0% found this document useful (0 votes)
81 views9 pages

Kinetics of Dry Neutralization of Dodecyl-Benzenesulfonic Acid With Respect To Detergent Granulation

1) The document examines the kinetics of the "dry neutralization" reaction between dodecyl-benzenesulfonic acid (HLAS) and sodium carbonate, which produces sodium dodecyl-benzenesulfonate (NaLAS). 2) The study measures the wetting kinetics of HLAS droplets on sodium carbonate pellets at different temperatures, as well as the reaction rates based on carbon dioxide gas evolved. 3) A numerical model is used to evaluate kinetic parameters like temperature-dependent reaction rate constants and apparent diffusion coefficients, to determine if reaction or diffusion is rate-limiting in the neutralization process.

Uploaded by

Tops10J
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
81 views9 pages

Kinetics of Dry Neutralization of Dodecyl-Benzenesulfonic Acid With Respect To Detergent Granulation

1) The document examines the kinetics of the "dry neutralization" reaction between dodecyl-benzenesulfonic acid (HLAS) and sodium carbonate, which produces sodium dodecyl-benzenesulfonate (NaLAS). 2) The study measures the wetting kinetics of HLAS droplets on sodium carbonate pellets at different temperatures, as well as the reaction rates based on carbon dioxide gas evolved. 3) A numerical model is used to evaluate kinetic parameters like temperature-dependent reaction rate constants and apparent diffusion coefficients, to determine if reaction or diffusion is rate-limiting in the neutralization process.

Uploaded by

Tops10J
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 9

ARTICLE

pubs.acs.org/IECR

Kinetics of Dry Neutralization of Dodecyl-Benzenesulfonic Acid with


Respect to Detergent Granulation
Marek Sch€ongut, Zdenek Grof, and Frantisek Stepanek*
Department of Chemical Engineering, Institute of Chemical Technology, Prague, Technicka 5, 166 28 Prague 6, Czech Republic

ABSTRACT: Sodium dodecyl-benzenesulfonate, a common anionic surfactant used in powder detergents, is a product of the so-
called “dry neutralization” reaction between dodecyl-benzenesulfonic acid and sodium carbonate. The acid acts not only as a reactant
but also as a binder in a granulation process; the carbonate particle wetting by the acid droplets and the neutralization reaction occur
simultaneously. In this work the reaction and wetting kinetics have been investigated by a macroscopic method of observing the
behavior of an acid drop on a substrate pellet made from sodium carbonate. Wetting kinetics at three temperatures (20, 40, and
60 °C) were measured and expressed as a relationship between the three-phase contact line velocity and the instantaneous contact
angle. The reaction rates were evaluated from the volume of CO2 gas evolved during the reaction. The kinetic parameters
(temperature-dependent reaction rate constant and apparent diffusion coefficients) have been evaluated using a numerical model of
coupled reaction and diffusion processes in the passivation layer. The results indicate that because of diffusion limitations the
neutralization reaction takes place predominantly in the “wet” product layer rather than at the solidliquid interface.

1. INTRODUCTION is to correlate the properties of the final product with changes in


Sodium dodecyl-benzenesulfonate is a common anionic surfac- process conditions and feed material properties.20 What makes
tant, used in a wide range of detergents such as laundry washing this particular process complex to describe is the fact that agglom-
powders, dishwasher tablets, and other household or industrial eration and an exothermic chemical reaction occur simultaneously.
detergents.13 It is produced by the neutralization of dodecyl- Properties of the liquid bridges (e.g., viscosity) are changing
benzenesulfonic acid, which is often supplied as a mixture of C10C13 continuously with time and the process of wetting is directly
isomers and therefore denoted as HLAS in the literature, where dependent on a chemical reaction between the binder and carbonate
LAS stands for Linear Alkylbenzene Sulfonate (NaLAS then particles. At the same time, the reaction rate depends on the
denotes its sodium salt). NaLAS has earned its popularity through liquidsolid contact area. An accurate analysis of microscale
good washing efficiency, favorable cost/performance ratio, and wetting11,14,21 and reaction-diffusion processes is therefore cru-
environmental friendliness.4 If the base used for HLAS neutra- cial for understanding and rational design of reactive granulation.
lization is in the solid phase (in this case sodium carbonate) the Wettability does influence not only the degree and rate of
reaction is called “dry neutralization” and its stoichiometric equation binder spreading over particles but also the shape and strength of
can be written as liquid bridges.2224 Wetting kinetics can be evaluated from an
observation of a contact angle between the solid substrate and the
2HLAS þ Na2 CO3 f 2NaLAS þ H2 O þ CO2 ð1Þ liquid during wetting and the velocity of the three-phase contact
line.25,26 For a constant volume and homogeneous substrates,
Fully neutralized NaLAS is a hygroscopic5 soft-solid material which analytical expressions such as the Tanner’s law can be used for
can form various mesophases (e.g., liquid crystals of lamellar type) describing the wetting kinetics.27 Spreading accompanied by a
at room temperature.6 The phase behavior is illustrated in Figure 1. phase change or a chemical reaction in the liquid phase is more
Industrially, dry neutralization typically occurs in a reactive difficult to describe and often requires a numerical solution.28
granulator where HLAS is contacted with sodium carbonate powder Experimentally determined dependence between the three-
(and possibly other formulation components) under intense phase contact line velocity and the instantaneous contact angle
mixing, and granules are formed.710 In a typical granulator the is then necessary for model validation or parameter estimation.
acid is fed to the mixer and sprayed onto a fluidized or mechanically Beyond detergent granulation, the investigation of the kinetics of
agitated powder bed. With ongoing neutralization, a layer of reaction reactive wetting and noncatalytic solidliquid interfacial reac-
products grows on the surface of unreacted sodium carbonate tions is significant also in other diverse application areas29 such as
particles and forms bridges strong enough to bind the particles the fabrication of microfluidic devices, the dissolution of phar-
together in larger agglomerates.11 The final granule properties maceutical tablets,30 or the acid stimulation of porous carbonate
such as size distribution and porosity, as well as the overall degree reservoirs for permeability enhancement.31
of neutralization are determined by the elementary process of The aim of the present study is to quantify the reactive wetting
binder spreading on the primary particles (wetting) and binder kinetics of HLAS acid on a sodium carbonate substrate as a
solidification due to a chemical reaction.12
It has been shown both computationally13 and experimentally1416 Received: May 16, 2011
that the relative rates of these elementary processes jointly deter- Accepted: September 7, 2011
mine granule attributes such as porosity and subsequently influence Revised: September 4, 2011
their end-use behavior such as dissolution rate.1719 A challenge Published: September 07, 2011

