You are on page 1of 18

This article was downloaded by: [York University Libraries]

On: 18 October 2013, At: 21:08


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Chemical Engineering Communications


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/gcec20

THE FEASIBILITY OF CONTROLLING UNCONFINED


RELEASES OF TOXIC GASES BY LIQUID SPRAYING
a
VASILIS M. FTHENAKIS
a
Biomedical & Environmental Assessment Division Department of Applied Science ,
Brookhaven National Laboratory , Upton, New York, 11973
Published online: 28 Nov 2010.

To cite this article: VASILIS M. FTHENAKIS (1989) THE FEASIBILITY OF CONTROLLING UNCONFINED RELEASES OF TOXIC GASES
BY LIQUID SPRAYING, Chemical Engineering Communications, 83:1, 173-189, DOI: 10.1080/00986448908940661

To link to this article: http://dx.doi.org/10.1080/00986448908940661

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Chern. Eng. Comm. 1989, Vol. 83, pp. 173-189
Reprints available directly from the publisher.
Photocopying permitted by license only.
© 1989 Gordon and Breach Science Publishers S.A.
Printed in the United States of America

THE FEASIBILITY OF CONTROLLING


UNCONFINED RELEASES OF TOXIC GASES BY
LIQUID SPRAYING
VASILIS M. FTHENAKlS
Biomedical & Environmental Assessment Division
Downloaded by [York University Libraries] at 21:08 18 October 2013

Department of Applied Science


Brookhaven National Laboratory
Upton, New York 11973
(Received November 18, 1988)

The technical and theoretical feasibility of liquid spraying as a means to control accidental unconfined
releases of toxic gases is investigated. Experimental and field information on water spraying of
hazardous gases is analyzed, and the effectiveness of i) dilution, and ii) absorption is assessed. In
addition, a mass-transfer model is developed and used to determine theoretically the effectiveness of
liquid spraying in absorbing and inactivating unconfined releases of toxic gases. It is shown that the
dilution achieved by water spraying is insufficient for toxic gases which pose health hazards at very low
concentrations, whereas absorption and chemical reaction of the gas in the liquid phase can be a very
effective control.
KEYWORDS Toxic Gases Liquid Spraying

1. INTRODUCTION

Accidental releases of toxic, corrosive or flammable gases present serious hazards


to public health and safety as recent accidents in the chemical and process
industries have shown. Highly toxic gases, especially, are potentially harmful if
they are accidentally released in the environment, even at relatively small
quantities (Fthenakis et al., 1987; Moskowitz et al., 1987). Although preventing
accidental releases through the choice of inherently safer technologies, processes,
systems, and procedures is of the utmost importance, in spite of all precautions,
releases of vapors/gases may occur. Therefore, techniques to mitigate these
hazards should be in place. Releases into confined spaces may be controlled with
conventional pollution-control systems (e.g., chemical scrubbing, thermal in-
cineration) if they can be contained and diverted into control systems designed
for massive and transient releases (Fthenakis et al., 1988). Releases in open or
partially confined spaces, however, are very difficult to control; techniques for
amelioration in most cases are only partially effective. Such techniques include:
inactivating the gas by burning it, entraining it by secondary confinement or foam

Prepared for publication to the Chemical Engineering Communications.

(173)/459
460/[174] V.M. FfHENAKIS
spraying; dispersing it by blowing, heating, or water spraying; and entraining and
inactivating it by chemically reactive foams or liquid jets.
Burning may be applicable only to certain flammable gases. Secondary
confinement has the disadvantages of high cost, limitation to stationary systems,
the potential of flammable gases to form explosive mixtures inside the non-air-
tight secondary containment, and susceptibility to failure caused by the same
events that caused the initial leakage (e.g., earthquake, fire). Foam spraying
efficiently suppresses the vaporization of a liquid spill by covering it. Reactive
foam spraying can be effective if the contaminated air is sucked inside the
foam-generating machine (Godfrey, 1988); this is possible if a release occurs
inside a ventilated building, but it is difficult to do with an unconfined release
Downloaded by [York University Libraries] at 21:08 18 October 2013

which quickly spreads out.


Both steam jets and water sprays can "push" air within a gas to dilute and
disperse it. A water spray, however, consumes much less energy and has the
advantage of the ready availability of water. Dispersion of a gas also can be
achieved by big fans; but this option may not be as practical as dispersion by
water jet due to the higher cost required.
The dispersion/dilution capability of water sprays was studied extensively by
the British Health and Safety Executive (McQuaid, 1976; Moodie, 1985) and
others (Heskestand et al., 1976; Watts, 1976). These studies indicate that water
spraying enhances significantly the rate of dilution and dispersion of heavy gas
releases. However, only a limited effectiveness has been demonstrated in real
emergencies. The aim in most cases was to enhance the dispersion of a "heavy"
gas by the momentum of the water jet and entrained air, rather than trap and
absorb the gas within the liquid. In specific cases where water was used to absorb
the highly water soluble ammonia vapors, this technique was highly effective. In
this study the feasibility of liquid spraying applied i) in a "dilution" pattern, and
ii) in an "absorbing" pattern, is investigated.

