Sunteți pe pagina 1din 135

RIJKSUNIVERSITEIT GRONINGEN

Absorption of formaldehyde in water


Proefschrift
ter verkrijging van het doctoraat in de
Wiskunde en Natuurwetenschappen
aan de Rijksuniversiteit Groningen
op gezag van de
Rector Magnificus, dr. F. Zwarts,
in het openbaar te verdedigen op
vrijdag 6 juni 2003
om 14.15 uur
door
Jozef Gerhardus Maria Winkelman
geboren op 2 juli 1961
te Warnsveld
Promotores: Prof. dr. ir. L.P.B.M. Janssen
Prof. dr. ir. H.J. Heeres
Prof. dr. ir. A.A.C.M. Beenackers

Beoordelingscommissie: Prof. dr. A.A. Broekhuis


Prof. dr. P.D. Iedema
Prof. dr. ir. G.F. Versteeg
Aan mijn ouders
Acknowledgements
The research reported on in this thesis was supported financially by Dynea B.V., Delfzijl, The
Netherlands, a formaldehyde producer, and by the Technology Foundation (STW) in the
Netherlands.
Contents
Contents
1. Introduction 1
2. The kinetics of the dehydration of methylene glycol.
Abstract 5
Introduction 5
Experimental 7
Results 8
Discussion 14
Physical properties 16
Conclusions 17
3. Simultaneous absorption and/or desorption with reversible first-order chemical
reaction.
Abstract 19
Introduction 19
Absorption with first-order reversible reaction and desorption 20
Special and limiting cases 23
Applications 26
Conclusions 31
4. Kinetics and chemical equilibrium of the hydration of formaldehyde
Abstract 33
Introduction 33
Experimental 35
Results 37
Discussion 45
Physical properties 46
Conclusions 46
Addendum 47
5. Modeling and simulation of industrial formaldehyde absorbers.
Abstract 49
Introduction 49
Reactions in aqueous formaldehyde solutions 50
Development of absorber model 52
Mass transfer rates 53
Energy transfer rates 55
Physical and chemical parameters 56
Vapour liquid equilibria 56
Contents
Computational method 59
Results 61
Conclusions 66
6. Simulation of industrial formaldehyde absorbers: the behaviour of methanol and non-
equilibrium stage modelling.
Abstract 69
Introduction 69
Reactions in aqueous methanolic formaldehyde solutions 70
Vapour liquid equilibria in formaldehyde-water-methanol mixtures 73
Model development 74
Mass transfer rates 76
Energy transfer rates 78
Method of solution 79
Results 83
Conclusions 92
7. Epilogue 93
Symbols 95
References 99
Appendix A: Correlations for the density and viscosity of aqueous formaldehyde
solutions. 103
Appendix B: The equilibrium composition of aqueous methanolic formaldehyde
solutions. 117
Appendix C: The reaction order of formaldehyde in its hydration reaction. 123
List of publications 129
Samenvatting in het nederlands (Summary in Dutch) 131
Dankwoord 137
Chapter 1: Introduction
1
Chapter 1
Introduction
This thesis describes theoretical and experimental work on the absorption of formaldehyde
in water and the development of chemical engineering models for the description and
optimization of industrial formaldehyde absorbers. This introduction gives a short description of
the industrial formaldehyde production process, and of the formaldehyde absorption step therein.
The introduction ends with an outline of the other chapters and appendices.
Formaldehyde is an important base chemical in the process industry with a world
production rate of approximately 10 million metric tons annually (Weirauch, 1999). Historically,
one of the first important applications was in the production of artificial indigo. Nowadays, its
main applications are in the production of engineering plastics and resins, especially urea, phenol
and melamine resins, of which large quantities are used in the plywood and particle board
manufacturing industry, and also in the manufacturing of rubber, paper, fertilisers, explosives,
preservatives, etc. (Walker, 1964; Cancho et al., 1989).
Formaldehyde is industrially produced from methanol. The production is perfomed at
approximately atmospheric pressure. Three major steps can be identified, see Fig 1. In a first
step, the liquid methanol is vaporized into an air stream, and steam is added to the resulting
gaseous mixture. In a second step, the gaseous mixture is lead over a catalyst bed, where the
methanol is converted to formaldehyde via partial dehydrogenation and partial oxidation. The
temperature of the gaseous product increases to approximately 870 K because of the highly
exothermic character of the conversion of methanol to formaldehyde.
steam
air
methanol
tail gas
water
55 wt% formalin
VAPORIZER REACTOR ABSORBER
Fig. 1. Simple scheme of the formaldehyde production process.
Chapter 1: Introduction
2
To prevent thermal decomposition of formaldehyde, the gas stream is cooled directly after
passing over the catalyst. In a third step, the formaldehyde is absorbed in water in an absorption
column, because formaldehyde in its pure, gaseous form is highly unstable, and also because the
reactor product stream contains the formaldehyde diluted in other gases, mainly nitrogen. From
the absorber, the commercial product is obtained: an approximately 55% by weight solution of
formaldehyde in water, or formalin.
The design, operation and optimization of formaldehyde absorbers is complicated by a
number of factors, of which two important ones are the reactions in the liquid phase and the
exothermicity of the processes in the absorber. Formaldehyde absorbers operate less efficient
than could have been expected based on the good apparent solubility of formaldehyde in water.
The reason is that, in aqueous solutions, formaldehyde reacts with water to methylene glycol and
higher poly(oxymethylene) glycols via a series of reversible reactions
2 2 2 2
) OH ( CH O H O CH + (1)
O H H O) HO(CH H O) HO(CH (OH) CH
2 n 2 1 n 2 2 2
+ +

. (2)
The good apparent solubility of formaldehyde in water is actually the good solubility of
methylene glycol, and the capacity of the solution to accommodate poly(oxymethylene) glycols.
Formaldehyde itself, like most gases, is only sparingly soluble in water.
The rate of the hydration reaction (1) is relatively fast, causing chemical enhancement of the
gas-to-liquid transfer of formaldehyde. The formation rate of the higher poly(oxymethylene)
glycols is low, with reaction times in the order of tens of minutes to hours, depending on the
temperature.
The absorption of formaldehyde, and its consequent hydration, as well as the condensation
of steam are exothermic processes. Therefore, the temperature of the liquid increases as it flows
down the absorber.
Because of factors such as the ones mentioned above, formaldehyde absorbers generally are
divided into different absorption sections. Each of the absorption sections, or beds, is provided
with a relatively large, externally cooled liquid recirculation stream. A typical example of a
formaldehyde absorber is shown in Fig. 2.
This thesis
The major aim of this work is the development of reliable models that are capable, first, of
accurately describing the performance of current formaldehyde absorbers, second, of predicting
the influence of changing various operating parameters, and third, of optimising the performance
of the absorber columns towards formaldehyde absorption efficiency and capacity. To achieve
this goal, knowledge of the kinetics of the principal reaction (1) and the consecutive
polymerisation reactions (2) is of major importance. The kinetics of the latter, the formation of
the poly(oxymethylene) glycols, have been investigated extensively by other research groups
(p.e. Dankelman et al., 1988; Hasse & Maurer, 1991; Hahnenstein et al., 1994, 1995).
Chapter 1: Introduction
3
feed
product
tail gas
water
Fig. 2. Scheme of a typical formaldehyde absorber.
The investigations into the kinetics of the principal reaction are treated in the next three
chapters. Following these are two chapters on the development of chemical reaction engineering
models for formaldehyde absorbers, a concluding chapter, and some additional material.
Chapter 2 describes the experimental work on the determination of the dehydration rate of
methylene glycol, the reverse of reaction (1). Using a traditional wet chemistry methodology, the
dehydration rates where obtained from the measured formation rates of hydroxymethane
sulphonate from the reaction of formaldehyde with
- 2
3
SO , at various temperatures. The results
could be correlated to an Arrhenius type expression.
Chapter 1: Introduction
4
In Chapter 3, a theoretical treatment is presented of the problem of simultaneous
absorption and/or desorption of two components, accompanied by a first-order reversible liquid
phase reaction among the two. The analytical solutions developed here for the concentration
profiles in the mass transfer film and for the enhancement factors are used in the next chapter.
Chapter 4 describes the experimental determination of the kinetics of the hydration of
formaldehyde in water. The measurements are based on the chemically enhanced absorption of
formaldehyde gas into water in a stirred cell and mathematical modelling of the transfer process.
The temperature dependency of the hydration rate constant correlates well to an Arrhenius type
expression. From the results of this chapter, combined with those of chapter 2, the equilibrium
constant of the hydration of formaldehyde is obtained.
In Chapter 5 a model is developed for formaldehyde absorbers, based on differential
equations for the mass and energy balances in each phase. The resulting set of coupled boundary
value problems is solved by a semi-transient method.
In Chapter 6 the absorber model is extended with a description of the behaviour of
unconverted methanol that enters the absorber. Also modelled now are vaporisation and
reabsorption of methylene gylcol and hemiformal, the principal reaction products of
formaldehyde with water and methanol. The modelling is based here on a non-equilibrium stage
model.
Concluding remarks can be found in Chapter 7.
In the appendices, some additional material can be found: Appendix A presents the results
of experimental work on the determination of the viscosity of aqueous formaldehyde solutions
and correlations for the density and viscosity of aqueous formaldehyde solutions as a function of
the temperature and the strength of the solution; Appendix B elucidates some calculation
methods to obtain the equilibrium composition of solutions containing formaldehyde. Some
additional material to chapter 4, on the determination of the reaction order of formaldehyde in the
hydration reaction, is included in Appendix C.
Chapter 2: The kinetics of the dehydration of methylene glycol.
5
Chapter 2
The kinetics of the dehydration of methylene glycol.
Abstract
The kinetics of the dehydration of methylene glycol were measured under conditions
relevant in industrial formaldehyde absorbers (293 K T 333 K; 6.0 pH 7.8). The rapid
reaction between formaldehyde and SO
3
2
to hydroxymethane sulphonate (HMS

) was used as a
formaldehyde scavenger:
O H O CH (OH) CH
2 2 2 2
+
d
k
,

+
3 2
fast 2
3 2
)SO (O CH SO O CH ,
+
+
3 2
fast
3 2
(OH)SO CH H )SO (O CH .
At the experimental conditions, the rate determining step in the formation of HMS

appeared to
be the dehydration of methylene glycol. The observed reaction rate constant for the dehydration
of methylene glycol could be correlated as
1 / 6705 7
s 10 96 . 4

=
T
d
e k . The dehydration rate is
shown to be independent of the concentrations of both sulphite and hydroxide ions at the
conditions applied here.
Introduction
For a detailed model of the absorption process of formaldehyde, as well as in the further
applications of formaldehyde, the kinetics of the dehydration of methylene glycol (see eq. (1)),
are important.
The literature data on the dehydration rate of methylene glycol are limited to ambient
temperatures. Using various chemical scavengers, LeHnaff (1963) and Bell et al. (1966)
obtained k
d
= 4.510
-3
s
-1
at 293 K and 5.110
-3
s
-1
at 298 K, respectively. From the
formaldehyde production and subsequent reaction in the radiolysis of methanol in a flow
measurement system, Sutton and Downes (1972) calculated k
d
= 4.410
-3
s
-1
at 295 K. Los et al.
(1977) found k
d
= 5.710
-3
s
-1
at 298 K with pulse polarography. Funderburk et al. (1978)
obtained k
d
= 4.210
-3
s
-1
at 298 K using carbazides and hydrazine as trapping agents.
In this contribution we report on the reaction rate of the dehydration of methylene glycol,
characterized by the reaction rate constant k
d
, at the conditions prevailing in industrial
formaldehyde absorbers; at the wider temperature range of 293-333 K and at a pH between 6 and
7.
The experimental method applied uses the reaction of sodium sulphite with formaldehyde.
In aqueous solution sulphite ions react specifically with the carbonyl group of aldehydes or
ketones to form -hydroxy sulphonates (Blackadder and Hinshelwood, 1958). The sulphite ions
do not react with methylene glycol. The reaction of sulphite with formaldehyde is fast
Chapter 2: The kinetics of the dehydration of methylene glycol.
6
(LeHnaff, 1963; Boyce & Hoffmann, 1984), and under suitable conditions the formation rate of
the product, hydroxymethane sulphonate, CH
2
(OH)SO
3

(HMS

), is completely determined by
the dehydration rate of methylene glycol. Sulphite then is a trapping agent, or chemical
scavenger, of any formaldehyde produced by the dehydration of methylene glycol.
In the mechanism of the HMS

formation, the following reactions are relevant here


O H O CH ) OH ( CH
2 2 2 2
+
h
d
k
k
, (1)
+
+ H SO HSO
2
3
2
3
a
K
, (2)

+
3 2
2
2
2
3 2
SO ) O ( CH SO O CH
k
k
, (3)


+ + OH SO ) OH ( CH O H SO ) O ( CH
3 2
1
) HMS ( 2
2 3 2
a w
K K
. (4)
Srensen and Andersen (1970) studied the kinetics of the HMS

formation from sodium


sulphite and formaldehyde at 298 K in strongly alkaline aqueous solutions (pH 9-12). At these
high pH values they found the rate determining step of the overall reaction to be the dehydration
of methylene glycol catalysed by hydroxide ions, and obtained k
d
= 1.710
3
[OH

] s
-1
at 298 K.
The initial concentrations of formaldehyde, 20-38 mol m
-3
, and sulphite, 29-37 mol m
-3
, were of
comparable magnitudes.
Boyce and Hoffmann (1984) studied the rate of formation of HMS

from formaldehyde in
acidic solutions (pH = 0-3.5) at temperatures from 288 K to 313 K. They concluded that the rate
determining step in the HMS

formation involves the nucleophilic addition of bisulphite and/or


sulphite ions to the carbonyl group of formaldehyde. They also found bisulphite to be the
principal reactant at pH < 2, whereas at pH > 4 the reaction is carried by sulphite. The initial
concentrations of formaldehyde, 10-100 mol m
-3
, were considerably higher than the initial
sulphite + bisulphite concentrations, 0.25-1.25 mol m
-3
.
From the overall reaction equation, which results from adding eqs (1), (3) and (4), it is seen
that the formation of HMS

is accompanied by the release of hydroxide ions



+ + OH (OH)SO CH SO (OH) CH
3 2
2
3 2 2
. (5)
By using sodium bisulphite in combination with sodium sulphite, a buffering capacity is
introduced and the conditions can be chosen in such a way that the pH rises only modestly for a
considerable length of time as compared to the time needed to get close to the maximum
conversion of sulphite.
Chapter 2: The kinetics of the dehydration of methylene glycol.
7
Experimental
A stock solution containing approximately 1% by weight of formaldehyde was prepared by
dissolving paraformaldehyde (Janssen Chimica) in water and allowing to equilibrate several
days. The overall formaldehyde concentration in the stock was determined accurately with the
sulphite method (Walker, 1964). The low overall formaldehyde content in the stock solution
ensures that any polymeric forms of formaldehyde can be neglected (see Appendix B). Reagent
grade sodium bisulphite and sodium sulphite (Janssen Chimica) were used to prepare 100 ml
aliquots of buffer solutions of these components just prior to the experiments. Because carbon
dioxide can interfere with the pH measurements, and oxygen can induce oxidation of bisulphite
or sulphite, the water used in the experiments and in the preparation of the solutions was double
distilled, and boiled out and stored with nitrogen purging.
A Metrohm type E561/1 pH meter was used. The output signal from the pH meter was send
to a computer with a frequency of exactly 1 Hz for later analysis. The experiments were carried
out in a closed, double walled vessel, kept at the desired temperature by circulating water. The
vessel was equipped with a nitrogen inlet, for purging prior to the experiment, a pH electrode
and a septum sealed inlet for adding reagent solutions via syringes.
Before an experiment was started, the desired amount of buffer solution was added to 100
ml water in the stirred reaction vessel, and allowed to attain thermal equilibrium. Meanwhile, a
sample of the formaldehyde stock solution was brought to the same temperature in a separate
vessel also under nitrogen. The injection of a desired amount of this formaldehyde stock solution
into the reaction vessel marked the start of the experiment.
Experiments were performed at 5 temperatures ranging from 293 to 333 K. At each
temperature 8-10 experiments were performed with various sulphite/bisulphite ratios of the
buffer solution, and with various amounts of buffer solution and formaldehyde stock solution
added to the reaction vessel. This way, the following ranges of variation in initial concentrations
were realised: pH
0
= 6-7, [NaHSO
3
]
0
= 1.5-8.0 mol m
-3
, [Na
2
SO
3
]
0
= 0.3-5.2 mol m
-3
, and
[CH
2
(OH)
2
]
0
= 4-25 mol m
-3
. The concentrations were chosen in such a way that
[CH
2
(OH)
2
]
0
> S
tot
in all experiments, where S
tot
denotes the total concentration of sulphur:
0 3 2 0 3
] SO Na [ ] NaHSO [ + =
tot
S . (6)
It should be noted here that [NaHSO
3
]
0
and [Na
2
SO
3
]
0
are the initial molar concentrations of
reagentia used to make up the solution. They are not a priori identical with the true initial
concentrations of bisulphite and sulphite ions, [HSO
3

]
0
and [SO
3
2
]
0
, since these are determined
by the dissociation constants governing the system.
The build-up of pyrosulphate according to
O H O S HSO 2
2
2
5 2 3
+

is proportional to the square of the HSO
3

concentration (Hayon et al., 1972). However, the


equilibrium constant of the reaction is very small, and the S
2
O
5
2
concentration becomes
Chapter 2: The kinetics of the dehydration of methylene glycol.
8
negligible in solutions containing less than 50 mol m
-3
of sulphur (Golding, 1960; Deister et al.,
1986). Since the total sulphur concentrations employed here were always less than 15 mol m
-3
,
we can safely neglect any pyrosulphate formation.
Results
An example of the pH readings recorded during an experiment is shown in Fig. 1. The
profile is typical for all experimental results: after a gradual increase there is a sharp upturn of
the pH corresponding to a decrease of the concentration of hydrogen ions.
In this section we will first explain qualitatively the origin of the typical profiles of the
measured pH curves, and secondly show how quantitative kinetic information was extracted
from the experimental data.
12
10
8
6
0 20 40 60 80 100
t (s)
pH
Fig. 1. pH as a function of time for a typical experiment (no. 20.1): T = 293 K, [NaHSO
3
]
0
=
4.311 mol m
-3
, [Na
2
SO
3
]
0
= 1.146 mol m
-3
, and [CH
2
(OH)
2
]
0
= 22.84 mol m
-3
.
For each of the data points of an experiment, [H
+
] was used to calculate the concentrations
of the other relevant species, [H
2
SO
3
], [HSO
3

], [SO
3
2
], [HMS

], [OH

], [Na
+
] and [CH
2
(OH)
2
],
from the sulphur balance
] HMS [ ] SO [ ] HSO [ ] SO H [
2
3 3 3 2

+ + + =
tot
S , (7)
the sodium balance
] Na [ ] SO Na [ 2 ] NaHSO [
0 3 2 0 3
+
= + , (8)
Chapter 2: The kinetics of the dehydration of methylene glycol.
9
the carbon balance
] [HMS ] (OH) [CH ] (OH) [CH
2 2 0 2 2

+ = , (9)
the charge balance
] [OH ] [HMS ] 2[SO ] [HSO ] [Na ] [H
2
3 3
+ +
+ + + = + , (10)
the water ionisation equilibrium
W
K =
+
] OH ][ H [ , (11)
and the two sulphite dissociation equilibria
1
3 2
3
] SO H [
] H ][ HSO [
a
K =
+
, (12)
2
3
2
3
] HSO [
] H ][ SO [
a
K =

+
. (13)
In writing the carbon balance, the contribution of CH
2
O was neglected because the
equilibrium of eq. (1) is far to the left and the concentration of free formaldehyde is very low
(Zavitsas et al., 1970). The calculations showed that [H
2
SO
3
] was always negligible. Therefore
this species was neglected in the further considerations. For the remaining sulphur containing
components the concentrations vs. pH are shown in Fig. 2 for experiment no 20.1. It may be
noted that before the large pH jump, [SO
3
2
] remains constant. This was found to be the case for
all experiments, and can be explained by subtracting the sulphur balance, eq. (7), from the charge
balance, eq. (10), substituting eqs (6) and (8) for S
tot
and [Na
+
], and rearranging, giving
] H [
] H [ ] SO Na [ ] SO [
0 3 2
2
3
+
+
+ =
W
K
. (14)
Chapter 2: The kinetics of the dehydration of methylene glycol.
10
6 7 8 9
pH
5
4
3
2
1
0
HSO
3
-
SO
-
3
2
HMS
-
10 11
c
o
n
c
e
n
t
r
a
t
i
o
n


(
m
o
l
/
m
)
3
Fig. 2. Concentrations of sulphurous components vs. pH of experiment no 20.1, see Fig. 1.
Symbols: calculated points. Lines: for illustrative purposes only.
It can easily be shown that, for the range of [Na
2
SO
3
]
0
and pH
0
values employed here, in order to
observe a decrease of the sulphite concentration from [SO
3
2
]
0
with say 1%, an increase of the
pH is needed varying from pH > 9 at 293 K to pH > 8 at 333 K. The highest pH measured before
the large pH jump occurred was 7.8. It can therefore safely be assumed that [SO
3
2
] is constant
for pH < 7.8.
Now, the typical pH profile of Fig. 1 can be explained qualitatively. The SO
3
2
ions that
react with formaldehyde to HMS

are replenished with new ones obtained from the dissociation


of HSO
3

. The gradual decrease of [HSO


3

] results in a decrease of the buffering capacity of the


mixture and a gradual increase of the pH. At the point where all HSO
3

initially present is
consumed, the buffering capacity breaks down completely, resulting in the large pH jump. Any
data points beyond the pH jump, i.e. those with pH>8, are not used in the current analysis.
In writing the mechanism of the HMS

formation, eqs (1)-(4), it is assumed that the ion


equilibria, eqs (2) and (4), are established rapidly. The equilibrium of eq (4) is far to the right at
pH < 8, and the concentration of HMS
2
is negligible. Because SO
3
2
is the better nucleophile by
a factor of 10
5
compared to HSO
3

(Burnett, 1982), and based on the results of Boyce and


Hoffmann (1984), see introduction, the direct addition of bisulphite ions to formaldehyde was
neglected in the scheme of the reaction mechanism.
Since the equilibrium of the overall HMS

formation reaction, eq (5), is far to the right


(Dasgupta et al., 1980; Deister et al., 1986) the decomposition of HMS

, via the reverse of


reactions (4) and (3), can be neglected for the first part of the measured pH profiles, i.e. before
the pH jump occurs. The rate of HMS

formation then is
Chapter 2: The kinetics of the dehydration of methylene glycol.
11
] SO ][ O CH [
d
] HMS d[
2
3 2 2

= k
t
. (15)
Applying the steady state principle to [CH
2
O] in reactions (1) and (3) gives
] SO [
] ) OH ( CH [
] O CH [
2
3 2
2 2
2

+
=
k k
k
h
d
, (16)
where k
h
is the pseudo-first-order rate constant for the hydration of formaldehyde, see eq (1).
Substitution of eq (16) into eq (15) gives
] SO [
] ) OH ( CH ][ SO [
d
] ) OH ( CH d[
d
] HMS d[
2
3 2
2 2
2
3 2 2 2


+
= =
k k
k k
t t
h
d
. (17)
Before the pH jump, i.e. pH < 7.8, the sulphite concentration remains constant. This allows the
integration of eq (17)
t k
obs
=
0 2 2 2 2
] ) OH ( CH ln[ ] ) OH ( CH ln[ , (18)
where
] SO [
] SO [
2
3 2
2
3 2

+
=
k k
k k
k
h
d
obs
. (19)
The experimental data appear to be very well described by eq (18). An example is shown in Fig.
3 for the data also used in Figs 1 and 2.
Rewriting eq (19) as
d d
h
obs
k k
k k
k
1
] SO [
1 / 1
2
3
2
+ =

(20)
shows that the values of (k
h
/k
2
) and k
d
at each temperature should follow from the k
obs
data
obtained whilst varying [SO
3
2
]. However, at each temperature applied, a regression analysis of
the data according to eq (20) showed that the variation of 1/k
obs
with 1/[SO
3
2
]
was not significant (95% confidence level). A plot of the data obtained at 323 and 333 K
according to eq (20), shown in Fig. 4, illustrates this nicely. Apparently we have k
2
[SO
3
2
] >> k
h
, and eq (19) reduces to k
obs
= k
d
.
Chapter 2: The kinetics of the dehydration of methylene glycol.
12
3.2
3.1
3.0
2.9
2.8
0 10 20 30 40 50
t (s)
l
n
(




)
m
o
l
/
m
3
[
C
H
(
O
H
)
]
2
2
Fig. 3. Data of experiment no 20.1, see Fig. 1, plotted according to eq (18).
1/[SO ] (m /mol)
3
2- 3
0 1 2 3 4 5
1
/
k

(
s
)
o
b
s
10
2
10
1
1
333K
323K
Fig. 4. Data plotted according to eq (20). Symbols: 1/k
obs
obtained from eq (18) for a single experiment.
Lines: mean value of 1/k
obs
at a given temperature.
The reaction rate, eq (17), then reduces to
] (OH) CH [
d
] (OH) CH d[
d
] HMS d[
2 2
2 2
d
k
t t
= =

. (21)
This way, it is demonstrated that, under the experimental conditions applied, the rate of
formation of HMS

is completely controlled by the dehydration rate of methylene glycol.


Chapter 2: The kinetics of the dehydration of methylene glycol.
13
Regression of the reaction rate constants with the reciprocal temperature gave


=
T
k
d
6705
exp 10 96 . 4
7
s
-1
. (22)
Equation (22) describes the data obtained from the individual experiments with a mean absolute
relative deviation of 13%, see Fig. 5. From the temperature coefficient the activation energy was
found as E
a
= (55.8 2.7) kJ mol
-1
. Regression of the reaction rate constants according to the
transition-state theory resulted in H

= (53.2 2.7) kJ mol


-1
and S

= (-106.3 8.7) J mol


-1
K
-1
for the dehydration reaction.
2.7 2.9 3.1 3.3 3.5 3.7
10
10
10
-1
-2
-3
1
1/ T x 10 (K )
3 1
-
k

(
s
)
d
-
1
Fig. 5. Arrhenius plot of the methylene glycol dehydration rate constant k
d
. Symbols: mean value
of at least 8 experiments. Horizontal bars: standard deviation. Line: eq (22).
Chapter 2: The kinetics of the dehydration of methylene glycol.
14
Discussion
The reaction rate constant for the dehydration of methylene glycol obtained from eq (22) at
293 K, k
d
= 5.710
-3
s
-1
compares well with the literature data mentioned in the introduction.
However, the mechanism obtained here for the reaction of sulphite with formaldehyde in
aqueous solution, i.e. the rate limiting dehydration of methylene glycol, differs from both the
mechanism obtained by Srensen and Andersen (1970) and by Boyce and Hoffmann (1984), see
the introduction of this chapter.
Since the variation of k
d
with [SO
3
2
] appeared not to be significant, see Fig. 4, the addition
of SO
3
2
to formaldehyde can not be rate controlling. Probably, Boyce and Hoffmann (1984)
found otherwise because of the large excess of methylene glycol and the high acidities in their
experiments. Hydrogen ions are known to catalyse the dehydration of methylene glycol under
acidic conditions (Funderburk et al., 1978).
In order to check for the possible catalysis of the dehydration of methylene glycol by
hydroxide ions, which Srensen and Andersen (1970) found rate determining at their conditions,
eq (21) can be extended to
( ) ] (OH) CH [ ] OH [
d
] (OH) CH d[
2 2 OH ,
2 2
+ =
d d
k k
t
. (23)
Because this equation cannot be integrated analytically, the method of Himmelblau et al. (1967)
was used to fit the experimental data. Integration of eq (23) gives


=
i
t
d i d i
t k t k
0
OH , 0 2 2 2 2
d ] OH [ ] (OH) CH ln[ ] (OH) CH ln[ , (24)
where the subscript i denotes the ith data point from an experiment. The integral on the right
hand side of eq (24) was evaluated numerically for each data point. The trapezoidal rule suffices
for this purpose because the values of [OH

] vary smoothly, and data points were obtained every


second. The constants k
d
and k
d,OH
were calculated using multiple regression according to eq (24)
applied to all the data points from all the experiments at a specific temperature simultaneously.
The k
d,OH
values thus obtained appeared to decrease with increasing temperature, and
k
d,OH
[OH

]<<k
d
for all experiments. From this it follows that catalysis of the dehydration of
methylene glycol by hydroxide ions is not significant at the conditions used here. This
conclusion is consistent with the measurements of Funderburk et al. (1978) who showed that the
influence of OH

is significant only for pH > 8 at 298 K.


