Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

HAZARDOUS WASTE

Volume 1, Number 1, 1984


Mary Ann Liebert, Inc., Publisher
Pp. 41-65

Fluid BedCatalytic Oxidation: An


Underdeveloped Hazardous Waste
Disposal Technology
MICHAEL P. MANNING

Nahant Engineering, Nahant, MA 01908


and
Department of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139

ABSTRACT

Fluid bed incinerators have been used to dispose of a wide


variety of hazardous wastes including refinery waste streams, sur-
plus explosives, PCB-contaminated transformer oil and oxychlorina-
tion process wastes. Both inert and catalytic solids have been
employed separately and in mixed beds or sequential reactors. How-
ever, in most catalytic processes, the nature of the catalyst and
the process operating conditions are proprietary. In this paper,
the catalytic oxidation kinetics of three chlorinated hydrocarbons
are determined in an externally pumped recycle reactor with a com-

mercially available chromia on alumina catalyst. Conversion of the


chlorocarbon apecies in the reactor effluent are monitored by gas
chromatography for temperatures in the range 350 to 550CC. Regres-
sion of the measured catalytic oxidation rates yields rate expres-
sions which are first order in the chlorocarbon concentration, be-
tween zero and first order in oxygen concentration and exhibit some
inhibition by water. Some catalyst poisoning by halogen containing
species is observed. However, the poisoning phenomena can be mini-
mized by control of the H/Cl ratio in the mixed waste feed stream.

41
42 MANNING

BACKGROUND

Wastes from the manufacture of many chlorinated hydrocarbons


are presentlyclassified as hazardous. As such their disposal must
meet the criteria specified in the regulations promulgated under the
Resource Conservation and Recovery Act. While many wastes can be
incinerated with conventional technology, the incineration of chlor-
inated hydrocarbons presents several unique complications. First,
halogen containing materials are known to inhibit flame propagation,
have lower heats of combustion, and hence are more difficult to
oxidize completely. Secondly, the combustion products of the halo-
gen, mainly HC1 and some Cl_, are both corrosive and toxic and must
be scrubbed from the incinerator flue gas stream before discharge.
Catalytic techniques for oxidation of chlorinated hydrocarbons
while suffering from the same disadvantages appear attractive in
that two benefits are possible. The operating temperature of a
catalytic process could be more than 300°K lower than a comparable
thermal process with similar residence time and conversion. Since
the increased temperature in the thermal process is normally pro-
vided by burning fuel which represents up to 40% of the total oper-
ating cost, substantial cost savings are achievable. The decreased
temperature level also allows utilization of available corrosion
resistant alloys for process vessel construction rather than the
replacable ceramic linings which corrode rapidly in the thermal
furnace enclosures. Moreover, the possibility exists that in situ
halogen removal might be accomplished with a fluidized solid sorbent
such as lime, CaO, limestone, CaCO , or soda ash, Na CO as has been
demonstrated for sulfur and halogen containing wastes.

INTRODUCTION

Both catalytic processes and fluid bed reactor techniques have


been demonstrated separately as effective oxidation and waste dis-
posal technologies. In a few limited cases, the advantages of com-
bining catalytic oxidation with fluid bed reactor performance have
been reported. Hardison and Dowd (1) describe the development of
a pilot scale catalytic unit 0.91 m in diameter with a 15 cm bed
of proprietary catalyst. In the disposal of a refinery waste
stream, the unit oxidized 98% of the feed organic materials of an
operating temperature of 438°C and a catalyst contact time of about
150 milliseconds.
Similarly Becker and Wall (2,3,4) have described the operation
of Dorr-Oliver's full sized fluid bed incinerators. These non-cata-
lytic units are typically from four to twenty one feet in diameter
depending on capacity. As shown in Table 1, six of the units in-
stalled at petroleum refineries are utilized for disposal of waste
streams containing sizable amounts of water. The fluidized solids
are 20 to 80 mesh inert particulates, 1 to 1.8 m deep, such as sand
at startup. In operation, however, the solid composition changes
as waste containing various inorganic salts from the caustic washing
and other refinery processes accumulate in the reactor. The technol-
ogy exists for controlling the particulate compostion so that low
FLUID BED CATALYTIC OXIDATION 43