2. THE EFFECTIVENESS OF WATER SPRAYING

2.1 Dilution
Water sprays have been used mainly as a means of diluting a released gas by
inducing air flow; either fixed water spray installations or mobile systems (e.g.,
fire trucks, fire boats) have been used. The effectiveness of a water spray in
diluting the gas can be determined from the "air entrainment ratio", expressed as
the volume of entrained air per unit volume of water sprayed (cubic meters of air
per liter of water).
There have been several experimental investigations of the use of water sprays
against hazardous gas releases. Eggleston et al. (1976) reports air entrainment
ratios in the range of 2.5-6 m3/l for nozzle pressures in the range of 20-90 psig.
McQuaid (1976; 1983) analyzed series of laboratory data and reported that the
rate of air entrainment is related to basic parameters of the spray nozzle (i.e., the
water's flow rate and pressure at the nozzle, the tube's diameter, and water
CONTROLLING TOXIC GASES [17511461
density). According to his studies, air entrainment ratios may vary from 0.1 m 3 /l
to a maximum of 15 m 3 / 1; the maximum values were obtained for wide pattern
sprays at relatively low flow rate, high water pressure, and small tube diameter.
Hekestad et al. (1976) investigated both experimentally and theoretically the air
entrainment from solid-cone sprays that faced downwards. Their study showed
that for a fixed wide-pattern nozzle there is an increase of entrainment ratios with
distance, and droplet diameter. The air entrainment ratio increased from 6 m3 II
to 15 rrr'II as distance from the nozzle increased from 1.5 m to 5.8 m. The
variation with droplet size was slight, being only a 30% reduction with a 300%
increase in size.
In a series of full-scale tests Moodie (1985) studied the effect of meteorological
Downloaded by [York University Libraries] at 21:08 18 October 2013

conditions and different patterns of water sprays. His study showed that the
effectiveness of dilution is higher in slow winds, and that upward-pointing sprays
perform better than sprays pointing downward. In these field-tests, dilution ratios
(ratios of gas concentrations with and without the water sprays operating) were in
the range of 2 to 5 measured 10 to 20 m downstream of the spray. The author
suggested that "water sprays can achieve a worthwhile enhancement of the rate of
dispersion and dilution of heavy gas spills." This is certainly the case close to the
source of spraying. The questions arise, however, whether this dilution effect
lasts, and if it is sufficient to protect population living downwind from the release.
For this analysis the maximum air entrainment ratio of 15 m3 I I will be used to
provide an upper bound of dilution effectiveness. Under the assumption that only
ambient air free of gas is entrained by the spray and it "perfectly mixes" with the
gas, a max:imum water spray dilution ratio (DR) is estimated from the gas mass
balance equation as:
1';. (m/MW) + 1';nA
DR=-= (1)
You. (m/MW)
where m is the mass release rate of the toxic gas, MW its molecular weight, 1';.
and You, the molar fractions of the toxic gas before and after dilution, and A the
entrained air molar flow rate.
Figure 1 shows the variation of dilution ratios with gas concentrations for gases
of different molecular weight. Dilution ratios increase linearly with increasing
water flow rate, as shown in Figure 2. Higher water flows can be attained by a
series of water nozzles and nozzle stages. A standard fire-engine, for example,
can deliver through several sprays a total flow rate of about 60 I/s. Spray nozzles
which deliver about 2 lis of water flow can be spaced about 1 m apart from
complete coverage.
To assess, the effect of dilution further away from the point of application of the
spray, the atmospheric dispersion of an instantaneous release of 1 kg AsH3 was
modeled, with and without dilution, using the WHAZAN dense gas model
(Technica, 1986).t It was estimated that in calm and very slow wind (1 m/s)
conditions, the AsH 3 cloud 5 m downwind of the source is about 9 m wide with a
centerline concentration of 7.7 v%; in neutral and moderate wind (5 m/s)

t Technica Inc., Columbus OH.


462/[176] Y.M. FfHENAKIS
DILUTION RATIO (Yln/Youl)
6°r-;::====;-'"'-'-----------i
- C02: MW·44
50 ..... AsH3, MW o78

40
Downloaded by [York University Libraries] at 21:08 18 October 2013

3 e 5 6 7 B 9 10
CONCENTRATION IN (v'll)

FIGURE 1 Maximum' dilution produced by a water spray for gases of different molecular we1ht;
gas flow rate = 1 kg/s, water flow rate = 10 1Is. (. maximum refers to entrainment ratio of 15 m air
per liter of water.)

conditions, the width is about 4 m and the peak concentration is about 13 v%.
Spraying from a plane perpendicular to the wind direction with 21/s per meter
give dilution ratios of 70 and 55 for calm and neutral conditions. This dilution
reduces the concentration of the gas locally (Figure 3), but this effect disappears
farther downwind (Figure 4), so that the critical downwind distance correspond-
ing to the "immediately dangerous to life or health" (IDLH) concentration of
6 ppm, is about the same with and without spraying.
This result is not surprising since dilution does not alter the amount of mass
being transported downwind. The dilution provided by spraying, however, alters
the dispersion of a "dense" gas, by providing an early transition to neutrally
buoyant gas dispersion. This effect enhances the vertical diffusion while reduces
the horizontal dispersion resulting in decreased residence times and less horizon-
tal spreading of the cloud. Thus the total times of exposure and the areas covered

12° ;::.1~L=UT=I=O=N=R=AT='o=(y=in='=yo=u=t)::::;_----------1
o
1
- WalerfGas Rallo' I)

100 - weIer/Gas aeuc • 10


.••.. WeIer 108S Rollo· 20

80

60

40

20

o ~:::E:::::::::::=-~~~_~~-----,-_-,--------l
o 2 3 4 5 6 7 8 9 10
CONCENTRATION IN (v'l.)

