You are on page 1of 16

Pharmaceutical Development and Technology, 1:17–32, 2005

Copyright D 2005 Taylor & Francis Inc.


ISSN: 1083-7450 print / 1097-9867 online
DOI: 10.1081/PDT-200035869

The Nonsteady State Modeling of Freeze Drying: In-Process Product


Temperature and Moisture Content Mapping and Pharmaceutical
Product Quality Applications
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

M. J. Pikal, S. Cardon, Chandan Bhugra, F. Jameel, and S. Rambhatla


School of Pharmacy, University of Connecticut, Storrs, Connecticut, USA

W. J. Mascarenhas and H. U. Akay


Technalysis Inc., Indianapolis, Indiana, USA

properties, product properties, and numerical analysis parameters


Introduction. Theoretical models of the freeze-drying required for the finite element analysis. Most input data are either
process are potentially useful to guide the design of a freeze- available in the literature or may be easily estimated. Product
drying process as well as to obtain information not readily resistance to water vapor flow, mass transfer coefficients
accessible by direct experimentation, such as moisture distribution describing secondary drying, and container heat transfer coef-
and glass transition temperature, Tg, within a vial during ficients must normally be measured. Each element (i.e., each small
For personal use only.

processing. Previous models were either restricted to the steady subsystem of the product) may be assigned different values of
state and/or to one-dimensional problems. While such models are product resistance to accurately describe the nonlinear resistance
useful, the restrictions seriously limit applications of the theory. behavior often shown by real products. During primary drying, the
An earlier work from these laboratories presented a nonsteady chamber pressure and shelf temperature may be varied in steps.
state, two-dimensional model (which becomes a three-dimen- During secondary drying, the change in gas composition from pure
sional model with an axis of symmetry) of sublimation and water to mostly inert gas is calculated by the model from the
desorption that is quite versatile and allows the user to investigate a instantaneous water vapor flux and the input pumping capacity of
wide variety of heat and mass transfer problems in both primary the freeze dryer. Results. Comparison of the theoretical results
and secondary drying. The earlier treatment focused on the with the experiment data for a 3% sucrose formulation is generally
mathematical details of the finite element formulation of the satisfactory. Primary drying times agree within two hours, and
problem and on validation of the calculations. The objective of the product temperature vs. time curves in primary drying agree
the current study is to provide the physical rational for the choice within about ± 1°C. The residual moisture vs. time curve is
of boundary conditions, to validate the model by comparison of predicted by the theory within the likely experimental error, and
calculated results with experimental data, and to discuss several the lack of large variation in moisture within the vial (i.e., top vs.
representative pharmaceutical applications. To validate the model side vs. bottom) is also correctly predicted by theory. The
and evaluate its utility in studying distribution of moisture and theoretical calculations also provide the time variation of ‘‘Tg – T’’
glass transition temperature in a representative product, calcu- during both primary and secondary drying, where T is product
lations for a sucrose-based formulation were performed, and temperature and Tg is the glass transition temperature of the
selected results were compared with experimental data. Theo- product phase. The calculations demonstrate that with a secondary
retical Model. The model is based on a set of coupled differential drying protocol using a rapid ramp of shelf temperature, the
equations resulting from constraints imposed by conservation of product temperature does rise above Tg during early secondary
energy and mass, where numerical results are obtained using finite drying, perhaps being a factor in the phenomenon known as ‘‘cake
element analysis. Use of the model proceeds via a ‘‘modular shrinkage.’’ Conclusion. The theoretical results of in-process
TM
software package’’ supported by Technalysis Inc. (Passage / product temperature, primary drying time, and moisture content
Freeze Drying). This package allows the user to define the problem mapping and history are consistent with the experimental results,
by inputing shelf temperature, chamber pressure, container suggesting the theoretical model should be useful in process
development and ‘‘trouble-shooting’’ applications.

Received 14 January 2004, Accepted 27 March 2004. Keywords non-steady state modeling of lyophilization,
Address correspondence to M. J. Pikal, School of Pharmacy, temperature and moisture distribution in freeze
University of Connecticut, Storrs, CT 06269, USA. Fax: (860) drying, lyophilization process, freeze drying, heat
486-4998; E-mail: pikal@uconnvm.uconn.edu and mass transfer in freeze drying

17
Order reprints of this article at www.copyright.rightslink.com
18 M. J. Pikal et al.

INTRODUCTION temperature profiles for primary drying, and therefore,


is of great use in the rational design of a freeze-drying
Freeze drying[1] is a process by which a solution, process as well as trouble-shooting-applications, such a
normally aqueous, is frozen and the solid solvent (i.e., model does not address residual moisture in the dry
ice) is removed by direct sublimation during ‘‘primary layer nor can one obtain information for the nonsteady
drying’’ in a vacuum environment. While most of the state parts of primary drying, such as the time im-
solvent (i.e., water) does convert to the crystalline solid, mediately following a change in shelf temperature.
some solvent does not crystallize and must be removed Nonsteady-state models of sublimation and desorption
by desorption drying during ‘‘secondary drying.’’ While drying are needed for a complete description of the
the operational definition of secondary drying is normally drying process. Such models are based on a set of
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

given as the portion of the drying process after all ice has coupled differential equations and are therefore, much
been removed, it is well known[1] that secondary drying more complex.[3]
actually begins in a section of the product once ice is We have previously[4] described the mathematical
removed from that local section. At least for pharmaceu- details of an extension of the one-dimensional Liapis and
tical products, primary drying is normally the longest part Millman nonsteady-state sublimation:desorption mod-
of the process, often requiring days for completion. el.[3] This extension is a finite element analysis of
Because of the long process times and the high cost of drying and is a two-dimensional model that can address
production freeze drying plants, there is considerable spatial distribution of temperature and residual moisture
interest in developing processes that minimize process in a more realistic fashion. In addition, the extension
time and still maintain product quality. Often such differs from the Liapis:Millman model in the detail of
process development is accomplished empirically, and several of the boundary conditions, offering a more
it is an unfortunate fact that many manufacturing physical set of boundary conditions. The earlier treat-
processes are far from optimal. Furthermore, problems ment focused on the mathematical details of the finite
For personal use only.

often develop, particularly during scale-up, and solutions element formulation of the problem and on validation of
need to be found quickly. Again, empiricism is often the the calculation by comparison of one-dimensional results
only approach. on skim milk with the Liapis:Millman study. The
Theoretical modeling can provide a better under- objective of the current study is to provide the physical
standing of the impact of process and formulation rational for the choice of boundary conditions, to validate
variations on cycle time and product temperature history the model by comparison of calculated results with
and, thereby, facilitate process development. Likewise, experimental data, and to discuss several representative
trouble shooting manufacturing problems can be aided by pharmaceutical applications.
theoretical modeling approaches. Of course, experiments
are essential, but experiments whose design and interpre-
tation are guided by theory are often a much more MATERIALS AND METHODS
efficient use of time.
Fundamentally, optimizing and/or trouble shooting Materials
the drying portion of a freeze-drying process is an
exercise in coupled heat and mass transfer. For every Crystalline Sucrose was obtained from Sigma
gram of ice sublimed, about 2700 joules of heat must Aldrich Co. The vials used were 5 mL and 10 mL
be supplied and roughly 1500 liters of water vapor must tubing vials with a 20mm neck size designed for freeze
flow through the previously dried porous product or cake drying (Wheaton Company). For the 10 mL vials, the
and ultimately be converted to ice at the condenser. inner diameter was 22 mm, and the outer diameter was
For most of the primary drying stage, the process is 24 mm. The 5 mL vials had an inner diameter of
‘‘pseudo steady state,’’ meaning that as a good first 19.2 mm and an outer diameter of 21.6 mm. Stoppers
approximation, the heat input from the shelves to the were 20 mm finish gray butyl stoppers (West Pharma-
product exactly balances the heat removed by sublima- ceuticals) designed for lyophilization.
tion (i.e., the heat needed to change the temperature of
the product is negligible). Indeed, a simple steady-state
model can quantitatively describe primary drying as long Freeze Drying
as accurate experimental or estimated heat and mass
transfer coefficients are input into the model.[2] While Two different freeze-drying operations were per-
the simple steady-state model of sublimation drying[2] formed. In the first experiment (Process 1) 10 mL vials
does accurately predict drying times and product were filled with 1.55 mL of 5% w/v sucrose solution and
Nonsteady State Modeling of Freeze Drying 19

