You are on page 1of 13

Chemical Engineering Science 174 (2017) 374–386

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Systematic model identification and optimization-based active


polymorphic control of crystallization processes
E. Simone a,b, B. Szilagyi b,c, Z.K. Nagy b,c,⇑
a
School of Food Science and Nutrition, University of Leeds, Leeds LS2 9JT, UK
b
Department of Chemical Engineering, Loughborough University, Loughborough LE11 3TU, UK
c
Davidson School of Chemical Engineering, Purdue University, West Lafayette, IN 47907-2100, USA

h i g h l i g h t s

 A model-based active polymorphic control for batch crystallization is proposed.


 A high resolution finite volume method is used to solve population balance equations.
 Optimization was performed to estimate the best temperature profile.
 The dissolution cycle is beneficial to obtain larger crystals of the stable polymorph.

a r t i c l e i n f o a b s t r a c t

Article history: Polymorphism is an important issue in industrial crystallization, since polymorphs of the same com-
Received 31 May 2017 pound can present very different properties, such as solubility, melting point or density, influencing con-
Received in revised form 17 August 2017 siderably the manufacturability and bioavailability of the final product.
Accepted 16 September 2017
This work proposes a model-based active polymorphic control strategy that allows obtaining large
Available online 19 September 2017
crystals of the stable polymorph at the end of a batch crystallization process, even in the case of erro-
neous seeding or in situ nucleation of a mixture of both the stable and metastable forms. A novel system-
Keywords:
atic experimental design was applied to estimate the kinetic parameters of dissolution, growth and
Polymorphic control
Population balance equations
secondary nucleation of the stable and metastable polymorphs of the model compound (ortho-
Batch crystallization optimization aminobenzoic acid, OABA). Such experimental approach allows the determination of the studied kinetics
without any correlation between parameters during the estimation, and without the need of off-line
measurements of the crystal size distribution during the experiments.
The estimated kinetic parameters were used to build a population balance model for the calculation of
the optimal temperature profile needed, during a batch cooling crystallization process, for the (i) elimi-
nation of the metastable form crystals nucleated in situ or erroneously seeded and the (ii) maximisation
of the size of the crystals of the stable polymorph obtained at the end of the batch process.
Ó 2017 Elsevier Ltd. All rights reserved.

1. Introduction conditions, temperature, pH and the use of additives can determine


the polymorphic outcome of a cooling crystallization, while PAT
Polymorphs of the same compound can have different physical tools can be used to check the purity of the final product and con-
characteristics such as solubility, stability, melting point and, most trol its growth. ATR-FTIR, ATR-UV/Vis, in situ Raman and FBRM
importantly, bioavailability. For this reason both discovering new have been frequently used to control the growth of both stable
polymorphs and designing new control strategies to tailor the and metastable polymorphs through different control approaches.
polymorphic purity of the final product crystallized in industrial Recently a feedback control technique, the active polymorphic
processes is very important. The choice of solvent, supersaturation feedback control (APFC), was developed to select and grow the
desired polymorphic form of the crystallized compound (Simone
et al., 2014). In this strategy both Raman and ATR-UV/Vis spec-
⇑ Corresponding author at: Department of Chemical Engineering, Loughborough troscopy are used: the Raman probe can detect the nucleation or
University, Loughborough LE11 3TU, UK and Davidson School of Chemical seeding of a polymorphic mixture and it eliminates the metastable
Engineering, Purdue University, West Lafayette, IN 47907-2100, USA. form by triggering a controlled dissolution cycle.
E-mail address: zknagy@purdue.edu (Z.K. Nagy).

https://doi.org/10.1016/j.ces.2017.09.034
0009-2509/Ó 2017 Elsevier Ltd. All rights reserved.
E. Simone et al. / Chemical Engineering Science 174 (2017) 374–386 375

ATR-UV/Vis is instead used to control the crystallization condi- centration, crystal size distribution and polymorphic ratio; a finite
tions and allow only the growth of the stable form using supersat- elements method was used to solve the population balance equa-
uration control. Other control approaches proposed in the tion, PBE (using the software FEMLAB). Many different solution
literature either use the Raman system only to detect the forma- techniques were used to solve the PBE for the polymorphic trans-
tion of the unwanted polymorph and restart the crystallization formation of L-glutamic acid: moving pivot technique (Cornel et al.,
with a different cooling rate, or use only supersaturation control 2009), finite volume method in gPROMS (Ono et al., 2004) and the
in conjunction with a suitable seeding procedure to drive the sys- method of moments (Hermanto et al., 2007, 2009, 2011;
tem in the phase diagram in order to obtain the desired polymor- Sheikholeslamzadeh and Rohani, 2013). Despite working with the
phic form. A summary of recent research works on polymorphic same system the authors of the mentioned studies used different
control is shown in Table 1. types of equations to express the kinetics of the phenomena
The APFC strategy is a model-free approach, which was evalu- involved in the polymorphic transformation of L-glutamic acid.
ated for the cooling crystallization of ortho-aminobenzoic acid, All the authors found a good agreement between simulated and
and led to pure polymorphic forms in the case of unseeded crystal- experimental data, even when semi-empirical, simplified functions
lization processes where nucleation of polymorph mixtures were used.
occurred, or for seeded crystallization with contaminated seed Ono et al. included in the model only dissolution of the meta-
crystals containing an unwanted polymorph impurity (Simone stable form (Sherwood correlation), size-dependent growth and
et al., 2014). During the experiments performed, a partial dissolu- secondary nucleation of the stable form (semi-empirical function
tion of the desired form together with the elimination of the unde- of supersaturation and mass of crystals of the stable form in
sired form was observed. However, it is not clear whether such slurry). More phenomena were included in the models described
partial dissolution favours the attainment of larger crystals of the by Hermanto et al. (2011), Cornel et al. (2009) and
stable form at the end of the batch or not. A model based approach Sheikholeslamzadeh and Rohani (2013): primary nucleation and
can help understanding if the initial dissolution cycle improve or dissolution of the metastable form, secondary nucleation and
worsen the final size distribution of the crystals of the stable form growth of the stable form. Besides, less empirical correlations were
and how the temperature profile could be optimized in order to used in such studies compared to the model described by Ono et al.
maximize such distribution. The aim of this work is to develop a (2004). In fact, the growth kinetics of the stable and metastable
model-based active polymorphic control by determining the polymorphs of L-glutamic acid were found to be integration con-
kinetic parameters of the growth and polymorphic transformation trolled and of the birth-and-spread type, with the exception of
of ortho-aminobenzoic acid through properly designed experi- the studies performed by Hermanto and co-workers, where a
ments, and then by simulating and optimizing the batch crystal- power-law function was used to express the growth rate of the
lization process in order to control both size and polymorphic metastable form. A Sherwood correlation was used to estimate
purity of the final crystals. the dissolution of the metastable form in all the referenced works.
Parameter estimation and modelling of polymorphic transfor- The functions used to express the nucleation rates for both the
mation has rarely been performed because of the complexity of stable and metastable forms of L-glutamic acid were different in
the phenomenon, which involves two steps: dissolution of the the mentioned studies: Cornel et al. (2009) and
metastable form and nucleation and growth of the stable one Sheikholeslamzadeh and Rohani (2013) used primary nucleation
(Cardew and Davey, 1985). exponential functions to describe the primary nucleation of the
A first example of population balance applied to a polymorphic metastable form, while Hermanto et al. (2011) employed a simpler
transformation was the study on the conversion of citric acid from equation as a function of supersaturation and the third moment
the anhydrate to its monohydrate form (Fevotte et al., 2007). calculated for the metastable form. The kinetic of secondary nucle-
Seeded isothermal experiments were conducted to estimate the ation of the stable form of L-glutamic acid was expressed with a
kinetic parameters for the growth and nucleation of the monohy- semi-empirical function only of the mass of metastable crystals
drate form as well as for the dissolution of the anhydrate form. by Cornel et al. (2009) and of the mass of both the stable and meta-
Power-law relationships were used to express dissolution and stable crystals by Hermanto et al. (2011) Sheikholeslamzadeh and
growth as a function of supersaturation, while secondary nucle- Rohani (2013) instead employed a two-terms expression to esti-
ation was expressed as a function of supersaturation as well as of mate both the heterogeneous (exponential nucleation function)
the concentration of crystals of the stable form present in suspen- and surface secondary nucleation (as a function of the second
sion. Raman and image analysis were used to measure solute con- moment of the metastable crystals) of stable L-glutamic acid.