TABLE 1

Fluid Bed Incineration Systems at Six Refineries

Plant Reactor
_Design_Size
Waste Rate _

Company and Inside


_Plant Location_ 1,000 ton/day wt% water Diameter,m
Mobil Oil,
Worth (Karlsruhe), Germany 40 95 4.0
Gulf Oil,
Bertonico (Lodi), Italy 10 93.5 2.2
Gulf Eastern,
Bantry Bay, Ireland 5 95 1.0
Stanic (Esso) ,

Bari, Italy 38 87 4.3


Stanic (Esso),
Livorno, Italy 84 83 6.0
American Oil,
Whiting, Ind. 338 85-95 8.5

TABLE 2

Hazardous Wastes Tested in Dorr-Oliver Incinerator

Waste A Phenolic Waste B Methyl Methacrylate


- -

Compound Concentration Compound Concentration


Ethanol 0.06 Methanol 4.9
Acetone 0.07 Acetone 0.7
2-Butanone (MEK) 0.05 Méthylène chloride 0.7
Butanol 0.01 2-Butanone 1.1

2-Ethoxyethanol 0.07 Methyl proponoate 0.4


Toluene 0.3 Methyl methacrylate 33.9
Xylene (two isomers) 0.6 2-Ethoxy ethanol 0.4
Isopropyl benzene 0.8 Toluene 1.1
Phenol 3.7 Xylene 0.9
o-Cresol 0.9 2-Ethoxy ethyl acetate 0.4
m-Cresol 1.9 Phenol 12.6
water 86.0 Cresols 3.4
inorganic solids 5.5 inorganic solids 1.7
water 38.0
44 MANNING

metling eutectic mixtures are avoided but generally requires operation at


low chloride contents (3). In the absence of chlorine in the waste feed
the reactors normally operate at 815°C in the fluid bed to ensure com-
plete combustion of organic materials. In the presence of chlorides,
however, eutectic mixtures form between NaCl and other solids, such as
Na„CO , which can melt at temperatures as low as 623°C. When this oc-
curs, the solids become sticky, aggregate into larger clumps and the
bed of solids can no longer be fluidized.
In 1976, Arthur D. Little and TRW conducted a test burn of
two candidate hazardous wastes in a 7.6 m diameter fluid bed manu-
factured by Dorr-Oliver (5). Two waste streams, both containing
fairly large amounts of water were incinerated. Waste stream A,
shown in table 2, is generated from the caustic scrubbing of gaso-
line to remove phenol and hydrogen sulfide. Due to the size of
the petroleum refining industry, this phenol waste stream is esti-
mated to be generated at a rate of over 50 million kg/yr. Waste
stream B, contains residual methyl methacrylate monomer such as
might be generated from the manufacture or application of surface-
coating resins. It is estimated that the methyl methacrylate
waste is generated at a rate of 1 to 10 million kg/yr. Because of
the elevated water contents, these wastes were blended volumetricly
with supplementary number 2 fuel oil in the ratios of 2.3 to 3.0
and 2.0 to 2.6 for wastes A and B respectively. The temperatures
measured in the freeboard space above the fluid bed ranged from
815°C to 900°C while the average fluid bed temperatures were some
50°C to 150°C lower. Obviously, some of the observed waste oxi-
dation is occuring homogeneously in the freeboard space and not
entirely in the fluid bed region. Measured waste destruction
efficiencies of greater than 99.999% were reported for both wastes
during the test burn program.
Ziegler e_t al. (6) avoided the problems associated with eutect-
ic mixtures by employing two fluid beds in series operated at
lower temperatures. In a laboratory scale test, the first reactor
contained a fluidized bed of sodium carbonate which acted as a
sorbent by reacting with the HC1 produced during combustion of
solid poly vinyl chloride wastes. The second fluid bed was filled
with a chromia catalyst to oxidize any organic materials which were
not completely oxidized in the first, non-catalytic, fluid bed.
Both beds were operated at 550°C, well below the NaCl-Na CO eutect-
ic temperature. The process originally conceived by Ziegler et al.
has been further developed by researchers at Rockwell into a pilot
scale process capable of handling 82 kg/hr in two reactors, 0.15 m
and 0.30 m in diameter. In 1981, Rockwell announced that the two
stage fluid bed process had achieved a 99.9999% destruction effic-
iency for PCB1s during a test burn in which a sample of transform-
er coolant (52% by weight PCB and 48% trichlorobenzene) at temper-
atures below 695°C (7).
A similar development effort arrived at combining fluid bed
incineration and catalytic oxidation was reported by Carroll et al.
(8). The development effort was the result of Army research into
safe, environmentally sound methods for disposing of waste munitions
including the explosive TNT and RDX. A single fluid bed 0.15 m in
FLUID BED CATALYTIC OXIDATION 45