r 2011 American Chemical Society 11576 dx.doi.org/10.1021/ie201047r | Ind. Eng. Chem. Res. 2011, 50, 11576–11584
Industrial & Engineering Chemistry Research ARTICLE

Figure 1. Layer of a liquid-crystalline phase of NaLAS/H2O present


between HLAS (dark bottom layer) and a Na2CO3/H2O solution (top
clear layer).

Figure 2. Scheme of an acid drop spreading over a substrate pellet.

function of temperature, and simultaneously to evaluate the


kinetics of the dry neutralization reaction. By means of a
numerical model of coupled reaction and diffusion, the rate- Figure 3. SEM image of the surface of a freshly prepared Na2CO3 pellet.
limiting step of the neutralization reaction will be determined and
the relevant parameters (temperature-dependent rate constants
and apparent diffusion coefficients) evaluated.

2. METHODOLOGY
2.1. Experimental Section. To quantify the rates of elemen-
tary processes that take place during reactive wetting, it is necessary
to decouple them from other phenomena that may occur during
reactive granulation such as particle agglomeration and surface
renewal due to granule breakage. To investigate wetting kinetics,
the three-phase contact line velocity and the instantaneous
contact angle must be observable, and to quantify the rate of a
heterogeneous chemical reaction, the solidliquid contact area
and the extent of reaction must be measurable. Conditions Figure 4. Experimental setup: (a) light source, (b) heater, (c) digital
camera, (d) heated cell.
prevailing in a mechanically agitated powder bed do not facilitate
such direct observation, therefore the following approach has
been taken. calculation of solidliquid contact area from the measured base
A small drop of HLAS (purchased from Acros Organics, diameter of the spreading droplet.
mixture of C10C13 isomers) with a volume of approximately Prior to their use, tablets were kept in a desiccator to avoid
4 μL was placed on a substrate pellet (Figure 2) which has been absorption of atmospheric moisture. During an experiment the
prepared from anhydrous sodium carbonate fine powder (purchased tablet was placed into an insulated measurement cell with
from Lach-ner, mean particle size 14 μm) and compacted on a controlled temperature (Figure 4), and the acid was dropped
Carver pellet press at 7000 psi resulting in an essentially nonporous using a standard medical syringe equipped with a needle of 0.8
tablet with a microscopically smooth surface. A SEM image of a mm internal diameter. Both the tablet and the syringe were kept
freshly prepared unreacted tablet can be seen in Figure 3. The in the measurement cell for a sufficiently long time to allow their
absence of surface roughness or porosity is important to allow the temperatures to equilibrate. Although the neutralization reaction
11577 dx.doi.org/10.1021/ie201047r |Ind. Eng. Chem. Res. 2011, 50, 11576–11584
Industrial & Engineering Chemistry Research ARTICLE