FIGURE 2 Maximum dilution produced by spraying water on AsH) release; 1 kgls gas release;
15m3 / 1entrainment ratio.
CONTROLLING TOXIC GASES [177)1463
CONCENTRATION (v %)
14[-;=======;---------1
- F, , mlS,no dilution
12 ._ •. F, 1 m/a.ouuucn-rc
- D, 5 m/s.no dilution
10 •••. 0.5 m/s.dllutlon-55
Downloaded by [York University Libraries] at 21:08 18 October 2013

o L::::~'.:±.'~~~~_""",,",,":-_.J
5 10 15 20 25 30 35 40 45 50
DISTANCE (m)

FIGURE 3 Local dispersion of 1 kg AsH, with and without dilution; ground level concentrations
versus downwind distance.

by the toxic gas cloud can be reduced by dilution, but these effects alone do not
provide an effective means of control.

2.2 Absorption and Chemical Reaction


Absorption in water can permanently remove a toxic gas from the cloud and
provide a means of local and far-field protection, as was recently demonstrated in
field tests (If hydrogen fluoride (HF) and ammonia (NH 3 ) releases.
Open-field and laboratory experiments on HF spraying with water were
described by Blewid et al., (1987). In the open-field tests HF was released from a
liquid source and controlled by water spraying at approximately 120 ft from the
source. Wind speeds of 5.4 to 6.8 m/s and Pasquill atmospheric stability class Cor
D were prevailing, and the plume width was 50 to 60 ft at this point. The spray

CONCENTRATION (ppm)
lOO°ir========,-----------l
- F, , m/e.nc dilution
- - F, 1 m/S,dllullono70
800 - D. 5 mlS,no dilution
. - . D, 5 m/s,dllullon-Sfi

600

IDLH
fI ppm 0 L_~::::::3"'":::-,...;;;;;:;;;;;;:;==:====~=~=d
50 100 150 200 250 300 :350 400 450 500 550
DISTANCE (m)

FIGURE 4 Dispersion of 1 kg AsH, with and without dilution; ground level concentrations versus
downwind distance.
464/[178] V.M. FfHENAKIS

TABLE I

Water spraying of HF spills (from Blewitt, Hague, 'I al., 1987)

WaterlHF mass ratio Effectiveness (%)

7 8.7
10 37.8
19 44.6
22 54.7
43 69.7
64 78.8
Downloaded by [York University Libraries] at 21:08 18 October 2013

Other parameters (personal communication, W.J. Hague, 1988).


Single nozzle: 90--110' pattern.
Nozzle Dow rate: 0.2-1.5 gpm.
Droplet average size: (250-300) 10- 6 m.
Chamber height: 0.61 m.
Ar~a of vaporization: 0.28 m •
2

system comprised two nozzle spray headers, of 25 nozzles each, placed 5 ft apart.
The first header was directed upwards from ground level, and the second was
directed downwards from a height of about 12 ft. Fog-type nozzles were used with
a mean drop size of 250 microns delivering a total water flow rate of
approximately 700 gallons per minute (a ratio of 21/1 water to HF mass). An
effectiveness of 36 to 49% was determined from atmospheric and deposition
sampling at 300, 1000 and 3000 m. At these distances the dilution effect had been
diminished and therefore the measurements showed the actual effectiveness of
HF removal.
For the laboratory experiments, a flow-through chamber with air velocity of
approximately 3 mls was used. The effectiveness of HF removal was measured for
several water/HF ratios, by analysis of collected spent spray liquor. These data
are listed in Table I.
A very informative paper on NH 3 control with water sprays by Greiner (1984)
gives a detailed procedure for water spraying developed for firefighters respond-
ing to ammonia vapor releases. In this procedure, which is based on field tests

TABLE lIa

Data on water spraying of NH 3 releases'

Release conditions Two-phase


Ambient temperatue ('C) 5-32
Storage vapor pressure (psi) 80-115
Discharge pipe diameter (inch) 0.75 & 1
Total release (kg) 16--22.7
Duration of release (s) 30-40
Water Dow rate (gpm) lOOper nozzle; 6 nozzles
Duration of spraying (s) 40-50
Wind speed (m/s) . 2.2-5 (ideal conditions)
Location of nozzles 30-45 m from source of leak
• From personal communication with M.L. Greiner, 1988.
CONTROLLING TOXIC GASES [179]/465
TABLE lIb

Calculated effectiveness of water spraying on NH 3 releases'

Estirnatedt" peak concentration


without spraying at 53 m (ppm) 4000
Assumed concentration behind
sprays (53 m from source) 200 (discomfort level)
Estimated" effectiveness of
NH 3 removal (%) 95

• Under the conditions described in Table IIa .


