Production of Calcium Magnesium Acetate (CMA) From Dilute Aqueous Solutions Acetic Acid
Production of Calcium Magnesium Acetate (CMA) From Dilute Aqueous Solutions Acetic Acid
Production of Calcium Magnesium Acetate (CMA) From Dilute Aqueous Solutions Acetic Acid
Daniel B. Leineweber
Master of Science
(Chemical Engineering)
at the
University of Wisconsin-Madison
December 25, 2002
Abstract
i
lution of CMA, and, more generally, that it is possible to achieve a wide
range of other product compositions if desired. Feasibility is confirmed by an
experimental continuous dissolution process, and the performance of three
different neutralizer materials is compared: type S dolime and selectively
calcined dolomite are both found to be adequate choices for future applica-
tions; however, the use of uncalcined dolomite is not recommended due to
its very low reactivity. Selectively calcined dolomite provides the option of
producing a high-magnesium CMA with a calcium to magnesium mole ra-
tio of 2:3, which might be suitable for crystallization as a double salt; this
finding makes the almost unknown selectively calcined material particularly
interesting. The preliminary design of a process for the production of CMA
from wood hydrolyzate (waste sugars) and selectively calcined dolomite is
presented, and an economic analysis shows that this new process compares
well to previously proposed ones.
ii
Acknowledgements
iii
and support is greatly appreciated.
Several Undergraduate Students have worked with me on experiments
during their ChE 424 Summer Lab, and D. Kotowsky has helped me with
the preparation of selectively calcined dolomite during the fall semester.
Last but not least I also want to thank my parents in Germany for their
continuous support and my girl friend Elizabeth for her patience and love.
iv
Contents
3 Equilibrium Models 32
3.1 Nonideal Effects in Aquatic Solutions . . . . . . . . . . . . . . 33
3.2 The Hydroxide System . . . . . . . . . . . . . . . . . . . . . . 37
3.3 The Carbonate System . . . . . . . . . . . . . . . . . . . . . . 49
v
4.2.1 Carbonates: Calcite, Magnesite, Dolomite . . . . . . . 67
4.2.2 Oxide Minerals: CaO, MgO . . . . . . . . . . . . . . . 83
4.3 The Dynamic Model and Its Limitations . . . . . . . . . . . . 89
4.3.1 Outline of the Process Model . . . . . . . . . . . . . . 90
4.3.2 Model Uncertainties and Conclusions . . . . . . . . . . 105
vi
List of Tables
vii
A.4 Concentrations of impurities . . . . . . . . . . . . . . . . . . . 156
viii
List of Figures
ix
5.5 Block flow diagram of acetic acid pulping . . . . . . . . . . . . 131
5.6 Mass balance for production of 1:1 CMA . . . . . . . . . . . . 134
5.7 Mass balance for production of 2:3 CMA . . . . . . . . . . . . 136
5.8 Process flow diagram for CMA production . . . . . . . . . . . 137
x
Notation
Dimensions are given in terms of mass (M ), length (L), time (t), temperature
(T ), and amount of substance (N ).
Latin Letters
A area, L2
A constant in Debye-Hückel formula, L3/2 N −1/2
A constant in Davies’ formula, dimensionless
a adjustable size parameter in Debye-Hückel formula, L
Ai molar activity of species i, N L−3
Āi effective surface area for species i, L2
B constant in Debye-Hückel formula, L1/2 N −1/2
b constant in Davies’ formula, dimensionless
C molar concentration in bulk solution, N L−3
C (s) molar concentration at surface layer, N L−3
Ceq equilibrium concentration, N L−3
(s)
Ceq equilibrium concentration at surface layer, N L−3
Ci molar concentration of species i, N L−3
D coefficient of diffusion, L2 t−1
xi
D dilution rate, t−1
G Gibbs free energy, M L2 t−2
∆G molar free energy change of reaction, M L2 t−2 N −1
∆G0 standard free energy change of reaction, M L2 t−2 N −1
F volumetric flow rate, L3 t−1
I ionic strength, N L−3
J molar flux, N t−1 L−2
JC molar flux due to surface reaction, N t−1 L−2
JT molar flux due to transport process, N t−1 L−2
K equilibrium constant, dimension varies
Kc concentration equilibrium constant, dimension varies
Ksp solubility product, dimension varies
ki rate constant of chemical reaction i, dimension varies
kC rate constant of heterogenous chemical reaction, dimension varies
kT mass transport coefficient, Lt−1
M molar mean molecular weight, M N −1
Mi molecular weight of species i, M N −1
N amount of substance, N
Ni amount of species i, N
Nx amount of biomass, N
n empirical reaction order, dimensionless
P pressure, M L−1 t−2
Pi partial pressure of component i, M L−1 t−2
Q reaction quotient, dimension varies
R gas constant, M L2 t−2 T −1 N −1
R net rate of heterogenous reaction, N t−1 L−2
xii
Rf forward rate of heterogenous reaction, N t−1 L−2
Rb backward rate of heterogenous reaction, N t−1 L−2
Si symbol for species i, dimensionless
T absolute temperature, T
TOT total (analytical) concentration, N L−3
t time, t
V volume, L3
xi mole fraction of species i, dimensionless
z rectangular coordinate, L
zi charge number of species i, dimensionless
Greek Letters
αP growth related constant for product formation, dimensionless
αS growth related constant for substrate uptake, dimensionless
βP non-growth related constant for product formation, t−1
βS non-growth related constant for substrate uptake, t−1
δ thickness of diffusion layer (Nernst theory), L
γi activity coefficient of species i, dimensionless
µ growth rate, t−1
µi chemical potential of species i, M L2 t−2 N −1
µ0i standard free energy of species i, M L−2 t−2 N −1
νi stoichiometric coefficient of species i, dimensionless
ξi mass fraction of species i, dimensionless
xiii
Subscripts and Superscrips
(aq) aqueous
b backward
C chemical reaction
dol dolomite
eq equilibrium
f feed
f forward
(g) gaseous
P product
S substrate
(s) solid
(s) surface layer
sp solubility product
T transport process
T total (analytical)
x biomass
xiv
Chapter 1
1
2
There is also concern about possible effects on human health caused by high
levels of sodium in drinking water. It has been estimated that the total annual
damage resulting from the use of sodium chloride as road deicer is at least
15 times the cost for purchase and application of the salt [Cho91]. However,
these economic losses must be related to the benefits of road salting like
reduced number of accidents, or reduced business losses due to transportation
delays—economically the benefits still outweigh the losses. Hence it pays to
keep highways snow- and ice-free in winter, but it would pay even more if
this could be done without the adverse effects.
In the early 1970s an extensive research program was initiated by the
Federal Highway Administration (FHWA) in order to investigate the effects
of road salt and find ways to minimize the damage. As a result many im-
provements were made to bridge construction, e.g. sealed concrete bridge
decks, epoxy-coated reinforcing bars, and cathodic protection, which are ex-
pected to slow down bridge deterioration considerably. Measures to reduce
the amount of salt needed for the desired melting effect were also successfully
implemented, mainly through improved application techniques that made it
possible to get the salt to the proper location in the proper amount and
to keep it there. But all of this could not really solve the problems associ-
3
ated with the use of sodium chloride, thus a search for alternative deicing
chemicals was begun.
In 1976 the FHWA awarded a contract for the study of alternative de-
icers to Bjorksten Research Laboratories, Inc. in Madison, WI. This study
was conducted by Dunn and Schenk who became the “inventors” of CMA
[Dun80]. Starting from the periodic table they first eliminated all chemical
elements that were radioactive, toxic or expensive, and by considering deicing
properties, corrosivity, etc. of the compounds that could be made from the
remaining ones they were able to eliminate several more candidates. Only
nine elements were left as possible constituents of new deicing chemicals:
H, C, N, O, Na, Mg, P, K, and Ca. Compounds containing these elements
were grouped in organic compounds, inorganic salts, and mixed salts (con-
sisting of an inorganic and an organic component), and solubility, eutectic
temperature, solution pH, and estimated cost were evaluated for each com-
pound. From the continued elimination process finally two candidate deicers
remained, methanol and calcium magnesium acetate. Both were studied
extensively in lab and field experiments to assess their deicing properties,
potential corrosive effects, and toxicity. Later methanol was also eliminated
because of its flammability, solvent nature, and poor persistency on road
surfaces, thus leaving only CMA for further consideration.
At this point the FHWA and the state highway agencies directed their
efforts towards a comprehensive project for the complete evaluation of CMA.
Numerous institutions were involved in those studies1 made during the 1980s.
General characteristics and deicing properties of CMA were determined, its
practical application was tested in several field trials, production methods
1
For a list of references, see [Cho91].
4
• less harmful to soil, groundwater, human health, plant and animal life,
universities, and states are continuing to search for economically feasible ways
of manufacturing CMA, and to further improve the deicing and handling
properties of the material.
Still another potential application for CMA has been proposed recently
[Wis91, Chapters 9–14]: it can be used as an additive to coal combustion,
where it catalyzes the combustion and at the same time acts as a “sulfur
grabber”, forming solid calcium sulfate and thus substantially reducing sulfur
dioxide in the stack gases. The coal merely has to be impregnated with a
CMA solution before combustion. Used in this manner, CMA would help to
prevent acid rain, and would also enhance the efficiency of coal-fired boilers.3
In conclusion, CMA is a very interesting new chemical for environmental
applications—both of its proposed major uses can be expected to have a
large positive impact on the environment.
Table 1.1: Properties of CaAc2 and MgAc2 ; solubilities from [Lin65], all
other properties from [Mer83].
tures with water as reported by [Dun80] are −15◦ C for CaAc2 and −30◦ C
for MgAc2 compared to −21.1◦ C for NaCl; the corresponding phase diagrams
are shown in Figures 1.1 and 1.2. From this data one should expect bet-
ter deicing properties for high-magnesium CMA or even consider the use of
pure magnesium acetate. In fact the evaluation of eutectic data for CMA
solutions of various Ca:Mg mole ratios revealed that optimum freezing point
depression was obtained for a Ca:Mg ratio of 3:7 over a wide range of CMA
concentrations. However, under field conditions this 3:7 CMA tended to be
much slower in its ice melting rate than the originally proposed 1:1 CMA, so
finally the 1:1 ratio was recommended again as the optimum deicer composi-
tion [Sch91]. This last remark may not apply to the double salt mentioned
above—[Tod90] claim for their CMA unique properties that are not com-
parable to those of a physical mixture with the same composition.
Probably as important as the chemical composition is the physical form
of the deicer. As Gancy points out [Gan84a, Gan84b], dense and course
particles are required for effective ice removal, because it is sufficient to just
break up the ice layer by “drilling holes” down to the pavement and weak-
ening the pavement-ice interface. The ice layer then fractures and is finally
cleared away by the ongoing traffic. Hence it would be an enormous waste
of deicer material to use a fine powder or a brine in an attempt to melt all
the ice on the road surface. The most preferred physical form for a CMA
deicer is the hard, coarse, and non-friable flake which can be produced from
wet CMA using a drum pelletizer or similar equipment.
8
10
CaAc2 – H2 O liquidus
measured data 3
CaAc2 – H2 O eutectic line
I
03 measured data 2
NaCl – H2 O eutectic +
3
3
3
3
3 3
−10 II
temp. 2 3
(◦ C) 2 2 2 2
2
−20 +
III
−30
−40
0 10 20 30 40 50
CaAc2 conc. (wt%)
Figure 1.1: Phase diagram for CaAc2 – H2 O. In region I there is only liquid
CaAc2 solution present, in region II there is water ice and liquid CaAc2 solu-
tion, and in region III there are two solid phases, water ice and CaAc2 ·2H2 O.
Eutectic of NaCl – H2 O included for comparison. Adapted from [Dun80].
9
10
MgAc2 – H2 O liquidus
measured data 3
I MgAc2 – H2 O eutectic line
03 measured data 2
NaCl – H2 O eutectic +
3
3
−10
temp. 3
(◦ C) 3
II
−20 +
3
2
−30 2 2 2
2 III
−40
0 10 20 30 40 50
MgAc2 conc. (wt%)
10
11
than with most other organic acids [Wag78]. Vinegar, a dilute aque-
ous solution of acetic acid (approximately 5%), has been used as a food
acidulant and preservative for thousands of years. Today acetic acid
is also widely used in the chemical industry as a solvent and as a raw
material for many organic syntheses, e.g. the manufacture of vinyl ac-
etate and cellulose acetate. The acetic acid production capacity in the
United States was 1.6 million tons in 1990, while only about 1.1 million
tons were actually consumed [Bus90]. At the present time the main
production routes for acetic acid are liquid-phase oxidation of n-butane,
and methanol carbonylation, both with feedstocks derived from natu-
ral gas or petroleum [Bus90, Wag78]. Alternative processes include
the destructive distillation of wood, and the fermentation of ethanol or
sugars, but except for the production of vinegar which is still made by
fermentation, these routes were considered no longer competitive when
cheap feedstocks based on natural gas and petroleum were introduced
following World War II. However, more recently there is again a grow-
ing interest in producing acetic acid from renewable biomass sources
like corn or even organic wastes, and much research has been done in
this field [Bus90, Gho85, Sch82b, Wan78]. The price1 for synthetic
glacial acetic acid was between $ 500 and $ 600 per ton in 1990.
