You are on page 1of 7

4236 Ind. Eng. Chem. Res.

1997, 36, 4236-4242

Recovery of Calcium Carbonate and Hydrogen Sulfide from Waste


Calcium Sulfide

Mark W. Brooks† and Scott Lynn*


Department of Chemical Engineering and Lawrence Berkeley National Laboratory, University of California,
Berkeley, California 94720-1462

The removal of H2S from hot coal gas using a limestone-based sorbent generates a large quantity
of calcium sulfide waste (e.g., 67 tons/day from the gasification of 1000 tons/day of 3% S coal)
that cannot be landfilled because of potential H2S evolution and sulfide leaching. In the process
proposed here, the CaS is dissolved by reaction with H2S complexed with aqueous methyldi-
ethanolamine (MDEA) or other alkanolamines. In the second step, the highly soluble Ca(HS)2
reacts readily with CO2, also complexed with aqueous MDEA, to precipitate pure CaCO3 of
uniform crystal size and to form MDEA-complexed H2S in solution. The solution from step 2 is
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

splitshalf is recycled to step 1 and half is sent to a stripper to recover H2S and from there to a
column where the needed CO2 is absorbed from a combustion gas.
Downloaded via MANHATTAN COLG on September 14, 2022 at 14:27:30 (UTC).

Introduction and Previous Work These characteristics make limestone an attractive


sorbent for H2S removal from coal gas. However, CaS
Currently, most of the coal consumed in power
cannot be placed in a landfill directly because of the
generation is burned to produce high-pressure steam,
danger of H2S evolution or leaching of the sulfide ions
which is expanded through a steam turbine system to
into the ground water. Practical possibilities for CaS
drive electric generators. The last few decades have
treatment include converting the sulfide back to CaCO3
seen intense research on cleaner and more efficient coal-
or CaO, thus regenerating the H2S sorbent, and direct
fired power generation. The technique of gasifying coal
oxidation of the sulfide to sulfate, which would allow
prior to combustion has received a great deal of this
safe disposal of the waste.
research attention.
Direct oxidation of the CaS with air produces CaO
A benefit of coal gasification at high temperature and
by the following reaction:
pressure is the greater ease of sulfur removal. The
sulfur is in the form of H2S and can be removed by a
solid sorbent with relatively little cooling of the gas. By 2CaS + 3O2 f 2CaO + 2SO2 (2)
contrast, following combustion the sulfur is in the form
of SO2, is greatly diluted, and is at near-atmospheric Unfortunately, if sulfate formation is to be prevented,
pressure. This combination increases cleaning costs. the reaction must be conducted above 1400 °C
Of the many possible solid sorbents, the least expen- (Schwerdtfeger and Barin, 1993). However, after five
sive is limestone. Fenouil and Lynn (1995a,b) found cycles their CaO retained only 15% of its initial H2S-
that limestone, which is typically about 97 wt % CaCO3, sorption capacity, probably as a result of sintering.
can reduce H2S to a level of 100-200 ppm, well below Keairns et al. (1974) proposed reversing the sorption
that mandated by the Clean Air Act, with conversion reaction at a lower temperature:
of 90+% of the calcium to CaS:
CaO + H2S f CaS + H2O (1) CaS + H2O + CO2 f CaCO3 + H2S (3)

The sorption is done in a separate step, just ahead of a However, they reported that, after 21 cycles, sintering
combustion turbine and within 25 °C of the calcination reduced the conversion of CaCO3 to CaS from 69% to
temperature, which occurs at 800-900 °C and depends 13%.
on pressure and CO2 content. (Limestone is frequently The argument could be made that the cost of lime-
mixed with the coal in the gasifier, but there is a stone is so small there is no need to regenerate the
significant loss in both calcium conversion and H2S sorbent. This may be the case, but CaS is considered
reduction and a corresponding increase in ash volume.) hazardous and must be rendered inert before disposal.
Limestone is much cheaper than zinc-based sorbents Schwerdtfeger and Barin (1993) studied oxidation to
(even with allowance for recycling the latter). Moreover, CaSO4 but concluded that it was not possible to convert
its optimal temperature being 200-300 °C higher allows CaS quantitatively to CaSO4 in a single-step process.
an improved thermal efficiency of power generation. They found that at elevated temperatures (>1100 °C)
Even when the lower H2S level obtainable with zinc- the sulfation reaction was fast but that some of the CaS
based sorbents is deemed essential (10-20 ppm), bulk was converted to CaO with the release of SO2. Evolu-
desulfurization with limestone followed by polishing tion of SO2 was avoided at lower temperatures, but the
with a zinc-based sorbent may prove to be economically oxidation was too slow for an industrial process.
attractive. Van der Ham et al. (1996) proposed a complex
reaction sequence that featured reaction (1) followed by
* Author to whom correspondence is addressed. Telephone:
(510) 642-1634. FAX: (510) 642-4778. E-mail: Lynn@ CaS + 2SO2 h CaSO4 + S2 (4)
cchem.berkeley.edu.
† Present address: Exxon Research and Engineering, P.O.