Table 1
Control strategies used by researchers to control the crystallization of polymorphic compounds (Simone et al., 2014).

Control approach Reference


Seeding the desired form between the solubility curve and the metastable limit line in order to avoid the nucleation Threlfall (2000) and Beckmann (2000)
of the other form and keep cooling until the supersaturation is consumed
Finding the correct amount of seed above which secondary nucleation of the metastable form is suppressed and Doki et al. (2003)
solution-mediated transformation is avoided in a cooling crystallization
Seeding during a cooling crystallization and using focused beam reflectance measurement (FBRM) in combination Doki et al. (2004)
with ATR-FTIR to check the total counts and the supersaturation in order to reach the desired size of the crystals
and eliminate the fines via dissolution
Temperature control and concentration control for the conversion of the metastable form of a polymorph to its Hermanto et al. (2007), Kee et al. (2009a, 2009b),
stable form (simulation and experimental work) Hermanto et al. (2009)
Seeding and growth of the metastable form during a cooling crystallization performing supersaturation control Kee et al. (2009), Chew et al. (2007)
Combination of anti-solvent and cooling crystallization to obtain the desired polymorph Minamisono and Takiyama (2013)
Feedback control of reactive crystallization in a semi-batch precipitation using MID-IR or Raman, ATR-FTIR and a Qu et al. (2009) and Alatalo et al. (2010)
pH-meter
Control of Polymorphism in Continuous Crystallization via Mixed Suspension Mixed Product Removal Systems Lai et al. (2015)
Cascade Design: estimation of the optimal operating conditions to crystallize one specific polymorph
376 E. Simone et al. / Chemical Engineering Science 174 (2017) 374–386

The solution-mediated transformation of DL-methionine poly- ied in this work is the one from the metastable form II to the stable
morphs was modelled by Wantha and Flood (2013) using the form I in a solution of 90% water and 10% IPA, below 50 °C (see
method of moments to solve the PBEs. In this work semi- Fig. 1a and b). OABA (>98% Form I, Sigma-Aldrich), isopropyl alco-
empirical functions of the supersaturation were used to express hol (99.97%, Fisher Scientific) and ultrapure water obtained via a
the growth kinetics of both forms and the dissolution kinetics of Millipore ultra-pure water system, were used. Experiments were
the metastable polymorph; a primary nucleation exponential func- performed in a 400 ml jacketed vessel; a PT-100 temperature
tion was used to estimate nucleation of the stable polymorph. probe connected to a Huber Ministat 230 thermoregulator was
Scholl et al. (2006) solved the PBEs using the commercial soft- used to control the temperature. A RXN1 Raman analyser with
ware PARCIVAL for the parameter estimation of the kinetics of immersion probe and 785 nm laser (Kaiser with iC Raman 4.1 soft-
transformation of L-glutamic acid (Scholl et al. (2006)). The model ware) was used to detect different polymorphs in suspension,
used included the kinetics of heterogeneous nucleation of the while a MSC621 Carl Zeiss UV/Vis (in-house LabView software)
metastable form (exponential primary nucleation type function), with Hellma ATR (type 661.822-UV) probe was used to determine
size-independent growth rates of both the stable and metastable the solute concentration. A Malvern Mastersizer 2000 was used to
forms (integration controlled and birth and spread type of func- determine the crystal size distribution at the beginning and during
tions), dissolution of the metastable form (Sherwood correlation) the experiments.
and heterogeneous and surface nucleation of the stable form. A The mean and the standard deviation of the crystal size distri-
similar model was used to describe the polymorphic transforma- butions measured using the Mastersizer, were used to calculate a
tion of Buspirone hydrochloride from the metastable form II to Gaussian curve that approximate the experimental data. This
stable form I (Trifkovic et al., 2012). Such model was solved using approximation was necessary to avoid the overestimation of fine
the methods of moments. particles in the measured samples, due to the non-spherical shape
More recently, the methods of characteristics was used to of the OABA crystals of both polymorphic forms. In fact, the volume
describe the behaviour of the a and b forms of para- and number distributions of needles and flat crystals measured by
aminobenzoic acid in a two stages MSMPR reactor (Lai et al., laser diffraction can show a large number of fine crystals (or even a
2015). The authors included in the model the growth of both stable bimodal shape) simply because of the orientation of elongated par-
and metastable forms (size-independent and surface integration ticles on their shortest side during the measurement (Su et al.,
controlled) and their secondary nucleation (semi-empirical equa- 2017).
tion as a function of the mass of crystals in suspension). A previously developed calibration model (Simone et al., 2014)
As explained in the previous paragraph, population balance was used to determine solute concentration from ATR-UV/Vis
models in the literature can include or not primary nucleation of while specific Raman peaks for form I and II were tracked during
both the stable and the metastable form but all of them include the experiments to estimate the rate of transformation and check
secondary nucleation of the stable form, expressed with semi- the composition of the slurry during the experiments performed
empirical functions, primary heterogeneous nucleation exponen- with the metastable polymorph. Furthermore, the initial seeds
tials or with two-terms functions including both heterogeneous were analysed with a Raman microscope (DXR Raman, Ther-
primary nucleation and surface secondary nucleation. In fact, sec- mofisher) in order to check their purity.
ondary nucleation of the stable form and dissolution of the meta- The solubility curves for both form I and II between 10 and
stable polymorph are the key mechanisms happening during a 40 °C were estimated using an the ATR-UV/Vis probe (interpolating
polymorphic transformation (Cardew and Davey, 1985).
Only one theoretical study considered the presence of sec-
ondary nucleation of the metastable form and analysed its effect (a)
on the transformation time and the concentration profile (Kobari 100 µm
et al., 2014). However, the presence of secondary nucleation of
the metastable form can be neglected if, during the crystallization
process, the solute concentration is very close, or below the solu-
bility of the metastable form. Thus, the supersaturation is too
low to allow secondary nucleation.
The parameters necessary to define and model the active poly-
morphic control of ortho-aminobenzoic acid are: (i) dissolution
kinetics for both forms, (ii) growth kinetics for both forms, (iii) sec-
ondary nucleation of the stable form (during transformation), and
(iv) primary nucleation of the stable form (during transformation
and after seeding far from the solubility curve). The estimated
parameters were then validated and applied to an optimization
problem in order to design batch cooling crystallization processes
(b)
that allow the growth of large crystals of the stable polymorph 100 µm
even in case of erroneous seeding or in situ nucleation of a mixture
of the stable and metastable forms. In conclusion, the model-based
active polymorphic control (mbAPC) proposed in this work repre-
sents a useful approach for the correct design of batch crystalliza-
tion processes for polymorphic systems.