loaded onto the shelf of a laboratory scale Durastop tory freeze dryer used in these studies, the ‘‘vial/shelf’’
freeze dryer (Kinetics Thermal Systems). The freeze- heat transfer coefficients should be only on the order of
drying procedure involved a normal freezing cycle where 10% less than the vial heat transfer coefficient.[2,7] For the
the shelf temperature was first lowered to 5°C at a ramp 10 mL vials, the heat transfer coefficient was determined
rate of 2.5°C/min and held for 15 min, then lowered to from the process data for the freeze drying runs with
 5°C and held for another 15 mins, and then lowered to sucrose. Knowing the fill volume and the solution
– 40°C at a ramp rate of 1°C/min and held for 60 min. concentration, one may calculate the total amount of
During primary drying, the shelf temperature was raised water removed during freeze drying, which then allows
to – 20°C and the chamber pressure was set to 100 mTorr the calculation of the total heat required to sublime this
and held until the end of primary drying (nearly 13 amount of water, DQ, from the known heat of sublima-
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

hours). During secondary drying, the shelf temperature tion. The duration of primary drying, Dt, is estimated by
was raised to 33°C at a ramp rate of 0.28°C/min and held the product thermocouple response. Primary drying is
until the end of the run. In the second operation (Process taken as the total time between the decrease of chamber
2), 5 mL vials were used, and 2 mL of 5% w/v sucrose pressure that marks the start of primary drying and the
solution was filled into each vial. While the freezing time when all product thermocouples in vials in the
process was similar to the previous run, a lower shelf interior of the vial array have reached (within 1°C of) the
temperature of –30°C was maintained during primary shelf temperature. Mass loss from the start of pump down
drying. The ramp rate towards secondary drying was until the pressure approaches the control pressure (100
increased to 0.5° C/min. mTorr) is negligible and is ignored here. Thus, the vial
heat transfer coefficient is calculated from the average
sublimation rate, DQ/Dt, the time average temperature
difference between shelf and product, h Ts –Tpi, and the
Dry Layer Resistance Measurement
area of the vial, Av, as evaluated from the vial outer
For personal use only.

diameter, from the expression,


Dry layer resistance measurements were made using
the manometric temperature measurement (MTM) proce-
dure.[5,6] The procedure consists of quickly closing the DQ=Dt
Kv ¼
valve connecting the drying chamber with the condenser Av hTs  Tp i
chamber, monitoring the pressure vs. time data generated,
and fitting the pressure rise data to the theoretical
This procedure yields data that is of useful accuracy
equations describing the pressure rise. This procedure
even though the heterogeneity in drying behavior causes
gives the dry layer resistance, essentially from how fast
some ambiguity in defining the end of primary drying.
the pressure increases initially, and also gives the
The heat transfer coefficient for the 5 mL vials was
temperature of the ice at the sublimation interface.
determined using a procedure similar to the one
described previously for the 10 mL vials, except that
in these experiments, the freeze-drying runs were
Vial Heat Transfer Coefficient Measurement stopped well short of the end of primary drying, and
the mass of ice sublimed was calculated as the
While vial heat transfer coefficients are most difference between the initial and final masses of a
accurately determined with special experiments,[7] data selected group of vials. This procedure was carried out
with acceptable accuracy can also be obtained from during freeze drying of several products at several
routine freeze-drying experiments; and it was these latter chamber pressures spanning 100 mTorr, and the value
procedures we used in this research. We also note that, for 100 mTorr was interpolated from smoothed raw data.
since in the context of both the experiments reported here For the 10mL vials used, the value of vial heat transfer
and in the theory, the ‘‘vial heat transfer coefficient’’ is coefficient was found to be: 104Kv = 3.0± 0.1 cal s 1
based upon the shelf temperature ‘‘set point,’’ rather than cm 2 K 1, where the uncertainly represents the standard
on shelf-surface temperature. Thus, the vial heat transfer error for three measurements. For the 5 mL vials, a
coefficients discussed and reported here really reflect value of 3.7± 0.1 was found for 104Kv. These values are
resistance to heat transfer from shelf surface to vial lower than were previously found for tubing vials.[7]
bottom plus resistance to heat transfer from heat transfer While some of this difference is due to the fact that the
fluid to shelf surface. That is, in the context of the present vial heat transfer coefficient data reported here also
work, the vial heat transfer coefficients are really vial/ includes thermal resistance in the shelf, the values
shelf heat transfer coefficients. However, for the labora- obtained do seem low, particularly for the 10 mL vials.
20 M. J. Pikal et al.

We have no explanation except to note the poor vial RESULTS AND DISCUSSION
bottom:shelf contact with these vials.
The Theoretical Model

Residual Moisture The Mass and Heat Transfer Equations

To obtain the kinetics of secondary drying, samples Here, we give only the general forms of the heat
were removed from the freeze-dryer shelf at several times and mass transfer equations with an ‘‘intuitive’’
during secondary drying using a sample extractor. The interpretation. The notation used is defined in the
sealed samples were then assayed for water content using ‘‘nomenclature’’ section. The reader is referred to the
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

a coulometric Karl Fischer Titrator (Orion, AF7LC). literature[4] for additional mathematical detail. In a finite
Samples were dissolved in 1 mL of ‘‘dry’’ methanol and element model, the physical system is divided up into a
about 0.5 mL of solution was injected into the titrator. number of subsystems or ‘‘elements,’’ most being
Knowing the weight of the sample and the amount of roughly rectangular in geometry, and the heat and mass
water in dry methanol, the water content in the sample transfer equations are solved for each element. In our
could be determined. case, each element can, in principle, have different
physical properties. The assembly of elements define a
‘‘grid,’’ which in our case is a two-dimensional grid
Water Sorption Isotherms comprising on the order of 100 elements, which with an
axis of symmetry can represent a three-dimensional
Roughly 10 mg of sample was loaded onto a quartz cylindrical container such as a vial. There is a boundary
pan enclosed in a high vacuum electronic microbalance representing the sublimation interface that separates
apparatus (Sartorius). The apparatus consists of a ‘‘dry region’’ elements from the ‘‘frozen region’’
For personal use only.