•• By using the WHAZAN dense gas model to simulate the atmospheric dispersion of the
instantaneous release of 22.7 kg of NH 3 •
Downloaded by [York University Libraries] at 21:08 18 October 2013

and real accidents, the gas is sprayed from downwind and the sides. A "capture"
zone is created by having four water hoses with fog nozzles positioned downwind
in a semi-circular pattern and allowing ammonia vapor to drift into this pocket.
For normall wind conditions (5 to 15 mph) a good working distance was 125 to
150ft from the point of release; shorter distances (about 75 ft) were needed for
releases in stronger or very slow winds. This positioning, as well as a wide and
rapid rotation of the nozzles, were found essential for effective control. In the
most successful tests the concentration of NH 3 behind the firefighters was reduced
to a level that caused no general discomfort (probably about 200 ppm). For these
cases an effectiveness of 95% was estimated by using the WHAZAN dense gas
dispersion model. The field data are presented in Table lIa and our calculations
in Table lIb.

3. THEORETICAL EFFECTIVENESS
The above studies indicate that water spraying can effectively control unconfined
releases of hydrogen fluoride and ammonia, which are highly water soluble gases.
There is no quantitative information on the effectiveness of this control option
with other gases. Prugh (1987) has qualitatively assessed the effectiveness of
water curtains on several gases, based on their values of Henry law constant.
Prugh (1986) has also presented mass balance equations to calculate the
theoretical effectiveness of water curtains from water and gas flow rates, initial
gas concentration and Henry's law constant. Although useful for an estimate of a
maximum effectiveness, these equations do not describe the effect of several
parameters (e.g., droplet diameter, spray pattern, gas and liquid mass-transfer
coefficients) which are important in the design of an efficient system.
There is substantial literature, however, on modeling momentum and mass
transfer processes within drops or sprays which is applicable to this study. This
literature was developed mostly for studying rain scavenging of atmospheric
pollutants (mass-transfer into single drops: Hales, 1972; Walcek and Pruppacher,
1984; Schwartz, 1986), spray dynamics (Schlichting, 1979; Yeung, 1982), momen-
tum and mass-transfer into drops and streams of drops (Ramachandran, 1985;
Lindhjem, 1987), and momentum, mass and heat transfer in spray drying systems
466/[180J V.M. FfHENAKIS
(Crowe, 1980). For the purpose of assessing the feasibility of liquid spraying, a
simple mass-transfer model suffices. Existing models are not readily applicable
and they require involved numerical solutions. Therefore, an analytical model,
which uses relationships from the rain scavenging literature, is developed from
basic principles.

3.1 Basic Considerations


Water solubilities of several heavier-than-air toxic gases are listed in Table Ill.
Both solubility and Henry's law constant describe the equilibrium gas con-
centration in a well mixed drop. These type of data can be used only if the
Downloaded by [York University Libraries] at 21:08 18 October 2013

equilibrium concentration is established inside the drop during the time it travels
within the gas cloud. Therefore, time and space scales are important in
determining effectiveness of control in relation to rates of mass-transfer and
chemical reaction.
To determine if phase and chemical equilibrium can be attained, we consider
the characteristic times for the mass-transfer SUb-processes which take place in gas
absorption by a falling water drop (Table IV). These times are known reasonably
well from experimental data and steady-state solutions of the mass-transfer
equations.
Table IV shows that for all processes except liquid-phase diffusion, the
mass-transfer into a falling drop is very rapid compared to the fall time of the
drop; this allows the assumption that the equilibrium concentration of the gas
within each drop is reached during the time interval the drop is in contact with
the gas. This assumption can be made for sprays that provide high turbulence and
mixing, where the concentration of the entrained gas does not change significantly

TABLE.II1

Solubilities in water of several toxic gases

Solubility' Henry's law constant" IDLW Density


Gas (g/l00 em") (atm/molar fraction) (ppm) (air = 1)

HF 0.002 20 Id
HCI 82.3 0.015 100 Id
NH, 89.9 0.75 500 Id
S02 22.8 13 100 2.4
CH,NCO 6.7 30 20 2.0
H 2Se 0.94 480 2 3.7
H 2S 0.65 540 300 1.2
CO 2 0.35 1400 50000 2.9
AsH, 0.07 7200 6 2.7

a in water at ere.
• soluble in all proportions.
b mean value at 1 v% and at the IDLH concentration (Prugh, 1987).
c IDLH, the concentration immediately dangerous to life or health, represents a maximum level at
which escape could be made within 30 min without any escape-impairing symptoms or any irreversible
health effects.
d when released in air these gases may form heavier-than-air clouds due to aerosol formation (NH,),
or polymerization (HF).
CONTROLLING TOXIC GASES [181]/467
TABLE IV

Characteristic times pertinent to mass transfer of gases into aqueous droplets

Characteristic Equivalent timet


Process time Reference (s)

Gas-phase diffusion d 2/12Dg Schwartz, 1986 0.0003


Aqueous-phase diffusion d 2/41f2D, Schwartz, 1986 6
Convective mixing with
internal circulation 25d/v, Slinn, 1984 0.0064
1st order reaction 2c/vd Schwartz, 1986
Fast ionization reaction Slinn,1984 <0.001
Drop free-fall hl»;
Downloaded by [York University Libraries] at 21:08 18 October 2013

from h =0.6m 0.5


from h = 1.5 m 1.25

t Assumed values: Dg = 0.23 em 2/s; DI = 1.5E-05 cm 2/s; d = 0.05 em; u, = (4OOOd)/s.