The same conclusion can be drawn from inspecting the relative residuals of the k
d
values,
by comparing the values obtained from eq (22) with those obtained from the individual
experiments, as a function of the measured initial pH, see Fig. 6. The residuals were calculated as
Chapter 2: The kinetics of the dehydration of methylene glycol.
15
% 100
) (
) ( ) (
residual
experiment
experiment eq(22)

=
d
d d
k
k k
. (25)
If the dehydration rate would be susceptible to catalysis by hydroxide ions at the conditions
employed here, a systematic deviation would be expected in Fig. 6 from more positive residuals
at lower pH, i.e. low [OH

], to more negative ones at higher pH, with higher [OH

]. The absence
of such a trend indicates the absence of catalysis by hydroxide ions in this pH region.
Similarly, any catalysis of the dehydration of methylene glycol by hydrogen ions, as
observed under acidic conditions (Funderburk et al. ,1978), can be excluded in the p|H region
considered her, based on Fig. 6.
The observation of Srensen and Andersen (1970) that the HMS

formation rate is
completely determined by k
d,OH
[OH

] can be attributed to the much higher alkalinity of their


solutions, where the pH ranged from 9 to 12.
6 7
pH
6.5
100
50
0
-50
-100
r
e
s
i
d
u
a
l

o
f

k

(
%
)
d
+13%
-13%
Fig. 6. Relative residuals of the k
d
values from individual experiments
compared to the value obtained from eq (22).
Physical properties
All the calculations mentioned in this chapter were corrected for non-ideal behaviour. This
section explains how and why these corrections were performed.
Because of the electrical interactions between ionic species, even quite dilute electrolyte
solutions exhibit non-ideal behaviour. Thermodynamic treatments then require the use of
activities in stead of concentrations for the charged species. Here, activity coefficients were used
Chapter 2: The kinetics of the dehydration of methylene glycol.
16
to correct for the deviations from ideal behaviour, allowing the thermodynamic dissociation
equilibrium constants to be expressed as a function of the temperature only.
The temperature dependence of the thermodynamic equilibrium constant for the
dissociation of H
2
SO
3
, K
a1
th
, was obtained by regression of the data measured by Devze and
Rumpf (1964)

= =
+
+
08 . 4
2009
exp
] SO H [
] H ][ HSO [
) H ( )
3
HSO (
3 2
3
1
T
K
th
a
. (26)
Similarly, K
a2
th
, for the dissociation of HSO
3

, was obtained from the data measured by Hayon


et al. (1972)

= =

+
39 . 14
1400
exp
] HSO [
] H ][ SO [
)
3
HSO (
) H ( )
2
3
SO (
3
2
3
2
T
K
th
a


. (27)
The thermodynamic dissociation constant of water was taken from Stumm and Morgan (1970,
p.76)
T
T
K
th
w
= =
+
+
01706 . 0
99 . 4470
0875 . 12 ) ] OH ][ H log([ log
) OH ( ) H (
10 10
. (28)
The concentration based equilibrium constants K
a1
, K
a2
and K
w
of eqs (11)-(13) were obtained by
correcting the corresponding thermodynamic equilibrium constants for the influence of the ionic
strength according to the Gntelberg modification of the Debye-Hckel limiting law (Stumm &
Morgan, 1970, p.83)
5 . 0
5 . 0
5 . 1
2
6
1 ) (
10 20 . 4 ln

+
=
T
z
i
i
, (29)
where is the dielectric constant of water, varying from = 80.14 at 293 K down to = 66.78 at
333 K (Handbook of Chemistry and Physics), z
i
denotes the charge number of ion i, and the ionic
strength, , is given by


=
i
i i
z C ) 10 (
2
1
2 3
, (30)
where the summation runs over all ionic species.
Because of the non-linearity of the equations, the concentrations of the components and the
concentration based equilibrium constants had to be calculated using an iterative method. A
Chapter 2: The kinetics of the dehydration of methylene glycol.
17
successive approximation method, initialised by taking all activity coefficients equal to 1.0,
proved to be efficient for this purpose.
Conclusions
For the nearly neutral conditions employed here, pH-range 6.0-7.8, the rate determining step in
the reaction of sulphite with formaldehyde in aqueous sulphite bisulphite buffer solutions is the
dehydration of methylene glycol. The dehydration rate appeared to be independent of
concentrations of both sulphite, [SO
3
2
], and hydroxide, [OH

], ions as opposed to the results


obtained by Boyce and Hoffmann (1984) at very acidic conditions, pH-range 0-3.5, and by
Srensen and Andersen (1970) at more alkaline conditions, pH-range 9-12.
The methylene glycol dehydration rate constant obtained in this work at 293 K agrees well
with the values obtained by other authors at 293-298 K. The experimental k
d
values at
temperatures ranging from 293 to 333 K could be described with an Arrhenius-type equation,
from which an activation energy of E
a
= (55.8 2.7) kJ mol
-1
was obtained.
Chapter 3: Simultaneous absorption and/or desorption with reversible first-order chemical reaction.
19
Chapter 3
Simultaneous absorption and/or desorption with
reversible first-order chemical reaction.
Abstract
The problem of gas absorption accompanied by a first order reversible reaction, producing a
volatile reaction product, is considered. A general analytical solution is developed for the film
model, resulting in explicit relations for the concentration profiles in the film and for the mass
transfer enhancement factors. The solution is not restricted to equal diffusivities, and is also
applicable to the simultaneous absorption and reversible reaction of two gases. Several limiting
cases are derived. It is shown that enhancement factors are possible with values larger than one,
but also smaller than one, and even negative. The latter occurs in the industrially important
absorption of formaldehyde in aqueous solution.
Introduction
In the description of interphase mass transfer accompanied by chemical reactions, it is
common practice to incorporate the effects of reactions on the mass transfer rates in enhancement
factors. In the extensive literature on mass transfer enhancement by chemical reaction, many
contributions can be found on analytical and approximate analytical expressions for the
enhancement factors for a number of reaction stoichiometries, with single or multiple reactions
and with irreversible or reversible reactions [e.g. Westerterp et al. (1984, Chap. 7 and 8) and the
references therein]. The greater part of this literature is devoted to the absorption of a reacting
species from a gaseous phase to a liquid phase, where it reacts to non-volatile products. For this
type of operation, the enhancement factor for the transferred component, defined as Eq (12) has a
lower limit of one.
Recently, Landau (1992) published analytical solutions for the case of first order reactions,
irreversible and reversible, involving a non-volatile reactant present in the liquid phase only, and
desorption of the volatile reaction product. Here also, the enhancement factor for the transferring
species has a minimum of one. In fact, by linking the description of the processes in the
interfacial region to those in the bulk, Landau showed that the enhancement of product
desorption is negligible if the reactant originates completely from the liquid.
A different situation arises if two gases absorb in a liquid, accompanied by reversible
reactions. Cornelisse et al. (1980) numerically simulated the absorption of H
2
S and CO
2
in an
amine solution accompanied by complex reversible reactions, and showed that, under certain
conditions, negative enhancement factors are possible for such systems. As far as we know,
Cornelisse et al. were the first to point out the possibility of negative enhancement factors.
Chapter 3: Simultaneous absorption and/or desorption with reversible first-order chemical reaction.
20
In our work on the simulation of the absorption of formaldehyde in water we came across a
very simple system displaying negative enhancement factors: the absorption of a gas,
accompanied by a first order reversible reaction and partial desorption of the reaction product.
The purpose of this contribution is to analyse this type of absorption kinetics.
Absorption with first-order reversible reaction and desorption
Gas absorption accompanied by a first order reversible reaction and partial desorption of the
reaction product
) ( ) ( A A l g (1)
) ( ) ( P A l l (2)
) ( ) ( P P g l (3)
with the reaction rate equation
) (
1
K
C
C k R
P
A A
= (4)
can be described according to the film model by the following equations for diffusion with
reaction:
A
A
A
R
x
C
D =
2
2
,
d
d
l
(5)
A
P
P
R
x
C
D =
2
2
,
d
d
l
(6)
and boundary conditions
I
P P
I
A A
C C C C x = = = , : 0 (7)
P P A A
C C C C x = = = , : . (8)
The concentrations in the bulk liquid are not necessarily at equilibrium. Equations (5) and (6) can
be solved analytically for the concentrations of A and P in the film:
Chapter 3: Simultaneous absorption and/or desorption with reversible first-order chemical reaction.
22
Substitution of eq (13) into (15) gives
R
R
R
P
I
P
P A
P
I
P
A
I
A
R
R
P
v K
C C
C C K
C C
C C K
v E




] tanh[
) (
] cosh[
1
1 1
) ( ] tanh[
1
1
+
(

(
(

+
(
(

= . (16)
From eqs (12) and (14) it is seen that the sign of E depends on the direction of mass
transfer, i.e. the sign of
0
) /d (d
= x
x C , and on the sign of the concentration difference over the
film, ) ( C C
I
. If the combined effects of mass transfer and reaction cause the concentration of
a component at the interface to be equal to the concentration in the liquid bulk, an asymptote is
found for the enhancement factor of that component. Small deviations from this situation result
in either positive or negative enhancement factors. We will return to this remarkable behaviour
below.
For
l l , , P A
D D = (i.e. v = 1) a solution of eqs (5)-(8) was presented earlier by Shah and
Sharma (1976). Their eq (269) gives a relation for the desorption rate of the product P through
the gas-liquid interface, ) (
,
I
P P P P
C C E k
l
, which contains E
P
implicitly. However, their
equation seems to contain some misprints because calculations using it gave erroneous results.
In engineering calculations, the bulk phase concentrations are usually available from
macroscopic balances. A combination of eqs (13) and (15) with the rate balances
) ( ) / (
, , , A
I
A A A A
I
A g A A g
C C E k m C C k =
l
(17)
) ( ) / (
, , , P
I
P P P P
I
P g P P g
C C E k m C C k =
l
(18)
allows the calculation of the unknown interface concentrations and enhancement factors from the
known bulk phase concentrations. If the gas phase resistance to mass transfer is negligible so that
mass transfer is completely liquid-film-controlled, eqs (17) and (18) can be replaced by
g A A
I
A
C m C
,
= and
g P P
I
P
C m C
,
= .
From the general scheme outlined above, several interesting special cases can be derived.
Here we will look at some of these in detail.
Chapter 3: Simultaneous absorption and/or desorption with reversible first-order chemical reaction.
21
v K
C C v x
v K
C vC x
v K
C C K
x
v K
C KC
x
C
P A
I
P
I
A
P A
R
R
I
P
I
A
R
R
A
+
+
+
+
+
+
+

+
+


=
) ( ) 1 (
] sinh[
] sinh[
] sinh[
) 1 ( sinh

(9)
and
)] ( ) ( [ ) (
P
I
P A
I
A A
I
A
I
P P
C C C C v
x
C C v C C + + =

. (10)
The parameters
R
and v are defined as
l
l
l ,
,
,
1
and
) (
P
A
A
R
D
D
v
K D
v K k
=
+
= . (11)
The enhancement factor for component A for a non-instantaneous reaction defined as
) (
d
d
,
0
,
A
I
A A
x
A
A
A
C C k
x
C
D
E

|
|
.
|

\
|

=
l
l
(12)
is found from Eq (9) as
R
R
R
A
I
A
P A
A
I
A
P
I
P
R
R
A
v K
C C
C C K
C C
C C
K
E




] tanh[
) (
] cosh[
1
1
] tanh[
1
1
+
(

(
(

+
(
(

+ = . (13)
Similarly, the enhancement factor of component P for a non-instantaneous reaction, defined as
) (
d
d
,
0
,
P
I
P P
x
P
P
P
C C k
x
C
D
E

|
|
.
|

\
|

=
l
l
(14)
is found from eqs (10), (12) and (14) as
P
I
P
A
I
A
A P
C C
C C
E v E

+ = ) 1 ( 1 (15)
Chapter 3: Simultaneous absorption and/or desorption with reversible first-order chemical reaction.
23
Special and limiting cases
Instantaneous reversible reaction, volatile reactant and product
If the reaction is instantaneous, the rates of the forward and backward reactions of eq (2)
have infinite values, and chemical equilibrium will prevail everywhere in the liquid. For this
situation, the right-hand sides of eqs (5) and (6) become undefined. Elimination of the reaction
term
A
R from (5) and (6) the gives
0
2
2
2
2
= +
dx
C d
D
dx
C d
D
P
P
A
A
.
With the equilibrium condition
A P
KC C = and the boundary condition (7), this equation can
easily be solved for
A
C in the liquid film:
) (
A
I
A
I
A A
C C
x
C C =

. (19)
A balance at the interface states that the amounts of A and P arriving at the interface from the gas
phase equal the amounts of A and P that diffuse from the interface into the liquid film:
0 0 , , , ,
) ( ) ( ) / ( ) / (
= =
= +
x
P
P x
A
A P
I
P g P P g A
I
A g A A g
dx
dC
D
dx
dC
D m C C k m C C k .(20)
Substitution of eqs (17)-(19) and the equilibrium condition
A P
KC C = into eq. (20) gives the
relation between the enhancement factors
A
E and
P
E :
K
E v E
A P
1
) 1 ( 1 + = . (21)
Now, with a known set of bulk phase concentrations, the interface concentrations and
enhancement factors can be calculated from eqs (17), (18) and (21) together with the equilibrium
condition
I
A
I
P
KC C = .
Volatile reactant, nonvolatile reaction product
For the case where the reaction product P is nonvolatile, the mass transfer enhancement
factor for P is not defined anymore. For this case, from eqs (10) and (12) and the condition
0 ) /d (d
0
=
= x P
x C , an expression is found for
I
P
C
) )( 1 (
A
I
A A P
I
P
C C E v C C + = . (22)
Chapter 3: Simultaneous absorption and/or desorption with reversible first-order chemical reaction.
24
Substitution of eq (22) into (13) results in an equation for E
A
which is only valid if
0 ) /d (d
0
=
= x P
x C :
R
R
R
A
I
A
P A
A
v
K
C C v
C C K
v
K
E



] tanh[
1
] cosh[
1
1
) (
1
+
(

(
(

+ +
= . (23)
This expression is similar to the one obtained by Onda et al. (1970). If also the reaction is
instantaneous, taking the limit for
R
of eq (23) gives
v
K
E
A
+ =1 (24)
which is contained in the solution given for this case by Astarita and Savage (1980).
Nonvolatile reactant, volatile reaction product
Another special case is the situation where the product P is volatile, but where the reactant
A originates completely from the liquid phase and is nonvolatile, 0 ) /d (d
0
=
= x A
x C . Then we
have from eq (10)
)] ( ) ( [
1
d
d
0
P
I
P A
I
A
x
P
C C C C v
x
C
+

= |
.
|

\
|
=

(25)
Substitution of eq (25) into (14) gives
P
I
P
A
I
A
P
C C
C C
v E

+ =1 (26)
Furthermore, from eq (9) together with 0 ) /d (d
0
=
= x A
x C an expression for
I
A
C is found, which
is substituted into eq (26) to find
R
R
R
P
I
P
P A
P
v
K
C C
C C K
v
K
E



] tanh[
] cosh[
1
1 1
+
(

(
(

+
= . (27)
Chapter 3: Simultaneous absorption and/or desorption with reversible first-order chemical reaction.
25
If the reaction is instantaneous, there is chemical equilibrium in the liquid:
I
A
I
P
KC C = and
A P
C K C = . For this case eq (27) reduces to:
K
v
E
P
+ =1 . (28)
This equation was derived by Westerterp et al. (1984, p.423). It is also implicitly present in the
expression obtained by Shah and Sharma (1976) for the desorption flux of P, but their eq (61)
unfortunately contains a misprint.
Irreversible reaction
The first order irreversible reaction can be regarded as a limiting case of the reversible
reaction, with K going to infinity. While eq (15) is still valid, E
A
can be found for this case by
taking the limit of eq (13). In this way we indeed get the classical result:
A
I
A
R A
I
A
R
R
A
K
A
C C
C C
E E

= =

] cosh[ /
] tanh[
) ( lim
,
,
,
,

(29)
with
l ,
1
,
) ( lim
A
R
K
R
D
k
= =

. (30)
Simultaneous absorption and reaction of two gases
Recently, van Swaaij and Versteeg (1992) presented a review on gas-liquid mass transfer
with reversible reactions. For the simultaneous absorption and reaction of two gases they noted
that reversible reactions have not been studied so far. For this case, if the reaction between the
two absorbed gases is first-order-reversible according to the scheme
) ( ) ( A A l g (31)
) ( ) ( P P l g (32)
) ( ) ( P A l l (33)
it may be noted that the problem is also stated by eqs (5)-(8), and that these equations are
independent of the direction of mass transfer. Also the solution in terms of the concentration
profiles (9) and (10) and the enhancement factors (13) and (16) are independent of the direction
of mass transfer. Thus, for the case of absorption of two gases, A and P, accompanied by a first
order reversible reaction in the liquid, the mass transfer rates and enhancement factors can be
obtained from eqs (13) and (16) combined with eqs (17) and (18).
Chapter 3: Simultaneous absorption and/or desorption with reversible first-order chemical reaction.
27

R
=0.1
2
4.47
10
50

x/
0 0.2 0.4 0.6 0.8 1
1.0
0.8
0.6
0.4
0.2
0.0
-0.2
C C
A A
-
m C
A A,g
Fig. 2. Dimensionless concentration of A in the liquid film.
Parameter values as mention in example 1.

R
=
2
4.47
10
50
x/
0 0.2 0.4 0.6 0.8 1
1.0
0.8
0.6
0.4
0.2
0.0
-0.2
v m C
A A,g
C C
P P
-
0.1
Fig. 3. Dimensionless concentration of P in the liquid film.
Parameter values as mention in example 1.
These are shown in Figs 2 and 3 as dimensionless concentration differences vs. the depth in
the liquid film, for various
R
values. At the critical value of 47 . 4 =
R
, the value of
I
P
C equals
that of
P
C . This results in the asymptotic behaviour of
P
E because ) (
, P i P
C C is in the
denominator of eq. (14). However, note that although
I
P
C equals
P
C , there is still desorption of
P as indicated by the positive gradient
P
C at the interface. At larger values of
R
the reaction is
fast enough to make
I
P
C larger than
P
C , and E
P
becomes negative. At smaller values of
R
, the
Chapter 3: Simultaneous absorption and/or desorption with reversible first-order chemical reaction.
28
production of P by reaction cannot keep up with the vaporisation of P:
I
P
C becomes smaller than
P
C and
P
E is positive. In Figs 2 and 3, the lines indicated with =
R
denote the limiting
instantaneous case where the concentration profiles are given by eq. (19) and the equilibrium
condition
A P
KC C = .
The influence of the volatility of the reaction product P on
A
E is shown in Fig. 4, using the
same set of parameters except for the value of ) /(
, , P l P P g
k m k . The line marked with
) /(
, , P l P P g
k m k = 0 denotes the limiting case of volatile reactant with a nonvolatile reaction
product. The dimensionless the concentration profiles of P vs. the depth in the liquid film are
shown in Fig. 5 for a relatively fast reaction ( 10 =
R
). It may be noted that the positive gradient
of P at the interface, and therefore the desorption rate of P, decreases with a decreasing volatility
of P. Also, in the limiting case of a nonvolatile reaction product P, we indeed have
0
) / (
= x P
dx dC = 0.
1 10 10 10 10
2 3 4

R
E
A
10
2
10
1
2
k
g,P
m k
P l,P
=0
10
20
Fig. 4. The influence of the volatility of P on E
A
. Parameter values as mention in example 1.
Chapter 3: Simultaneous absorption and/or desorption with reversible first-order chemical reaction.
26
Applications
In this section we will illustrate the remarkable behaviour of gas absorption accompanied by a
reversible first order reaction, producing a volatile reaction product. The first two examples show
the asymptotic behaviour of the enhancement factor for product desorption and for reactant
absorption with selected sets of parameters. The third example describes an industrially
important case of negative enhancement.
Example 1
The following values are chosen for the parameters: 0
,
=
g P
C mol m
-3
, ) /(
, , A l A A g
k m k =
) /(
, , P l P P g
k m k = 10 , ) /(
, g A A A
C m C = 04 . 0 , ) /(
, g A A P
C vm C = 2 . 0 , and 0 . 5 / = v K . Figure 1
show the variation of
A
E and
P
E with
R
calculated from eqs (13), (15), (17) and (18).
A
E
increases monotonously with increasing
R
until the limiting instantaneous value of

R
A
E

) (
= 37 . 73 is reached for very large
R
.
P
E displays an asymptote at 47 . 4 =
R
. If 47 . 4 <
R

then we have 1
P
E , but surprisingly, if 47 . 4 >
R
then
P
E is always negative, with a limiting
value of

R
p
E

) ( = 47 . 13 for very large
R
. This remarkable behaviour can be understood
by inspection of the concentration profiles in the liquid film.
100
50
0
-50
-100
0.1 1 10 10 10 10
2 3 4

R
E , E
A P
P
A
P
Fig. 1. Variation of E
A
and E
P
with
R
. Parameter
values as mention in example 1.
Chapter 3: Simultaneous absorption and/or desorption with reversible first-order chemical reaction.
29
x/
0 0.2 0.4 0.6 0.8 1
2.5
2.0
1.5
1.0
0.5
0.0
v m C
A A,g
C C
P P
-
2
k
g,P
m k
P l,P
=0
10
20
Fig. 5. The influence of the volatility of P on C
P
in the liquid film. Parameter values as mention in example
1.
0.1 1 10 10 10 10
2 3 4

R
E
A
10
10
10
4
3
2
10
1
1
10
10
2
10
3
K/ = v
Fig. 6. The influence of the reversibility of the reaction on E
A
.
Parameter values as mention in example 1.
The influence of the reversibility of the reaction, as indicated by the value of v K / , on
A
E
is shown in Fig. 6. Again, 0
,
=
g P
C mol m
-3
, ) /(
, , A l A A g
k m k = ) /(
, , P l P P g
k m k = 10 and
) /(
, g A A P
C vm C = 2 . 0 were used. The liquid bulk phase concentration of A was taken in
chemical equilibrium with that of P.
A
E increases with v K / until the limiting case of a first
order irreversible reaction is reached for v K / , where
A
E is given by eq. (29).
Chapter 3: Simultaneous absorption and/or desorption with reversible first-order chemical reaction.
30
Example 2
The same set of parameters as in example 1 is used here, i.e. 0
,
=
g P
C mol m
-3
, ) /(
, , A l A A g
k m k
= ) /(
, , P l P P g
k m k = 10 and 0 . 5 / = v K . The only difference being that now the concentration of
A in the gas phase is lowered in such a way that we now have ) /(
, g A A A
C m C = 5 . 0 and
) /(
, g A A P
C vm C = 5 . 2 . This change in the conditions drastically alters the variation of
A
E and
P
E with
R
, as illustrated in Fig. 7. Now it is the enhancement factor of the absorbing gas,
A
E ,
that displays an asymptote. This occurs at 5 . 17 =
R
. The dimensionless concentration profiles
in the liquid film associated with 5 . 17 =
R
are shown in Fig. 8. The value of
I
A
C equals that of
A
C , but there is clearly absorption of A at the interface as is revealed by the initially large
negative gradient of
A
C in the film. At large values of
R
, the absorption of A cannot keep up
with the disappearance of A due to reaction in the film; the reaction product acts as a sink for
component A, and
I
A
C becomes smaller than
A
C which causes negative enhancement factors for
A. On the other hand, if 5 . 17 <
R
, then the mass transfer of A is relatively faster, and
A
E is
positive because
A
I
A
C C < . The limiting instantaneous reaction enhancement factors for this
situation are 5 . 26 ) ( =

R
A
E

and 5 . 6 ) ( =

R
P
E

. The linear profiles of
A
C and
P
C in
the liquid film for
R
, obtained from eq. (19) and the equilibrium condition
A P
KC C = ,
are also shown in Fig. 8.
50
25
0
-25
-50
0.1 1 10 10 10 10
2 3 4

R
E , E
A P
A
P
A
Fig. 7. Variation of E
A
and E
P
with
R
. Parameter
values as mention in example 2.
Chapter 3: Simultaneous absorption and/or desorption with reversible first-order chemical reaction.
31

R
=17.5

x/
0 0.2 0.4 0.6 0.8 1
C C
A A
-
m C
A A,g
0.0
-0.1
-0.2
-0.3
-0.4
0.0
-0.5
-1.0
-1.5
-2.0

R
=17.5

v m C
A A,g
C C
P P
-
Fig. 8. Dimensionless concentration of A and P in the
liquid film. Parameter values as mention in example 2.
Practical example
Formaldehyde is commercially produced by gas phase oxidation of methanol with air. The
gaseous reactor product stream contains formaldehyde and is fed to an absorber to dissolve the
formaldehyde in water. In water, formaldehyde (A) reacts reversible to methylene glycol (P).
Because of the large excess of water, this reaction can be considered here as pseudo first order.
At the bottom part of the absorber, the entering gas does not yet contain any methylene glycol
( 0
,
=
g P
C mol m
-3
). For this region of the absorber, the following parameter values are typical
for an industrial formaldehyde absorber:, ) /(
, , A l A A g
k m k 95 . 0 = , ) /(
, , P l P P g
k m k
3
10 38 . 3

= ,
v K / 2 . 21 = , 4 . 1 = , ) /(
, g A A A
C m C 24 . 0 = , ) /(
, g A A P
C vm C 14 . 5 = . From eqs (13), (15),
(17) and (18) the enhancement factors are calculated as 55 . 1 =
A
E and 13 . 0 =
P
E , indicating
that negative enhancement factors may occur in the absorption of formaldehyde.
Conclusion
A new analytical solution is presented for the concentration profiles and mass transfer
enhancement in the general case of gas absorption accompanied by a first order reversible
reaction, producing a partly volatile product. The solution is not restricted to equal diffusivities.
From this general solution, several known limiting cases are readily derived. The solution is also
applicable for the case of the simultaneous absorption and first order reversible reaction of two
gases. It is shown that enhancement factors are possible with values not only larger than one but
also smaller than one, and even negative enhancement factors are possible. The latter occurs in
the industrially important absorption of formaldehyde in aqueous solution.
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
33
Chapter 4
Kinetics and chemical equilibrium of the hydration of formaldehyde
Abstract
The reaction rate of the hydration of formaldehyde is obtained from the measured,
chemically enhanced absorption rate of formaldehyde gas into water in a stirred cell with a plane
gas liquid interface, and mathematically modelling of the transfer processes. Experiments were
performed at the conditions prevailing in industrial formaldehyde absorbers, i.e. at temperatures
of 293-333 K and at pH values between 5 and 7. The observed rate constants could be correlated
as k
h
= 2.0410
5
e
-2936/T
s
-1
. Using the results, and the dehydration reaction rate constant,
obtained previously at similar conditions, the chemical equilibrium constant for the hydration is
obtained as K
h
= e
3769/T-5.494
.
Introduction
Formaldehyde is an important industrial intermediate in the manufacturing of resins,
plastics, adhesives and many other products. Because formaldehyde is unstable in its pure,
gaseous state it is usually produced as an aqueous solution. In such a solution, formaldehyde is
almost completely hydrated to methylene glycol,
2 2 2 2
) (OH CH
k
k
O H O CH
d
h
+ . (1)
Methylene glycol, in turn, depending on the strength of the solution, may polymerize to form a
series of polyoxymethylene glycols,
O H H O CH HO H O CH HO OH CH
n n 2 2 1 2 2 2
) ( ) ( ) ( + +