As will be seen later, equilibria and kinetics of these reactions are highly
dependent on pH and tend to be quite different for calcium and magnesium,
2
According to a personal communication with Dunn, the type N material suffers from
the low reactivity of the hard-burned magnesium component.
15
respectively. Generally the hydroxides react much faster with acetic acid
than the carbonates.
The ways of reacting acetic acid with dolime or dolomite to make CMA,
and the subsequent steps necessary to get a dry product in the form
of flakes or pellets have been extensively studied, and many of them
are patented [Gan83, Gan84a, Gan84b, Gan86, Gan87b, Tod90,
Rip86]. Chevron Chemical Co. began commercial production of their
CMA-deicer ICE-B-GONTM in 1985, and they are now one of the main
producers of CMA. The selling price for ICE-B-GONTM was $ 657
16
per ton in 1989; it has been estimated that on a large production scale
a price between $ 400 and $ 450 per ton could possibly be reached—still
too much to really consider complete replacement of sodium chloride
by CMA, since rock salt is available for between $ 25 and $ 50 per ton.
In order to obtain an affordable deicing product it has been suggested
to produce CMA-coated sand or substitute part of the acetate by cheap
chloride [Gan87a].
calcium acetate (or some other acetate salt) and easily decom-
posable ammonium compounds like ammonium hydroxide. This
solution is first sent into a scrubbing column to recover the ex-
pensive ammonia and then through a multi-effect evaporator and
dryer, where the water is boiled away and a dry acetate salt is
obtained as final product. In the scrubbing tower, the broth is
contacted with the vapor from the first effect of the evaporator;
the vapor leaving the top of the scrubbing tower serves as the
heating medium in the second effect. It contains about 96% of
the ammonia which can be directly recycled back to the fermen-
tor for reuse as neutralizer.
1. The actual yields of acetic acid are often more than 85% (g acetic acid
per g glucose) for homoacetate fermentations, compared to only 50%
for the vinegar process.
2. Homoacetogenic bacteria are able to produce acetic acid not only from
C6 sugars like glucose, but also from some C5 sugars like xylose that
would represent a significant portion of the sugars produced by hydrol-
ysis of lignocellulosic feedstocks. In the vinegar process only glucose is
metabolized.
or in sum,
C6 H12 O6 −→ 3CH3 COOH. (2.8)
Since acetic acid is the only fermentation product, a relatively “clean” solu-
tion of CMA could be obtained without need for expensive separation tech-
niques. If organic wastes containing cellulose and hemicellulose are to be used
as a feedstock, these materials must be transformed into a mixture of soluble
sugars by hydrolysis under mild conditions before they can be fermented.
Wang et al. found that C. thermoaceticum produces acetic acid not only
during growth, but also when cell growth ceases [Wan78]. They used the fol-
lowing model to describe the growth-related and non-growth-related aspects
of product formation,
dNP
= (αP µ + βP )Nx , (2.10)
dt
where NP denotes the produced amount of acetic acid in moles, µ the spe-
cific growth rate, and Nx the total biomass in moles, based on the average
elemental composition of cells.4 Since for exponential growth
dNx
= µNx , (2.11)
dt
equation (2.10) can be rewritten as
dNP dNx
= αP + β P Nx , (2.12)
dt dt
4
Often the composition CH1.8 O0.5 N0.16 S0.0045 P0.0055 is used for biomass, corresponding
to a molecular weight of 24.4 g mol−1 [Rie91].
22
i.e., the total acetic acid production rate depends linearly on both the biomass
growth rate and the amount of biomass already present; the corresponding
terms in equation (2.10) are often referred to as growth term and maintenance
term. The linear relationship was found to be valid for specific growth rates in
the range of 0 to 0.15 h−1 with αP = 0.8 (mole acetic acid per mole cells) and
βP = 0.065 h−1 (mole acetic acid per mole cells per hour) at pH 7.0 [Wan78].
In a very similar manner, the substrate consumption can be modeled by the
expression
dNS
= (αS µ + βS )Nx ,
− (2.13)
dt
where NS denotes the amount of substrate consumed in moles, and the overall
yield of product on substrate (“actual” yield) is then given by
ov dNP /dt
YPS =− . (2.14)
dNS /dt
ov
A typical value for YPS on a mole basis is 2.55 (mole acetic acid per mole glu-
cose), or, equivalently, 0.85 on a weight basis (g acetic acid per g glucose).5
From these results it can be concluded that growth is not absolutely re-
quired for acetic acid production—a high productivity could also be achieved
through high cell density at minimal growth using immobilized cells or cell
recycling.
Direct anaerobic digestion of cellulosic waste by a mixed culture of acid-
formers is another process that has been considered for CMA production
[Tra91, Tra90, Wis88]. The organisms are found in stable manure, pond
mud, and sewage sludge; some of them—called cellulolytic bacteria—are able
to hydrolyze cellulose or hemicellulose, while others depend on the solu-
ble sugars and alcohols provided by the metabolic activity of the former.
5
This shows that it is very important to indicate on which basis the yield has been
calculated, although it is a dimensionless quantity.
23
Concentration of
2+
Ca Mg2+ Ac− Total CMA
tolerance levels 7 12 50
ions even higher levels of acetate, magnesium, and calcium might be toler-
ated). Calcium has a relatively high toxicity and would definitely be one of
the factors limiting the attainable concentration of CMA. As shown in Ta-
ble 2.1, it should be theoretically possible to obtain a CMA concentration of
at least 50 g l−1 using this improved strain; at 60 g l−1 the calcium tolerance
would be slightly exceeded, and at 70 g l−1 both calcium and acetate would
be limiting. For these calculations, a 1:1 molar ratio of Ca:Mg was assumed.
Experimentally, 32 g l−1 of a 1:3 CMA were obtained by Wiegel et al. in
their pH controlled batch fermentation [Wie91]. Thus it still is somewhat
optimistic—but not too unrealistic—to assume that it will be possible to at-
26
tain a CMA concentration of about 50 g l−1 (4.8 wt%) with currently available
strains of C. thermoaceticum. This assumption shall be used throughout the
rest of this study, although it falls short of the target concentration stated
above. Possibly, an organism might be genetically engineered that possesses
a higher tolerance to calcium and acetate in order to meet the target. It
has been also proposed to use halophilic organisms from the Dead Sea that
can naturally tolerate the conditions of highly concentrated ionic solutions,
in particular high levels of calcium ion; actually such organisms have been
successfully weaned from sodium chloride and encouraged to produce acetic
acid [Hud88]. But keeping in mind that it is quite difficult to obtain high
product concentrations with continuous processes, 7–10 wt% does not seem
to be a reasonable goal at the moment.
Different studies have shown that C. thermoaceticum is far more sensi-
tive to the undissociated acetic acid than the acetate ion [Sch82b, Wan84].
While acetate ion can be tolerated at concentrations of up to 48 g l−1 , total
growth inhibition by undissociated acetic acid already occurs when its con-
centration is as low as 2.8 g l−1 [Wan84]. It is quite instructive to calculate
the maximum attainable total acetic acid concentration (undissociated acetic
acid plus acetate ion) as a function of pH by the Henderson-Hasselbalch
equation
[Ac− ]
log = pH − pKa (2.15)
[HAc]
and the mole balance
assuming that neither of the two respective inhibitory concentrations for ac-
etate ion and undissociated acid may be exceeded. Then the maximum total
27
pH Concentration of
HAc Ac− HAc plus Ac−
substrate CO2
6
?
¾
dolime
slurry
fermentor 6
(pH 6–6.8)
CO2 - - pH control
unit
Figure 2.1: Fed batch fermentation process for CMA production, pH is con-
trolled by addition of a dolime slurry; process as used by [Wie91].
tion of HAc production. A simple pH-controlled fed batch process for CMA
production as the one used by Wiegel et al. [Wie91] is shown in Figure 2.1,
a continuous process could be controlled in a similar way.
After completion of the fermentation, the pH of the broth must be ad-
justed to pH 8–9, because traces of free acetic acid in the product would
be damaging to concrete. This pH adjustment could be done by adding an
extra amount of lime, but the sluggish reaction makes it difficult to obtain
a uniform and stable product. For this reason, it has been suggested to use
a small amount of potassium hydroxide instead of lime for the “fine tuning”
[Gan84a]. At the same time, the desired 1:1 mole ratio of Ca:Mg should
be achieved in the solution. If there remain undissolved solids, they may or
may not be removed before drying the product. Remaining particles could
actually improve deicing properties, since they would enhance friction.
29
These observations were obviously made on the basis of simple model sys-
tems and equilibrium calculations—“real” fermentation experiments showed
remarkably different results. As part of their CMA-project Wiegel et al.
produced 150 lb of CMA by fermentation [Wie91]. They grew C. ther-
moaceticum in fed batch fermentation on hydrolyzed corn starch with bub-
bling of CO2 through the medium; the temperature was maintained at about
60◦ C; the pH was kept in the range of pH 6–6.8 by addition of a dolime
slurry during fermentation, and it was finally adjusted to about pH 8 using
an extra amount of dolime; the duration was 5–6 days. In contrast to the
results cited above they found [Wie91, page 384]:
For instance, they reported final soluble concentrations of 58.8 mM for cal-
cium and 168.0 mM for magnesium, while the total concentrations for these
elements were 142.09 mM and 179.25 mM, respectively. They did not check
the nature of the precipitates (e.g. hydroxide, carbonate, phosphate), and
they did not offer an explanation. Among the possible reasons for these dif-
ferences are temperature effects, the precipitation of calcium carbonate (due
to CO2 bubbling) or less significantly, the precipitation of calcium phosphate
(because the medium contained a certain amount of phosphate). It should
be also pointed out that the fermentation is necessarily a non-equilibrium
process and that ultimately a dynamic approach must be taken in order to
explain the process behaviour.
31
Equilibrium Models
32
33
for each species, while the so-called equilibrium constant method depends on
information about the independent reactions that take place in the system
and their equilibrium constants. Since stability constants are readily avail-
able for most complexes and solids of interest in aquatic systems [Mar89],
the second technique is used more frequently in this context.
“Small” chemical equilibrium problems can be solved by hand using a
methodology that allows to make significant simplifications to the original
set of nonlinear equations and also gives some insight on how the system
would behave under altered conditions [Mor83]. More complex problems
involving many different species and reactions must be solved numerically on
a computer; appropriate software for the study of aquatic systems has been
developed. Most of these programs use the Newton-Raphson algorithm
to solve the set of nonlinear equations for equilibrium composition. The
sophisticated ones like MINEQL [Mor72] can handle nonideality corrections
necessary for concentrated systems and can even find the correct set of solids
in cases where it is unknown which solid phases are present at equilibrium.
Since nonideal effects must be considered for the systems to be discussed, a
brief introduction to nonideality corrections is in place.
µi = µ0i + RT ln Ci (3.2)
µi = µ0i + RT ln γi Ci . (3.3)
0 = ν 1 S1 + ν 2 S2 + ν 3 S3 + · · · (3.7)
The constant term ∆G0 is called the standard free energy change of the reac-
tion, and Q is the reaction quotient. At equilibrium, ∆G = 0 and therefore
∆G0
à !
Q = exp − =K (3.10)
RT
Aνi i = K,
Y
(3.11)
i
37
where the activity coefficients have been brought to the other side. The term
on the right side of (3.12)
γi−νi
Y
Kc = K (3.13)
i
Recipe 1:
• 0.5 M HAc
The problem was solved “almost” by hand using only a programmable pocket
calculator; essentially the same solution was obtained from the TITRATOR
software package which is available on MS-DOS machines [Cab87]. The
species considered to be present at equilibrium were H2 O, H+ , OH− , HAc,
Ac− , Ca2+ , CaOH+ , Ca(OH)2 (s), CaAc+ , Mg2+ , MgOH+ , Mg(OH)2 (s), and
MgAc+ . Among these species, the following independent reactions can be
written (equilibrium constants from [Mar89], at zero ionic strength and
25◦ C):
H2 O = H+ + OH− , log K = −14.0
HAc = H+ + Ac− , log K = −4.76
CaOH+ = Ca2+ + OH− , log K = −1.15
Ca(OH)2 (s) = Ca2+ + 2OH− , log K = −5.19
(3.14)
CaAc+ = Ca2+ + Ac− , log K = −1.2
MgOH+ = Mg2+ + OH− , log K = −2.6
Mg(OH)2 (s) = Mg2+ + 2OH− , log K = −11.1
MgAc+ = Mg2+ + Ac− , log K = −1.3
As principal components, the species H2 O, H+ , Ac− , Ca(OH)2 , and Mg(OH)2
were chosen. The number of components (five) is in fact equal to the number
of species (thirteen) minus the number of independent reactions (eight) since
each reaction defines a stoichiometric relationship among species which allows
to express one of them as a formula of the others, i.e. for each species it is
39
H+ 1
OH− −1 −14.0 −13.7
HAc 1 1 +4.76 +4.46
Ac− 1
Ca2+ 2 1 +22.81 +23.12
CaOH+ 1 1 +9.96 +9.96
Ca(OH)2 (s) 1
CaAc+ 2 1 1 +24.01 +23.71
Mg2+ 2 1 +16.9 +17.2
MgOH+ 1 1 +5.5 +5.5
Mg(OH)2 (s) 1
MgAc+ 2 1 1 +18.2 +17.9
Recipe TOT X
HAc 1 1 0.5 M
Ca(OH)2 (s) 1 0.3 M
Mg(OH)2 (s) 1 0.3 M
41
and from the rows, the mass laws for the species are (assuming zero ionic
strength)
In general, these mass laws are valid for activities rather than concentrations,
but it is possible to retain the given convenient form at ionic strength greater
than zero by simply using the concentration equilibrium constants Kc defined
in equation (3.13) instead of the standard constants K. The activities of solid
species are always fixed to one, hence they do not appear in (3.19)–(3.26).