with height. Liquid-phase diffusion is much slower than the expected fall times;
however, for the considered drop sizes (e.g., >0.01 em), convective mixing,
rather than diffusion is the mass-transfer mechanism in the liquid phase.
If the dissolved gas reacts chemically, further considerations must be made. In
general, chemical reactions in the liquid-phase increase both the absorption rate
and the gas-holding capacity. Each gas-liquid system must be considered
individually since reactions and reaction rates differ for each system. Due to the
complexity of the problem only some general inferences are made here. For
chemical reactions which are sufficiently fast so that chemical equilibrium occurs
within a short time, the problem is reduced to defining the physical and chemical
equilibrium of the system. The ionization reactions of both NH 3 and HF provide
examples of very fast reactions:

N~OH ~ NHt + OW
HF~H++F-

which have forward and reverse reaction rate constants in the order of 10+5 S-1
and 1011 (1 mol" S-I) correspondingly (Slinn, 1984).
The rate of NH; formation is

d[NHt] = k (NH OH) (2)


dt f •

At equilibrium the forward and reverse reaction rates are equal, thus:
kf[NH.OH) = k,[NH;][OW] (3)
where [NHt]eq = [OH-]eq = 10- 7 (from the dissociation of H 20 ) . The time it
takes for equilibrium to be attained is estimated from Eq. (2) to be 10-' s.
Similar considerations allow one to determine if other types of reactions are
sufficiently fast for equilibrium to be attained during the short times of interest.
Extremely fast reactions can be either pseudo-first order or second order
reactions with a sufficiently large excess of liquid-phase reagent. If chemical
468/[182] Y.M. FfHENAKIS
equilibrium can be assumed, then the problem of quantifying the effect of
chemical reaction on the overall mass-transfer rate becomes easier, since a
considerable amount of work has been done on the effect of equilibrium reactions
on mass transfer (Chang and Rochelle, 1982; Perry and Green, 1984). These
studies give formulas to calculate mass-transfer enhancement factors for several
types of reactions, that can be used to adjust the mass-transfer and the
equilibrium coefficients in the model presented below.

3.2 A Model of Absorption and Chemical Reaction


Gas absorption into liquid droplets involves transport of the gas (solute)
Downloaded by [York University Libraries] at 21:08 18 October 2013

molecules: i) through the surrounding air to the surface of the droplet; ii) through
the gas-liquid interface; and iii) into the interior of the drop, and possible
chemical transformations.
These mass transfer sub-processes are coupled and for dilute systems can be
represented by the equations
N = Kg(Y - YO) = K,(X O- X)
A (4a)
or
(4b)
where NA is the average molar flux of solute A passing through the gas-liquid
interface, and Y and X denote molar fractions of A in the gas and liquid phases;
the subscript i refers to interface conditions, and the superscript 0 to phase
equilibrium conditions.
Kg can be related to kg and k, by the equation
II Kg = l/kg + mlk, (5)
where m denotes the slope of the equilibrium curve in the region between X and
Xi'
m = (yo - Y;)/(X - Xi)

3.2.1 Mass Transfer Balances Consider a downpointing nozzle, spraying on a


dispersed gas. The gas stream at any point in the spray contains G total moles/(s)
(em" of cross section), consisting of solute A of molar fraction Y, and essentially
insoluble air. Similarly, the liquid stream contains L total moles/(s) (ern"), made
of absorbed A of molar fraction X, and of nonvolatile solvent. The turbulent
action of the spray mixes the gas thoroughly, and therefore the concentration of
the solute A in the gas phase can be assumed uniform in the z direction (Edwards
and Huang, 1977); thus Y= Y(x) only.
The differential mass-transfer equation for a control volume of height h and of
width dx, is:
(6)

where G and L are assumed constant within the control volume. The coordinate z
is in the jet direction, and x is in the radial direction.
CONTROLLING TOXIC GASES [183]1469
The gas molar flow can be correlated with the volumetric mass transfer
coefficient Kga by means of Eq. (4a) as:
x)
-G dix = Kga[Y(x) - Y;:'(x)] (7)

where a represents the interfacial area per unit volume, and Y;:'(x) an average
value of the molar fraction of A in the gas-phase in equilibrium with the liquid.
Thus:

Y;:'(x) = ~ f Y*(x, z) dz (8)


Downloaded by [York University Libraries] at 21:08 18 October 2013

Y*, the equilibrium molar fraction of A in the gas-phase in equilibrium with the
liquid of molar fraction X, is a function of z, since X is a function of z; and:
X i x , z)
Y*(x,z)=Y*(x,O)+
lX(x.O)
mdX(x,z) (9)

but since the liquid is free of A at z = O,


Y*(x. 0) = X(x, 0) =0
Thus Eq. (9) becomes:
Y*(x. z) = mX(x, z) (10)
By integrati.ng Eq. (6) with respect to z we obtain X as:
GdY(x)
- L ~z=X(x,z)-X(x,O)=X(x,z) (11)

and Eq. (8) becomes by means of (10) and (11):