Box 101, Florham Park, NJ 07932. and


S0888-5885(97)00151-6 CCC: $14.00 © 1997 American Chemical Society
Ind. Eng. Chem. Res., Vol. 36, No. 10, 1997 4237
CaS + 3CaSO4 h 4CaO + 4SO2 (5) MDEA‚H2S + CaS f MDEA + Ca(HS)2 (8)

in which sulfidation (reaction (1)) is best carried out MDEA‚H2CO3 + MDEA + Ca(HS)2 f
near 850 °C, reaction (4) near 700 °C, and reaction (5) 2MDEA‚H2S + CaCO3 (9)
at 1100-1200 °C. They postulated that reaction (5)
occurs in the eutectic melt formed by CaS and CaSO4
above 1100 °C. Although such a mechanism would In reaction (8), MDEA has complexed H2S, keeping it
normally lead to sintering, van der Ham et al. reported in solution and readily available for reaction with solid
a drop in capacity of only about 35% after eight cycles. CaS. The products of this liquid-solid reaction are free
However, their equipment limited the temperature for MDEA and a solution of Ca(HS)2. (Calcium hydrosul-
reaction (5) to 980 °C, so the loss in capacity may fide is very soluble in water, 248.7 g/L at 20 °C (Seidell,
thereby have been minimized. 1940).) Carbon dioxide, bound in solution by MDEA as
All of the proposed regeneration and disposal schemes the bicarbonate, then reacts in a separate step with the
described above produce dilute streams of SO2 requiring aqueous Ca(HS)2 to precipitate CaCO3 and produce H2S,
further processing and/or encounter sintering problems which is complexed by the free MDEA. A portion of the
that hinder further cycles of regeneration. The produc- MDEA‚H2S stream is used to produce more Ca(HS)2 via
tion of fresh CaCO3 crystals from an aqueous slurry of reaction (8), and the remaining portion is stripped to
CaS would eliminate both problems. As early as 1869, produce a concentrated H2S gas, which may be con-
Bechamp studied reaction (3) between CO2 and solid verted to elemental sulfur by known technology. The
CaS particles suspended in water (Gmelins Handbuch stripped MDEA solution is reloaded with CO2 in an
der Anorganishen Chemie, 1957). However, the reaction absorber, with a combustion gas as the CO2 source, and
was incomplete due to the formation of a CaCO3 layer recycled. This system has several advantages compared
around the CaS. to other CaS-regeneration techniques:
Neither CO2 nor CaS is very soluble in water. (1) “Fresh” CaCO3 is obtained after each regeneration
Calcium sulfide has a solubility of less than 1 g/L at 25 cycle. This eliminates crystal-structure problems such
°C, and CO2 at 1 atm of partial pressure has a solubility as sintering and attrition that are experienced in many
of only 1.7 g/L at 20 °C (Perry and Green, 1984). high-temperature sorption/regeneration cycles. Pure
Reaction (3) occurs between the dissolved CO2 and solid CaCO3 has other uses as a product but is also a suitable
CaS at the liquid-solid interface and is limited by how material for landfill.
fast the CO2 can dissolve and diffuse. The production (2) A concentrated stream of H2S is generated from
of CaCO3 at this interface eventually encapsulates the the stripping operation, which facilitates its conversion
remaining CaS, preventing further reaction. to elemental sulfur. Utilization of CO2 is correspond-
Biswas et al. (1976) achieved 98% carbonation of 150- ingly high.
µm CaS particles in an aqueous slurry by controlling (3) The higher H2S and CO2 solubilities increase
the CO2 flow, bubble size, and slurry depth. By using reaction rates and decrease reactor volumes.
roughly 3 times the stoichiometric amount of CO2, they (4) The formation of the highly soluble Ca(HS)2
produced a gas containing about 33% H2S, just high eliminates CaS encapsulation during carbonation, re-
enough for sulfur recovery in a Claus plant. They sulting in nearly stoichiometric conversion of CaS to
proposed that a series of reactions occurs: reaction (3) CaCO3.
followed by The goal of this research was to study the critical
steps in this method for converting CaS into CaCO3 and
CaS + H2S f Ca(HS)2 (6) H2S. The effect of CaS particle size and prior CaS
oxidation on the kinetics of Ca(HS)2 generation, reaction
and (8), was determined. Reaction (9) was studied by
determining the effects of MDEA‚H2CO3 and Ca(HS)2
Ca(HS)2 + CO2 + H2O f CaCO3 + 2H2S (7) concentrations on the crystal-size distribution of the
precipitated crystals and the amount of excess CO2
Thus, CaS reacts initially with CO2, with the formation required for complete Ca(HS)2 conversion. The data
of aqueous H2S and the precipitation of CaCO3. The were then used to design a complete process for the
aqueous H2S that does not escape into the gas phase production of an enriched H2S stream and a relatively
reacts with more CaS to form the highly soluble Ca- pure (>95%) calcium carbonate product from waste CaS.
(HS)2, which can then react in solution with CO2 to
precipitate CaCO3 and produce more H2S, perpetuating Experimental Equipment and Procedures
the reaction. The rate at which Ca(HS)2 and CaCO3 are
formed is dictated by how fast the CO2 can dissolve and The limestone used in the experiments described
diffuse to a reaction site and also by the concentration below was provided by Great Lakes Calcium Corp.,
of aqueous H2S that can be maintained. A process in Green Bay, WI; Table 1 shows the average chemical
which the aqueous solubilities of both CO2 and H2S were composition. The MDEA was supplied by Union Car-
significantly increased would greatly improve the rates bide and had a purity of 99 wt %, with the balance being
of reactions (3), (6), and (7), the utilization of the CO2, water. The CaS was obtained from three sources. One
and the concentration of the product H2S gas. batch was made from limestone by the procedure
In the research presented below, aqueous methyldi- described below. The second was a sample of technical-
ethanolamine (MDEA) is used to increase the solubili- grade CaS supplied by Fisher Scientific Co. that had
ties of CO2 and H2S by forming the corresponding been exposed to air periodically over a period of several
substituted ammonium complexes. Since alkanol- years. The third was a fresh sample of reagent-grade
amines are weaker bases than CaS, solubilizing CO2 (“Alpha”) CaS supplied by Johnson Mathey Co. Table
and H2S in this way does not inhibit their reactions with 2 gives the chemical compositions of these calcium
CaS and Ca(HS)2. The reactions are as follows: sulfide samples.
4238 Ind. Eng. Chem. Res., Vol. 36, No. 10, 1997