To obtain a first rough estimation for the equilibrium composition, the
uncorrected mass laws (3.19)–(3.26) were substituted into the mole bal-
ances (3.15)–(3.18), resulting in four nonlinear equations for the component
concentrations [H+ ], [Ac− ], [Ca(OH)2 (s)], and [Mg(OH)2 (s)]. It was possi-
ble to further eliminate [Ac− ] from the TOT H balance (3.15) using equa-
tion (3.16); thus only one equation of one unknown had to be solved numeri-
cally, which could be easily done with the built-in equation solver of a pocket
calculator. The result is shown in Table 3.2 (first column). From this data
42
Table 3.2: Equilibrium speciation for the hydroxide system (Recipe 1). pH
corresponds to negative log of hydrogen ion activity rather than concentra-
tion.
the ionic strength I could be calculated: the value found was I = 0.35 M,
indicating a fairly concentrated solution.
For a more accurate result, ionic strength corrections were required. Ap-
proximated values for the activity coefficients γi at I = 0.35 M were obtained
from Davies’ formula (3.6), which gave γi = 0.70 and γi = 0.24 for ions hav-
ing single and double charge, respectively. Then the concentration equilib-
rium constants Kc could be determined according to (3.13); their log values
are given in Table 3.1. Using the corrected constants, a slightly different
equilibrium composition was found (Table 3.2, second column). The differ-
ences were smaller than expected, in particular for the total soluble calcium
and magnesium concentrations, both of which did not change significantly.
The resulting new ionic strength was I = 0.48 M, its increase was mainly
due to the increase of [Ca2+ ] (double charge!) at the expense of [CaAc+ ].
It was not necessary to make another iteration using I = 0.48 M, because
the activity coefficients have almost constant values in the range I = 0.3
to 0.7 M, so that no further significant changes were expected, taking into
account the restricted accuracy of such calculations.
In order to confirm the validity of these calculations, the equilibrium
composition for Recipe 1 was also determined experimentally. The system
was prepared from deionized water and reagent-grade HAc, Ca(OH)2 , and
Mg(OH)2 . Stirring and sparging with nitrogen were provided during equi-
libration; the pH of the solution was monitored. After approximately half
an hour the pH had stabilized, and it was assumed that the system had
reached equilibrium. The undissolved solids were removed by filtration, and
the sample solution was analyzed for total soluble calcium and magnesium by
inductively coupled plasma emission spectroscopy (ICP). It was found that
44
pH and total soluble calcium had been predicted with an error of less than
10%, but the concentration of magnesium had been grossly underestimated
by almost two orders of magnitude (see Table 3.2, third column1 ). One pos-
sible explanation for this significant deviation might be that the system had
not yet reached “true” equilibrium, i.e., the solution was still oversaturated
with respect to magnesium species.
Very similar values were obtained for industrial-grade type S dolime,
supplied by The Western Lime & Cement Co. in West Bend, Wisconsin:
pH 12.47, 0.258 M total soluble calcium, and 4.03 · 10−5 M total soluble
magnesium. Obviously the impurities2 did not have a strong influence on
the equilibrium speciation.
The theoretical as well as the experimental results show that in ac-
cordance with the observations by [Mar85], only a negligable amount of
Mg(OH)2 (s) dissolved because of the high pH, and essentially a solution of
calcium acetate was produced, not one of CMA. However, it must be pointed
out that in the context of the proposed fermentation process a pH > 12 will
never be reached. This leads to the more important question what happens,
if there is no excess of Ca(OH)2 (s) in the system, and the pH is allowed to
drop far enough for dissolution of Mg(OH)2 (s).
An answer to this question was found by simulating a titration of calcium
and magnesium hydroxide with acetic acid, i.e., by determining the equilib-
rium composition depending on the total amount of acetic acid added to
the system. These calculations were made using the TITRATOR program.
Unfortunately it turned out that the available option of “automatic” ionic
1
The original data on the ppm scale can be found in Appendix A.1.
2
See Appendix B.1 for a chemical analysis of the lime.
45
strength corrections did not work properly, so zero ionic strength had to be
assumed. In order to restrict the error introduced at higher ionic strength,
only half the hydroxide concentrations of Recipe 1 were chosen according to
Recipe 2:
The results are visualized separately for solid species and total soluble con-
centrations in Figures 3.1 and 3.2; the equilibrium concentration of H+ has
been included in both diagrams. There are two interesting points, where
some of the concentrations change dramatically: one at TOT HAc = 0.3 M,
where just all Ca(OH)2 (s) is dissolved, and another at TOT HAc = 0.6 M,
where also the last bit of Mg(OH)2 (s) has disappeared. As long as there is
Ca(OH)2 (s) present, a pH > 12 is maintained, and the solution phase con-
tains only traces of magnesium species. After Ca(OH)2 (s) has disappeared,
the pH drops to a value of about pH 9–10 and again remains almost constant
while the dissolution of Mg(OH)2 (s) takes place; finally it drops further to
the region of pH 4–6 (the pKa of HAc is 4.76, i.e. a pH of about this value
would be reached when the concentrations of acetate ion and undissociated
acetic acid become equal).
Several real titration experiments were performed with the individual
hydroxides and the mixed system according to Recipe 2. After each addition
of HAc, the pH was allowed to stabilize before a reading was taken and the
titration was continued. The time required for a stable pH reading was less
46
0
×××××
×××
×
−2 ×
−4
3
33
−6 3
log C
−8
3333333
333
−10 3
−12 33333
333333 Ca(OH)2 (s) ×
Mg(OH)2 (s)
H+ 3
−14
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
TOT HAc
×× ××××××××××××××××××××
× ××
−2 ×
−4
3
33
−6 3
log C
−8
3333333
333
−10 3
−12 33333
333333 total soluble Ca ×
total soluble Mg
H+ 3
−14
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
TOT HAc
than 20 sec for Ca(OH)2 , but as long as 10 min for Mg(OH)2 , indicating that
Mg(OH)2 reacted significantly slower than Ca(OH)2 . A good agreement was
found between the obtained experimental curves of pH versus TOT HAc and
the simulation results stated above: Ca(OH)2 (s) dissolved at pH > 12 and
Mg(OH)2 (s) at pH 9–10.
Therefore, no thermodynamic limitation should be expected for the dis-
solution reaction of Mg(OH)2 (s) at any pH < 9, and the experiment showed
that although Mg(OH)2 (s) dissolved slowly compared to Ca(OH)2 (s) the dis-
solution rate was still considerable. This result is in contrast to the observa-
tion reported by [Mar85] that “MgO becomes essentially inert at pH > 6”.
Since [Mar85] used unslaked dolomitic lime CaO · MgO rather than highly
hydrated (type S) dolime Ca(OH)2 · Mg(OH)2 , their problem was most likely
due to hard-burned (sintered) MgO with low reactivity, hence the exergonic
reaction
MgO(s) + H2 O −→ Mg(OH)2 (s)
proceeded very slowy at pH > 6. Apart from using type S dolime which is
more expensive than quicklime and normal hydrated lime, two other possi-
bilities to overcome this problem would be
must proceed to the right until Ca(OH)2 (s) is exhausted if PCO2 is fixed
[Mor83]. The same is true for the corresponding magnesium compounds;
again the reaction
Recipe 3:
• 0.5 M HAc
H+ 1
OH− −1 −14.0 −13.7
HAc 1 1 +4.76 +4.46
Ac− 1
CO2 (g) 1
H2 CO∗3 1 −1.5 −1.5
HCO− 3 −1 1 −7.8 −7.5
CO2−
3 −2 1 −18.1 −17.2
Ca2+ 2 1 −1 +9.75 +10.06
CaOH+ 1 1 −1 −3.1 −3.1
CaAc+ 2 1 1 −1 +10.95 +10.65
CaHCO+ 3 1 1 +3.21 +3.21
CaCO3 1 −5.15 −5.15
CaCO3 (s) 1
Mg2+ 2 1 −1 +10.64 +10.95
MgOH+ 1 1 −1 −0.76 −0.76
MgAc+ 2 1 1 −1 +11.94 +11.64
MgHCO+ 3 1 1 +3.85 +3.85
MgCO3 1 −4.06 −4.06
MgCO3 (s) 1
HAc 1 1 0.5 M
CaCO3 (s) 1 0.3 M
MgCO3 (s) 1 0.3 M
CO2 (g) 1 ?
52
Table 3.4: Equilibrium speciation for the carbonate system (Recipe 3). pH
corresponds to hydrogen ion activity rather than concentration.
Recipe 4:
using TITRATOR; Figures 3.3 and 3.4 show the resulting titration curves.
Initially the pH has a value of about pH 8, and it only slightly decreases while
MgCO3 (s) and then CaCO3 (s) dissolve. After both solids have disappeared, it
drops to the region of pH 4–6. There are no big concentration “jumps” at the
point TOT HAc = 0.3 M as was the case for the hydroxides—the reactivities
of MgCO3 (s) and CaCO3 (s) are different, but not very much. Because of the
slow reaction, no real titration experiment was performed; this would have
required an automated titrator.
An experiment using the double carbonate dolomite CaMg(CO3 )2 instead
of a mixture of CaCO3 and MgCO3 showed quite different results: the reac-
tivity was much lower than for the system defined by Recipe 3 (the pH was
still below pH 6 after several hours of equilibration), and very surprisingly,
more calcium than magnesium was found in the filtered sample taken after
24 hours. The concentrations were 0.121 M and 0.0854 M for calcium and
magnesium, respectively. Again, an excess of base was used (about 0.6 M
dolomite), and the material was finely pulverized. It is difficult to find an
explanation for this strange behaviour; first of all, the system probably did
not reach true equilibrium, and perhaps the limestone contained not only
dolomite, but also some calcite which could possibly dissolve much faster
55
0
××××××××××××××××
×××
×
×
−2 ×
−4
3
33
−6 3
log C
33
333333 3333333
−8 333333
3
−10
−12
CaCO3 (s) ×
MgCO3 (s)
H+ 3
−14
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
TOT HAc
0
××××
×× ×××××
××
×
−2 × ×××××
× ×××
×
×
−4
3
33
−6 3
log C
33
333333 3333333
−8 333333
3
−10
−12
total soluble Ca ×
total soluble Mg
H+ 3
−14
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
TOT HAc
59
60
There was just one single publication found that specifically dealt with the
dissolution of limestone in acetic acid [Sua84], and nothing about the reac-
tion of acetic acid with lime. Probably some of the more general models could
be applied to the production of CMA, but difficulties are to be expected,
since these models were found using CO2 -water systems in combination with
strong mineral acids like HCl or H2 SO4 . Hence it would be necessary to per-
form experiments in order to verify model structure and parameters. Most
likely the dynamic model will turn out to be a system of differential algebraic
equations (DAEs)—some differential equations describing the heterogenous
surface reactions combined with a set of algebraic equations (mole balances
and mass laws) which determine the speciation in the bulk fluid.
Another question is whether the design of the process can be based on
such a dynamic model which will necessarily involve large uncertainties (liter-
61
ature data on the dissolution rate of one specific compound at a given pH and
temperature may differ by an order of magnitude or more among different
sources). It may be possible to overcome this problem by a “clever” process
design that does not require exact knowledge of the dissolution kinetics.
Of these, steps 1 and 5 are tranport processes, steps 2–4 are chemical pro-
cesses. Depending on which step is the slowest and therefore rate determin-
ing, the overall rate may be either
boundary bulk
solid layer solution
6
Ceq a
b
C
- z
Figure 4.1: Concentration profiles near the interface for (a) transport con-
trolled, (b) chemically controlled, and (c) intermediate-type heterogenous
reactions.
• of intermediate type (general case), if both rates are of the same order
of magnitude and hence the observed rate is determined by a function
of the two.
Figure 4.1 shows the corresponding concentration profiles near the interface
for a dissolution reaction involving only one solute and the solvent. In case
of a chemically controlled reaction the bulk concentration C of the solute
extends right to the surface, and there is no boundary layer. In both other
cases the concentration at the interface is not the same as in the bulk so-
lution; there must exist a boundary layer with a concentration gradient. If
the observed rate is transport controlled then equilibrium is established at
63
the interface, i.e., the surface concentration is equal to the equilibrium con-
centration Ceq . In the general case the surface concentration has some value
between C and Ceq .
