copyright and use of this thesis · iii | page lactose) to 21.5 ± 0.5 m2 g-1.the typical specific...

187
COPYRIGHT AND USE OF THIS THESIS This thesis must be used in accordance with the provisions of the Copyright Act 1968. Reproduction of material protected by copyright may be an infringement of copyright and copyright owners may be entitled to take legal action against persons who infringe their copyright. Section 51 (2) of the Copyright Act permits an authorized officer of a university library or archives to provide a copy (by communication or otherwise) of an unpublished thesis kept in the library or archives, to a person who satisfies the authorized officer that he or she requires the reproduction for the purposes of research or study. The Copyright Act grants the creator of a work a number of moral rights, specifically the right of attribution, the right against false attribution and the right of integrity. You may infringe the author’s moral rights if you: - fail to acknowledge the author of this thesis if you quote sections from the work - attribute this thesis to another author - subject this thesis to derogatory treatment which may prejudice the author’s reputation For further information contact the University’s Director of Copyright Services sydney.edu.au/copyright

Upload: others

Post on 15-May-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

Copyright and use of this thesis

This thesis must be used in accordance with the provisions of the Copyright Act 1968.

Reproduction of material protected by copyright may be an infringement of copyright and copyright owners may be entitled to take legal action against persons who infringe their copyright.

Section 51 (2) of the Copyright Act permits an authorized officer of a university library or archives to provide a copy (by communication or otherwise) of an unpublished thesis kept in the library or archives, to a person who satisfies the authorized officer that he or she requires the reproduction for the purposes of research or study.

The Copyright Act grants the creator of a work a number of moral rights, specifically the right of attribution, the right against false attribution and the right of integrity.

You may infringe the author’s moral rights if you:

- fail to acknowledge the author of this thesis if you quote sections from the work

- attribute this thesis to another author

- subject this thesis to derogatory treatment which may prejudice the author’s reputation

For further information contact the University’s Director of Copyright Services

sydney.edu.au/copyright

Page 2: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

Enhancements in Spray-Dried Powder Functionality,

Using Multistage Fluidized-Bed Drying and a New

Templating Process

A thesis submitted in fulfilment of the requirements for the degree of

Doctor of Philosophy

by

Morteza Saffari

Supervisor: Prof. Timothy Langrish

April 2016

School of Chemical and Biomolecular Engineering

Page 3: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

i | P a g e

Declaration

This thesis is a presentation of my original research work. Wherever contributions of others

are involved, every effort is made to indicate this clearly, with due reference to the literature,

and acknowledgement of collaborative research and discussions. The work was done under

the guidance of Professor Timothy Langrish, at the University of Sydney, Australia.

Morteza Saffari

April, 2016.

Page 4: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

ii | P a g e

SUMMARY

Spray drying is a technique that is widely applied to produce dry powders from their liquid or

slurry states in the food and pharmaceutical industries. Due to the rapid removal of water in

this technique, amorphous solids may be formed. Spray-dried forms of dairy powders may

have several advantages, such as greater solubility, dissolution rates and surface areas.

However, their molecular structures are relatively disordered, and they are

thermodynamically-unstable solids, which may undergo undesirable changes, such as caking

after production.

To overcome this problem, a fluidized-bed dryer has been used in this study to improve the

degree of lactose crystallinity of sweet whey. Powder was fed at different rates for each

fluidization process at different sets of process temperature, humidity, and residence time. It

has been found that, with controlling the operating conditions in a continuous multi-stage

fluidized bed dryer, whey powders can have different degrees of lactose crystallinity between

10% and 20% with reasonable overall processing times of 30 minutes. Gravimetric moisture

sorption tests were performed on the degree of lactose crystallinity for the spray-dried

powders and the processed powders and compared with X-ray powder diffraction, Fourier

transform infrared spectroscopy, and Raman spectroscopy analyses. Experimental results

using all these methods confirm that that there were significant reductions in the amorphicity

of lactose at different humidities and temperatures and suggested that the overall extent of

lactose crystallinity for processed whey powders has been between 10% and 20% within the

processes studied.