is exothermic, the thermal mass of the tablet and the unreacted


volume of the droplet is large compared to the evolved heat, and
the system can be regarded as isothermal. This assumption was
confirmed by monitoring the tablet surface temperature near the
three-phase contact line by an infrared thermometer throughout
several experiments.
The distance between the needle tip and the pellet surface
was 15 mm, which is small enough to neglect the effect of
gravity on the droplet deposition and its spreading on the
substrate, leaving only capillary forces to be taken into account
(low Weber number deposition). With these conditions, a
spherical cap approximation can be used for the droplet shape
fitting during wetting. The spreading experiment itself takes
approximately 30 s, which was found to be long enough for the
contact line between the drop and the pellet to arrest (contact
line pinning). In the course of spreading, droplet images
were recorded using a Canon EOS 40D camera with Sigma
Figure 5. Model of the passivation layer with concentration profiles and
EX 105 mm macro-lens, at a frequency of approximately 6 boundary conditions.
frames per second.
The images were subsequently cropped and adjusted in products NaLAS (further denoted C) and H2O (denoted D)
the image processing software ImageJ and merged into a single separating the pure reactants, HLAS acid (denoted A) and the
file suitable for further image analysis. The droplet shape was solid base Na2CO3 (further denoted B), cf. Figure 2. In line
evaluated by the Sessile Drop Fit method (SDF) using the Drop with the macroscopic observation (Figure 1), the passivation
Shape Analysis software from Kruss, Inc. The desired parameters, layer consisting of the NaLAS-H2O liquid crystals is assumed
namely, drop base diameter, drop height, drop volume, and con- to be separated from the unreacted HLAS acid by a sharp
tact angle, were thus obtained for each point in time. The velocity interface that is not affected by diffusion of the time course of
of the three-phase contact line was then calculated from the time the experiment. As the passivation layer is just a small fraction
change of the droplet base diameter. The relationship between of the whole droplet, no CO2 presence is assumed within the
the contact line velocity and the dynamic contact angle was used passivation layer in the form of bubbles. This assumption is
for the description of wetting kinetics. based on the fact that compared with the highly viscous liquid
Reaction kinetics and passivation layer growth were calculated crystalline phase of the NaLAS-H2O product layer, a more
from the volume of evolved CO2 (increase in the overall droplet favorable environment for bubble nucleation at a given CO2
volume with time). During the reaction rate estimation, the following supersaturation is the lower-viscosity bulk HLAS droplet
simplifying assumptions were used: (i) ideal gas equation of state, above the passivation layer.
(ii) the volume of CO2 evolved equals the increase in the drop Upon the initial contact of the acid and the carbonate surface,
volume (i.e., negligible contribution of dissolved CO2), (iii) all an interfacial reaction can be expected, but as soon as some
evolved CO2 forms bubbles with atmospheric pressure that do reaction products form, the location of the reaction front is a
not leave the droplet interior, (iv) the net volume change attribu- priori unknown. For further reaction to occur, either HLAS or
table to the consumption of acid and formation of the passivation dissolved Na2CO3 (or both) must diffuse into the product layer
layer is negligible. Under these assumptions, the reaction kinetics where the neutralization reaction can take place. Interfacial
can be estimated from the measured change in droplet volume. reaction at either end of the product layer would occur in the
The extent of reaction (eq 1) obtained by experiment ξexp is limiting case of a very fast diffusion of one reactant across the
therefore layer or a very low solubility of the other reactant in the layer.
Otherwise, the neutralization reaction must occur in the volume
dξexp dnCO2 p dV p V ðti Þ  V ðti1 Þ of the product layer, with both reactants (HLAS and dissolved
¼ ¼ z ð2Þ
dt dt RT dt RT ti  ti1 sodium carbonate) diffusing against each other, cf. Figure 5.
Assuming that the density of the product layer is constant and
using the additivity of molar volumes for NaLAS and water
where ti1 and ti are two subsequent points in time, V is the components, the following equation can be written for the layer
measured droplet volume, p is atmospheric pressure, T is growth rate:
temperature, and R is the molar gas constant.
The assumption that no CO2 gas bubbles leave the droplet for !
dL 3 2 MC 1 MD dξ
the duration of the experiment was found to be valid for all three ¼ þ ð3Þ
temperatures reported in this work (20, 40, and 60 °C). When dt A 3 FC 3 FD dt
the experiment was attempted at 80 °C, the viscosity of HLAS
acid became too low and the partial pressure of CO2 too high to where L is the layer thickness, ξ is the extent of reaction, MC, MD,
prevent the bubbles from escaping the droplet, meaning that FC, and FD are molar weights and densities of NaLAS and water,
drop shape analysis could no longer be used as a method for respectively. The contact area A = πr2 is assumed constant as L is
measuring the reaction rate. small compared to the droplet height and base radius r. Note that
2.2. Mathematical Model. A pseudo-steady state model the volumetric effect of diffusing reactants is neglected in eq 3.
describing the growth of a product (passivation) layer has been The overall neutralization rate is obtained by integration over
formulated. The layer is composed of the neutralization reaction the product layer and adding the contribution of possible interfacial
11578 dx.doi.org/10.1021/ie201047r |Ind. Eng. Chem. Res. 2011, 50, 11576–11584
Industrial & Engineering Chemistry Research ARTICLE