The classical theory of heterogenous reactions was formulated by Nernst
at the beginning of the century. He assumed that the chemical processes are
always much faster than the transport processes, that there is no concen-
tration gradient within the bulk solution in well-stirred systems, but the
concentration varies linearly with distance (measured normal to the surface)
in a thin boundary layer adhering to the solid surface, and that the thickness
of this layer is only a function of the stirring rate and geometry of the system,
but does not depend on the coefficient of diffusion of the solute, the viscos-
ity, and the temperature [Bir52]. In general, none of these assumptions is
justified, but the theory provided a useful approximation at least for many
transport controlled reactions. Consider a solid of surface area A being in
contact with a volume V of solution. Then it can be shown by Fick’s law
and a mass balance that
dC DA dC
=− , (4.1)
dt V dz
where dC/dt is the rate of change of concentration in the bulk solution, dC/dz
is the concentration gradient normal to the surface and D is the coefficient of
diffusion of the solute. Assuming equilibrium at the interface and linearity,
the gradient can be written
dC C − Ceq
= , (4.2)
dz δ
where Ceq is the equilibrium concentration and δ the thickness of the diffusion
layer; by substitution in (4.1),
dC DA
= (Ceq − C). (4.3)
dt Vδ
64
This is a very simple first order rate equation, but reasonable agreement was
found for many heterogenous reactions. Careful analyses have shown that
the values obtained for δ are too large to be interpreted as above, and that
the linearity assumption for the gradient is far from reality. As more and
more chemically controlled reactions were discovered, the need to generalize
from the transport-limited case was realized.
In order to extend the Nernst theory to reactions that are at least partly
controlled by surface reaction, the equilibrium assumption must be given up,
and an additional rate equation for the chemical reaction is required [Plu76].
The molar flux JT from the interface to the bulk solution can be expressed
as
JT = kT (C (s) − C), (4.4)
where C is again the bulk fluid concentration and C (s) is the concentration
at the outer edge of a very thin surface layer right at the interface, see Fig-
ure 4.2. The constant kT is the mass transport coefficient for the solute and
is determined by fluid properties and surface geometry. The concentration
difference C (s) − C acts as a driving force for the mass transport from the
interface to the bulk solution through the diffusion boundary layer. Equa-
tion (4.4) is a good approximation for relatively small fluxes and is valid only
if no homogenous reactions involving the solute occur within the boundary
layer. On the other hand, it is often possible to express the rate of surface
reaction by an empirical relation of the form
(s)
JC = kC (Ceq − C (s) )n , (4.5)
where JC is the molar flux into the surface layer that results from the chemical
(s)
reaction characterized by rate constant kC and empirical order n, and Ceq
65
diffusion
boundary bulk
solid layer solution
C (s) C
- -
JC JT
surface
layer
Figure 4.2: Molar fluxes due to surface reaction and mass transfer for an
intermediate-type heterogenous reaction.
J = kT (C (s) − C) = kC (Ceq
(s)
− C (s) )n , (4.6)
the limiting case of reaction control. Similarly, the transport controlled case
is obtained for kC À kT . The rate of change of concentration in the bulk
solution for dissolution from a surface of area A into a fluid of volume V is
related to the flux J by
dC A
= J (4.10)
dt V
as long as no homogenous reactions with the solute occur in the bulk fluid.
This allows to write the rate equation
dC A kT kC
= (C (s) − C) (4.11)
dt V kT + kC eq
for first order, intermediate-type reactions. In some cases, C may denote
not the bulk concentration of the dissolving species itself, but of some other
species in the system which reacts with the solid and is depleted, e.g., H+
for the dissolution of a metal in acid. However, it must be pointed out
that the equations discussed here can be applied only to very simple disso-
lution processes involving just one significant heterogenous reaction and no
homogenous reactions in the boundary layer or bulk solution. As will be seen
in the next section, surface processes involving more than one species, paral-
lel heterogenous reactions, and accompanying homogenous reactions actually
occur; the corresponding models are of course more complicated.
point because of their relevance for CMA production. From these models,
approximate dissolution rates will be calculated for the range of pH 5–7 in
order to give an idea of the orders of magnitude to be expected. The two
classes of potential raw materials, carbonate minerals and oxide minerals,
are discussed separately. More detailed information about the dissolution of
carbonates is available than about that of oxides, especially in the higher pH
range.
Soon it was realized that there exist different mechanisms for calcite disso-
lution depending on the pH regime, and the partial pressure of CO2 . The
saturation of the bulk solution with respect to Ca2+ and CO2−
3 has also an
68
important influence, since far from equilibrium the overall dissolution rate is
determined solely by the forward rate (the backward rate is neglegibly small),
while near equilibrium the forward and backward rate are of the same order
of magnitude. In some cases, experimental conditions may also play a role,
especially hydrodynamics and surface geometry of the system. For instance,
it has been shown only recently that the dissolution of calcite at pH < 4
is not necessarily transport controlled as supposed by most authors, but
rather limited by the rate of surface reaction under conditions of high mass
transport [Com90a].
The first comprehensive model of the dissolution process was the one de-
veloped by Plummer et al. [Plu78, Plu79]. In their experiments, they
used two different fractions of crushed and sized Iceland spar (estimated
surface area of the particles 44.5 cm2 /g and 96.5 cm2 /g, respectively) in a
well-stirred system at fixed PCO2 (provided by bubbling with CO2 , N2 , or a
mixture of both) and constant temperature. A “pH-stat” method was used
to study dissolution far from equilibrium at low pH, while “free drift” experi-
ments were made to examine systems nearer equilibrium. In the overlapping
region of the two methods, the obtained results were in agreement.
During the pH-stat runs, a constant specified pH was maintained through
controlled addition of a standard HCl solution using an automated titrator;
the rate of addition of standard solution dVHCl /dt was recorded. Then the
rate of calcite dissolution R could be estimated from
CHCl dVHCl
R = 0.5 , (4.12)
A dt
where CHCl is the molarity of the standard HCl solution and A the surface
area of the solid particles. This relation is justified, because at constant
2−
pH and PCO2 the activities of H2 CO∗3 , HCO−
3 , and CO3 are constant in
69
solution. Hence for each mole of CaCO3 dissolved, two moles of hydrogen
ions are consumed, and one mole of CO2 leaves the system with the bubbled
gas phase; i.e., the rate of calcite dissolution is essentially half the rate of
“hydrogen ion addition”.
In case of the free drift runs, the pH was allowed to vary as the reaction
proceeded to near equilibrium at constant PCO2 and temperature. Measure-
ment of the pH as a function of time provided a means to follow the reac-
tion, since the composition of the bulk solution at each point of time could
be calculated from pH and PCO2 using an iterative algorithm (similar to the
algorithms discussed in Chapter 3). The rate of calcite dissolution had to be
computed stepwise from this data by means of a central difference formula.
By analysis of their pH-stat data, Plummer et al. found the following
empirical expression for the forward rate of dissolution
where Ai is the activity of species i in the bulk fluid, and ki are rate con-
stants dependent on temperature. The dimensions used in [Plu78] are
M = mol l−1 = mmol cm−3 for the activities Ai , cm s−1 for the rate con-
stants ki , and hence mmol cm−2 s−1 for the rate Rf itself. Throughout this
section, another convention shall be adopted in order to make the models of
different authors better comparable and avoid confusion: all rates are mea-
sured in mol cm−2 s−1 , activities have units of M = mol l−1 , and first order
rate constants are assigned dimensions of mol cm−2 s−1 M−1 = l cm−2 s−1 ,
i.e., their numerical values differ from [Plu78] simply by a factor of 1000.
The temperature dependence of the rate constants at temperatures between
70
Table 4.1: Summary of empirical rate constants for the dissolution of cal-
cite and magnesite; the corresponding rate expressions are given by equa-
tions (4.19) and (4.29). First order constants k1 , k2 , k3 and second order con-
stants k−3 , k4 have dimensions of mol cm−2 s−1 M−1 and mol cm−2 s−1 M−2 ,
respectively.
Calcite Magnesite
2. At moderate to high pH and high PCO2 , the rate shows a linear depen-
dence on the bulk concentration of dissolved CO2 .
These regions in the pH,PCO2 plane, as well as an area where the forward rate
depends significantly on more than one mechanism, are shown in Figure 4.3.
In addition, Rf was found to be dependent on the rate of stirring in regime 1,
but not in the other two regions. This led to the conclusion that the forward
rate is controlled by H+ transport at pH < 5, but essentially controlled by
surface reaction at high pH.
Data from the free drift experiments showed that the backward rate Rb
depends linearly on the activity product of Ca2+ and HCO−
3 (Rb was calcu-
lated as the difference between the forward rate Rf and the observed rate R)
and hence could be described by
where k4 (dimensions mol cm−2 s−1 M−2 ) was found to be a function not
only of temperature, but also of PCO2 . From equations (4.13) and (4.18) it
follows that the net rate R is given by
1
k1 AH+ = k2 AH2 CO∗3 + k3 AH2 O
k2 AH2 CO∗3 = k1 AH+ + k3 AH2 O
k3 AH2 O = k1 AH+ + k2 AH2 CO∗3
0.8
0.6 H2 CO∗3
PCO2
H+
(atm)
0.4
0.2
H2 O
0
3 4 5 6 7 8 9 10
pH
Figure 4.3: The rate of forward reaction is dominated by one of the terms
in equation (4.13) in the three regions labeled H+ , H2 CO∗3 , and H2 O. Along
the lines, one term just balances the other two, and in the area enclosed the
forward rate depends significantly on more than one mechanism. Adapted
from [Plu78].
73
and further assumed that reaction (4.20) is fast and therefore limited by
mass transport, while reactions (4.21) and (4.22) are much slower, and their
rates are chemically controlled. On basis of these assumptions, the surface
layer activities of H2 CO3 and H2 O are equal to the corresponding bulk fluid
activities, and the activities of all other species in the surface layer (including
H+ ) are determined by calcite and carbonate species equilibria. For instance,
(s) (s)
AH+ is given by the calcite saturation value at AH2 CO3 = AH2 CO3 and may
be calculated using an equilibrium model. A formal kinetic analysis of the
reactions (4.20)–(4.22) leads to a rate expression that is equivalent to the
empirical equation (4.19), and defines the rate constant k4 as
(s) (s)
k4 = k40 + k400 AHCO− + k4000 AOH− , (4.23)
3
(s)
where Ai are the surface layer activities of species i, and the new constants
k40 , k400 , and k4000 can be expressed in terms of the forward rate constants and
some equilibrium constants [Plu78, Plu79]. The agreement between the-
oretical and observed values of k4 was satisfactory over some range of CO2
partial pressures.
Plummer et al. had to make the crucial assumption that the partial
pressure of CO2 at the surface has the same value as in the bulk fluid (this is
(s)
equivalent to AH2 CO3 = AH2 CO3 since AH2 CO3 is proportional to AH2 CO∗3 which
in term depends linearly on PCO2 by Henry’s law)—an assumption that
74
may not hold under certain conditions. In an attempt to verify their model
by comparison with several other studies in the literature, Plummer et al.
themselves found discrepancies of more than one order of magnitude, espe-
cially under conditions of low PCO2 [Plu79]. More recently, the model has
been criticized by Compton and co-workers as conveying only limited mech-
anistic information, because the empirical rate expression (4.19) is based on
bulk fluid activities instead of surface layer activities, and the experimental
conditions do not allow a well-defined, calculable and controllable transport
of the reactants to the interface [Com90a, Com90b]. Although the model
seems to agree quite well with the original set of data, its applicability to
data obtained under different experimental conditions is strictly limited, in-
dicating that there must be a significant dependence upon specific properties
of the experimental method applied.
A completely different experimental approach was used in the study by
Compton et al. which has already been cited [Com90a, Com90b]. They
proposed a new strategy for studying surface reactions and applied their
sophisticated method to calcite dissolution: a calcite crystal forms part of
one wall of a rectangular flow cell, through which a standard HCl solution
is pumped under laminar flow conditions. Immediately downstream of the
crystal, an appropriate detector electrode is placed in the wall, measuring
either directly the concentration of Ca2+ or the remaining level of the reactant
H+ as a function of flow rate. This technique has several advantages, for
instance
Two different pH regimes were examined: At pH < 4, the data was described
best by first-order heterogenous kinetics according to the rate law
where R is the dissolution rate with units mol cm−2 s−1 , [H+ ](s) denotes the
surface concentration of H+ in M, and the rate constant was found to be
k1 = (4.3 ± 1.5) · 10−5 mol cm−2 s−1 M−1 . Compton et al. were able to
avoid transport control during all their experiments, down to very low values
of pH. Under high pH conditions, i.e., at pH > 7, and in the near absence of
CO2 , they chose the following rate law as that which best fits the observed
data
R = k − k 0 [Ca2+ ](s) [CO2− (s)
3 ] , (4.25)
where R has the same units as above and k 0 = k/Ksp with the solubility
product of calcium carbonate Ksp . Again, [Ca2+ ](s) and [CO2−
3 ]
(s)
are sur-
face concentrations, and the zero order rate constant k is dependent upon
surface morphology (especially roughness). The value found for a surface of
Iceland spar, polished with a succession of grit sizes down to 0.25 µm, was
k = 9.5 · 10−11 mol cm−2 s−1 . It can be shown that in the absence of CO2
and at pH > 7 Plummer’s rate expression (4.19) reduces to the same ki-
netic form as (4.25) under the assumption that dissolution ceases when the
calcite solubility product is reached [Com90b]. However, equation (4.25)
76
refers to surface concentrations rather than bulk phase quantities, and one
may conclude that it is more physically significant, i.e., closer related to the
“real” mechanism of surface reaction. On the other hand, model equations
like (4.19) are of great practical value, since in most practical cases the sur-
face concentrations can neither be measured nor calculated, and it is not
possible to eliminate the influence of transport processes.