Y* (x) = _ hmG dY(x) (12)


m 2L dx
Then from Eq. (7):
hm G)
dY(G +K a +'K aY=O (13)
dx g 2L g

which has the simple solution


Y(x) = Y(O)e- bx (14)
where

3.2.2 Estimation of mass transfer parameters The gas-phase mass transfer


coefficient, kg, can be determined by the following expression which is valid for
scavenging of gas into free-falling water droplets of diameter ranging from 0.04 to
2.5 mm (Beard and Pruppacher, 1971; Pruppacher and Rasmussen, 1979).
Sh = 1.61 + 0.718 Re"? Sc'" (15)
470/[184] Y.M. FfHENAKIS
where
Sh = kgd Re = vzd Sc = ~
DAc V DA
kg, as obtained from Eq. (15), depends on the concentration of the gas; for
concentrations of about 10v% this equation gives values of kg which result to a
good fit of the experimental data, whereas for low concentrations it leads to
serious underestimates of the data. Since Eq. (15) does not account for the
enhancement of gas-phase mass-transfer due to turbulence, and in order to fit the
experimental data, a value of kg an order of magnitude higher than the one
predicted from Eq. (15), was used.
The liquid-phase mass transfer coefficient, k t , depends on film effects at the
Downloaded by [York University Libraries] at 21:08 18 October 2013

interface, internal circulation, and liquid-phase chemical reactions. Even without


the chemical reaction effects, there is a considerable disagreement between
predictions of theoretical models (Altwicker and Lindhjem, 1988). For two
limiting cases of mass-transfer without chemical reaction, Hales (1972) has
estimated k, for single falling drops, as:
For rapid convection: k, = very large, (resulting to Kg = kg)
For zero convention (diffusion controlled transfer): k, = 1ODAc/d, which
approaches zero for low gas concentrations.
In both the HF and NH 3 tests rapid convection was attained; therefore the
assumption of Kg = kg was adopted in modeling these tests.
The slope of the equilibrium curve, m, is equal to the constant, H, for
gas-liquid systems that obey Henry's law, and approaches zero for systems that
react rapidly and irreversibly.
The interfactial area per unit volume, a, can be calculated by the volumetric
flow rate of the liquid (Q), the drop's surface to volume ratio, the average
exposure time of each drop in the gas (r), and the volume of the system (V):
6 1.4 h
a=Q-dt/V where t=-- (16)
Vz

3.3 Model Results


The model was used to simulate the tests on HF and NH 3 previously discussed.
As seen in Figure 5, the model slightly overestimates the HF laboratory results
for mass ratio 7 and underestimates them for ratios 10 to 43. A possible
explanation for this deviation is that the air temperature and, subsequently, the
vaporization rate, was not constant during these experiments. This would result in
different initial concentrations that may significantly alter the effectiveness
estimated by the model. Unfortunately there are no data available to test this
explanation. For the 21/1 water to HF mass ratio the model prediction is in the
range of the open-field data.
The same model was used to simulate the ammonia field tests. Spraying from
the sides rather than from the top was crudely accounted for by adjusting the
drop-gas contact time and interfacial area to include both the horizontal motion
and vertical fall of a drop. The results of the simulation are shown in Figure 6;
these results agree with the estimated field effectiveness for a value of m between
o and the value of Henry's law constant, 0.75.
CONTROLLING TOXIC GASES [185]/471

10°1
~E-;:F=FE=C=T='=VE=N:::E=S=S=(%=J=:::;- l
- Model aeauue
.c. Lebor atcr v Tests
80
"* Open-Field Tests

80

40

20
Downloaded by [York University Libraries] at 21:08 18 October 2013

oL-_~--~-~--~ _ _~_ _-'-_---J


o 10 20 30 40 00 60 70
Water /HF Mass Ratio
FIGURE 5 Effectiveness of water spraying on HF releases. Model parameters: ,
K =
0.002 moles/cm 2/s, d = 0.03 em, v = 150 cml«, Y(O) = 0.003.

- m-H'075
- moO
'* Field las I

oL-_ _ ~ __ ~_~ ~ __ ~ _ _---J


o as 1 1.5 2 2.5 3
CONCENTRATION (v%)
FIGURE 6 Effectiveness of water spraying on NH 3 releases. Model parameters: K, =
0.002 moles/cm 2/s, d = 0.03 em, v = 150cm/s, water/gas mass ratio = 69.

100 ~E._FF_E:..C:..T_'V"E;.N-,E_S_S_(,-%.:.I ...,

80

60

40

20

o fL:==-r::...-~-,.-~~F-"''-r-~O;='-'--T==;==
o 2 3 4 5 a 8 g 10
CONCENTRATIOt-1 (v%l

FIGURE 7 The effect of gas solubility on the effectiveness of water spraying. Model parameters:
K, = 0.OO2moles/cm2/s, d = 0.03 em, v = 150cm/s, water/gas mass ratio = 43.
472/[186] Y.M. FTHENAKIS

~-===:::====::::-=="l
100 r-EF_F.:.EC.:..T::'V:::E::N::ES::S:::I:;%;,J
1

80

80
... "
Downloaded by [York University Libraries] at 21:08 18 October 2013

oL--~ __ ~ __ ~ __ ~ __ ~ _ _---.J
o 234 5 8
CONCENTRATION {v~l
FIGURE 8 The effect of drop size on the effectiveness of water spraying. Model parameters:
K. = 0.002 moles/em 2/s, m = 0.002, v = 150em/s, water/gas mass ratio = 43.