. (2)
In the design of formaldehyde absorption and distillation processes, as well as in
downstream processing, the kinetics and chemical equilibria of both reactions are important. The
research group of Maurer recently studied the kinetics and chemical equilibria of the
polyoxymethylene glycol formation (Hasse & Maurer, 1991; Hahnenstein, Hasse, Kreiter &
Maurer, 1994; Hahnenstein, Albert, Hasse, Kreiter & Maurer, 1995)
In the open literature, only two entries with experimental data on the reaction rate constant
of the hydration of aqueous formaldehyde, k
h
, were encountered. Schecker and Schulz (1969)
obtained formaldehyde hydration rates from temperature jump experiments and measurement of
the approach of the formaldehyde concentration to the new equilibrium value with UV-
absorption. From experiments at 298-333 K they obtained k
h
= 7800e
-1913/T
s
-1
at pH = 4-7.
Sutton and Downes (1972) obtained k
h
= 9.8 s
-1
at 295 K from radiolysis of aqueous solutions of
methanol containing oxygen and semicarbazide hydrochloride in a flow system.
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
34
Most of the literature data on the chemical equilibrium constant of the hydration of
formaldehyde, K
h
, were obtained from the carbonyl specific UV-absorption at approximately
290 nm (Bieber & Trmpler, 1947; Iliceto, 1954; Landqvist, 1955; Gruen & McTigue, 1963;
Siling & Akselrod, 1968; Schecker & Schulz, 1969; Zavitsas, Coffiner, Wiseman & Zavitsas,
1970). At this wavelength an electron from a non-bonding oxygen n-orbital is promoted to an
anti-bonding *-orbital of the carbonyl double bond (e.g. Calvert & Pitts, 1966). From their
measurements mentioned above, Sutton and Downes (1972) could calculate also K
h
at 295 K.
Valenta (1960) used oscillographic polarography with pulses of short duration. Then, unhydrated
formaldehyde is the only reducible species and its concentration could be obtained. Finally,
Bryant and Thompson (1971) derived an expression for K
h
from a consistent set of, partly
experimental, thermochemical data. The values of K
h
obtained by the various authors show a
considerable spreading; differences of more than a factor 3 are found. The reaction enthalpy for
the hydration obtained from the sources mentioned varies from -21.4 to -39.4 kJ mol
-1
.
In previous work,
h
K was usually determined from the concentrations of free formaldehyde
and methylene glycol in aqueous solutions, i.e. from the equilibrium value of
F MG
C C / . We
think that much of the spreading of the reported
h
K data in the literature can be explained from
the, occasionally unrecognized, complications in establishing these concentrations. In aqueous
solutions, the concentration of unhydrated formaldehyde is very low because the equilibrium (1)
is far to the right. In addition, the molar extinction coefficient of formaldehyde in UV-absorption
is small and its value for aqueous solutions is not known. In more concentrated solutions (say
above 1 M) the amount of unhydrated formaldehyde is higher, but the then the polymerization
reactions (2) become significant and the methylene glycol concentration is not known anymore.
Also, with more concentrated solutions, UV-absorption measurements are hampered by
substantial nonspecific absorbance. Finally, commercial formaldehyde solutions often contain
considerable amounts of methanol for stabilization; a fact that was not always recognized in older
literature.
In this contribution, we report on the reaction rate of the hydration of formaldehyde,
characterized by the rate constant k
h
, at conditions prevailing in industrial formaldehyde
absorbers, i.e. at temperatures of 293-333 K and at pH values between 5 and 7. Using these
results, and the dehydration reaction rate constant, k
d
, obtained previously at similar conditions
(Winkelman, Ottens and Beenackers, 2000), the chemical equilibrium constant for the hydration
is obtained as
d h h
k k K / = . The
h
K data obtained this way, as the ratio of measured reaction
rates rather than as the ratio of concentrations, are believed to be more reliable for the reasons
mentioned in the section above.
The experimental technique for obtaining the hydration rate is based on the measurement of
the chemically enhanced absorption rate of formaldehyde gas into water. In general, chemical
enhancement occurs if the absorption of a gaseous component into a liquid is accompanied by a
chemical reaction of that component in the liquid. If the reaction is fast enough, then the
concentration of the component is reduced in the liquid already near the gas-liquid interface. This
results in a larger gradient, and thus a larger flux, of the component, when compared to the
gradient and flux without any reaction. The phenomenon is characterized by the so-called
enhancement factor,
i
E , which is defined as
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
35
reaction without ) (
reaction with ) (
0
0
=
=

x i
x i
i
J
J
E , (3)
where both fluxes at the gas-liquid interface,
0
) (
= x i
J , are based on the same concentration
difference over the liquid film ) (
, , , l l i IF i
C C (see e.g. chapter VII in Westerterp, Van Swaaij &
Beenackers, 1984).
Formaldehyde absorption experiments are carried out in a stirred cell reactor. The water in
the reactor is stirred with a constant, but limited intensity, in such a way that the liquid surface
always remains flat, and the value of the gas-liquid interfacial area remains accurately known.
This way, the formaldehyde gas to liquid molar flux can be obtained from observed mass transfer
rate, which, in turn, allows the calculation of the liquid phase chemical enhancement factor for
formaldehyde mass transfer and the reaction rate of formaldehyde hydration, respectively, from
mathematical modeling of the transfer processes.
The calculation of the desired hydration reaction rate from the observed mass transfer rate
requires accurate knowledge of both the gas and liquid phase mass transfer coefficients.
Therefore, prior to the kinetic measurements, the mass transfer coefficients prevailing in the
stirred cell reactor were accurately measured.
Experimental
The experimental set-up for the kinetic measurements was build around a stirred cell
reactor, see Fig. 1. The double walled glass reactor, 0.08 m diameter, 0.1 m height, was equipped
with 4 baffles. Stirring was provided by a 0.046 m 8-bladed turbine stirrer in the gas phase, and a
0.050 m flat blade stirrer in the liquid phase on the same shaft, driven via a magnetic coupling.
The liquid surface remained perfectly flat for stirring rates of up to 22 Hz. Beyond this limit,
some rippling and heightening of water against the baffles was observed. Nitrogen was passed
through a saturation column, approximately 1 m in height, filled with a 32 wt.% aqueous
formaldehyde solution, and flowed either via the head space of the reactor, or directly via a by-
pass, to the analysis unit. The signals of the temperature control units, pressure measurement,
flow measurement and stirring rate were stored into a computer. The composition of the solution
in the saturation column was accurately determined using the sodium sulphite method (Walker,
1964).
The formaldehyde concentration of the gas stream entering and leaving the reactor were
measured with UV-spectrophotometry using the carbonyl specific absorption around 295 nm. For
this purpose, a special measuring unit was constructed by Macam Photometrics Ltd.. UV-light
was obtained from a low pressure deuterium discharge lamp with a stabilised power supply and
reference intensity measurement for stable radiation output. Further, the unit was provided with
collimating and alignment optics, a 2 nm band width monochromator, a special long-path
measuring cuvet with temperature control, and a photomultiplier tube. The measuring cuvet, with
a volume of only 4010
-6
m
3
, was equipped with two flat mirrors and a concave mirror, in such a
way that a total optical path length of 1 m was achieved by 12 passes of the light through the
cuvet, see Fig. 2. This way, the low extinction coefficient of the absorption band of gaseous
formaldehyde was compensated for, and accurate measurements could be made.
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
36
1
2
3
4
Fig. 1. Experimental set-up, 1: gas supply, 2: saturation column,
3: stirred cell, 4: to analysis unit.
In a typical kinetic experiment, the reactor was filled with the desired amount of distilled
water, and the system, including the saturation column, the UV absorption measuring cuvet, and
all connections, was allowed to equilibrate to the desired temperature. Next, the nitrogen flow to
the saturation column was set to the desired rate, and passed directly to the analysis unit, via the
by-pass, to measure the reactor inlet formaldehyde concentration. The stirrer was switched to the
desired rate, and the flow was led through the reactor until the achievement of a pseudo steady
state, as indicated by a constant UV-absorption reading from the spectrophotometer, marked the
end of the experiment after a few minutes. This way, with an experiment, the hydration was
measured at a single pseudo steady state point.
to photo-
mutiplier
from mono-
chromator
Fig. 2. Measuring cuvet for UV-absorption and optical path of the light.
For the liquid phase mass transfer measurements, the gas supply was switched to CO
2
, and a
vacuum pump was connected to the gas phase reactor outlet. Before an experiment, the water was
degassed for 15 minutes by lowering the pressure to just above the water vapour pressure at the
temperature employed. Next the pressure was lowered even further, and the water allowed to boil
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
37
for a short time, to drive out any gasses remaining. An experiment was started by pressurising the
reactor with CO
2
within a few seconds to approximately 0.12 MPa, closing the inlet valve, and
recording the pressure decrease at a rate of exactly 1 Hz until the pressure decrease had
diminished.
Gas phase mass transfer coefficients were obtained by measuring the evaporation rates of
pure liquids into an inert gas stream. For this purpose, a continuous N
2
, CO
2
or He gas supply
was used, while the gas phase reactor outlet was now connected to 3 cold traps in series, cooled
with liquid nitrogen, or, when CO
2
gas was used, a mixture of water, ice and CaCl
2
at
approximately 258 K. A fourth trap, filled with CaCl
2
, prevented any moisture from the
environment to enter. The amount of evaporated liquid was obtained from the weight increase of
the cold traps. The following combinations of gases and pure liquids were used: N
2
with water,
octane and ethanol; CO
2
with water; He with butyl ether, acetone, anhydrous proprionic acid and
butanol.
Results
Liquid phase mass transfer coefficients
The liquid phase mass transfer coefficients were determined by monitoring the pressure
drop during the absorption of CO
2
in water, whilst operating both gas and liquid phases in batch
mode. The variation of the CO2 partial pressure with time, obtained from molar balances for CO
2
over both phases, is given by
l
l
l
l l
V
t
V
V m
S k
V m
V
P
P
V m
V
g
CO
CO
g
CO
CO
CO
g
) 1 ( ) 1 ( ln
2
2
2
2
2
0
+ =

+ . (4)
where the distribution coefficient,
2
CO
m , was taken from Versteeg and Van Swaaij (1988).
Pressure readings were taken every second, for a total monitoring time varying from about half
an hour to one hour. The individual experiments could be described by equation (4) very well:
the mean absolute relative residual (MARR) of the calculated and observed pressures never
exceeded 0.1%.
The experimental
l
k -values were correlated using dimensionless groups, resulting in
% . Sc Re . Sh
. .
5 4 178 0
36 0 58 0
=
l l l
, (5)
where a 4-fold variation of Sc and a 22-fold variation of Re were achieved by measuring at
different temperatures and stirring rates, see Fig. 3.
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
38
10
4
10
5
Re
L
10
100
S
h
L
/
S
c
L 0
.
3
6
Fig. 3. Liquid phase mass transfer coefficients. Symbols:
experimental data. Line: Eq. (5).
Gas phase mass transfer coefficients
The gas phase mass transfer coefficients were obtained from the rate of evaporation of pure
liquids into carrier gases. This way, any liquid side resistance against mass transfer was
eliminated. The
g
k values were calculated by solving simultaneously the balance equations for
the vapour content of the carrier gas stream,
tot i
i
in
g v
i
sat
i i g
P P
P
p p S k
/ 1
) (
,
,

=

(6)
and the weight of condensed vapour accumulated in the cold traps,
RT P P
P
t M
W
tot i
i
in
g v
i
) / 1 (
,

=

(7)
for
i g
k
,
and
i
P . No influence of the gas flow rate on
g
k was observed for any of the
components. The
g
k data were also correlated via dimensionless groups:
% Sc Re . Sh
.
g
.
g g
11 163 0
50 0 70 0
= . (8)
By using eight combinations of carrier gases and liquids, and various stirring rates, a 23-fold
variation of Sc and a 50-fold variation of Re were achieved, see Fig. 4.
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
39
10
2
10
3
10
4
100
10
1
S
h
/
S
c
g
g 0
.
5
0
Re
g
N
N
N
CO
He
He
He
He
2
2
2
2
- water
- octane
- ethanol
- water
- butyl ether
- acetone
- prop. anhydr.
- butanol
Fig. 4. Gas phase mass transfer coefficients. Symbols:
experimental data. Line: Eq. (8).
UV-analysis of formaldehyde
The formaldehyde concentration in the gas stream was obtained spectrophotometrically
from the intensity of the characteristic absorption band of the carbonyl group at a wavelength of
295 nm. Simple aldehydes, p.e. acetaldehyde, propionaldehyde and butyraldehydes, show nearly
continuous absorption spectra in this wavelength region, and obey Beer's law (Calvert & Pitts,
1966; Mller & Schurath, 1983)
C I I l = ) / ln(
0
. (9)
However, because of its simple structure, the absorption band of formaldehyde shows
considerably more vibrational and rotational structure, and has a discrete line structure. Because
the spectral line widths are narrower than the resolution of 2 nm used in the measurements, the
absorbance varies nonlinear with the formaldehyde concentration. This has been reported
previously by Moortgat, Seiler and Warneck (1983), Mller and Schurath (1983) and Rogers
(1990).
Mller and Schurath (1983) measured the absorption of gaseous formaldehyde in a 2.48 m
cell in the concentration range of 0-0.12 mol m
-3
. They correlated their results according to
) 702 . 2 186 . 2 ( ) / ln(
0
C C I I = l . This correlation is illustrated in Fig. 5, together with our
measured data, showing a good agreement. The gas phase formaldehyde concentrations of Fig. 5
are the ones in the gas stream entering the reactor, and are calculated from the composition of the
solution in the saturation column, using the vapour-liquid equilibrium model of developed by the
research group of Maurer (Maurer, 1986; Albert, Garca, Kuhnert, Peschla, & Maurer, 2000), and
assuming that the gas leaving the saturation column has reached physical equilibrium with the
liquid.
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
40
0.0 0.05 0.10 0.15 0.20
0.4
0.3
0.2
0.1
0.0
C
F,g
[mol/m ]
3
l
n
(
I
/
I
)
0
Fig. 5. Absorbance vs. gaseous formaldehyde concentration. Symbols: this work,
horizontal bars: standard deviation. Line: Measured by Mller and Schurath
(1983) at 12 . 0 0
,
=
g F
C mol/m
3
and extrapolated to 0.2 mol/m
3
.
The absorbances vs. the formaldehyde gas concentrations for the entire concentration range
employed here are shown in Fig. 6. They could be correlated with an expression similar to the
one obtained by Mller and Schurath, and which is also shown in Fig. 6:
) 488 . 0 652 . 1 ( ) / ln(
0
C C I I = l (10)
0.0 0.2 0.4 0.6 0.8
C
F,g
[mol/m ]
3
l
n
(
I
/
I
)
0
1.2
0.8
0.4
0.0
Fig. 6. Absorbance vs. gaseous formaldehyde concentration. Symbols: this work,
horizontal bars: standard deviation. Line: Eq. (10).
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
41
Kinetic measurements
The gaseous phase above an aqueous formaldehyde solution contains formaldehyde gas,
water vapour, and also methylene glycol vapour (Maurer, 1986). Therefore, the gas stream
entering the stirred cell from the saturation column will contain these three components, and the
absorption of formaldehyde in the reactor is accompanied by absorption of methylene glycol.
Since the vapour pressure of water above an aqueous formaldehyde solution is lower than above
pure water at the same temperature, some evaporation of water will also occur
) ( 2 ) ( 2 l
O CH O CH
g
, (11)
) (
2 2
) (
2 2
) ( ) (
l
OH CH OH CH
g
, (12)
) ( 2 ) ( 2 g
O H O H
l
. (13)
Once absorbed, formaldehyde will be hydrated according to the reversible reaction (1). Since
water is present in large excess, the hydration reaction can be characterised by a first-order rate
constant, k
h
(Bell, 1966):
) (
,
,
h
MG
F h F
K
C
C k R
l
l
= (14)
The reaction rates of the polyoxymethylene glycol formation reactions, Eq. (2), are very low.
Furthermore, if the overall concentration of formaldehyde is low (say below 1 wt.%), the
equilibrium constants of these reactions do not favour the formation of polyoxymethylene glycols
and the solution will contain formaldehyde and methylene glycol. Therefore, any
polyoxymethylene glycol formation is neglected in this work.
The experiments are evaluated using the two-film model. This model for mass transfer is
based on the assumption that near the interface in the liquid phase a stagnant film exists, of
thickness
l l l
k D / = , where any transport of the components is by diffusion only. Similarly, in
the gas phase a film is present at the interface of a thickness
g g g
k D / = . Using the film model
for mass transfer, the following equations describe the processes in the stirred cell
) , , , ( ) (
2 , , , , 0
N W MG F i C C S J
g i g v
in
g i
in
g v x i
= =
=
, (15)
) , , ( ) / ( ) (
, , , , 0
W MG F i m C C k J
i IF i g i i g x i
= =
= l
, (16)
) , ( ) ( ) (
, , , , 0
MG F i C C E k J
i IF i i i x i
= =
= l l l
. (17)
For the general case of absorption and/or desorption of two gases accompanied by a first-order
reversible reaction in the liquid, Winkelman and Beenackers (1993) derived analytical solutions
for the enhancement factors of the components involved. Translated to the notation used here,
their equation for the enhancement factor of methylene glycol reads
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
42
l l
l l
, , ,
, , ,
) 1 ( 1
MG IF MG
F IF F
F MG
C C
C C
E v E

+ = . (18)
With the measured
g F
C
,
, and with 0
2
=
N
J and
l l , , , W IF W
C C = , the set of Eqs (15)-(18)
can be solved for the other gasphase concentrations, gas flow rate leaving the cell, interface
concentrations, and mass transfer enhancement factors. In the calculations, it was assumed that
the formaldehyde and methylene glycol concentrations in the bulk of the liquid are negligible
compared to the interface concentrations, see the addendum at the end of this chapter. The film
thickness was obtained from Eq (5) using
l l l
k D / = , and varied in the experiments from
approximately 50 to 130 m. The observed reactor outlet gas phase formaldehyde concentrations
of the individual experiments increased with increasing gas flow rates, and decreased with
increasing stirring rates and temperatures. In Fig. 7 the ratio of the reactor outlet and inlet
formaldehyde concentrations is plotted versus the quantity ) /( ) (
3
,
T N
in
g v
, which was chosen
intuitively for illustrative reasons only, to reduce the scattering of the data in the plot.
0.8
0.6
0.4
0.2
0.0
( )
/(NT)

v,g
in 3
10 10 10
-19 -18 -17
C
/
C
F
,
g
F
,
g
i
n
Fig. 7. Ratio of reactor outlet and inlet gas phase formaldehyde concentrations obtained experimentally.
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
43
The enhancement factor of formaldehyde calculated from the experimental data, as
described in the previous paragraph, can be equated to the one obtained from the differential
equations for diffusion with parallel reaction in the liquid according to the film model
) 0 ( ) (
2
2
, l l
= x
K
C
C k
dx
C d
D
h
MG
F h
F
F
, (19)
) 0 ( ) (
2
2
, l l
= x
K
C
C k
dx
C d
D
h
MG
F h
MG
MG
. (20)
An analytical solution of (19)-(20) for the E
F
, with the assumption of negligible liquid bulk
phase concentrations, reads (Winkelman & Beenackers, 1993)
v K
C
C
K
E
h
IF F
IF MG
h
R
R
F
+

+ =
) )( 1
] tanh[
(
1
, ,
, ,
l
l

, (21)
where the reaction factor,
R
, is defined by
h F
h h
R
K D
v K k ) ( +
=
l
. (22)
With the values of E
F
obtained from Eqs (15)-(18), the reaction rate constants, k
h
, were calculated
numerically from (21)-(22), where the equilibrium constant K
h
was written as K
h
= k
h
/k
d
, and k
d
was taken from Winkelman, Ottens and Beenackers (2000). The values of
R
obtained
numerically ranged from 3.7 to 11.4.
Finally, regression of the reaction rate constants with the reciprocal temperature gave
1 5
)
2936
exp( 10 04 . 2

= s
T
k
h
. (23)
The mean absolute relative deviation between the data from the individual experiments and from
Eq. (23) is 9.6%, see Fig. 8. From the temperature coefficient the activation energy was found as
E
a
= (24.4 2.7) kJ mol
-1
. Regression of the results according to the transition-state theory
resulted in H

=(21.8 2.7) kJ mol


-1
and S

=(-152.0 9.5) J mol


-1
K
-1
. The enhancement
factors obtained from Eqs (21)-(23) are plotted against the ones obtained from the experimental
data and Eqs (15)-(18) in Fig. 9.
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
44
2.8 3.0 3.2 3.4 3.6
10
30
50
5
3
1000/T
k
h

[
1
/
s
]
Fig. 8. Arrhenius plot of the formaldehyde hydration rate constant k
h
.
Symbols: : this work, mean of 8 to 12 experiments, horizontal bars:
standard deviation; : Sutton & Downes (1972). Solid Line: this
work, Eq. (23). Dashed line: Schecker & Schulz (1969).
0 5 10 15
(E )
F observed
(
E
)
F
c
a
l
c
u
l
a
t
e
d
15
10
5
0
Fig. 9. Parity plot of the formaldehyde enhancement factors.
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
45
With the hydration rate according to Eq. (23), and the dehydration rate obtained previously
(Winkelman, Ottens and Beenackers 2000) the equilibrium constant for the hydration of
formaldehyde can be obtained as
) 494 . 5
3769
exp( = =
T k
k
K
d
h
h
. (24)
Equation (24) is illustrated in Fig. 10. From the temperature coefficient in Eq. (24) the reaction
enthalpy of the hydration was obtained as H = -31.4 kJ mol
-1
.
The experimental data obtained here can also be used to actually establish the reaction order
of formaldehyde in the hydration reaction, see Appendix C. It turns out that the reaction is indeed
first order in formaldehyde under the experimental conditions applied.
2.9 3.0 3.1 3.2 3.3 3.4 3.5
1000/T
K
h
1000
500
100
300
3000
1
2
3
4
5
Fig. 10. Van 't Hoff plot of the formaldehyde hydration chemical equilibrium constant.
Solid line: this work, Eq. (24). Symbols: : Landqvist (1955); : Bieber & Trmpler
(1947); : Valenta (1960); : Iliceto (1954); : Sutton & Downes (1972). Dashed lines:
1: Schecker & Schulz (1969); 2: Zavitsas et al. (1970); 3: Gruen &McTigue (1963); 4:
Bryant & Thompson (1971); 5: Siling & Akselrod (1968).
Discussion
The relation for the liquid phase mass transfer coefficients, Eq. (5) is of a form often
encountered in the literature. Usually, the exponent of
l
Sc is taken as 1/3 based on theoretical
reasons (Westerterp, Van Swaaij & Beenackers, 1984). The exponent of 0.36, obtained from
fitting the data, is in good agreement with this theoretical value. The value of the exponent of
l
Re obtained here, 0.58, is at the lower end of the wide range encountered in the literature for
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
46
this type of reactor: from 0.5 to 1.0. This may be attributed to the specific design of our liquid
phase stirrer.
Data on gas-side mass transfer coefficients in stirred cell reactors are more scarce. Tamir,
Merchuk and Virkar (1979) and Yadav and Sharma (1979) investigated the influence of the
diffusivity on
g
k in stirred cell reactors. For
n
g i g
D k
,
they obtained exponents n = 0.632 and n
= 0.487-0.518, respectively. Our exponent of 0.5 of
g
Sc is in good agreement with these data.
The reaction rate constants of this work are in reasonable agreement with the results of
Schecker and Schulz (1969), see Fig. 8. However, the activation energy of the hydration,
obtained here as 24.4 kJ mol
-1
, is substantially higher compared to the value of 15.9 kJ mol
-1
obtained by Schecker and Schultz. The agreement of Eq. (22) with the value established by
Sutton and Downes (1972) is excellent.
The chemical equilibrium constant for the hydration of formaldehyde, obtained here as Eq.
(24), appears to be within the range of data and relations found in the literature, see Fig. 10. A
thorough comparison however is not possible due to considerable spreading of the literature data
as mentioned before. Also, the heat of reaction, H = -31.4 kJ mol
-1
, appears to be in the wide
range data, from -21.4 to -39.4 kJ mol
-1
, encountered in the literature.
Physical properties
Pure component properties were taken from Daubert and Danner (1985), or predicted using
the methods given by Reid, Prausnitz and Poling (1988). The distribution coefficient and
diffusivity of CO
2
in water were taken from Versteeg and Van Swaaij (1988). Mixture properties
were calculated using the mixing rules given by Reid et al. (1988). The density and viscosity of
aqueous formaldehyde solutions were taken from Winkelman and Beenackers (2000). Diffusion
coefficients in water were calculated with the Wilke-Chang method (Reid et al., 1988), where the
volume of formaldehyde at its normal boiling point was taken form Daubert and Danner (1985)
and that of methylene glycol was obtained with the Le Bas method (Reid, et al., 1988). The
distribution coefficients m
F
, m
MG
and m
W
were calculated with the model of Maurer for vapour-
liquid equilibria of aqueous formaldehyde solutions (Maurer, 1986; Albert, Garca, Kuhnert,
Peschla, and Maurer, 2000).
Conclusions
The reaction rate of the hydration of formaldehyde is obtained from measuring the
chemically enhanced absorption of formaldehyde gas into water in a stirred cell with a plane gas
liquid interface, and mathematically modelling of the transfer processes. At the conditions
prevailing in industrial formaldehyde absorbers, i.e. at temperatures of 293-333 K and at pH
values between 5 and 7, the rate is found as k
h
= 2.0410
5
e
-2936/T
s
-1
.
Using these results, and the dehydration reaction rate constant, k
d
, obtained previously at similar
conditions, the chemical equilibrium constant for the hydration is obtained as K
h
= e
3769/T-5.494
.
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
47
Addendum
In the calculation of the hydration rate constants, the bulk liquid concentrations of
formaldehyde and methylene glycol were neglected. In this addendum, we take a closer look at
the influence of this assumption by modelling the system without neglecting the bulk liquid
concentrations, and comparing the results to the ones previously obtained.
Model equations
During a kinetic experiment, gas flow continuously through the headspace of the stirred
cell, while the liquid is in batch mode. The transient component balances over the headspace and
the liquid bulk read
) , , ( ) (
0 , , , ,
,
W MG F i J S C C
dt
C d
V
x i g i g v
in
g i
in
g v
g i
g
= =
=
(A-1)
) , ( ) (
,
MG F i J S R V n
dt
C d
V
x i F i
i
= + =
= l
l
l
(A-2)
with the initial conditions
.
, , , , , ,
; 0 : 0
sat
g W g W MG F g MG g F
C C C C C C t = = = = = =
l l
. (A-3)
In Eqs (A-1) and (A-2), x denotes the distance into the liquid, thus
0
) (
= x
J denotes the flux of a
component at the gas-liquid interface and
= x
J) ( denotes the flux into the liquid bulk at the
interface of the liquid film and bulk,
F
R is the rate of the hydration reaction in the liquid bulk,
see Eq (14), and the stoichiometric coefficients,
i
n , are given by 1 =
F
n and 1 =
MG
n .
The rate constant
h
k was determined for each experiment by adjusting it until
g F
C
,
,
obtained by integration of Eqs (A-1)-(A3) over the experimental run time, was equal to its
measured value.
The Fourier times for diffusion in the gas and liquid film (typically of the order of 0.1 s) are
much smaller than the time scale at which variation of the bulk concentrations take place (the gas
phase residence time was of the order of 10 to 20 s). Therefore, during the integration, the fluxes
of formaldehyde and methylene glycol at the interface are calculated by solving simultaneously
Eqs (16)-(18) and the expression for the enhancement factor of formaldehyde, which reads in this
case (Winkelman and Beenackers, 1993):
R
R
h
R F IF F
MG F h
F IF F
MG IF MG
h
R
R
F
v K
C C
C C K
C C
C C
K
E




] tanh[
) (
] cosh[
1
1
] tanh[
1
1
, , ,
, ,
, , ,
, , ,
+

+ =
l l
l l
l l
l l
(A-4)
Chapter 4: Kinetics and chemical equilibrium of the hydration of formaldehyde
48
where
R
is given by Eq (22).
The fluxes from the liquid film into the liquid bulk are obtained from the analytical
expressions for the concentrations in the film (Winkelman and Beenackers, 1993) by
differentiation according to =
= x i
J ) (
=

x i i
dx dC D ) / (
,l
, giving
)
] cosh[
1
1 (
] tanh[
) (
) ( ) (
) ( ) (
, , , , , ,
, 0
R
R
R
h
MG IF MG F IF F h
F x F x F
v K
C C C C K
k J J


+
+ +
=
= =
l l l l
l
(A-5)
and
] [ ) ( ) (
, , ,
, , , ,
v
C C
C C k J J
MG IF MG
F IF F F x F x MG
l l
l l l

+ + =
= =
(A-6)
Results
The transient model described above was solved numerically for
h
k for each experiment.
When compared to the
h
k data obtained Eqs (15)-(18) and (21), the differences were small. On
average, the rates obtained here were 0.48 % larger, with a maximum of 1.02 %. Since this is
well within the estimated experimental uncertainty, we conclude that the influence of the bulk
concentration of formaldehyde and methylene glycol can safely be neglected in the evaluation of
the experiments.
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
49
Chapter 5
Modelling and simulation of industrial formaldehyde absorbers
Abstract
The industrially important process of formaldehyde absorption in water constitutes a case of
multicomponent mass transfer with multiple reactions and considerable heat effects. A stable
solution algorithm is developed to simulate the performance of industrial absorbers for this
process using a differential model. Good agreement with practice was achieved. Using the model,
the conditions of one of the absorbers of Dynea B.V. were optimized, leading to considerable
methanol savings.
Introduction
Formaldehyde, CH
2
O, is produced on a large scale as a raw material for a great variety of
end products. Its industrial production starts from methanol. Air and vaporized methanol,
combined with steam and recycled gas, are passed over hot silver grains, at ambient pressure.
Here the methanol is converted to formaldehyde by partial oxidation and by reduction at about
870 K. To prevent thermal decomposition of formaldehyde, the gases are cooled immediately
after the catalyst bed.
The reactor product gas stream has a temperature of 420 K and consists typically of N
2
(50%), H
2
(15%), water vapor (20%) and formaldehyde (15%). Minor amounts of by-products
and unreacted methanol are neglected in this study. This stream is passed through a partial
condenser, where the temperature is reduced to 328 K and part of the water vapor and
formaldehyde are condensed. The resulting mixed gas-liquid stream is subsequently fed to the
absorber.
Because it is impossible to handle in its pure, gaseous form, formaldehyde is almost
exclusively produced and processed as an aqueous solution: formalin. The latter is obtained
commercially by absorbing the gases leaving the reactor in water.
The goals in optimizing the absorber performance seem conflicting. On the one hand the
formaldehyde content of the tail gas should be minimized. On the other hand however, the
formaldehyde concentration in the product liquid leaving the absorber should be as high as
possible, thereby reducing the absorbing ability of the liquid in the column.
A scheme of the absorber studied is shown in Fig. 1. The gaseous part of the two-phase
stream entering at the bottom of the column passes upwards through two beds, randomly filled
with modern high performance Pall-ring like packing. The tail gas is partly recirculated to the
reactor. Make up water enters at the top of the column and flows downward, meanwhile taking
up heat and absorbing formaldehyde and water from the gas stream. Each of the absorption beds
has an external liquid recirculation with heat exchangers. Just below the top bed, the absorber is
equipped with a partial draw off tray to provide a buffer for the upper liquid recirculation pump.
At the bottom of the column a liquid layer is kept as a buffer for the lower liquid reciculation
pump.
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
50
feed
product
tail gas
water
Fig. 1. Scheme of a formaldehyde absorber.
Reactions in aqueous formaldehyde solutions
Besides direct heat transfer between gas and liquid, and the mass transfer of water and
formaldehyde, a number of reactions have to be considered in modeling the performance of
formaldehyde absorbers. In formalin the dissolved formaldehyde is present principally in the
form of methylene glycol, CH
2
(OH)
2
, and a series of low molecular weight polyoxymethylene
glycols, HO(CH
2
O)
n
H (e.g. Walker, 1964). As an example the equilibrium composition of an
aqueous 50 wt.% formaldehyde solution is shown in Fig. 2.
Although the concentration of the unhydrated monomeric formaldehyde is well under 0.1%
even in concentrated solutions, all dissolved aldehyde remains available for chemical reaction in
downstream processing because of the reversibility of the reactions.
The following reactions take place in the absorber:
hydration
2 2 2 2
) OH ( CH O H O CH + (1)
with kinetics
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
51
) / (
1
WF W F 1 h h
K C C C k r = (2)
were
eq
W F
WF
h
C C
C
K

=
1
(3)
polymerization reactions
O H H O) HO(CH H O) HO(CH (OH) CH
2 n 2 1 n 2 2 2
+ +