Still another problem arises from the effect of impurities: it is known
that trace amounts of inhibitors (such as phosphate) can cause large changes
in rate without significantly altering the thermodynamic properties of the
system [Ber74, Plu79]. Unfortunately, there is no simple way of predicting
the total effect that an impurity or combination of impurities may have on the
rate of reaction, since theoretical models are not available to date. This shows
the difficulty in developing a useful general model of calcite dissolution—
very often a specific empirical relation for the system in question might be
preferred.
steady state. The output solution was collected and analyzed for dissolved
species, and also the pH was measured. This allowed to determine the disso-
lution rate by multiplying the concentration of cations with the flow rate of
the output solution and dividing by the estimated surface area of the solid.
The partial pressure of CO2 could not be fixed at some arbitrary value like
in the experiments above, but it could be calculated from the measured data
(bulk fluid concentrations and pH). The experiments showed that both calcite
and magnesite have the same general rate dependence on pH, but the rate
of dissolution of magnesite is about three to four orders of magnitude lower
than that of calcite, see Figure 4.4. The authors proposed a reaction model
similar to that of Plummer et al. to describe their data on the dissolution
of one-component carbonates,
where M represents either Ca or Mg. From these three reaction steps they
derived the rate expression
where the contribution of the first two reactions to the backward rate has
already been omitted since it is not significant. The rate constants ki were
determined from the experimental results; their numerical values are given
in Table 4.1 for both calcite and magnesite.
For dolomite as a two-component carbonate, Chou et al. found the fol-
lowing empirical rate equation,
−7
calcite ×
magnesite
dolomite 3
−8
×
3
××
×
−9 3
×
×
3
××
−10 ××
×× ×
3 ×
××
×
log R−11 ×
33
3 ××
3
33 3
3 33
33
−12 3
−13
−14
−15
2 4 6 8 10
pH
Figure 4.4: Dissolution rates vs. pH for calcite, dolomite, and magnesite;
experimental data obtained with continuous fluidized bed reactor at 25◦ C.
Adapted from [Cho89].
79
Dolomite
which only describes the forward rate. After a very short initial stage of
dissolution, calcium and magnesium are released stoichiometrically, i.e., in
a 1:1 mole ratio.1 An important difference in behaviour compared to the
other carbonates is that the reaction order n is a fractional number less than
one. The best fit of experimental data was obtained for n = 0.75 and the rate
constants given in Table 4.2. There is not yet a really convincing explanation
for the fractional order of reaction—Chou et al. assume that the formation
1
Since initially Ca is released faster than Mg, the dolomite surface acquires a Ca:Mg
mole ratio that is smaller than that of the bulk solid. This affects only a few atomic layers
on the surface [Bus82].
80
Rb = k4 AHCO−3 , (4.31)
and they were able to show that this backward rate is significant even far
from equilibrium. This is one important reason for the very low net rate
of dissolution, particularly at pH > 6 as H+ attack becomes less dominant.
The dissolution reaction essentially stops quite far from equilibrium, and
the amount of time necessary to reach “true” equilibrium might be so large,
that this will never be observed in laboratory or nature. Rate constants for
the temperatures 25◦ C and 55◦ C can be found in Table 4.2, but it must
be pointed out that these constants are highly dependent on the source of
dolomite. Busenberg and Plummer studied eight different dolomite speci-
mens of both sedimentary and hydrothermal origin and found that the latter
dissolved significantly slower at least in the low temperature regime. The
rate constants given here refer to a microcrystalline sedimentary dolomite
from Wisconsin. There is also a significant increase in dissolution rate with
increasing temperature: dolomite dissolves five to ten times faster at 45–60◦ C
than at 25◦ C.
81
The use of crushed or powdered limestone to neutralize acidic wastes like coal
mine drainage and pickling liquors has been investigated by several workers
since limestone is the least expensive neutralizing agent available [Bar76,
Hoa45, Par65, Pea75]. There have been some difficulties in the practical
application of limestone neutralization, because
• the active limestone surface may be fouled by oil, grease, or sulfate and
phosphate precipitates, thereby disrupting dissolution,
Out of these reasons, lime is more commonly used for the treatment of acidic
wastes. Nevertheless, limestone neutralization has been successfully applied,
especially in cases where the waste water does not contain too much sulfuric
acid. Parsons described design and installation of an upflow limestone
neutralization plant; after some simple modifications like increased flowrates
in the limestone bed and removal of fouling substances from the input stream,
satisfactory operation of the system was achieved [Par65].
Hoak et al. proposed a combined limestone-lime treatment of spent pick-
ling liquors: pulverized high-calcium limestone is used to raise the pH to
about pH 6 and precipitate part of the metals, then the treatment is com-
pleted with lime. Dolomitic limestones are practically useless for this purpose
due to their very low reactivity. As a rule of thumb, the rate of reaction with
pickle liquor was found to be inversely proportional to the percentage of mag-
82
nesium carbonate in the limestone, e.g., a limestone with 5 wt% MgCO3 has
only about 1/5 of the normal reaction rate [Hoa45].
Barton and Vatanatham presented a simple model for the kinetics
of limestone neutralization [Bar76]. They examined the dissolution rate of
calcite particles in 0.01 N sulfuric acid; the pH was allowed to drift freely.
Particles with almost uniform initial size were used, and a shrinking particle
model was applied to describe the change in surface area. All the reactions
in the bulk phase were assumed to be at equilibrium, and for the dissolution
rate the following expression—valid in the range between pH 2 and 6—was
found which is first order with respect to H+ concentration2
dN
− = Ak ([H+ ] − [H+ ]eq )
dt
1/3
6kM N0
= N 2/3 ([H+ ] − [H+ ]eq ) . (4.32)
ρD0
In this equation,
in the given pH range. Since their dynamic model is not valid at pH > 6, it
cannot be applied to the production of CMA at near-neutral pH. The same
is true for the model proposed by Pearson et al., who examined the use
of crushed limestone for the treatment of coal mine drainage and derived
practical rules for the design of such neutralization processes [Pea75].
The only study that investigated the dissolution of calcite and dolomite
in acetic acid was done by Suarez and Wood in the context of soil analysis
[Sua84]. They measured the buildup of CO2 -pressure in a sealed reaction
vessel in order to determine calcite surface area and content in soil sam-
ples (surface area and total calcite were determined from maximum slope
of pressure vs. time and final pressure, respectively). A buffered solution
containing sodium acetate and acetic acid was used in these experiments;
therefore, the pH remained nearly constant at about pH 4.3 during the reac-
tion. Under these conditions, calcite dissolves approximately 70 times faster
than dolomite. This is but the only useful result since the publication con-
tains neither absolute rate data nor a theoretical model.
Regime II: Transport control is found also in the extreme cases where the
dissolution rate is very high, or the concentration of a reactant is very
low. The first case occurs at rates greater than 5 · 10−8 mol cm−2 s−1 ,
e.g., with CaO dissolving at pH 1–2, and the diffusion of dissolved
species away from the surface is rate limiting. The second case can be
observed with ionic oxides at pH > 6; then the rate is controlled by
arrival of H+ .
86
Regime III: In the intermediate pH regime (pH 2–6), the dissolution rate
is mainly reaction controlled, and it typically ranges between 10−9 and
10−12 mol cm−2 s−1 . The activation energy in this regime is usually
greater than 65 kJ mol−1 (compared to only 15–20 kJ mol−1 for dif-
fusion control). However, the results of MacDonald et al., who in-
vestigated MgO dissolution in dilute sulfuric acid at pH 3–5 using a
rotating disc technique, demonstrate the existence of mixed kinetic
control, with transport control playing a more and more dominant role
at higher temperatures [Mac71].
These different kinetic regimes combined with the necessary distinction be-
tween initial and advanced stages of dissolution clearly complicated the in-
terpretation of experimental results, and, in fact, the available literature is
not very conclusive. There are several conflicting results reported for the dis-
solution of MgO, and it is hard to find anything at all about CaO dissolution
kinetics, since in this case experiments are further complicated by the very
high reactivity of CaO (which is about one order of magnitude higher than
that of MgO).
Briefly, the main results can be stated as follows: During the initial stage
of dissolution, both CaO and MgO show a significant rise in rate despite
decreasing pH under free drift conditions; this initial rise is steeper for CaO
than for MgO. The effect has been attributed to surface roughening due to
initial solution attack [Seg88, Seg78], but is still not very well understood
and shall not be discussed in further detail here. For the advanced dissolu-
tion regime of MgO, most authors found a linear relationship between the
logarithm of rate R and pH, but both the slopes (ranging from −0.5 to −1)
and the absolute values differ greatly [Seg88]. Figure 4.5 gives a good im-
87
pression of the variation in published rates for MgO at 20–25◦ C for pH 2–6.
The advanced kinetics of CaO are reported to show similar features to MgO
but at rates about ten times greater [Seg88].
At the moment there seems to be no way to explain these apparent exper-
imental contradictions. Some of the difficulties may arise from uncertainties
in estimating the effective surface area (by gas absorption or geometrical
considerations), changes in particle size distribution during dissolution (es-
pecially in the case of powders), and the use of background electrolytes in
some of the experiments. Vermilyea found that certain substances act as
inhibitors (e.g. periodate) or accelerators (e.g. phosphate, acetate3 ) in the dis-
solution of magnesium hydroxide [Ver69]. Another possible effect that may
occur in technical applications of calcium hydroxide as neutralizing agent is
the precipitation of calcium sulfate (gypsum) on the surface of the calcium
hydroxide particles, if sulfuric acid is present in the solution. Even a very
thin film of gypsum has been shown to significantly decrease the dissolution
rate [Bog85].
From these results one must conclude that it is even more difficult to
develop a kinetic model for the dissolution of CaO and MgO than it was
for the carbonates. As already stated at the beginning of this section, it is
therefore necessary to rely on rather crude estimates for the dissolution rates
in the interesting pH range, and to verify these estimates through subsequent
experiments under “real” process conditions. Reasonable extrapolated values
for the rates at pH 6 would be 1 · 10−9 and 1 · 10−10 mol cm−2 s−1 for CaO
and MgO, respectively. These values may be compared to similar estimates
3
This is particularly interesting, since both acetate and phosphate would be present in
a CMA fermentation.
88
−7
MgO smoke crystals at 25◦ C
sized, sieved powder at 20◦ C
rotating disk at 20◦ C
rotating disk at 20◦ C
−8
−9
log R
−10
−11
−12
1 2 3 4 5 6 7
pH
Figure 4.5: Summary of published dissolution rates vs. pH for MgO, in units
mol cm−2 s−1 . The data was collected from various sources by Segall et
al.; diagram adapted from [Seg88].
89
for the carbonate minerals. Table 4.3 gives a summary of dissolution rates at
pH 5, 6, and 7 for all the minerals discussed. The rates cover a range of almost
four orders of magnitude from the slowest dissolving mineral (magnesite) to
the fastest (CaO). It might be interesting to observe that the rates of calcite
and MgO are relatively high and of the same order of magnitude. Since such
a mixture can be easily prepared from dolomite by selective calcination, it
could have some potential as neutralizing agent in CMA production.
1. The liquid phase comprises water and all dissolved species, but not
the suspended cells and particles. It may be characterized by the vol-
ume V and the total concentrations of dissolved substrate, calcium,
91
gas discharge
µ
gas phase
(PCO2 )
6
CO2
µ I
HAc S Ca, Mg, CO2
ª R
cells solids phase
(µ, Nx ) (Ni , Āi )
µ
R R
cell loss solids feed solids discharge
Figure 4.6: Overall structure of the model system with definition of phases,
streams, and relevant dynamic variables.
92
• The reactions involving solid species are omitted because solids are not
considered to be part of the equilibrium model any more (there is no
equilibrium between liquid and solid phase). This reduces the number
of species as well as the number of independent reactions by two.
Table 4.4: Tableau-representation of the equilibrium model used for the liquid
phase. Equilibrium constants at 25◦ C and zero ionic strength.
H+ 1
OH− −1 −14.0
HAc 1
Ac− −1 1 −4.76
CO2 (g) 1
H2 CO∗3 1 −1.5
HCO− 3 −1 1 −7.8
CO2−
3 −2 1 −18.1
Ca2+ 1
CaOH+ −1 1 −12.85
CaAc+ −1 1 1 −3.56
CaHCO+ 3 −1 1 1 −6.54
CaCO3 −2 1 1 −14.9
Mg2+ 1
MgOH+ −1 1 −11.4
MgAc+ −1 1 1 −3.46
MgHCO+ 3 −1 1 1 −6.79
MgCO3 −2 1 1 −14.7
T
HAc 1 CAc
T
CaCO3 (s)/Ca(OH)2 (s) −2 1 1/0 CCa
T
MgCO3 (s)/Mg(OH)2 (s) −2 1 1/0 CMg
CO2 (g) 1 ?