2. Materials and methods

The model compound used for the experiments is ortho-


aminobenzoic acid (OABA), which has three known polymorphic Fig. 1. Micrographs of OABA crystals: (a) prismatic form I, and (b) needle-like form
forms (Jiang et al., 2010b, 2010a, 2008). The transformation stud- II.
E. Simone et al. / Chemical Engineering Science 174 (2017) 374–386 377

data from a slow heating profile). Despite the system being enan- DI and DII are the dissolution rates of the two forms and GI and
tiontropic (Jiang et al., 2010b, 2010a, 2008), in the used tempera- GII the growth rates of form I and II defined as:
ture interval the two OABA polymorphs can be considered hlmi  
EdI
monotropically related. The formulas used for the solubility DI ¼ kdI ð1  SI ÞdI exp ð8Þ
s RT
(Simone et al., 2014) are:
hlmi
C eq;I ¼ 1:267  105 T 2  2:283  104 T þ 4:105  103 ð1Þ DII
EII
¼ kdII ð1  SII ÞdII expð
Þð1 þ LII Þ ð9Þ
s RT
5 2 4 3
C eq;II ¼ 1:299  10 T  2:082  10 T þ 4:808  10 ð2Þ hlmi  
EgI
GI ¼ kgI ðSI  1ÞgI exp ð10Þ
with the temperature T expressed in °C and the solubility C eq;I and s RT
C eq;II calculated in g/g solvent. The solubility of form II and form I
hlmi EgII
have been interpolated with polynomial functions and not with a
GII ¼ kgII ðSII  1ÞgII expð Þð1 þ LII Þ ð11Þ
Van’t Hoff type equation to keep consistency with our previous s RT
experimental APFC study (Simone et al., 2014). In such paper poly-
with the supersaturation defined as SI;II ¼ C satI;II
C
.
nomials were used as this is the only type of equation that can be
currently input in the in-house software (CryPRINS) to perform Two types of nucleation of the stable form were estimated: (i)
supersaturation control during batch crystallization experiments. primary nucleation after seeding of a mixture of polymorphs if
cI > csatII ; (ii) secondary nucleation during polymorphic transfor-
mation (if cI > csatI ). The different types of nucleation can be
2.1. Population balance model and solution
described by the equation:
  " #    
For the description of a particle population, let us introduce the # b Eb Es
s
monovariate number density function f ðL; tÞdL, which expresses BI ¼ k b 2
exp  þ k s ð1  SI Þ exp 
m3 sec ðlogSI Þ RT RT
the number of crystals within the L; L þ dL crystal size domain (L
expressed in mm) in a unit volume of suspension in the t time ð12Þ
moment (expressed in seconds). Then, population balance equa- A high resolution finite volume method (HR-FVM) was used to solve
tions can be used to predict and simulate polymorphic transforma- the model-equations (Gunawan et al., 2004). The basic idea of
tions considering one equation for each polymorph. Three main HR-FVM is the discretization of the continuous population density
mechanisms must be considered during a transformation: nucle- m
function; denoting with h the size and k the time interval, f l is
ation and growth of the more stable form and the dissolution of
the approximate (discrete) population density function defined as:
the less stable polymorph. In the mbAPC also dissolution of form
Z lh
I must be considered and estimated. Indicating with the index II m 1
fl  f ðL; mkÞdL ð13Þ
the parameters of the metastable form of OABA, and with I the h ðl1Þh
ones of the stable one, the PBE for the studied system, using the
simplified f ðL; tÞ ? f notations, are: where m and l are integers such that m P 0 and N P l P 1 and N
stands for the mesh size (i.e. the number of discretization points).
@f II @ðDII f II Þ Then, the population balance Eq. (6) reduces to a system of alge-
¼ ð3Þ
@t @L braic equations:
for dissolution of form II, mþ1 m k m m 
fl ¼ f l  Gl f l  Gl1 f l1
@f II @ðGII f II Þ   h    
þ ¼0 ð4Þ kGl kGl  m m kGl1 kGl1  m m 
@t @L  1 f lþ1  f l /l  1 f l  f l1 /l1
2h h 2h h
for growth of form II, k
þ b B ð14Þ
@f I @ðDI f I Þ h
¼ ð5Þ
@t @L In Eq. (14) b is a binary existence variable with values {0,1} which
for dissolution of form I, and controls the existence of nucleation. In this PBE formulation, b ¼ 1
if l ¼ 1 (nucleus size) and is 0 otherwise. It is worth noticing that the
@f I @ðGI f I Þ same Eq. (14) equation is used for growth and dissolution stages,
þ ¼ BI dðL  L0 Þ ð6Þ
@t @L treating the dissolution as negative growth and keeping in mind
for growth and nucleation of form I: where f II and f I are the average that the nucleation rate is 0 if the solution is undersaturated.
number density functions of the metastable and stable form of /l ¼ f ðhl Þ is the flux limiter function and hl is the ratio of consecu-
OABA and dðL  L0 Þ is the Dirac delta function (d ¼ 1 if L ¼ L0 ¼ 0 tive gradients:
R þ1
and d ¼ 0 if L–L0 with 1 dðxÞdx ¼ 1). m
f l  f l1
m