diameter and 30.5 m high was packed with a combination of inert alum-
ina and catalytic nickel oxide particles approximately 500 um in
diameter. The fluid bed was operated at temperatures between 850
and 950°C and achieved better than 99.99% destruction efficiency
of the explosive components while keeping the NO emissions, princi-
pally N0„ in the 200 to 400 ppm range.
The largest scale catalytic oxidation process for chlorinated
hydrocarbons in the B.F. Goodrich CATOXID process (9). Some details
of the process, the identity of the catalyst in particular, are
considered proprietary and have not been released. As shown in
Figure 1, chlorinated byproduct from the manufacture of vinyl chlor-
ide monomer, 70% Cl, 3.5% H, 26.5% C is fed to the fluid bed unit
which operates at temperatures below 540°C. A typical feedstock
contains dichloroethylene, trichloroethane, carbon tetrachloride,
chloroform, and dichlorobutene. The resulting product consisting
essentially of HC1, HO, N„, CO and C0„ is recycled to the
02,
balanced oxychlorination used to produce vinyl chloride monomer.
During process operation, heat is recovered from the reactor off
gases and used to generate steam by extracting the heat from the
fluidized solids. The economics of the catalytic process are ob-
viously favorable since it managed to displace an existing conven-
tional thermal incineration process in mid 1974. At that time the
thermal incineration equipment was only five years old but was
requiring higher than anticipated operating and maintenance costs.
Since the production capacity of the B.F. Goodrich Calvert City
plant has been reported to be about 500 million kg/yr of vinyl
chloride monomer, the operating capacity of the catalytic fluid
bed incineration process is on the order of 10 million kg/yr of
chlorinated wastes.
Unfortunately, because of the competitive and proprietary
nature of the catalytic processes being developed, relatively
little information has been published on the kinetics of the cata-
lytic oxidation which has hindered an assessment of the potential
of this process. The present paper seeks to provide some prelimi-
ary results on the kinetics and a discussion of the advantages and
disadvantages of disposal technology.

EXPERIMENTAL

Recycle Reactor and Flow Apparatus


A flow diagram of the experimental reaction apparatus is shown
in Figure 2. The reactor feed, air and chlorocarbon, is premixed
using precision flow controllers for the gaseous components and a
thermostatted saturator for liquid chlorocarbons. The parallel
arrangement of air and air-chlorocarbon flow lines allows the
reactor feed composition in the entire range from stoichiometric
to trace amounts of chlorocarbon in the air. The saturator as shown
in Figure 3, is a Pyrex container with an internal fritted surface
to disperse the air throughout the liquid chlorocarbon. A stainless
steel sponge is used as a packing material within the saturator to
46 MANNING

HYDaOCctN CHLCSlDE
CASàON OX,DE5
nitrogen Oxygen

ilfl COMPRESSOR

Figure 1. Flow Diagram of B.F. Goodrich CATOXID process

Oven -^

Byposs Streor

Rotometer Gas
j-Ty Cool in
Chromatograph
kX^mÀti Sam.phr.g Valve?
—OO-
-©-
K Recycle Strcom