Although the model oversimplifies the complex drop-gas interactions, it allows


us to draw some tentative conclusions on the influence of several important
parameters on spray effectiveness. The effect of solubility is shown in Figure 7.
For gases which are only moderately soluble in water (e.g., SOz), spraying is only
marginally effective, whereas for very low solubility gases (e.g., HzS) water
spraying is ineffective. The effect of the variation of water flow rate has been
shown in the HF tests (Figure 5). Different nozzle characteristics can be described
by different median drop size, initial drop velocity, and spray pattern. The effect
of variation in drop diameter is shown in Figure 8; smaller drops result to higher
effectiveness, due to increased times of interaction and increased interfacial areas.
The model, however, does not describe the increased internal circulation that
larger drops experience. This effect increases the liquid mass-transfer coefficient
and may counterbalance the better mixing that smaller drops provide.

4. DISCUSSION
Both the experimental work reviewed and our theoretical calculations show that
water spraying can be very effective for gases which are highly soluble and react
in the liquid phase (e.g., NH 3 and HF). This control option is not restricted to
water soluble gases. A number of highly toxic gases which are only slightly
absorbed by water (e.g., HzS, HzSe, SiF.) react very fast with dilute alkaline
solutions, while other gases (e.g., AsH 3 • PH 3 , BzH6) react fast with oxidizers
(e.g. NaOCI). The liquid-phase chemical reactions of these gases result in an
increased solubility and rate of absorption. The same considerations that were
discussed earlier about absorption and chemical reaction in pure water apply to
aqueous solutions. Our model can be used for these gas-liquid systems if the
chemical reactions can be established, and their effect on mass transfer can be
quantified using mass-transfer enhancement factors. Work in this area is in
progress and, as a next step, the model will be developed to include these effects.
CONTROLLING TOXIC GASES [187]/473
The use of liquid spraying may raise some concerns about the contamination of
the water supply and protection of personnel, especially if chemical solutions are
used. Spraying the liquid in the storage yard of a plant should not present any
problems of water contamination provided that the area is paved and drained into
a liquid collection basin where the liquid can be further treated. Personnel
dealing with a toxic gas emergency with liquid spraying will have hazardous
environmental protection suits which totally cover them. Materials such as
rubber, PVC, and neoprene protect the body from the irritating and caustic
effects of both the gas and the liquid solution. Positive pressure, self-contained
breathing apparatus will provide respiratory, eye, and face protection from toxic
gas or vapors (Greiner, 1984).
Downloaded by [York University Libraries] at 21:08 18 October 2013

In most eases, however, action by personnel may not be feasible due to the
short response times required. Atmospheric dispersion analysis showed that even
a heavier-than-air gas when released from a pressurized cylinder it will disperse
quickly to significant distances. Therefore stationary control systems that are
automatically triggered by an accident must be used. Potential use of portable
control equipment is probably limited to a few specific cases such as slowly
vaporizing liquid spills, partially confined gas releases, and releases of liquefied
gas in cold weather.

5. CONCLUSIONS

A critical review of the water spraying literature along with a study of


atmospheric dispersion, showed that dilution of a dispersed gas by water spraying
is completely ineffective in reducing public health hazards from releases of highly
toxic gases. Dilution by water spraying may be an effective control measure only
for releases of flammable gases where a local drastic decrease of the concentra-
tion could prevent ignition.
On the other hand, absorption of a gas by liquid spraying accompanied by
chemical reaction in the liquid phase, can be a very effective control option for
unconfined releases of toxic gases. The effectiveness of this option on releases of
ammonia and hydrogen fluoride have been demonstrated by field-experiments.
Analysis of the field results and theoretical considerations identified several
influential parameters. High liquid flow-rate and gas solubility, small drop
diameter, high mass-transfer coefficients, and chemical reactions, all are factors
which enhance gas absorption by a spray. Characteristic times of mass-transfer
and chemical reaction can be sufficiently fast so that simplified assumptions of
phase and chemical equilibrium are applicable.

ACKNOWLEDGEMENTS

I am grateful to Professor Victor Zakkay for his encouragement and advice. I also
wish to thank D.N. Blewitt, M.L. Greiner, W.J. Hague, W.R. Prugh, and S.E.
Schwartz for the information so kindly they provided. This work has been
474/[188] V.M. FfHENAKIS
partially supported from the Photovoltaic Energy Technology Division, Conser-
vation and Renewable Energy, U.S. Department of Energy, and I thank P.D.
Moskowitz, A. Bulawka and L.D. Hamilton for this support.

NOMENCLATURE

a interfacial area per unit area, cm- I


c "bulk" gas concentration, moles em"?
d drop median diameter, cm
diffusivity, cm2 S-I
Downloaded by [York University Libraries] at 21:08 18 October 2013

D
h height, em
H Henry's law constant, dimensionless
Henry's law constant for water vapor, 0.023
forward reaction rate constant
gas-phase mass-transfer coefficient, moles ern"? s-t
liquid-phase mass-transfer coefficient, moles em"? S-1
overall mass-transfer coefficient based on the gas-phase driving force,
moles em"? S-1
k, reverse reaction rate constant
m slope of equilibrium curve, dimensionless
Re Reynolds number, dimensionless
Sc Schmidt number, dimensionless
Sh Sherwood number, dimensionless
gas-drop contact time, s
v drop velocity, cm s-t
x denotes horizontal direction
z denotes vertical direction
X molar fraction of component A in the liquid-phase
Y molar fraction of component A in the gas-phase
Y' molar fraction of component A in the gas-phase in equilibrium with A in
the liquid-phase
Y;:' average value of Y' in the vertical control volume
v kinematic viscosity, em" s- t