) 2 (
max
n n (4)
with kinetics
) / (
1 1
n W WF WF WF n n
K C C C C k r
n n
=

(5)
were
eq
WF WF
W WF
n
n
n
C C
C C
K

=
1 1
. (6)
Here, n
max
denotes the largest polymer considered. The concentration of the polyoxymethylene
glycols decreases rapidly with increasing molecular weight, even for concentrated solutions (Fig.
2). Therefore, the largest polymer considered here is WF
10
(n
max
= 10). This way, there are ten
reactions in the liquid between twelve components.
The reaction rate of a species can be represented as

=
=
max
1
,
n
j
j j i i
r R . (7)
From eq (7) and the stoichiometry of reactions (1) and (4) it follows
1
r R
F
= (8)
max
...
3 2 1 n W
r r r r R + + + + = (9)
max 1
... 2
4 3 2 1 n WF
r r r r r R = (10)
) 2 (
max 1
n n r r R
n n WF
n
< =
+
(11)
) (
max
n n r R
n WF
n
= = . (12)
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
52
1
10
10
10
10
-1
-2
-3
-4
F W 1 2 3 4 5 6 7 8 9 10
WF
n
x
Fig. 2. Equilibrium molar fractions of F, W, and WF
n
(n=1..10)
in 50 wt% formalin at 343 K.
Development of absorber model
The methods found in the literature for the modeling of packed columns generally belong to
either of two types. In the first type the packed height is divided into a number of segments.
Within each segment the conditions are supposed to be uniform in both phases. This type of
models can be subdivided into equilibrium stage models (p.e. King, 1980), where the streams
leaving each stage are assumed to be in equilibrium with each other and departures from this
assumption are accounted for by one of several types of stage efficiencies, and nonequilibrium
stage models (p.e. Krishnamurty & Taylor, 1985), where material and energy balance relations
for each phase are solved simultaneously with mass and energy transfer rate equations.
In the second type of models, differential mass and energy balances for each phase are
written for a small section of packing, and the differential equations are numerically integrated
over the total packed height (p.e. Hitch et al., 1986).
Our model for the formaldehyde absorber belongs to the second type. In a subsequent paper
a stage model will be introduced for this type of column.
Since our first goal is the simulation of existing industrial formaldehyde absorbers, the
height of the packing in the absorption beds is fixed. Major assumptions of the model are: (1) the
absorption beds operate adiabatic; (2) the packing is fully wetted, therefore the interfacial area is
the same for heat and mass transfer; (3) heat and mass transfer relations are based on the
resistances in series model; (4) the counter current gas and liquid streams in the absorption beds
are in plug flow; (5) gas to liquid mass transfer of N
2
and H
2
is negligible because of their low
solubility in formalin; (6) the liquid in the partial draw off tray and at the bottom of the column
(the buffers for the recirculation pumps) is ideally mixed and heat and mass transfer to this liquid
is negligible.
With these assumptions, the component balances for the gas and liquid phases read
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
53
) , (
,
W F i aS J
dz
dv
g i
i
= = (13)
) .. 1 ; , , (
max ,
n n WF W F i S R aS J
dz
dl
n i i
i
= = =
l l
. (14)
The energy balances for the gas and liquid phases are
aS q
dz
dT
Cp v
g
g
i
g i i
=

) (
,
(15)

=
j
j R j
i
i i
H r S aS q
dz
dT
Cp l ) ( ) (
, l l
l
l
. (16)
The component and energy balances for the liquid on the partial draw off tray and at the bottom
of the column read
i B
in
i
out
i
R V l l + = (17)

=
j
j R j B
i
i
in
i
in out
H r V Cp l T T ) ( ) (
,l
. (18)
Mass transfer rates
The gas phase mass fluxes are calculated from
) , ( ) (
, , , ,
W F i C C k J
I
g i g i i g g i
= = (19)
were the interfacial concentrations are coupled by
) , (
, ,
W F i C m C
I
g i i
I
i
= =
l
. (20)
The fluxes on both sides of the interface are equal:
I
i g i
J J
l , ,
= . (21)
In the liquid, the diffusional transport is accompanied by chemical reactions, which causes mass
transfer enhancement. Therefore the enhancement factor, E
i
, is incorporated in the fluxes at the
liquid side of the interface
) (
, , , , l l l l i
I
i i i
I
i
C C E k J = . (22)
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
54
Also, the fluxes into the liquid bulk,
l , i
J , become different from the fluxes at the interface,
I
i
J
l ,
.
In a preliminary study this situation was assessed by solving the differential equations for mass
transfer with reactions in the film
) .. 1 ; , , (
max
2
2
,
n n WF W F i R
dx
C d
D
n i
i
i
= = =
l
. (23)
Assuming the polyoxymethylene glycols are nonvolatile, the boundary conditions of eq (23) are
0 , , : 0
,
,
,
,
= = = =
dx
dC
D
J
dx
dC
D
J
dx
dC
x
n
WF
W
I
W
W
F
I
F
F
l
l
l
l
(24)
l l ,
:
i i
C C x = = . (25)
The eqs (19)-(25) were solved numerically for a variety of conditions expected to prevail in
formaldehyde absorbers, using a fourth order Runge-Kutta method in combination with a
shooting method. The results showed that the polymerization reactions (4) are by far too slow to
have any influence on the diffusion fluxes, and the gradients of the concentrations of the higher
polyoxymethylene glycols in the film are negligible,
) .. 2 (
max ,
) 0 ( n n C C
n n
WF
x
WF
= =
l l
. (26)
On the other hand, the hydration reaction (1) causes significant mass transfer enhancement.
In column simulations it is not convenient to calculate the fluxes from the numerical
integration procedure described above. Therefore, an approximate method was developed, similar
to the method of Onda et al. (1970). The differential equations (23) for the formaldehyde and
methylene glycol concentrations in the film are linearized by taking the water concentration in
the film equal to that in the bulk. The resulting set of equations can be solved analytically to give
expressions for E
F
and
l , F
J

R
R
W h
F
WF
R
F
I
F
WF F W h
F
WF
W
F
WF
F
C K
D
D
C C
C C C K
D
D
C K
D
D
E



] tanh[
1
)
] cosh[
1
1 ( 1
,
,
,
, ,
, , ,
,
,
, 1
,
,
1
1 1 1
l
l
l
l l
l l l
l
l
l
l
l
+


+ +
= (27)
[
]
I
F F
I
F F WF
I
F W h WF
WF F W h WF
R
R
W h
F
WF
R I
F F
J C C k C C C K k
C C C K k
C K
D
D
J J
l l l l l l l l
l l l l
l
l
l
l l
, , , , , , , ,
, , , ,
,
,
,
, ,
) ( ) (
) (
] tanh[
] cosh[
1
1
1 1
1 1
1
+ +

(28)
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
55
with
) (
,
,
,
,
1
h WF
F
W
F
h
R
K D
D
C
D
k
l
l
l
l
l
+ = . (29)
Given the bulk concentrations in both phases, all the fluxes can now be calculated iteratively
from eqs (19)-(22), (27)-(29) and the balances
) (
, , , , l l l l F
I
F
I
W W
J J J J = (30)
) (
, , ,
1
l l l F
I
F WF
J J J = . (31)
The mass transfer rates calculated this way proved almost identical to the ones obtained by
numerically solving eqs (19)-(25).
Energy transfer rates
The energy transfer rates contain a conductive and a convective part
) ( ) (
, ,
l or g j H J q E
i
j
T
j i j i j j
= + =

. (32)
The film model of simultaneous mass and energy transfer leads to (p.e. Krishna and Taylor,
1986)
) ( ) ( l or g j T T A h q
I
j f j j
= = (33)
where A
f
is the Ackermann heat transfer correction factor for non-zero mass transfer fluxes in
phase j
1
=
f
C
f
f
e
C
A (34)
) (
, ,
l or g j
h
Cp J
C
j
i
j i j i
f
= =

. (35)
The interfacial temperature,
I
T , follows from a balance around the interfacial region,
l
E E
g
= :

+ = +
i
T
i i
i
g
T
g i g i g
H J q H J q ) (
, ,
) (
, , l l l l
. (36)
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
56
This balance can be rewritten as
h R F
I
F
i
I
i i
i
i vap
I
g g i g i g
H J J
T T Cp J H T T Cp J q q
) )( (
) ( ] ) ( [
, ,
, , , , ,
+
+ + + =

l l
l l l l
. (37)
Physical and chemical parameters
The mass transfer coefficients,
g i
k
,
and
l , i
k , and the specific area, a, were calculated from Onda
et al. (1968). The partial liquid hold up,
l
, was taken from Otake and Kunigita (1958). The heat
transfer coefficients,
g
h and
l
h , were evaluated from the mass transfer coefficients using the
Chilton-Colburn analogy.
To predict the heat and mass transfer coefficients, pure component and mixture physical
properties are needed. The pure component properties were taken from Daubert and Danner
(1985), or if not available, calculated using the predictive methods recommended by Reid et al.
(1988). Densities and viscosities of the liquid were taken from Appendix A. Other mixture
properties were calculated from the mixing rules recommended by Reid et al. (1988).
The reaction rate and equilibrium constant of the hydration reaction,
h
k and
h
K , were
taken from the results reported in Chapters 2 and 4 of this thesis. The reaction rate and the
equilibrium constants of the polymerization reactions, k
j
and K
j
(j=2..n
max
), were measured by
Dankelman et al. (1988) over the temperature range of 293-353 K.
Vapor liquid equilibria
Generally, at low pressure, vapor phase non-ideality can be neglected, and vapor-liquid equilibria
can be described with
s
i i i i
P x P y = (38)
where the liquid phase non-ideality is accounted for by the activity coefficient,
i
, which usually
is temperature and composition dependent.
The interpretation of published vapor-liquid equilibrium data for the formaldehyde-water
system is complicated by the fact that the analytical methods, used by various authors, do not
distinguish between different forms of formaldehyde present in the system, and the experimental
results are presented in terms of overall compositions.
To find vapor liquid equilibrium relationships for formaldehyde and water, experimental data
were used of Kogan et al. (1977), Maurer (1986) and Hasse (1990). First the true molar fractions
were calculated from the reported overall composition and the equilibrium relations (3) and (6).
After this, for each experimental point, the quantity
s
i i
P was calculated.
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
57
25
20
15
10
8
6
4
2
0
0 20 40 60 80
W [wt%]
F

F F
P
S
[MPa]
293 K
313 K
323 K
343 K
353 K
363 K
Fig. 3. Values of
F
P
F
S
. Lines: calculated from eq (49). Symbols: experimental data,
( , ): Kogan (1977), ( , ): Maurer (1986), ( , ): Hasse (1990).
Finally,
s
F F
P and
s
W W
P were smoothed as a function of T and liquid phase composition
] 10 9198 . 7
10 0723 . 3 10 3681 . 7
86 . 3146
4822 . 25 exp[
2 5
2 5
F
F F
s
F F
W
W T W
T
P



+ =
(39)
]
65 . 74
75 . 3140
0428 . 22 exp[

=
T
P
s
W W
. (40)
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
58
100
5
0
0 20 40 60 80
W [wt%]
F

W W
P
S
[kPa]
293 K
313 K
323 K
343 K
353 K
363 K
10
80
60
40
Fig. 5. Values of
W
P
W
S
. Lines: calculated from eq (49). Symbols: experimental data,
( , ): Kogan (1977), ( , ): Maurer (1986), ( , ): Hasse (1990).
Equations (39) and (40) are illustrated in Figs 3 and 4. The deviation of the calculated lines from
the experimental points is generally within the experimental accuracy limits, the mean deviations
are 3.4% for formaldehyde and 1.5% for water. The equilibrium ratio m
i
(eq 20) can now easily
be calculated from
s
i i
tot
i
P
RT C
m

l ,
= . (41)
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
59
Computational method
The eqs (13)-(16) form a set of 16 coupled non-linear differential equations which describe an
absorption bed. If the entering gas and liquid streams are specified, a two point boundary value
problem results. Several strategies for solving this type of problems for absorbers have been put
forward.
Treybal (1969) used a shooting method to model single solute systems. This method starts
by assigning trial values to the outlet gas stream conditions, and calculating the outlet liquid
conditions from overall balances. The numerical integration from bottom to top has to be
repeated until convergency is achieved on the trial values. Feintuch and Treybal (1978) extended
this model to multicomponent systems, but reported convergence problems due to equations
becoming mathematical indeterminate in one of several nested iteration loops. Kelly et al. (1984)
used the shooting method to model the physical absorption of acid gases in methanol.
Another way to solve the set of equations is by dynamic simulation. Von Stockar and Wilke
(1977), using a relaxation technique developed by Stilchmair (1972) and Bourne at al. (1974) for
plate columns, could avoid convergence problems by simulating the packed column start-up and
integrating the model equations with respect to time up to steady state. Hitch et al. (1987) also
developed an unsteady-state solution algorithm. Besides the advantage of providing information
on transient behaviour, they state that the relaxation method can handle complex systems without
the convergence problems encountered with other methods. This type of method is highly stable
because it follows the actual behaviour of start-up procedures.
Srivastava and Joseph (1984) solved the two-point boundary value problem associated with
packed separation columns by using orthogonal collocation for the spatial discretisation. Though
the method worked well for the steady state calculation of a binary system, with multicomponent
systems no steady state solution could be obtained directly. A dynamic simulation had to be
carried out to find the steady state as the asymptote of the transient response.
All methods mentioned above have been applied for single absorption beds, with specified
entering gas and liquid streams. The formaldehyde absorber considered here, however, consists
of two packed beds (Fig. 1). A further complicating factor, from the problem solving point of
view, is the presence of the liquid recirculations around each packed bed. Figure 5 shows the
streams that have to be considered in modeling the formaldehyde absorber. Specified are the
input streams 1A, 7 and 10, and the temperature and total mass flow of the liquid recycle streams
5A and 5B. The actual entering streams of the packed beds are a priori unknown.
Because of its superior convergency and stability characteristics, a transient approach was
used to solve the equations that describe the absorber. Initially, the composition of the liquid in
both packed beds, and of the liquid streams shown in Fig. 5, are all set equal to that of the make-
up stream 1A (pure water). The liquid temperatures in and around each bed are initially set equal
to the corresponding liquid recycle stream temperatures, which are input parameters.
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
60
tail gas
9
make up
top
bed
1A
2A 5A
3A 4A
8
7
10
2B
1B
3B 4B
5B
6
bottom
bed
gas feed
liquid feed
formalin
gas stream
liquid stream
Fig. 5. Flow diagram of the absorber of Fig. 1.
After initialization, a two-step cycle procedure is started. The first step starts with the
specified conditions of the gas feed stream 7, to integrate the eqs (13) and (15) up the bottom
bed. The resulting molar flows and temperature of stream 8 are used as the starting
values to integrate eqs (13) and (15) up the top bed. During the integrations, the interphase fluxes
are calculated using the stored liquid phase conditions. The calculated values of
g
T , v
F
and v
W
in
the absorption beds are stored. In the second step, new liquid phase conditions are calculated.
Starting with the specified conditions of stream 1A and the old values of stream 5A, the molar
flows and temperature of stream 2A are calculated, and used as starting values for the integration
of eqs (14) and (16) down the top bed. During this integration, the calculated liquid phase
conditions are used in conjunction with the stored gas phase conditions to obtain the mass and
heat fluxes. Next, eqs (17) and (18) are used to calculate stream 4A. The molar flows of stream
5A follow from the molar fractions, and the specified total mass flow rate of stream 5A. The
temperature of 1B is set equal to that of 4A, and the molar flow rates follow from the differences
between 4A and 5A. Now, the same sequence is repeated for the lower section of the
formaldehyde absorber, to arrive at stream 6.
The sequential update of gas and liquid phase conditions is repeated until the differences
between successively calculated molar flow rates and temperatures have fallen below preset
criteria, and steady state is reached, see Fig. 6. In practice 40 to 60 iterations were required,
depending somewhat on the operating conditions.
In contrast to the unsteady-state algorithms found in the literature, this algorithm does not
need a separate routine for the calculation of time derivatives and advancement in the time
direction. Furthermore, our algorithm contains only one level of nested iteration (which is the
calculation of the interfacial conditions necessary to obtain the fluxes in the packed beds, and the
calculation of the streams 4A and 4B from the non-linear algebraic eqs (17) and (18)).
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
61
start
input: feed streams, mass flow rates
initialize liquid phase conditions
store gas phase conditions
integrate gas phase equations up the packed
beds using Runge-Kutta
and temperature of liquid recycles
integrate liquid phase equations down the
packed beds and calculate liquid streams
store liquid phase conditions
convergence achieved?
stop
Fig. 6. Logic flow diagram for computing the absorber.
Results
A typical set of calculated temperature and vapor phase molar flow profiles is shown in Fig.
7. The presence of two packed absorption beds is clearly revealed in this figure. The occurence of
temperature maxima, found here in the bottom bed, is often encountered in exothermic
absorption with solvent evaporation (e.g. experimental observations of Raal and Khurana (1973)
and Bourne et al. (1974), and calculations of Stockar and Wilke (1977) and Krishnamurty and
Taylor (1986)). The liquid flowing down from the upper part of the bottom bed increases in
temperature due to the heats of absorption and reaction. But lower in the bottom bed, the liquid
meets an unsaturated gas flow, and the heat effect of solvent evaporation causes a drop in
l
T .
Because
l
T T
in
g
< the gas temperature rises while flowing upwards, until a maximum is reached.
Above the maximum
g
T exceeds
l
T , with the former decreasing because the energy transfer is
now directed towards the liquid.
The features of the temperature profiles are reflected in the profile of the molar flow rate of
water in the gas phase, see Fig. 7. The mass transfer of formaldehyde , on the other hand, is
always directed towards the liquid and is not influenced much by the changes in temperature.
Therefore, the gas phase molar flow rate of formaldehyde is monotonously decreasing. The
profiles of the liquid phase molar flows are not very exciting due to the large liquid recycle ratio's
used in formaldehyde absorbers.
The influence of several operating parameters on the performance of the absorber was
investigated. Important output parameters are the temperature rise of the liquid in each of the
packed beds,
l
T , the relative amount of formaldehyde that leaves the absorber with
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
62
vapor
liquid
water
formaldehyde
v
v
in
T
[C]
70
60
50
40
2
1
0
0.0 0.5 1.0
(bottom) (top)
relative height
Fig. 7. Typical calculated profiles.
the tail gas,
in
F
out
F
v v / , and the composition of the liquid streams leaving each of the packed beds,
calculated here as the overall weight percentage formaldehyde, W
F
.
The influence of the recycle ratio's is shown in Figs 8 and 9. R is defined here as the ratio of
the liquid mass flow rates of 5A and 1B for the top bed, and 5B and 6 for the bottom bed (see
Fig. 5). Increasing the recycle ratio increases the absorber efficiency, i.e. reduces
in
F
out
F
v v / . This
is caused by a combination of more favourable hydrodynamics and lower mean temperatures. For
example, increasing R
b
has virtually no influence on
b
T
, l
, but decreases the internal maximum
of the temperature profile.
Figures 10 and 11 show that increasing the temperature of a liquid recycle has a negative
effect on the absorber efficiency. Increase of T
t
Rec
causes a decrease of the amount of
formaldehyde absorbed in the top bed, but an even larger decrease of the amount of water
absorbed in this bed. This results in a reduction of the temperature rise of the liquid in the top
bed, and a larger weight percentage formaldehyde in the liquid leaving both beds, see Fig. 10.
The overall temperature rise of the liquid in the bottom bed with increasing T
b
Rec
becomes even
negative, see Fig. 11, due to the increasing water evaporation. This is of course a hypothetical
situation, because it would mean that the heat exchanger in the recycle would have to heat up
instead of cooling the liquid stream.
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
63
top-bed
top-bed
bottom-bed
bottom-bed
T
L
W
F
[Wt%]
v
F
out
v
F
in
W
F
[Wt%]
57
56
55
54
R
t
40 60 80 100
43
41
39
37
16
12
8
4
0
0.08
0.06
0.04
0.02
0
Fig. 8. Influence of
t
R on
l
T ,
in
F
out
F
v v / and
F
W .
top-bed
top-bed
bottom-bed
bottom-bed
T
L
W
F
[Wt%]
v
F
out
v
F
in
W
F
[Wt%]
57
56
55
54
R
b
40 20 30
43
41
39
37
16
12
8
4
0
0.08
0.06
0.04
0.02
0
Fig. 9. Influence of
b
R on
l
T ,
in
F
out
F
v v / and
F
W .
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
64
top-bed
top-bed
bottom-bed
bottom-bed
T
L
W
F
[Wt%]
v
F
out
v
F
in
W
F
[Wt%]
57
56
55
54
T
t
Rec
40 25 30
43
41
39
37
10
8
6
4
2
0
0.08
0.06
0.04
0.02
0
35 45
Fig. 10. Influence of
c
t
T
Re
on
l
T ,
in
F
out
F
v v / and
F
W .
top-bed
top-bed
bottom-bed
bottom-bed
T
L
W
F
[Wt%]
v
F
out
v
F
in
W
F
[Wt%]
57
56
55
54
T
b
Rec
70 55 65
43
41
39
37
15
10
5
0
0.08
0.06
0.04
0.02
0
60
-5
Fig. 11. Influence of
c
b
T
Re
on
l
T ,
in
F
out
F
v v / and
F
W .
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
65
top-bed
top-bed
bottom-bed
bottom-bed
T
L
W
F
[Wt%]
v
F
out
v
F
in
60
50
40
30
20
10
3 0 2
10
8
6
4
2
0
0.08
0.06
0.04
0.02
0
1
make-up [m /hr]
3
Fig. 12. Influence of the amount of make-up water on
l
T ,
in
F
out
F
v v / and
F
W .
Increasing the amount of make-up water has a positive influence on the absorber efficiency,
see Fig. 12. This is caused by a reduction of the formaldehyde content of the liquid, which
reduces the backpressure from the liquid phase. However, this reduction in W
F
is often
undesirable because of extra storage and transport costs.
A higher pressure increases the driving forces for absorption, thus reducing
in
F
out
F
v v / , see
Fig. 13. However, the extra water absorption is even higher, so that W
F
still will be reduced. In
practice, this can be compensated for by reducing the amount of make-up water.
Finally, the influence of varying the total feed rate (streams 7 and 10 in Fig. 5) is illustrated
in Fig. 14 for both a constant and a proportionally increased make-up water supply. With
increasing column loads, T
t
L
increases while the absorption efficiency ) / 1 (
in
F
out
F
v v goes
down. With a constant make-up, not only a decrease but surprisingly also an increase in the
relative feed rates results in a diluted liquid product. The latter effect is caused by a stronger
reduction of formaldehyde absorption efficiency relative to water: a maximum of W
F
results.
The model was tested by simulating industrial absorbers of two plants of Dynea B.V., The
Netherlands: one absorber with a configuration as shown in Fig. 1, and another absorber with
three packed beds and a slightly different configuration. In both cases, using the actual operating
parameters, the model could very well predict the performance of the columns.
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
66
top-bed
top-bed
bottom-bed
bottom-bed
T
L
W
F
[Wt%]
v
F
out
v
F
in
W
F
[Wt%]
57
56
55
54
1.0 1.8
43
41
39
37
10
8
6
4
2
0
0.08
0.06
0.04
0.02
0
1.4
P [bar]
Fig. 13. Influence of the total pressure on
l
T ,
in
F
out
F
v v / and
F
W .
Next, the model was used to optimize the performance of the first absorber, by simultaneously
varying the parameters shown in Figs 8-12 and 14. This resulted in a savings of 1% in the costs
of methanol (the basic material for making formaldehyde) for this particular formaldehyde plant.
Conclusions
The absorption of formaldehyde is an important step in the industrial formalin production. In the
absorber column, the process of formaldehyde and water absorption is accompanied by a number
of reactions in the liquid phase and considerable heat effects, necessitating separate liquid
recycles with external heat exchangers. For the simulation of this type of column we developed a
model based on the appropriate differential equations, without using HETP or HTU concepts.
The model is completely predictive. The convergence problems often encountered with this type
of complex modeling could be avoided by using a stable, semi-transient solution-algorithm.
Chapter 5: Modeling and simulation of industrial formaldehyde absorbers
67
3
2
1
0
0.08
0.04
0
56
55
54
53
0.5 1.0 1.5
15
10
5
0
45
35
25
bottom bed
bottom bed
top bed
top bed
constant make-up
make-up pro rata
T
L
T
L
v
F
out
v
F
in
W
F
(wt%)
W
F
(wt%)
Fig. 14. Influence of the relative feed rate on
l
T ,
in
F
out
F
v v / and
F
W .
Our results from simulations with varying process parameters suggest favourable absorber
performance with high liquid recycle ratio's (Figs 8 and 9), low temperatures of the recirculated
liquid (Figs 10 and 11) and a large amount of make-up water (Fig. 12), although the latter may be
limited by a minimum desired formaldehyde content of the product stream.
The suggestions mentioned above have been tested in practice and have led to a more
efficient absorption column.
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
69
Chapter 6
Simulation of industrial formaldehyde absorbers:
the behaviour of methanol and non-equilibrium stage modelling.
Abstract
A model is presented for the commercially important formaldehyde absorption in the
presence of methanol. Incorporated in the model are a large number of liquid phase reactions, gas
liquid heat transfer, and mass transfer with reaction of water, formaldehyde, methylene glycol,
methanol and hemiformal. The evaporation of water, methylene glycol and hemiformal in the
lower part of the column creates an internal circulation of these components. In this part of the
column negative enhancement factors are obtained for mass transfer with reaction of methylene
glycol and hemiformal. This indicates that approximate methods for calculating mass transfer
enhancement factors due to reaction might fail for absorption with chemical reaction and
simultaneous product desorption. The performance of an industrially applied packed column with
liquid recirculations is simulated using a non-equilibrium stage model. A solution algorithm is
developed and carefully described. The influence of a number of process parameters on the
behaviour of methanolic species in the absorber is investigated.
Introduction
Formaldehyde is industrially produced by partial oxidation and dehydrogenation of
vaporised methanol in air over a solid catalyst at approximately atmospheric pressure. The
reactor product gas stream consists of water vapour, formaldehyde and some unreacted methanol
in an inert matrix of nitrogen, hydrogen and carbon dioxide (minor amounts of by-products are
neglected in this study). This gas mixture is passed through a partial condenser, where the
temperature is reduced to 328 K and part of the water vapour and formaldehyde are condensed.
The resulting stream is subsequently fed to the absorber to extract the formaldehyde from the gas,
and to obtain the commercial product: a concentrated aqueous formaldehyde solution containing
some not-converted methanol.
A scheme of the absorber studied is shown in Fig. 1. The gas stream entering at the bottom
of the column passes upwards trough two packed beds, randomly filled with a modern high
performance Pall-ring like packing. Make-up water enters the top of the column and flows
downward, meanwhile exchanging heat and mass with the gas stream. Each of the absorption
beds is equipped with an external liquid recirculation with heat exchangers. Just below each bed
an amount of liquid is kept to provide buffers for the liquid recirculation pumps.
In a previous paper (Winkelman et al., 1992) a model for the formaldehyde absorber was
developed, based on differential equations for the mass and energy balances in each phase. The
resulting set of coupled boundary value problems was solved by a semi-transient method. The
presence of methanol in the absorber was neglected, and only formaldehyde and water were
assumed to be absorbed or desorbed. In this contribution the behaviour of methanol in the
absorption process is fully incorporated in the model. Also included in the model are vaporisation
and re-absorption of methylene glycol and hemiformal, which are the primary reaction products
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
70
feed
product
tail gas
water
Fig. 1. Scheme of the formaldehyde absorber.
of formaldehyde with water and methanol, and which are formed in the liquid. The absorber is
modelled using a non-equilibrium stage model.
Reactions in aqueous methanolic formaldehyde solutions
Besides heat and mass transfer between gas and liquid, a number of reactions have to be
considered in modelling the performance of formaldehyde absorbers because formaldehyde
reacts with both water and methanol. In aqueous solutions the dissolved formaldehyde (F) reacts
fast with water (W) to form methylene glycol, CH
1
(OH)
1
, denoted by WF
1

2 2 2 2
(OH) CH O H O CH + . (1)
The reaction rate is given by
) / (
1
1 h WF W F h
K C C C k r = , (2)
where K
h
is the chemical equilibrium constant for this hydration, defined as
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
71
eq
W F
WF
h
C C
C
K

=
1
. (3)
The equilibrium of reaction (1) is far to the right, which means that in an aqueous solution the
concentration of free formaldehyde is very low.
Methylene glycol, formed by reaction (1), slowly polymerises to form a series of low
molecular weight poly(oxymethylene) glycols, HO(CHO)
n
H, denoted here by WF
n

O H H O) HO(CH (OH) CH H O) HO(CH
2 n 2 2 2 1 n 2
+ +

) .. 2 (
max
n n = , (4)
with reaction rates
) / (
1 1
n W WF WF WF n n
K C C C C k r
n n
=

) .. 2 (
max
n n = , (5)
and the equilibrium constants
eq
WF WF
W WF
n
C C
C C
K
N
n

=
1 1
) .. 2 (
max
n n = . (6)
If methanol (M) is present in the solution, formaldehyde reacts with it in a similar manner as with
water, producing hemiformal, CH
3
OCH
2
OH (MF
1
)
OH OCH CH OH CH O CH
2 3 3 2
+ , (7)
with the reaction rate
( )
1 1 1
/
1
KM C C C km rm
MF M F
= , (8)
where the equilibrium constant is defined as
eq
M F
MF
C C
C
KM

=
1
1
. (9)
The formed hemiformal also polymerises slowly to a series of polymers, higher hemiformals in
this case, CH
3
O(CH
2
O)
n
H, denoted by MF
n