CH2CI2
Saturator

A ^r Flow Cont rollers Prchcotcr and


-
Heated
Line
Sampling

Reactor in
Shut off valve
Sond both
Needle valve

Figure 2. Flow diagram of ExDerimental Apparatus


FLUID BED CATALYTIC OXIDATION 47

Gas Inlet

Gas Outlet

Glass to Meta I
Seal
Con nee tion
to Valve

Tef Ion Bushing


Ace Thread
Connec tor

Stainless Steel
S ponge

Glass Fritted
Disk

Figure 3. Chlorocarbon saturator


4S MANNING

facilitate both gas liquid mass transfer and entrained mist elimi-
nation. In some experiments, water was added to the feed using a
similar saturator on the air line.
As shown in Figure 2, the premixed chlorocarbon air mixture
then flows to a recycle reactor flow loop. That is, the incoming
feed stream is mixed at the feed point with material that is re-
cycled from the reactor outlet by a Teflon recirculation pump. The
recycle stream flow rate can be regulated by an inline valve and
monitored on a rotameter. After mixing, the feed and recycle
streams flow through a preheater and into the catalytic reactor
shown in Figure 4. The preheater coil is a 3.5 m coil of 6 mm
quartz tubing wound helically around the central reactor tube.
The reactor tube is a vertical 22 mm i.d. quartz tube which holds
the catalyst. Both the preheater tube and reactor tube are im-
mersed in a constant temperature fluidized sand bath which is reg-
ulated to + 1°K.
The catalyst is mounted in the reactor tube as shown in Figure
5. A small indentation in the wall of the quartz tube supports a
thin slab of alumina honeycomb monolith. The catalyst is then
placed on the monolith. A 3 mm o.d. quartz thermocouple is posi-
tioned immediately above the catalyst. The catalyst is introduced
into and removed from the reactor through a removable threaded con-
nection at the top of the reactor tube as shown in Figure 3.
The gases leaving the reactor flow into a cooling coil con-
tained in a 370°K oven. The gases are thus cooled to a temperature
low enough to be within the operating limits of recycle pump but
high enough to prevent condensation of the oxidation product,
water.
All elements of the recycle loop and flow lines from the feed
point to the gas analysis section are made of teflon, quartz or
pyrex.

Sampling and Analytical Apparatus


Samples of the reactor feed and effluent streams were obtained
with a 10 port Hastalloy C sampling valve from the reactor effluent
line. In order to sample the feed gas, the bypass valves and line
shown in Figure 2 were used to divert the gas flow around the recy-
cle reactor loop. The sample was injected into a Perkin Elmer
Sigma 115 gas Chromatograph equipped with column switching valves,
heated auxiliary valve box, and both thermal conductivity and flame
ionization detectors.
The microcomputer in the Chromatograph is used to sequence the
columns through which carrier gas flows at various stages of the
analysis. The gas filling the 0.5 ml sampling loop is injected into
a Tenax column by the 10 port Valco valve. The light gases elute
rapidly from the Tenax column into the Porapak N and Molecular
Sieve columns. Nitrogen, oxygen, and carbon monoxide are trapped
in the Molecular Sieve by a switching valve so that the column is
bypassed. Shortly thereafter, the Porapak N column is isolated
by switching a second valve, trapping the carbon dioxide. Hydrogen
FLUID BED CATALYTIC OXIDATION 49

Modified Ace Thraad


Con nee tor

Tef Ion Bushing

i—»"0

Reactor Inlet

Preheat c r Co>l

Catalyst Bed
Loe at ion

Figure 4. Quartz reactor and preheat coil


50 MANNING

Quartz Thermocouple
Well

Catalyst
Pellets
Alum i na Monol i t h
Indentation
in Quor tz

Reac tant
Flow

Figure 5. Detail of Catalyst Support


FLUID BED CATALYTIC OXIDATION 51

chloride, water and chlorine then elute from the Tenax column and
are detected by the thermal conductivity detector. Next, the se-
cond valve is switched so that helium again flows through the Pora-
pak N causing carbon dioxide to elute to the detector. This pro-
cess is repeated with the Molecular Sieve column and the light
gases are detected. Finally, both the Porapak N and the Molecular
Sieve columns are bypassed so the chlorinated hydrocarbons can
elute from the Tenax column to be detected by the FID and TCD de-
tectors. Chromatograms showing the output of the two detectors
are depicted in Figure 6.

Chromatogram peak areas were converted to molar composition


on an absolute basis by calibration with known mixtures
containing
combinations of the components.

Catalyst
The catalyst tested was a commercially available A1_0 sup-
ported Cr„0 (Strem Chemical #24-0200) with properties as listed
in Table 3. Approximately 2 grams of catalyst were used in each
experiment, corresponding to a catalyst volume of approximately
0.5 ml. The catalyst was maintained in an air atmosphere during
preheating to reaction temperature at the start of each run. No
other preactivation procedures were employed.