Subscripts
A denotes gaseous component A
denotes interface
g denotes gas-phase
I denotes liquid-phase
CONTROLLING TOXIC GASES [189]/475
REFERENCE.S

Altwicker, E.R., and Lindhjem, C.E., Absorption of gases into drops, AIChE Journal, 34(2),
329-332 (1988).
Beard, K.V., and Pruppacher, H.R., A wind tunnel investigation of the rate of evaporation of water
drops falling at terminal velocity in air, J. Almas. Sc. 28, 1455-1464 (1971).
Blewitt, D.N., Yohn, I.F., Koopman, R.P., Brown, T.e., and Hague, W.I., Effectiveness of water
sprays on mitigating anhydrous hydrofluoric acid releases, Proceedings of the International
Conference on Vapor Cloud Modeling, pp. 155-171, Nov 2-4, 1987, published by the
American Institute of Chemical Engineers, (ed. 1. Woodward).
Chang-Shih, c., and Rochelle, T., Mass transfer enhanced by equilibrium reactions, Ind. Eng. Chern.
Fundam., 21, 379-385 (1982).
Crowe, C.T., Numerical Models for Gas-Particle Flows, Polyphase Flow and Transport Technology,
ASME Symposium, 7-9 (1980).
Downloaded by [York University Libraries] at 21:08 18 October 2013

Edwards, W.M., and Huang, 1'., The Kellogg-Weir air quality control system, Chem. Eng. Progress,
73(8), 64-65 (1987).
Eggleston, L.A. et 01., Water spray to reduce vapor cloud, AIChE Loss Prevention, 10,31 (1976).
Fthenakis, V.M., Moskowitz, P.O., and Hamilton, L.D., Personal safety in thin film photovoltaic cell
industries, Solar Cells, 19,269-281 (1986-1987).
Fthenakis, V.:M., Moskowitz, P.O., and Sproull, R.D., Control of accidental releases of hydrogen
selenide and hydrogen sulphide in the manufacture of photovoltaic cells: a feasibility study, J.
Loss Prevention, 1(4),206--212 (1988).
Greiner, M.L.; Emergency Response Procedures for Anhydrous Ammonia Vapor Releases,
Plant/Operations Progress, 3(2),66-71 (1984).
Godfrey, K.A., Foam over troubled toxics, Civil Engineering, (Feb. 1988).
Hales, I.M., Fundamentals of the theory of gas scavenging by rain, Almos. Enuir., 6, 635-659,
(1972).
Heskestad, G., Kunga, H.C., and Todtenkopf, N.F., Air entrainment into water sprays and spray
curtains, ASME Winter Annual Meeting, Paper 76-WA/FE-40, 1976.
Lindhjem, C.E., Absorption of gases by drops and streams of drops, Ph.D. Thesis, Rensselaer
Polytech. 1nst., Troy, NY (1987).
McQuaid, I., The design of water-spray ventilators, J. of Occup. Accidents, 1,9-19 (1976).
McQuaid, I., The scope for reduction of the hazard of flammable or toxic gas plumes, J. of Occup.
Accidents, 5, 135-141 (1983).
Moodie, K., The use of water spray barriers to disperse spills of heavy gases, PlantlOperations
Progress, 4(4),234-241 (1985).
Moskowitz, P.O., Fthenakis, V.M., Hamilton, L.D., and Lee, J.C., Public health issues in
photovoltaic energy systems: An overview of concerns, Solar Cells, 19,287-299, (1986-1987).
Perry, R.H., and Green, D.W., Perry's Chemical Engineers' Handbook, 6th Ed., (1984).
Prugh, W.R.. Mitigation of vapor cloud hazards: Part II. Limiting the quantity released and
countermeasures for releases, Plant/Operations Progress, 5(3), 169-174 (1986).
Prugh, W.R., Hazard Reduction Engineering, Inc., private communication, 1987.
Pruppacher, H. R., and Rasmussen, R., A wind investigation of the rate of evaporation of large water
drops falling at terminal velocity in air, J. Almas. Sci., 36, 1255-1260 (1979).
Ramachandran, S., Mathematical modeling and computer simulation of mass transfer in simple
multiple drop systems, Ph.D. Thesis, Rensselaer Polytech. Inst., Troy, NY (1985).
Schlichting, H., Boundary Layer Theor, 7th Ed., McGraw Hill, New York, 1979.
Slinn, W.G.N., Precipitation scavenging, in Atmospheric Science and Power Production, DOE/TIC·
27601, 1984.
Schwartz, S.E., Mass-transport considerations pertinent to aqueous phase reactions of gases in
liquid-water clouds, NATO ASI Series, Vol. G6, Chemistry of Multiphase Atmospheric
Systems, edited by W. Jaeschke, Springer-Verlag Berlin, 1986.
Walcek, C.I., and Pruppacher, H.R. On the scavenging of S02 by clouds and raindrops, J. Atmos.
Chemistry, 1,269-289 (1984).
Watts Jr., I.W., Effects of water spray on unconfined flammable gas, AIChE Loss Prevention, 10,
48-54 (1976).
Yeung, W.-S., Similarity analysis of gas-liquid spray systems, Journal of Applied Mechanics, 49,
687-690 (1982).

You might also like