microbalance and thermostats for regulating the temper- elements. As drying proceeds, the grid of elements is
ature of the sample (T1) and of water (T2) contained in a regenerated periodically to form more dry elements at
tube connected by a side arm to the sample tube.[8] A the expense of frozen elements.
stepwise variation in the temperature of the T2 bath A schematic of the freeze-drying model, but without
enabled a stepwise change in solvent partial pressure. For defining the grid, is given in Figure 1. Heat flows into the
each step change in partial pressure, the equilibrium water bottom of the sample from the shelf by radiation heat
content of water was determined from the sample mass transfer, conduction at the points of contact between the
recorded under equilibrium conditions (i.e., when the vial and the shelf, and by gas phase heat transfer that is
sample mass remained invariant with time over several sometimes denoted as ‘‘convection’’ but which, in reality,
hours). Gravimetric data at equilibrium for each particular is conduction.[7] We also allow for heat transfer to the top
temperature and relative humidity were used to evaluate of the product and the sides of the container, such as
the desorption isotherms. The reader is referred to the would occur by radiation. In general, the model allows for
literature[8] for additional experimental details. Data were both water flux, Nw, and inert gas flux, Nin, through the
obtained over a range of relative humidities at 25°C (0 to dry layer, but we find that in all freeze-drying applications
21% relative humidity), 0°C (0 to 41% relative humidity), of pharmaceutical interest, one may ignore the inert gas
and  13.4°C (0 to 60% relative humidity). In addition, a flux without introducing significant error. The net flux of
data point corresponding to equilibrium between ice at water vapor is, of course, from the sublimation interface,
– 35°C and the sucrose freeze concentrate was evaluated shown here with slight curvature as found by calculation
based upon the following considerations. We find the and experiment.[7] The mass transfer equations in the dry
value of Tg’ in frozen aqueous sucrose is  35°C. At layer are given by Eqs. 1 – 4. Eq. 1 describes water flow in
 35°C, the activity of water is 0.70 (i.e., 70% relative the dry layer where the first term on the left hand side of
humidity), and based upon our analysis of literature glass the equation represents the change in water in the gas
transition data for sucrose-water systems (to be described phase, the second term gives the change in water in the
later), the glass transition is – 35°C at a concentration of solid phase, and the term on the right hand side is the flux
water in sucrose of 17.5%, w/w. Thus, to be able to of water out of the system. This equation essentially
provide a self-consistent description of the water expresses the conservation of mass for water.
desorption characteristics and glass transition character-
istics in the region near Tg’, we use data points of –35°C,
70% relative humidity, and 17.5% water in the construc- @Cw;g @Cw;s
e þ rI ¼ r  Nw ½1
tion of the water desorption isotherm. @t @t
Nonsteady State Modeling of Freeze Drying 21
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

Figure 1. Schematic of the freeze-drying model showing the sublimation interface and the modes of heat transfer.

The rate of change in water content of the solute phase is where Cw,g is the concentrations of water in the gas phase,
assumed to be proportional to the difference between the P is the total pressure, and the constants, ki (i = 1, 2, 3, 4),
water content and the water content that would represent are mass transfer coefficients defined earlier in the
For personal use only.

equilibrium with the surrounding water activity, aw, at nomenclature section. The terms on the right hand side
temperature T, denoted C*(aw,T), of Eq. 4 represent the contributions of both diffusion and
bulk fluid flow, the latter being mostly Knudsen flow (i.e.,
@Cw; s free molecular flow) in the usual pharmaceutical freeze-
¼ kg ðCw; s  C *ðaw ; TÞÞ ½2
@t drying application. Since the vapor composition in the
cake and in the drying chamber during any practical
where kg is a ‘‘rate constant’’ assumed to exhibit primary drying situation is nearly 100% water [7] one may
Arrhenius temperature dependence. C* must be deter- ignore diffusion in the dry layer during primary drying.
mined by experiment as a function of relative humidity Further, since the rate of secondary drying is normally
and temperature. C* as a function of water activity (or determined by rate processes that occur in the solid state
relative humidity) at selected temperatures is nothing other (i.e., solid-state diffusion) or at the solid:vapor boundary
than the water sorption isotherm as a function of (i.e., evaporation),[10] and since the evolution of the vapor
temperature. C* is normally described by either the composition above the dry layer is set by the flux
GAB Equation,[9] or a generalized combination of the dependent boundary condition (to be discussed later), one
Fruendlich isotherm and the Langmuir isotherm which we could ignore diffusion even in secondary drying without
write in the form, introducing serious error. Using the experimental obser-
  vation that the composition of vapor is essentially all
   n A  Pw
Ef 1 Pw L Pw0 water, Eq. 4 reduces to the simplified expression,
C * ¼ Af 0  exp  0
þ  
T Pw 1 þ BL  PPw0
w ½3  
    Mw C01 Pw Mw
Ef 2 Ef 3 Nw ¼   Kw 1 þ rPw   Kw rPw
AL ¼ AL0  exp ; BL ¼ BL0  exp RT Kw mmx RT
T T
½5
where Af0, Efi (i = 0, 2, 3), and n, AL0, and BL0 are
constants. Pw is the partial pressure of water, and P0w is the where the term proportional to C01, which is a viscous
vapor pressure of pure water at the sample temperature. flow term, is small compared to unity and may be
The water flux, Nw, is given by neglected. Thus, the only mass transfer coefficient that is
needed to address ‘‘normal’’ freeze-drying problems is
Nw ¼ k1 rCw; g  k2 Cw; g rP ½4 the Knudsen flow mass transfer coefficient, Kw.
22 M. J. Pikal et al.

Heat transfer in the dry layer is described by, transfer by gas phase conduction, increases as the
chamber pressure increases, and is normally the largest
@T @Cw;s single mode of heat transfer.[7] Thus, h1 and h3 would
rI CpI ¼ kI r2 T þ rI DHv  Cp;g rðNt TÞ ½6
@t @T normally be radiation heat transfer while h2 represents
heat transfer from the shelf to the bottom of the vial and
where the term on the left is the heat used to increase the would include all three heat transfer mechanisms.
local temperature, the first term on the right is the net heat For maximum accuracy in the modeling, hi are
flow into the local dry layer region, the second term is the evaluated from the experimental heat transfer coefficient,
heat needed to desorb water from the solute, and the last Kv, which is defined by,
term on the right hand side is the heat carried away from
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

the region by gas flow out. Eq. 6 is essentially the dQ


conservation of energy equation for a local dry layer ¼ Av Kv ðTs  Tp Þ ½10
dt
region. Heat transfer in the frozen layer is described by
where Av is the cross sectional area of the vial based upon
@T the outer diameter, Ts is the shelf surface temperature, and
rII CPII ¼ kII r2 T ½7 Tp is the product temperature in the bottom center of the
@t
vial. The vial heat transfer coefficient may be written in
where the term on the left is the heat used to increase the the form,
local temperature, and the term on the right is the net heat
flow through the frozen layer region. Kp P
Kv ¼ KK þ ½11
At the sublimation interface, conservation of energy 1 þ KD P
requires
where KK is a constant representing the sum of the
    radiation contributions (top and bottom) and the ‘‘con-
@T @T
For personal use only.