Table 1. Limestone Composition (wt %) (Great Lakes


Calcium Corp., Green Bay, WI)
CaCO3 MgCO3 SiO2 Fe2O3 Al2O3 S
97.8 1.63 0.28 0.15 0.13 0.01

Table 2. Calcium Sulfide Compositions (wt %) and


Particle Sizes
source CaS CaSO4 CaCO3 MgCO3 SiO2 Fe2O3 Al2O3 size (µm)
Alfa 98 2 <0.1 <0.1 <0.1 <0.1 <43
Towler 95 3 <1 1 <1 <1 <1 74-88
reactor 95 3 <1 1 <1 <1 <1 1-2 mm
Fisher 74 12 5 5 2 1 1 74-88

Wet analytical techniques were used for the analysis


of solid CaS and to determine the concentrations of
MDEA, MDEA‚H2S, MDEA‚CO2, and Ca(HS)2 in solu-
tion. The purity of solid CaS and the concentration of
MDEA‚H2S solutions were determined by iodimetric
titration (Jeffery et al., 1989). The concentrations of the Figure 1. Bubble reactor flow sheet.
separate Ca(HS)2 and MDEA solutions were determined
by acidimetric titration with 1 M HCl. Analyses of °C, the particles assayed 95 wt % CaS. Before the
MDEA, MDEA‚H2S, and Ca(HS)2 mixtures were deter- sample was removed, the reactor was cooled to room
mined by first analyzing for the total sulfur content temperature under a purge of nitrogen.
iodimetrically and then titrating a separate sample with Bubble Reactor. An experimental apparatus was
1 M HCl. Although three different basic compounds built to absorb the acid gases H2S and CO2 with MDEA.
were present, their pKa’s were sufficiently similar that The apparatus, illustrated in Figure 1, was constructed
only one inflection point could be determined accurately. of a 250-mL Erlenmeyer flask with a side arm. Ad-
This point occurred around a pH of 4.5. Since the ditional details are given by Brooks (1995).
MDEA concentration in the samples was always known, In a typical experiment, 100 mL of a 60 wt % MDEA
the excess acid required to reach the endpoint was due solution was placed in the bubble reactor. The pH of
to the presence of Ca(HS)2. The H2S concentration could the solution was initially found to be 10.5. After purging
then be determined by the difference from the total the reactor with nitrogen, a flow of acid gas was
sulfur content. The amount of CO2 complexed with established. The solution was gently mixed with a 2-cm
MDEA in solution was determined gravimetrically by stirring bar. The flow rates were monitored by cali-
reacting with excess aqueous Ba(OH)2 to form solid brated rotameters, and the effluent gas was scrubbed
BaCO3. The solid was filtered, dried, and weighed. with a 2 M NaOH solution as indicated in Figure 1. For
Differential Reactor. The differential quartz reac- the CO2-loading case, the CO2 flow rate was established
tor designed and built by Towler (1992) and described at 2 mL/s. After 3 h, the pH reached a constant value
by Fenouil and Lynn (1995a) was used for the high- of 8.95. The time and flow correspond to 1.75 times the
temperature production of CaS from limestone. The stoichiometric amount of CO2 required for MDEA
design of the reactor allows a continuous flow of gas to saturation. A sample of the solution was then taken
be heated to reactor temperature, to pass by the and acid-titrated to determine how much water had
limestone sample, and then to exit. The limestone was been lost from solution during CO2 bubbling. The
first calcined to CaO at about 850 °C and then reacted MDEA concentration would typically increase from 5.04
with H2S to form CaS. (Sulfidation occurs much more to 5.28 M. To determine the CO2 loading, a sample of
rapidly and completely with CaO than with CaCO3 MDEA‚CO2 solution was reacted with Ba(OH)2 solution,
(Fenouil and Lynn, 1995a,b).) as noted above. The amount of CO2 in solution was
In these experiments, roughly 10 g of millimeter-sized determined with a precision of (0.3%. At a typical CO2
limestone particles, contained in a small quartz con- loading, 21% of the MDEA was found to be complexed.
tainer, was placed in the quartz reactor. The position The procedure for loading the MDEA with H2S was
of the limestone sample was located at the maximum virtually identical to the loading with CO2. The nitro-