OH CH H O) O(CH CH OH OCH CH H O) O(CH CH
3 n 2 3 2 3 1 - n 2 3
+ + ) .. 2 (
max
n n = (10)
with reaction rates
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
72
( )
n MF M MF MF n n
KM C C C C km rm
n n
/
1 1
=

) .. 2 (
max
n n = , (11)
and equilibrium constants
eq
MF MF
M MF
n
C C
C C
KM
n
n

=
1 1
) .. 2 (
max
n n = . (12)
Here, n
max
denotes the largest polymer. In Fig. 2 the equilibrium molar fractions are shown of an
aqueous solution containing 55 wt % formaldehyde and 1 wt % methanol. The concentrations of
the higher poly(oxymethylene) glycols and hemiformals decrease rapidly with increasing
molecular weight. Therefore the largest polymers considered here are WF
10
and MF
10
, with n
max
= 10 for both types of polymers. This way, there are 20 chemical reactions in the liquid phase,
and the total number of liquid phase components amounts to 23. The production rates of the
individual species are found from

=
+ =
max
1
, ,
) (
n
k
k k i k k i i
rm m r R , (13)
where
k i,
denotes the stoichiometric coefficient of component i in the reaction forming the k-th
poly(oxymethylene) glycol (negative for reactants and positive for reaction products). Likewise,
k i
m
,
denotes the stoichiometric coefficient in the formation of the k-th hemiformal. From eq
(13) and the stoichiometry of the reactions (1), (4), (7) and (10) it follows
1
10
10
10
-1
-2
-3
x
i
F
W
M
i=1
i=1
2
3
4
5
6
2
3
4
5
6
7
MF
i
WF
i
Fig. 2. Equilibrium molar fractions in an aqueous solution containing 50% by weight
formaldehyde and 1% by weight methanol, at 333 K.
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
73
1 1
rm r R
F
= , (14)
max
2 1 n W
r r r R + + + = K , (15)
max 1
3 2 1
2
n WF
r r r r R = K , (16)
) 2 (
max 1
n n r r R
n n WF
n
< =
+
, (17)
) (
max
n n r R
n WF
n
= = , (18)
max
2 1 n M
rm rm rm R + + + = K , (19)
max 1
3 2 1
2
n MF
rm rm rm rm R = K , (20)
) 2 (
max 1
n n rm rm R
n n MF
n
< =
+
, (21)
) (
max
n n rm R
n MF
n
= = . (22)
The concentration of the unhydrated monomeric formaldehyde is very low due to the
reactions mentioned above: well under 1%, even in concentrated solutions. However, the total
amount of dissolved aldehyde remains available for chemical reactions in downstream processing
because of the reversibility of the reactions.
Vapour-liquid equilibria in formaldehyde-water-methanol mixtures
Several models have been put forward to describe the vapour-liquid equilibria in pseudo
binary formaldehyde-water systems (e.g.: Kogan et al., 1977; Brandani et al., 1980). A method to
model the vapour-liquid equilibria in pseudo ternary formaldehyde-water-methanol mixtures has
been presented by the group of Maurer (Maurer 1986; Albert et al., 2000). This method is used
here to calculate the vapour liquid equilibria at the gas-liquid interface in the absorber.
Several substances in a formaldehyde-water-methanol mixture can vaporise. These are not
only the monomeric formaldehyde, water and methanol, but also the first reaction products of
formaldehyde with water and methanol which are methylene glycol and hemiformal,
respectively. The higher polymers always remain in the liquid phase because of their high boiling
points and negligible vapour pressure (Maurer, 1986). With the method of Maurer, the
thermodynamic equilibrium of the vapour-liquid system is calculated from the overall
composition of the liquid (e.g. weight percentages formaldehyde and methanol) using chemical
equilibrium conditions and overall composition balances in the liquid phase, combined with the
physical equilibria for the components that can vaporise,
) , , , , (
1 1
MF M WF W F i P x P y
s
i i i tot i
= = , (23)
where the activity coefficients,
i
, are calculated with the UNIFAC method (Gmehling et al.,
1982). In the column simulations, however, the liquid phase in a stage is not at chemical
equilibrium. To calculate the gas phase concentrations at the interface, in equilibrium with the
liquid phase concentrations at the interface, we proceed in the following way. The equilibrium
molar fractions are calculated from the overall composition of the liquid according to the method
of Maurer. Once the equilibrium molar fractions are known, the activity coefficients and partial
pressures can be calculated. These partial pressures are then corrected for the deviations of the
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
74
actual liquid phase molar fractions at the interface from those calculated from the chemical
equilibrium conditions. Here it is assumed that the activity coefficients do not vary substantially
with the change in composition from chemical equilibrium to the actual composition at the
interface.
Model development
Simulation of continuous absorption processes is often based on stage models. The column
is assumed to consist of a sequence of stages, each representing a section of the packing. Within
each stage the temperature and composition of the gas and liquid phases are assumed to be
constant. In this context, equilibrium stage models are widely used: the streams leaving a stage
are assumed to be in equilibrium with each other and departures from this assumption are
accounted for by a stage efficiency.
Krishnamurthy and Taylor (1985a,b) pointed out several drawbacks of the equilibrium stage
approach for separation processes. They developed a non-equilibrium stage model, where
material and energy balances for each phase are solved simultaneously with the mass and energy
transfer rate equations. In modelling the formaldehyde absorber, we followed the same approach.
A schematic representation of a non-equilibrium stage is shown in Fig. 3. The packed beds
shown in Fig. 1 consist each of a number of such stages, see Fig. 4. Vapour and liquid streams
from adjacent stages are brought into contact on the stage and are allowed to exchange mass and
energy across their common interface. The model of a stage consists of material and energy
balances for each phase and rate equations for inter-phase mass and energy transfer.
Z
j
H
v
T
g
i
g
j
j
j
H
v
T
g
i
g
j +1
j +1
j +1
H
T
l
l
j -1
j -1
j -1
l
i
H
T
l
l
j
j
j
l
i
H T
l l
j ,F j ,F j ,F
l
i
H T
l l
j ,D j ,D j ,D
l
i J
i ,l
j
q
l
j
q
g
j
J
i ,g
j
A
j
R
i
j
gas liquid
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
75
Fig. 3. Representation of a non-equilibrium stage.
v
i
N+1
v
i
1
l
i
1,F
l
i
1,F
l
i
N,D
l
i
N,D
l
i
N,F
l
i
N
l
i
0
T
T
1
2
T
T
N-1
N
.
.
.
.
B
B
1
2
B
B
N-1
N
.
.
.
.
Fig. 4. Absorber layout for non-equilibrium stage model.
The balance equations are denoted by . The component balance for component i on stage j
reads for the gas phase
0
,
1
,
= +
+ j j
g i
j
i
j
i
j
g i
A j v v , (24)
and for the liquid phase
0
, ,
,
1
,
= +
j D
i
j F
i
j j
i
j j j
i
j
i
j
i
j
i
l l Z SR A J l l
l l l
. (25)
The energy balance on stage j for the gas phase is given by
0
, ,
1
,
1
, ,
= + +

+ +
i
j
g i
j
g i
j j
g
j
i
j
g i
j
i
i
j
g i
j
i
j
g E
H J A q A H v H v , (26)
and for the liquid phase
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
76
0
,
,
, ,
,
,
, ,
1
,
1
, ,
= +




i
j D
i
j D
i
i
j F
i
j F
i
i
j
i
j
i
j j j
i
j
i
j
i
i
j
i
j
i
j
E
H l H l
H J A A q H l H l
l l
l l l l l l
. (27)
With the assumption of constant heat capacities of the components over the temperature range of
adjacent stages, the energy balance eqs (26) and (27) are rewritten in terms of temperatures and
molar flows. For the gas phase this gives
0 ) (
,
1 1
,
= +

+ + j
g
j
i
g i
j
i
j
g
j
g
j
g E
q A Cp v T T . (28)
Similarly, for the liquid phase it follows
( ) 0 ) ( ) (
) ( ) (
,
, ,
,
1 1
,
= +
+



k
k R k k R k
j j
i
i
j F
i
j F j j j
i
i
j
i
j j j
E
Hm rm H r Z S
Cp l T T A q Cp l T T

l l l l l l l l
. (29)
The lowest stage in each of the packed sections represents the buffer for the liquid recirculation
pumps (see Figs 1 and 4). The interfacial area is very small here, and in the model it is assumed
to be zero.
Mass transfer rates
For clarity, the subscript j, indicating the stage number, has been dropped from all symbols
throughout this section. The calculation of the fluxes is based on the two film concept, with the
positive direction defined from the gas to the liquid phase. The gas phase mass transfer rates are
given by
) , , , , ( ) (
1 1 , , , ,
MF M WF W F i C C k J
I
g i g i i g g i
= = , (30)
where the gas phase concentrations at the interface,
I
g i
C
,
, are coupled to those in the liquid phase
by
) , , , , (
1 1 , ,
MF M WF W F i C m C
I
g i i
I
i
= =
l
. (31)
For the calculation of the equilibrium ratios, m
i
, see the section on vapour-liquid equilibria. The
fluxes on either side of the interface are equal:
) , , , , (
1 1 , ,
MF M WF W F i J J
I
i g i
= =
l
. (32)
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
77
At the liquid side of the interface, the diffusional transport of the transferred components is
accompanied by chemical reactions. This causes enhancement of the mass transfer rates. Also,
the fluxes into the liquid bulk,
l , i
J , differ from those at the interface,
I
i
J
l ,
. In a previous paper,
we showed the polymerisation reactions to be too slow to have any influence on the diffusion
fluxes in the film, and the gradients of the concentrations of the higher polymers in the liquid
film to be negligible (Winkelman et al., 1992). So, the hydration of formaldehyde, eq. (1), and
the hemiformal formation, eq. (7), are the only reactions affecting the fluxes of the transferred
components.
To account for these parallel reactions in the liquid phase, the film model is applied
) , , , , , 0 (
1 1 1 1 , 1 1 ,
2
2
MF M WF W F i x rm m r
dx
C d
D
i i
i
i
= = , (33)
with the boundary conditions
I
i
i
i
J
dx
dC
D x
l ,
: 0 = = , (34)
l ,
:
i i
C C x = = . (35)
The set of eqs (30)-(35) can not be solved analytically because of the non-linearity of the
reaction rates in eqs (33). Therefore, an iterative shooting method was used to calculate the
interfacial concentrations and the mass fluxes, given the bulk phase concentrations. From an
initial guess of the interfacial concentrations, the gradients at the interface were calculated with
eqs (30)-(32) and (34), and the differential eqs (33) were numerically integrated from 0 = x to
= x using a fourth order Runge-Kutta method. The interfacial concentrations were repeatedly
updated, using a multi-dimensional Newton-Raphson method, until the obtained concentrations
at = x match the liquid phase bulk concentrations. From the numerically calculated gradients
at = x , the flux into the liquid bulk is obtained
=

=
x
i
i i
dx
dC
D J
l ,
. (36)
The numerical effort of the above procedure is greatly reduced by noting the mutual
dependency of several interfacial concentrations and fluxes. This can be understood from
considering methanol and hemiformal as an example. Addition of eqs (33) for these two
components results in
0
2
2
2
2
1
1
= +
dx
C d
D
dx
C d
D
MF
MF
M
M
. (37)
Integrating twice while applying boundary conditions (34) and (35) gives an explicit relation for
the interfacial concentration of methanol as a function of the interfacial concentration of
hemiformal:
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
78
] / /[ )] (
) / ( [
, , ,
,
,
,
, , , , , , ,
1
1
1
1
1
1 1
M M g M MF
I
MF
MF
MF
I
MF
g MF MF g M M g M M g
I
M
m k k C C k
m C C k C k C k C
+
+ + =
l l
l
l
l
l l l
. (38)
Similar equations are obtained for
I
W
C
l ,
as a function of
I
WF
C
l ,
1
, and for
I
F
C
l ,
as a function of
I
MF
C
l ,
1
and
I
WF
C
l ,
1
. Thus, if the values for
I
WF
C
l ,
1
and
I
MF
C
l ,
1
are chosen or updated, the
values of
I
F
C
l ,
,
I
W
C
l ,
and
I
M
C
l ,
can be calculated from eq (38) and its analogs. This way, the
problem of mass transfer of five components, with two parallel reactions can be solved by
iteration on two interfacial concentrations only.
Energy transfer rates
The film model gives the following expressions for the conductive heat fluxes from the gas
phase,
g
q , and into the liquid phase,
l
q , (Krishna and Taylor, 1986)
) (
,
I
g g f g g
T T A h q = , (39)
) (
l l l
T T h q
I
= . (40)
The Ackermann factor, A
f
, corrects the conductive energy transfer rate in the gas phase for non-
zero mass transfer rates.

=
i
g i g i
g
Cf
f
Cp J
h
Cf
e
Cf
A
, ,
1
where ,
1
. (41)
For the liquid phase, this correction is negligible.
If heat and mass transfer occur simultaneously, the total energy transfer rate contains a
conductive and a convective contribution on either side of the interface. From a balance around
the interface it follows that the total energy fluxes out of the gas phase and into the liquid phase
must be equal

+ = +
i
i i
i
g i g i g
H J q H J q
l l l , , , ,
, (42)
where the summations are over all transferred species. Expressing the enthalpies in terms of heat
capacities and temperature differences, and introducing heats of vaporisation and reaction gives
) )( ( ) )( (
) ( ) (
1 , , , 1 , , ,
, , , , , ,
1 1 1 1
R g MF MF R g WF WF
i
i i
I
i
i vap g i
i
g i g i
I
g g
Hm J J H J J
Cp J T T H J Cp J T T q q
+ +
+ + + =

l l
l l l l
. (43)
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
79
In deriving eq (43), it is assumed that the variation of the heat capacities is negligible if the
temperature changes from T
g
to T
I
and from T
I
to
l
T and that the heats of reaction of both the
hydration of formaldehyde and the hemiformal formation in the film are liberated at the interface.
Further, the reaction products of these two reactions are taken as key components, thus allowing
the reaction rates in the film to be expressed as flux differences of these key components. From
eqs (39)-(43), both the interfacial temperature and the values of the heat transfer rates,
g
q and
l
q , were calculated.
Method of solution.
Newton's method for simultaneous correction was used to solve the model because it is
more effective than tearing algorithms (Krishnamurthy and Taylor, 1985a). The total number of
unknown bulk phase variables on a stage j is 30: 5 gas phase molar flow rates, the gas phase
temperature, 23 liquid phase molar flow rates and the liquid phase temperature. These are stored
in a vector (
j
X )
j MF MF M WF WF W F g MF M WF W F
T
j
T l l l l l l l T v v v v v X ) , , , , , , , , , ( ) (
10 1 10 1 1 1
, , l
K K = . (44)
Other quantities, such as the mass and energy transfer rates and the temperature and
concentrations at the gas-liquid interface, are functions of the bulk phase variables, and are
therefore not considered as independent variables in the solution process. The unknown variables
in (
j
X ) must be found by solving the component balance eqs (24) and (25), and the energy
balance eqs (28) and (29). These equations are ordered in a vector (
j
)
j E MF MF M WF WF W F
g E g MF g M g WF g W g F
T
j
) , , , , ,
, , , , , , ( ) (
, , , , , , , ,
, , , , , ,
10 1 10 1
1 1
l l l l l l l l
K K
=
. (45)
Specified quantities are the molar flow rates and the temperature of the feed to the last stage, the
make-up water molar flow rate and its temperature, the mass flow rates of the two liquid
recirculation streams and the temperatures of the recirculation streams entering the first stage of
each of the two packed beds.
From initial estimates, the variables are repeatedly updated with Newton's method, using the
equation
current current next
) ( ] ) ( ) ][( [ = X X J , (46)
where (X)
current
and (X)
next
denote the current and next estimate for the vector (X) which contains
all bulk phase variables
T
n
T
X X X )) ( ),.., (( ) (
1
= , (47)
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
80
() denotes the vector of current discrepancies of the set of equations to be solved and whose
solution is given by 0 ) ( =
) ) ( ) (( ) (
1
T
n
T T
= K , (48)
and [J] denotes the Jacobian matrix with elements
j
i
j i
X
J


=
,
. (49)
The variables and equations are grouped in such a way that the Jacobian matrix has the well
known block tridiagonal structure, shown in Fig. 5. The top bed of the absorber contributes to the
Jacobian the submatrices [S]
j
, [T]
j
and [U]
j
, which contain the partial derivatives of the functions
pertaining to stage j with respect to the variables of the stages j-1, j, and j+1, respectively. In a
similar way, the submatrices [A]
j
, [B]
j
and [C]
j
originate from the bottom bed in the absorber.
Formally, the submatrices have dimensions (3030), but [S]
j
and [A]
j
are very sparse. The
solution algorithm is adapted to use this sparseness, in order to save on computer storage
requirements and calculation time. Because of the presence of the liquid recirculations, the
composition of the liquid on the first stage of each absorption bed depends partly on the
composition of the last stage. This is reflected in the presence of the two off-diagonal
submatrices R
T
and R
B
in Fig. 5.
Most of the partial derivatives in the submatrices [T]
j
and [B]
j
of the Jacobian are too
complicated to calculate analytically, and are therefore obtained from finite difference
approximations. The entries of all the other submatrices are obtained from analytical expressions.
T
1
S
2
U
1
T
2
U
2
R
T
S
N-1
T
N-1
U
N-1
S
N
T
N
U
N
R
B
A
2
B
2
C
2
A
1
B
1
C
1
A
N-1
B
N-1
C
N-1
A
N
B
N
Fig. 5. Structure of the Jacobian matrix for the formaldehyde absorber.
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
81
Despite the presence of the two off-diagonal submatrices, the matrix generalisation of the
Thomas algorithm can be used to solve eq. (46) for (X)
next
. With this algorithm, the Jacobian
matrix is first converted to the so-called upper diagonal form by a recursive elimination. Most of
the sparsety of the Jacobian is preserved during the elimination process. Additional entries appear
only in the columns between the submatrices [R
T
] and [T
N
], and between [R
B
] and [B
N
]. From the
upper triangular form, the solution for (X)
next
is found by repeated back substitution.
Unfortunately, the solution method described above showed a poor convergence (e.g. large
computation time) or did not result in a solution at all. The latter particularly if the initial
estimates of the variables were not very close to the final solution. The reason for this
unsatisfactory behaviour is not completely clear, but possibly part of the problem is caused by the
great difference in the orders of magnitude of the terms in the energy balances as compared to
those in the mass balances. Therefore, the solution method was changed in such a way that the
energy balance equations and the temperatures where skipped from eq. (46). The remaining
modified eq. (46) now contains only the component balance equations and the molar flow rates,
and is solved in an inner loop
j MF MF M WF WF W F MF M WF W F
T
j
l l l l l l l v v v v v X ) , , , , , , , ( ) (
10 1 10 1 1 1
, ,
loop
inner
K K = (50)
j MF MF M WF WF W F
g MF g M g WF g W g F
T
j
) , , , ,
, , , , , ( ) (
, , , , , , ,
, , , , ,
loop
inner
10 1 10 1
1 1
l l l l l l l
K K
=
. (51)
The inner loop iterations continued until the sum of square residuals of the mass balances, SSR
M
,
satisfied the condition
( )
2 6 2
, ,
2
, ,
] / [ 10 ) ( ) ( s mol SSR
j i
g j i g j i M

< + =

. (52)
Subsequently, the temperatures were found by solving the energy balances in an outer loop using
eq. (46) with (X
j
) and (
j
) now defined as
j g
T
j
T T X ) , ( ) (
loop
outer
l
= , (53)
j E g E
T
j
) , ( ) (
, ,
loop
outer
l
= . (54)
Convergency of the energy balances was supposed to be obtained once the sum of square
residuals of the energy balance equations, SSR
E
, satisfied the condition
( )
2 3 2
,
2
,
] / [ 10 ) ( ) ( s J SSR
j
g E g E E
< + =

. (55)
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
82
The solution algorithm is summarised in Fig. 6. The calculations are initialised by assigning
guessed values to the molar flow rates and to the temperatures at the first stage, and by simply
taking the conditions at the other stages identical to the first stage. Although the initialisation is
not very sophisticated, the algorithm converges readily to a solution defined by eqs (52) and (55).
Once calculated solutions were available for some situations, these were used as starting
values, to initialise the calculations for other sets of operating parameters. This leads to a
considerable reduction of the calculation time.
The transfer coefficients and other physical and most of the chemical properties needed in
the calculations were obtained as described by Winkelman et al. (1992). The chemical
equilibrium constants for reactions involving formaldehyde and methanol where taken from
Maurer (1986). Because no open literature is available on the rates of reactions involving
methanol these rates were taken equal to the rates of the corresponding reactions involving water,
i.e.: km
i
= k
i
(i = 1..n
max
).
START
input parameters
initialize variables
calculate ( ) and [ ] J
for inner loop
update flowrates
SSR <10
M
-6
calculate ( ) and [J]
for outer loop
update temperatures
SSR <1
E
0
3
END
No
o
u
t
e
r

l
o
o
p
i
n
n
e
r

l
o
o
p
No
Fig. 6. Algorithm for computing absorber performance.
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
83
Results
Fig. 7 shows a typical example of the convergence behaviour of the double-loop solution
algorithm. In this figure, the values of SSR
M
calculated with inner loop iterations are plotted
against the values of SSR
E
from the outer loop. The calculations were initialised by using the
same guessed molar flow rates and temperatures for all the stages. To cut calculation time, the
value of SSR
E
was not calculated during the inner loop calculations. In constructing Fig. 7, the
value of SSR
E
during the inner loop iterations was therefore assumed to equal the one calculated
from the next outer loop iteration. It is seen that the number of inner loop iterations per outer
loop iteration reduces steadily from four to just one as the conditions approach the final solution.
Fig. 7 also illustrates that the improvement per individual iteration is greater the closer we are to
the final solution (up to several decades in terms of SSR-values). This is typical for Newton's
method.
The concentrations necessary for the calculations of the mass transfer fluxes from eqs (30)-(38)
can be calculated from the molar flow rates in several ways (Krishnamurthy and Taylor,
1985a,b). The simplest option is to assume constant bulk compositions, and to calculate the
concentrations from the molar flow rates leaving the stage. If the bulk compositions are assumed
to vary linearly, then the concentrations have to be calculated from the average of the molar flow
rates entering and leaving the stage. For the simulation of packed columns, Krishnamurthy and
Taylor (1985c) obtained the best results using the average composition for the bulk vapour
j
g tot
i
j
i
j
i
i
j
i
j
i j
g i
C
v
v
v
v
C
,
1
1
,
5 . 0

+ =

+
+
, (56)
11
0.4
0.02
4x10
-4
2x10
-5
END
START
10
2
10
6
10
10
10
14
10
10
10
10
10
10
10
5
0
-5
-10
-15
SSR /(W)
E
2
SSR
E
(mol/s)
2
Fig. 7. Typical convergence path. Numbers indicate the largest temperature correction
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
84
in either phase with an outer loop iteration step. Dashed lines: inner
loop iterations. Solid lines: outer loop iterations.
20 40 60 80 100
0
0.24
0.23
0.22
0.21
0.20
total number of stages
v
M
out
v
M
in
Fig. 8. Relative amount of vapour phase methanol leaving the bottom-bed vs. the
total number of stages applied; ( ): fluxes based on average vapour
composition; ( ): fluxes based on exit vapour composition.
and using the exit composition for the bulk liquid
j
tot
i
j
i
j
i j
i
C
l
l
C
l l , ,

= . (57)
We, therefore, applied the same method. The influence of the method of calculating the vapour
bulk composition is illustrated in Fig. 8, where, as an example, the relative methanol vapour
phase molar flow rate leaving the bottom-bed is shown as function of the total number of stages
applied. The bulk liquid phase composition was always calculated from the exit liquid molar
flows. Using the stage exit molar flows to calculate the fluxes leads to lower driving forces for
mass transfer, and therefore less absorption of methanol. The effect of the total number of stages
on the solution was studied by varying this number while always allocating 50% to the top- and
the bottom-bed, respectively. The many calculated output variables (molar flow rates and
temperatures) all asymptotically approached a constant value with increasing the number of
stages. Again, Fig. 8 may serve as a typical example.
Generally, the variation in the results appeared to become negligible above 60 stages.
Therefore all remaining calculations were performed with 60 stages (30 in each absorption bed).
A typical set of calculated bulk vapour and liquid temperature profiles is shown in Fig. 9.
The cooler gas stream entering the bottom-bed is quickly heated up to the liquid temperature
because of a large heat transfer capability and a much lower heat capacity of the gas stream
relative to the liquid stream. Similarly, the warm gas stream entering the top-bed from the
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
85
bottom-bed is rapidly cooled down to the liquid temperature level. As the liquid flows
0 1 0.25 0.5 0.75
relative height
65
64
63
62
61
60
44
42
40
38
36
34
T
[C]
T
[C]
Fig. 9. Typical temperature profiles in the absorber; ( ):liquid; ( ):gas.
downwards, its temperature rises, predominantly due to the heat effects connected with
condensation and reaction. In the lower part of the bottom-bed, solvent evaporation results in a
temperature maximum. Interestingly, the temperature profiles obtained here using a non-
equilibrium stage model are quantitatively similar to those obtained previously with a differential
model (Winkelman et al., 1992).
After inspection of the results, it turned out that the Ackermann correction factor for the
vapour phase, A
f,g
, deviates very little from unity. This can be easily understood. For mass
transfer to have a significant influence on the vapour phase conductive energy flux, say more
than 5% (A
f,g
0.95), it follows from eq. (41) that Cf > 0.1. With typical values for the vapour
phase heat capacity Cp
i,g
40 J/(mol K) and the heat transfer coefficient h
g
40 J/(m
2
K s) this
results in a condition for the fluxes: 1 . 0
,

v i
J mol/(m
2
s), which is larger than can be
expected in absorption columns for the solubilities typical in formaldehyde absorption. Further,
the temperature differences in the liquid film,
j j I
T T
l

,
, are generally well below 0.1 K because
of the low resistance against heat transfer in the liquid phase. According to these observations,
the following assumptions are justified: 1 =
f
A and
j j I
T T
l
=
,
. This allows for considerable
simplifications in the calculation of the energy transfer rates because apparently there is no need
to calculate the interfacial temperature. This way, eqs (39)-(43) reduce to
) (
l
T T h q
g g g
= , (58)
) )( ( ) )( (
) (
1 , , , 1 , , ,
, , , ,
1 1 1 1
R g WF WF R g WF WF
i
i vap g i
i
g i g i g g
H J J H J J
H J Cp J T T q q
+ +
+ + =

l l
l l
. (59)
Fig. 10 shows typical profiles of relative vapour phase molar flow rates in the formaldehyde
absorption column. Because of the large liquid/gas ratio, the liquid phase composition does not
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
86
change much with the stage number of the packed section and is therefore not shown in Fig. 10.
v
i
v
i
in
0 1 0.25 0.5 0.75
relative height
2
1
0
i=W
i=F
i=M
Fig. 10. Typical profiles of relative vapour phase molar flow rates.
Similar to the temperature profiles, also these profiles clearly reveal the division of de packed
height in two absorption beds. The gas feed to the absorber is undersaturated with water. In the
lower part of the bottom bed this leads to desorption of water. In the upper part of the bottom-
bed, but even more in the top-bed, the water vapour condenses. This way an internal circulation
of water is created. The mass transfer of both formaldehyde and methanol is always directed
towards the liquid. Again, the profiles obtained here for formaldehyde and water using a non-
equilibrium stage model are very similar to those obtained previously using a differential model.
The relative vapour molar flow rates of the reaction products methylene glycol and
hemiformal are sown in Fig. 11. These components are not present in the gas feed, and similar to
water, they are desorbed in the lower part of the bottom-bed and re-absorbed higher in the
column.
v
i
v
F
0 1 0.25 0.5 0.75
relative height
0.03
0.02
0.01
0.00
i=WF
1
i=MF
1
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
87
Fig. 11. Typical profiles of relative molar flow rates of the vaporised reaction products.
v v
WF MF
1 1
+
v
F
0 1 0.25 0.5 0.75
relative height
0.10
0.05
0.00
v
MF
1
v
M
1.5
1.0
0.5
0.0
Fig. 12. Fraction of converted formaldehyde and methanol in the gas phase
relative to the uncombined, free components.
Fig. 12 shows that the amount of formaldehyde that is transported in the gas phase in associated
form is relatively low compared to the amount of free formaldehyde: it always accounts for less
than 10% of the total amount of formaldehyde in the gas phase. On the other hand, the amount of
hemiformal is substantial compared to the amount of free methanol in the gas phase (in the
bottom-bed it becomes even the larger one of the two). But both the amounts of methanol and
hemiformal are small relative to formaldehyde and water.
The numerical evaluation of the fluxes, as described in the section on mass transfer rates,
allows for the calculation of the mass transfer enhancement factors, E
i