However, crystalline powders may have some undesirable properties, such as lower

dissolution rates. The creation of highly-porous crystalline powders with better functional

(free flowing), and storage (non-caking) properties, would be helpful in the food industry and

for pharmaceutical formulations, especially for controlled drug delivery and release

applications. Thus, a new templating method has been developed to create spray-dried

powders with high porosity and controlled characteristics, while maintaining the high

stability of a highly crystalline powder. In this technique, solutions containing core materials

such as lactose/mannitol and ethanol-soluble food-grade acids as templating agents were

spray dried. The resulting powders were then washed with ethanol to remove the templating

agent and produce porous frameworks of lactose/mannitol. There were significant

improvements in the surface areas, from 0.3 ± 0.2 m2 g-1 (for conventional spray-dried

Page 5: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

iii | P a g e

lactose) to 21.5 ± 0.5 m2 g-1. The typical specific surface areas of mannitol powders have

been reported to be 0.3 m2 g-1 (Babu and Nangia, 2011; Ho et al., 2012) and this templating

method resulted in a mannitol network with significant porosity and a high surface area of

10.5 ± 0.5 m2 g-1.

It has been found that altering the concentration of templating acids was very effective in

changing the degrees of lactose crystallinity in the spray-dried products and the Brunauer,

Emmett and Teller (BET) surface areas of the resultant porous materials. Therefore, the effect

of different acids with various acidity indices, such as boric acid, ascorbic acid, and lactic

acid, has been studied on the crystallinity of the spray-dried powder, on the process yield, and

the templating performance of these acids in the production of high-porosity lactose particles.

The results of present study suggested that there was a link between the extent of lactose

crystallization in the spray-dried powders and the BET surface areas for the ethanol-washed

particles. Increasing the degree of crystallinity for the spray-dried powders at high acid

concentrations decreased the amounts of acid molecules incorporated in the lactose structure,

leading to lower BET surface areas.

It was found that increasing the templating-acid concentrations significantly increased the

degrees of crystallinity for lactose in spray-dried powders. Textural properties, such as the

surface area of the resultant lactose, decreased considerably by increasing the concentrations

of different templating acids, due to increase in the lactose crystallinity for the final spray-

dried products. The yields from spray drying were also decreased significantly at higher

concentrations of templating acids. For instance, when using lactic acid as a templating acid a

change in lactic acid concentration from 1 w/w % to 3 w/w % showed a reduction in the

surface areas from 14.9 ± 0.9 m2 g-1 to about 9.5 ± 0.8 m2 g-1 and the yield of the spray-drying

process altered from 71% ± 2% to about 44% ± 1%. Likewise, in the case of using citric acid

as a templating acid, it was observed that adding 1% w/w citric acid as a templating material

significantly increased the BET surface area of lactose to 20.8 ± 0.9 m2 g-1. However, adding

further citric acid content (3 w/w%) decreased the BET surface area of lactose to 12.3 ± 0.8

m2 g-1 and the yield of the process significantly decreased from 71% ± 2% to 16% ± 4%. The

results suggested that using templating acids, with lower glass-transition temperatures than

that for lactose, decreased the overall glass-transition temperature of the spray-dried mixture

according to the Gordon–Taylor equation. Lower glass-transition temperatures increase the

temperature difference (T-Tg), which in turn decreases the crystallization time as suggested

Page 6: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

iv | P a g e

by the Williams–Landel–Ferry equation, leading to higher degrees of lactose crystallinity for

the spray-dried particles. Therefore, the porosities and the BET surface areas of the powders

decreased as the amorphous content decreased, which is connected with the decrease in the

amount of incorporated templating-acid molecules, due to the increase in the degree of

crystallinity for the spray-dried lactose powders. However, the main reason for this yield

behavior is due to having lower glass-transition temperatures in the powders and also greater

product moisture contents at higher templating-acid concentrations, which make the particles

stickier. The optimum concentration of lactic acid, citric acid, and ascorbic acid were found

to be 1 w/w % with the BET surface area of 14.9 ± 0.9 m2 g-1, 20.8 ± 0.9 m2 g-1, and 10.5 ±

0.6 m2 g-1, respectively while that for the spray-dried lactose with boric acid as a templating

agent was 3 w/w % with the BET surface area of 18.5 ± 0.8 m2 g-1.

These engineered particles were then used as porous excipients in an adsorption technique to

evaluate the potential of this technique for improvement of the dissolution characteristics of

poorly water-soluble drugs and for content uniformity enhancement by incorporating the drug

molecules onto this porous structure. Indomethacin, nifedipine, and acetaminophen as model

drugs were studied. This method significantly improved the dissolution rate of these poorly

water-soluble drugs due to the nanoconfinement of drug crystals inside the porous structure

and the high solubility of porous mannitol, with 80% drug release within the first fifteen

minutes for the drug-loaded samples. Tuneable drug loadings have also been achieved with

low variability for the drug content with all drug-loaded samples, indicating the ability of this

method to produce highly-uniform blends. The results of drug loading for nifedipine showed

an increase from 3.2±0.1 w/w % for a 0.08 M drug solution to 9.1±0.3 w/w % drug loading

for a 0.16 M drug solution, while indomethacin had slightly better performance for the

adsorption process, with 4.1±0.2 w/w % and 12.6±0.4 w/w % for 0.08 M and 0.16 M

concentrations of indomethacin, respectively, in the final formulation. These blends were also

very uniform, with a percentage relative standard deviation of less than 4% for drug-loaded

mannitol in both nifedipine and indomethacin.