In order to close the system of equations that characterize the hl ¼ m m ð15Þ


f lþ1  f l
presented model the liquid phase mass balance is required (tem-
perature is the controlled variable, therefore, the energy balance The Van Leer flux limiter of Eq. (16) has been successfully applied in
is not necessary): the solution of population balance equations thus is adopted in this
Z 1 Z 1  study too (Gunawan et al., 2004).
dc
¼ 3kv qC GI L2 f I dL þ GII L2 f II dL ð7Þ jhl j þ hl
dt 0 0 /ðhl Þ ¼ ð16Þ
1 þ jhl j
where kv is the volume shape factor and qC stands for the crystal
density. The mass balance Eq. (7) refers to the case when both pop- Note that the numerical apparatus Eqs. (13)–(16) applies for both
m m m m
ulations are growing and it considers 0 nucleus size. The equation populations: f l ! f I;l (with GI;l ; DI;l and BI ) and f l ! f II;l (with
remains valid for dissolution by applying simply DI /DII dissolution GII;l ; DII;l and BII ¼ 0). The resulted algebraic equation systems are
rates. solved simultaneously.
378 E. Simone et al. / Chemical Engineering Science 174 (2017) 374–386

The time step is recalculated in all iterations to satisfy the through desupersaturation experiments with low seeds loading at
Courant-Friedrichs-Lewy (CFL) criterion and the numerical system high supersaturation.
is stable if CFL  1. This approach has two main advantages: (i) a correlation
between the estimated parameters is avoided because only one
k phenomenon occurs in each set of experiments; (ii) only concen-
CFL ¼ maxfGI;l ; GII;l ; DI;l ; DII;l g ð17Þ
h tration data and the initial crystal size distribution (CSD) are
needed for the estimation, sampling is not necessary to estimate
Practically the CFL is fixed and k is expressed from Eq. (17). Finally,
the CSD during the crystallization process (Besenhard et al.,
the solute mass balance takes the form:
2015). Therefore, in this work the concentration profile is the only
X
N measured output used for the fitting procedure.
2 m m
cmþ1 ¼ cm  3kv qC h L2 ½f I;l GI;l þ f II;l GII;l  ð18Þ The present systematic approach allows the determination of
l¼1 the necessary kinetics parameters using only limited CSD data,
which can often be unreliable. In fact, such data cannot be easily
Similarly to the Eq. (14), the mass balance Eq. (18) is applicable for obtained online by standard process analytical technologies (e.g.
dissolution as well, involving the dissolution rate for undersatu- FBRM) and is often estimated via off-line techniques such as opti-
rated conditions. An extended version of the CrySiV function cal or scanning electron microscopy or laser diffraction. The need
(Szilagyi and Nagy, 2016) was used to efficiently solve the equation of sampling and the off-line nature of the traditional CSD measure-
system and a combination of Evolution Strategy with Covariance ment techniques lower the accuracy and reliability of the collected
Matrix Adaptation (ES-CMA) global optimization algorithm data. A detailed list of the experiments performed and their condi-
(Hansen et al., 2003) and Matlab’s nlinfit function (Levenberg- tions is shown in Table 2. Some of the experiments reported in
Marquardt algorithm) was employed to estimate the parameters Table 2 could be conducted consecutively in the same solution:
and the confidence intervals. growth of the metastable form can be estimated by a seeded exper-
iment that can then be used to estimate secondary nucleation of
2.2. Systematic experimental design for the model identification the stable form by just letting the metastable form transform. In
these cases, sampling at the beginning of the transformation is nec-
Experiments were planned carefully in order to simplify the essary to estimate the initial crystal size distribution to use in the
estimation of the kinetic parameters: the different phenomena parameter estimation. In particular, the kinetic parameters of
were isolated as shown in Fig. 2. Growth and dissolution for both growth of form II (kgII , EgII and g II ) and secondary nucleation of form
forms were estimated through seeded saturation or desupersatura- I (s,ks and Es ) were estimated from the same isothermal experi-
tion experiments. The secondary nucleation of the stable form was ments, number 11–14, each one conducted at a different tempera-
estimated through isothermal transformation experiments and ture. All data points before the start of the nucleation of the stable
using the dissolution and growth kinetics already estimated. form were used to estimate the growth of the metastable form II,
Finally, secondary nucleation of form I after seeding was evaluated while data collected after the appearance of the stable form were

Fig. 2. Schematic of the experimental design used for the parameters estimation. Growth and dissolution for both forms were estimated through seeded saturation or
desupersaturation experiments. The secondary nucleation of the stable form was estimated through isothermal transformation experiments and using the dissolution and
growth kinetics already estimated. Finally, secondary nucleation of form I after seeding was evaluated through desupersaturation experiments with low seeds loading at high
supersaturation.
E. Simone et al. / Chemical Engineering Science 174 (2017) 374–386 379

Table 2
Description of the conditions used in the different experiments to determine the kinetics parameters of OABA.

Parameter estimated Experimental conditions Experiment number



kdI , EdI and dI Four isothermal experiments (10; 15; 25 and 35 C). Seeds were added to the solvent in the 1–4
amount necessary to have a saturated solution after the complete dissolution
kdII , EdII and dII Four isothermal experiments (10; 15; 25 and 35  C). Seeds were added to the solvent in the 5–8
amount necessary to have a saturated solution. Raman spectroscopy was used to check the
absence of polymorphic transformation during dissolution
kgI , EgI and g I Seeds (10% of the total solute) were added to a saturated solution at 40  C. A slow linear 9–10
cooling was then applied to avoid nucleation. Two experiments at different cooling rates were
performed (0:1 and 0:05  C=min)

kgII , EgII and g II Four isothermal experiments (10; 15; 25 and 35 C). Seeds of the metastable form (10% of the 11–14 (Data points until the nucleation of
total solute) were added to supersaturated solutions (5  C of supersaturation) the stable form started)
s,ks and Es Isothermal seeded polymorphic transformation experiments at four different temperatures 11–14 (Data points after the nucleation of
(10; 15; 25 and 35  C) the stable form started)
b, kb and Eb Seeding of a mixture of polymorphs at saturated condition for form II (about 40  C) and cooling 15–17
at 1  C=min (three experiments, amount of seeds of 10% of the total solute)