temperature of the reactor, which was then set to 850 gen purge flow was replaced by an H2S flow of 0.6 mL/
°C, while a nitrogen flow of 0.55 mL/s was established. s. As the reaction proceeded, the pH was monitored and
The calcination temperature in an atmosphere of nitro- reached a constant of 8.05 after 7 h. The degree of H2S
gen is below 850 °C: loading was typically 70% of the MDEA and was
determined with a precision of (1.9%. The system was
CaCO3 f CaO + CO2 (10) monitored for leaks continuously when using H2S.
These acid-gas-loaded solutions were stored in tightly
Reaction was complete after 3 h at 850 °C; the stones sealed bottles under a fume hood, but acid gas still
turned from gray to white. Weight measurements were evolved upon opening. This required the loadings to be
used to determine when conversion was complete. redetermined before each set of experiments.
After calcination, sulfidation was conducted by lower- Calcium Hydrosulfide Generation. Calcium hy-
ing the reactor temperature to 800 °C and establishing drosulfide was produced via the liquid-solid reaction
the flows of N2 and H2S at 0.55 and 0.45 mL/s, described in reaction (8). In a typical experiment, a CaS
respectively. The reaction could be monitored qualita- sample of known purity and particle size was added to
tively by color change going from white to gray. A 3 mL of an MDEA‚H2S solution of known concentration
sample was then taken and titrated iodimetrically to and loading in a 10-mL reactor flask. The CaS was kept
determine the sulfur content quantitatively with a as the limiting reagent, with 35% excess H2S. After
precision of (0.2 wt %. After 8 h of sulfidation at 800 capping the flask, the slurry was agitated vigorously on
Ind. Eng. Chem. Res., Vol. 36, No. 10, 1997 4239
a magnetic stirrer. The reactor was placed in a ther-
mostated water bath kept at 20 °C. After a measured
amount of time, the reaction was stopped by vacuum
filtration of the unreacted CaS from the liquid. The CaS
was then washed, dried under nitrogen, and weighed,
and the sulfide content was determined with a precision
of (0.2 wt %. From this information the conversion of
CaS to Ca(HS)2 could be determined with a precision
of (0.5%. A check for experimental accuracy was done
by comparing the molar ratio of Ca(HS)2 found in the
filtrate to the decrease of CaS remaining in the solid.
The values ranged from 0.93 to 0.99.
Calcium Carbonate Precipitation. Experiments
to precipitate calcium carbonate via reaction (9) were
performed by mixing Ca(HS)2 and MDEA‚H2CO3 solu-
tions. The resulting CaCO3 crystals were separated
from the liquid, washed, and dried. The crystal size was
measured with an optical microscope.
In a typical experiment, 4 mL of an MDEA‚H2CO3
solution and 6 mL of a Ca(HS)2 solution were simulta-
neously added to a mixing flask at 22 °C. The concen-
trations of each solution were known, and Ca(HS)2 was
kept as the limiting reagent. Addition of these two
Figure 2. Effect of particle size on hydrosulfidation of CaS derived
solutions to the mixing flask was done in 2 min. from limestone.
Precipitation of CaCO3 crystals occurred immediately
upon mixing of the solutions. Gentle stirring of the
resulting CaCO3 slurry was continued for 5 min to
ensure complete reaction. The CaCO3 crystals were
then separated from solution by vacuum filtration,
washed, dried, and weighed. The conversion of Ca(HS)2
to CaCO3 was determined with a precision of (1.8%.
The crystal-size range of the CaCO3 product was
determined with a Cambridge Instruments Stereo Zoom
Light microscope. The structure of the crystals was also
analyzed with a scanning electron microscope. The
experiment was repeated for different Ca(HS)2 concen-
trations to determine if there was any effect on crystal
size. In addition, the amount of excess CO2 was reduced
to near-stoichiometric amounts to determine the effect
on Ca(HS)2 conversion.