reactions at the boundaries Table 1. List of Model Parameter Values Used


Z L
dξ model parameter symbol value
 R ¼A kCA ðxÞ CB ðxÞ dx þ AðRs, A þ Rs, B Þ
dt 0
HLAS (component A)
ð4Þ stoichiometric coefficient νA 2
Here k is the reaction rate constant, CA and CB are HLAS and molecular weight MA 326.49 g mol1
Na2CO3 concentrations in the product layer, and Rs,i repre- density FA 1200 kg m3
sent rates of reactions taking place at each boundary. At this diffusivity ratio DA/DB 0.53
point, let us neglect both surface terms and assume that the Na2CO3 (component B)
neutralization reaction takes place only in the bulk of the
stoichiometric coefficient νB 1
product layer. This approximation may seem to be a crude
one; however, it is actually justified as will be explained later in molecular weight MB 105.99 g mol1
the Results section. density FB 2540 kg m3
The reaction-diffusion equations representing the mass bal- saturated concentration cB0 2076 mol m3
ance of HLAS and Na2CO3 need to be solved to obtain their NaLAS (component C)
concentration profiles in the product layer (Figure 5). If the molecular weight MC 348.48 g mol1
characteristic time of the layer growth is sufficiently long compared
density FC ≈ 1000 kg m3
to the characteristic time of concentration profile evolution
within the layer, then the pseudo-steady state assumption can H2O (component D)
be used, and the mass balance equations reduce to a boundary molecular weight MD 18.02 g mol1
value problem for the following two ODEs: density FD 1000 kg m3
d2 C i
0 ¼ Di  jνi jkCA CB i ∈ fA, Bg ð5Þ The overall neutralization reaction rate, eq 4, written in terms of
dx2 the dimensionless concentrations becomes
where DA and DB are diffusion coefficients of HLAS and Na2CO3, Z 1
respectively, in the product (NaLAS/water mixture) and νi is the R ¼ ALkCA0 CB0 C ^ B ð^xÞ d^x
^ A ð^xÞ C ð12Þ
stoichiometric coefficient. The boundary conditions are 0

HLAS side : CA ðx ¼ 0Þ ¼ CA0 CB ðx ¼ 0Þ ¼ 0 ð6Þ The model is completed by substituting R into eq 3 to obtain
Z 1
dL 3kCA0 CB0 ^ B ð^xÞ d^x
^ A ð^xÞ C
 ¼ L C ð13Þ
dCA  dt CCD 0
Na2 CO3 side :  ¼ 0 CB ðx ¼ LÞ ¼ CB0 , ð7Þ
dx  In the last equation, the constant CCD is the molar density of the
x¼L
NaLAS/water mixture
where CA0 = FA/MA and CB0 is the concentration of saturated
1 2 MC 1 MD
Na2CO3 solution in the product mixture. The zero-flux boundary ¼ þ ð14Þ
condition at x = L (Na2CO3 side) follows from the condition of CCD 3 FC 3 FD
no diffusion into the solid phase, and the boundary condition at
(x = 0) is a definition of the interface between unreacted acid and 2.2.2. Estimation of Model Parameters from Experimental
the product layer (i.e., all dissolved carbonate is consumed and its Data. The model in dimensionless form has three independent
concentration therefore drops to zero). parameters: Da2A/L2, Da2B/L2, and 3kCA0CB0/CCD. As the satura-
2.2.1. Mathematical Model in Dimensionless Form. Let us tion concentrations of all components can be estimated relatively
define dimensionless concentrations C ^ A = CA/CA0 and C ^B =
easily (cf. Table 1), the unknowns left to be found by fitting
CB/CB0, and a dimensionless length ^x = x/L. Equation 5 then experimental data are k, DA and DB. To further reduce the
becomes number of unknowns, the ratio between the diffusion coefficients
^i
d2 C DA/DB was assumed constant and equal to 0.53. This ratio is
0¼ ^A C
 Da2i ðLÞC ^B i ∈ fA, Bg ð8Þ based on a scaling expressed by the StokesEinstein equation
d^x2
 
with boundary conditions DA MB FA ð1=3Þ
≈ ð15Þ
DB MA FB
HLAS side : ^ A ð^x ¼ 0Þ ¼ 1
C ^ B ð^x ¼ 0Þ ¼ 0
C ð9Þ
 The growth of the passivation layer is obtained by integrating
^ A 
dC
Na2 CO3 side :  ¼0 ^ B ð^x ¼ 1Þ ¼ 1
C ð10Þ eq 13. The boundary value problem (eq 8) has to be solved to
d^x  obtain the concentration profiles required for the evaluation of
^x ¼ 1
the integral reaction rate from eq 13. The value of the integral in
and two Damk€ohler numbers, which are ratios between the eq 13 depends on Da2A and Da2B and therefore changes with
reaction and the diffusion rates, respectively, growing layer thickness L. The passivation layer thickness as
function of time was evaluated from experimental data using the
jνA jkCB0 L2 jνB jkCA0 L2
Da2A ðLÞ ¼ Da2B ðLÞ ¼ ð11Þ combination of eq 2 and eq 3. The actual values of the two un-
DA DB known parameters (rate constant k and diffusion coefficient DA)
11579 dx.doi.org/10.1021/ie201047r |Ind. Eng. Chem. Res. 2011, 50, 11576–11584
Industrial & Engineering Chemistry Research ARTICLE

Figure 6. Example of droplet shape evolution during reactive wetting at 20 °C. Note that the CO2 bubbles remain trapped within the droplet.

