Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total...

59
Numerical analysis of in- flight freezing droplets: Application to novel particle engineering technology Author names and affiliations Andrew Tait a , Jonathan G. M. Lee a , Bruce R. Williams b and Gary A. Montague c a Biopharmaceutical Bioprocessing Technology Centre, School of Chemical Engineering and Advanced Materials, Newcastle University, Newcastle upon Tyne, Tyne and Wear, NE1 7RU b Accelyo Limited, 35 Gypsy Lane, Nunthorpe, Middlesbrough, Tees Valley, TS7 0DU c School of Science, Engineering and Design, Teesside University, Middlesbrough, Tees Valley, TS1 3BX Corresponding author Andrew Tait [email protected] Biopharmaceutical Bioprocessing Technology Centre, Page 1 of 59 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19

Transcript of Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total...

Page 1: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Numerical analysis of in-flight freezing droplets: Application to novel particle

engineering technology

Author names and affiliations

Andrew Tait a, Jonathan G. M. Lee a, Bruce R. Williams b and Gary A. Montague c

a Biopharmaceutical Bioprocessing Technology Centre, School of Chemical Engineering and

Advanced Materials, Newcastle University, Newcastle upon Tyne, Tyne and Wear, NE1 7RU

b Accelyo Limited, 35 Gypsy Lane, Nunthorpe, Middlesbrough, Tees Valley, TS7 0DU

c School of Science, Engineering and Design, Teesside University, Middlesbrough, Tees

Valley, TS1 3BX

Corresponding author

Andrew Tait

[email protected]

Biopharmaceutical Bioprocessing Technology Centre,

School of Chemical Engineering and Advanced Materials,

Newcastle University,

Newcastle upon Tyne,

Tyne and Wear,

NE1 7RU

United Kingdom

+44 (0) 7841656281

Page 1 of 39

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

Page 2: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Abstract

The freezing of a stream of free-falling monodispersed droplets was simulated through the

development of numerical models in this work. Prediction of the freezing time and

temperature transition of a single droplet is beneficial for optimisation of novel continuous

spray freeze drying (cSFD) processes. Estimations of the vertical free-falling distance of the

droplets in a slip stream and predictions of the chances of droplet coalescence greatly

enhance process understanding and can be leveraged to direct equipment design and process

development. A design space of droplet diameters in the range from 100 m to 400 m and

ambient temperatures from -120°C to -40°C was explored. The rate of supercooling within

the design space was predicted to range from 48 to 830°C s-1 depending on the ambient

temperature and droplet size. A comparison of the vertical free-falling distances of solitary

droplets and streams of droplets at different temperatures showed that the terminal velocity of

a vertically falling stream of droplets is always in excess of the terminal velocity of a solitary

droplet of the same size. A difference of 1.35 m was predicted for the free-falling distance of

a 400 m droplet compared to a stream of droplets at -42°C. A comparison between flow

rates for consecutively generated 100 m droplets showed that droplet coalescence was

predicted at 0.05 Lh-1, whilst at 0.02 Lh-1 a separation distance of 23 m was maintained thus

preventing coalescence.

Keywords

Continuous Spray Freeze Drying; Freeze Drying; Lyophilisation; Freezing Transitions;

Freezing; Numerical Modelling; Spray Freezing

Abbreviations

cSFD, continuous spray-freeze-drying

Page 2 of 39

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

Page 3: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

ICH, International Council for Harmonisation of Technical Requirements for

Pharmaceuticals for Human Use

NDF, numerical differentiation formula

NDS, new drug substances

NMR, nuclear magnetic resonance

VSVO, variable-step variable-order

Introduction

Freeze-drying is both a time-consuming and energy-intensive process with cycles typically

taking days or even weeks 1. A shorter cycle has the advantage of higher throughput for a

given freeze dryer, enabling the use of fewer freeze dryers for a given amount of product 2.

The risk of product contamination or batch failure increases with lengthy cycle times, and

manufacturing under sterile conditions is cost intensive. Scalability from lab- to production-

scale is complex and requires characterisation of the equipment for robust scale-up 3–6, such

as shelf temperature characterisation and surface mapping 7. Batch non-uniformity is an

ongoing concern, as the vials may exhibit different time evolutions due to radiation from the

wall of the chamber, temperature gradients on the heating shelf, vapour fluid dynamics, and

non-uniform inert distribution 8,9.

Continuous processing is becoming established in the food industry to address these issues

with processes that combine frozen microsphere production by spray dispersion with

dynamic bulk freeze-drying. Process intensification by reducing the volume scale from

millilitre filled vials to nanolitre monodispersed droplets increases the achievable heat and

mass transfer capabilities. The increase in surface area to volume ratio results in rapid

freezing and drying. This continuous spray-freeze-drying (cSFD) process is a novel particle

engineering technology for the production of powders, as the freezing rate can be controlled

Page 3 of 39

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

Page 4: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

to determine the internal structure of the particles. The cSFD process uses liquid nitrogen

vapour as the ambient gas.

The aim of this study was to develop numerical models that would simulate the freezing of a

stream of monodispersed droplets in a cSFD process. The numerical models were extended

from the foundation works of Hindmarsh et al. to accommodate the differences in processing

conditions 10–12. Droplet vertical free-falling distance and droplet coalescence predictions

were incorporated into the numerical models to enhance process understanding, guide

equipment design, size and scale, and establish product processing boundaries.

Background

The temperature transition of a freezing aqueous solution droplet is important in determining

the morphology of the final dry product 10–12. The governing equations for the four stages of

the freezing process are discontinuous and can be defined as:

Supercooling: The solution is cooled from the initial state to a temperature below the

equilibrium freezing point Tf until nucleation occurs.

Nucleation and recalescence: Sufficient supercooling has occurred for spontaneous

nucleation at Tn. The supercooling drives rapid kinetic crystal growth from the nuclei,

where an abrupt temperature rise is apparent as this growth liberates the latent heat of

fusion. Recalescence terminates when the supercooling is exhausted and the

equilibrium freezing temperature Tf is reached.

Freezing: Further growth of the solid phase is governed by the rate of heat transfer to

the environment. Freezing continues until the solution has completely solidified.

Tempering: The temperature of the frozen solution reduces to a steady-state value

near that of the ambient temperature Ta.

The ability to predict the freezing process is important as changes to the product formulation

can affect the ice nucleation temperature 13; freezing point depression and freeze

Page 4 of 39

69

70

71

72

73

74

75

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

Page 5: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

concentration of solutes 14; and the crystal growth rate during recalescence 15. Understanding

of how these changes affect the stages of freezing can lead to greater process control and

reduce the risk of unfrozen product accumulation in equipment or non-uniform product

within a batch.

Numerical modelling of the solidification of droplets has been well investigated in the

literature for droplets of water 16–18, food solutions 19, and metals 20–25. A key limitation of

these studies was the lack of experimental data to validate the model performance. However,

works by 10,11 addressed this issue by obtaining empirical data for suspended single droplets

on the junction of a thermocouple for both pure water and sucrose solutions. The application

of nuclear magnetic resonance (NMR) by Hindmarsh et al. also proved advantageous for the

verification of theoretical solidification models 12.