97
I 1/2
à !
ln γi = −Azi2 − bI , (4.47)
1 + I 1/2
1X 2
I= z Ci , (4.48)
2 i i
in the way it was discussed in Section 4.1 (see equations (3.4) and (3.6) for
further explanation). The pH is defined as the negative logarithm of H+ -
activity, i.e., it can be calculated using
2−
TOT H = [H+ ] − [OH− ] − [Ac− ] − [HCO−
3 ] − 2[CO3 ]
Dissolution of Solids
T T
Differential equations for the remaining dynamic variables CCa and CMg can
be found using the results on dissolution kinetics from the beginning of this
chapter. First, dolomite shall be considered as neutralizing agent—the cor-
responding liquid phase mole balances for calcium and magnesium yield
T
dCCa 1 ³ T
´
= Ādol Rdol − F CCa (4.57)
dt V
T
dCMg 1 ³ T
´
= Ādol Rdol − F CMg , (4.58)
dt V
101
where Ādol denotes the effective surface area of dolomite and Rdol is obtained
by combining equations (4.30) and (4.31) to
Since calcium and magnesium are released at the same rate Rdol after a very
short initial stage of dissolution, one should expect that
T T
CCa = CMg (4.60)
dNdol
= (feed)dol − Ādol Rdol − (discharge)dol . (4.61)
dt
Much more difficult to quantify is the effective surface area Ādol since it
depends on the particle size distribution which in the most general case
would have to be modeled using a population balance approach. The size of
each individual particle depends on its initial size, on the time it entered the
system, and on the conditions it met during its lifetime, i.e., on its entire
“history”. A simple shrinking sphere model like the one used by Barton
and Vatanatham [Bar76] is certainly not adequate because it is based on
the assumption that all particles start dissolving at the same time, and there
is no continuous feed of “fresh” particles. Population balance models have
been applied to processes like crystallization, but they involve rather complex
mathematics, hence using such a sophisticated approach is not justified in our
case considering the other serious limitations of our model. However, under
steady state conditions, the effective surface area as well as the amount of
102
2/3
Ādol = cNdol , (4.62)
where again Āi denotes the effective surface area of solid species i. The
carbonate rates may be derived from equation (4.29)
using the proper sets of rate constants ki for CaCO3 and MgCO3 , respec-
tively. The hydroxide rates can be estimated from the linear log R vs. pH
relationship discussed in Section 4.2.2 which leads to expressions of the form
where the constants c1 and c2 can be determined for Ca(OH)2 and Mg(OH)2
from the data presented in Table 4.3. Of course, the simple rate expres-
sions (4.67) are only very crude approximations, and more accurate relations
would definitely be desirable. The effective surface area Āi and the total
103
amount Ni of each solid species i behave in the same way as it was explained
above for dolomite, so no further discussion is necessary here. Ideally, ĀCaCO3
and ĀMgCO3 are zero for type S dolime, and ĀCa(OH)2 and ĀMgCO3 are zero
for selectively calcined dolomite, but this is only true as long as there is
no precipitation of carbonates or direct transformation of solid hydroxides
into carbonates. Especially in case of selectively calcined dolomite, the CO2
generated by dissolution of CaCO3 might react with Mg(OH)2 to form some
MgCO3 . Furthermore, it is unlikely that the solid rawmaterials are “pure”,
i.e., selectively calcined dolomite will almost certainly contain Ca(OH) 2 and
MgCO3 , while type S dolime might contain some CaCO3 and MgCO3 , al-
though in this case the residual carbonates can be effectively eliminated by
grit removal after slaking. As can be seen from the rate expressions, now
generally
T T
CCa 6= CMg , (4.68)
1. The underlying empirical expressions for the dissolution rates are not
necessarily adequate for the chemical and hydromechanical conditions
encountered in the continuous bioreactor.
effective surface area Āi of the various solids, because this would have
been too much effort considering the limited reliability of other model
parts.
Out of these reasons, the results obtained from the model should not be
expected to carry too much information about the real process, even if con-
ditions of steady state are assumed.
where the vector x contains all the dynamic variables of the model, while
y consists of the remaining algebraic variables. DAEs arise naturally in
many applications, and their study is an active field in applied mathematics.
Numerical methods to solve DAEs directly—i.e., without transformation into
a system of ordinary differential eqations (ODEs)—have been developed since
the early 1970s, and reliable algorithms exist. DAEs present numerical and
analytical difficulties which do not occur with ODEs, for instance the initial
values must be consistent with the algebraic constraints (4.70). An excellent
review of recent developments in theory and numerical methods for DAEs is
provided by Brenan et al. [Bre89].
105
• the assumption that hydroxides are not directly transformed into car-
bonates by CO2 .
Nevertheless, the model was helpful to gain some insight into the dynamic
structure of the process and to determine important variables and their in-
terdependencies. The two main implications for practical applications are:
106
2. It will be much more difficult to maintain the proposed 1:1 mole ra-
tio of Ca:Mg (or any other fixed mole ratio, if desired). A solution to
this problem will have to be found especially for type S dolime, since
Ca(OH)2 dissolves about ten times as fast as Mg(OH)2 . For selectively
calcined dolomite, the dissolution rates derived in Section 4.2.2 seem
to indicate that CaCO3 and Mg(OH)2 dissolve at rates of at least the
same order of magnitude under process conditions (see Table 4.3). In
case of dolomite only the initial dissolution stage could cause difficul-
ties, because ultimately calcium and magnesium should be released
stoichiometrically. If very fine particles are used in order to achieve a
large surface area, this initial non-stoichiometric release could actually
have a significant effect on product composition.
Various fermentation processes for CMA production have been proposed, but
none of them has been commercialized or even tested at a pilot plant level
yet. This is kind of surprising since the authors have all claimed signifi-
cant cost advantages for their processes compared to the conventional route
which is currently used for small scale production. A real breakthrough in
the commercialization of CMA just did not happen, probably for following
reasons:
107
108
tivity and higher final product concentrations. People tend to wait for
these possible improvements rather than to take the risk of an early
investment.
On the other hand, a case can be made also for rather small alternative CMA
plants, at least with respect to the next five to ten years, since it is not very
likely that large plants will be built during this time. There definitely is a
small market for CMA at prices of $ 500 per ton or even higher—for instance,
it can be sold to the State Highway Administrations for use on bridges, and
to environmentally concerned private consumers who would probably like to
use it on their sidewalks and driveways—and if a small scale fermentation
process can actually do better than the conventional route, it should really
be worth trying. One possibility to initiate such a small scale production of
CMA would be to establish CMA as a new byproduct of the pulp and paper
industry, made from unpurified waste sugars which otherwise would just be
burned. Quite much research has been done on new organosolv wood pulping
methods recently, and one of the most promising processes of this kind uses
concentrated acetic acid as cooking liquor. From the spent liquor of this
acetic acid pulping process an aqueous solution containing xylose, glucose,
and some acetic acid could be obtained after recovering other byproducts
and most of the acetic acid—a “waste” stream that would be ideally suited
for fermentation to CMA. The preliminary design for a CMA plant based on
109
this feedstock will be presented at the end of the chapter together with an
economic analysis.
None of the previously published CMA fermentation processes offered a
solution to the mole ratio problem which provided the starting point for this
study. In the next section, it will be shown that an amazingly simple solution
exists. First, the key features of a continuous process capable of producing
the desired mole ratio of Ca:Mg will be discussed, then the experimental
realization of such a process will be described, and results will be discussed
for three different types of raw materials (type S dolime, selectively calcined
dolomite, and dolomite). Finally, the potential of the new process as a whole
and of each of the raw materials will be assessed.
– there is still not very much experience with large scale applications
of membrane filters.
However, for the experiments that will be described in the next section,
a cross flow filter was used since it could be obtained more easily and
also provided simpler operation than a continuous centrifuge.
There might be two reasons for not recycling all solids—to prevent the
112
buildup of insolubles like silica which are always present in lime and lime-
stone, and to influence product composition, if the desired mole ratio does
not match the composition of the feed solids. This can be done by discharg-
ing solids directly with the product stream (a certain amount of solids in
the product could actually improve the deicing properties) or by splitting
the retentate stream and thus sending only part of it back to the fermentor
(again the solids could be added to the deicer product, and it would actually
be preferable to do that after the evaporation step in order to avoid scaling
problems in the evaporator system). In practice, it would be most desirable
to specifically separate out unwanted components from the recycle stream
rather than simply split the whole stream into two substreams of the same
composition. It is not unlikely that such a specific separation can be achieved
since there will probably be sufficient differences in size and density among
the solid components.
How will the three different kinds of solid base perform in the proposed
new process? Ultimately, this question can be answered only based on exper-
imental results, but a preliminary assessment of these raw materials can be
made by considering their general properties and the rate data from Chap-
ter 4:
1. In type S dolime, high dissolution rates for both minerals Ca(OH)2 and
Mg(OH)2 are combined with a very large effective surface area due to
small particles (most of them in the range 2–5 µm) of porous structure.
Therefore, a high overall reactivity and only a relatively small accumu-
lation of solids is expected. Although the dissolution rate of Ca(OH) 2 is
about one order of magnitude higher than that of Mg(OH)2 , a station-
ary state should be established quickly, because Mg(OH)2 still dissolves
113
2. Selectively calcined dolomite will react more slowly, but the dissolution
rates for both CaCO3 and MgO should still be relatively high and of
the same order of magnitude. The material has to be ground to obtain
a powder because it does not decompose automatically when mixed
with water. Even in finely pulverized form it will probably have a
smaller effective surface area than type S dolime due to the less porous
structure. Although the MgO should be only light-burned and reactive,
it is difficult to predict whether the step
MgO + H2 O −→ Mg(OH)2
is rate determining or not, i.e., whether the MgO will dissolve at es-
sentially the same rate as the Mg(OH)2 from type S dolime or rather
more slowly. From the rate data in Table 4.3 it would be expected that
the CaCO3 dissolves slightly faster than the MgO, but the difference
(factor two) is small compared to possible errors in the data, and the
opposite might be true. As a matter of fact, if MgO would turn out
to be the faster dissolving component, this would be a very interest-
ing property of the material, since it would allow the production of a
high magnesium CMA by discharging part of the undissolved solids.1
Compared to type S dolime, the necessary accumulation of solids in
the system will be higher, but it is not expected to be excessive. Since
1
As discussed in Section 1.2, there exists a CMA double salt with a Ca:Mg mole ratio
between 2:3 and 3:7 which may have superior deicing properties compared to the 1:1 CMA.
114
the dissolution rates of CaCO3 and MgO should not differ too much,
the steady state is supposed to be established after a relatively short
period of time.
Unfortunately, the temperature controller failed during the first hour of op-
eration, so that all experiments had to be done at 26–31◦ C (without tem-
perature control), since time constraints did not allow waiting for repair or
replacement of the broken unit. A schematic diagram of the experimental
system is given in Figure 5.1; some less important parts (temperature control,
volume control, stirrer, and fraction collector) are not shown.2 At the begin-
ning of each experiment, the dissolver (“fermentor”) was filled with V = 1.5 l
deionized water. The recirculation loop (pump P2 ) was operated at relatively
high flow rates of 0.27–0.43 l min−1 in an attempt to prevent the buildup
2
For a complete list of equipment, see Appendix C.
116
CO2
6
P1 - ¾
P4 neutralizing
HAc feed agent
(40 g l−1 )
6
(slurry)
dissolver
(“fermentor”)
- pH control
unit
¾ ¾
6
P2 - cross flow -
? filter
P3 - CMA product
? (≈ 50 g l−1 )
Figure 5.1: Schematic diagram of the experimental system used for continu-
ous production of CMA under steady state conditions.
117
of a filter cake in the cross flow filter, and the contents of the dissolver was
well mixed by stirring at 800–1000 rpm. The feed and the product stream
had exactly the same flow rate F = 0.15 l h−1 ; this was accomplished by
using one pump drive with two identical peristaltic pump heads P1 and P3 .
Nevertheless, a volume control was required, because some additional water
was introduced with the neutralizer slurry. Whenever the liquid level in the
dissolver became too high, excess fluid was removed automatically by a pump
operating parallel to pump P3 (not shown in the diagram). Without these
additional streams the dilution rate would have been D = F/V = 0.1 h−1 ;
the actual value was slightly higher. The pH was controlled at a setpoint
of about pH 6 using a rather unsophisticated on/off-controller which actu-
ated the neutralizer feed pump P4 . A high speed stirrer was used for the
slurry storage vessel to prevent settling of the particles. The system was
operated at atmospheric pressure; several openings in the headplate of the
dissolver provided a means for the discharge of CO2 generated by carbonate
dissolution.
The feed stream contained 40 g l−1 acetic acid, thus simulating an acetic
acid productivity of 4 g l−1 h−1 in the “fermentor”. Under conditions of
steady state, this translates into approximately 50 g l−1 CMA in the prod-
uct stream, if the dilution by the slurry feed can be neglected. The total
concentration of acetate ion in the dissolver CAc can be described by
f
dCAc /dt = D(CAc − CAc )
(5.1)
CAc (0) = 0,
f
where CAc is the feed concentration of acetate ion, and D = F/V the dilution
rate. Equation (5.1) has the solution
f
CAc (t) = CAc (1 − e−Dt ), (5.2)
118
f
i.e., the steady state acetate concentration CAc will be approached very closely
after 40–50 h of operation. Therefore, a duration of 50 h was chosen for the
experiments, and it was hoped that this amount of time would also allow the
solids phase to get close to steady state.
Hourly samples of the solids-free product solution were taken with an
automatic fraction collector; these samples were later analyzed for calcium,
magnesium, and several other dissolved ions by inductively coupled plasma
emission spectroscopy (ICP). From the analyses, the total concentration of
CMA in the product solution and the Ca:Mg mole ratio could be determined
as functions of time.
The cross flow filter with pore size 1.2 µm performed well in retaining even
the finest lime particles—a perfectly clear product solution was obtained.