TABLE 3

Catalyst Specifications

Type Cr„0 supported on A1.0

(Strem Chemical #24-0200)


Loading 12.5 wt percent Cr

Surface Area 62 m2/g (BET)


Pore Volume 0.3 ml/g
Size 0.40 cm x 0.40 cm pellets

RESULTS

Typical run
CH^Cl,, Oxidation
-

The results of a typical run are shown in table 4. In run 6,


the feed stream contained 0.318 mole % CH Cl and 3.89% HO with
the balance air as shown in the first column. The composition of
the reactor effluent at various times during the next six hours
is shown in the following columns. From examination of the di-
« MANNING

-1-
25 «s

TIME (KIN)

Figure 6. Sample Chromatogram


FLUID BED CATALYTIC OXIDATION 53

TABLE 4

Composition of Feed and Effluent Streams (Volume Percent)

Time (hr) Feed 0.530 2.500 3.620 4.570

Hydrogen Chloride 0.000 4.472 4.605 4.705 4.404

Water 3.890 4.372 3.940 3.738 3.535

Chlorine 0.000 0.000 0.000 0.000 0.000

Carbon Dioxide 0.024 0.211 0.191 0.212 0.192

Oxygen + Argon 20.780 20.543 20.635 20.353 20.363

Nitrogen 73.510 73.257 73.753 74.579 73.837

Carbon Monoxide 0.000 0.000 0.000 0.000 0.000

Dichloromethane 0.318 0.091 0.087 0.087 0.096

Tricholomethane 0.000 0.007 0.006 0.007 0.006

Volume percent water in feed stream as determined


from the average water vapor pressure 2.4%. =

Reactor operating conditions: pressure 16.4 psia


temperature 497°C
feed rate 370 ml/min
catalyst charge 2.00 g.

chloromethane mole fraction during the experiment, the conversion


(fractional disappearance of dichloromethane) appears to be stable
at 72%. For this run at 77°K, the nominal catalyst contact time
is approximately 115 ms.
From the tabulated data one problem area is also apparent
the reported HC1 concentration exceeds that which is possible at
-

complete conversion and hence does not satisfy a material balance


for chlorine. During our experimentation we noted that the thermal
conductivity response factor for HC1 seemed to vary by a factor of
up to 3.0 depending on the water content of gas samples. Since the
calibration standards employed were all anhydrous, the resulting
response factors consistently yielded HC1 concentrations higher than
the actual values. In later experiments, HC1 was analyzed by
aqueous titration for comparative purposes. New methods for cali-
brating for HC1 are presently being studied in our laboratories.
54 MANNING

Data Analysis
a reactor configured as shown in Figure 2, where a large
For
portion of the gas stream leaving the reactor is pumped by a recycle
pump back to the inlet of the reactor and mixed with a small amount
of fresh feed, a material balance will indicate that the reactor
inlet and outlet compositions will be only differentially different.
Similar arguments or an energy balance will show that the temperature
is also essentially constant through the reactor. Thus a material
balance can be formulated as

F x. =
F x. . r. V
i,out i,m -
i

where F is the molar flow rate of a key component (e.g., N„) ,

x is the mole ratio of component i to the key,


r. is the volumetric rate of appearance of i,
and V is the reactor volume. Rearrangement gives
F r
r. =

V
(x. x. . )
^
x -

i,out - i,in
Thus, for any data point such as those shown in Table 4, the
rate ofdisappearance of the chlorocarbon can be calculated. This
rate isexpected to be a function of the known, uniform composition
and temperature. Thus, the resulting rates can be formulated as a
function of the temperature and composition by using standard regres-
sion techniques. In particular, the rate of dichloromethane disap-
pearance was formulated as an Arrhenius power law expression of the
form
E
a b
("
ri ko lï} PCH2C12 P02
=
6XP

resulting correlation is presented graphically in Figure 7.