Page 7: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

v | P a g e

ACKNOWLEDGEMENTS

I would like to thank and express my sincere gratitude to my supervisor Professor Timothy

Langrish for his kind supervision, direction, support, and inspiration which helped me to be

motivated, develop research abilities, and complete this study in due time. His patience and

time to teach and guide me are much appreciated.

I would like to thank University of Sydney for the postgraduate scholarship and the financial

support.

The facilities and the scientific and technical assistance of the Australian Microscopy &

Microanalysis Research Facility at the Electron Microscope Unit, The University of Sydney

are gratefully acknowledged.

Finally, I would like to thank my parents and family members, especially my dear nephew

Amirali for being supportive. Their sacrifice, inspiration, support, and encouragements

helped me to continue this job and reach this far.

Morteza Saffari

April, 2016.

Page 8: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

vi | P a g e

TABLE OF CONTENTS

SUMMARY ............................................................................................................................. II

ACKNOWLEDGEMENTS ................................................................................................... V

TABLE OF CONTENTS ..................................................................................................... VI

LIST OF FIGURES ................................................................................................................ X

LIST OF TABLES ............................................................................................................. XIII

NOMENCLATURE ........................................................................................................... XIV Brunauer, Emmett and Teller ............................................................................................................................. xiv Barrett-Joyner-Halenda ...................................................................................................................................... xiv

CHAPTER 1 INTRODUCTION .......................................................................................... 1

1.1 Spray Drying ............................................................................................................................................... 1

1.2 Enhancements in Spray-Dried Powder Functionality ............................................................................ 1

1.3 Application of New Engineered Particles with High Porosities and Controlled Characteristics ..... 2

1.4 Thesis Structure .......................................................................................................................................... 4

CHAPTER 2 LITERATURE REVIEW .............................................................................. 8

2.1 Amorphous and Crystalline Powders ....................................................................................................... 8

2.2 Sticking and Caking Mechanism of Amorphous Lactose ...................................................................... 8

2.3 Crystallization Kinetics ............................................................................................................................ 10 2.3.1 Modelling the Kinetics of Solid-Phase Crystallization with the Williams–Landel–Ferry Equation ... 11 2.3.2 Modelling the Kinetics of Solid-Phase Crystallization with the Avrami Equation .............................. 12

2.4 The Controlled Crystallization Processes .............................................................................................. 13 2.4.1 Modified Spray-Drying Technique ........................................................................................................ 14 2.4.2 Crystallization in Fluidised-Bed Dryers/Crystallizers .......................................................................... 15

2.5 Types and Composition of Whey ............................................................................................................ 18

2.6 Effect of Acidity on Lactose Crystallinity .............................................................................................. 19

2.7 Disadvantages of Crystalline Powders ................................................................................................... 20

2.8 A New Templating Method to Produce Highly Crystalline Powder with Significant Porosity and High Surface Area .................................................................................................................................................. 21

Page 9: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

vii | P a g e

2.9 Mannitol as an Alternative to Lactose ................................................................................................... 25

2.10 Highly Porous Excipients as a Solution for Poor Solubility and Poor Content Uniformity in Solid Dosage Form ........................................................................................................................................................... 27

CHAPTER 3 MATERIALS, EQUIPMENT AND METHODS ...................................... 30

3.1 Materials .................................................................................................................................................... 30

3.2 Equipment ................................................................................................................................................. 31 3.2.1 Spray Dryer ............................................................................................................................................ 31 3.2.2 Fluidized-Bed Dryer ............................................................................................................................... 32

3.3 Powder Characterization ......................................................................................................................... 35

3.4 Measurements of Degree of Crystallization and Thermal Transitions .............................................. 37 3.4.1 Moisture Sorption Tests ......................................................................................................................... 37 3.4.2 X-Ray Diffraction (XRD) ...................................................................................................................... 38 3.4.3 Modulated Differential Scanning Calorimetry (MDSC) ....................................................................... 38 3.4.4 Raman Spectroscopy .............................................................................................................................. 39 3.4.5 Fourier Transform Infrared Spectroscopy (FTIR) ................................................................................. 39