) (
, , ,
0
l l l i
I
i i
x
i
i
i
C C k
dx
dC
D
E

=
=
. (60)
Remarkable results are obtained for the reaction products methylene glycol and hemiformal,
see Fig. 13. The enhancement factor of hemiformal is smaller than one in many of the stages.
Methylene glycol has even negative enhancement factors in some of the stages. This
phenomenon is not some numerical peculiarity, but results from the combined action of mass
transfer and reaction. This becomes clear by taking a closer look at the concentration profiles of
methylene glycol in the stagnant liquid film adjacent to the gas-liquid interface. These profiles
are shown in Fig. 14 for the stages 50 to 54 (the stage numbers are also indicated in Fig. 13). The
profiles were obtained from the numerical method described in the section on mass transfer rates.
In these stages, the positive gradient of the concentration near x = 0 indicates that the flux of
methylene glycol at the interface is directed towards the gas phase. Therefore the numerator of eq
(60) is always negative here. However, the concentration difference of methylene glycol over the
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
88
film in the denominator of eq (60),
l
l
,
, 1
1
WF
I
WF
C C , changes sign in going from stage 52 to 53.
E
WF
1
0 1 0.25 0.5 0.75
relative height
E
MF
1
5
0
-5
-10
-15
-20
2.0
1.5
1.0
0.5
0.0
54
50
Fig. 13. Mass transfer enhancement factors of the reaction products;
( ): methylene glycol; ( ): hemiformal.
C C
WF
1
, WF
1
, l l

mol/m
3
1
0
-1
0 1 0.25 0.5 0.75
x/
stage 50
51
52
53
54
Fig. 14. Concentration profiles of methylene glycol in the film at
the interface for the stages 50-54.
For the stage numbers 53 and 54 the denominator is negative, resulting in positive enhancement
factors, whereas for the stage numbers 50-52 the denominator is positive, resulting in negative
enhancement factors. Recently, the possible occurrence of negative enhancement factors was
shown also by analytically solving the case of mass transfer with a single first-order reversible
reaction (Winkelman and Beenackers, 1993).
In a previous contribution, we already discussed the influence of several operating
parameters on the temperature rise of the liquid in the packed beds, on the formaldehyde
absorption efficiency and on the overall weight percentage of formaldehyde in the liquid leaving
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
89
the packed beds (Winkelman et al., 1992). New aspects here are the role of methanol and
hemiformal. Below, we focus on these aspects only.
Fig. 15 illustrates the influence of the temperature of the liquid recycle stream of the top-
bed,
Rec
t
T . Increasing
Rec
t
T results in higher partial pressures of methanol and hemiformal, and
therefore in increasing relative molar flows of these components in the gas stream leaving the
top-bed and thus in less absorbed methanol in the liquid leaving the bottom-bed. With increasing
Rec
b
T the vapour pressures of methanol and hemiformal in the bottom-bed increase. This causes
increased levels of these components in the top-bed, and higher amounts leaving the absorber
with the gas stream, see Fig. 16. Although the higher amounts of absorbed methanolic
compounds enter the bottom-bed with the liquid flowing down from the top-bed, this cannot
completely compensate for the decreased absorption of methanol and hemiformal in the bottom-
bed and the overall amount of methanol leaving the bottom-bed with the liquid stream decreases.
0.14
0.12
0.10
0.08
0.06
0.04
v
i
out
v
M
in
25 30 35 40 45
T [C]
t
Rec
W
M
[wt%]
2.0
1.9
1.8
1.7
i=M
i=MF
1
bottom
Fig. 15. Influence of
Rec
t
T on the relative amounts of methanolic compounds in
the exit gas and in the exit liquid stream.
0.12
0.10
0.08
0.06
v
i
out
v
M
in
55 60 65 70 75
T [C]
b
Rec
W
M
[wt%]
2.0
1.9
1.8
1.7
i=M
i=MF
1
bottom
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
90
Fig. 16. Influence of
Rec
b
T on the relative amounts of methanolic compounds in
the exit gas and in the exit liquid stream.
0.12
0.10
0.08
0.06
v
i
out
v
M
in
30 40 50 60 70 80 90
R
t
W
M
[wt%]
1.95
1.90
1.85
1.80
i=M
i=MF
1
bottom
Fig. 17. Influence of R
t
on the relative amounts of methanolic compounds
in the exit gas and in the exit liquid stream.
v
i
out
v
M
in
20 25 30 35 40
R
b
W
M
[wt%]
1.95
1.90
1.85
1.80
i=M
i=MF
1
bottom
0.11
0.10
0.09
0.08
0.07
0.06
Fig. 18. Influence of R
b
on the relative amounts of methanolic compounds in
the exit gas and in the exit liquid stream.
The influence of
t
R , the amount of liquid recycled around the top-bed, is illustrated in Fig.
17. The amounts of methanol and hemiformal leaving the absorber with the gas stream decrease
with increasing
t
R , while at the same time the weight percentage of methanol in the liquid
leaving the absorber at the bottom is almost constant. This might seem conflicting, but is caused
mainly by a decrease of the temperature rise of the liquid in the top-bed with increasing
t
R . The
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
91
result is even more formaldehyde and water absorption, and therefore an increase of the total
amount of liquid entering the bottom-bed from the top-bed which tends to dilute the methanolic
species. Similarly, the temperature rise of the liquid in the bottom-bed decreases with increasing
b
R , leading to more formaldehyde and water absorption in the bottom-bed, dilution of the
methanol and lower weight percentages of methanol in the liquid leaving the absorber, see Fig.
18. On the other hand, the increased absorption of formaldehyde and methanol in the bottom-bed
results in lower amounts of these components absorbed in the top-bed. This reduces the liquid
flow in the top-bed, resulting in a higher temperature rise, increased vapour pressures of
methanol and hemiformal, and higher amounts of these components leaving the top-bed with the
gas stream.
Interesting phenomena were observed if the amount of make-up water is varied. In both
absorption beds, the amount of methanol in the gas stream increases with increasing amounts of
make-up water, whereas the amount of hemiformal decreases. This is caused by a shift in the
relative amounts of reaction products in the liquid phase. Due to the lower overall amounts of
formaldehyde present in the liquid, less methanol combines with formaldehyde to form
hemiformal, and more methanol is present in the uncombined form. This results in a higher
partial pressure of methanol and a lower partial pressure of hemiformal. The amounts of these
components leaving the absorber with the gas stream react correspondingly to increasing
amounts of make-up water, see Fig. 19.
v
i
out
v
M
in
0 2 4 6
W
M
[wt%]
2.5
2.0
1.5
1.0
i=M
i=MF
1
bottom
0.12
0.10
0.08
0.06
0.04
relative amount of make up water
Fig. 19. Influence of the relative amount of make-up water on the relative amounts
of methanolic compounds in the exit gas and in the exit liquid stream.
Chapter 6: Simulation of absorbers: the behaviour of methanol and non-equilibrium stage modeling.
92
Conclusions.
The industrially important process of formaldehyde absorption in the presence of methanol is
simulated using a non-equilibrium stage model. The model takes into account the mass transfer
of formaldehyde, water and methanol, and their primary reaction products methylene glycol and
hemiformal. In the liquid a large number of reactions take place, giving rise to numerous liquid
phase components.
The double loop solution algorithm, solving the component balances in an inner loop and
the energy balances in an outer loop, proved to be stable and converged readily despite a simple
initialisation and a large number of equations. The non-zero mass flux correction of the
conductive energy fluxes proved to be negligible. Similarly, it was found that the resistance
against energy transfer in the liquid phase is negligible under practical conditions. In this type of
modelling it is useful to check whether these simplifications are possible, because they reduce the
calculation effort significantly.
The evaporation of water, methylene glycol and hemiformal in the lower part of the column
creates an internal circulation of these components. In this part of the column negative
enhancement factors are obtained for mass transfer with reaction of methylene glycol and
hemiformal. This indicates that approximate methods for calculating mass transfer enhancement
factors due to reaction might fail for absorption with chemical reaction and simultaneous product
desorption.
Chapter 7: Epilogue.
93
Chapter 7
Epilogue
The research reported in this thesis was instigated upon a request from Dynea B.V.
concerning the absorbers in their formaldehyde plants. Since formaldehyde is a bulk chemical, its
cost price is critical, and the production facilities need constant adjustments to increase the
efficiency and to reduce the costs. But despite the need for a remunerative production process,
the absorption columns didn't operate optimally for many years. They were bottlenecking
production levels, hampering efficiency increases, and preventing reduction of the consumption
of methanol, which is the raw material in the production of formaldehyde. Internally, tests were
conducted with the absorbers aimed at getting insight into the influence of various process
parameters on the performance, without much success though. With the lapse of time, some tests
were even repeated, without the results getting any better. It turned out that the number of
possible variations and process parameters was too large, and their influence on the absorber
performance too complicated to be able to achieve an understanding, let alone optimisation, this
way.
The strategy opted for to tackle the absorber problems was the development of reaction
engineering models from first principles, to describe the performance of the columns and the
effects of variations in the operating parameters on the efficiencies.
From an initial survey of the literature, concerning the availability of physical en chemical
data on the reactions in, and properties of aqueous formaldehyde solutions, it was concluded that
experimental work was needed on the kinetics of the main reaction, the hydration of
formaldehyde, and the density and viscosity of the solutions. Consequently, the kinetics of the
dehydration of methylene glycol were measured, the influence, in general, of a reversible
reaction on gas liquid mass transfer was studied theoretically, the kinetics and equilibrium of the
hydration were measured, and the density and viscosity of aqueous formaldehyde solutions were
determined, see Chapters 2, 3 and 4 and Appendix A of this thesis, respectively. Data on the rate
and equilibrium of the relatively slow formation of the series of poly oxymethylene glycols, and
on the vapour liquid equilibria pertaining to formaldehyde containing systems, are readily
available from the literature, especially from the publications by the research group of Maurer
(e.g. Maurer, 1986; Hasse, 1990; Hahnenstein et.al, 1994, 1995; Albert et.al, 2000).
Various reaction engineering models were developed, see Chapters 5 and 6 of this thesis,
and translated to computer code. They were adapted to, p.e., the different absorption column
configurations that were operational. In a first instance, the models were successfully tested for
their ability to describe the current performance of the absorbers. Next, the effect of variations of
the operational parameters on the column performances was investigated. This resulted in a new
optimised set of operational conditions and an increased efficiency of the absorber performance.
Chapter 7: Epilogue.
94
The advantages of the absorber models are threefold. The model predicts exactly the
consequences of any alterations in the tuning, and the process-engineers now have complete
control over the column performance. Because of the higher efficiency of the absorbers, the same
formaldehyde production level can now be achieved with less methanol consumption and
considerable savings. An advantage of a more psychological nature is the ending of internal
discussions for years on absorber operation among the engineers, and the clarity and univocal
nature of the instructions that can now be given to the operators in the plant.
The results obtained here are not just confined to usage in the modelling and optimisation of
formaldehyde absorbers. They were also, for example, successfully applied in the modelling of
industrial formaldehyde-methanol-water distillation, where methanol is separated from
formaldehyde water mixtures. In general, with the availability now of a complete description of
aqueous formaldehyde systems, the results are of use in the design of many separation processes
involving formaldehyde.
Symbols
95
Symbols
a interfacial area, m
-1

f
A Ackermann heat transfer correction factor
j
A interfacial area in stage j, m
2

C concentration, mol m
-3

Cp
i
molar heat capacity of i, J mol
-1
K
-1

D diffusivity, m
2
s
-1

d stirrer diameter, m
E energy flux, W m
-2

E chemical enhancement factor for mass transfer, -
H
i
molar enthalpy of i, J mol
-1

j R
H ) ( reaction enthalpy of reaction j, J mol
-1

i vap
H
,
molar heat of vaporization of i, J mol
-1

h heat transfer coefficient, W m
-2
K
-1

HMS

hydroxymethane sulphonate, CH
2
(OH)SO
3


J molar flux, mol m
-2
s
-1

K chemical equilibrium constant, -
K
a1
, K
a2
acid dissociation constants of H
2
SO
3
and HSO
3

, respectively, in terms of
concentrations, mol m
-3

K
h
chemical equilibrium constant of the formaldehyde hydration, -
K
w
dissociation constant of water in terms of concentrations, (mol m
-3
)
2

k reaction rate constant, m
3
mol
-1
s
-1

1
k first order reaction rate constant, s
-1

k
2
rate constant for the reaction of formaldehyde with
2
3
SO , m
3
mol
-1
s
-1

h d
k k , rate constants for the dehydration of methylene glycol, and the pseudo-first-order
hydration of formaldehyde, respectively, s
-1

g
k gas phase mass transfer coefficient, m s
-1

l
k liquid phase mass transfer coefficient, m s
-1

l optical path length, m
l
i
liquid phase component molar flow rate, mol s
-1

M molar weight, kg mol
-1

m distribution coefficient (
g
C C /
l
, at equilibrium), -
N stirrer rate, s
-1

max
n number of CHO segments in the largest polymers considered
P pressure, Pa
s
i
P saturated vapor pressure of i, Pa
q conductive heat flux, W m
-2

R ideal gas constant, J mol
-1
K
-1

R
i
production rate by reaction of i, mol m
-3
s
-1

Symbols
96
R
t
, R
b
liquid recycle ratio for the top and bottom bed, -
Re Reynolds number, /
2
N d , -
j
r rate of reaction j, mol m
-3
s
-1

S Chapter 4: gas liquid interfacial area, m
2

S cross sectional area of the column, m
2

S
tot
total concentration of sulphur, mol m
-3

Sc Schmidt number, D / , -
Sh Sherwood number,
g g g
D d k / or
l l l
D d k / , -.
E
SSR sum of square residuals of the energy balances, W
M
SSR sum of square residuals of the mass balances, mol s
-1

T temperature, K
T
l
temperature rise of the liquid, K
t time, s
V volume, m
3

V
B
volume of liquid on the partial draw off tray at the bottom of the column, m
3

v diffusivity ratio (=
l l , ,
/
P A
D D in Chapter 3; =
l l , ,
/
MG F
D D in Chapter 4), -
v
i
gas phase molar flow rate of i, mol s
-1

W weight
W
F
overall weight percentage of formaldehyde in the liquid, wt.%
W
M
overall weight percentage of methanol in the liquid, wt.%
x liquid phase molar fraction, -; distance in the film, m
x
~
overall liquid phase molar fraction, -
y gas phase molar fraction, -
Z height of packing, m
z charge number of ionic species, -
j
Z height of stage j, m
Greek letters
activity coefficient
film thickness, m
extinction coefficient, m
2
mol
-1
; dielectric constant of water
l
partial liquid holdup, -
viscosity, Pa s
ionic strength

i,j
stoichiometric coefficient of component i in reaction j, -
density, kg m
3

R
reaction factor, -
g , v
gas phase volumetric flow rate, m
3
s
-1

vector of discrepancies of the set of balance equations
Subscripts
0 initial
2,3,.. polymerization reactions
Symbols
97
A reactant
b bottom bed
F formaldehyde
g gas phase
h hydration reaction
i component i
j reaction j
l liquid phase
M methanol
n
MF hemiformal, CH
2
O(CH
2
O)
n
H,
max
1 n n
MG methylene glycol
max maximum
n number of CH
2
O segments in a polyoxymethylene glycol
P reaction product
t top bed
tot total
W water
WF
n
polyoxymethylene glycol with n CH
2
O segments
irreversible reaction, limit for K
Superscripts

(overbar) bulk conditions


D liquid draw off
F liquid feed
I interface
in incoming or feed conditions
j stage j
out outgoing stream
Rec liquid recirculation stream
sat saturated
th thermodynamic equilibrium constant based on activities
References
99
References
Albert, M., Garca, B.C., Kuhnert, C., Peschla, R., & Maurer, G. (2000). Vapor-liquid equilibrium of
aqueous solutions of formaldehyde and methanol. A.I.Ch.E. J., 46, 1676-1687.
Allan, J.M. & Teja, A.S. (1991). Correlation and prediction of the viscosity of defined and undefined
hydrocarbon liquids. Can. J. Chem. Eng., 69, 986-991.
Astarita, G., & Savage, D.W. (1980). Gas absorption and desorption with reversible instantaneous chemical
reaction. Chem. Engng Sci., 35, 1755-1764.
Bell, R.P. (1966). The reversible hydration of carbonyl compounds. In Adv. Phys. Org. Chem., Gold, V. Ed.,
Vol.4, 1-29, London: Academic Press.
Bell, R.P., Evans, F.R.S., & Evans, P.G. (1966). Kinetics of the dehydration of methylene glycol in aqueous
solution. Proc. Roy. Soc. (London) Ser.A, 291, 297-323.
Bieber, R. & Trmpler, G. (1947). An`genherte spektrographische Bestimmung der Hydratations-
gleichgewichtskonstanten wssriger Formaldehydlsungen. Helv. Chim. Acta, 30, 1860-1865.
Blackadder, D.A., & Hinshelwood, C. (1958). The kinetics of the decomposition of the addition compounds
formed by sodium bisulphite and a series of aldehydes and ketones. Part I & II. J. Chem. Soc. 1958,
2720-2727; 2728-2734.
Bohne, D., Fischer, S. & Obermeier, E. (1984). Thermal conductivity, density, viscosity and Prandtl-
numbers of ethylene glycol-water mixtures. Ber. Bunsen-Ges. Phys. Chem., 88, 739-742.
Bourne, J.R., Von Stockar, U., & Coggan, G.C. (1974). Gas absorption with heat effects. I. A new
computational method. Ind. Eng. Chem. Process Des. Dev., 13, 115-123.
Boyce, S.D., & Hoffmann, M.R. (1984). Kinetics and mechanism of the formation of
hydroxymethanesulfonic acid at low pH. J. Phys. Chem. 88, 4740-4746.
Brandani, V., Di Giacomo, G. & Foscolo, P.U. (1980). Isothermal vapor-liquid equilibria for the water-
formaldehyde system. A predictive thermodynamic model. Ind. Eng. Chem. Process Des. Dev., 19, 179.
Bryant, W.M.D., & Thompson, J.B. (1971). Chemical thermodynamics of polymerization of formaldehyde
in an aqueous environment. J. Polym. Sci. A-1, 9, 2523-2540.
Burnett, M.G. (1982). The mechanism of the formaldehyde clock reaction. J. Chem. Ed. 59, 160-162.
Calvert, J.G., & Pitts, J.N., (1966). Photochemistry. New York: John Wiley & Sons, Inc.
Cancho, J.C., Sanchez, A.M.E., Tena, A.A., & Moreno, M.C. (1989). Evolution and current methods for
obtaining formaldehyde. Chem. Biochem. Eng. Q., 3, 51-55.
Chase, J.D. (1984). The qualification of pure component physical property data. Chem. Eng. Prog., 4, 63-
67.
Cornelisse, R., Beenackers, A.A.C.M., Van Beckum, F.P.H., & van Swaaij, W.P.M. (1980). Numerical
calculation of simultaneous mass transfer of two gases accompanied by complex reversible reactions.
Chem. Engng Sci., 35, 1245-1260.
Dankelman, W., Muizenbelt, W.J., & Leentjes, H. (1988). AKZO Engineering, internal report.
Dasgupta, P.K., DeCesare, K., & Ullrey, J.C. (1980). Determination of atmospheric sulfur dioxide without
tetrachloromercurate(II) and the mechanism of the Schiff reaction. Anal. Chem., 52, 1912-1922.
Daubert, T.E., & Danner, R.D. (1985). Data Compilation Tables of Properties of Pure Compounds. New
York: American Institute of Chemical Engineers.
Deister, U., Neeb, R., Helas, G., & Warneck, P. (1986). Temperature dependence of the equilibrium
CH
2
(OH)
2
+ H
2
SO
3

= CH
2
(OH)SO
3

+ H
2
O in aqueous solution. J. Phys. Chem., 90, 3213-3217.
Devze, D., & Rumpf, P. (1964). tude spectrophotomtrique des solutions aqueuses de gaz sulfureux dans
divers tampons acides. Compt. Rend., 258, 6135-6138.
Feintuch, H.M., & Treybal, R.E. (1978). The design of adiabatic packed towers for gas absorption and
stripping. Ind. Eng. Chem. Process Des. Dev., 17, 505.
Funderburk, L.H., Aldwin, L., & Jencks, W.P. (1978). Mechanisms of general acid and base catalysis of the
reaction of water and alcohols with formaldehyde. J. Am. Chem. Soc., 100, 5444-5459.
Gmehling, J., Rasmussen, P. & Fredenslund, A. (1982). Vapor-liquid equilibria by UNIFAC group
contribution. Revision and extension 2. Ind. Eng. Chem. Proc. Des. Dev., 21, 118.
References
100
Golding, R.M. (1960). Ultraviolet absorption studies of the bisulphite-pyrosulphite equilibrium. J. Chem.
Soc., 3711-3716.
Gruen, L.C., & McTigue, P.T. (1963). Hydration equilibria of aliphatic aldehydes in H
2
O and D
2
O. J.
Chem. Soc., 5217-5223.
Hahnenstein, I., Hasse, H., Kreitler, C.G. & Maurer, G. (1994).
1
H- and
13
C-NMR spectroscopic study of
chemical equilibria in solutions of formaldehyde in water, deuterium oxide, and methanol. Ind. Eng.
Chem. Res., 33, 1022-1029.
Hahnenstein, I., Albert, M., Hasse, H., Kreitler, C.G., & Maurer, G. (1995). NMR spectroscopic study of
reaction kinetics of formaldehyde polymer formation in water, deuterium oxide, and methanol. Ind. Eng.
Chem. Res., 34, 440-450.
Hasse, H. (1990). Dampf-Flussigkeits-Gleichgewichte, Enthalpien und Reaktionskinetik in
formaldehydhaltigen Mischungen. PhD Thesis, Universitt Kaiserslautern, Germany.
Hasse, H., & Maurer, G. (1991). Kinetics of the poly(oxymethylene) glycol formation in aqueous
formaldehyde solutions. Ind. Eng. Chem. Res., 30, 2195-2200.
Hasse, H. & Maurer, G. (1991). Vapor-liquid equilibrium of formaldehyde-containing mixtures at
temperatures below 320 K. Fluid Phase Equilibria, 64, 185.
Hayon, E., Treinin, A., & Wilf, J. (1972). Electronic spectra, photochemistry, and autoxidation mechnism of
the sulfite-bisulfite-pyrosulfite systems. The SO
2

, SO
3

, SO
4

and SO
5

radicals. J. Amer. Chem. Soc.,


94, 47-57.
Himmelblau, D.M., Jones, C.R., & Bischoff, K.B. (1967). Determination of rate constants for complex
kinetic models. Ind. Eng. Chem. Fundam., 6, 539-543.
Hitch, D.M., Rousseau, R.W., & Ferell, J.K. (1986). Simulation of continuous separation processes:
multicomponent, adiabatic absorption. Ind. Eng. Chem. Process Des. Dev., 25, 699.
Hitch, D.M., Rousseau, R.W., & Ferell, J.K. (1987). Simulation of continuous separation processes:
unsteady state multicomponent, adiabatic absorption. Ind. Eng. Chem. Process Des. Dev., 26, 1092.
Iliceto, A. (1954). Sul sistema acque-formaldeide. Nota VI. Equilibri della fasi liquida e gassosa. Gazz.
Chim. Ital., 84, 536-552.
Kelly, R.M., Rousseau, R.W., & Ferrell, J.K. (1984). Design of packed, adiabatic absorbers: physical
absorptuion of acid gases in methanol. Ind. Eng. Chem. Process Des. Dev., 23, 102.
King, C.J. (1980). Separation Processes, 2nd ed., New York: McGraw-Hill.
Kirk-Othmer Encyclopedia of Chemical Technology (1994). 4th ed. Kroschwitz, J.I., Exec. Ed., Howe-
Grant, M., Ed., John Wiley & Sons, New York.
Kogan, L.V., Blazhin, Y.M., Ogorodnikov, S.K., & Kafarov, V.V. (1977). Liquid-vapor equilibrium in the
system formaldehyde-water. Zhur. Prikl. Khim., 50, 2682-2687.
Krishna, R., & Taylor, R. (1986). Multicomponent mass transfer: theory and applications. In: Handbook of
Heat and Mass Transfer, Vol. 2, Chap. 7, pp. 259-432, Houston: Gulf Publ. Company.
Krishnamurthy, R. & Taylor, R. (1985a). A non-equilibrium stage model of multicomponent separation
processes. Part I: Model description and method of solution. A.I.Ch.E. J., 31, 449.
Krishnamurthy, R. & Taylor, R. (1985b). A non-equilibrium stage model of multicomponent separation
processes. Part II: Comparison with experiment. A.I.Ch.E. J., 31, 456.
Krishnamurthy, R. & Taylor, R. (1985c). Simulation of packed distillation and absorption columns. Ind.
Eng. Chem. Process Des. Dev., 24, 513.
Krishnamurty, R., & Taylor, R. (1986). Absorber simulation and design using a nonequilibrium stage
model. Can. J. Chem. Eng., 64, 96.
Landau, J. (1992). Desorption with chemical reaction. Chem. Engng Sci., 47, 1601-1606.
Landqvist, N. (1955). On the polarography of formaldehyde. Acta Chem. Scand., 9, 867-892.
Lee, R.J. & Teja, A.S. (1990). Viscosities of poly(ethylene glycols). J. Chem. Eng. Data, 35, 385-387.
LeHnaff, P. (1963). Sur la vitesse de dshydratation du mthylne-glycol en formaldhyde. Compt. Rend.,
250, 1752-1754.
References
101
Lileev, A.A., Poblinkov, D.B., Lyashchenko, A.K. & Shepotko, M.L. (1982). Study of the effects of some
polar additives on water according to data of various structure-sensitive methods. Deposited Doc.
VINITI 3101-82 (Russian Institute of Scientific and Technological Information).
Los, J.M., Roeleveld, L.F., & Wetsema, B.J.C. (1977). Pulse polarography. Part X. Formaldehyde hydration
in aqueous acetate and phophate buffer solutions. J. Electroanal. Chem. Interfacial Electrochem., 75,
819-837.
Maurer, G. (1986). Vapor-liquid equilibrium of formaldehyde- and water-containing multicomponent
mixtures. A.I.Ch.E. J., 32, 932-948.
Moortgat, G.K., Seiler, W., & Warneck, P.J. (1983). Photodissociation of HCHO in air: CO and H
2
quantum
yields at 220 and 300 K. J. Chem. Phys., 78, 1185-1190.
Mller, R.E., & Schurath, U. (1983). Generation of formaldehyde in test atmospheres with low
concentrations of hydrogen and carbon monoxide. Anal. Chem., 55, 1440-1442.
Nhaesi, A.H. & Asfour, A.A. (1998). Prediction of the McAllister model parameters from pure component
properties of regular binary mixtures. Ind. Eng. Chem. Res., 37, 4893-4897.
Onda, K., Takeuchi, H., & Okumoto, Y. (1968). Mass transfer coefficients between gas and liquid phases in
packed columns. J. Chem. Eng. Japan, 1, 56-62.
Onda, K., Sada, E., Kobayashi, T., & Fujine, M. (1970). Gas absorption accompanied by complex chemical
reactions - I Reversible chemical reaction. Chem. Engng Sci., 25, 753-760.
Otake, T., & Kunigita, E. (1958). Mixing characteristics of irrigated packed towers. Kagaku Kogaku, 22,
144.
Perry, R.H., Green, D.W. & Maloney, J.O. (1984). Perrys Chemical Engineers Handbook, McGraw-Hill,
New York.
Raal, J.D., & Khurana, M.K. (1973). Gas absorption with large heat effects in packed columns. Can. J.
Chem. Eng., 51, 162.
Reid, C.R., Prausnitz, J.M., & Sherwood, T.K. (1977). The Properties of Gases and Liuids. 3rd ed., New
York: McGraw-Hill.
Reid, R.C., Prausnitz, J.M., & Poling, B.E. (1988). The properties of Gases and Liquids. 4th ed., New York:
McGraw-Hill.
Rogers, J.D. (1990). Ultraviolet absorption cross sections and atmospheric photodissociation rate constants
of formaldehyde. J. Phys. Chem., 94, 4011-4015.
Schecker, H.G., & Schulz, G. (1969). Untersuchungen zur Hydratationskinetik von Formaldehyd in
wriger Lsungen. Z. Phys. Chem. NF, 65, 221-224.
Secor, R.M., & Beutler, J.A. (1967). Penetration theory for diffusion accompanied by a reversible chemical
reaction with generalized kinetics. A.I.Ch.E. J., 13, 365-373.
Shah, Y.T., & Sharma, M.M. (1976). Desorption with or without chemical reaction. Trans. Instn Chem.
Engrs, 54, 1-41.
Siling, M.I., & Akselrod, B.Ya. (1968). Spectrophotometric determination of equilibrium constants of the
hydration and protonation of formaldehyde. Russ. J. Phys. Chem., 42, 1479-1482.
Skelding, A.A. & Ashbolt, R.F. (1959). New data for the densiometric determination of methanol in
formalin. Chem. Ind., 1959, 213-218.
Soliman, K. & Marschall, E. (1990). Viscosity of selected binary, ternary and quartenary liquid mixtures. J.
Chem. Eng. Data, 35, 375-381.
Srensen, P.E., & Andersen, V.S. (1970). The formaldehyde-hydrogen sulphite system in alkaline aqueous
solution. Kinetics, mechanisms, and equilibria. Acta Chem. Scand., 24, 1301-1306.
Srivastava, R.K., & Joseph, B. (1984). Simulation of packed-bed separation processes using orthogonal
collocation. Comp. Chem. Eng., 8, 43.
Stilchmair, J. (1972). Untersuchungen zum dynamischen Verhalten einer Absorptions-Bodenkolonne.
Chem. Ing. Techn., 44, 411.
Stockar, U. v., & Wilke, C.R. (1977). Rigorous and short-cut design calculations for gas absorption
involving large heat effects. 1. A new computational method for packed gas absorbers. Ind. Eng. Chem.
Fundam., 16, 88-93.
References
102
Stumm, W., & Morgan, J.J. (1970). Aquatic Chemistry, Wiley, New York.
Sutton, H.C., & Downes, T.M. (1972). Rate of hydration of formaldehyde in aqueous solution. J. Chem.
Soc., Chem. Comm., 1-2.
Tamir , A., Merchuck, J.C., & Virkar, P.D. (1979). Effect of diffusivity on gas-side mass transfer
coefficient. Chem. Engng Sci., 34, 1077.
Treybal, R.E. (1969). Gas absorption and stripping in packed towers. Ind. Eng. Chem., 61, 36.
Valenta, P. (1960). Oszillographische Strom-Spannungs-Kurven III. Untersugung des Formaldehyds in
gepuffertem Milieu. Collect. Czech.. Chem. Commun., 25, 853-861.
Van Swaaij, W.P.M., & Versteeg, G.F. (1992). Mass transfer accompanied with complex reversible
chemical reactions in gas-liquid systems: an overview. Chem. Engng Sci., 47, 3181-3195.
Versteeg, G.F., Blauhoff, P.M.M. & Van Swaaij, W.P.M. (1987). The effect of diffusivity on gas-liquid
mass transfer in stirred vessels. Experiments at atmospheric and elevated pressures. Chem. Engng Sci. ,
42, 1103-1119.
Versteeg, G.F., & Van Swaaij, W.P.M. (1988). Solubility and diffusivity of acid gases (CO
2
and N
2
O) in
aqueous alkanolamine solutions. J. Chem. Eng. Data, 33, 29-34.
Walker, J.F. (1964). Formaldehyde. 3rd ed., ACS Monograph Series, Reinhold, New York.
Weirauch, W. (1999). 2000 to be tough for methanol producers. Hydrocarbon Processing, 78(11), 21.
Westerterp, K.R., Van Swaaij, W.P.M., & Beenackers, A.A.C.M. (1984). Chemical Reactor Design and
Operation. 2nd ed., John Wiley & Sons Ltd, New York.
Winkelman, J.G.M., Sijbring, H., Beenackers, A.A.C.M. & De Vries, E.T. (1992). Modeling and simulation
of industrial formaldehyde absorbers. Chem. Engng Sci., 47, 3785.
Winkelman, J.G.M., & Beenackers, A.A.C.M. (1993). Simultaneous absorption and desorption with
reversible first-order chemical reaction: analytical solution and negative enhancement factors. Chem.
Engng Sci., 48, 2951-2955.
Winkelman, J.G.M., Ottens, M., & Beenackers, A.A.C.M. (2000). The kinetics of the dehydration of
methylene glycol. Chem. Engng Sci., 55, 2065-2071.
Winkelman, J.G.M., & Beenackers, A.A.C.M. (2000). Correlations for the density and viscosity of aqueous
formaldehyde solutions. Ind. Eng. Chem. Res., 39, 557-562.
Winkelman, J.G.M., Voorwinde, O., Ottens, M., Beenackers, A.A.C.M. & Janssen, L.P.B.M. (2002). The
kinetics of the hydration of formaldehyde. Chem. Engng Sci., 57, 4067-4076.
Yadav, G.D. & Sharma, M.M. (1979). Effect of diffusivity on true gas-side mass transfer coefficients in a
model stirred contactor with a plane liquid interface. Chem. Engng Sci., 34, 1423-1424.
Zavitsas, A.A., Coffiner, M., Wiseman, T., & Zavitsas, L.R. (1970). The reversible hydration of
formaldehyde. Thermodynamic parameters. J. Phys. Chem., 74, 2746-2750.
Appendix A: Correlations for the density and viscosity of aqueous formaldehyde solutions.
103
Appendix A
Correlations for the density and viscosity of aqueous
formaldehyde solutions
Abstract
Empirical correlations are presented for the density and viscosity of aqueous formaldehyde
solutions as a function of temperature (T) and overall weight percentage of formaldehyde (W
F
).
Experimental density data from the literature, at T = 288-338 K and with W
F
= 1.6-50 wt %, are
described with an average absolute residual (AAR) of 0.14%. Experimental viscosity data, both
new and from the literature, at T = 288-333 K and with W
F
= 1.6-50 wt %, are described with an
AAR of 1.8%. The residuals of the correlations are free of trending effects as a function of T and
W
F
. It is shown that both properties can be described using liquid mixture correlation methods
from the literature with almost the same accuracy relative to the empirical correlations.
Introduction
Formaldehyde is an important industrial base chemical. One of the key steps in its
production is the absorption of gaseous formaldehyde in water, usually in a packed absorber. The
performance of the absorbers depends on the process operation variables, such as the
temperature, the pressure, and the flow rates, and on the hydrodynamic properties of the packing,
such as the mass-transfer coefficients, the specific interfacial area, and the liquid phase hold-up
in the packing. Therefore, in modelling, design, and optimisation calculations of the
formaldehyde absorbers, the hydrodynamic properties have to be evaluated. In the literature these
parameters are usually correlated to, among other things, the liquid-phase physical properties,
especially the density and viscosity. Also, in the specification sheets of packing manufacturers,
the performance of the packing types is often given as a function of these liquid-phase properties,
along with various flow-rate parameters.
Two literature sources were found giving correlations for the density of aqueous
formaldehyde solutions,
m
. Walker (1964) gives a correlation for
m
as a function of the strength
of the solution, W
F
, which is valid at 291 K only,
) 291 ( 3 10 00 . 1 ) (
3
291
K T W
F K m
= + = , (1)
and the temperature coefficients in the range of T = 288-303 K for W
F
= 15 and 45 wt %, from
which, using eq (1), the following correlations can be obtained:
) 303 288 %, 15 ( ) 291 ( 2 . 0 1045 ) (
% 15
K T wt W T
F wt m
= = + = , (2)
) 303 288 %, 45 ( ) 291 ( 4 . 0 1135 ) (
% 45
K T wt W T
F wt m
= = + = . (3)
The Kirk-Othmer Encyclopedia of Chemical Technology (1994) presents a correlation for