Experimental Results
1. Effect of CaS Particle Size on Rate of Hydro-
sulfidation. Conversion-versus-time data were gath-
ered for 95 wt % CaS particles of two different particle
diameters. The CaS was made by the high-temperature
sulfidation of limestone. Figure 2 illustrates the data
for particles with diameters of 1-2 mm and also of 74- Figure 3. Effect of CaS oxidation on CaS hydrosulfidation.
88 µm. Figure 2 shows that the rate of CaS hydrosul-
fidation increases as the size of the particles decreases, prior to or as part of the CaS/H2S reactor to decrease
as would be expected from the increase in the external the reaction time for CaS digestion.
surface area available for reaction. Particles in the size 2. Effect of Prior CaS Oxidation on Hydrosul-
range of 1-2 mm required 120 min to achieve a fidation Reaction. The effect of CaS oxidation and
fractional conversion of 0.85-0.90. This is in contrast impurities on conversion is illustrated in Figure 3. The
to particles in the size range of 74-88 µm, which CaS particles had purities of 74, 95, and 98 wt %. Most
required only 15 min to reach the same conversion. The of the impurity was the result of oxidation of CaS to
porosities of the samples were not measured separately the sulfate resulting from exposure to air. Some inerts
but were assumed to be similar. No attempt was made such as MgCO3 and SiO2 were also present. The
to model the dissolution process mathematically because particles were reacted with a 3.5 M solution of
of the high fraction of CaS (∼95 wt %) in these particles. MDEA‚H2S; the conversion of CaS to Ca(HS)2 was
The matrix of insoluble residue remaining after the CaS measured versus time. Figure 3 shows how the rate of
was dissolved would be too irregular and variable for hydrosulfidation of CaS decreases as the purity of the
such modeling to be either meaningful or useful. CaS decreases. It appears that the decrease in the rate
A moving-bed sorption system would use millimeter- of conversion was caused by the encapsulation of inner
sized limestone particles for H2S removal and produce deposits of CaS by CaSO4. This is most evident in the
CaS particles of the same size. To produce CaCO3 from conversion of the 98 wt % CaS, which reached complete
waste CaS on an industrial scale, a CaS-crushing conversion in 4 min. The high rate of reaction of this
operation (such as a ball mill) might be installed either material also reflects the smaller particle size. The 98
4240 Ind. Eng. Chem. Res., Vol. 36, No. 10, 1997