3.5 Characterization of Drug-Loaded Powders .......................................................................................... 40 3.5.1 Determination of Drug Loading and Blend Uniformity ........................................................................ 40 3.5.2 Dissolution Experiments ........................................................................................................................ 40 3.5.3 Tableting Properties ............................................................................................................................... 42

3.6 Particle Sizing and Morphology Studies ................................................................................................ 42 3.6.1 Scanning Electron Microscope (SEM) .................................................................................................. 42 3.6.2 Particle Size Analysis ............................................................................................................................. 43 3.6.3 BET Surface Area Analysis ................................................................................................................... 43

3.7 Other Items of Methods and Equipments .............................................................................................. 43 3.7.1 Moisture Content Measurement ............................................................................................................. 43 3.7.2 pH of Aqueous Solution ......................................................................................................................... 43 3.7.3 Statistics .................................................................................................................................................. 43

CHAPTER 4 COMBINED DRYING AND CRYSTALLIZATION IN A CONTINUOUS MULTI-STAGE FLUIDIZED-BED DRYER ......................................... 45

4.1 Sample Preparation .................................................................................................................................. 46 4.1.1 Fresh Spray-Dried Whey Powder .......................................................................................................... 46 4.1.2 Crystalline Whey Powders ..................................................................................................................... 46

4.2 Process Conditions .................................................................................................................................... 47

4.3 Results and Discussion ............................................................................................................................. 48 4.3.1 Moisture Sorption Tests ......................................................................................................................... 51 4.3.2 X-Ray Diffraction (XRD) ...................................................................................................................... 52 4.3.3 Estimating the Degree of Surface Crystallinity Using Raman and FTIR ............................................. 55 4.3.4 Particle Size Analysis ............................................................................................................................. 61 4.3.5 Surface Morphology ............................................................................................................................... 62

Page 10: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

viii | P a g e

4.4 Conclusions ................................................................................................................................................ 64

CHAPTER 5 EFFECT OF ACIDITY ON IN-PROCESS CRYSTALLIZATION OF LACTOSE DURING SPRAY DRYING .............................................................................. 66

5.1 Materials and Methods ............................................................................................................................ 67 5.1.1 Sample Preparation ................................................................................................................................ 67 5.1.2 Operating Conditions for Spray Drying ................................................................................................ 67

5.2 Results and Discussion ............................................................................................................................. 68 5.2.1 Yield from Spray Drying ....................................................................................................................... 68 5.2.2 Assessment of the Particle Crystallinity by Moisture Sorption Tests ................................................... 72 5.2.3 Further Assessment of the Particle Crystallinity by MDSC ................................................................. 74 5.2.4 X-Ray Diffraction (XRD) ...................................................................................................................... 77

5.3 Conclusions ................................................................................................................................................ 80

CHAPTER 6 A NEW TEMPLATING APPROACH FOR HIGH-POROSITY SPRAY-DRIED PARTICLE PRODUCTION ................................................................................... 82

6.1 Materials and Methods ............................................................................................................................ 83 6.1.1 Sample Preparation ................................................................................................................................ 83 6.1.2 Operating Conditions for Spray Drying ................................................................................................ 83

6.2 Results and Discussion ............................................................................................................................. 85 6.2.1 Yield from Spray Drying ....................................................................................................................... 85 6.2.2 Assessment of the Degree of Crystallinity for the Lactose in Spray-Dried Powders ........................... 90

6.3 Conclusions .............................................................................................................................................. 101

CHAPTER 7 FORMATION OF HIGH POROSITY MANNITOL PARTICLES BY A NEW TEMPLATING PROCESS ...................................................................................... 102

7.1 Materials and Methods .......................................................................................................................... 102 7.1.1 Sample Preparation .............................................................................................................................. 102 7.1.2 Operating Conditions for Spray Drying .............................................................................................. 103

7.2 Results and Discussion ........................................................................................................................... 103 7.2.1 Yield from Spray Drying ..................................................................................................................... 103 7.2.2 BET Specific Surface Area .................................................................................................................. 107 7.2.3 SEM Micrographs ................................................................................................................................ 108 7.2.4 Fourier Transform Infrared Spectroscopy (FTIR) ............................................................................... 111

7.3 Conclusions .............................................................................................................................................. 113

CHAPTER 8 HIGHLY POROUS EXCIPIENTS AS A SOLUTION FOR POOR SOLUBILITY AND POOR CONTENT UNIFORMITY IN SOLID DOSAGE FORMS 114

8.1 Materials and Methods .......................................................................................................................... 114