 kI þ kII ¼ Ntn Cp;g T  Nwn DHs ½8 tact’’ contribution. Experimental values of Kv are
@n I @n II normally reported for vials in the interior of an array of
where the first term on the left is the heat flow to the other vials where there is essentially no side radiation (i.e.,
interface from the dry region, and the second term is the h3 = 0). KD is a constant proportional to the average
heat flow to the interface from the frozen region. The first separation distance between the bottom of the vial and the
term on the right hand side is the heat removed by gas shelf, and KP is a constant that appears to be essentially the
flow from the interface, and the second term is the heat same for all vials.[7] Experimental values of Kv have been
removed by sublimation of ice, which is normally the reported in units of ‘‘cal s 1cm 2K 1,’’ where pressure in
largest term on the right. Eq. 11 is in Torr.[7] Here, the value of Kp = 0.00332 cal s 1
cm 2 Torr 1, and the top radiation term (i.e., analogous to
Heat Transfer Boundary Conditions h1) is 0.84 10 4 cal s 1 cm 2 for all vials. To obtain the
value of h2 from experimental data we must first calculate
The heat transfer boundary conditions are given by, the portion of the experimental Kv that represents only
bottom heat transfer, denoted Kvb, by subtracting the top
q1 ¼ h1 ðT  Tup Þ top surface radiation term (0.84  10 4) from the experimental Kv.a
Secondly, heat transfer in the freeze-drying model
q2 ¼ h2 ðT  Ts Þ bottom surface ½9
a
q3 ¼ h3 ðT  Tw Þ side surface This procedure assumes the ‘‘conventional’’ description[7] of
radiation heat transfer in terms of a radiation heat transfer
where the units of the heat transfer coefficients, hi, are coefficient which is given in units of cal/sec, as 1  10 4ei (Ts –
power/area K (i.e., Kw/m2K) and qi are heat fluxes in Tp), where ei is the relevant emissivity, and the numerical factor,
power/area (i.e., KW/m2). In the low pressure environ- 10 4, is an approximation for ‘‘4 < T3>,’’ consistent with a mean
ment characteristic of freeze drying, heat transfer temperature of 264 K. In our studies, the mean temperature (i.e.,
coefficients contain contributions from three ‘‘mecha- of shelf and product) in primary drying is about  30°C and
about + 10°C in early secondary drying, giving 263 K as a mean.
nisms’’: 1) conduction through the solid state at the points
However, this simplification does partition too much of the heat
of contact between the vial and the shelf, 2) radiation heat
transfer to top radiation during primary drying and too little in
transfer from dryer surfaces (i.e., shelves above the vials secondary drying. Overall heat transfer is not impacted. The
and under the vials and the dryer walls), and 3) heat only impact on the present study is in the distribution of heat
transfer via collisions of gas molecules with the hot transfer between top and bottom, with the vial ‘‘bottom heat
surface (i.e., shelf) and the cold surface (i.e., bottom of transfer coefficient, Kvb’’ being in error by about  7% in
vial).[7] The latter mechanism which is essentially heat primary drying and + 7% in early secondary drying.
Nonsteady State Modeling of Freeze Drying 23

Figure 2. Schematic of elements along vial bottom and assignment of heat transfer coefficients, h21.
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

(Figure 1) is based upon heat flow through the cross- and then rearranged to yield
sectional area of the product. Thus, the relationship be-
 
tween Kvb (cals 1 cm 2 K 1 units) and h2 (MKS units) is, h2c
h21 ¼ 1:385  hh2 i  1  0:278  ½15
  hh2 i
Av
hh2 i ¼ 41:876   Kvb ½12
Ap
with < h2> being given in terms of the experimental
where <h2> is the ‘‘average’’ heat transfer coefficient for quantity, Kvb by Eq. 12. The heat transfer coefficient for
heat flow across the vial bottom. The number, 41.876, elements 10 and 11 is calculated from the expression for
comes from the units’ conversion. The ratio of the vial the vial-bottom heat transfer coefficient, Kvb, where the
‘‘outer’’ area to vial inner or ‘‘product’’ area, Av/Ap, is separation distance is set equal to zero (i.e., KD = 0),
about 1.2– 1.3 for most tubing vials. The heat transfer
coefficient for the top boundary, h1, is given in terms of the Kvb ðKD ¼ 0Þ
emissivity for top radiation, ev (= 0.84), h2c ¼ hh2 i ½16
Kvb
For personal use only.

 
Av
h1 ¼ 41:876   ev  104 ½13 If the pressure dependence of the vial heat transfer
Ap coefficient has not been experimentally determined, the
While Eq. 11 could be used to specify the heat transfer value of KD will not be known. Here, we simply use the
coefficients all across the bottom boundary, we generally approximation that for tubing vials, h2c/ < h2> 1.5. With
choose to specify heat transfer coefficients that recognize this procedure, we force the overall heat transfer
the variation in heat transfer coefficient with variation in coefficient to correspond to the experimental value but
the separation distance between shelf and vial bottom yet allow spatial variation in heat transfer in the radial
according to the schematic shown in Figure 2 for a grid that direction to match physical reality.
has 12 elements in the radial direction (i.e., 12 elements
from the center of the vial to the edge). Generalization to The Flux Dependent Pressure Boundary Condition:
‘‘n’’ elements in the radial direction is obvious. Here, each Evaluation of P w
element has ‘‘radial length’’ (1/12) r, where r is the radius
of the product cake. As an approximation to the actual The theoretical model allows the drying process to be
contour for a vial bottom, we take elements 1 –9 and divided into a number of steps, with the chamber pressure
element 12 to have a roughly equal separation between the and shelf temperature held constant throughout each
vial bottom and the shelf, and have the same heat transfer individual step. As a boundary condition, it is necessary to
coefficient, h21. Elements 10 and 11 are taken to be in fix the value of the partial pressure of water above the dry
contact with the shelf (i.e., heat transfer coefficient layer, Pw. In conventional freeze drying practice, the
corresponding to KD [Eq. 10] = 0, and have the heat chamber pressure is held constant throughout the process.
transfer coefficient, h2c. The total heat flow into the vial, However, the composition of the gas in the drying
q2 = pr2 < h2>, is given by the sum of the heat transfer into chamber, and therefore, above the dry cake, changes
the various elements, which may be expressed in the form, during secondary drying from nearly all water vapor to
" nearly all inert gas as the product dries.[1] Thus, even at
 2 2  2 #
2 9 11 9 constant chamber pressure, the partial pressure of water
pr  hh2 i ¼ p r h21 þ p r  r  h2c changes with time during drying. In the theoretical model,
12 12 12
the partial pressure of inerts, Pin, is calculated as a
"  2 #
11 function of time during secondary drying from a
þ p r2  r  h21 ½14 knowledge of the total water flux coming from the
12
sample normalized to the total area of the dry cake (i.e.,
24 M. J. Pikal et al.

the area of the interface between the dry layer and the The molar flux of inerts may be written in the form,
‘‘outside’’) and the pumping characteristics of the freeze
dryer. The following analysis develops the appropriate ðPc  Jw  R0c Þ
Jin ¼ ½24
equations by an analysis of mass transfer through three ½R0c þ R0P ð1 þ k  Jw Þ
external barriers, or resistances.
which may, as a good approximation, be simplifiedb as
The barrier to total mass flow from the drying
chamber (i.e., outside the dry layer) to the condenser Pc
chamber, expressed in molar units, R’c , is given by Jin ffi ½25
R0P
ðPc  PCD Þ Note that the partial pressure of inerts is given by,
R0c ¼
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

½17
ðJw þ Jin Þ
Jin
Pin ¼ Pc  ½26
where PCD is the total pressure in the condenser chamber ðJin þ Jw Þ
and Jw and Jin are the molar fluxes of water and inerts,
respectively, in mol m 2 s 1. The resistance to flow from and the mole fraction of inerts in the drying chamber,
condenser chamber to surface of condenser plates, R’CD, is Xin, is given by,

ðPCD ice Jin


w  Pw Þ Xin ¼ ½27
R0CD ¼ ½18 ðJin þ Jw Þ
Jw
where Pice Calculations proceed by evaluation of Jw from the model
w is the vapor pressure of ice at the surface
temperature of the condenser plates. This pressure is and Jin from Eq. 24 [or Eq. 25]. Thus, Pin may be
normally small compared to all other pressures in the evaluated from Eq. 26, with Pw then being evaluated as
system, and we may set it equal to zero with little error. the difference, Pw = Pc Pin. The calculations start (in
early primary drying) with Pw = Pc, and thereafter the nth
For personal use only.