To determine the dependence of the conversion on the


amount of excess CO2 in solution, the CO2 was reduced
to an excess of only 4%. These experiments also
resulted in a yield of CaCO3 of about 0.96. Thus high
conversions could still be attained at near-stoichiometric
levels of CO2. The CaCO3 crystal size was measured
and found to remain in the range of 5-10 µm. Experi-
ments were also conducted with the addition of 5-10-
µm CaCO3 seed crystals. The molar quantity of seed
crystals added was equivalent to 25% of the Ca(HS)2 in
solution. It was reasoned that if there were CaCO3
crystals present, the fresh CaCO3 would prefer to
precipitate on the existing crystals rather than nucleate
fresh crystals. However, it was found that the addition
of seed crystals had no effect on the resulting range of
CaCO3 crystal sizes.
The next set of experiments was conducted with a
Figure 4. Calcium carbonate crystals. decrease in Ca(HS)2 concentration to 0.104 M; the
MDEA‚CO2 concentration remained constant. It was
wt % CaS was a freshly purchased storeroom chemical thought that decreasing the Ca(HS)2 concentration
with roughly half the particle size of the other CaS might decrease the rate of formation of CaCO3 crystals
particles tested. Because the quartz reactor was un- and result in fewer but larger crystals. However, the
available at the time, it was not possible to prepare resulting crystal size again remained in the range of
fresh, unoxidized CaS from limestone for these experi- 5-10 µm.
ments. The settling velocity of the CaCO3 crystals was
In an industrial process, the CaS content of the determined in water at 22 °C and found to be about 7.5
material leaving the H2S-sorption step would be about × 10-3 cm/s. This corresponds to a particle size of 9.5
80 wt %. The impurities would consist mainly of µm according to Stokes’ law, which agrees approxi-
unreacted CaO, together with magnesium and silicon mately with the size found microscopically. At 20 °C
compounds. The CaO would not hinder the hydrosul- the nominal viscosities of water and a 60 wt % MDEA
fidation reaction since CaO will also react with H2S to solution are 1 and 23.7 mPa‚s, respectively (Bryan
form Ca(HS)2. The CaS stream would be cooled and Research and Engineering, 1994). Thus, the settling
transported to the hydrosulfidation step while minimiz- velocity of the CaCO3 particles in the 60 wt % MDEA
ing contact with the air to prevent excessive formation solution should be of the order of 3 × 10-4 cm/s. Since
of CaSO4 by oxidation. Using a liquid-flooded ball mill the lowest settling velocity recommended for separation
as the reactor would eliminate any encapsulation of CaS by sedimentation is 10-2 cm/s, separation by filtration
that does exist by simultaneously reacting the CaS and or centrifugation would be used.
crushing away the inert material. Process Design and Evaluation. A process based
3. Precipitation of Calcium Carbonate. Solu- on the results above was designed to treat the CaS
tions of Ca(HS)2 and MDEA‚H2CO3 were simulta- waste from a coal-gasification plant processing 1000
neously added to a mixing flask at 22 °C. The quantities metric tons of 3 wt % sulfur coal/day, i.e., 67 tons/day
and concentrations of each solution were known, and of CaS. Assuming that 80% of the CaCO3 in the feed
MDEA‚H2CO3 was kept in excess. The CaCO3 precipi- forms CaS and the remainder is calcined, the flow of
tated immediately upon mixing of the two solutions. CaO from the process would be 13 tons/day. As noted
Initial experiments were conducted with a 0.67 M Ca- above, this flow of CaO forms Ca(HS)2 upon reaction
(HS)2 solution and a 60 wt % MDEA solution that was with H2S, leading to the consumption of additional CO2
21% loaded with CO2. The ratio of the solutions gave a in the carbonation step. Inert material in the limestone
35% excess of CO2 upon mixing. During these experi- will pass through undissolved.
ments, the yield of CaCO3 was determined to be 0.96- Figure 5 illustrates the process-flow diagram for the
0.99; the loss was ascribed to experimental uncertainty CaS-treatment system. The simulation program
rather than slow kinetics. The photographic images of TSWEET (Bryan, 1995) was utilized for the design of
crystals of CaCO3 were then measured and found to be the absorber and stripper sections of this design.
in the range of 5-10 µm. Details of the process parameters used in the simulation
Figure 4 shows the CaCO3 crystals as filmed with a are given by Brooks (1995). The process starts with the
scanning electron microscope (SEM). The SEM photo- feed of hot (∼200 °C) millimeter-sized CaS waste
graph shows two different CaCO3 crystal structures. particles along with an 8 mol % excess of MDEA‚H2S
Most of the crystals have a roughly spherical shape with solution to CSTR 8, where a slurry forms and the
diameters of 5-10 µm. The remaining crystals are hydrosulfidation reaction begins. (Since the MDEA
slightly smaller and have a rhombohedral crystal struc- solution also contains some complexed CO2, some Ca-
ture of about the same size as the spheroids. All of the CO3 also forms in CSTR 8.) The residence time in CSTR
crystals have a relatively narrow size distribution. 8 (and for CSTR’s 6 and 7 and Reactor 9) was set at 5
There is a striking absence of very fine material. Both min, giving a volume for each of about 6.0 m3. The
crystal structures are suitable for separation by filtra- partially reacted CaS slurry is then fed to a ball mill
tion. The known crystal structures of CaCO3 (calcite) (or other liquid-filled grinder), Reactor 9, where the
are hexagonal and rhombohedral. The presence of the remaining CaS particles are simultaneously crushed
spheroidal particles may be caused by impurities in and reacted with MDEA‚H2S to form Ca(HS)2. Reactor
solution. The SEM photograph has a magnification of 9 and CSTR 8 combine to achieve a 99% conversion of
2800, and the white line is 3.57 µm long. CaS to Ca(HS)2. The assumption that freshly-formed
Ind. Eng. Chem. Res., Vol. 36, No. 10, 1997 4241

