Chemical Engineering Science: Marcello Murru, Giovanni Giorgio, Sara Montomoli, Francois Ricard, Frantisek Stepanek
Chemical Engineering Science: Marcello Murru, Giovanni Giorgio, Sara Montomoli, Francois Ricard, Frantisek Stepanek
a r t i c l e i n f o abstract
Article history: An integrated methodology for the scale-up of vacuum contact drying with intermittent agitation is
Received 10 May 2010 described in this work. The methodology combines a mathematical model of vacuum contact drying,
Received in revised form based on differential transient heat and energy balances, and a small-scale experimental apparatus for
7 April 2011
model validation and parameter estimation. The validated model was used for the estimation of drying
Accepted 28 June 2011
Available online 5 July 2011
times of six different pharmaceutical compounds at the pilot and manufacturing scale over a range of
drying conditions – pressure 15–200 mbar, temperature 45–70 1C, solvents: acetone, water, methanol,
Keywords: n-propanol, and isopropyl acetate. The mean difference between predicted and actual drying times for
Drying the six compounds was less than 9%, which is considered a significant improvement over current semi-
Heat transfer
empirical approaches to vacuum contact drying scale-up.
Mass transfer
Crown Copyright & 2011 Published by Elsevier Ltd. All rights reserved.
Mathematical modelling
Scale-up
Pharmaceutical processing
0009-2509/$ - see front matter Crown Copyright & 2011 Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2011.06.059
5046 M. Murru et al. / Chemical Engineering Science 66 (2011) 5045–5054
aim of this work was to validate the model for several real
pharmaceutical compounds under a range of operating conditions
(pressure, temperature) and for different solvents, and to demon-
strate its use in an integrated methodology for process scale-up
that involves parameter identification at the laboratory scale and
the prediction of drying times at the pilot or manufacturing scale.
The model is also shown to be easily extensible to vacuum drying
with intermittent agitation of the filter cake.
surface diffusion or desolvation are important, the model would Since WS is assumed to be constant and WL is a state variable
have to be extended to take these phenomena into account. and therefore known in each time and location, WG can be
calculated as a complement to one. The local dependence of the
2.2. Governing equations effective thermal conductivity on the solvent content is calculated
from the thermal conductivity of the dry and fully liquid-satu-
Based on the above assumptions, the governing system of rated porous medium (filter cake) using an expression proposed
partial differential equations is composed of a mass balance in by Kohout et al. (2004)
the liquid phase and an energy balance on the cake (liquid and WL
solid), which are coupled by a kinetic expression for the evapo- leff ¼ ldry þ ðlwet ldry Þ ð7Þ
1WS
ration rate.
Liquid-phase mass balance: Let us now turn to the description of liquid-phase mass
transfer. Assuming capillary flow to be the dominant mechanism
@WL _ LG
m
¼ rUðDL rWL Þ ð1Þ of liquid phase migration in the filter cake, the mass flux of the
@t rL
liquid phase is given by Darcy’s law
Energy balance: rL
jL ¼ kðWL ÞrpL ð8Þ
@T mL
/rCp S _ LG DHvap
¼ rUðleff rTÞm ð2Þ
@t
where k(WL) (m2) is the relative permeability, mL (Pa s) is the
In the above equations, WL (dimensionless) is the liquid phase liquid viscosity, and pL (Pa) is the liquid-phase pressure, related to
volume fraction, T (K) is temperature, t (s) is time, DL (m2 s 1) is the gas-phase pressure through the capillary pressure, pC:
the apparent liquid phase diffusion coefficient (more on this
pC ðWL Þ ¼ pG pL ð9Þ
coefficient below), leff (W m 1 K 1) is the effective thermal
conductivity, rL (kg m 3) is the liquid phase density, rCp The capillary pressure is generally a non-linear, decreasing
3 1 1
(J m K ) is the mean volumetric heat capacity, DHvap (J kg ) function of the moisture content. After substitution into Eq. (8)
is the latent heat of vaporisation, and m _ LG (kg m 3 s 1) is the and differentiation, we obtain the Fick-like moisture transport
local evaporation rate. We postulate that the evaporation rate is term occurring in Eq. (1), which is also known as the Richards
proportional to the difference between the equilibrium and the equation in the porous media literature (e.g. Daian and Saliba,
current vapour pressure, i.e., 1993).