The ‘‘condenser resistance,’’ R’CD , is given in terms of the


partial pressure of inerts in the condenser chamber, PCD time step uses the previous calculated value of Pw.
in , by
The validity of Eqs. 25 –27 and this approach for
evaluation of vapor composition in the freeze dryer is
R0CD ¼ k  PCD
in ½19
demonstrated by Figure 3. Here, a double reciprocal plot,
based on Eq. 27, is given for a freeze drying run with 5%
where k is a constant characteristic of the freeze dryer.[2]
sucrose (Process 1). Such a plot should be a straight line
The resistance to pumping the inerts out of the condenser
with an intercept of 1.0, as observed (Figure 3), with the
chamber with the vacuum pump, denoted, R’p . is given by
slope giving Jin, from which the value R0P may be
PCD calculated from Eq. 25. The value found is 1.56 105 Pa s
R0P ¼ in
½20 m2 mol 1. The mass transfer constants are somewhat
Jin
specific for the dryer and product area loaded.b
After some algebra with the defining equations Eqs. 17 –19,
one may develop the working equations, which allow
Evaluation of Mass Transfer Coefficient for Water
calculation of Pin. First, the molar flux of water is related to
Vapor Transport
the mass flux by,
1000 Mass transfer in the dry layer is dominated by
Jw ¼  Nw ½21 Knudsen flow[1,2,12] , so the most critical mass transfer
Mw
coefficient is the Knudsen flow constant, Kw. The
where Mw is the molecular weight of water. The resistance Knudsen flow coefficient may be written in the form [3,4]
of the chamber may be written as a function of pressure in
the form[2,7] ,
Kw ¼ f C1 ðTtop þ Tinterface Þ1=2 ½28
1
R0c ¼ ½22
b1  P
where b1 is a constant and P is the mean pressure which b
Values of the constants for other dryers are as follows:
may be calculated from, R’p = 3.84  10 5 Pa s m 2 mol  1 ; k = 1.368 s m 2 mol  1 ,
( sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi) b1 = 6.59  10 3 mol (Pa) 2 s 1 m 2. The values of R’p and b1
Pc 2Jw are for a normal production freeze dryer.[2] The value for k is
P ¼  1þ 1 ½23 taken from experiments done with a small laboratory freeze
2 b1 P2c
dryer.[7]
Nonsteady State Modeling of Freeze Drying 25

where dm/dt is the sublimation rate, Ap is the cross-


sectional area of the dry layer, P0 is the vapor pressure of
ice, Pv is the pressure above the dry layer, and R ^ is the
p
area normalized product resistance. The usual units for R ^
p
2 1 ^
are cm hr Torr g . We find that Rp increases with an
increase in dry layer thickness, but not always in linear
fashion and the extrapolated value of R ^ at zero product
p
thickness is normally significantly higher than zero.[1] In
fact, the variation of resistance with thickness normally
fits an empirical expression of the form,
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

^ ¼ R þ A1 l
R p 0 ½30
1 þ A2 l
Figure 3. Mole fraction of inerts in secondary drying:
Comparison of values calculated from Eqs. 25 – 27 with where R0, A1, and A2 are constants and l is the thickness
experimental measurements from an electronic moisture sensor of the dry layer. Except for possible units differences,
(Ondyne). Values of Jw are evaluated from residual moisture vs. Kw is essentially the reciprocal of the derivative of R ^
p
time data for bovine somatotropin (BST) for a 1 cm fill depth
with respect to dry layer thickness. Note that Kw is a
(data from Ref. [11]). Freeze drying was with a Virtis 25 SRC-X ^ is linear in
constant throughout the dry layer only if R p
lab scale freeze dryer. Freeze drying conditions are: chamber
pressure = 0.1 Torr, shelf temperature = + 12°C, product temper- dry layer thickness.
ature  28°C in steady state primary drying. Product descrip- Calculation of the heterogeneity factor in such a way
as to maintain correspondence with the experimental R ^
tion is: 5% solids (BST) in solution with 0.19 g solids per vial p
after drying, maximum cake thickness = 1 cm., area of cake vs. l data may be done with the following procedure. For
each row, calculate the value of Kw for that row by,c
For personal use only.

surface = 3.8 cm2. The symbols are experimental data while the
straight line is a linear fit to the data.
0:0292
Kwi ¼ ½31
ðdRp =dlÞi
where Ttop and Tinterface are the temperatures (K) of the
top of the dry layer and the sublimation interface,
where the numerical value of 0.0292 represents units
respectively, C1 is a constant dependent on the structure ^ and cm for
conversion from the ‘‘usual’’ units used for R p
of the porous system, and f is the ‘‘heterogeneity
dry layer thickness to the mks units used in the theoretical
factor.’’ The heterogeneity factor allows for variation in
model for Kw. We next calculate an ‘‘average’’ value of
pore size and Knudsen flow constant as a function of
the Knudsen permeability, denoted Kwo, from the average
position in the cake. Thus, at an extreme, the model can
value of the derivative of Rp with respect to thickness, l,
accommodate a unique value of Kw for each element ^ /dli
denoted hdR p lmax, by the equation,
through assignment of a unique value for the heteroge-
neity factor, f, for each element. On the other extreme, if 0:0292
the permeability is constant throughout the system, all Kw0 ¼ ½32
^
hdRp =dlil max
elements would have the same value of Kw, and we
would set the heterogeneity factor equal to one. In The heterogeneity factor for row i is then calculated as,
practice, permeability usually does not vary greatly in
the radial direction, but significant variations in per- Kwi
fi ¼ ½33
meability occur as one moves from the top to the bottom Kw0
of the dry layer, so reality dictates we assign different
heterogeneity factors for each row of elements. We now
Evaluation of Desorption Parameters
consider how to set the heterogeneity factor for each row
to reproduce the experimental mass transfer behavior in
Experimental values of C* at various water
the Knudsen permeability format.
activities and temperatures were determined as described
We normally[1] experimentally describe mass transfer
earlier, the obtained experimental data were then fitted
in the dry layer in terms of product resistance, where
resistance is essentially defined by,
c
Dividing the dry layer into sections or ‘‘elements,’’ it is
dm ðP0  Pv Þ straightforward to show that the steady-state mass transfer rate is
¼ Ap ½29 ½PðlþDlÞPðlÞ
dt R^ given by, dm
dt ¼ ^ ^ ¼ C  Ap  ½PðlþDlÞPðlÞ
Dl  Kwi .
p ½ Rp ðlþDlÞ Rp ðlÞ
26 M. J. Pikal et al.