It was found that the addition of sucrose significantly influenced the mass transfer rates from

unfrozen and frozen droplets 11. Higher sucrose concentrations resulted in lower mass transfer

rates from both the unfrozen and frozen droplets, and significantly lower rates when

compared to water for frozen droplets. The mass transfer rate was also found to be lower at

reduced temperature for frozen droplets. The experimental data demonstrated a total droplet

mass reduction of up to 5% within the first minute but there was negligible observable

change within the timescale of this numerical study. Around 1.2% of the total heat transfer

was attributed to mass transfer during numerical modelling of the supercooling, freezing and

tempering stages of a 1% w/v sucrose solution droplet with a diameter of 100 μm. Based on

this numerical prediction and the experimental observations of Hindmarsh et al. 11, it was

assumed that mass transfer could be neglected as there was limited potential to improve

model accuracy from increased complexity. Collection of further experimental data to

account for the reduced droplet size, temperature and ambient gas composition was beyond

the scope of this work. Experimental validation of the mass transfer rates was limited by the

Page 5 of 39

94

95

96

97

98

99

100

101

102

103

104

105

106

107

108

109

110

111

112

113

114

115

116

117

118

Page 6: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

sensitivity of measurement equipment. A droplet suspended on the junction of a

thermocouple would demand a rapid response time and a very small wire diameter to

minimise the impact on the droplet. The inclusion of a thermocouple into the droplet would

act as a nucleation site, which would affect the temperature transition profile compared with a

stream of free-falling droplets. Thus, an improved experimental method is required for

determining the mass transfer within the temperature and droplet size ranges explored within

this numerical model.

The numerical analyses were performed under conditions described by small Biot numbers

(Bi < 0.1), so that a lumped model for the droplet temperature could be used that did not

account for heat conduction in the interior of the droplet and assumed that there was a

uniform temperature distribution within the droplet during cooling and freezing 11,12. The Biot

number is defined as Bi=h Lc/k d where h is the heat transfer coefficient, Lc = Vd/Sd is the

characteristic length, and kd is the thermal conductivity of the droplet, as given in Table 1

where w per is the weight fraction of the solution and the temperature T is in Kelvin. The

correlation factor k d for thermal conductivity calculations 26 were consistent with the

experimental results within ± 0.012 27.

The freezing stage can be solved most simply by balancing the rate of heat transfer from the

droplet with the internal heat transfer. The external heat fluxes govern the freezing of the

remaining liquid following the formation of the initial solid volume during recalescence. The

heat transfer within the droplet is dependent on the assumed form of the initial solid volume

produced following nucleation with the possibilities being:

Inward moving boundary: A solid shell on the surface which grows inwardly towards

the centre of the droplet. External heat fluxes conduct heat through the solid shell to

the droplet surface.

Page 6 of 39

119

120

121

122

123

124

125

126

127

128

129

130

131

132

133

134

135

136

137

138

139

140

141

142

Page 7: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Outward moving boundary: A solid core at the centre of the droplet which grows

outwardly to the surface. External heat fluxes conduct heat through the undercooled

liquid to the droplet centre.

Heat balance: The solid volume fraction is uniformly dispersed within the droplet.

The solid fraction increases uniformly throughout the droplet as the latent heat of

fusion is removed by external heat transfer.

The inward moving boundary, outward moving boundary, and heat balance numerical models

were found to result in reasonable predictions of the experimental freezing times 10. The

internal heat balance of the droplet was accurately solved within the experimental uncertainty

by employing a simple heat balance model for each of the liquid cooling, freezing, and solid

cooling stages.

The scanning electron microscopy images presented in Figure 1 further support the freezing

process assumption, as the thin shell and lack of visible dendrite growth in the internal

structure of the fractured particles demonstrates minimal freeze concentration, which is a

result of rapid ice growth speeds inhibiting solute diffusion 28. The purity of ice crystals

formed in ice crystallisation is dependent on growth rate. If the ice growth rate is low enough

for the solute to diffuse away from the advancing ice front, then pure ice is formed, and the

solution becomes more concentrated until the eutectic point is reached. Conversely, if the

growth rate is so high that all solutes are entrapped in the ice then this leads to impure ice

crystals and no enrichment of the solution 29. This phenomenon explains the increase in

primary drying times experienced for freezing at faster rates, as the impure ice crystals

impede vapour flow during sublimation, thus reducing the driving force behind primary

drying.

Figure 1. Scanning electron micrographs of skimmed milk (10% w/w) particles produced by spray freezing into

vapour over liquid: (a) particle size distribution, (b) single particle, (c) particle surface, and (d) fractured

particle.

Page 7 of 39

143

144

145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162

163

164

165

166

167

168

Page 8: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Numerical models

The temperature transition of the droplet can be solved by balancing the internal energy with

the energy removed by convective heat transfer qh and thermal radiation qr from the droplet

surface. The governing equations of the freezing process describing the internal heat balance

of the droplet and the physical properties are discontinuous for the four stages. Separate

models were formulated for the supercooling, nucleation and recalescence, freezing, and

tempering stages. Model parameters are presented in Table 1.

Surface heat fluxes

The convective heat flux qh from the droplet surface and the thermal radiation heat flux qr are

respectively described by qh=h (T d−Ta ) and qr=εσ (T d4−T a

4 ) where Td is the droplet

temperature, Ta is the ambient temperature, and ε and σ are the emissivity for thermal

radiation and the Stefan Boltzmann constant, respectively 30. The ambient temperature for

radiation is assumed to be the liquid nitrogen vapour temperature as the surrounding surfaces

for the cSFD equipment are thermally conductive and well insulated. The correlation for the

Nusselt number Nu was rearranged to estimate the heat transfer coefficient 31 as h, where r is

the droplet radius and the thermal conductivity of nitrogen gas is k. For flow past a sphere,

the Reynolds number is given by ℜ=2 r ρa v /μa where v is the relative velocity, μa is the

ambient viscosity, and ρa is the ambient density 30, where Pa is the ambient pressure, M ( N 2 )

is the molar mass of nitrogen, and R is the universal gas constant. The ambient viscosity μa is

determined from the Sutherland's formula 32, as follows:

μa=μa ,0( T a ,0+SN

T a+SN)( T a

T a , 0)

32

Equation 1

Page 8 of 39

169

170

171

172

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

Page 9: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

where μa , 0 is the reference viscosity, T a ,0 is the reference temperature, and SN is the

Sutherland constant for nitrogen gas. The Sutherland formula, Equation 1, is valid in the

region from 100 K to 1900 K.

The relative velocity v was taken with respect to the terminal velocity of the droplet v t, which

was determined from the Galileo number 33, as follows:

Ga={ 18ℜ , Ga0 ≤ 3.618ℜ+2.7 ℜ1.687 , 3.6<Ga0≤ 105

13 ℜ2 , Ga0>105

Equation 2

where the initial Galileo number is defined as:

Ga0=(2 r )3 ρa ( ρl−ρa ) g

μa2

Equation 3

Balancing the initial Galileo number (Equation 3) with the Galileo number (Equation 2),

substituting in the Reynolds number, and rearranging with respect to the terminal velocity v t

enables the Reynolds number to be regressed. The density of supercooled water ρl 34 is

determined from:

ρl=(∑n=0

6

AnT r , ln )×103

Equation 4

where A0 to A6 are fitting parameters, as given in Table 1, and T r ,l is the relative temperature.

The Prandtl number Pr used in determining the heat transfer coefficient is given by

Pr=c p , a μa/k where the ambient specific heat capacity c p , a is determined from

c p , a=B1 eB 2T a+B3 eB 4 Ta where B1, B2, B3 and B4 are fitting parameters 35, as given in Table 1.