The
The regression indicates that the rate of oxidation is first order
in chlorocarbon and half order in oxygen concentration. In view of
the range of 0„ concentrations which was limited to between 9 and
20 mole percent, the half order dependence with respect to 0. may
not be statistically significant. More data at reduced oxygen
levels are required to verify this dependence.
There appears to be some scatter in the data at the various
temperature levels shown in Figure 7. As the scatter appeared to
be most directly affected by water concentration, a second correla-
tion was performed in which the oxygen order was set arbitrarily
to zero and the order with respect to water was determined. The
resulting improved correlation is shown in Figure 8. The resulting
kinetic expression is

1
dNCH Cl
tH2L12 16,900. -0.35
, SQ p
r =

-
w~c -dE-= 3'89 exp (_
,
T~} pch9ci, Vo
p

¿LI

where r =
rate of chlorocarbon oxidation, g moles/s-g catalyst
FLUID BED CATALYTIC OXIDATION 55

CM
O
CL
c
O in -1 3.0
h- o
u i
z
Z>
LL Ü
CM
z X -14.0 h
o u

5
ÜJ
£
S So -15.0 h
O u
U

-16.0
1.3 1.4 1.5 1.6
RECIPROCAL TEMPERATURE,103/T,°K"1
Figure 7. Plot of logarithm of the Rate Expression Derived from the Partial
Pressures of Oxygen and Dichloromethane.

O
'M
I
O
0- -10.0 -

r-
u
z
r>
ti. Ö
i

_C\J
U -11.0 h
C\|
I
u
CL
üJ
ce
ce
O
U -12.0 h
o
u

INVERSE TEMPERATURE, KT/T, K


Figure 8. Plot of the logarithm of the Rate Expression Derived from the
Partial Pressures of Oxygen, Dichloromethane.
56 MANNING

P =
component partial pressures, torr
R =
gas constant, 1.987 cal/mole °K
T =
reactor temperature, °K

The form of the rate expression and trends in the dichloro-


methane oxidation data are similar to those presented by other
investigators for oxidation of non halogenated hydrocarbons and
carbon monoxide over supported chromia catalysts. Indeed, as part
of automobile exhaust catalyst research extensive investigation of
these types of catalysts were performed. Others including Yao (10),
Yao and Kummer (11,12) Varghese and Wolf (13) have measured the
rates of oxidation of CO and several olefins such as propylene and
butène reported oxidation rates which were first order in the spe-
cies being oxidized, near zero order in oxygen, and "inhibited" by
water vapor. In addition, Shelef et al. (14) and Laidig et^ al.
(15) reported inhibition by HC1 and competition between various
oxidizable species in the presence of halogen. With the uncer-
tainty in our present HC1 analysis we cannot determine the extent
of HC1 inhibition, if any. The competitive reactions are apparent
as shown in Figure 9. In this figure, the ratio of the CO to C0„
mole fraction is shown as a function of reactor operating temper-
ature. Because of the nature of the recycle this ratio is also
the ratio of the rates of production of the two species. At the
moment, the exact reaction network in the oxidation of dichloro-
methane is unknown, but the following two step oxidation is plaus-
ible.

CH Cl + 1/2 CO + 2 HC1
02 rate =
* r

CO + 1/0, ->
C02 rate =
r

Thus the ratio of the rates of production of CO to CO- would


be expectedto vary in the same manner as

rl
"

r2 ri
—-

r
- —

r
-1 =
f(T, x.)
i
2 2
That is, the shift in selectivity may well reflect the compe
tition between CO and CH Cl for similar sites on the oxidation
catalyst.
Other competing reactions may occur on the same surface. For
example, as shown in Table 4, some trichloromethane is also found
in the product indicating that chlorination of the dichloro species
is also occuring simultaneously. In other experiments, traces of
tetrachloromethane were also found. In general, however, less than
a few percent of the dichloromethane is converted to more chlori-
nated species rather than to the carbon oxides. Bhat (16) has shown
that HC1 can increase the alkylation activity of silica alumina
catalysts. This would be expected to lead to the formation of chlor-
inated C species on the catalyst support; however, to date no such
alkylation activity has been detected.
FLUID BED CATALYTIC OXIDATION 57