8
< kvap rL ðp pG Þ=pG if WL 40
n
> jL ¼ rL DL ðWL ÞrWL ð10Þ
m_ LG ¼ 0 if WL r0 ð3Þ
>
:0 The apparent moisture diffusivity in the unsaturated porous
if pn r pG
medium
where kvap (s 1) is a rate constant, pG (Pa) is the actual local
kðWL Þ @pc ðWL Þ
vapour pressure, and pn(T) is the vapour pressure that would be in DL ð W L Þ ¼ ð11Þ
mL @ WL
equilibrium with the local temperature. pn(T) can be evaluated
from the Antoine equation is an a priori unknown function that combines the product of
permeability and the derivative of capillary pressure with respect
B
log10 pn ¼ A ð4Þ to moisture content. Various expressions for the dependence of
C þT
the relative permeability and the capillary pressure on solvent
The constants A, B and C can be found for common solvents e.g. content in an unsaturated porous medium have been proposed in
in Smallwood (2006). Due to the assumption of zero resistance to the literature (see e.g. Mahadevan et al., 2006); their parameters
vapour phase mass transfer in the present model, the vapour have to be determined empirically from experimental data. In this
phase pressure pG is constant and equal to the head-space work, we use a simple two-parameter linear relationship of the
pressure in all points of the filter cake. form
The kinetic expression for the evaporation rate m _ LG has been (
introduced mainly for numerical reasons; if the value of kvap is aðWL WnL Þ if WL Z WnL
DL ¼ ð12Þ
if WL o WL
n
chosen sufficiently large then the local vapour–liquid equilibrium 0
is essentially instantaneous and the solution becomes indepen- n
where WL is the residual saturation (for which the capillary
dent of kvap as vaporisation is not the rate-limiting step. However,
pressure diverges) and a is a proportionality constant. Both
the governing set of PDE’s (Eqs. (1)–(2)) can still be integrated as a
adjustable parameters have to be determined by fitting labora-
differential rather than a differential-algebraic system. It should
tory-scale experimental data; however, as material properties
be noted though, that in a general case – e.g. when drying porous
they should be preserved upon scale-up.
particles, hygroscopic materials, solvates or other systems with
intra-particle mass-transfer – the rate-limiting step may well be
the local evaporation rate and in that case the rate constant kvap 2.3. Boundary conditions
would have to be treated as an adjustable parameter with an
influence on the overall drying rate. The governing PDE’s (Eqs. (1) and (2)) were integrated
To complete the set of governing equations, the mean volu- numerically on a spatially 2D domain in cylindrical coordinates,
metric heat capacity is calculated from the pure components and taking advantage of the axial symmetry of the Nutsche filter dryer
the phase volume fractions according to (cf. Fig. 1). The COMSOL Multiphysics finite element software
package (http://www.comsol.com) was used for the numerical
rCp ¼ WL rL Cp,L þ WS rS Cp,S þ WG rG Cp,G ð5Þ
integration of the PDE’s. The following boundary conditions were
where ri and Cp,i are the density and the specific heat capacity, applied:
respectively, of each phase. The phase volume fractions must at r ¼0 (axis):
satisfy
@WL @T
¼ ¼0 ð13Þ
WL þ WS þ WG ¼ 1 ð6Þ @r @r
5048 M. Murru et al. / Chemical Engineering Science 66 (2011) 5045–5054
@T @T and
leff ¼ hq ðTh TÞ or leff ¼ hq ðTh TÞ ð14bÞ Z
@r @z 1
W~ L ¼ WL dV ð19Þ
at z¼H (top of filter cake): V V
by single crystal X-ray diffraction or by determining the unit cell measuring cylinder and calculated from its definition:
and space group symmetry from a powder X-ray diffraction mS
pattern. Density is related to the unit cell volume by the following rb ¼ ð21Þ
Vb
equation:
where mS is the mass of the powder sample and Vb the bulk
Mw Z volume occupied by it.