The thermal conductivity k of nitrogen gas can be evaluated using the method for pure non-

polar gases (linear molecules) at low pressure 36, as follows:

Page 9 of 39

191

192

193

194

195

196

197

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

Page 10: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

k=cv ,a μa

M N [1.30+( Rcv ,a )(1.7614−0.3523

T r , a )]Equation 5

The reduced ambient temperature T r , a=T a /Tc where T c is the critical temperature 37. The

ideal gas isochoric heat capacity cv , a is given by cv , a=c p0−R. The ideal gas isobaric heat

capacity c p0 for nitrogen is determined from:

c p0=R (C1

T a3 +

C2

T a2 +

C3

T a+C4+C5 T a+C6 Ta

2+C7 Ta3+

C8 (C9/T a )2eC9 /T a

(eC9/Ta−1 )2 )Equation 6

where C1 to C9 are fitting parameters 38, as given in Table 1.

Supercooling

The temperature transition for the liquid cooling and supercooling of the liquid droplet is

determined by the heat balance:

dT d

dt=

Sd

c p ,l ρl V d(qh+qr )

Equation 7

where Sd and V d are the surface area and volume of the droplet, respectively. The specific

heat capacity of supercooled water c p ,l 39, used in Equation 7, is given by

c p ,l=(D1 eD2 T d+D3 eD4 T d ) /M ( H2O ) where D1, D2, D3 and D4 are fitting parameters, as given

in Table 1, and M ( H2O ) is the molar mass of water.

Nucleation and recalescence

The temperature transition for a droplet with a high degree of supercooling occurs extremely

rapidly at the time of nucleation until the equilibrium freezing temperature is reached. The

initial frozen volume V s0 produced during the recalescence stage can be estimated using the

experimentally validated Stefan number Ste 40.

Page 10 of 39

213

214

215

216

217

218

219

220

221

222

223

224

225

226

227

228

229

230

231

232

233

234

Page 11: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

The Stefan number is a dimensionless representation of the latent to sensible heat in a system.

Assuming that the liquid instantaneously rises to the equilibrium freezing temperature, the

Stefan number will represent the solid fraction formed during recalescence. The Stefan

number can be adjusted to compensate for the density difference between the solid and liquid

phases. The solid fraction V s0 formed from recalescence is given by:

V s0=V d

c p ,l ρl

ρ s

( T f−T n )L f

Equation 8

where Lf is the latent heat of fusion 30, the nucleation temperature T n is assumed to be the

homogeneous nucleation temperature of water, 235 K 41, and T f is the equilibrium freezing

temperature T f =273.15−Δ T f . The freezing point depression, ΔT f , is defined as the

difference between the equilibrium freezing point of the pure solvent and the freezing point

of the solution, which is calculated from Blagden's Law for ideal solutionsΔT f =K f ∙ b ∙ i

where b is the molality, and i is the van't Hoff factor, which must be equal to 1 for diluted

solutions, as this simple relationship does not include the nature of the solution. The

cryoscopic constant for water K f is dependent on the properties of the solvent, not the solute.

The molality is given by b=( ρs V s)/M s ρl V l where M s is the solute molar mass and V s is the

solute volume. The solid phase density ρ s was evaluated from the density of ice below 273.15

K, and is defined as:

ρ s=∑n=1

5

En T d(5−n)

Equation 9

where E1 to E5 are fitting parameters 35, as given in Table 1. The solid phase density used was

that of ice due to the insolubility of common bulk excipients in ice on a micro-scale 11. When

the nucleation temperature T n is reached, the droplet is assumed to instantly change to the

equilibrium freezing temperature T f and have partially frozen volume V s0.

Page 11 of 39

235

236

237

238

239

240

241

242

243

244

245

246

247

248

249

250

251

252

253

254

255

256

257

258

Page 12: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Freezing

The freezing stage can be modelled most simply with an energy balance, assuming a uniform

temperature within the droplet, by balancing the external heat transfer with the amount of

latent heat required to be removed in order to completely freeze the mass of water in the

droplet 16:

dV s

dt=

Sd

L f ρs(qh+qr )

Equation 10

The heat balance equation was solved with a forward difference time step. For the time

period t+dt, the change in the solid volume fraction d V s was calculated from the total heat

flux from the droplet for the time step dt . Blagden's Law was employed at each iteration to

calculate the freezing point depression with the increased solute concentration within the

liquid phase. The solvent mass is given by ms=ρsolvent V l where the liquid volume fraction

V l=V d−V sf−V s was adjusted with the change in the solid volume fraction V sf to calculate

the molality required to evaluate the equilibrium freezing temperature at each time step. The

phase change calculations were terminated when the droplet was fully frozen at V l=0.

Tempering

The temperature transition in the solid cooling or tempering stage of the solid frozen droplet

is determined by a heat balance similar to Equation 7 for the supercooling stage:

d T d

dt=

Sd

c p , s ρsV d( qh+qr )

Equation 11

where the solid phase parameters replace the liquid phase specific heat capacity and density.

The specific heat capacity of ice c p , s 35 is given by c p , s=F1T d3+F2T d

2+F3T d+F4 where F1 to

F4 are fitting parameters, as given in Table 1.

Page 12 of 39

259

260

261

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

277

278

279

280

281

Page 13: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Droplet vertical free-falling distance

Extension of the single droplet numerical model to a system of monodispersed droplets gives

rise to directly applicable solutions in cSFD processes. Study of solitary droplets alone is

beneficial to enhance understanding of the product, such as morphology, and processing

conditions, such as the temperature transition of a freezing droplet. Numerical evaluation of

the vertical free-falling distance travelled by monodispersed droplets is valuable in equipment

design and determining processing boundaries. The total droplet vertical free-falling distance

dd is defined as:

dd=∑i

v t ,i Δt i

Equation 12

The iterative time step Δt i is determined from the solutions of the supercooling, freezing and

tempering stages. The terminal droplet stream velocity v t ,i 42 is evaluated as:

v t=12 [st +(st

2+4 A )12 ]

Equation 13

The droplet and ambient nitrogen streams approach a terminal condition where the drop

velocity and peak average nitrogen velocity have constant values. In the terminal state:

A= m gπ μa a (1− (k 2−k3/2 )

k1)(1−

ρa

ρd )Equation 14

where k1=0.52034, k 2=0.20992, k3=0.31934 and a=ln (2 ) are constants 42, and m is the

mass flow rate of droplets. The standard terminal velocity st of a solitary falling sphere is

given by:

st=29

( ρd−ρa )μa

g ( Dd

2 )2

Equation 15

Page 13 of 39

282

283

284

285

286

287

288

289

290

291

292

293

294

295

296

297

298

299

300

301

302

303

304

Page 14: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

where μa is the ambient viscosity.

Droplet coalescence

The velocity of a stream of vertically free-falling monodispersed droplets generated with jet

velocity will adjust to the terminal droplet stream velocity during flight. It was assumed that

the jet velocity is within the Rayleigh jet breakup regime and is equal to the initial droplet

velocity. If the droplet jet velocity v j is greater than the terminal droplet stream velocity, v t

then there is a possibility that the droplets will coalesce before reaching terminal velocity.

Droplet coalescence is assumed to occur when the separation distance between two

consecutively generated droplets is equal to zero. The separation distance S between two

vertically free-falling droplets d1 and d2 is defined as the pitch less the droplet diameter,

where Dd=Dd1=Dd2.