m
, which reads (slightly modified to yield consistent units)
Appendix A: Correlations for the density and viscosity of aqueous formaldehyde solutions.
104
)] 328 ( 10 55 . 0 0 . 1 )][ 45 ( 3 1119 [
3
T W
F m
+ + =

. (4)
No information is given on the accuracy of eq (4), nor on the temperature and concentration
range for which it is valid. The same source also presents a correlation of the viscosity of
aqueous formaldehyde solutions,
m
, which slightly modified reads
)] 15 . 273 ( 024 . 0 039 . 0 28 . 1 [ 10
3
+ =

T W
F m
. (5)
Equation (5) is valid for rather concentrated solutions, W
F
= 30-50 wt %, and for T = 298-313 K.
No information is given on the accuracy of eq (5).
In this contribution the results of a study to correlate the available literature data on
m
and

m
as a function of T and W
F
are reported. Also, the results of a series of viscosity measurements
of aqueous formaldehyde solutions are reported.
Density
Three literature sources were found reporting data on
m
; see Table 1. Lileev et al. (1982)
specified the strength of the solutions in terms of the overall formaldehyde molar fraction,
F
x
~
,
from which we calculated W
F
, because these units were used by the other authors mentioned in
Table 1:
% 100
)
~
1 (
~
~

+
=
W F F F
F F
F
M x M x
M x
W . (6)
The experimental results show that
m
varies with T and W
F
and that always
W m
> (for
0 >
F
W ). Fig. 1 shows the density difference ) (
W m
as a function of W
F
, with
W
from
Perry et al. (1984) It shows that a considerable fraction of the observed variation of
m
can be
accounted for by introducing a linear dependency of ) (
W m
on W
F
. Least-squares
regression of the data accordingly, followed by an analysis of the residuals,
i
, defined as
% 100
) (
) ( ) (
exp
exp


=
i
m
m calc m
i


, (7)
showed that the residuals have a clear trend as a function of T, indicating an inadequacy in the
relation which makes extrapolation unreliable outside the applied experimental conditions. The
residuals tend to increase monotonically with increasing T, justifying the introduction of an
additional temperature-dependent parameter. Using multiple regression the following equation
was thus obtained:
F W m
W T) 10 8166 . 6 0950 . 5 (
3
+ = . (8)
Appendix A: Correlations for the density and viscosity of aqueous formaldehyde solutions.
105
Table 1. Literature data on the density of aqueous formaldehyde solutions.
Ref. T (K) W
F
(wt %)
m
(kg m
-3
) data points
1 291-338 2-50 1005.4-1570.0 27
3 288, 298 6-43 1018.4-1135.4 16
4 288, 298, 308 1.6-17 997.4-1045.4 15
all
data
288-338 1.6-50 997.4-1570.0 58
0 20 40 60
160
120
80
40
0

m
W



(
k
g
/
m
)
3
W (wt%)
F
Walker
Skelding & Ashbolt
Lileev et al.
Fig. 1. Density difference between aqueous formaldehyde solutions and water,
both at the same temperature.
m
: from the literature sources indicated.

W
: Perry et al. (1984).
When eq (8) was applied to a 15 wt % solution, at 288 T 303 K, the difference with eq (2)
(Walker, 1964) was always less than 0.1%. Similarly, with a 45 wt % solution, at the same
temperatures, the difference between eqs (8) and (3) (Walker, 1964) was no more than 0.3%.
For eq (8) an average absolute residual (AAR) of 0.14% was found, with a maximum
absolute residual (MAX) of 0.69%. The AAR is calculated from

=
=
n
i
i
n
1
1
AAR . (9)
More importantly, however, the residuals obtained with eq (8) do not show any systematic
variation with T or W
F
; see Figs 2 and 3. Therefore, eq (8) is a reliable empirical equation for
m
.
With eq (4) (Kirk-Othmer Encyclopedia of Chemical Technology, 1994) an AAR of 0.42%
(MAX of 1.8%) was observed, which is 3 times as high as the value of eq (8).
Appendix A: Correlations for the density and viscosity of aqueous formaldehyde solutions.
106
280 300 320 340
T (K)
r
e
l
a
t
i
v
e

r
e
s
i
d
u
a
l
s

(
%
)
4
2
0
-2
-4
Fig. 2. Relative residuals of the empirical density correlation (8) as a function of T.
Symbols: see Fig. 1.
0 20 40 60
W (wt%)
F
r
e
l
a
t
i
v
e

r
e
s
i
d
u
a
l
s

(
%
)
4
2
0
-2
-4
Fig. 3. Relative residuals of the empirical density correlation (8) as a function of W
F
.
Symbols: see Fig. 1.
Many literature methods for the calculation of liquid mixture densities use the critical
properties and acentric factors, i.e., vapour pressure vs. temperature correlations, of the
individual components (p.e. Reid et al., 1988) and are therefore not suitable here because the
required properties of the higher poly(oxymethylene) glycols (POMs) are unknown. Although
Amagats law originally holds strictly only for mixtures of ideal gases, it is also recommended
for the calculation of liquid densities of mixtures of similar components (Perry et al., 1984),

=
i
i i m
V x V . (10)
Appendix A: Correlations for the density and viscosity of aqueous formaldehyde solutions.
107
To apply eq (10), the composition of the liquid has to be considered. In aqueous solutions,
formaldehyde is hydrated to methylene glycol and a series of POMs:
2 2 2 2
(OH) CH O H O CH + , (11)
) .. 1 ( O H O) HO(CH (OH) CH O)H HO(CH
2 1 i 2 2 2 2
= + +
+
i . (12)
Methylene glycol and the POMs only exist in formaldehyde solutions. They cannot be isolated in
a pure form, and their pure-component properties cannot be measured directly.
The equilibrium of eq (11) is far to the right and the concentration of free formaldehyde in
aqueous solutions is negligible compared to those of methylene glycol and the higher POMs.
Then, for the formaldehyde-water system, Amagats law can be written as

=
+ + =
1
) (
i
WFi F W W W m
x iV V x V V , (13)
where it is assumed that the molar volumes of the POMs can be written as the sum of the
volumes of the constituent groups. The subscript WFi denotes HO(CH
2
O)
i
H, i.e., the component
consisting stoichiometrically of water and i formaldehyde units.
With the molar balance
1
1
= +

= i
WFi w
x x , (14)
the overall formaldehyde balance

=
+ +
=
1
1
) 1 (
~
i
WFi W
i
WFi
F
x i x
ix
x , (15)
and the substitution / M V = , eq (13) can be rewritten as
F
F
F
W
W
m
m
V
x
x M M
~
1
~

+ =

. (16)
Because every molecule in the solution is either a free water molecule or a water molecule
chemically bonded to one or more formaldehyde units, the true total concentration in the solution
is equal to the overall water concentration. Therefore, the true mean molar weight of the solution,
M
m
, can be obtained as
) 100 / ( 1
F
W
m
W
M
M

= , (17)
Appendix A: Correlations for the density and viscosity of aqueous formaldehyde solutions.
108
and the model equation for the density of aqueous formaldehyde solutions, from inserting eqs (6)
and (17) in eq (16) and rewriting, becomes
F F W F F
F W
m
V W M W
M

+
=
) 100 (
100
. (18)
The only parameter in eq (18) to be determined from the experimental data is the molar volume
of the CH
2
O groups in the POM molecules, V
F
. By taking V
F
constant, an AAR of 0.45% was
obtained (MAX of 1.7%). Not surprisingly, however, taking V
F
constant resulted in a clear trend
of the residuals of eq (18) as a function of T, varying in the expected direction, i.e., from negative
values at the lower temperatures to positive values at the higher temperatures.
Because of the clear trend of the residuals, a second parameter to account for the influence
of T on V
F
seems justified. Least-squares analysis of the experimental data according to eq (18)
resulted in the following optimum parameters for V
F
:
T V
F
6 3
10 59 . 30 10 709 . 12

+ = . (19)
Figs 4 and 5 illustrate the relative residuals of
m
calculated with eqs (18) and (19) as a function
of W
F
and T. No trend in the residuals was found. Here, an AAR of 0.22% was found (MAX of
0.69%).
At first glance, eqs (8) and (18) might seem paradoxical: eq (8) correlates
m
linearly with
W
F
, while eq (18) correlates 1/
m
similarly. This is not a true inconsistency because the
coefficient of W
F
is positive in eq (8), resulting in an increase of
m
with an increase of W
F
,
whereas the overall coefficient of W
F
in the denominator of eq (18) is negative, giving the same
direction of variation of
m
with W
F
.
280 300 320 340
T (K)
r
e
l
a
t
i
v
e

r
e
s
i
d
u
a
l
s

(
%
)
4
2
0
-2
-4
Fig. 4. Relative residuals of the density correlation obtained from Anmagats law (eq 18)
as a function of the temperature. Symbols: see Fig. 1.
Appendix A: Correlations for the density and viscosity of aqueous formaldehyde solutions.
109
0 20 40 60
W (wt%)
F
r
e
l
a
t
i
v
e

r
e
s
i
d
u
a
l
s

(
%
)
4
2
0
-2
-4
Fig. 5. Relative residuals of the density correlation obtained from Anmagats law (eq 18)
as a function of W
F
. Symbols: see Fig. 1.
Viscosity
Table 2 summarizes literature data on the viscosity of aqueous formaldehyde solutions.
Because this data set is rather limited, we performed additional viscosity measurements with a
Schott automated viscosity meter (described in more detail by Soliman & Marschall, 1990) The
solutions were prepared by dissolving a desired amount of paraformaldehyde (Janssen Chimica)
in distilled water. By keeping high efflux times (120-360 s), the error due to kinetic energy was
assumed negligible. Although the vapour pressure of pure formaldehyde at the highest
temperature of the measurements, 325 K, is more than 1.1 MPa (Reid et al., 1988), its
concentration is so low because of the reactions (11) and (12) that the formaldehyde vapour
pressure over a 33 wt % solution is only approximately 1 kPa (Maurer, 1986). Thus, the
influence of possible evaporation of formaldehyde on the measurements is neglected. The
viscometer was calibrated at each temperature using pure water. The absolute viscosity was
determined from the measured kinematic viscosity using the density obtained from eq (8). The
results are shown in Table 3, where each data point is the mean of three measurements whose
flow times were within 0.15 s. The total uncertainty of the viscosity data was estimated to be
1.5%.
Table 2. Literature data on the viscosity of aqueous formaldehyde solutions.
T (K) W
F
(wt %)
3
10
m
(Pa s)
data points
1 298, 333 5-50 0.54-1.87 16
4 288, 298, 308 1.6-17 0.7487-1.6086 15
all
data
288-333 1.6-50 0.54-1.87 31
Appendix A: Correlations for the density and viscosity of aqueous formaldehyde solutions.
110
Table 3. New experimental data on the viscosity of aqueous formaldehyde solutions.
3
10
m
(Pa s)
W
F
(wt %) T = 297.85 K T = 307.15 K T = 318.05 K T = 325.25 K
5 1.0295 0.8417 0.6830 0.6011
15 1.2853 1.0537 0.8470 0.7465
25 1.6377 1.3365 1.0619 0.9279
33 2.0456 1.6483 1.3129 1.1277
Over a wide temperature range, the logarithm of the kinematic viscosity, /, uses to
correlate linearly with 1/T for pure liquids (Reid et al., 1988). This appears also to hold for
formaldehyde solutions for a constant W
F
. The influence of the composition could be accounted
for by correlating ) / ln(
m m m
M linearly both to 1/T and W
F
. Finally, from an analysis of the
residuals it was found that an additional term, linearly with T, was needed to obtain a correlation
free of trending effects of the residuals. The empirical correlation developed this way is
T W
T M
F
m m
m
0404 . 0 10 36 . 9
5644
90 . 47 ln
3
+ + + =

, (20)
with an AAR of 1.8% (MAX of 7.5%) for 288 T 333 K.
m
and M
m
are obtained from eqs (8)
and (17), respectively. Equation (20) is illustrated in Fig. 6. The residuals of eq (20) did not show
any clear trend as a function of W
F
or T; see Figs 7 and 8.
0 20 40 60
W (wt%)
F
100
80
60
40
20
Walker
Lileev c.s.
Perry c.s.
this work


M
m
m
m
x
1
0
9
k
m
o
l
.
m
k
g
.
s
2
(
)
Fig. 6.
m
/
m
M
m
as a function of W
F
for various T. Symbols:
m
from the sources indicated,

m
and M
m
from eqs (8) and 17, respectively. Lines:
m
/
m
M
m
calculated with the
empirical viscosity correlation (20).
Appendix A: Correlations for the density and viscosity of aqueous formaldehyde solutions.
111
0 20 40 60
W (wt%)
F
20
10
0
-10
-20
r
e
s
i
d
u
a
l

(
%
)
Fig. 7. Relative residuals of the empirical viscosity correlation (20) as a function of W
F
.
Symbols: see Fig. 6.
20
10
0
-10
-20
r
e
l
a
t
i
v
e

r
e
s
i
d
u
a
l

(
%
)
280 300 320 340
T (K)
Fig. 8. Relative residuals of the empirical viscosity correlation (20) as a function of T.
Symbols: see Fig. 6.
In addition to eq (20), we also tested an Antoine-type of temperature dependency,
augmented with a linear term in W
F
, i.e., ln(
m
/
m
M
m
) = p
1
+p
2
/(T+p
3
)+p
4
W
F
. After optimization
of the parameters using nonlinear regression, the same AAR (1.8%) was observed; however,
MAX was somewhat larger (8.6%) as compared to eq (20).
Appendix A: Correlations for the density and viscosity of aqueous formaldehyde solutions.
112
The methods found in the literature for obtaining the viscosity of liquid mixtures are often
based on the mole fraction average of the logarithms of the pure-component viscosities, extended
with various types of correction factors (Perry et al., 1984; Reid et al., 1988). Applying mole
fraction averaging to the formaldehyde-water system gives

=
+ =
1
ln ln
i
WFi WFi W w m
x x . (21)
In the literature it is shown that for various homologous series the logarithm of the pure-
component viscosities varies linearly with the molecular size (p.e. Chase, 1984; Allan & Teja,
1991; Nhaesi & Asfour, 1998). This concept cannot be tested directly for methylene glycol and
the POMs, because they cannot be obtained in pure form.
However, experimental viscosity data are available for the closely related series of ethylene
glycol and the poly(ethylene glycols) HO(CH
2
CH
2
O)
i
H or PEG
i
. Here, we will use these data just
to illustrate the concept before returning attention to the aqueous formaldehyde solutions. We
found that for 294 T 333 K the viscosities of PEG
i
can be described by
ib a
i
PEG
+ = ln , (22)
with T a / 3406 60 . 15 + = and T b / 8 . 132 1925 . 0 + = . Fig. 9 shows experimental data of the
viscosities of PEG
i
(i=1..6) and the straight lines calculated with eq 22. Although the viscosity of
the monomer, ethylene glycol, deviates somewhat, the overall agreement is satisfactory
considering the simplicity of the correlation.
10
10
10
-1
-2
-3
1 2 3 4 5 6
molecular size i


(
P
a
.
s
)
Lee & Teja
Bohne c.s.
Fig. 9. Viscosity of poly(ethylene glycols) as a function of the molecular size i at various T.
Symbols: experimental data from the sources indicated.
Lines: calculated with eq (22).
Appendix A: Correlations for the density and viscosity of aqueous formaldehyde solutions.
113
Assuming this concept also applies to the series of methylene glycol and the higher POMs gives
iB A
WFi
+ = ln . (23)
With eqs (14), (15) and (23), eq (21) can be rewritten as
B
x
x
A x x
F
F
W W W m
~
1
~
) 1 ( ln ln

+ + = . (24)
Least-squares regression of the experimental data to eq (24) resulted in the following parameter
values:
T A / 7174 97 . 17 = , T B / 5048 72 . 14 + = . (25)
The true molar fraction of water, x
W
, in the solutions was calculated by solving the equilibrium
equations for the reactions (12)
2
2
1
2
K
x
x x
WF
W WF
= , (26)
) 3 (
3
1 1
=

i K
x x
x x
WF WFi
W WFi
, (27)
simultaneously with the balances (14) and (15), where
F
x
~
was obtained from eq 6. The
equilibrium constants K
2
and K
3
for the formaldehyde-water system were taken from
Hahnenstein et al. (1994)
0 20 40 60
W (wt%)
F

m
x
1
0
(
P
a
.
s
)
3
2
1
0.5
Fig. 10. Viscosity of aqueous formaldehyde solutions. Symbols: experimental data,
see Fig. 6. Lines: calculated with eq (24).
Appendix A: Correlations for the density and viscosity of aqueous formaldehyde solutions.
114
The accuracy of eq (24) is comparable to that of eq (20) (AAR of 2.0% and MAX of 7.6%).
The correlation is illustrated in Fig. 10. The residuals do not show any trend as a function of T or
W
F
as shown in Figs (11) and (12).
When applied to all of the data, the errors of eq (5) (Walker, 1964) for the viscosity of
aqueous formaldehyde solutions were large (AAR of 15.6% and MAX of 93%). When only the
data within the ranges of W
F
= 30-50 wt % and T = 298-313 K were considered, an AAR of 3.5%
(MAX of 6.2%) was obtained, thereby demonstrating the more limited applicability of eq (5).
0 20 40 60
W (wt%)
F
20
10
0
-10
-20
r
e
s
i
d
u
a
l

(
%
)
Fig. 11. Relative residuals of the viscosity correlation (24) as a function of W
F
.
Symbols: see Fig. 6.
20
10
0
-10
-20
r
e
l
a
t
i
v
e

r
e
s
i
d
u
a
l

(
%
)
280 300 320 340
T (K)
Fig. 12. Relative residuals of the viscosity correlation (24) as a function of T.
Symbols: see Fig. 6.
Appendix A: Correlations for the density and viscosity of aqueous formaldehyde solutions.
115
Conclusions
The density and viscosity of aqueous formaldehyde solutions can be accurately and reliably
obtained as a function of the temperature and the strength of the solution with the simple
empirical correlations obtained here. The empirical density correlation (eq 8) employs two
adjustable parameters that were optimised using three literature sources of density data. The
empirical viscosity correlation (eq 20) has four coefficients that were optimised using two
literature data sources together with a series of new additional measurements.
The residuals of the correlations presented are free of trending effects as a function of both
the temperature and the weight percentage of formaldehyde. Therefore, we conclude that the
correlations can be used reliably in engineering calculations with a small extrapolation to cover
the entire range of conditions prevailing in formaldehyde absorbers, i.e., 280 T 340 K and 0
W
F
60 wt %.
A mixture density correlation method from the literature, where the molar volume of the
mixture is obtained as the molar fraction average of the pure-component molar volumes,
appeared to represent the data with almost the same accuracy. In this case two coefficients were
fitted to the data, to correlate the molar volume of the CH
2
O groups linearly to the temperature.
Similarly, a literature method for liquid mixture viscosities, where the logarithm of the
pure-component viscosities are molar fraction averaged, resulted in almost the same accuracy
relative to the empirical correlation. In this case, it was assumed that the logarithm of the
viscosities of the homologous series of methylene glycol and the higher POMs varies linearly
with the molecular size of the components. This way, the molar fraction average method contains
two temperature-dependent parameters, i.e., four adjustable coefficients.
At first glance it seems surprising that the empirical relations for the density (eq 8) and
viscosity (eq 20) both result in somewhat lower AAR values as compared to the relations that
were arrived at starting from methods found in the literature (eqs 18, 19 and 24, respectively),
even though in both cases the same number of coefficients were adjusted to the experimental
data. This may reflect, however, the difficulties still encountered at present in the development of
theory applicable to estimating liquid mixture properties.
Appendix B: Equilibrium molar fractions in aqueous methanolic formaldehyde solutions
117
Appendix B
Equilibrium molar fractions in aqueous methanolic formaldehyde
solutions
In aqueous mixtures of formaldehyde and methanol, the equilibrium composition is determined
by a series of reactions:
2 2 2 2
(OH) CH O H O CH + , (1)
O H H O) HO(CH (OH) 2CH
2 2 2 2 2
+ , (2)
) 3 (i O H H O) HO(CH (OH) CH H O) HO(CH
2 i 2 2 2 1 - i 2
= + + L , (3)
OH OCH CH OH CH O CH
2 3 3 2
+ , (4)
) 2 (i
OH CH H O) O(CH CH OH OCH CH H O) O(CH CH
3 i 2 3 2 3 1 - i 2 3
=
+ +
L
. (5)
The equilibrium conditions, in terms of molar fractions, for the reactions read
W F
WF
x x
x
K
1
1
= , (6)
2
2
1
2
WF
W WF
x
x x
K = , (7)
) 3 (
1 1
3
= =

L i
x x
x x
K
WF WF
W WF
i
i
, (8)
M F
MF
x x
x
KM
1
1
= , (9)
) 2 (
1 1
2
= =

L i
x x
x x
KM
MF MF
M MF
i
i
. (10)
In addition, three overall balances combine the overall molar fractions,
F
x
~
,
W
x
~
and
M
x
~
, with
the true molar fractions in the mixture:

=
+ +
+ +
=
1 1
1 1
1
~
i
MF
i
WF
i
MF
i
WF F
F
i i
i i
ix ix
ix ix x
x , (11)
Appendix B: Equilibrium molar fractions in aqueous methanolic formaldehyde solutions
118

=
+ +
+
=
1 1
1
1
~
i
MF
i
WF
i
WF W
W
i i
i
ix ix
x x
x , (12)

=
+ +
+
=
1 1
1
1
~
i
MF
i
WF
i
MF M
M
i i
i
ix ix
x x
x . (13)
The set of eqs (6)-(13), in principle, provides enough information to calculate all molar
fractions. However, we have an infinite number of eqs (8) and (10), and the summations in (11)-
(13) have no upper limit. These problems can be overcome by using some simple properties from
the theory of power series. Thus, by substitution of the molar fractions of the reaction products
obtained from (7)-(10), the summations in (11)-(13) can be written as
1
1
1
WF
i
WF
x S x
i
=

=
,
v K
v K
S
3
2
1
1
1

+ = , (14)
1
2
1
WF
i
WF
x S ix
i
=

=
,
2
3
2
3
2
2
) 1 ( 1
1
v K
v K
v K
v K
S

+ = , (15)
M
i
MF
x Sm x
i
1
1
=

=
,
v KM KM K
v KM
Sm
2 1 1
1
1

= , (16)
M
i
MF
x Sm ix
i
2
1
=

=
,
2
2 1 1
1 1
2
) ( v KM KM K
v K KM
Sm

= . (17)
where the quantities S
1
, S
2
, Sm
1
and Sm
2
are introduced for ease of notation. With (14)-(17) the
overall balances, eqs (11)-(13) can be rewritten as
) 1 ( )
~
1 (
) (
~
) 1 )(
~
(
1 2 1
1 2 1 1
1
Sm S x K
K v Sm x Sm v x K
x
F
M F
WF
+
+ +
= , (18)
2 1
2
~
1
) 1 (
~
1
Sm x Sm
x S x
x
M
WF M
M
+
+
= , (19)
M W WF W W W
x Sm x x S S x x x
2 1 2
~
)
~
(
~
1
+ + = , (20)
where v is defined as
W
WF
x
x
v
1
= . (21)
Appendix B: Equilibrium molar fractions in aqueous methanolic formaldehyde solutions
119
The set of eqs (18)-(21) can easily be solved for
1
WF
x ,
M
x and
W
x by iteration on v, where
v is limited to 1 0 < v . The other molar fractions,
F
x , ) 2 ( i x
i
WF
and ) 1 ( i x
i
MF
, can be
obtained from (6)-(10) straight forward. Figure 1 illustrates the smooth variation of the ratio v
with the overall formaldehyde molar fraction. Here,
1
K was taken from Winkelman et. al (2002),
2
K ,
3
K and
2
KM were taken from Hahnenstein et. al (1995), and
1
KM was obtained by
multiplying
WM
K (Hahnenstein et. al, 1995) and
1
K , where
WM
K the equilibrium constant is of
the reaction O H OH OCH CH OH CH (OH) CH
2 2 3 3 2 2
+ = + .
0.0 0.1 0.2 0.3 0.4 0.5
0.20
0.16
0.12
0.08
0.04
0.00
x
~
F
v
T
Fig. 1. Variation of the ratio v with
F
x
~
at temperatures of 300, 320
and 340 K. Solid lines:
0
~
=
M
x
; dotted lines:
05 . 0
~
=
M
x
.
Simplifications
1. Free formaldehyde is not important
If the very small molar fraction of free formaldehyde is not important, then eq (18) reduces to
) 1 )(
~
1 (
~
) 1 (
~
1 2
2 1
1
Sm x S
x Sm Sm x
x
F
M F
WF
+
+
= . (22)
The system now consists of eqs (19)-(22), and can be solved in the same way as before, by
iteration on v.
Appendix B: Equilibrium molar fractions in aqueous methanolic formaldehyde solutions
120
2. No methanol present
If the mixture does not contain any methanol then of course
M
x
~
,
M
x and ) 1 ( i x
i
MF
all are
zero, and eqs (18) and (20) for obtaining
1
WF
x and
W
x reduce to
2 1
1
)
~
1 (
)
~
(
1
S x K
v x K
x
F
F
WF