rS ¼ ð20Þ Thermal conductivity and heat capacity of both loose and
NA V0
compact powder were measured by the transient hot wire
where rS is the true density of the material, Mw is the molecular method. The measurement was easily carried out using the KD2
mass, Z is the number of molecules in the unit cell, NA is Avogadro’s Pro thermal properties analyser (Decagon Devices, Pullman, USA)
number, and V0 is the volume of the unit cell. The value of density equipped with a 3 cm dual needle sensor. The Arctic silver paste
calculated with this method was used in the simulation and for the was applied on the sensor to allow for a better thermal contact
calculation of the cake void and solid volume fractions. between the probe and the powder.
The bulk density of the compressed powder (filter cake) has to
be measured in situ during the drying experiments after blow- 3.2. Laboratory-scale experiments
down. This value should be used for static drying case as the cake
remains compressed during drying without mechanical agitation. Laboratory-scale static drying experiments are a crucial com-
However, in case the cake is agitated, a bulk density of the loose ponent of the overall scale-up methodology as they allow the
powder is appropriate due to ‘‘aeration’’ effect of the agitation. For fitting of the liquid effective diffusivity parameters. The experi-
the loose powder the bulk density was obtained simply by ments were carried out in a 500 ml glass filter dryer whose set-up
measuring the weight and the volume of the powder in a configuration is shown schematically in Fig. 2. In order to obtain a
drying curve, the vapour leaving the jacketed vessel (the dryer)
was condensed using a condenser inserted in the vacuum line
before the vacuum pump equipped with a vacuum controller. The
mass of the collected condensate as function of time was recorded.
The temperature of the cake was measured by thermocouples and
all the equipment was connected to a computer with data
acquisition software for all the experiments. An overhead stirrer
was embedded in the filter cake to facilitate agitation when
required.
Before the vacuum drying was started, the cake depth was
measured in order to know the value of H for boundary condi-
tions, and the initial moisture content of the cake – the loss on
drying (LOD) – was determined for the initial conditions.
The actual vacuum drying experiment then involves setting
the desired heating jacket temperature and the vacuum level, and
starting the agitator if required (intermittent or continuous). The
mass of condensate as function of time is recorded by means of
computer data logging Apart from the mass of condensate, other
parameters that are logged are: jacket temperature, head-space
pressure, cake temperature (noting the exact position of the
Fig. 2. Set-up of a 500 ml glass filter dryer used in the lab scale experiments. thermocouples in the cake). The cake depth before and after
Table 1
Summary of scale-up predictions and actual drying times at pilot plant or manufacturing scale.
Compound, cake volume, Predicted drying Actual drying a (m2 s 1), WnL (dimensionless), pG, solvent, Th
agitation mode time (h) (model) time (h) (experiment) ldry (W m 1 K 1)
drying, and the mass of solid charged and discharged are also 60
manually recorded.
55
40
4.1. Model-based parametric sensitivity study
35
A preliminary study on the effect of the cake parameters on
30
drying time was conducted in order to identify the most sensitive
parameters and the required accuracy of their measurements. The 25
parametric sensitivity study was conducted on static drying cases
varying the cake properties in the following ranges: 20
1000 1250 1500 1750 2000 2250 2500 2750 3000
Thermal conductivity of the cake 0–10–6 W m 1 K 1 Cp [J Kg-1 K-1]
Apparent diffusion coefficient 0.02–0.2 m2 s 1
Solid heat capacity 1000–3000 J kg 1 K 1 Fig. 5. Effect of solid heat capacity on drying time, simulation results. Cake radius
40 cm, cake height 10 cm. Solid apparent diffusion coefficient 0.1 m2/s, thermal
conductivity 0.168 W m 1 K 1. Solvent isopropyl acetate and initial solvent
The values used for the study are within a typical range for content 15% w/w.