The droplet vertical free-falling distance is denoted by δ d. The separation distance S at time t

can then be derived from the cumulative free-falling distances ∑ δd 1 and ∑ δd 2

travelled by

droplets d1 and d2, respectively:

S (t )=∑i=0

t

δ d1 ,i−∑i=0

t

δ d2 ,i−Dd

Equation 16

where ∑ δd 1−∑ δ d2

is the pitch between droplets d1 and d2. If monodispersed droplets are

generated at a frequency of Fd then droplet d2 will be generated 1/ Fd s after droplet d1,

which is termed the fixed time step Δt Fd. Applying this fixed time step to Equation 16 and

eliminating the cumulative free-falling distance results in S ( t )=δ d ( t )−Dd.

When S (t )=0 then it is assumed that the droplets begin to coalesce. The droplet vertical free-

falling distance is defined as δ d ( t )=v d (t ) ∆ tF d. The droplet velocity vd ( t ) is calculated from the

Page 14 of 39

305

306

307

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

325

326

Page 15: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

preceding droplet velocity, acceleration and free-falling distance at time t−∆tF d. Therefore,

the droplet velocity is given by vd (t )=[vd2 ( t−∆ tF d )+2 ad δd ( t−∆ tF d )]

12 .

The initial droplet velocity vd (0 ) is defined as the jet velocity v j and the initial free-falling

distance is evaluated as δ d (0 )=λ where λ is the spatial disturbance wavelength. The

acceleration of the free-falling droplet is ad (t )=[wd−R ( t ) ] /md where wd is the droplet weight

and md is the droplet mass. The aerodynamic resistance R ( t )=12

A X ρa vd2 (t ) CD ( t ) where ρa is

the ambient density, AX=π ( Dd/2 )2 is the cross-sectional area of the droplet, and CD (t ) is the

drag coefficient:

CD (t )= 8ℜ ( 2+3κ

1+κ )Equation 17

where κ=μa/μ is the viscosity ratio of the dispersed phase to the continuous phase 43.

Table 1. Fitting parameters.

Numerical solutions

The numerical model was programmed in MATLAB R2015b (The MathWorks Inc., USA).

The user inputs to the numerical model are displayed in Table 2. The stiff differential

equations were evaluated with the ode15s multistep solver, which is a variable-step variable-

order (VSVO) solver based on the numerical differentiation formulas (NDFs) of orders 1 to 5

and requires the solutions at several preceding time points to compute the solution.

Table 2. Numerical model: user inputs.

Results

Biot number

The Biot numbers for droplet diameters ranging from 100 m to 400 m explored in this

study for the extremes of the ambient temperature range from 153 K to 233 K were evaluated

Page 15 of 39

327

328

329

330

331

332

333

334

335

336

337

338

339

340

341

342

343

344

345

346

347

348

349

Page 16: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

to be below the Bi<0.1 condition. The 400 m droplet at the coldest ambient temperature

approaches the limit of the assumption of uniform temperature distribution. Larger droplets at

colder ambient temperatures would exceed the small Biot number assumption and

incorporation of heat conduction within the droplet would be required for prediction under

these conditions.

Temperature transition

The temperature transition of a freezing droplet is one of the outputs of the numerical model.

The temperature transition from an initial temperature of 293 K to a final temperature of 253

K, with a homogeneous nucleation temperature of 235 K and an equilibrium freezing

temperature of 273 K was studied. The results showed clearly defined liquid (supercooling)

and solid (tempering) cooling stages.

The recalescence stage following nucleation differed from the general case, as the deeply

supercooled liquid droplet temperature instantaneously rose from the homogeneous

nucleation temperature of water to the equilibrium freezing point when the latent heat of

supercooling was released. The assumption of an instantaneous temperature rise in the

numerical model is attributed to the deep supercooling which drives the rapid temperature

transition.

The freezing stage was solved with an adjusted heat balance model which accommodated the

freezing point depression as the solute concentration increased in the unfrozen solution.

During rapid ice crystal growth, the solute is entrapped in the ice matrix, as solute diffusion

away from the water/ice interface is inhibited, which significantly reduces the effects of

freeze concentration.

A sensitivity analysis was conducted to assess the assumption of homogeneous nucleation at

253 K. The total time taken to completely freeze and temper a droplet with a diameter of 200

m was considered. Two ambient nitrogen temperatures of 123 K and 223 K, which

Page 16 of 39

350

351

352

353

354

355

356

357

358

359

360

361

362

363

364

365

366

367

368

369

370

371

372

373

374

Page 17: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

represent a typical operating range of cSFD equipment, were used in the numerical analysis.

The final tempered droplet temperature was taken at 0.5 K above the ambient temperature in

each instance. The time and percentage variation at 123 K was 0.78 s± 1.5 %, and at 223 K

was 1.00 s± 3.1%.

The effect of nucleation temperature on the freezing time was shown for two different

ambient temperatures 10, and the experimental results demonstrated that the freezing time was

only significantly influenced by the nucleation temperature if the ambient temperature was

close to the minimum nucleation temperature. These results are in agreement with the results

of the sensitivity analysis, as the freezing time variation increases as the ambient temperature

approached the homogeneous nucleation temperature.

Rate of supercooling

The predicted rates of supercooling were calculated for a range of ambient liquid nitrogen

vapour temperatures for four droplet sizes. The results of the numerical analysis are presented

in Figure 2. As can be seen, the rate of supercooling is affected by both the ambient

temperature and droplet size, which emphasises the potential for improved process control

compared to batch lyophilisation, where only the supercooling rate can be controlled within a

much more limited range. Rapid supercooling at pilot-scale for batch processes is relatively

easy to achieve with pre-cooled shelves. However, at industrial scale, freezing at the same

rates is unrealistic due to issues with product preparation, such as filling and loading times.

The product morphology is dependent on the structure formed during the freezing stage.

Typical batch freeze-drying cycles are cooled at a rate of 0.5 K min-1 to maximise the ice

crystal size upon nucleation. This reduces the primary drying time due to increased mass

transfer through the pores as the ice is sublimed.

Figure 2. Rate of supercooling for droplet diameters of 100 m (––), 200 m (- - -), 300 m (⋯), and

400 m (- ∙ -).

Page 17 of 39

375

376

377

378

379

380

381

382

383

384

385

386

387

388

389

390

391

392

393

394

395

396

397

398

399

Page 18: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

A comparison can be made between the intensified freeze-drying process with rapid liquid

nitrogen freezing of droplets and the traditional batch freeze-drying process in vials using a

metric of mass transfer. The limit to mass transfer is described by the dry layer resistance

Rp=A p ( P i−Pc )/ (dm /dt ) where Pi is the vapour pressure of the sublimating ice, Pc is the

chamber pressure, Ap is the cross-sectional area of the vial, and dm /dt is the sublimation rate

of ice, which is experimentally determined 44. Assuming that the processing conditions are the

same, then a large reduction in the area from the cross-section of a vial to the surface area of

a droplet, would result in a much lower dry layer resistance. However, the sublimation rate is

dependent on the ice crystal structure of the product.

The sublimation rate for a 7.5% (w/w) trehalose solution frozen by ramped cooling on a shelf

at both slow (0.1 K min-1) and moderate (1.67 K min-1) rates, and by liquid nitrogen

quenching at a rapid rate (108 K min-1) was compared 45. The results show little difference in

the sublimation rate for the slow and moderate freezing, however, the rapid freezing

demonstrated a significantly longer primary drying time. The mass transfer is limited due to

the small pores formed from the ice crystal structure during rapid freezing, which is reflected

in the reduced sublimation rate seen in the results. Applying these observations to the

sublimation rate would result in an increased dry layer resistance for a rapidly frozen product.