Figure 5. Process configuration for recovery of H2S and CaCO3 from CaS.

CaS would react fully within the allotted residence time Table 3. Operating Parameters for Stripper
of 5 min in each of the two reactors would have to be ideal actual tray spacing diameter reboiler % H2S
demonstrated in a pilot plant. stages stages (m) (m) duty (kW) stripped
A fraction of the stream leaving Reactor 9 may be 10 20 0.61 0.63 1296 95.5
recycled through a cooler back to CSTR 8, if needed, for
temperature control. If one wishes to prevent inert Table 4. Operating Parameters for Absorber
solids from being mixed with the CaCO3 to be formed ideal actual tray spacing diameter gas flow % CO2
in CSTR 6, the slurry of those solids suspended in the stages stages (m) (m) rate (m3/h) absorbed
solution of Ca(HS)2 and free MDEA leaving Reactor 9 10 30 0.51 0.81 592 94
may be pumped to Filter 10. (The inert solids would
absorber. Typically, a combustion gas would be used
consist of small amounts of CaCO3 and CaS along with
as the source of CO2 for the process. This stream,
gangue such as MgO and SiO2 coming from the original
comprised of about 10% CO2, would be cooled and
limestone.)
compressed before being fed to the absorber. The
Either directly or by way of Filter 10, the solution of
absorption process benefits from operation at higher
Ca(HS)2 plus MDEA leaving Reactor 9 goes to CSTR 6,
pressures, so the operating pressure of the absorber is
where it reacts with MDEA‚H2CO3 solution and CaCO3
set by an optimization that will be site-specific. The
precipitates. The relative quantity of MDEA‚H2CO3
CO2 loading of the MDEA solution leaving the absorber
solution fed to CSTR 6 is a 5-10 mol % excess relative
was specified to be 30-35%, which requires operating
to Ca(HS)2 to ensure substantially complete precipita-
the absorber at the inlet pressure of the combustion
tion of calcium ion. The resulting CaCO3 slurry is then
turbine. This CO2-rich MDEA stream is fed to CSTR 6
fed to a second reactor, CSTR 7, where the reaction
for the CaCO3 precipitation reaction. Table 4 gives the
reaches completion before the slurry is pumped to Filter
operating parameters for the absorber. Operation of the
11 for CaCO3 recovery. The CaCO3 is filtered and
absorber at a relatively low pressure might lead to the
washed free of mother liquor using condensate from the
choice of diethanolamine (or another alkanolamine) as
stripper condenser. A system that includes Filter 10
sorbent to improve the kinetics of CO2 absorption.
would produce 116 tons of high-purity CaCO3/day from
Filter 11 for the flows above.
Conclusions
The solution leaving Filter 11 is free of Ca(HS)2 but
contains MDEA that is about 25% loaded with H2S and In the process described above, a stream of raw, solid
1-2% loaded with CO2 at a temperature of about 60 CaS is dissolved into an aqueous solution by reaction
°C. Roughly half of this solution is pumped to CSTR 8, with MDEA-complexed H2S. The Ca(HS)2 so formed is
and the remainder is pumped to a stripping column. The converted by reaction with MDEA-complexed CO2 into
stripper produces a concentrated gas stream (90-95% relatively pure CaCO3, regenerating MDEA-complexed
H2S) and a lean solution of MDEA. Operating param- H2S. Part of the latter is stripped, forming a concen-
eters for the stripper are given in Table 3. trated stream of H2S. The process utilizes MDEA (or
The MDEA stream coming from the bottom of the another alkanolamine) to increase the concentrations
stripper is cooled from 135 to 75 °C by heat exchange of CO2 and H2S that can be dissolved in an aqueous
with the H2S-rich feed. The MDEA is then cooled solution, thereby greatly improving the kinetics of both
further with cooling water before being fed to the reactions.
4242 Ind. Eng. Chem. Res., Vol. 36, No. 10, 1997