to the generalized sorption equation [Eq. 3], and the


parameters obtained are shown in Table 1. The
isotherms reproduced by these parameters fit the data
available reasonably well, but there were systematic
deviations (Figure 4). The deviations seem to be beyond
anticipated experimental error and probably reflect the
inability of the sorption equation to fit the data, even
though a large number of parameters were used. Note
that this treatment is only intended to reproduce and
extrapolate the experimental observations and yield a
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

self-consistent parameterization where the value of C* at


 35°C (i.e., Tg’) is forced to be 0.175 g/g, so that the
composition of the maximally freeze-concentrated solute
will be the equilibrium water content at Tg’. No physical Figure 4. Water desorption isotherms for amorphous sucrose
at selected temperatures: C* as a function of water activity and
meaning of the parameters evaluated is intended. The
temperature. The symbols represent experimental data (open
desorption parameter kg, was calculated using data
circle =  35°C, triangles =  13.4°C, squares = 0°C, and closed
obtained from one of the experiments denoted, Process circles = 25°C. The solid lines represent the corresponding data
1. To estimate the desorption coefficient kg, vials were calculated from the ‘‘best fit’’ parameters of Eq. 3.
removed from the freeze dryer at several time intervals
during the secondary drying stage of the freeze-drying
process. The concentration of unfrozen water C(t) was expected, kg did exhibit an Arrhenius temperature de-
assayed using a Karl-Fischer titrator. For each sampling, pendence (Figure 5).
values of the partial pressure of water vapor, Pw were
For personal use only.

calculated from the dew point measurement and the Finite Element Analysis Solution
vapor pressure of supercooled water, P0w was calculated
from the product temperature and the known relationship The preprocessor module of the finite element
between temperature and vapor pressure. These values analysis software involves the use of input parameters
were then used to calculate C*, i.e., from Eq. 3. At each to create the model for the freeze-drying analysis. A list of
time, the value of C(t), C*, and dC/dt (obtained by input parameters required to run the model is shown in
numerical differentiation) were calculated. Using Eq. 2, Table 2. While some parameters are standard handbook
the desorption parameter, kg, was then calculated. As values and do not change as the application changes,
some others are available by estimation with negligible
compromise on accuracy. For the effective heat capacity,
heat conductivity, and density of the dried material (i.e.,
Table 1
solute), one may use the corresponding property of any
Parameters for the generalized sorption model (Eq. 3) and
similar material, or the Passage default value,d with
parameters for a modified Gordon-Taylor equation describing
glass transition temperature as a function of water content, acceptable accuracy. Since the gas is always mostly water
shown in the table. The modified Gordon-Taylor equation vapor during primary drying, when heat transfer effects
was fit to literature data[13 – 15] over a wide range of are most critical, the standard values for water vapor may
water contents from zero to 25% be used for the heat capacity and heat conductivity of the
gas in the dry layer (i.e., the Passage default valuesb). The
½KTg  Tg2  Cw  ðKTg  Tg2  Tg1 Þ gases are the same in nearly all freeze-drying applications
Tg ¼ þ a1  Cw  ð1  Cw Þ
½Cw  ð1  KTg Þ þ KTg (i.e., water and nitrogen), so for the free gas mutual
þ a2  Cw2  ð1  Cw Þ
diffusivity constant, one may use the default value.b With
essentially no loss in accuracy, the values of heat
Parameter Value Parameter Value capacity, heat conductivity, and density for the frozen
layer may be taken as the weighted average of dry solid
Tg2 348.2 AL0 6.42  10 5 and ice values. Input parameters such as internal mass
Tg1 135 BL0 7.88  106
transfer coefficient for desorption, kg, and Knudsen flow
KTg 0.0920 Af0 4.26  105
a1 480.8 Ef1 1929
a2  1224 n 0.3 d
Ef2, Ef3  5028 The default values are the values from the skim milk problem
studied by Millman, Liapis, and Marchello.[3]
Nonsteady State Modeling of Freeze Drying 27
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

Figure 5. Arrhenius plot of the desorption constant evaluated from secondary drying data for Process 1. A fit to the Arrhenius
equation results in an activation energy Ea of 8,136 J/mol, and a preexponential factor of 3.34.10 3 s1.

constant, Kw, may be estimated but are best determined estimated or measured vial heat transfer coefficients, Kv,
from experiments on the system of interest, sucrose in this as described earlier.
case, using procedures described earlier. Heat transfer The solver or the processor module performs the
parameters, hi (i = 1, 2, 3), may be calculated from finite element analysis of the model using input files
For personal use only.

Table 2
Input parameters and property data for the model

Property Description Value

C0 1 Constant for viscous flow, depends on cake structure 0 (term not used)
C1 Constant for Knudsen flow, depends on cake structure 4.64  10 4
C2 Constant for diffusivity in cake, depends on structure 0.311
CpI Effective heat capacity of dry layer (kJ/Kg.K) 0.2595
CpII Effective heat capacity of frozen layer (kJ/Kg.K) 1.94
Cpg Effective heat capacity of gas in dry layer (kJ/Kg.K) 2.01
C0s, w Initial concentration of sorbed water(kg water/kg solid) 0.2059
k1 Constant for mass transfer C2D0w, inKw/(C2D0w, in+ KmxP)
k2 Constant for mass transfer KwKin/(C2D0w,in+ KmxPT) + C01/mmx
kI Effective thermal conductivity of dry layer 8.826  10 8*P + 2.706  10 5
kII Effective thermal conductivity of frozen layer 0.002763
kg Mass transfer coefficient for desorption (s 1) see Figure 5
Kw Knudsen diffusivity fC1(Ttop+ Tintf)^0.5
Min Molecular weight of inert gas (kg/kg – mole) 28.02
Mw Molecular weight of water (kg/kg – mole) 18.016
R Gas law constant g (J/kg – mole.K) 8314.0
DHs Heat of sublimation of ice (kJ/kg) 2834.6
DHv Heat of vaporization of bound water (kJ/kg) 2499.6
e Void fraction in the dried region 0.95
mmx Viscosity of the binary mixture of water vapor and inert gas term not used
rI Density of dry region (kg/m3) 55.0
rII Density of frozen region (kg/m3) 927.1
R0 Constant term in dry layer resistance equation (Eq. 30) 0.2
A1 Constant in dry layer resistance equation (Eq. 30) 67.9 (Process 1); 14.6 (Process 2)
A2 Constant in dry layer resistance equation (Eq. 30) 26.2 (Process 1); 1.51 (Process 2)
h1 Heat transfer coefficient for top radiation (MKS units) 0.0042
h21 Heat transfer coefficient for shelf-vial bottom (MKS units) 0.00873(Process 1), 0.01223 (Process 2)
h2c Heat transfer coefficient for shelf-vial bottom at ‘‘contact’’ (MKS units) 0.01622(Process 1), 0.02272 (Process 2)
28 M. J. Pikal et al.