Figure 8. Change of drop volume with time evaluated from a wetting


Figure 7. Change of base diameter and instantaneous contact angle experiment at 20 °C.
evaluated from a wetting experiment at 20 °C.

were then found by minimization of the sum of squared errors 3.1. Wetting Kinetics. The wetting phase starts with the drop
between experimental and modeled data for a period of 30 s from touching the substrate and ends with the contact line arrest. While
the moment when an increase of the droplet was detected. The the chemical reaction (thus the associated drop volume increase)
time interval of 30 s was chosen so as to provide a sufficiently long continues beyond this point, the three-phase contact line pinning
time series for statistically robust evaluation of the adjustable occurs after a relatively short time presumably because of the layer
parameters. of a semisolid liquid crystalline phase that forms as a reaction
product. The contact line velocity was calculated as the time
derivative of the drop base radius using the finite difference formula
3. RESULTS AND DISCUSSION
dr rðti Þ  rðti1 Þ
The evolution of droplet shape in the course of one experi- U ¼ z ð16Þ
ment is shown in Figure 6. Note the gradual emergence of CO2 dt ti  ti1
bubbles in the droplet volume. The droplet base radius, droplet where U is the contact line velocity, r is the droplet radius, and t is
height and the instantaneous contact angle were evaluated at time. The wetting kinetics equation is usually defined as a relation-
each point in time. The dependence of the contact line velocity ship between the contact line velocity and the instantaneous contact
and dynamic contact angle on time (Figure 7) was used for the angle. For nonreactive wetting of an ideal surface this relationship is
wetting kinetics characterization, while the time dependence of often found to be linear when the contact angle approaches its
drop volume (Figure 8), which was calculated from the base equilibrium value. Wetting of nonhomogeneous substrates or wetting
radius and droplet height based on the spherical cap approx- coupled with other phenomena (heat transfer, reactive wetting, etc.)
imation,21 was used for the evaluation of reaction kinetics. can generally follow more complex relationships; however, the system
Experiments were carried out at 20, 40, and 60 °C, and at least considered in this work was also found to satisfy a linear relationship
five repeats were done at each temperature (depending on the
statistical spread of the data). UðθÞ ¼ αðθ  θ0 Þ ð17Þ

11580 dx.doi.org/10.1021/ie201047r |Ind. Eng. Chem. Res. 2011, 50, 11576–11584


Industrial & Engineering Chemistry Research ARTICLE

where α is the spreading rate prefactor and θ0 is the value of contact was 5.0 s, 3.5 s, and 0.2 s for 20, 40, and 60 °C, respectively. Each
angle at contact line arrest. Note that θ0 cannot be called “equilib- experimental data series was used for finding the rate constant k
rium” contact angle as equilibrium in the thermodynamic sense is not and diffusion coefficient DA according to the procedure de-
yet reached because of the ongoing chemical reaction. The two scribed in Sec. 2.2.2 above. An example of fitting the experimental
unknown parameters, α and θ0, were found by linear regression of data (in this case at 60 °C) by the numerical model is shown in
experimental data obtained from the droplet shape evolution at each Figure 11. A summary of the mean values of the kinetic
temperature. Figure 9 shows an example of one such experiment at parameters and their standard deviations at each temperature is
20 °C. The mean values and standard deviations of the regression provided in Table 2.
parameters α and θ0 based on at least five repeats of each measure- From these experimental data it is clear that the reaction
ment at 20, 40, and 60 °C are given in Table 2. The times of contact rate constant k increases with increasing temperature, and as
line arrest were 16.3 ( 3.0 s, 10.8 ( 4.2 s, and 5.4 ( 1.1 s for 20, 40, Figure 12 shows, this increase is in line with the Arrhenius law.
and 60 °C, respectively. On the other hand, the estimated diffusion coefficient DA does
The data show that increasing temperature leads to a faster
spreading over the substrate, which can be explained by the fact
that the viscosity of the acid decreases with increasing tempera-
ture. The asymptotic contact angle drops significantly when
temperature is increased from 20 to 40 °C but then remains
more-or-less constant, considering the experimental error. This
is consistent with the fact that the wettability of most substrates
generally increases with increasing temperature. The reason why
the decrease of the contact angle does not continue (or may even
be slightly reversed) beyond 40 °C in this case can be attributed
to the chemical reaction. The evolving CO2 gas leads to an increase
in the droplet volume, which manifests itself as an increase in the
apparent contact angle when the contact line is pinned. This
process competes with the decrease of contact angle which would
occur if there were no reaction. The observed value of θ0 is the
result of a combination of these two competing processes, and
therefore a faster reaction at higher temperatures may prevent
the contact angle from reaching a lower value.
3.2. Reaction Kinetics. Experimental values of the passivation Figure 10. Experimental data on passivation layer growth at different
layer growth at different temperatures, calculated from the temperatures. The time is relative from the on-set of reaction at each
droplet volume using eqs 2 and 3 are summarized in Figure 10. temperature.
The time is taken as a relative time from the moment when the
droplet volume started to increase (induction time). This time

Figure 9. Estimation of wetting kinetics parameters for experimental Figure 11. Experimental data at 60 °C fitted by the numerical model.
data measured at 20 °C. The time is relative from the on-set of reaction.