However, there were problems with the accumulation of a filter cake in all
experiments, and it was not possible to prevent this by increasing the flow
rate in the recirculation loop (perhaps it just could not be increased enough).
Shaking of the filter capsule from time to time allowed recirculation of part
of the accumulated solids, but new cake formed quickly. The amount of
solids “trapped” in the cake (and hence essentially removed from the active,
dissolving part of the solids phase) was significant compared to the total
amount of solids present in the system. As will be seen from the discussion
of results, it was mainly this effect of “quasi solids removal” which made it
difficult to approach the steady state. Another problem that occurred only
in the experiment with selectively calcined dolomite was the clogging of the
neutralizer feed mechanism due to fast settling of the slurry. One time the
neutralizer feed was cut off for several hours, which produced quite interesting
effects, but of course seriously disturbed the course of the experiment from
119
the way it was planned. Time constraints made it impossible to repeat the
experiment with an improved system; therefore, the presumed undisturbed
behaviour had to be extrapolated from the available data.
Discussion of Results
The first experiment was done with type S dolime supplied by The Western
Lime & Cement Co. in West Bend, Wisconsin; it had a Ca:Mg mole ratio
of 1.04:1.3 A slurry was prepared from one part dry lime and two parts deion-
ized water. The total amount of dolime consumed during 50 h of operation
was about 170 g; therefore, approximately 340 g of water were introduced
with the slurry—this amounts to less than 5% of the total feed of water,
thus increasing the dilution rate only slightly. As already mentioned, the
temperature controller failed during the first hour after start up, and it was
decided to continue the experiment without temperature control. The tem-
perature fell to the range of 26–31◦ C, about 5 K above ambient temperature
due to the heat generated by stirrer and pumps. During the first 10 min
of the experiment, the pH varied between pH 4.87 and 9.81, i.e., there was
some overshooting each time when dolime was added, but after 25 min it
had already stabilized in the range of pH 6.00–6.11. As more solids built
up in the system and the total acetate concentration increased, the range
narrowed further so that an essentially constant value of pH 6.02 could be
maintained. Starting after a few hours of operation, a brownish, gel-like filter
cake accumulated in the filter capsule which could not be removed during
the experiment. Backflushing at high flowrates had to be applied afterwards
in order to clean the filter for the next experiment.
3
Analyses of this material and the mole ratio calculation can be found in Appendix B.1.
120
2.5
CMA from type S dolime 3
2 3
Ca:Mg 1.5
3
mole 333 33 3 3 3 3
3 3 3 3
ratio 1 33
0.5
0
0 5 10 15 20 25 30 35 40 45 50
time (h)
Figure 5.2: Mole ratio vs. time for CMA produced from type S dolime;
determined from ICP analyses of the product samples.
The Ca:Mg mole ratio in the product was calculated for some of the
hourly samples; the results are shown in Figure 5.2.4 As expected, the
calcium hydroxide turned out to dissolve much faster than the magnesium
hydroxide—at the beginning the Ca:Mg mole ratio in the product was as high
as 2:1. Thereafter, magnesium hydroxide started to build up in the dissolving
solids phase, and the mole ratio decreased rapidly. After only seven hours
of operation, the theoretical value 1.04:1 was reached, although there had
already been some minor accumulation of filter cake which caused the curve
to turn up slightly between hour 3 and 5. Obviously, serious cake accumula-
tion started just before hour 8, because there suddenly was a jump towards
4
Original results of the sample analyses are given in Appendix A.2.
121
higher calcium content, and for the rest of the experiment the mole ratio
remained in the range between 1.20:1 and 1.35:1. The filter cake contained
high amounts of magnesium, alumina, and iron (which was responsible for
the brown color). The amount of solids that made up the solids phase in the
dissolver was very small (probably less than 5 g l−1 ) due to the fast reaction
of both minerals. At the end of the experiment, the concentration of CMA
in the product stream was about 45 g l−1 . Dry CMA was recovered from 2 l
product solution collected during the last 13 h of operation.
In the second experiment, selectively calcined dolomite was used as neu-
tralizer. The uncalcined stone was again supplied by The Western Lime
& Cement Co.; it had a Ca:Mg mole ratio of 1.03:1, i.e., it was essentially
pure dolomite.5 Samples of this stone were calcined for half an hour at
about 700◦ C in an electrical furnace. Ideally, only the magnesium part of
the dolomite should be decomposed under these conditions according to the
reaction
CaMg(CO3 )2 −→ CaCO3 + MgO + CO2 ↑, (5.3)
and the theoretical weight loss can be calculated from the amount of CO2 re-
leased. For the selective calcination of pure dolomite, the theoretical weight
loss is 23.9%, and if all carbonate is decomposed (full calcination), the weight
loss would be 47.7% instead. The actual weight losses of the calcined sam-
ples used in this experiment ranged from 17.7% to 23.6%, indicating that
there remained a small portion of unreacted dolomite. Before the slurry was
prepared, all samples were mixed and pulverized; the powder was screened
using a 35 mesh sieve. A total amount of 377.4 g dry powder was obtained.
The first batch of slurry was prepared from one part selectively calcined ma-
5
A chemical analysis can be found in Appendix B.2.
122
terial and two parts deionized water. No observable reaction occurred when
the water was added—this confirmed that the material was practically free
of CaO which would have been hydrated in a fast and highly exothermal
reaction. Unfortunately, in this experiment the slurry feed mechanism did
not work as well as it had done with the type S dolime because the solids
tended to settle in the tubing. In an attempt to prevent clogging, an ele-
vated neutralizer storage vessel was used, and the slurry was diluted with
more water. Thus a relatively high amount of about 1.18 l water was added
to the system with the slurry, corresponding to almost 14% of the total feed
of water (this caused a significant increase of the overall dilution rate, and a
more dilute product solution was obtained). But still the clogging problem
was not solved completely, and one time the neutralizer feed was blocked for
several hours before the failure was discovered. Definitely, it would be a bet-
ter idea to feed the dry powder or a putty instead of the slurry using a screw
conveyor or similar piece of equipment when repeating the experiment. The
theoretical amount of selectively calcined dolomite which had to be dissolved
during 50 h of operation was 175.5 g, but the actual amount fed into the
system was higher, since much more undissolved solids accumulated than in
the first experiment due to the lower reactivity of the material. Right after
the experiment had been started, the pH fell from about pH 7 to 5.07, and
then started rising slowly, as slurry was added continuously. About 10 min
later the setpoint of pH 6.00 was reached. Enough solids had built up in
the dissolver to maintain the required overall neutralization rate at around
pH 6, and the neutralizer feed became discontinuous. There was no signif-
icant overshooting like with type S dolime, the pH just varied in the range
of pH 5.96–6.03. Within a few hours of operation this range narrowed fur-
123
2.5
CMA from selectively calcined dolomite 2
2
Ca:Mg 1.5 2
mole
ratio 1 2
2 2
2 2 2
2 2
0.5 2222 2
222222 2 2 2 2
0
0 5 10 15 20 25 30 35 40 45 50
time (h)
Figure 5.3: Mole ratio vs. time for CMA produced from selectively calcined
dolomite; determined from ICP analyses of the product samples.
ther due to the buffering effect of the increasing acetate concentration. The
problem with the accumulation of filter cake started even earlier than in the
first experiment, although the cake was less “sticky” and could be removed
more easily during operation. Again, some nonideality in the behaviour of
the system had to be expected.
Figure 5.3 shows the Ca:Mg mole ratio for the product during the course
of the experiment.6 The two most important observations are:
1. The magnesium oxide clearly dissolved faster than the calcium carbon-
ate at pH 6, in contradiction to the rate estimates from Section 4.2.2
(Table 4.3). After eleven hours of operation the Ca:Mg mole ratio was
6
See Appendix A.2 for the original results of the analyses.
124
mole ratio below the start up level. Therefore, “quasi solids removal”
might be responsible for delaying the upturn towards steady state after
hour 11, but definitely not for the observed minimum itself. A likely
explanation for this strange behaviour is that the dissolution rate of
the magnesium oxide significantly increased during the first hours of
operation. As briefly mentioned in Section 4.2.2, acetate acts as an
accelerator in the dissolution of magnesium hydroxide, and the same
might be true for bicarbonate. The concentrations of both acetate and
and bicarbonate increased after start up,7 hence an accelerating effect
might very well occur. Since for selectively calcined dolomite the disso-
lution rates are of the same order of magnitude, even a relatively small
acceleration of magnesium oxide dissolution is observable which is not
necessarily true for type S dolime due to the much higher reactivity of
calcium hydroxide.
The chemical analysis of the solids phase showed that it contained calcium
and magnesium in a molar ratio of 3.07:1, and also significant amounts of
sulfur, alumina, and iron. The solids concentration in the dissolver was
approximately 45 g l−1 —higher than in the first experiment, but not unrea-
sonably high. A final CMA concentration of about 38 g l−1 was reached, and
again dry CMA was recovered from 2 l product solution.
In the third experiment, the same dolomitic limestone was used as in the
second experiment, but without the calcining. The stone was just pulverized
(less than 35 mesh), and a slurry was prepared from one part dolomite and
about three parts water. Some difficulties with a high accumulation of solids
7
The bicarbonate formed by dissolution of calcium carbonate is likely to accumulate
to a certain level due to slow carbon dioxide exsolution.
126
had been expected, but it was still surprising how soon the system had to
be shut down because of complete failure. Immediately after the experiment
had been started, the pH dropped to pH 4.70, then stabilized and started
rising slowly as limestone was continuously fed into the system. However,
about 20 min later there was already an excessive buildup of dolomite in
the filter capsule, and the pH had just reached pH 5.40—still far below the
setpoint. The solids concentration was so high that even shaking of the filter
capsule could not prevent further accumulation of cake. Only 35 min after
start up the cross flow filter got clogged, and the recirculation pump had to
be turned off. The acetic acid feed was stopped as well, only the pH con-
trol was left operating. It took another 10 min of continuous limestone feed,
before the setpoint pH was reached. Clearly, the uncalcined material was
too slow to maintain pH 6 at a reasonable concentration of solids under the
chosen process conditions. Perhaps the performance would have been better
at higher temperature (60◦ C) and for a smaller particle size. High solids con-
centrations could probably be handled using a more sophisticated equipment
(e.g., a centrifuge instead of the filter), but the other two materials which
can do the job at much lower accumulations would definitely be preferable
in the context of a fermentation process. An analysis of the product solution
collected during the first 35 min of operation showed that it had a Ca:Mg
mole ratio of 1.12:1; another sample taken at pH 8.14 after solids and liquid
in the dissolver had equilibrated for several hours had an even higher ratio of
1.50:1. These results confirm again the significance of the initial dissolution
stage which already had been found in the equilibrium experiment described
in Section 3.3.
Based on the results of these experiments, a better evaluation of the pro-
127
2.5
CMA from type S dolime 3
CMA from selectively calcined dolomite 2
2 3
Ca:Mg 1.5
mole 3
ratio
33
1 333333333333333333333
22222
2 2 2 2222
0.5 2 2 22
2 22
222222
0
0 5 10 15 20 25 30 35 40 45 50
time (h)
posed process and the tested raw materials could be made. It should be
possible to avoid the adverse effects of filter cake accumulation by using a
continuous centrifuge instead of the filter—such a system will be able to op-
erate under conditions of steady state for long periods of time. Although the
experimental system described above was far from ideal, it performed well
enough to confirm the feasibility of the new process. A more ideal system
would probably have shown a behaviour similar to that given in Figure 5.4;
the curves for both type S dolime and selectively calcined dolomite were
obtained from the available data by correcting for the effects of cake accu-
mulation and control system failure. The type S curve approaches the 1:1
steady state mole ratio from above, and the curve for the selectively calcined
material comes from below (it initially turns down due to acceleration of
128
magnesium oxide dissolution, and then turns up again because of the accu-
mulation of calcium carbonate in the solids phase). A curve for uncalcined
dolomite would probably look quite similar to the type S curve, if this ma-
terial could be successfully applied in spite of its low reactivity.
The potential of the three raw materials tested can now be assessed in
the following way:
In the next section, the proposed process will be incorporated in the prelim-
inary design of a fermentation plant for the production of CMA from wood
hydrolyzate, and the economics of such a plant will be briefly discussed.
Then a summary of the advantages of this new process compared to other
proposed processes will be given.
wood chips
?
acetic acid - - pulp
(from recovery) digestion
?
acetic acid,
evaporation - byproducts
and drying
(to recovery)
?
- lignin - sugar solution
water precipitation (to fermentation)
?
lignin
Figure 5.5: Simple block flow diagram of the acetic acid pulping process; the
solvent recovery system which is essential to make this process economically
viable is not shown in the diagram.
132
• The pulping process has a capacity of 100 t per day pulp. About 330 t
per day hardwood chips (40 wt% moisture) are required as feed, and
it is assumed that among other byproducts 50 t per day sugars can be
recovered for fermentation to CMA.
• Of the sugars fed to the fermentor, 95% are utilized by the organism,
and only 5% are lost with the product stream (and form part of the
deicer product). For 1 g of sugar consumed by the organism, 0.85 g
of acetic acid are produced, i.e., the yield on a weight basis is 85%.
Therefore, 0.95 × 0.85 × 50 t = 40 t per day acetic acid are produced;
that corresponds to 50 t per day pure CMA product.
• Again, a value of 5 g l−1 h−1 is used for the CMA production rate per
unit volume of the fermentor, and hence a fermentor volume of 416 m3
is required.