It was determined experimentally that the CaS Bryan Research and Engineering. TSWEET Simulation Program,
particle size and purity greatly influence the rate of CaS Version 95.0; Bryan Research and Engineering: Bryan, TX,
hydrosulfidation. As the CaS particle size decreases 1995.
Fenouil, L. A.; Lynn, S. Study of Calcium-Based Sorbents of High
and the purity increases, the rate of hydrosulfidation Temperature H2S Removal. 1. Kinetics of H2S Sorption by
increases. As the purity of the CaS particles decreases Uncalcined Limestone. Ind. Eng. Chem. Res. 1995a, 34, 2324-
(particularly by partial oxidation to sulfate), the impuri- 2333.
ties encapsulate interior regions of CaS and slow the Fenouil, L. A.; Lynn, S. Study of Calcium-Based Sorbents of High
hydrosulfidation reaction. Temperature H2S Removal. 2. Kinetics of H2S Sorption by
The precipitation of CaCO3 from the reaction between Calcined Limestone. Ind. Eng. Chem. Res. 1995b, 34, 2334-
2342.
Ca(HS)2 and MDEA‚CO2 occurs very rapidly, and reac- Gmelins Handbuch der Anorganishen Chemie; Ca 8 [B] (Auflage
tion is essentially complete with only a 5% stoichiomet- Calcium); Verlag Chemie GMBH: Weinheim, 1957; p 645.
ric excess of CO2. The CaCO3 crystals produced have a Jeffery, G. H.; et al. Vogel’s Textbook of Quantitative Chemical
relatively narrow crystal-size distribution centered at Analysis, 5th ed.; John Wiley & Sons, Inc.: New York, 1989;
5-10 µm, with a striking absence of fines. The addition pp 389-398.
of seed crystals to this precipitation reaction had no Keairns, D. L.; et al. Sulphur Emission Control with Limestone/
Dolomite in Advanced Fossil Fuel Processing System. Environ-
effect on the crystal size nor did varying the Ca(HS)2 mental Aspects of Fuel Conversion Technology; St. Louis, MO,
concentration or the reactant stoichiometry. 1974; Research Triangle Institute: Research Triangle Park, NC,
The use to which the CaCO3 would be put will vary 1974.
with the circumstances. At worst it would form an Perry, R. H.; Green, D. Perry’s Chemical Engineers’ Handbook,
innoccuous landfill. Pelletization and reuse as sorbent 6th ed.; McGraw-Hill Book Company: New York, 1984.
for H2S may prove practical. The highest value would Schwerdtfeger, K.; Barin, I. Problems in Hot Desulfurization of
Coal Gas with Lime. Erdoel Kohle, Erdgas, Petrochem. 1993,
be derived from maximizing its purity and size charac- 46 (3), 103-110.
teristics to generate a pigment or a chemical raw Seidell, A. Solubilities of Inorganic and Organic Compounds; D.
material. Streams of raw SrS or BaS could be converted Van Nostrand Company, Inc.: New York, 1940.
to their respective carbonates by analogous chemistry. Towler, G. P. Synthesis and Development of Processes for the
Recovery of Sulfur from Acid Gases. Ph.D. Dissertation, Uni-
versity of California at Berkeley, Berkeley, CA, 1992.
Acknowledgment Van der Ham, A. G. J.; Heesink, A. B. M.; Prins, W.; van Swaaij,
This research was funded by the Morgantown Energy W. P. M. Proposal for a Regenerative High-Temperature Process
for Coal Gas Cleanup with Calcined Limestone. Ind. Eng. Chem.
Technology Center through the U.S. Department of Res. 1996, 35, 1487-1495.
Energy under Contract DE-AC03-76SF00098.
Received for review February 12, 1997
Literature Cited Revised manuscript received June 16, 1997
Accepted June 17, 1997X
Biswas, S. C.; Sabharwal, V. P.; Dutta, B. K. Hydrogen Sulphide
from Reduced Gypsum. Fert. Technol. 1976, 13 (4), 255-258. IE970151J
Brooks, M. W. Recovery of Hydrogen Sulfide and Calcium Carbon-
ate from Waste Calcium Sulfide. M.S. Thesis, Department of
X Abstract published in Advance ACS Abstracts, August 15,
Chemical Engineering, University of California at Berkeley,
Berkeley, CA, 1995. 1997.

You might also like