60 60
55
50
50
Drying time [h]
Drying time [h]
45 40
40 30
35
20
30
25 10
20 0
0 2 4 6 8 10 12 0 50 100 150 200 250
10-7 * Apparent diff. coeff. [m2/s] Headspace pressure [mbar]
Fig. 3. Effect of diffusion coefficient on drying time, simulation results. Cake radius Fig. 6. Effect of head-space pressure on drying time, simulation results. Cake
40 cm, cake height 10 cm. Solid effective thermal conductivity 0.168 W m 1 K 1, radius 40 cm, cake height 10 cm. Solid apparent diffusion coefficient 0.1 m2/s,
solid heat capacity 2060 J kg 1 K 1. Solvent isopropyl acetate and initial solvent thermal conductivity 0.168 W m 1, solid heat capacity 2060 J kg 1 K 1. Solvent
content 15% w/w. isopropyl acetate and initial solvent content 15% w/w.
M. Murru et al. / Chemical Engineering Science 66 (2011) 5045–5054 5051
n
scale). The procedure to regress the parameters a and WL can be 0.25
shown for one of the compounds (GSKC#1) with 1-propanol
solvent, studied as an example. The drying curve resulting from a
0.2
static drying experiment at lab scale (10 cm diameter, 4 cm cake
depth) at Th ¼50 1C and pG ¼ 40 mbar is shown in Fig. 8. The drying
model was run while systematically varying the values of a and WL
n 0.15
in order to minimise the integral difference between the experi-
mental and calculated data points. 0.1
n
The optimised parameters a and WL were then used for drying
simulation at different scales and conditions provided that the 0.05
cakes had the same physical properties. Before scale-up, the data
were verified using another laboratory-scale experiment with the 0
same compound and solvent, but dried under different conditions 0 1000 2000 3000 4000 5000
(Th ¼80 1C and pG ¼30 mbar). The experimental and the calculated Time [sec]
The model was run at the pilot or manufacturing scale using ldry for the uncompressed (loose) powder was used. Table 1
the same parameters used for the lab scale runs, just by changing summarises the results obtained from the simulations detailing
the geometry (radius and depth) of the cake. In case of inter- the conditions (pressure and temperature), the scale (cake
mittent agitation the value of the effective thermal conductivity volume) and the values of the moisture transport parameters a
n
and WL used in the simulations.
As can be seen in Table 1, although the predicted and
0.35 measured drying times do deviate to some extent, these differ-
ences are very small compared to prevailing practice where semi-
0.3 empirical methods of drying end-point estimation often lead to
Solvent content w/w
Fig. 11. Solvent content and temperature profiles for GSKC#1, static drying— Fig. 12. Apparent moisture diffusion coefficient and local evaporation rate profiles
simulation results. for GSKC#1, static drying—simulation results.
M. Murru et al. / Chemical Engineering Science 66 (2011) 5045–5054 5053
0.35 305
0.3 300
0.25 295
Temperature [K]
0.2 290
0.15 285
0.1 280
0.05 275
0 270
0 0.5 1 1.5 2 2.5 3 3.5 4
Time [hour]
Fig. 13. Simulation results for intermittent agitated drying at pilot plant scale (31 L cake) of GSKC#4.
0.049 the dry layer, and moisture migration by capillary flow from the
10 min wet to the dry region (cf. Lehmann et al., 2008).
0.042 30 min As seen in Fig. 10, the drying rate in a static drying case
1 hour decreases significantly with time, resulting in very long drying
Solvent content w/w
0.035 static time at large scale. This behaviour is more evident at larger scale
because of the relatively larger proportion of the dry section of
0.028 the cake in contact with the wall that is offering high heat transfer
resistance (lower surface to volume ratio after scale-up). Agitation
0.021 is crucial at large scale in order to reduce drying time by
effectively contacting the wet powder with the wall.
0.014