The surface area to volume ratio for the droplet sizes studied was in the range from 15000 m-1

to 60000 m-1 and for a 20 mL filled vial was 141 m-1. Even with the increased dry layer

resistance due to the product morphology, the overall 1000-fold surface area reduction from

the cross-sectional area of a vial to the surface area of a droplet would significantly reduce

the dry layer resistance, thereby reducing the overall drying time.

Droplet vertical free-falling distance

The free-falling distances were calculated based on five ambient liquid nitrogen vapour

temperatures. The capacity to predict the free-falling distance travelled by droplets provides

Page 18 of 39

400

401

402

403

404

405

406

407

408

409

410

411

412

413

414

415

416

417

418

419

420

421

422

423

424

Page 19: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

an indication as to whether the droplets will be sufficiently solidified within the equipment

limits to prevent accumulation of unfrozen product.

It was found that if the droplet frequency and jet velocity are low, then the motion of a

droplet stream is an approximation of a solitary droplet 42. Conversely, if the droplet

frequency and jet velocity are high, then the droplet stream accelerates, and the solitary

droplet decelerates. Experimental results demonstrated that the entrainment of air by a stream

of droplets provides a sheath, in the form of an air jet, through which the droplet stream can

slip relative to the surrounding medium 42. The terminal velocity of a vertically falling stream

of droplets is always in excess of the terminal velocity of a solitary droplet of the same size.

These findings support the results shown in Figure 3, which presents a comparison between

the vertical free-falling distances of solitary droplets and streams of droplets at different

ambient temperatures. Figure 3 demonstrates that colder ambient temperatures result in

smaller vertical free-falling distances before droplets are sufficiently solidified.

Figure 3. Vertical free-falling distance for sufficient droplet solidification evaluated for a solitary droplet

at ambient temperatures of 153 K (- - -) and 233 K (- · -) and a droplet stream at ambient temperatures

of 153 K (––) and 233 K (· · ·) with droplet diameters in the range from 100 m to 400 m.

Droplet coalescence

Droplet separation distance will increase as droplet velocity increases from jet velocity v j to

terminal droplet stream velocity v t. If v j<v t then the vertical free-falling distance D1, at

which point the droplet velocity becomes terminal, must not exceed the vertical free-falling

distance D2, where the separation distance equals zero and the droplets coalesce. Droplets

with diameters in the range of 200 m to 400 m were shown not to coalesce for a flow rate

of 0.05 Lh-1, as the terminal droplet stream velocity v t exceeded the jet velocity v j. At the

same flow rate, 100 m droplets were shown to have a higher initial jet velocity, which

would result in the droplets slowing to the terminal droplet stream velocity during flight and

Page 19 of 39

425

426

427

428

429

430

431

432

433

434

435

436

437

438

439

440

441

442

443

444

445

446

447

448

449

Page 20: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

potentially causing coalescence. Since D1 (11.61 mm) was greater than D2 (8.61 mm), it was

predicted that the droplets would coalesce. Reducing the flow rate for the 100 m droplets to

0.02 Lh-1 was shown to slightly increase the vertical free-falling distance D2 (8.92 mm) but

significantly reduce the vertical free-falling distance where the terminal droplet stream

velocity is reached (6.77 mm). This resulted in the estimation that the droplets would not

coalesce.

A comparison between the two flow rates of 0.02 Lh-1 and 0.05 Lh-1 for consecutively

generated 100 m droplets is shown in Figure 4. The separation distance S is equal to zero

for the 0.05 Lh-1 flow rate of droplets at a vertical free-falling distance of 8.61 mm, at which

point the droplets coalesce. For the 0.02 Lh-1 flow rate of droplets, the terminal droplet stream

velocity is reached at a vertical free-falling distance of 6.77 mm thus preventing coalescence

by maintaining a separation distance of 22.85 m.

Figure 4. Separation distance S between two consecutively generated 100 m droplets at flow rates of 0.02 Lh-1

(––) and 0.05 Lh-1 (- - -). When S=0 then it is assumed that the droplets begin to coalesce. Terminal drop

stream velocity is reached (⋯) for the 0.02 Lh-1 droplets before coalescence occurs.

Discussion

Coupling the intensified freezing process with an ability to adjust the ambient temperature

and droplet size to achieve the desired product morphology is beneficial to produce novel

powders and for making improvements to existing products. Spray-freezing has been applied

as a novel particle engineering technology to enhance aqueous dissolution of poorly water

soluble drugs 46–48. Novel particle engineering overcomes the limitations of conventional size

reduction techniques, such as milling, which exhibits impeded control of particle size, shape,

morphology, surface properties and electrostatic charge. Conventional methods of size

reduction are less efficient due to high energy demands, risk of thermal or chemical

degradation of the product, and non-uniform particle size distribution.

Page 20 of 39

450

451

452

453

454

455

456

457

458

459

460

461

462

463

464

465

466

467

468

469

470

471

472

473

474

Page 21: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Low water solubility of a high percentage of biopharmaceuticals is a big issue for

pharmaceutical applications due to the resulting low absorption and hence limited therapeutic

efficacy 49. Overcoming this issue is important as about 40 % of the compounds being

developed by the pharmaceutical industry are poorly water soluble 50,51. Various efforts have

been made to address the problem by delivery in nanocarriers and by nanoparticles

engineering. Drug nanoparticles can be produced by a range of top-down and bottom-up

approaches. The size of large drug nanoparticles can be reduced by high pressure

homogenisation, milling, or microfluidisation for the top-down approach, or for the bottom-

up approach, the drug nanoparticles are formed from molecules in an aqueous, organic, or

aqueous-organic co-solvent solutions, aqueous-organic emulsions, or suspensions 49.

Nanoparticle engineering enables manufacturing of poorly water-soluble drugs into

nanoparticles alone, or incorporation with a combination of pharmaceutical excipients. The

use of these processes has dramatically improved in vitro dissolution rates and in vivo

bioavailabilities of many poorly water-soluble drugs 52.

The dissolution of poorly water-soluble APIs is the rate-limiting step to absorption, it is

important to improve the wetting and dissolution properties to thereby enhance absorption

and bioavailability 48. cSFD processes can be employed to produce nanoparticles with an

amorphous structure, high surface area, and enhanced wettability to greatly improve

dissolution. The feed solution for a cSFD process can be produced by dissolving a

hydrophobic API and/or hydrophilic excipients in a solution. Some poorly water-soluble

APIs have relatively low solubility in aqueous-organic co-solvent solutions, however,

resulting in low solution loading 48. The use of organic solvents, such as acetonitrile, ethanol,

acetone, methanol, and methylene chloride, can greatly increase the solubility of hydrophobic

APIs, which results in enhanced drug loading in the feed solution.

Page 21 of 39

475

476

477

478

479

480

481

482

483

484

485

486

487

488

489

490

491

492

493

494

495

496

497

498

Page 22: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

The benefits of freeze-drying droplets with diameters in the ¿400 m range are not just

purely processing-related, as a number of product-specific applications have been identified

within the literature. From a regulatory perspective, the ICH Q6A guidance demonstrates the

criticality that particle size may have to dissolution, solubility, or bioavailability of certain

drug products. The ability to process a drug product and the stability, content uniformity, and

product appearance are also all affected by particle size. For some new drug substances

(NDS) intended for use in solid or suspension drug products, particle size can have a

significant effect on dissolution rates, bioavailability, and stability.