used for the estimation of the kinetics of secondary nucleation of Table 3


such polymorph. Kinetic parameters estimated from the performed experiments.
The results of one of the combined experiments are shown in Parameter Value Upper limit of the 95% Lower limit of the 95%
Fig. 3. Growth of metastable form and secondary nucleation of interval of confidence interval of confidence
the stable form at 10 °C are measured. The first 4000 s of the exper- kdI ½lsm 3.45  108 1.38  108 5.6  108
iment were used, together with the other three isothermal growth kJ
EdI ½mol  48,091 49,587 46,595
experiments, to estimate growth of form II while the remaining dI [–] 0.65 0.69 0.61
time was used to estimate secondary nucleation of form I. Another kdII ½lsm 204 101 307
important piece of information shown in Fig. 3 is that the system kJ
EdII ½mol  24,276 25,545 23,001
can be considered neither growth nor dissolution controlled as in dII [–] 0.87 0.91 0.83
the case of previously studied compounds. kJ
kgI ½mol 6.20  1013 1.39  1014 8.84  1012
kJ
EgI ½mol  90,628 93,378 87,877
g I [–] 0.82 0.88 0.75
3. Results and discussion kJ 0.00232 0.00449 0.00014
kgII ½mol
kJ
EgII ½mol  9466 11,388 7445
3.1. Parameters estimation and validation
g II [–] 0.41 0.47 0.36
ks ½m3#sec 7.92  104 4.84  106 1.34  102
The kinetic parameters estimated from all the experiments are kJ 42,323 24,254 60,300
Es ½mol
shown in Table 3 while Figs. 4–7 show the simulated and experi- s [–] 3.95 4.86 3.04
mental data for dissolution and growth of form I and II at different kb ½m3#sec 6.02  1042 1.85  1070 2.69  1014
conditions, as well as the two types of nucleation. In order to val- kJ
Eb ½mol 179,480 348,390 164,390
idate the parameters estimated a leave-one-out cross-validation b [–] 0.1a 0.0129 0.054
was performed using all the available experiments. a
Lower bound of searching domain.
Fig. 4 shows, for each dissolution experiment performed with
form I, the experimental data for concentration together with the
best fit (obtained using all the available experiments) and the cal-
culated concentration for the cross validation (calculated using the for the growth of form I and, finally, in Fig. 7 for the growth of form
parameters obtained by leaving that experiment out of the II. The simulated concentrations for the dissolution of both forms
estimation). seem to follow well the experimental values and the 95% confi-
The same experimental, fitted and cross-validation concentra- dence interval for all the estimated values are narrow (as shown
tions are shown in Fig. 5 for the dissolution of form II, in Fig. 6 in the third and fourth column of Table 3).

Fig. 3. Results for growth of form II and transformation experiment at 10 °C.


380 E. Simone et al. / Chemical Engineering Science 174 (2017) 374–386

x 10-3 x 10-3
3.5 3.5
Concentration (g/g solvent)
3

Concentration (g/g solvent)


3

2.5 2.5

2 2

1.5 1.5

1 1
Simulated Simulated
0.5 Cross-validation 0.5 Cross-validation
Experimental Experimental
0 0
0 200 400 600 800 0 100 200 300 400
Time (s) Time (s)
(a) (b)

x 10-3
7 0.012
Concentration (g/g solvent)

Concentration (g/g solvent)


6
0.01

5
0.008
4
0.006
3

0.004
2
Simulated Simulated
1 Cross-validation 0.002
Cross validation
Experimental Experimental
0 0
0 100 200 300 400 0 100 200 300 400
Time (s) Time (s)
(c) (d)

Fig. 4. Dissolution experiments for form I: (a) Experimental, fitted and cross-validation concentration profile for the dissolution of form I at 10 °C; (b) Experimental, fitted and
cross-validation concentration profile for the dissolution of form I at 15 °C; (c) Experimental, fitted and cross-validation concentration profile for the dissolution of form I at
25 °C; (d) Experimental, fitted and cross-validation concentration profile for the dissolution of form I at 35 °C.

The cross validation still follows reasonably well the experi- 2016). Fig. 8 shows the simulated primary and secondary nucle-
mental data with the exception of the dissolution of form I at 10 ation for form I.
and 35 °C (see Fig. 4a and d). In these two cases, the trends of The difference between simulated and experimental concentra-
the cross-validation concentration slightly differ from the experi- tion is higher compared to the experiments with growth and disso-
mental values. There are two possible reasons for this behaviour; lution of both forms and the 95% confidence intervals are also
the first is simply the approximation of the crystal size distribution broader (as shown in Table 3). This is due to the difficulty in esti-
to a Gaussian function that might generate an error in the evalua- mating the kinetics parameters for a stochastic process such as
tion of the initial crystal size distribution for these two specific nucleation and also because of the limited number of experimental
experiments. The second reason might be a reduced sensitivity to data available.
the temperature change during the estimation performed leaving One of the APFC experiments performed (Simone et al., 2014)
out the highest and lowest temperatures, due to the smaller tem- was used to validate the set of parameters estimated. Seeding
perature range in which the parameters are calculated (20 and and dissolution cycle were simulated using the initial conditions
15 °C instead of 25 °C for the cross-validation of the experiments shown in Table 4.
at 15 and 25 °C). Fig. 6a also shows a deviation of the calculated The mean l and the sigma r of the crystal size distribution for
cross-validation concentration from the experimental values, espe- the validation experiment were estimated as follows:
cially at the beginning of the profile. In this case, the difference is
most certainly due to an experimental error on the determination Lmax þ Lmin
l¼ ð19Þ
of the initial crystal size distribution or to the approximation of the 2
distribution itself with a Gaussian function, as the deviation is
located close to the initial period. Lmax  Lmin
r¼ ð20Þ
However, the presence of an estimation error due to an impre- 2
cise initial crystal size distribution is not surprising considering where Lmax and Lmin are the maximum and minimum sizes of the
how difficult is to obtain good and reliable measurement of the sieves used to separate the seeds (the Malvern Mastersizer was
crystal size distribution with standard techniques such as the Mal- not used for this sample).
vern Mastersizer or 2D image analysis (Su et al., 2017; Ma et al., The results of the validation experiment are shown in Figs. 9
and 10.
E. Simone et al. / Chemical Engineering Science 174 (2017) 374–386 381

-3
x 10 x 10
-3
4 5

Concentration (g/g solvent) 3.5 4.5

Concentration (g/g solvent)


4
3
3.5
2.5
3
2 2.5

1.5 2

1.5
1
Simulated 1 Simulated
0.5 Cross-validation Cross-validation
Experimental 0.5 Experimental
0 0
0 100 200 300 400 500 0 100 200 300 400 500
Time (s) Time (s)
(a) (b)
-3
x 10
8 0.012

7
0.01
Concentration (g/g solvent)

6 Concentration (g/g solvent)


0.008
5

4 0.006

3
0.004
2
Simulated
0.002 Simulated
1 Cross-validation
Cross-validation
Experimental Experimental
0 0
0 100 200 300 400 500 0 100 200 300 400 500 600
Time (s) Time (s)
(c) (d)

Fig. 5. Dissolution experiments for form II: (a) Experimental, fitted and cross-validation concentration profile for the dissolution of form II at 10 °C; (b) Experimental, fitted
and cross-validation concentration profile for the dissolution of form II at 15 °C; (c) Experimental, fitted and cross-validation concentration profile for the dissolution of form
II at 25 °C; (d) Experimental, fitted and cross-validation concentration profile for the dissolution of form II at 35 °C.