Table 2. Wetting and Reaction Kinetics Parameters Evaluated from Experiments


T/°C α/(mm s1 deg1) θ0/deg k/(m3 mol1 s1) DA/(m2 s1)

20 (2.83 ( 0.41)  102 67.4 ( 5.4 (6.05 ( 1.13)  105 (1.88 ( 0.74)  1014
2 4
40 (3.31 ( 0.70)  10 40.2 ( 5.5 (4.73 ( 3.47)  10 (3.04 ( 0.83)  1014
2 3
60 (4.42 ( 0.95)  10 48.8 ( 5.7 (1.78 ( 0.41)  10 (1.48 ( 0.37)  1014

11581 dx.doi.org/10.1021/ie201047r |Ind. Eng. Chem. Res. 2011, 50, 11576–11584


Industrial & Engineering Chemistry Research ARTICLE

Figure 12. Dependence of the reaction rate constant on temperature


(Arrhenius plot).

not appear to follow any obvious pattern as temperature changes.


Although diffusion coefficient on its own should slightly increase
with temperature, the phase behavior of the passivation layer
made from sodium salt liquid crystals is more complex, and it can
act to either enhance or suppress any underlying trends in diffusion
coefficients at various temperature intervals. In any case, it can be
concluded that compared to the exponential rise of the rate constant
with temperature, the diffusion coefficients remain essentially
unchanged. This in turn means that with increasing temperature,
diffusion limitations can be expected to play a more significant
role in the overall progress of the reaction, as will be analyzed in
more detail in the following section.
From the granulation point of view, diffusion limitations may
result in an incomplete neutralization of the original HLAS acid, Figure 13. Solutions of eq 8 for Da2A/Da2B = 2.71 and parametrized by
which is undesirable. Conversion can be improved by providing a Da2A = 10a, where a was varied in the range < 2;5 > with stepsize 0.5.
stoichiometric excess of carbonate, by intense mechanical agita- Data corresponding to Da2A = 1 are highlighted as a dashed line. (a)
tion for surface renewal, by reducing the HLAS droplet size, or by Concentration profile of HLAS (red color) and Na2CO3 (blue color).
other methods that enhance its spreading (e.g., viscosity reduction). (b) Profiles of the local neutralization rate.
3.3. Parametric Analysis of the Model. To gain a deeper
understanding of the rate-limiting step under different condi-
tions and also to verify that the initial model assumptions are
consistent with the experimentally observed behavior, let us
discuss the HLAS and Na2CO3 concentration profiles in the
product layer for a range of Da numbers as shown in Figure 13a.
Two limiting cases can be identified: (i) reaction-limited regime
for low Da numbers and (ii) diffusion-limited regime for high Da
numbers. The neutralization rate depends on the product of both
acid and base concentrations in the layer, as shown in Figure 13b.
In the reaction-limited regime (i.e., fast diffusion and/or slow
reaction), the reaction takes place throughout the layer as both
species can diffuse at a sufficient rate. However when the Damk€ohler
number is increased toward the diffusion-limited regime, the
reaction becomes localized in a narrower section within the layer R
where both diffusing species meet and react instantaneously Figure 14. Dependence of the integral I = 10C ^ AC
^ B dx on Da2A. Points
(large Da means fast reaction compared to diffusion). The exact denoted by circles on the line correspond to individual profiles shown in
position of the reaction peak depends on the ratio Da2A/Da2B. Figure 13.
The value of Da2A at which the neutralization reaction becomes
hindered by mass transport limitations can be identified by plotting kinetically limited regime (Da2A < 0.1), transition regime (0.1 <
the dependence of the integral appearing in eq 13 on the Damk€ohler Da2A < 10), and diffusion-limited regime (Da2A > 10). Let us recall
number, as shown in Figure 14. The dependence reveals a typi- that on deriving the model, terms accounting for the contribution
cal sigmoidal shape, which can be divided into three regions: of possible surface reactions in eq 4 were neglected. As is evident
11582 dx.doi.org/10.1021/ie201047r |Ind. Eng. Chem. Res. 2011, 50, 11576–11584
Industrial & Engineering Chemistry Research ARTICLE