= , (23)
1
)
~
(
~
1 2 WF W W W
x S S x x x + = . (24)
For this case, Fig. 2 illustrates the relative amount of methylene glycol with increasing overall
formaldehyde content in aqueous solutions. The figure shows that at low concentrations, say
below 1 mmol/l, virtually all the formaldehyde is present as methylene glycol, and the amount of
poly oxymethylene glycols is negligible.
0.0 0.1 0.2 0.3 0.4 0.5
1.0
0.8
0.6
0.4
0.2
0.0
x
~
F
T
x
~
F
x
WF
1
Fig. 2. The relative amount of methylene glycol at
temperatures of 300, 320 and 340 K.
If also the very small molar fraction of free formaldehyde is not important, p.e. in the calculation
of the viscosity (see Appendix A), then the ratio v can be obtained from the cubic equation
0
4 3
2
2
3
1
= + + + a v a v a v a , (25)
with the coefficients
)
~
2 1 )( (
2 3 3 1 F
x K K K a = , (26)
) )( 2
~
3 ( ) 1 (
~
2 3 3 3 2
K K x K x K a
F F
+ = , (27)
) 1 (
~
2 1
3 3
K x a
F
= , (28)
F
x a
~
4
= . (29)
Appendix B: Equilibrium molar fractions in aqueous methanolic formaldehyde solutions
121
The physically significant root of eq (24) can easily be identified: either the cubic has only one
real root, or the cubic has only one root in the correct region, i.e., 1 0 < v . The true molar
fractions of water and methylene glycol are now obtained from
) 2 ( ) 1 (
) 1 ( ) 1 (
~
1
~
1
3 2
2
3
3 2
2
3
v K v K v K
v K v K v K
x
x
x
F
F
W
+
+

= , (30)
v x x
W WF
=
1
. (31)
A further remark: if the overall molar fractions of formaldehyde and water are exactly equal, i.e.
5 . 0
~ ~
= =
W F
x x , then the cubic equation (25) degenerates to a quadratic one in v, and true molar
fractions of water and methylene glycol can be obtained directly from
3 2
2 1
1
1
K K
x
WF
+ +
= , (32)
) 1 ( 1
2
1
K x x
WF W
+ = . (33)
Appendix C: The reaction order of formaldehyde in the hydration.
123
Appendix C
The reaction order of formaldehyde in its hydration reaction.
Method
The measurements described in Chapter 4 can be used to obtain the reaction order of
formaldehyde in the hydration, as well as the reaction rate constant. Note that the experimental
conditions and measured data allowed for the calculation of the interface concentrations and the
enhancement factors, i.e. the gradients at the interface, of formaldehyde and methylene glycol
without any information on the kinetics of the hydration reaction. These quantities are indexed
here as observed.
To establish the reaction order of formaldehyde in the hydration, the equations for diffusion
with parallel reaction in the liquid film are used
F
F
F
R
dx
C d
D =
2
2
) 0 ( x , (1)
F
MG
MG
R
dx
C d
D =
2
2
) 0 ( x , (2)
with the boundary conditions
observed IF MG x MG observed IF F x F
C C C C ) ( ) ( ; ) ( ) (
, 0 , 0
= =
= =
, (3)
MG x MG F x F
C C C C = =
= =
) ( ; ) ( . (4)
Here, the rate of the reaction is written as
MG d
n
F h F
C k C k R
F
= ) ( , (5)
where
F
n denotes the reaction order of formaldehyde.
The additional condition
observed x
F
x
F
dx
dC
dx
dC
, 0 0
) ( ) (
= =
= , (6)
allows the determination of the reaction rate constant,
h
k . The gradient of methylene glycol at
the interface is not independent, but is determined by the one of formaldehyde and the interface
concentrations. This can easily be seen by adding eqs (1) and (2), integrating twice, and applying
boundary conditions (3) and (4), giving
Appendix C: The reaction order of formaldehyde in the hydration.
124
] ) ( [ ) (
, , , , MG IF MG F IF F F IF F IF MG MG
C C C C v
x
C C v C C + + =

, (7)
and
] ) ( [
1
) ( ) (
, , 0 0 MG IF MG F IF F x
F
x
MG
C C C C v
dx
dC
v
dx
dC
+ =
= =

. (8)
Thus, for individual experiments we have no further information available to determine
F
n .
Therefore, the following strategy was adopted. For a given value of
F
n , the reaction rate
constants, ) (
F
n
h
k , were calculated for all experiments by solving eqs (1)-(7) (see below). Next,
the individual rate constants were fitted to an Arrhenius type expression
RT E
F
n
h
a
e k k
/
) (

= , (9)
and the mean absolute relative residual, marr, of the reaction rates was calculated
i
all
eriments
i
F
n
h
F
n
h F
n
h
k
k k
marr

=

=
exp
1
) (
) ( ) (

. (10)
This procedure was repeated for
F
n values ranging from 0.0 to 2.0.
Analytical and approximate analytical solutions
In general, the equations (1)-(7) can be solved numerically only. However, an approximated
analytical solution for the enhancement factor can be obtained by linearization of
F
n
F
C ) (
according to ) 1 /( ) ( 2
1
,
+

F F
n
IF F
n C C
F
. Argument for this linearization is found in the solution
for irreversible nth order kinetics, which, to a good approximation, equals the solution for first
order kinetics, provided that the reaction rate constant, k, is replaced by ) 1 /( 2
1
+

n kC
n
IF
(Westerterp, Van Swaaij & Beenackers, 1984). This way, the enhancement factor is very similar
to the analytical solution obtained by Winkelman & Beenackers (1993) for first order reversible
reactions,
'
] ' tanh[
) ' (
)
] ' cosh[
1
1 (
'
) ' )(
'
] ' tanh[
1 (
1
, ,
,




v K
C C
C C K
C C
C C
K
E
F IF F
MG F
F IF F
MG IF MG
F
+


+ = , (11)
Appendix C: The reaction order of formaldehyde in the hydration.
125
however, here with
'
) ' ( '
'
K D
v K k
F
+
= , (12)
1
,
) (
1
2
'

+
=
F
n
IF F
F
h
C
n
k
k , (13)
1
,
) (
) 1 (
2 '
'

+
= =
F
n
IF F
d F
h
d
C
k n
k
k
k
K . (14)
For a given value of
F
n , the reaction rate constants were obtained from eqs (11)-(14)
iteratively, using the experimental data and the observed values of the formaldehyde enhance-
ment factors. The simple secant iteration method proved adequate for this purpose.
Note that for a first order reaction in formaldehyde, i.e. 1 =
F
n , eqs (11)-(14) represent the
exact analytical solution. The only other case that allows for an exact solution is the zero order
reaction, 0 =
F
n , where the enhancement factor is given by
0
0
,
,
0
0
, 0
0
] tanh[
)
] tanh[
1 (
/
)
] cosh[
1
1 (
1 ) (
=
=
=
=
=
=

+ =
F
F
F
F
F
F
n
n
F IF F
MG IF MG
n
n
F IF F
MG d h
n
n F
v
C C
C C
C C
C k k
E

(15)
where
MG
d
n
D
k
F

=0
. (16)
Numerical solution
Equation (1), with
F
R given by (5),
MG
C by (7), and the boundary conditions by (3), (4)
and (6), was solved for the formaldehyde concentration profile and the reaction rate constant
simultaneously by replacing the differential equation by finite difference equations on a grid of
mesh points on the interval ) 0 ( x . Here,
j F
C ) ( denotes
F
C at mesh point j, i.e. at
) ( x j x = , where N j L 1 = and N x / ) ( = .
A finite difference approximation of eq (1) with second order accuracy reads
j F
j F j F j F
F
R
x
C C C
D ) (
) (
) ( ) ( 2 ) (
2
1 1
=

+
+
) 1 1 ( = N j L . (17)
Appendix C: The reaction order of formaldehyde in the hydration.
126
The boundary conditions (3) and (4) give two more equations
IF F F
C C
, 0
) ( = , (18)
F N F
C C = ) ( . (19)
An additional equation is obtained from a second order Taylor series approximation of
1
) (
F
C :
0
2
2 2
0 0 1
) (
2
) (
) )( ( ) ( ) (
dx
C d x
dx
dC
x C C
F F
F F

+ + = . (20)
The second derivative in (20) is equal to
F F
D R / ) (
0
, see eq (1), while the first derivative is set
equal to the observed gradient at the interface.
The 2 + N equations (17)-(20) can be solved for the unknowns
h
k and ) 0 ( ) ( N j C
j F
L = .
Because of the nonlinearity in the reaction rates
j F
R ) ( Newton-Raphson iteration was used. For
this purpose, the equations, labeled by
j
F , are written as
0
) (
2
) (
) )( ( ) ( ) (
0
2
, 0 0 1 1
=

=
=
F
F
observed x
F
F F
D
R x
dx
dC
x C C F , (21)
0 ) (
, 0 0
= =
IF F F
C C F , (22)
0 ) (
) (
) ( ) ( 2 ) (
2
1 1
=

+ =
+ j F
F
j F j F j F j
R
D
x
C C C F ) 1 1 ( = N j L , (23)
0 ) ( = =
F N F N
C C F . (24)
The vector of unknowns
T
N F F h
C C k ] ) ( , , ) ( , [
0
L = y is updated with a correction y , i.e.
y y y + =
current new
, until convergence is achieved, where the vector of corrections y is
obtained from the matrix equation

=
= =

N
k
j k
k
j
N j F y
y
F
1
) 1 ( L . (25)
Results
The calculated marr data from eq (10), obtained with the approximate analytical solution
and with the numerical method, are shown in Fig. 1 below as a function of the reaction order of
formaldehyde,
F
n . The data show a clear minimum around 1 =
F
n , allowing the conclusion that
the hydration is indeed of the first order in formaldehyde.
Appendix C: The reaction order of formaldehyde in the hydration.
127
A second conclusion is that the results obtained with the approximate analytical solution
method are virtually identical to those obtained from the numerical method. Therefore, at least at
the circumstances considered here, the approximate analytical method is suitable for calculating
mass transfer enhancement factors.
Fig. 1. Marr of the reaction rate constants, see eq (10), vs. the order of formaldehyde in the hydration.
Line: numerical solution; symbols: approximate analytical solution.
0.0 0.4 0.8 1.2 1.6 2.0
40
30
20
10
0
reaction order of formaldehyde
m
a
r
r

[
%
]
List of publications
129
List of publications
The following publications originated from this work:
J.G.M. Winkelman, H. Sijbring, A.A.C.M. Beenackers & E.T. De Vries (1992).
Modeling and simulation of industrial formaldehyde absorbers.
Chemical Engineering Science, 47, 3785.
(included as Chapter 5)
J.G.M. Winkelman, S.J. Brodsky, & A.A.C.M. Beenackers (1992).
Effects of unequal diffusivities on enhancement factors for reversible reactions: numerical
solutions and comparison with DeCourseys method.
Chemical Engineering Science, 48, 2951-2955.
J.G.M. Winkelman & A.A.C.M. Beenackers (1993).
Simultaneous absorption and desorption with reversible first-order chemical reaction:
analytical solution and negative enhancement factors.
Chemical Engineering Science, 48, 2951-2955.
(included as Chapter 3)
J.G.M. Winkelman, M. Ottens & A.A.C.M. Beenackers (2000).
The kinetics of the dehydration of methylene glycol.
Chemical Engineering Science, 55, 2065-2071.
(included as Chapter 2)
J.G.M. Winkelman & A.A.C.M. Beenackers (2000).
Correlations for the density and viscosity of aqueous formaldehyde solutions.
Industrial & Engineering Chememistry Research, 39, 557-562.
(included as Appendix A)
J.G.M. Winkelman, O. Voorwinde, M. Ottens, A.A.C.M. Beenackers & L.P.B.M. Janssen
(2002).
The kinetics and chemical equilibrium of the hydration of formaldehyde.
Chemical Engineering Science, 57, 4067-4076.
(included as Chapter 4)
Samenvatting in het Nederlands
131
Samenvatting in het Nederlands
Deze dissertatie beschrijft theoretisch en experimenteel werk aan de absorptie van formaldehyde
in water. Met resultaten hiervan zijn chemisch-technische modellen ontwikkeld voor de
beschrijving en optimalisatie van industrile formaldehydeabsorbeurs. Deze samenvatting geeft
eerst algemene informatie over formaldehyde, en de commercile productie ervan. Daarna wordt
ingegaan op het doel van dit werk, en vervolgens het uitgevoerde onderzoek en de resultaten.
Formaldehyde
Formaldehyde is een belangrijke grondstof in de chemische industrie. Het is met name een
hoofdbestanddeel van veel soorten kunststoffen en -harsen. Daarnaast worden kleinere
hoeveelheden formaldehyde gebruikt als ontsmettings- en conserveringsmiddel (sterk water),
en bij de productie van rubbers, speciale betonsoorten, explosieven, meststoffen, genees-
middelen, papier, etc. In 2000 werd wereldwijd ongeveer 10 miljoen ton formaldehyde
geproduceerd.
In zuivere vorm is formaldehyde een gas. Het kan echter niet in zuivere vorm worden opgeslagen
of vervoerd omdat formaldehydegas niet stabiel is. Het zuivere gas reageert snel, bijvoorbeeld
met de wand van een container waarin het is opgeslagen, tot een onbruikbare vaste stof. Daarom
wordt formaldehyde vrijwel uitsluitend geproduceerd, verhandeld en vervoerd als een oplossing
van het gas in water.
Industrile productie van formaldehyde
Bij de industrile productie van formaldehyde is methanol, ofwel methylalcohol, de grondstof.
De methanol wordt verdampt en gemengd met lucht. Dit gasmengsel gaat naar een reactor.
Methanoldamp reageert in de reactor met zuurstof uit de lucht tot formaldehyde. Het gas dat uit
de reactor komt, bestaat voornamelijk uit stikstof en formaldehyde. Soms bevat het gas ook nog
een hoeveelheid niet-omgezette methanol. Dit gasmengsel wordt naar een absorbeur geleid
waarin het formaldehydegas oplost in water. Hierbij onstaat het commercile product: een gecon-
centreerde oplossing van formaldehyde in water, ofwel formaline. In de praktijk wordt formaline
vaak verhandeld met een sterkte van 37 of 55 gewichtsprocent formaldehyde.
methanol-
verdamper
reactor absorbeur
formaline
afgas lucht
methanol
water
Fig. 1. Belangrijke stappen in de industrile productie van formaline.
Samenvatting in het Nederlands
132
Absorbeur
Een absorbeur wordt gebruikt om een gasvormige stof op te lossen in een vloeistof. Het gas
wordt onderaan het kolomvormige apparaat naar binnen geleid, en stroomt naar boven. De
vloeistof wordt bovenaan gentroduceerd, en stroomt neerwaarts. Om zoveel mogelijk gas op te
lossen is het voordelig om de neerstromende vloeistof en het omhoog stromende gas intensief
met elkaar in contact te brengen. Dit kan onder meer worden bereikt door de absorbeur te vullen
met een pakking. In de praktijk wordt vaak een gestorte pakking gebruikt: de kolom wordt
gevuld met een willekeurige stapeling van bijvoorbeeld bolletjes, ringen, of n van de vele
andere vormen die commercieel beschikbaar zijn. Door de vloeistof over de pakking te versprei-
den ontstaat een groot contactoppervlak tussen de vloeistof en het gas.
In het onderste deel van een absorbeur wordt de oplossing verzameld, zodat deze kan worden
afgevoerd voor verdere verwerking of opslag. Bij formaldehydeabsorbeurs wordt een deel van de
oplossing die beneden aankomt weer teruggevoerd naar de bovenzijde van de absorbeur. De
vloeistof loopt dan nogmaals door de kolom, en absorbeert meer gas. Dit leidt tot een meer
geconcentreerde oplossing. In figuur 2 is een voorbeeld te zien van een absorbeur met een
vloeistofterugvoer. De pakking is in de figuur schematisch aangeduid met bolletjes.
afgas
water
gas
oplossing
koel-
water
Fig. 2. Een absorbeur met n absorptiebed, en extern gekoelde vloeistofterugvoer.
Samenvatting in het Nederlands
133
Over het algemeen werken absorbeurs efficinter als het gas goed oplost in de vloeistof, en
minder efficint als de oplosbaarheid van het gas klein is. Formaldehyde is schijnbaar goed
oplosbaar in water, gezien het feit dat oplossingen met wel 55 gewichtsprocent formaldehyde
gangbaar zijn. Maar de formaldehydeabsorbeurs opereren minder efficint dan op grond van deze
goede schijnbare oplosbaarheid verwacht mag worden. De belangrijkste reden hiervoor is dat in
de oplossing het formaldehyde met water reageert tot methyleenglycol. Deze hydratatiereactie is
een evenwichtsreactie: methyleenglycol kan ook weer terugreageren tot formaldehyde en water.
Op zijn beurt kan methyleenglycol weer verder reageren tot een serie polymere vormen van
formaldehyde, de zogenaamde poly-oxy-methyleenglycolen. Deze polymerisatiereacties zijn
langzaam, en het zijn ook weer evenwichtsreacties, zie figuur 3.
methyleen-
formaldehyde
h
ydratatie
d
e
h
ydrata
tie
water
glycol
methyleen-
glycol
p
o
ly
merisa
tie
d
e
p
o
lymerisa
t
ie
poly-oxy-
Fig. 3. Reacties van formaldehyde opgelost in water.
De goede schijnbare oplosbaarheid van formaldehyde in water is dus eigenlijk de goede
oplosbaarheid van methyleenglycol en de capaciteit van de oplossing om poly-oxy-methyleen-
glycolen op te nemen. Formaldehyde zelf is, zoals de meeste gassen, slecht oplosbaar in water.
De hydratatiereactie is relatief snel, en bij evenwicht is de hoeveelheid vrij formaldehyde veel
kleiner dan de hoeveelheid methyleenglycol: het hydratatie-evenwicht ligt sterk aan de zijde van
methyleenglycol. Dit zorgt ervoor dat de overdracht van formaldehyde vanuit het gas naar de
vloeistof chemisch versneld is. Formaldehydeabsorbeurs werken dus minder efficint dan op
basis van de goede schijnbare oplosbaarheid verwacht mag worden, maar meer efficint dan
wanneer het formaldehyde niet zou reageren in de oplossing.
Een verdere complicerende factor is dat bij de absorptie van formaldehyde en de daaropvolgende
hydratatiereactie veel warmte vrij komt. Ook bevat het gas dat de absorbeur binnenkomt een
hoeveelheid stoom die, vooral bij condensatie, veel warmte afgeeft . De temperatuur van de neer-
stromende vloeistof neemt hierdoor toe. De warmte die vrijkomt wordt afgevoerd door de
vloeistofterugvoer te koelen met koelwater (zie fig. 2).
Onder meer vanwege dit type complicaties zijn formaldehydeabsorbeurs vaak onderverdeeld in
meerdere secties of absorptiebedden. Elke sectie heeft dan een vloeistofverdeler, -opvang en -
terugvoer met externe koeling. De vloeistof uit de bodem van de ene sectie wordt voor een deel
teruggevoerd naar de top van dezelfde sectie. De rest wordt naar de top van de volgende sectie
geleid. Iets dergelijks geldt voor de gasstroom: het gas dat uit de top van een sectie treedt, wordt
Samenvatting in het Nederlands
134
aan de onderzijde van een vorige sectie ingevoerd. Er ontstaat een patroon zoals gellustreerd is
voor een tweetal secties in figuur. 4. In de praktijk zijn de secties boven elkaar geplaatst, in n
enkele kolom, zodat de installatie compact blijft.
afgas
water
gas
koel-
water
oplossing
koel-
water
Fig. 4. Een schakeling van twee absorptiebedden, elk met een extern gekoelde vloeistofterugvoer.
Dit proefschrift
Het belangrijkste streven van het onderzoek is de ontwikkeling van betrouwbare chemisch-
technische modellen voor industrile formaldehydeabsorbeurs. Het doel hiervan is drieledig. Ten
eerste het nauwkeurig beschrijven van het gedrag van bestaande formaldehydeabsorbeurs.
Vervolgens het voorspellen van de invloed die veranderingen in de manier van bedrijven hebben
op de prestatie van de absorbeurs. En ten slotte het optimaliseren van de absorptie-efficintie en -
capaciteit van formaldehyde in de absorbeurs. Om deze doelstellingen te bereiken is het van
belang om de snelheid te kennen van de reacties in de oplossing (zie fig. 3) bij verschillende
concentraties en temperaturen. Anders gezegd: de kinetiek van de reacties moet bekend zijn.
De kinetiek van de polymerisatie- en depolymerisatiereacties is uitgebreid onderzocht en in de
literatuur beschreven door andere researchgroepen. Over de reactiesnelheid van de hydratatie van
formaldehyde en de dehydratatie van methyleenglycol is slechts fragmentarisch gepubliceerd. Er
zijn een handjevol data bekend, over het algemeen slechts bij kamertemperatuur. De gegevens
die in de literatuur zijn te vinden over de chemische evenwichtsconstante van de hydratatie
Samenvatting in het Nederlands
135
vertonen een grote spreiding: als we verschillende publicaties met elkaar vergelijken dan zit daar
soms meer dan een factor 3 verschil tussen. Daarom zijn de kinetiek en het chemisch evenwicht
van de hydratatie/ dehydratatiereactie nader onderzocht. Dit onderzoek en de resultaten ervan
worden, na een inleidend hoofdstuk, beschreven in de hoofdstukken 2, 3 en 4 van dit proefschrift.
In de hoofdstukken 5 en 6 worden modellen ontwikkeld voor de simulatie en optimalisatie van
industrile formaldehydeabsorbeurs. In een drietal appendices wordt onderzoek naar de fysische
eigenschappen van formaline en enig aanvullend materiaal op de eerdere hoofdstukken
gepresenteerd.
In hoofdstuk 2 is het onderzoek aan de orde naar de reactiesnelheid van de dehydratatie van
methyleenglycol. Hierbij is gebruik gemaakt van de snelle reactie van sulfiet met formaldehyde
waarbij hydroxide-ionen vrijkomen. De concentratie van de hydroxide-ionen kan eenvoudig
worden gemeten. In een oplossing van methyleenglycol en sulfiet wordt formaldehyde gevormd
via de dehydratatiereactie. Dit formaldehyde reageert zeer snel met sulfiet. De toename van de
concentratie van hydroxide-ionen is dan een maat voor de snelheid van de dehydratatie van
methyleenglycol. De reactiesnelheid van de dehydratatie is op deze wijze gemeten bij
temperaturen die relevant zijn voor formaldehydeabsorbeurs (circa van 20 tot 60
o
C).
Hoofdstuk 3 geeft een theoretische verhandeling over de absorptie en/of desorptie van twee
componenten samen met een evenwichtsreactie in de vloeistof tussen die twee. De analytische
oplossingen die hier worden afgeleid geven de absorptie- en/of desorptiesnelheid als functie van
chemische en hydrodynamische eigenschappen van het systeem en de concentraties van de
componenten. De theoretische resultaten die hier zijn bereikt worden toegepast in het volgende
hoofdstuk.
De bepaling van de kinetiek van de hydratatie van formaldehyde wordt beschreven in hoofdstuk
4. De metingen zijn gebaseerd op de chemisch versnelde absorptie van formaldehyde in water en
het mathematische model van hoofdstuk 3 voor dit proces. Als een gasstroom met daarin form-
aldehydegas over water wordt geleid, dan zal het formaldehyde in het water gaan oplossen. Als
het opgeloste formaldehyde niet zou reageren, dan zou een bepaalde mate van verzadiging
optreden aan het oppervlak van de vloeistof. Deze mate van verzadiging is mede bepalend voor
de oplossnelheid. Nu echter het opgeloste formaldehyde wegreageert via de hydratatiereactie, zal
de mate van verzadiging aan het vloeistofoppervlak veel kleiner zijn. Dit resulteert in een hogere
oplossnelheid dan wanneer de reactie niet zou optreden: de absorptie is chemisch versneld. De
mate waarin de absorptie chemisch versneld wordt, is afhankelijk van de reactiesnelheid. Door de
absorptiesnelheid te meten kan via een wiskundig model de reactiesnelheid worden berekend. Op
deze manier is de hydratatiesnelheid van formaldehyde gemeten over het traject van 20 tot 60
o
C.
Met deze resultaten, en die van hoofdstuk 2, is ook de chemische evenwichtsconstante bepaald
over hetzelfde temperatuurtraject.
De industrile absorptie van formaldehyde in water wordt gekenschetst door simultane stof-
overdracht van meerdere componenten, gepaard met meerdere simultane reacties in de vloeistof
en aanzienlijke warmte-effecten. In hoofdstuk 5 wordt een model gepresenteerd waarmee dit
proces wordt gesimuleerd. Het model is gebaseerd op differentiaalvergelijkingen voor de stof- en
energiebalansen in de gasfase en in de vloeistoffase. Voor de resulterende set van gekoppelde
grenswaardeproblemen is een stabiele oplosmethode ontwikkeld via een semi-tijdsafhankelijke
Samenvatting in het Nederlands
136
benadering. Het model bleek zeer goed in staat om het gedrag van bestaande formaldehyde-
absorbeurs te simuleren. Met het model zijn vervolgens de procescondities van een absorbeur bij
Dynea B.V. geoptimaliseerd waardoor een hogere capaciteit en een aanzienlijke besparing van de
grondstof methanol zijn bereikt.
In hoofdstuk 6 wordt het absorbeurmodel uitgebreid met de beschrijving van het gedrag van niet-
omgezet methanol dat via de gasstroom de absorbeur binnenkomt. In het model zijn een groot
aantal reacties in de vloeistoffase opgenomen, alsmede warmte-uitwisseling tussen de gas- en
vloeistofstroom, en chemisch versnelde stofoverdracht van formaldehyde, water, methyleen-
glycol, methanol en hemiformal. Het gedrag van industrile formaldehydeabsorbeurs met extern
gekoelde vloeistofrecirculatie wordt gesimuleerd met het model van niet-evenwichts-trappen. Dit
model is toegepast om de invloed te onderzoeken die de verschillende procesomstandigheden
hebben op het gedrag van methanol in de absorbeurs.
De appendices geven enig aanvullend materiaal. Zo worden in appendix A de resultaten
gepresenteerd van de experimentele bepaling van de viscositeit van formaline bij verschillende
temperaturen en verschillende formaldehydegehaltes. Ook worden hier correlaties gegeven om de
viscositeit en dichtheid van formaline nauwkeurig te berekenen als functie van de temperatuur en
de concentratie. In appendix B worden enkele methoden toegelicht voor de berekening van de
evenwichtsamenstelling (zie fig. 3) van oplossingen van formaldehyde in water, en van
formaldehyde en methanol in water. Appendix C ten slotte bevat additioneel materiaal bij
hoofdstuk 4. Hier wordt aangetoond dat de hydratatiereactie eerste orde in formaldehyde is,
ofwel dat de reactiesnelheid van de hydratatie recht evenredig met de concentratie van form-
aldehyde is.
137
Dankwoord
Iedereen die aan dit proefschrift heeft meegewerkt wil ik van harte bedanken. Enkele mensen wil
ik hier speciaal noemen. Allereerst Prof. dr. ir. A.A.C.M. (Ton) Beenackers, die mijn leermeester
is geweest. Zijn enthousiasme en vakkennis waren van onschatbare waarde, niet alleen voor het
promotieonderzoek, maar ook tijdens de vele projecten waar we nadien samen bij betrokken
waren. Zijn plotselinge heengaan blijft ontstellend. Prof. dr. ir. L.P.B.M. Janssen en Prof. dr. ir.
H.J. Heeres dank ik voor hun bereidheid de promotie af te ronden en voor hun suggesties ten
aanzien van de uiteindelijke tekst. De leden van de leescommissie, Prof. dr. A.A. Broekhuis,
Prof. dr. P.D. Iedema en Prof. dr. ir. G.F. Versteeg, dank ik voor hun aandacht en snelle
beoordeling. Dynea B.V. heeft dit werk niet alleen financieel ondersteund; zeer nuttig waren ook
de uren die ik doorbracht met dhr S. Doorn en dhr E.T. de Vries.
Veel werk is gedaan door de afstudeerders Hermen Steven Giesbert, Henk Sijbring, Marcel
Ottens en Olaf Voorwinde. Hun bijdragen waren onmisbaar voor de totstandkoming van dit
proefschrift. Evenals die van Luuk Balt voor zijn prachtige ontwerpen van de meetopstellingen;
Oetse Staal voor zijn inbreng in de automatisering; Karel van der West en Jaap Struik voor het
betere sleutel- en schroefwerk; Rob Cornelissen voor de ware kunstwerken in glas; en Jan Henk
Marsman voor wie elk analyseprobleem weer een uitdaging bleek. Ten slotte dank ik mijn
familie en vrienden voor hun directe en indirecte bijdragen bij de totstandkoming van dit
proefschrift, in het bijzonder Piety Groeneveld, Wiveca Jongeneel en Anthony Runia.

S-ar putea să vă placă și