Conclusions

This work has presented numerical models to simulate the freezing of a stream of free-falling

monodispersed droplets for application in a cSFD process. It was demonstrated that being

able to predict the freezing times and temperature transition of a single droplet is essential for

optimisation of spray-freezing processes 12. Estimating the vertical free-falling distances of

monodispersed droplets in a slip stream and predicting the chances of droplet coalescence can

greatly enhance process understanding and guide equipment development. A design space of

droplet diameters in the range from 100 m to 400 m and ambient temperatures from 153 K

to 233 K was explored. Further experimental work is required to verify the validity of the

vertical free-falling distance and droplet coalescence numerical models. Simulation of hail

formation has been the focus of previous studies 53 which has application for the freezing of

in-flight droplets. A subject of future study would be to combine the numerical models for

freezing of a stream of free-falling droplets presented in this work with droplet generation

models in a computational fluid dynamics (CFD) simulation within the design space

described above for droplet diameters and ambient temperatures.

The small Biot number condition was satisfied for the experimental design space. Therefore,

a uniform temperature distribution was assumed within the droplet during supercooling and

Page 22 of 39

499

500

501

502

503

504

505

506

507

508

509

510

511

512

513

514

515

516

517

518

519

520

521

522

523

Page 23: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

freezing, which did not account for heat conduction within the interior of the droplet. It was

found that the assumption of a uniform temperature profile within a droplet was empirically

justified and accurate for droplets with small Biot numbers10,11.

The temperature transitions of freezing droplets were evaluated for 1% lactose solutions

within the experimental design space from an initial temperature of 293 K to a final droplet

temperature of 253 K, with a homogeneous nucleation temperature of 235 K. An adjusted

heat balance model was used to solve the freezing stage, which accommodated the freezing

point depression as the solute concentration increased in the unfrozen solution. The effects of

freeze concentration were significantly reduced as solute diffusion away from the water/ice

interface is inhibited during rapid ice crystal growth. The solute is entrapped in an ice matrix

and no solute shell is formed, as demonstrated by the scanning electron microscopy (SEM)

images presented in Figure 1. It was illustrated that the surface morphology of droplets can be

controlled with the addition of excipients and/or changing the freezing conditions 54. Most

significantly, the rate of supercooling offers a means of controlling the surface properties of a

powder produced by cSFD. It was noted that if the formulation contains a sugar, then an outer

surface layer may be formed which can affect the quality of the powder. It was demonstrated

that a high freezing rate with high supercooling inhibited the formation of a sucrose layer on

the outer surface of the droplets, which could reduce the stickiness and increase the

flowability of the powder 54.

The assumption of homogeneous nucleation resulted from the rapid rates of supercooling

experienced by small droplets in cSFD processes and the minimal heterogeneous nucleation

sites present in the filtered precursor solutions. The reduced variability from homogeneous

nucleation compared to heterogeneous nucleation is beneficial for cGMP processes and

minimising batch to batch variation. A sensitivity analysis concluded that an approximate

Page 23 of 39

524

525

526

527

528

529

530

531

532

533

534

535

536

537

538

539

540

541

542

543

544

545

546

547

Page 24: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

± 3% variation in total freezing time is experienced for ambient temperatures close to the

nucleation temperature, which is well within the computational error.

The rates of supercooling were determined within the experimental design space for a 1%

lactose solution. Superior process control is afforded to cSFD processes compared to batch

freeze-drying, as the rate of supercooling is affected by both the ambient temperature and the

droplet size. The supercooling rate is determined by the temperature alone for batch

lyophilisation processes within a much more limited range. Issues with rapid freezing in

batch lyophilisation processes become more prevalent at industrial scale, where product

preparation becomes a limiting factor. The product morphology is dependent on the structure

formed during the freezing stage. Further studies would be necessary to determine the

relationship between the degree of supercooling and the ice crystal structure formed at rapid

freezing rates.

The design of equipment for cSFD processes can be guided by the estimation of droplet

vertical free-falling distances. The slip stream effect and phase changes within the droplets

during freezing result in droplet velocity fluctuations during flight. Ensuring that droplets are

sufficiently frozen within the equipment limits can prevent the accumulation of unfrozen

product leading to downstream blockages. Estimation of vertical free-falling distances

evaluated from solitary droplet calculations alone were shown to under predict the total

distance travelled. Entrainment of the surrounding medium by a stream of droplets provides a

sheath, in the form of a jet, through which the droplet stream can slip. Therefore, the terminal

velocity of a vertically falling stream of droplets is always more than the terminal velocity of

a solitary droplet of the same diameter. The numerical results could influence equipment

design and process control by increasing the height to accommodate larger droplet diameters,

reducing the ambient temperature when processing larger droplet diameters, or focusing on

smaller droplet diameters. Consideration of the vertical free-falling distance alone is not

Page 24 of 39

548

549

550

551

552

553

554

555

556

557

558

559

560

561

562

563

564

565

566

567

568

569

570

571

572

Page 25: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

sufficient to prevent unfrozen product accumulation, as droplet velocity changes, from the

initial jet velocity to the terminal droplet stream velocity, can lead to coalescence. For the 100

m diameter droplets, coalescence was predicted for a droplet mass flow rate of 0.05 Lh-1.

Reducing the flow rate to 0.02 Lh-1 resulted in a droplet separation distance of 22.85 m,

which would prevent coalescence.

Acknowledgements

This work was supported by the Engineering and Physical Sciences Research Council

(EPSRC) [Grant Number EP/G037620/1].

References

(1) Tang, X.; Pikal, M. J. Design of Freeze-Drying Processes for Pharmaceuticals:

Practical Advice. Pharm. Res. 2004, 21 (2), 191–200.

(2) Pikal, M. J. Freeze-Drying of Proteins. Part I: Process Design. BioPharm 1990, 3 (8),

18–27.

(3) Rambhatla, S.; Pikal, M. J. Heat and Mass Transfer Scale-up Issues during Freeze-

Drying, I: Atypical Radiation and the Edge Vial Effect. AAPS PharmSciTech 2003, 4

(2), 22–31.

(4) Rambhatla, S.; Ramot, R.; Bhugra, C.; Pikal, M. J. Heat and Mass Transfer Scale-up

Issues during Freeze Drying: Ii. Control and Characterisation of the Degree of

Supercooling. AAPS PharmSciTech 2004, 5 (4), 54–62.

(5) Rambhatla, S.; Tchessalov, S.; Pikal, M. J. Heat and Mass Transfer Scale-up Issues

during Freeze-Drying, Iii: Control and Characterisation of Dryer Differences via

Operational Qualification Tests. AAPS PharmSciTech 2006, 7 (2), E61–E70.

(6) Pikal, M. J.; Bogner, R.; Mudhivarthi, V.; Sharma, P.; Sane, P. Freeze-Drying Process

Development and Scale-up: Scale-up of Edge Vial versus Centre Vial Heat Transfer

Coefficients, Kv. J. Pharm. Sci. 2016, 105 (11), 3333–3343.

Page 25 of 39

573

574

575

576

577

578

579

580

581

582

583

584

585

586

587

588

589

590

591

592

593

594

595

596

597

Page 26: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

(7) Biddlecombe, J. Advancements in Lyophilised Drug Product Development and Scale-

Up. In Pharmaceuticals and Biopharmaceuticals: Lyophilisation; The SMi Group:

London, 2014.