At 20 °C the onset of the reaction becomes visible after an


induction time of about 5 s, while at 60 °C the onset of the reaction
is practically immediate. A mathematical model taking into
account the reaction and diffusion in the passivation layer was
developed and used for the evaluation of kinetic parameters from
experimental data. The obtained rate constant was found to follow
the Arrhenius law, and the validity of the model assumptions was
further confirmed by analyzing the concentration profiles and
localization of the reaction front in the passivation layer. With the
exception of the very early phases of the reaction, the reaction
front was found to be contained within the product layer (a mixture
of NaLAS salt and H2O, with dissolved HLAS acid and sodium
carbonate base). Our results indicate that although the base that
participates in the “dry” neutralization reaction initially starts as a
Figure 15. Dependence of Da2A on the layer thickness L calculated using
values of k and DA obtained from the experimental data. Horizontal lines
solid, the reaction itself actually occurs as a homogeneous reaction
mark the diffusion-limited (top), transition (middle) and reaction- within the “wet” product phase rather than as a heterogeneous
limited (bottom) regimes. reaction at the solidliquid interface. The reaction and wetting
kinetics obtained by this work will be used as input parameters
from Figure 13, in the reaction-limited regime (limit of low Da for a particle-level model of granule microstructure formation
number) both reactants are still present at opposite borders of during reactive granulation.
the passivation layer, suggesting that surface reactions could in
principle take place. On the other hand, as soon as the passivation ’ AUTHOR INFORMATION
layer thickness, L, increases above approximately 0.1 μm and the Corresponding Author
system enters the transition regime, the concentration of each *E-mail: [email protected].
reactant at the interface with the bulk of the other reactant significantly
decreases. Finally, in the diffusion-limited regime the interfacial
reactions can be ruled out as each reactant is consumed within ’ ACKNOWLEDGMENT
the passivation layer well before it can reach the opposite end,
which is also manifested by the localized peak of reaction rate This work has been supported by the Czech Science Founda-
(cf. Figure 13). In this case the passivation layer effectively sep- tion (104/08/H055) and by the grant of the Specific University
arates both reactants, and the surface reaction terms Rs,i therefore Research (MSMT no. 21/2011). F.S. would like to acknowledge
become insignificant. support by the European Research Council (project no. 200580-
From this analysis it can be concluded that the present model Chobotix).
would probably underestimate the reaction rate at the very early
stages when the passivation layer thickness is small (low Da ’ REFERENCES
number) and interfacial reactions may be present, but should (1) de Groot, W. H.; Adami, I.; Moretti, G. F. The Manufacture of
describes the reaction rate accurately at later stages after the product Modern Detergent Powders; H. de Groot Academic Publisher: Wassenaar,
layer is sufficiently thick for diffusion to be the rate-limiting step. The Netherlands, 1995.
To assess whether the use of this model was indeed justified in (2) Scheibel, J. J. The Evolution of Anionic Surfactant Technology to
our case, the dependence of Da2A on the layer thickness L was meet the Requirements of the Laundry Detergent Industry. J. Surfactants
calculated using k and DA parameters found by fitting experi- Deterg. 2004, 7, 319.
mental data, and it is shown in Figure 15. The horizontal lines in (3) Yu, Y.; Jin, Z.; Bayly, A. E. Development of Surfactants and
Figure 15 delimit the reaction- and diffusion-limited regimes Builders in Detergent Formulations. Chin. J. Chem. Eng. 2008, 16, 517.
(4) Mungray, A. K.; Kumar, P. Fate of Linear Alkylbenzene Sulfo-
together with a transition regime between them. Comparing the
nates in the Environment: A Review. Int. Biodeterior. Biodegrad. 2009,
values of L with those seen in Figure 10 it can be concluded that 63, 981.
the majority of the experimental data points indeed lie in the (5) Cohen, L.; Soto, F.; Roberts, D. W.; Emery, D. Hygroscopicity of
transition or the diffusion-limited regions where the model is valid. Linear Alkylbenzene Sulfonate Powders. Tenside, Surfactants, Deterg.
2003, 40, 1.
4. CONCLUSIONS (6) Ramaraji, S. M.; Carroll, B. J.; Chambers, J. G.; Tiddy, G. J. T.
The Liquid Crystalline Phases formed by Linear-dodecylbenzene Sul-
From the experiments it appears that the kinetics of reactive phonic Acid during Neutralization with Sodium Carbonate. Colloids
wetting of sodium carbonate by HLAS has two phases. The first Surf., A 2006, 288, 77.
phase is associated with rapid spreading, meaning a fast decrease (7) Appel, P. W. Modern Methods of Detergent Manufacture. J. Surf.
of the drop height and contact angle, and an increase of the drop Deterg. 2000, 3, 395.
base diameter until an asymptotic value is reached at which the (8) Borchers, G. Design and Manufacturing of Solid Detergent
three-phase contact line is pinned. The wetting kinetics was Products. J. Surf. Deterg. 2005, 8, 123.
(9) Smulders, E.; Rybinsky, W.; Sung, E.; R€ahse, W.; Steber, J.;
found to follow a linear relationship between the contact line
Wiebel, F.; Nordskog, A. Laundry Detergents; Wiley: Weinheim,
velocity and the instantaneous contact angle. The spreading rate Germany, 2007.
systematically increased with increasing temperature. During (10) Germana, S.; Simons, S. J. R.; Bonsall, J. Reactive Binders in
the second (reaction) phase, which overlaps with the wetting phase Detergent Granulation: Understanding the Relationship between Bin-
to some extent, the drop volume increase becomes clearly visible der Phase Changes and Granule Growth under different Conditions of
because of carbon dioxide formation by the neutralization reaction. relative humidity. IEC Res. 2008, 47, 6450.