Fig. 6. Growth experiments for form I: (a) Experimental, fitted and cross-validation concentration profile for the growth of form I in a desupersaturation experiment
performed with 0.1 °C/min cooling rate; (b) Experimental, fitted and cross-validation concentration profile for the growth of form I in a desupersaturation experiment
performed with 0.05 °C/min cooling rate.

Despite the difficulty in obtaining a good estimation for the of the first moment of form II is similar to the corresponding
nucleation kinetics, the simulated concentration for the validation Raman signal. A discrepancy is present in the cooling section and
experiments correctly follows the experimental data and the trend it is probably due to the uncertainty in the estimation of primary
382 E. Simone et al. / Chemical Engineering Science 174 (2017) 374–386

x 10-3 x 10-3
4.7 5.4
Cross-validation Simulated
Simulated 5.3
Concentration (g/g solvent)
4.6 Cross-validation

Concentration (g/g solvent)


Experimental Experimental
5.2
4.5
5.1

4.4 5

4.3 4.9

4.8
4.2
4.7

4.1 4.6
0 500 1000 1500 2000 2500 3000 0 500 1000 1500 2000 2500
Time (s) Time (s)
(a) (b)
-3
x 10
8.3 0.0146
Simulated
Simulated
Cross-validation
Concentration (g/g solvent)

Concentration (g/g solvent)


8.2 0.0144 Cross-validation
Experimental
Experimental

8.1 0.0142

8 0.014

7.9 0.0138

7.8 0.0136

7.7 0.0134
0 500 1000 1500 2000 2500 0 1000 2000 3000 4000
Time (s) Time (s)
(c) (d)

Fig. 7. Growth experiments for form II: (a) Experimental, fitted and cross-validation concentration profile for the growth of form II at 10 °C; (b) Experimental, fitted and cross-
validation concentration profile for the growth of form II at 15 °C; (c) Experimental, fitted and cross-validation concentration profile for the growth of form II at 25 °C; (d)
Experimental, fitted and cross-validation concentration profile for the growth of form II at 35 °C.

nucleation (the simulated first moment of form I is higher than the The batch time was discretized in fifty time intervals of equal dura-
actual one, and therefore, the growth is overestimated). tion and the temperature profile optimization was performed by
Fig. 10 shows the third moments of form I and II compared to applying the ES-CMA global optimization algorithm. The results
the Raman signals of the specific peaks of those forms. The Raman were further refined by performing a second optimization using
intensity is proportional to the amount of solid in suspension and, the global optimizer’s crude optimum as starting point, applying
therefore, can be directly compared with the third moments of the the Matlab fmincon function (SQP algorithm). The initial tempera-
two polymorphs. In fact, the third moment of the distribution (l3 Þ ture of seeding was fixed at around 37 °C. The problem is formu-
is proportional to the specific volume of crystal population (V c ) and lated as follows:
is one of the infinite moments of CSD, generally defined as:
Z minTðkÞ ðLI;end Þ ð22Þ
1
k
lk ¼ L f ðL; tÞdL ! V c ¼ kv l3 ð21Þ Subject to:
0

This means that the model is able to simulate and predict well the dT
0:5 6 6 0:5 ð23Þ
APFC dissolution cycle so it is suitable for optimization. Form II dt
slightly grows during the first seconds after seeding and then is
fully dissolved by the heating cycle. The amount of form I increases 11 6 T 6 45 ð24Þ
because of nucleation, then decreases during the dissolution cycle
because of partial dissolution of form I with form II and then C end 6 C max;end ¼ 0:005 g=g solvent ð25Þ
increases again due to growth.
lII1;end ¼ 0 ð26Þ
3.2. Process optimization for polymorphic crystallization
where T is the temperature defined in °C, dT dt
the heating/cooling
Optimization was performed using the kinetic parameters to rates in °C/min, C end the solute concentration at the end of the batch
find the optimal temperature profile that eliminates form II and (g/g solvent) and lII1;end the second moment of form II at the end of
maximizes the size of the crystals of form I at the end of the batch. the batch.
E. Simone et al. / Chemical Engineering Science 174 (2017) 374–386 383

x 10-3
14 0.016
Simulated
12 Cross-validation
Concentration (g/g solvent)
0.015

Concentration (g/g solvent)


Experimental

10 0.014

8 0.013

6 0.012

4 Simulated
0.011
Cross-validation
Experimental
2 0.01
0 2000 4000 6000 8000 10000 0 100 200 300 400
Time (s) Time (s)
(a) (b)

0.018 0.016

Simulated
0.017 Cross-validation 0.015
Concentration (g/g solvent)

Concentration (g/g solvent)


Experimental
0.016
0.014

0.015
0.013
0.014
0.012
0.013

0.011 Experimental
0.012 Simulated
Cross-validation
0.011 0.01
0 100 200 300 400 0 100 200 300 400
Time (s) Time (s)
(c) (d)

Fig. 8. (a) Secondary nucleation of form I during transformation (four isothermal experiments): continuous line is simulated and dots are experimental data (b–d). Primary
nucleation of form I after seeding (three experiments, similar conditions): continuous line is simulated and dots are experimental data.

Table 4
Initial conditions for the model validation shown in Figs. 9
and 10.