0 12 3
CARBON MONOXIDE /CARBON DIOXIDE
RATIO

Figure 9. Temperature Variation of Carbon Oxide Selectivity


5X MANNING

Experiments of the same type have also been performed for di, tri,
and perchloroethylene. The correlation of the associated rate con-
stants for tri and perchloroethylene oxidation are shown in Figures
10 and 11 as a function of inverse temperature and a comparative
plot of the predicted and measured reaction rates at the various ex-
perimental conditions is shown in figures 12 and 13 respectively.
As can be seen the oxidation rates are again first order in chloro-
carbon and half order to near zero order in oxygen concentration.
However, the oxidation of perchloroethylene shows no inhibition by
water in the range of partial pressures where inhibition was found
for the other chlorocarbons. The activation energies, 10,300 cal/
mole, for oxidation of the perchlorethylene is significantly less
than the activation energies for less chlorinated dichloromethane
(16,900 cal/mole) and trichlorethylene (24,700 cal/mole). While it
may be these three chlorocarbons do not behave similarly as sug-
gested by Laiding e_t al. (1981), the disagreement may also be due
to the use of the power law kinetic rate formulation rather than a
Langmuir-Hinschelwood formulation in analyzing the data.
Poisoning of Catalysts
In experiments, in an integral tubular reactor described pre-
viously (17), we also examined the long term stability of the
chromia on alumina catalysts. As shown in Figure 14 the conversion
of C„C1. obtained in a short packed bed, once-through reactor with
a dry air stream dropped with time on stream. The dependence of
the decay in conversion suggests that the poisoning is brought
about by the chlorocarbon, i.e., the faster the chlorocarbon is fed
to the bed or the higher the chlorocarbon concentration, the more
rapid the decay of catalytic activity. During this same period a
red deposit was collected at the reactor outlet tube. Analysis of
the catalyst after the experiment also showed chromium loss from
the catalyst. It is possible that in the absence of water, the
C„C1 and oxidation product Cl converted the Cr„0 to red chromium
oxychloride, CrO-Cl Addition of water to the feed stream is shown
.

in Figure 15 to fiait the decay in catalyst activity probably by both


shifting the oxidation product to HC1 via the Deacon reaction
2 HO + Cl Î 2 HC1 + 1/2 02
and/or by hydrolysis of the chromium oxychloride.

CONCLUSIONS

Chromia based catalysts have been shown to be effective oxida-


tion catalysts for several chlorocarbons. The observed oxidation
rates are first order in the chlorocarbon species and vary between
zero and first order in oxygen. Reaction rates appear to be slightly
inhibited by water as shown by a small negative order dependence
with respect to this species. Although some catalyst poisoning
FLUID BED CATALYTIC OXIDATION 59

1 OO

z
<
r- 0.0 0
l/l
z
O
u
UJ
< -I.OOh
cr

O
>:
i -2.0 0
cr
<
O
O
-3.0 0
1.2 1.4 1.6 1.8
RECIPROCAL TEMPERATURE , 10J/T, K

Figure 10. Arrhenius Plot for C9HCI3 Oxidation at Various Temperatures.


Rate expression is given by
24,800 0.5 -0.28
2.72 x 10 exp (- RT
C2HC13 H20
60 MANNING

0.0 Df

1.00 1.10 1.20 1.30 1.40 1.50


1
RECIPROCAL TEMPERATURE 103/T, °K

Figure 11. Arrhenius Plot for CCI Oxidation at Various Temperatu


Rate Expression is given by

r
r- = K ¿a
5.68 x in2 exp
10 10,300.
(--)
/ „
p p 0.12
u-J-i
-
RT
C2C14 r02
FLUID BED CATALYTIC OXIDATION hi

_ro
U 12.0
X
CM
U
LA
CJ '
100
O D
£

ro
O


t-
<

Z
LU
Z
et
LU
Q_
X
U

00 2.0 4.0 6.0 8.0 10.0 12.0


3 ----CpH
g moles CI-»
CALCULATED RATE,(x10J) g cot
*-
hr
-

Figure 12. Comparison of Measured C HC1 Oxidation Rates and Calculated


Rates from Regression
62 MANNING

50.0

a 4 0.0 h
O o
£ u

30.0 h

ÜJ
r-
<
ce 20.0h

z
LO
10.0 h
Ct
LU
a
x
üJ

0.0 10.0 20.0 30.0 40.0 50.0


CALCULATED RATERO4) 9molc*C2C'4
g cot -
hr

Figure 13. Comparison of Measured C Cl Oxidation Rates and Calculated


Rates from Regression
FLUID BED CATALYTIC OXIDATION 63

C2 CI4 IN FEED ( Mole •/. )

10 20 30
CATALYST TIME ON STREAM, HOURS

Figure 14. Catalyst Deactivation During Oxidation of Dry C Cl,

H/CI RATIO

-
100
o
1/1
a
>
z
O
u

U
i\l
u

10 20 30 40 50 60
CATALYST TIME ON STREAM, HOURS

Figure 15. Effect of Water in Feed on Catalyst Deactivation.