• Although the sugar feed stream might already contain some nutrients in
addition to the substrate itself (e.g. minerals from the wood), and others
might be supplied by the lime, it is most likely that at least additional
sources of nitrogen, sulfur, and phosphorus have to be provided in the
medium for growth of the organism. These nutrients are assumed to
be already added to the sugar feed stream.
Ca:Mg Ca:Mg
1:1.12 5:1 CO2 6.51
? - CaCO3 2.30
H2 O 1, 005.67 MgO 0.18
HAc 2.18
CaAc2 23.46
MgAc2 23.71 Ca(OH)2 1.35
? ?
neutralizer tank (pH 8–9)
Ca:Mg
1:1
?
H2 O 1, 006.32
CaAc2 26.34
MgAc2 23.71
Figure 5.6: Mass balance for the production of a 1:1 CMA from selectively
calcined dolomite; all figures are given in units of metric tons per day.
135
In both cases, 2.50 t per day sugars would also be contained in the
deicer. Therefore, the total amount of deicer material produced would
be 55.03 t per day in case 1 and 64.51 t per day in case 2 if all discharged
solids were added to the product.
136
Ca:Mg Ca:Mg
1:1.74 4:1 CO2 5.08
? - CaCO3 11.36
H2 O 1, 005.67 MgO 1.14
HAc 2.18
CaAc2 18.18
MgAc2 28.45 Ca(OH)2 1.35
? ?
neutralizer tank (pH 8–9)
Ca:Mg
2:3
?
H2 O 1, 006.32
CaAc2 21.06
MgAc2 28.45
Figure 5.7: Mass balance for the production of a 2:3 CMA from selectively
calcined dolomite; all figures are given in units of metric tons per day.
selectively
sugar carbon calcined
feed dioxide water dolomite
calcium
hydroxide
5 LP
4 1
steam
cond.
LP
steam
cond. dry CMA
product
Figure 5.8: Process flow diagram for small-scale CMA production from waste sugars and selectively
calcined dolomite.
137
138
A process flow diagram is given in Figure 5.8; the process flow through all
of the important pieces of equipment shall be briefly described at this point.
The fresh medium containing the sugars and additional nutrients is first
heated to 60◦ C and then enters the 416 m3 fermentor which is a continuous-
flow stirred-tank reactor (CSTR). A heating coil—or, alternatively, a steam
jacket—provide a means to maintain the process temperature inside the fer-
mentor. Selectively calcined dolomite is purchased in crushed form; it is first
pulverized using a grinder and then mixed with a very small amount of wa-
ter. The resulting putty is directly fed to the fermentor by a screw conveyor
which is actuated by the pH controller. Fermentation broth containing prod-
uct, undissolved lime solids, and cells is sent to a disc nozzle type centrifuge
capable of producing an overflow essentially free of solids and cells while dis-
charging these continuously along with a variable portion of the liquid phase
in the underflow. The solids concentration in the underflow does not have to
be exceptionally high, as a matter of fact, it would probably be preferable
to maintain a high flowrate at relatively low solids concentration to avoid
settling problems. A second, smaller centrifuge is used to separate out the
fraction of lime solids which has to be discharged; essentially all cells and
the remaining lime solids are recycled back to the fermentor with the over-
flow. The overflow of the first centrifuge is sent to a 40 m3 neutralizer tank,
where a small amount of calcium hydroxide is added in order to raise the
product pH to about pH 8–9. Complete reaction of the calcium base in the
neutralizer can be assumed, and the essentially solids-free product stream is
then sent to a multi-effect evaporator for concentration of the CMA to about
40 wt%. The available total temperature difference is constrained by the de-
composition temperature of CMA which is about 160◦ C. For instance, if the
139
evaporator was operated in the range between 60◦ C and 140◦ C, the number
of effects would be limited to five assuming a minimum approach temperature
of 16 K. The design could be further improved by using an adiabatic flash in
order to bring the temperature of the evaporator feed down to about 45◦ C
(this would widen the available temperature difference for the evaporator
and at the same time already remove some of the water), and the maxi-
mum number of effects could be increased by thermal or mechanical vapor
recompression. These possibilities shall not be considered in this preliminary
design, although they could significantly reduce the cost of evaporation. A
five-effect evaporator with a steam efficiency of 4.4 is assumed, i.e., each ton
of low pressure steam at 3.614 bar, 140◦ C evaporates 4.4 tons of water. The
concentrated CMA solution is then mixed with the solids discharged from
the second centrifuge and sent to a steam heated drum dryer/flaker, where
the dry deicer product is recovered in the form of coarse flakes.
At this point, the major design specifications shall be summarized, and
the results shall be given in both metric and U.S. units:
– 24.0 t (26.5 ton) per day and 32.1 t (35.4 ton) per day selectively
calcined dolomite as neutralizer for the production of 1:1 CMA
and 2:3 CMA, respectively,
– 1.35 t (1.49 ton) per day calcium hydroxide for the pH adjustment.
These specifications will be used in the next section as a basis for the eco-
nomic evaluation of the process.
• Carbon steel is used for the evaporator and drum dryer/flaker, but all
other major equipment items have to be made from stainless steel.
• Neither the waste acetic acid contained in the sugar feed stream nor
the additional amount of CMA product formed from this acid are con-
sidered in the calculations.
• The sugars are available at their fuel value, i.e. for about $ 25.00/t
($ 22.70/ton).
Table 5.1: Capital requirement for a small CMA plant with 64.5 t (71.1 ton)
per day output of 2:3 CMA deicer (including lime solids and sugars).
• Electric energy costs $ 0.06 per kWh; total power requirements for
stirrer, centrifuges, and pumps are estimated at 1,250 kW.
An estimate of the capital requirement for such a plant addition was made
based on the cost data provided by [Ulr84]; all prices were escalated to
1991 using the CE Plant Cost Index. Table 5.1 lists the prices found for the
major individual equipment items. These prices are bare module prices, i.e.,
installation is already included. Their sum has just to be multiplied by the
142
Table 5.2: Annual operating cost for the same CMA plant, assuming a
90% plant service factor, i.e., an output of 21,190 t (23,360 ton) per year
of 2:3 CMA deicer.
raw materials
sugar feedstock ($ 25/t) 410,600
other nutrients 317,800
selectively calcined dolomite, crushed ($ 50/t) 527,200
high-calcium hydrated lime ($ 60/t) 26,600
utilities
electricity ($ 0.06/kWh) 591,300
low pressure steam at 3.614 bar, 140◦ C ($ 5/t) 373,300
labor 500,000
fixed charges on capital 542,500
factor 1.18 to account for contingency and fee, yielding a total fixed capital
of $ 7,750,000.
Table 5.2 shows the annual operating costs for the production of 2:3 CMA
deicer, determined on the basis of a 90% plant service factor (21,290 t per
year output). If a 30% simple rate of return (ROR) before taxes is required
on the total fixed capital, the deicer product can be sold for $ 264/t or
$ 239/ton (at plant gate). This compares well to the cost estimates for other
fermentation processes, which range from about $ 260 to $500/ton, depending
on the feedstock and process scale. It also should be pointed out that this
low production cost could be achieved on a very small plant scale—all other
143
proposed CMA plants were much bigger; most of them were designed for a
capacity of 500–1,000 tons per day.
The new process for the production of a 2:3 CMA deicer from selectively
calcined dolomite has several advantages compared to previously proposed
processes (and even to the production of 1:1 CMA using the “same” process):
• Only inexpensive raw materials are used to make the 2:3 CMA which
is supposed to have superior deicing properties. In the conventional
Chevron process, expensive magnesia ($ 265–366/ton) has to be added
in order to achieve the 2:3 mole ratio, and together with the acetic acid
this contributes significantly to the high cost of the product (selling
price $ 657/ton).
• The solids discharge is roughly one third of the solids feed—this effec-
tively limits the undesired buildup of inerts in the solids phase.
Therefore, it may be concluded that the production of 2:3 CMA from se-
lectively calcined dolomite has the greatest potential and would definitely
justify further investigation.
Chapter 6
Analysis of Results
At this point, a brief summary of the most important results of this study
shall be given, and the major conclusions shall be discussed from the stand-
point of the original objectives of the research. This discussion will quite
naturally lead to the formulation of concrete recommendations for the prac-
tical application of some of the reported findings, and it will also lead to
specific suggestions for future work.
144
145
mole ratio), on whether or not it has been crystallized as a double salt, and
also on its physical form (coarseness and form of the flakes). From the avail-
able literature it was concluded that suitable compositions for a CMA deicer
would be either a mole ratio of 1:1 which most likely results in a physical
mixture of the two acetate salts, or some ratio between 3:7 and 4:6 which
may lead to crystallization as a double salt. For the other possible applica-
tion as combustion aid, a calcium acetate (CA) might actually be preferred
to CMA, since the magnesium does not react.
The raw materials for the production of CMA—acetic acid, and various
kinds of dolomitic lime or limestone—were discussed in Chapter 2, and it
was found that acetic acid cost is the key factor determining the cost of the
final product, since CMA contains almost 80 wt% acetate. Synthetic glacial
acetic derived from petroleum or natural gas costs more than $ 500 per ton,
hence CMA has to be expensive unless a cheaper source of acetate can be
found. Compared to the acid, the cost of the possible neutralizer materials is
almost insignificant, and a choice should be made based on their properties
and suitability for the process rather than on their cost.
Three basically different processes for the manufacture of CMA were re-
viewed,
In the carbonate system made from HAc, CaCO3 , and MgCO3 there was no
profound difference between the behaviour of the calcium and the magnesium
compounds; both carbonates dissolved up to about pH 8. Experiments were
also made with dolomite (which showed a significantly different behaviour,
e.g. much lower reactivity, so that no true equilibrium was established within
24 h) and with a CO2 -bubbled hydroxide system (which was essentially trans-
formed into a carbonate system within a few hours). From the latter result
it was concluded that CO2 bubbling is likely to have a strong influence on
product composition in a CMA fermentation.
The kinetics of the various dissolution reactions relevant for CMA pro-
duction were reviewed in Chapter 4. Rate data and models were collected
from the available literature for carbonates (calcite, magnesite, dolomite)
and oxide minerals (CaO, MgO). Included was also the discussion of several
148
its very low reactivity. Selectively calcined dolomite provides the option of
producing high-magnesium CMA without the addition of expensive magne-
sia; most interestingly, a 2:3 CMA can be made which might be suitable for
crystallization as a double salt. It has been claimed that this double salt has
superior deicing properties [Tod90]. The preliminary design of a small-scale
process for the production of CMA from wood hydrolyzate (waste sugars)
and selectively calcined dolomite was presented, and an economic analysis
showed that this new process compares well to previously proposed ones. It
was concluded that particularly the production of 2:3 CMA from selectively
calcined dolomite has much potential, and that further investigation of this
specific route would definitely be justified.
If the lab scale results with the organism are still promising, the process
could be scaled up to pilot plant level. Chances are slim that an acetic acid
pulping plant will be built at this time because of high overcapacities in the
pulp and paper industry. However, should the need for a new plant arise,
this would be a very interesting alternative to a conventional kraft plant, and
CMA production could be tested at the pilot scale. Here in Wisconsin, an
almost ideal location for such a plant would be the Green Bay area, since
it has a long tradition in the manufacture of pulp and paper, large natural
supplies of dolomite, and lots of ice and snow to melt during the winter.
Appendix A
M
xi = ξi (A.1)
Mi
151
152
Table A.1: Original results of the ICP analyses on six filtered solution samples
from equilibrium experiments; source: Soil & Plant Analysis Laboratory,
University of Wisconsin-Madison.
1
M=P −1 (A.2)
j Mj ξ j
is the molecular weight of the mixture. With the assumption that 1 l solution
contains 55.6 mol H2 O, the molar concentrations in Tables 3.2 and 3.4 were
obtained.
153
where xi are the mole fractions, ξi the mass fractions, and Mi the molecular
weights. Table A.4 shows the detected ranges of concentration (minimum
and maximum) for several other elements that were present as impurities in
the product solution.
154
Table A.2: Original results of the ICP analyses on CMA solution samples
made from type S dolime; source: Soil & Plant Analysis Laboratory, Univer-
sity of Wisconsin-Madison.
Table A.3: Original results of the ICP analyses on CMA solution samples
made from selectively calcined dolomite; source: Soil & Plant Analysis Lab-
oratory, University of Wisconsin-Madison.
Table A.4: Ranges of concentration detected for other elements that were
present as impurities in the CMA solution; source: Soil & Plant Analysis
Laboratory, University of Wisconsin-Madison.
where xi are the mole fractions, ξi the mass fractions, and Mi the molec-
ular weights, the mole ratio of Ca:Mg can easily be calculated—the value
found from the given data is 1.04:1, i.e., the material has almost the desired
composition.
157
158
Table B.1: Sieve analysis of type S dolime; source: The Western Lime &
Cement Co.
Table B.2: Chemical analysis of type S dolime; source: The Western Lime &
Cement Co.
The following equipment was used for the continuous dissolution experiment
described in section 5.1.2; brand names are given for identification purposes
only:
160
161
• cross flow filter capsule with 1.2 µm pores, 900 cm2 area (Gelman
Acroflux)
– slow pump (Mod. 7520-35) with two identical heads (Mod. 7013-
21)
162
163