Validation initial conditions parameters Value

Seeding temperature½ C 37.26


g
Solute concentration½g solvent  0.0151
Mass of seed crystals ½g 0.61
Form II in the seed crystals ½%w=w 60
Mass of solvent ½g 400
Form I sigma;rI ½lm 6
Form I mean;lI ½lm 69
Form II sigma;rII ½lm 225
Form II mean;lII ½lm 75

A 20 min stabilization time was applied: the final temperature


was kept constant to allow the consumption of the remained
supersaturation. The initial conditions used for the optimization
are shown in Table 5.
The results of the optimization (shown in Fig. 11a–c) demon- Fig. 9. Model validation: experimental and simulated data for an APFC experiment.
strate that a heating step is not only required to eliminate form Temperature plotted with experimental and simulated concentrations.
II but also allows larger crystal size of form I at the end of the
batch: imposing only cooling in the optimization code
   The value of the objective function calculated was quite low
0:5 min
C
6 dT
dt
6 0:001 C=min resulted in lower crystal size, l
compared to the experimental results (l1;Iopt ¼ 26:5 lm, corre-
although all the metastable form naturally converted to the stable 0;Iopt
l
one by the end of the batch. sponding to a l4;Iopt ¼ 44:9 lm versus around 100–150 mm obtained
3;Iopt
384 E. Simone et al. / Chemical Engineering Science 174 (2017) 374–386

800 2000

Raman signal intensity form II (a.u.)

Raman signal intensity form I (a.u.)


1800
700
1600
600
1400
500 1200
400 1000

300 800
600
200
400
100
200
0 0
0 500 1000 1500 2000 0 2000 4000 6000 8000 10000
Time (s) Time (s)
(a) (b)

(c) (d)

Fig. 10. Model validation: experimental and simulated data for an APFC experiment. Raman signal intensities for form I and II during the experiment (a and b) compared to
the simulated third moment of both polymorphs.

Table 5 end of the profile to reach the desired yield. Fig. 11b shows the
Initial conditions for the optimization. optimized temperature profile in the phase diagram: the solute
Validation initial conditions parameters Value concentration is kept below the solubility of the metastable form
Seeding temperature ½ C 37.26
to avoid its further nucleation and above the solubility of the stable
g
Solute concentration ½g solvent  0.015 form to allow its growth during the cooling phase.
Mass of solid ½g 0.6
Form II in the seed crystals ½%w=w 40
4. Conclusions
Mass of solvent ½g 400
Form I sigma;rI ½lm 10
Form I mean;lI ½lm 50 The active polymorphic feedback control (APFC) is a strategy
Form II sigma;rII ½lm 10 that detects and eliminates the metastable polymorph after nucle-
Form II mean;lII ½lm 50 ation of a mixture or contaminated seeding (Simone et al., 2014).
The approach uses a combination of Raman spectroscopy to detect
the metastable polymorph and trigger a dissolution cycle to elim-
experimentally at the end of the batch) but this is most probably inate it, and then applies ATR-UV/Vis spectroscopy to grow the
due to the uncertainty of the parameters estimated for nucleation remaining crystals of the stable form through supersaturation con-
of form I. The heating step at the very beginning of the optimal trol. Despite being very efficient in obtaining the pure stable poly-
temperature profile is not only has the effect of dissolving form morph, this model-free control does not lead to optimal crystal size
II, but it is also beneficial to improve the crystal size distribution distribution of the stable polymorph at the end of the batch. In fact,
of form I. the size distribution of the crystals of the stable form after the dis-
The presence of heating steps in optimized batch crystallization solution cycle is not controlled and it might not be the optimal to
processes was already found by other authors (Majumder and allow a good quality CSD at the end of the supersaturation control.
Nagy, 2013; Qamar et al., 2010; Yeom et al., 2013; Nagy et al., For this reason a model-based active polymorphic control
2011) as a result of the inclusion of the dissolution kinetics in (mbAPC), that allows both the elimination of the metastable form
the PBEs. In those cases heating can correct a non-optimal seeding and larger crystals of the stable form at the end of the batch, was
and allows a better final CSD. After the heating step in the optimal developed. The kinetic parameters that are needed to describe
profile calculated in this work the temperature is kept high in order the mbAPC for ortho-aminobenzoic acid (dissolution and growth
to allow growth of the form I crystal and then drops towards the of form I and II, secondary nucleation of form I) were estimated
E. Simone et al. / Chemical Engineering Science 174 (2017) 374–386 385

Fig. 11. (a) Optimal temperature profile plotted against time; (b) Optimal operating trajectory plotted in the phase diagram; (c) Optimal crystal size distribution (CSD)
obtained at the end of the batch.

and validated using the data from seeded experiments. A specific References
design of experiments was performed to estimate each parameter
separately and therefore, to avoid correlations between them, as Alatalo, H., Hatakka, H., Kohonen, J., Reinikainen, S.P., Louhi-Kultanen, M., 2010.
Process control and monitoring of reactive crystallization of L-glutamic acid.
well as to simplify the parameter estimation. All the parameters AIChE J. 56 (8), 1063–2076.
estimated presented a narrow 95 % confidence interval, apart from Beckmann, W., 2000. Seeding the desired polymorph: background, possibilities,
the nucleation of the stable form, probably because of the stochas- limitations, and case studies. Org. Process Res. Dev. 4, 372–383.
Besenhard, M.O., Chaudhury, A., Vetter, T., Ramachandran, R., Khinast, J.G., 2015.
tic nature of this phenomenon. Evaluation of parameter estimation methods for crystallization processes
After the parameter estimation, optimization was performed. It modeled via population balance equations. Chem. Eng. Res. Des. 94, 275–289.
was found that the dissolution cycle, normally induced by the Cardew, P.T., Davey, R.J., 1985. The kinetics of solvent-mediated phase
transformations. Proc. Roy. Soc. A 398, 415–428.
APFC, not only allows the elimination of the metastable form II,
Chew, J.W., Black, S., Chow, P.S., Tan, R.B.H., Carpenter, K.J., 2007. Stable
but it is also beneficial to obtain larger crystals of form I at the polymorphs: difficult to make and difficult to predict. Cryst. Eng. Commun. 9,
end of the batch. This is in accordance with experiments as well 128–130.
as with the results of other optimization studies where dissolution Cornel, J., Lindenberg, C., Mazzotti, M., 2009. Experimental characterization and
population balance modeling of the polymorph transformation of L-glutamic
was included in the model. acid. Cryst. Growth Des. 9 (1), 243–252.
In conclusion, the proposed mbAPC can be useful for the design Doki, N., Seki, H., Takano, K., Asatani, H., Yukota, M., Kubota, N., 2004. Process
of batch crystallization processes of polymorphic systems as it control of seeded batch cooling crystallization of the metastable a-form glycine
using in-situ ATR-FTIR spectrophotometer and an in-situ FBRM particle counter.
allows obtaining large crystals of the stable form, even in case of Cryst. Growth Des. 4 (5), 949–953.
in situ nucleation of a mixture of the stable and metastable poly- Doki, N., Yukota, M., Kido, K., Sasaki, S., Kubota, N., 2003. Reliable and selective
morphs or erroneous seeding. crystallization of the metastable a-form glycine by seeding. Cryst. Growth Des.
4 (1), 103–107.
Fevotte, G., Caillet, A., Nida, S.O., 2007. A population balance model of the solution-
mediated phase transition of citric acid. AIChE J. 53 (10).
Acknowledgements Gunawan, R., Fusman, I., Braatz, R.D., 2004. High resolution algorithms for
multidimensional population balance equations. AIChE J. 50 (14), 2738–2749.
Hansen, N., Mueller, S.D., Koumoutsakos, P., 2003. Reducing the time complexity of
Financial support provided by the European Research Council the derandomized evolution strategy with covariance matrix adaptation (CMA-
grant no. [280106-CrySys] is acknowledged. ES). Evol. Comput. 11 (1), 1–18.
386 E. Simone et al. / Chemical Engineering Science 174 (2017) 374–386