Table 3
Comparison of experimental and theoretical primary drying times and temperatures. Experimental drying times are from dew
point sensor data, and experimental temperatures are mean temperatures from the start of primary drying to the
sharp increase in temperature indicating that thermocouple has lost contact with ice

Average product temperature in


Primary drying time (hours) primary drying, 8C

Freeze drying runs Experiment Passage Experiment Passage

Process 1 13.0 12.7  36.0  36.5


Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

Process 2 37.2 39.4  36.2  35.5

created in the preprocessor. The primary freeze-drying represents a comparison of experimental and calculated
part of the solver solves the heat and mass transfer product temperatures during drying for Process 2. Good
equations associated with primary drying while the agreement is observed with the difference between
secondary drying module solves the corresponding predicted and experimental product temperatures being
coupled equations associated with secondary drying. less than 1°C in the initial part of primary drying and
The output of the finite element analysis calculations throughout secondary drying. Likewise, the mean temper-
includes time and spatial distribution of residual water atures in primary drying (Table 3) are in good agreement.
contents, product temperature, vapor pressure and the For reasons we do not fully understand, the approach
difference between the glass transition temperature and towards the shelf temperature at the end of primary drying
product temperature (Tg  T). is more gradual for thermocouple-measured product
temperatures as compared to the sharp increase in
For personal use only.

Validation of the Model simulated product temperatures at the end of primary


drying. Perhaps some of this difference may be attributed
Product Temperatures and Drying Times . Vali- to the heat capacity of the glass in the vial, which the
dation of the model was carried out by comparison of data calculations ignore, and our approximations for thermal
obtained from an experiment with data obtained by conductivity and heat capacity in the dry layer. There also
simulation using Passage software. A comparison of may be some residual ice in the vial well after the sharp
primary drying times predicted by Passage and those break to higher temperature. In support of this specula-
obtained from experiment for Processes 1 and 2 is shown tion, we note that the end of primary drying as measured
in Table 3. Experimental primary drying times were by dew point response only occurs well after the
determined from the electronic moisture sensor response temperature break.
(i.e., a sharp decrease in dew point indicates the end of
primary drying). Good agreement between experiment
and theory is observed. It should be noted that primary
drying times differ according to position on the shelf.
Experiments conducted in our laboratory indicate that
vials located closer to the freeze dryer door have higher
sublimation rates and lower primary drying times. A
plausible explanation for this effect is atypical radiation
heat transfer that edge vials receive due to their direct
view of a warmer surface. Both experimental and
simulation results reported here refer to typical vials
located in the interior of the vial array (i.e., vials that are
enclosed in a hexagonal arrangement).
Product temperatures recorded by thermocouple
measurements during freeze drying were compared with
product temperatures predicted by Passage (Table 3 and
Figure 6). Theoretical product temperatures were taken as
the average of 1  2 elements, one radial and two Figure 6. Comparison of product temperatures determined
lengthwise, so as to compare with the dimensions of the from experiment (solid triangles) and passage (open circles) for
product thermocouple located inside the vial. Figure 6 Process 2. Solid line represents the shelf temperature.
Nonsteady State Modeling of Freeze Drying 29

experimentally. Predicted product temperature profiles


during primary drying are shown for three different
locations in Figure 8a. Since the sublimation interface
moves from the top downwards with the top of the
material drying the earliest, there is about a 3° difference
in temperature between the core and the top of the sample.
There is about a 1° variation in temperature in the radial
direction. The difference between the glass transition
temperature and the product temperature (T  Tg) is an
important factor to be considered throughout the drying
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

process, but the glass transition temperature during


secondary drying is difficult to directly measure. Figure
8b show predicted profiles of (T  Tg) during the
secondary drying stage. Three different locations, the
top, the bottom, and the core of the sample were chosen.
Notice from Figure 8b that the product temperature is well
above the glass transition temperature for several hours
Figure 7. Comparison of residual water contents obtained
from experiment (open and closed circles) and theory (repre-
sented by the solid line) for Process 2. Experiment 1 (open
circles) represents data obtained from a freeze-drying experi-
ment without the use of radiation shields and experiment 2
(filled circles) are data obtained using aluminum foil on the
interior surface of the door as a radiation shield. Error bars
For personal use only.

represent standard deviations.

Residual Moisture vs. Time . Experimental resid-


ual water contents were determined from Karl Fischer
analysis of samples removed at different times during the
freeze-drying process. Figure 7 represents a comparison
of experimental and theoretical water contents with time
obtained for Process 2, (i.e., a validation run that was
operated using a different freeze drying protocol than did
Process 1). Good agreement is observed, although it does
appear that agreement would be better if the initial water
content were lower. That is, our choice of 17.5% water
in the freeze concentrate causes the theoretical results to
be high relative to the experiment at early times. How-
ever, the correspondence is close enough so that the
calculated glass transition temperature, whose accuracy
depends upon accuracy of moisture content, should have
useful accuracy.
It is important, also, to note that heterogeneity in
sublimation rates is more pronounced at lower shelf
temperatures during primary drying, and one can observe
that the variation in data was reduced significantly by the
use of aluminum foil as a radiation shield on the door of
the freeze dryer.
Figure 8. a) Spatial distribution of product temperature (T)
Spatial Distribution of Temperature and Mois- during primary drying, and b) (Tg  T) during secondary drying.
ture . The finite element analysis model can be used to Data are shown for three different positions; bottom center (bold
monitor the time and spatial distribution of several solid lines), top (bold dashed lines), and edge-center (thin dashed
important parameters that may be difficult to determine lines). Process conditions were as in Process 2.
30 M. J. Pikal et al.

during the secondary drying stage. In this particular ples, which were cut into three sections consisting of the
process, a ramp rate of 0.5 C/min was used to advance bottom, top, and side of the sample, show no significant
from primary drying to secondary drying. Thus, it may be difference between the sections (Figure 9b). This
important to use a low ramp rate towards secondary observation is in sharp contrast to both the experimental
drying to avoid partial product collapse, or ‘‘cake data obtained earlier for drying bovine somatotropin
shrinkage’’ that can occur if the product temperature (BST)[11] and the Passage simulation results for BST.[4]
exceeds the Tg. Experimental data for BST[11] showed very large moisture
Another important observation from the spatial variation with position in the cake. There was evidence
distribution of residual moisture during freeze drying of for significantly higher moisture content in the bottom
sucrose samples was the absence of large differences in section, and the sides section typically had the lowest
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

water content between the bottom, top, and sides (i.e., moisture content. The lower moisture in the side (or outer)
outer section, near the vial wall) of the sample (Figure section was attributed to preferential drying near the side,
9a). The theoretical results show only small differences likely caused by the cake shrinking away from the vial
between top and bottom early in secondary drying. Karl wall during primary drying. For BST, Passage simulation
Fischer analysis for water content of these sucrose sam- results[4] showed the following order in residual moisture
For personal use only.