11583 dx.doi.org/10.1021/ie201047r |Ind. Eng. Chem. Res. 2011, 50, 11576–11584


Industrial & Engineering Chemistry Research ARTICLE

(11) Germana, S.; Simons, S.; Bonsall, J.; Carroll, B. LAS Acid
Reactive Binder: Wettability and Adhesion Behaviour in Detergent
Granulation. Powder Technol. 2009, 189, 385.
(12) Palzer, S. Agglomeration of Pharmaceutical, Detergent, Chemical
and Food Powders - Similarities and Differences of Materials and
Processes. Powder Technol. 2011, 201, 2.
(13) Stepanek, F.; Ansari, M. A. Computer Simulation of Granule
Microstructure Formation. Chem. Eng. Sci. 2005, 60, 4019.
(14) Iveson, S. M.; Litster, J. D.; Hapgood, K.; Ennis, B. J. Nuclea-
tion, Growth and Breakage Phenomena in Agitated Wet Granulation
Processes: A Review. Powder Technol. 2001, 117, 3.
(15) Rajniak, P.; Mancinelli, C.; Chern, R.; Stepanek, F.; Farber, L.;
Hill, B. T. Experimental Study of Wet Granulation in Fluidized Bed:
Impact of the Binder Properties on the Granule Morphology. Int. J.
Pharm. 2007, 334, 92.
(16) Thielmann, F.; Naderi, M.; Ansari, M. A.; Stepanek, F. The
Effect of Primary Particle Surface Energy on Agglomeration Rate in
Fluidised-bed Wet Granulation. Powder Technol. 2008, 181, 160.
(17) Stepanek, F. Computer-Aided Product Design  Granule Dissolu-
tion. Chem. Eng. Res. Des. 2004, 82, 1458.
(18) Brielles, N.; Chantraine, F.; Viana, M.; Chulia, D.; Branlard, P.;
Rubinstenn, G.; Lequeux, F.; Mondain-Monval, O. Dissolution of a
Surfactant-containing Active Porous Material. J. Colloid Interface Sci.
2008, 328, 344.
(19) Ansari, M. A.; Stepanek, F. The Effect of Granule Microstruc-
ture on Dissolution Rate. Powder Technol. 2008, 181, 104.
(20) Rough, S. L.; Wilson, D. I.; York, D. W. Effect of Solids
Formulation on the Manufacture of High Shear Mixer Agglomerates.
Adv. Powder Technol. 2005, 16, 145.
(21) Stepanek, F.; Rajniak, P. Droplet Morphologies on Particles
with Macroscopic Surface Roughness. Langmuir 2006, 22, 917.
(22) Adams, M. J.; Perchard, V. The Cohesive Forces between
Particles with Interstitial Liquid. Int. Chem. Eng. Symp. Ser. 1985, 91.
(23) Lian, G.; Thornton, C.; Adams, M. J. A Theoretical Study of the
Liquid Bridge Forces between Two Rigid Spherical Bodies. J. Colloid
Interface Sci. 1993, 161, 138.
(24) Grof, Z.; Lawrence, C. J.; Stepanek, F. The Strength of Liquid
Bridges in Random Granular Materials. J. Colloid Interface Sci. 2008,
319, 182.
(25) Leger, L.; Silberzan, P. Spreading Drops on Solid Surfaces.
J. Phys.: Condens. Matter 1990, 2, SA421.
(26) Kumar, G.; Prabhu, K. N. Review of Non-reactive and Reactive
Wetting of Liquids on Surfaces. Adv. Colloid Interface Sci. 2007, 133, 61.
(27) Leger, L.; Joanny, J. F. Liquids Spreading. Rep. Prog. Phys. 1992,
55, 431.
(28) Zadrazil, A.; Stepanek, F.; Matar, O. K. Droplet Spreading,
Imbibition and Solidification on Porous Media. J. Fluid Mech. 2006,
562, 1.
(29) Craster, R. V.; Matar, O. K. Dynamics and Stability of Thin
Liquid Films. Rev. Mod. Phys. 2009, 81, 1131.
(30) Stepanek, F.; Loo, A.; Lim, T. S. Multiscale Modelling Meth-
odology for Virtual Prototyping of Effervescent Tablets. J. Pharm. Sci.
2006, 95, 1614.
(31) Fredd, C. N.; Fogler, H. S. Influence of Transport and Reaction
on Wormhole Formation in Porous Media. AIChE J. 1998, 44, 1933.

11584 dx.doi.org/10.1021/ie201047r |Ind. Eng. Chem. Res. 2011, 50, 11576–11584

You might also like