(8) Barresi, A. A.; Pisano, R.; Fissore, D.; Rasetto, V.; Velardi, S. A.; Vallan, A.; Parvis,

M.; Galan, M. Monitoring of the Primary Drying of a Lyophilization Process in Vials.

Chem. Eng. Process. Process Intensif. 2009, 48 (1), 408–423.

(9) Barresi, A. A.; Pisano, R.; Rasetto, V.; Fissore, D.; Marchisio, D. L. Model-Based

Monitoring and Control of Industrial Freeze-Drying Processes: Effect of Batch

Nonuniformity. Dry. Technol. 2010, 28 (5), 577–590.

(10) Hindmarsh, J. P.; Russell, A. B.; Chen, X. D. Experimental and Numerical Analysis of

the Temperature Transition of a Suspended Freezing Water Droplet. Int. J. Heat Mass

Transf. 2003, 46 (7), 1199–1213.

(11) Hindmarsh, J. P.; Russell, A. B.; Chen, X. D. Experimental and Numerical Analysis of

the Temperature Transition of a Freezing Food Solution Droplet. Chem. Eng. Sci.

2004, 59 (12), 2503–2515.

(12) Hindmarsh, J. P.; Wilson, D. I.; Johns, M. L.; Russell, A. B.; Chen, X. D. NMR

Verification of Single Droplet Freezing Models. AIChE J. 2005, 51 (10), 2640–2648.

(13) Roos, Y. H. Phase Transitions and Transformations in Food Systems. In Handbook of

Food Engineering; Heldman, D. R., Lund, D. B., Eds.; CRC Press: Boca Raton, FL,

2007; pp 287–352.

(14) Mellor, J. D. Fundamentals of Freeze Drying; Academic Press: London, 1978.

(15) Schwegman, J. J.; Hardwick, L. M.; Akers, M. J. Practical Formulation and Process

Development of Freeze Dried Products. Pharm. Dev. Technol. 2005, 10, 151–173.

(16) Liao, J.; Ng, K. Effect of Ice Nucleators on Snow Making and Spray Freezing. Ind.

Eng. Chem. Res. 1990, 29 (3), 361–366.

Page 26 of 39

598

599

600

601

602

603

604

605

606

607

608

609

610

611

612

613

614

615

616

617

618

619

620

621

622

Page 27: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

(17) Gao, W.; Smith, D. W.; Sego, D. C. Freezing Behavior of Freely Suspended Industrial

Wastewater Droplets. Cold Reg. Sci. Technol. 2000, 31 (1), 13–26.

(18) Feuillebois, F.; Lasek, A.; Creismeas, P.; Pigeonneau, F.; Szaniawski, A. Freezing of a

Subcooled Liquid Droplet. Journal of Colloid And Interface Science. 1995, pp 90–102.

(19) Wilson, H. A.; Singh, R. P. Numerical Simulation of Individual Quick Freezing of

Spherical Foods. Int. J. Refrig. 1993, 10, 149–155.

(20) Epstein, M.; Fauske, H. S. Kinetic and Heat Transfer Controlled Solidification of

Highly Supercooled Droplets. Int. J. Heat Mass Transf. 1993, 36, 2987–2995.

(21) Grant, P. S.; Cantor, B.; Katgerman, I. Modelling of Droplet Dynamic and Thermal

Histories during Spray Forming: I. Individual Droplet Behaviour. Acta Metall. Mater.

1993, 41, 3097–3108.

(22) Lee, E.; Ahn, S. Solidification Progress and Heat Transfer Analysis of Gas-Atomised

Alloy Droplets during Spray Forming. Acta Metall. Mater. 1994, 44, 3231–3243.

(23) McCoy, J. K.; Markworth, A. J.; Collings, R. S.; Brodkey, R. S. Cooling and

Solidification of Liquid-Metal Drops in a Gaseous Atmosphere. J. Mater. Sci. 1992,

27, 761–766.

(24) Su, Y. H.; Tsao, C. A. Modelling of Solidification of Molten Metal Droplet during

Atomisation. Metall. Mater. Trans. B 1997, 28, 1249–1255.

(25) Wang, G. X.; Matthys, E. F. Numerical Modelling of Phase Change and Heat Transfer

Rapid Solidification Processes: Use of Control Volume Integrals with Element

Subdivision. Int. J. Heat Mass Transf. 1992, 35, 141–153.

(26) Bubnik, Z.; Kadlec, P.; Urban, D.; Bruhns, M. Sugar Technologists Manual -

Chemical and Physical Data for Sugar Manufacturers and Users; Albert Bartens:

Berlin, Germany, 1995.

(27) Werner, M.; Baars, A.; Werner, F.; Delgado, A. Thermal Conductivity of Aqueous

Page 27 of 39

623

624

625

626

627

628

629

630

631

632

633

634

635

636

637

638

639

640

641

642

643

644

645

646

647

Page 28: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Sugar Solutions under High Pressure. Int. J. Thermophys. 2007, 28 (4), 1161–1180.

(28) Butler, M. F. Freeze Concentration of Solutes at the Ice/Solution Interface Studied by

Optical Interferometry. Cryst. Growth Des. 2002, 2 (6), 541–548.

(29) Hasan, M.; Louhi-Kultanen, M. Ice Growth Kinetics Modelling of Air-Cooled Layer

Crystallisation from Sodium Sulphate Solutions. Ice growth Kinet. Model. air-cooled

layer Cryst. from sodium sulphate Solut. 2015, 133, 44–53.

(30) Incropera, F. P.; Dewitt, D. P.; Bergman, T. L.; Lavine, A. S. Principles of Heat and

Mass Transfer, 7th ed.; John Wiley and Sons: Singapore, 2012.

(31) Ranz, W. E.; Marshall, W. R. Evaporation from Drops - Part 1. Chemical Engineering

Progress. 1952, pp 141–148.

(32) Smits, A. J.; Dussauge, J.-P. Turbulent Shear Layers in Supersonic Flow, 2nd ed.;

Spinger: New York, 2006.

(33) Richardson, J. F.; Harker, J. H.; Backhurst, J. R. Particle Technology and Separation

Processes, 5th ed.; Oxford, 2002.

(34) Hare, D. E.; Sorensen, C. M. The Density of Supercooled Water. II. Bulk Samples

Cooled to the Homogeneous Nucleation Limit. J. Chem. Phys. 1987, 87 (8), 4840–

4845.

(35) Weast, R. C. Handbook of Chemistry and Physics, 1st ed.; CRC Press: Boca Raton,

FL, 1988.

(36) Stiel, L. I.; Thodos, G. The Thermal Conductivity of Non-Polar Substances in the

Dense Gaseous and Liquid Regions. AIChE J. 1964, 10, 26–30.

(37) Green, D. W.; Perry, R. H. Chemical Engineering Handbook, 8th ed.; McGraw-Hill:

New York, 2008.

(38) Jacobsen, R. T.; Stewart, R. B. Thermodynamic Properties of Nitrogen Including

Liquid and Vapour Phases from 63 K to 2000 K with Pressures to 10,000 Bar. J. Phys.

Page 28 of 39

648

649

650

651

652

653

654

655

656

657

658

659

660

661

662

663

664

665

666

667

668

669

670

671

672

Page 29: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Chem. Ref. Data 1973, 2 (4), 757–922.

(39) Angell, C. A.; Oguni, M.; Sichina, W. J. Heat Capacity of Water at Extremes of

Supercooling and Superheating. J. Phys. Chem. 1982, 86 (6), 998–1002.