Figure 9. Spatial distribution of residual water contents during secondary drying. (a.) Calculated by Passage; (b.) Experimental
measurements. Symbols key: sides (vertical bars), bottom (solid bars) and top (horizontal bars). The experimental values were by
Karl Fischer analysis of sections of the cake using previously described procedures.[11] The calculated results refer to elements
located as follows: side (two elements from the edge, middle of cake), bottom (three elements up from bottom near center of vial),
and top (two elements down from top near center of vial). Error bars represent standard errors. The asterisk above the 9.8 hour time
point in (a.) indicates only one measurement available for that time point.
Nonsteady State Modeling of Freeze Drying 31

content, top <sides <bottom, with the difference between ev vial top emissivity
top and bottom being about a factor of 20. Cake shrinkage f heterogeneity factor
was not accounted for in the Passage model, so it is h heat transfer coefficient (kW/m2 K)
understandable that the theoretical results did not exactly Jw molar flux of water (mol/m2 s)
reproduce the experimental observations with regard to Jin molar flux of inerts (mol/m2 s)
lowest moisture in the sides section. However, it is kI thermal conductivity of dry layer(kW/mK)
interesting and significant that the qualitative experimen- kII thermal conductivity of frozen layer(kW/mK)
tal results of no large spatial variation in moisture for k1 bulk diffusivity constant C2D0w,inKw/(C2D0w,in
sucrose and large spatial variation in moisture for BST are + KmxP)
predicted by the calculations. k2, k 4 self diffusivity constant KwKin/(C2D0w,in+KmxP)+
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

(C01/mmax)
k3 bulk diffusivity constant C2D0w,in Kin/(C2D0w,in+
CONCLUSIONS KmxP)
kg mass transfer coefficient for desorption (s 1)
The theoretical results of in-process product temper- Kv vial heat transfer coefficient (cal/sec cm2 K)
ature, primary drying time, and residual moisture content Kvb vial bottom heat transfer coefficient (cal/s cm2 K)
variation, both spatial and temporal, are fully consistent Kw Knudsen diffusivity for water vapor C1 (RT/
with the experimental results, suggesting the theoretical Mw)0.5
model should be useful in process development and Kin Knudsen diffusivity for interts C1 (RT/Min)0.5
trouble-shooting applications. Of particular use, is the Kmx mean Knudsen diffusivity for binary gas mixture,
feature of spatial and temporal variation in residual (Kmx = ywKin+ yinKw)
moisture during secondary drying. Such data are ex- l thickness of dry layer, cm
tremely difficult to obtain by direct experiment and are M molecular weight (kg/kg mole)
For personal use only.

quite relevant to questions of collapse during secondary Nin molar flux of inerts (kg/m2sec)
drying as well as perhaps relevant to predictions of Nw molar flux of water (kg/m2sec)
protein stability during drying. That is, since the P total pressure
calculations can give the Tg history throughout the vial Pc chamber pressure (Pa)
during the secondary drying process, stability predictions Pcd condenser pressure (Pa)
can, perhaps, be made for those cases where T  Tg is an Pin partial pressure of inert gas (Pa)
important stability variable. P0 vapor pressure of ice (Torr)
Pv pressure above the dry layer (Torr)
Pw partial pressure of water vapor (Pa)
NOMENCLATURE q heat flux (kW/m2)
R gas law constant
A cross-sectional area of the dry layer, cm2 Rp area normalized product resistance (cm2 hr
C* equilibrium concentration of sorbed water (kg Torr/g)
water/kg solid) R’c chamber resistance
Cp heat capacity (kJ/kg K) RCD’ resistance to flow from condenser chamber to
Cw, s concentration of sorbed water (kg water/kg condenser plates
solid) t time (s)
C01 constant dependent only upon structure of po- T temperature (K)
rous medium and giving relative D’Arcy flow Tint temperature of the sublimation interface (K)
permeability Tp temperature of the product (K)
C1 constant dependent only upon structure of po- Ts shelf Temperature (K)
rous medium and giving relative Knudsen flow Ttop temperature of the top of the dry layer (K)
permeability Tup upper plate temperature (K)
C2 constant dependent only upon structure of po- Xin mole fraction of inerts
rous medium and giving the ratio of bulk dif- yw mole fraction of water vapor
fusivity within the porous medium to the free
gas bulk diffusivity (dimensionless) Greek symbols
Dw, in free gas mutual diffusivity in a binary mixture
of water vapor and inert gas DHs heat of sublimation of ice (kJ/kg)
D0w, in Dw, in P DHv heat of vaporization of bound water (kJ/kg)
32 M. J. Pikal et al.

e void fraction in the dried region as a method of monitoring product temperature


mmx viscosity of the binary mixture of water vapor and during lyophilization. PDA J. Pharm. Sci. Technol.
inert gas in the porous dried layer (kg/m.s) 1997, 5, 7 –16.
r density (kg/m3) 6. Tang, X.; Pikal, M.J. Evaluation of manometric
temperature measurement (MTM) in freeze drying.
Subscripts Part II: measurement of dry layer resistance, (under
review).
I dry region 7. Pikal, M.J.; Roy, M.L.; Shah, S. Mass and heat
II frozen region transfer in vial freeze-drying of pharmaceuticals:
n normal component role of the vial. J. Pharm. Sci. 1984, 73, 1224 –
Pharmaceutical Development and Technology Downloaded from informahealthcare.com by McGill University on 11/30/12

x x-component 1237.
y y-component 8. Pikal, M.J.; Lang, J.E.; Shah, S. Desolvation kinetics
of cefamandole sodium methanolate: the effect of
water vapor. Int. J. Pharm. 1983, 17, 237– 262.
ACKNOWLEDGMENTS 9. Zografi, G.; Kontny, M.J. The interaction of water
with cellulose-and starch derived pharmaceutical
This project was funded by a grant from the National excipients. Pharm. Res. 1986, 187 –194.
Science Foundation – Center for Pharmaceutical Process- 10. Pikal, M.J.; Shah, S.; Roy, M.L.; Putman, R. The
ing Research (NSF –CPPR). secondary drying stage of freeze drying: drying
kinetics as a function of temperature and chamber
pressure. Int. J. Pharm. 1990, 60, 203– 217.
REFERENCES 11. Pikal, M.J.; Shah, S. Intravial distribution of
moisture during the secondary drying stage of
For personal use only.

1. Pikal, M.J. Freeze drying. In Encyclopedia of freeze drying. PDA J. Pharm. Sci. Technol. 1997,
Pharmaceutical Technology, 6; Swarbick, J., Boylan, 51, 17 –24.
J., Eds.; Marcel Dekker, 2001. See also, the Dekker 12. Pikal, M.J.; Shah, S.; Senior, D.; Lang, J.E. Physical
web site, www.Dekker.com. The DOI number is, chemistry of freeze-drying: measurement of subli-
10.1081/E-EPT-100001712, pages 1299 –1326. mation rates for frozen aqueous solutions by a micro
2. Pikal, M.J. Use of laboratory data in freeze drying balance technique. J. Pharm. Sci. 1983, 72, 635 –
process design: heat and mass transfer coefficients 650.
and the computer simulation of freeze-drying. J. 13. Saleki-Gerhardt, A.; Zografi, G. Non-isothermal and
Parenter. Drug Assoc. 1985, 39 (No. 3), 115 – 138. isothermal crystallization of sucrose from the
3. Millman, M.J.; Liapis, A.I.; Marchello, J.M. An amorphous state. Pharm. Res. 1994, 11, 1166 –1173.
analysis of the lyophilization process using a 14. Hatley, R.H.M.; van den Berg, C.; Franks, F. The
sorption-sublimation model and various operational unfrozen water content of maximally freeze con-
policies. AICHE J. 1985, 3, 1594– 1604. centrated carbohydrate solutions: validity of the
4. Mascarenhas, W.J.; Akay, H.U.; Pikal, M.J. A methods used for its determinaton. Cryo-Lett. 1991,
computational model for finite element analysis of 12, 113 –124.
the freeze-drying process. Comput. Methods Appl. 15. Hatley, R.H.M.; Mant, A. Determination of the
Mech. Eng. 1997, 148, 105– 124. unfrozen water content of maximally freeze-con-
5. Milton, N.; Pikal, M.J.; Roy, M.L.; Nail, S.L. centrated carbohydrate solutions. Int. J. Biol.
Evaluation of manometric temperature measurement Macromol. 1993, 15, 227 –232.

You might also like