Hermanto, M.W., Braatz, R.D., Chiu, M.S., 2011. Integrated batch-to-batch and Nagy, Z.K., Aamir, E., Rielly, C.D., 2011. Internal fines removal using a population
nonlinear model predictive control for polymorphic transformation in balance model based control of crystal size distribution under dissolution,
pharmaceutical crystallization. AIChE J. 57 (4), 1008–1019. growth and nucleation mechanisms. Cryst. Growth Des. 11, 2205–2219.
Hermanto, M.W., Chiu, M.S., Braatz, R.D., 2009. Nonlinear model predictive control Ono, T., Kramer, H.J.M., Ter Horst, J.H., Jansens, P.J., 2004. Process modeling of the
for the polymorphic transformation of L-glutamic acid crystals. AIChE J. 55 (10), polymorphic transformation of L-glutamic acid. Cryst. Growth Des. 4 (6), 1161–
2631–2645. 1167.
Hermanto, M.W., Chiu, M.S., Woo, X.Y., Braatz, R.D., 2007. Robust optimal control of Qamar, S., Mukhtar, S., Seidel-Morgenstern, A., 2010. Efficient solution of a batch
polymorphic transformation in batch crystallization. AIChE J. 53 (10). crystallization model with fines dissolution. J. Cryst. Growth 312, 2936–2945.
Jiang, S., ter Horst, J.H., Jansens, P.J., 2010a. Control over polymorph formation of o- Qu, H., Alatalo, H., Hatakka, H., Kohonen, J., Louhi-Kultanen, M., Reinikainen, S.,
aminobenzoic acid. Cryst. Growth Des. 10, 2541–2547. Kallas, J., 2009. Raman and ATR FTIR spectroscopy in reactive crystallization:
Jiang, S., ter Horst, J.H., Jansens, P.J., 2010b. Mechanism and kinetics of polymorphic simultaneous monitoring of solute concentration and polymorphic state of the
transformation of o-aminobenzoic acid. Cryst. Growth Des. 10, 2123–2128. crystals. J. Cryst. Growth 311 (13), 3466–3475.
Jiang, S., ter Horst, J.H., Jansens, P.J., 2008. Concomitant polymorphism of Scholl, J., Bonalumi, D., Vicum, L., Mazzotti, M., 2006. In situ monitoring and
o-aminobenzoic acid in antisolvent crystallization. Cryst. Growth Des. 8 (1), modeling of the solvent-mediated polymorphic transformation of L-glutamic
37–43. acid. Cryst. Growth Des. 6 (4), 881–891.
Kee, N., Tan, R.B.H., Braatz, R.D., 2009a. Selective crystallization of the metastable a- Sheikholeslamzadeh, E., Rohani, S., 2013. Modeling and optimal control of solution
form of the L-glutamic acid using concentration feedback control. Cryst. Growth mediated polymorphic transformation of L-glutamic acid. Ind. Eng. Chem. Res.
Des. 9 (7), 3044–3051. 52 (2633), 2641.
Kee, N.C., Arendt, P.D., Tan, R.B.H., Braatz, R.D., 2009b. Selective crystallization of the Simone, E., Saleemi, A.N., Nagy, Z.K., 2014. Active polymorphic feedback control of
metastable anhydrate form in the enantiotropic pseudo-dimorph system of crystallization processes using a combined Raman and ATR-UV/Vis
L-phenilalanine using concentration feedback control. Cryst. Growth Des. 9 (7), spectroscopy approach. Cryst. Growth Des. 14 (4), 1839–1850.
3052–3061. Su, Q., Rielly, C.D., Powell, K.A., Nagy, Z.K., 2017. Mathematical modelling and
Kobari, M., Kubota, N., Hirasawa, I., 2014. A population balance model for solvent- experimental validation of a novel periodic flow crystallization using MSMPR
mediated polymorphic transformation in unseeded solutions. Cryst. Eng. crystallizers. AIChE J. 63, 1313–1327.
Commun. 16, 6049–6058. Szilagyi, B., Nagy, Z.K., 2016. Graphical Processing Unit (GPU) acceleration for
Lai, T.C., Cornevin, J., Ferguson, S., Li, N., Trout, B.L., Myerson, A.S., 2015. Control of numerical solution of population balance models using high resolution finite
polymorphism in continuous crystallization via mixed suspension mixed volume algorithm. Comput. Chem. Eng. 91, 167–181.
product removal systems cascade design. Cryst. Growth Des. 15 (7), 3374– Threlfall, T., 2000. Crystallization of polymorphs: thermodynamic insight into the
3382. role of solvent. Org. Process Res. Dev. 4, 384–390.
Ma, C.Y., Liu, J.J., Wang, X.Z., 2016. Measurement, modelling, and closed-loop Trifkovic, M., Rohani, S., Sheikhzadeh, M., 2012. Kinetics estimation and
control of crystal shape distribution: literature review and future perspectives. polymorphic transformation modeling of Buspirone hydrochloride. J.
Particuology 26, 1–18. Crystallization Process Technol. 2, 31–43.
Majumder, A., Nagy, Z.K., 2013. Fines removal in a continuous plug flow crystallizer Wantha, L., Flood, A., 2013. Population balance modeling of the solution-mediated
by optimal spatial temperature profiles with controlled dissolution. AIChE J. 59 transformation of DL-Methionine polymorphs. Chem. Eng. Technol. 36 (8),
(12), 4582–4594. 1313–1319.
Minamisono, T., Takiyama, H., 2013. Control of polymorphism in the anti-solvent Yeom, S., Yun, H., Yang, D.R., 2013. Optimization of temperature swing strategy for
crystallization with a particular temperature profile. J. Cryst. Growth 362, 135– selective cooling crystallization of a-form L-glutamic acid crystals. Korean J.
139. Chem. Eng. 30 (10), 1836–1842.

You might also like