64 MANNING

phenomena have been observed, control of the feed stoichiometry,


specifically the H/Cl ratio in the mixed feed permits extended
operation at only slightly reduced rates. The hydrogen in the feed
can be supplied from steam and does not have to be provided from a
more expensive hydrocarbon fuel.
This kinetic data for a readily available catalyst represents
the first available data for assessment of the potential applica-
tions. In view of the fact that 20 to 40 percent of the cost pre-
sently associated with thermal incineration of chlorinated wastes
is for auxiliary fuel, the reduced operating expenses associated
with the reduced temperature requirements of the catalytic process
are an economic advantage. There are design problems associated
with combustion of wastes containing dissolved inorganic solids
such as those described by Becker and Wall (2,3,4) in conjunction
with a catalytic process as described here. While this may limit
the potential applications at large scale commercial waste disposal
facilities which handle very diverse wastes; there are more numerous
but smaller potential applications in the chemical process industry.
EPA estimates that 85% of the hazardous waste generated in the US
is presently being treated on-site. In this environment, where the
waste streams are less variable in composition and volume, the ap-
plication of catalytic oxidation in fluid bed reactors offers more
immediate potential application with less engineering development.

ACKNOWLEDGMENTS

The author gratefully acknowledges the financial support of


the Environmental Protection Agency, Office of Research and Develop-
ment, Office of Exploratory Research (RD-675) for support under
grants 808296-01 and 810397-01.

REFERENCES

(1) Hardison, L.C., and Dowd, E.J., Chem. Eng. Prog., 73 (7), 31
(1977)
(2) Becker, K.P. and Wall, C.J., Hydrocarbon Proc, 54 (10), 88
(1975)
(3) Wall, C.J., Graves, J.T. and Roberts, E.J., Chem. Eng., 77.
April 14, 19 75
(4) Becker, K.P. and Wall, C.J., Chem. Eng. Prog. 72 (10) 61 (1976)

(5) Ackerman, D.G., Chausen, J.F., Johnson, R.J., and Zee, C.A.
"Destroying Chemical Wastes in Commercial Scale Incinerators,
Facility Report No. 3, Systems Technology, Inc." PB 265540,
November 1976.
FLUID BED CATALYTIC OXIDATION 65

(6) Ziegler, D.L., Johnson, A.J., and Meile, J.L., RFP-2016, Rocky
Flats Div., Dow Chemical June 1973.

(7) _, Chem. Eng. News, 34, July 20, 1981

(8) Carroll, J.W., Guinwan, T.L., Tuggler, R.M., Williams, K.E.


and Lillian, D.L., Am. Ind. Hyg. Assoc. J. 40, 147 (1979)

(9) Benson, J.S., Hydrocarbon Proc, 58 (10), 107 (1979)

(10) Yao, Y.-F. Y., J. Catalysis, 28, 139 (1973).


(11) Yao, Y.-F. Y. and Kummer, J.T., J. Catalysis, 28, 124 (1973).
(12) Yao, Y.-F. Y. and Kummer, J.T., J. Catalysis, 4_6_, 388 (1977).
(13) Varghese, P., and Wolf, E.E. J. Catalysis, 59, 100 (1979).
(14) Shelef, M. Otto, K., and Otto, N.C., Advances
, in Catalysis,
27, 311 (1978).
(15) Laidig, G. , Honicke, D., and Greisbaum, K., Erd. und Kohle,
34 (8), 329 (1981)

(16) Bhat, S .G. T.,. React. Kinet. Catal. Lett., 1J7_, 93 (1981).

(17) Manning, M.P., "Environmental Engineering and Pollution Control


Processes: State of the Art Seminar" EPA-600/9-82-018, Sept-
ember 1982 p. 200.

You might also like