNT Energy
Yield (Tonnes)
W Mass of vapor in cloud (kg)

ΔHc Heat of combustion of material


(Kcal/kg) for gasoline

f Explosive yield (efficiency)


factor (section 3.4.3 of FMDS)

40 ERI/NRC 09-203
GASOLINE (Liquid)

Tonnes kg
Then We = 0.065 64.6

From Regulatory Guide 1.91 R RG  17.85W 1/3

Where:

Radial Distance (m) where


RRG
pressure wave is less than 7
KPa (1 psi)
W Charge mass, (kg)

ERI Calc.
ERI Calc. (m) (ft) STP Calc. (ft)
Then RRG = 71.5 234.5 266

41 ERI/NRC 09-203
HYDROGEN (GAS)
ΔHc (BTU/lb) - H2 51,600.00 Taken from Table 1 - FM Datasheets

Vapor density of H2
(lbs/ft3) 0.005254 Density of air (0.074 lb/ft3) x Specific Gravity of H2 (0.071)

Mass in
Vapor Cloud
Material at Risk Amount of Hydrogen Released Units (lb) Assumptions
100,200 ft3 of hydrogen escapes confinement and is immediately
Hydrogen gas 100,200 ft3 526.5 mixed with air. The gas remains confined in a vapor cloud and is
not dispersed.

H co
We  W
From ERI TER  H of  TNT lb

Where:
We Mass of Equivalent TNT Energy
Yield (lbs)
W Mass of vapor in cloud (kg)
ΔHc Heat of combustion of material
(BTU/lb) for H2

lbs kg
Then We = 13582.431 6162.6

From Regulatory Guide R RG  17.85W 1/3


1.91

Where:

42 ERI/NRC 09-203
HYDROGEN (GAS)
Radial Distance (m) where
RRG
pressure wave is less than 7
KPa (1 psi)
W Charge mass, (kg)

ERI Calc.
ERI Calc. (m) (ft) STP Calc. (ft)
Then RRG = 326.3 1070.3 1047

43 ERI/NRC 09-203
Hydrazine (Liquid)

ΔHc (Kcal/kg) - Hydrazine 4,000.00 From the internet


Explosion efficiencies listed in paragraph 3.4.3 of the FMDS (page 22)
are based on historical evidence and on literature classification of
f - Explosive Yield Factor
materials by several different sources, including Dave J. Lewis of ICI
(Class III Materials
(England), Unconfined Vapor Cloud Explosion—Historical Perspective
FMDS)
and Predictive method Based on Incident Records, (1980) Prog.
0.15 Energy Comb. Sci., VC, pp 151-165.
Vapor density of
Hydrazine (lbs/ft3) 0.074 Density of air (0.074 lb/ft3) x Specific Gravity of Hydrazine (1.0)
Density of Hydrazine
(lbs/ft3) 62.93

Mass in
Vapor Cloud
Material at Risk Amount of Liquid Released Units (kg) Assumptions
Approximately 150 gallons of Hydrazine leaks from confinement and
spills to surrounding asphalt. The liquid is assumed to flash once it has
Hydrazine 1260 lbs 0.7 escaped confinement and is vaporized and concentrations in the
vapor are between the LFL and the UFL. The vapor remains confined
in a cloud and is not dispersed and detonates.

 H co f
Equation 7 from FMDS We  W
1.11x10 6
Where:
We Mass of Equivalent TNT Energy
Yield (Tonnes)
W Mass of vapor in cloud (kg)

ΔHc Heat of combustion of material


(Kcal/kg) for Hydrazine

44 ERI/NRC 09-203
Hydrazine (Liquid)

f Explosive yield (efficiency)


factor (section 3.4.3 of FMDS)

Tonnes kg
Then We = 3.63E-04 3.63E-01

From Regulatory Guide R RG  17.85W 1/3


1.91

Where:

Radial Distance (m) where


RRG
pressure wave is less than 7
KPa (1 psi)
W Charge mass, (kg)

ERI Calc.
ERI Calc. (m) (ft) STP Calc. (ft)
Then RRG = 12.7 41.8 86

45 ERI/NRC 09-203
Ethylene (Gas)

ΔHc (BTU/lb) - Ethylene 20,300.00 Taken from Table 1 - FM Datasheets

Vapor density of Ethylene


(lbs/ft3) 0.0932 Density of air (0.074 lb/ft3) x Specific Gravity of Ethylene (1.26)

Mass in
Vapor Cloud
Material at Risk Amount of Ethylene Released Units (lb) Assumptions
470,000 lbs of Ethylene gas escapes confinement and is immediately
Ethylene gas 470,000 lbs 4.70E+05 mixed with air. The gas remains confined in a vapor cloud and is not
dispersed.

H co
We  W
From ERI TER
 H of  TNT
Where:
Mass of Equivalent TNT Energy
We
Yield (lbs)
W
Mass of vapor in cloud (kg)

ΔHc Heat of combustion of material


(BTU/lb) for Ethylene

lbs kg
Then We = 4.77E+06 2.16E+06

From Regulatory Guide


R RG  17.85W 1/3
1.91

Where:

46 ERI/NRC 09-203
Ethylene (Gas)

Radial Distance (m) where


RRG pressure wave is less than 7
KPa (1 psi)
W Charge mass, (kg)

ERI Calc.
ERI Calc. (m) (ft) STP Calc. (ft)
Then RRG = 2297.8 7536.8 7575

47 ERI/NRC 09-203
Acetic Acid (Liquid)

ΔHc (Kcal/kg) - Acetic


Acid 3,137.00 From the internet
Explosion efficiencies listed in paragraph 3.4.3 of the FMDS (page 22)
are based on historical evidence and on literature classification of
f - Explosive Yield Factor materials by several different sources, including Dave J. Lewis of ICI
(Class II Materials FMDS) (England), Unconfined Vapor Cloud Explosion—Historical Perspective
and Predictive method Based on Incident Records, (1980) Prog.
0.1 Energy Comb. Sci., VC, pp 151-165.
Vapor density of Acetic
Acid (lbs/ft3) 0.078 Density of air (0.074 lb/ft3) x Specific Gravity of Acetic Acid (1.05)

Mass in
Vapor Cloud
Material at Risk Amount of Liquid Released Units (kg) Assumptions
500,000 gallons of acetic acid leaks from confinement and spills to
surrounding asphalt. The liquid is assumed to flash once it has
Acetic Acid 500,000 Gal 7628.2 escaped confinement and is vaporized and concentrations in the
vapor are between the LFL and the UFL. The vapor remains confined
in a cloud and is not dispersed and detonates.

 H co f
We  W
Equation 7 from FMDS 1.11x10 6

Where:
We Mass of Equivalent TNT Energy
Yield (Tonnes)
W Mass of vapor in cloud (kg)

ΔHc Heat of combustion of material


(Kcal/kg) for acetic acid

48 ERI/NRC 09-203
Acetic Acid (Liquid)

f Explosive yield (efficiency)


factor (section 3.4.3 of FMDS)

Tonnes kg
Then We = 2.156 2155.8

From Regulatory Guide


1.91 R RG  17.85W 1/3

Where:

Radial Distance (m) where


pressure wave is less than 7
RRG KPa (1 psi)
W Charge mass, (kg)

ERI Calc.
ERI Calc. (m) (ft) STP Calc. (ft)
Then RRG = 230.0 754.4 814

49 ERI/NRC 09-203
This page intentionally left blank

50 ERI/NRC 09-203

Das könnte Ihnen auch gefallen