(40) Hindmarsh, J. P.; Wilson, D. I.; Johns, M. L. Using Magnetic Resonance to Validate

Predictions of the Solid Fraction Formed during Recalescence of Freezing Drops. Int.

J. Heat Mass Transf. 2005, 48 (5), 1017–1021.

(41) Starr, F. W.; Angell, C. A.; Stanley, H. E. Prediction of Entropy and Dynamic

Properties of Water below the Homogeneous Nucleation Temperature. Physica A

2003, 323, 51–66.

(42) Arrowsmith, A.; Foster, P. J. The Motion of a Stream of Monosized Liquid Drops in

Air. Chem. Eng. J. 1973, 5 (3), 243–250.

(43) Clift, R.; Grace, J. R.; Weber, M. E. Bubbles, Drops and Particles; New York, 1978.

(44) Koganti, V. R.; Shalaev, E. Y.; Berry, M. R.; Osterberg, T.; Youssef, M.; Hiebert, D.

N.; Kanka, F. A.; Nolan, M.; Barrett, R.; Scalzo, G.; et al. Investigation of Design

Space for Freeze Drying: Use of Modelling for Primary Drying Segment of a Freeze-

Drying Cycle. AAPS PharmSciTech 2011, 12 (3), 854–861.

(45) Gieseler, H.; Lee, G. Influence of Different Cooling Rate on Cake Structure of Freeze

Dried Samples Measured by Microbalance Technique. In Controlled Release Society

(CRS) German Chapter Annual Meeting; Munich, Germany, 2003.

(46) Rogers, T. L.; Hu, J.; Yu, Z.; Johnston, K. P.; Williams, R. O. A Novel Particle

Engineering Technology: Spray-Freezing into Liquid. Int. J. Pharm. 2002, 242 (1–2),

93–100.

(47) Rogers, T. L.; Johnston, K. P.; Williams III, R. O. Physical Stability of Micronised

Powders Produced by Spray Freezing into Liquid (SFL) to Enhance the Dissolution of

an Insoluble Drug. Pharm. Dev. Technol. 2003, 8, 187–197.

Page 29 of 39

673

674

675

676

677

678

679

680

681

682

683

684

685

686

687

688

689

690

691

692

693

694

695

696

697

Page 30: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

(48) Hu, J.; Johnston, K. P.; Williams, R. O. Spray Freezing into Liquid (SFL) Particle

Engineering Technology to Enhance Dissolution of Poorly Water Soluble Drugs:

Organic Solvent versus Organic/Aqueous Co-Solvent Systems. Eur. J. Pharm. Sci.

2003, 20 (3), 295–303.

(49) Grant, N.; Zhang, H. Poorly Water-Soluble Drug Nanoparticles via an Emulsion-

Freeze-Drying Approach. J. Colloid Interface Sci. 2011, 356, 573–578.

(50) Radtke, M. Pure Drug Nanoparticles for the Formulation of Poorly Soluble Drugs.

New Drugs 2001, 3, 62–68.

(51) Lipinski, C. Poor Aqueous Solubility — an Industry Wide Problem in Drug Delivery.

Am. Pharm. Rev. 2002, 5, 82–85.

(52) Hu, J.; Johnston, K. P.; Williams III, R. O. Rapid Dissolving High Potency Danazol

Powders Produced by Spray Freezing into Liquid Process. Int. J. Pharm. 2004, 271,

145–154.

(53) Phillips, V. T. J.; Khain, A.; Benmoshe, N.; Ilotoviz, E.; Ryzhkov, A. Theory of Time-

Dependent Freezing. Part II: Scheme for Freezing Raindrops and Simulations by a

Cloud Model with Spectral Bin Microphysics. J. Atmos. Sci. 2014, 72 (1), 262–286.

(54) Hindmarsh, J. P.; Russell, A. B.; Chen, X. D. Fundamentals of the Spray Freezing of

Foods-Microstructure of Frozen Droplets. J. Food Eng. 2007, 78 (1), 136–150.

Page 30 of 39

698

699

700

701

702

703

704

705

706

707

708

709

710

711

712

713

714

715

716

717

Page 31: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Table 1. Model parameters.

k d A w per+B

A 5.466 ×10−8∙ T 2−1.176 ×10−5 ∙T−3.024 ×10−3

B −7.847 ×10−6 ∙T 2+1.976 × 10−3 ∙ T +0.563

ε 0.95

σ 5.67 ×10−8 W m-2 K-4

h (k /2 r )(2+0.6 Pr13 ℜ

12)

ρa Pa M ( N2 ) /T a R

Pa 101325 kg m-1 s-2

M ( N 2 ) 0.0 280134 kg mol-1

R 8.314462 J mol-1 K-1

μa , 0 17.81× 10−6 kg m-1 s-1

T a ,0 300.55 K

SN 111 K

T r ,l T d−273.15 K

T c 126.192 K

M ( H2O ) 0.01801528 kg mol-1

Ste c p ,l (T f −T n )/ Lf

Lf 333700 J kg-1

K f 1.853 K kg mol-1

V s ( ws ρl V s )/ ( ρs+w s [ ρl−ρs ])Density of supercooled water

n An

0 9.999 ×10−1

1 6.690 ×10−5

2 −8.486 ×10−6

3 1.518 ×10−7

4 −6.948 ×10−9

5 −3.645 ×10−10

6 −7.497 ×10−12

Ambient specific heat capacity

Page 31 of 39

718

Page 32: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

n Bn

1 3.298 ×103

2 −4.918× 10−2

3 1.051× 103

4 −3.507×10−5

Ideal gas isobaric heat capacity

n Cn

1 −7.352 ×102

2 3.422 ×101

3 −5.577 ×10−1

4 3.504 ×100

5 −1.734 ×10−5

6 1.747 ×10−8

7 −3.569×10−12

8 1.005 ×100

9 3.353 ×103

Specific heat capacity of supercooled water

n Dn

1 1.195 ×1013

2 −1.138 ×10−1

3 8.012 ×101

4 −2.193×10−4

Density of ice

n En

1 4.328 × 10−9

2 −2.657 ×10−6

3 1.923×10−4

4 −5.038×10−3

5 9.338 ×102

Specific heat capacity of ice

n Fn

1 6.128 ×10−5

2 −3.286×10−2

Page 32 of 39

Page 33: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

3 1.254 ×101

4 −1.168 ×102

Page 33 of 39

719

720

Page 34: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Table 2. Numerical model user inputs.

User input Value

Droplet radius 100 – 400 m

Ambient nitrogen temperature 153 – 233 K

Homogeneous nucleation temperature 235 K

Initial droplet temperature 293 K

Final droplet temperature 253 K

Solute molar mass a 0.3423 kg mol-1

Solute density a 1525 kg m-3

Solute weight percentage a 1%a Solute properties for lactose.

Figure 1a

Figure 1b

Page 34 of 39

721

722

723

724

725

726

Page 35: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Figure 1c

Page 35 of 39

727

728

Page 36: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Figure 1d

Page 36 of 39

729

730

Page 37: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Figure 2

Page 37 of 39

731

732

Page 38: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Figure 3

Figure 4

Page 38 of 39

733

734

735

736

Page 39: Teesside University's Research Portal - Author names and ... · Web viewAround 1.2% of the total heat transfer was attributed to mass transfer during numerical modelling of the supercooling,

Page 39 of 39

737

738