Page 11: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

ix | P a g e

8.1.1 Sample Preparation .............................................................................................................................. 114 8.1.2 Porous Mannitol Production and Drug Loading Process .................................................................... 114

8.2 Results and Discussion ........................................................................................................................... 115 8.2.1 Highly-Porous Mannitol Particles Using Sucrose as a Templating Agent ......................................... 115 8.2.2 Effect of Porous Mannitol on Drug Loading ....................................................................................... 118 8.2.3 Physical State of Loaded Drug ............................................................................................................ 123 8.2.4 Blend Uniformity ................................................................................................................................. 127 8.2.5 Morphological Characterization .......................................................................................................... 130 8.2.6 Dissolution Profiles .............................................................................................................................. 132 8.2.7 Tableting Properties of the Products .................................................................................................... 133

8.3 Conclusions .............................................................................................................................................. 135

CHAPTER 9 A NOVEL FORMULATION FOR SOLUBILITY AND CONTENT UNIFORMITY OF POORLY WATER-SOLUBLE DRUGS ......................................... 136

9.1 Porous Mannitol Production and Drug Loading Process .................................................................. 136

9.2 Results and Discussion ........................................................................................................................... 137 9.2.1 Analysis of Drug Loading .................................................................................................................... 137 9.2.2 Physical State of the Loaded Drugs ..................................................................................................... 138 9.2.3 Drug Content Uniformity ..................................................................................................................... 143 9.2.4 Microscopic Characterization .............................................................................................................. 144 9.2.5 Dissolution Profiles .............................................................................................................................. 146

9.3 Conclusions .............................................................................................................................................. 148

CHAPTER 10 CONCLUSIONS AND RECOMMENDATIONS .................................. 149

10.1 Significant Outcomes of This Study ..................................................................................................... 149 10.1.1 Crystallization in Fluidized-Bed Dryers/Crystallizer ..................................................................... 149 10.1.2 Effect of Acidity on Lactose Crystallinity ...................................................................................... 149 10.1.3 Producing Highly Crystalline Lactose with Significant Porosity and High Surface Area ............ 150 10.1.4 Producing Highly Porous Mannitol with High Surface Area ......................................................... 151 10.1.5 Effect of acidity on a new templating approach for high-porosity spray-dried particle production 152 10.1.6 Solubility and Content Uniformity Enhancement of Poorly Water-Soluble Drugs Using Highly-Porous Excipients .............................................................................................................................................. 154

10.2 Recommendations and Future Work ................................................................................................... 155

Page 12: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

x | P a g e

LIST OF FIGURES

Figure 2.1 Sticking and caking mechanism of amorphous lactose. .......................................... 9 Figure 2.2 Effect of humidity on increasing the crystallinity of lactose in “Humid Loop”. ... 15 Figure 2.3 The upper limit on the relative humidity at which partially- and fully-crystallized whey powders can be fluidized. In this figure, Vuataz [10] corresponds to the glass-transition curve found by Vuataz (2002) for milk-based powders. ......................................................... 16 Figure 2.4 The upper limit of fluidization as related to the relative humidity of process air, for different lactose base materials (Yazdanpanah and Langrish, 2011b). .............................. 17 Figure 2.5 The upper limit on the relative humidity for which different degrees of crystallinity in skim milk powder can be fluidized and the process pathways between different stages of crystallization-drying. A) Fluidization of conventional spray-dried powders in three stages, B) Fluidization of Humid Loop spray-dried powders in two stages (Yazdanpanah and Langrish, 2011a). ...................................................................................... 18 Figure 3.1 Schematic diagram of the BüchiB-290 spray dryer. .............................................. 32 Figure 3.2 Schematic diagram of the Pilot-Scale Fluidized Bed Dryer. ................................. 34 Figure 3.3 Pilot-Scale Fluidized Bed Dryer. ........................................................................... 35 Figure 4.1 Multi-stage processing scheme for the fluidized-bed dryer. ................................. 46 Figure 4.2 Results of the moisture sorption behaviour of the spray-dried whey powder compared with whey powder that was processed after being exposed to different conditions in the fluidized bed (sorption test at 25°C and 75% relative humidity) (Run 1). .................... 52 Figure 4.3 X-ray diffraction patterns for spray dried whey powder compared with FB processed whey powder (Run 3), 10%, and 20% crystalline whey powders. ......................... 55 Figure 4.4 Raman spectra of whey powder with varying degrees of lactose crystallinity. The intensity increases at 476 cm-1 with increasing lactose crystallinity. The band areas used for the calculations are shown in Figure 4.5. ................................................................................. 56 Figure 4.5 Ratio of the Raman band areas as a function of the amorphous lactose content for the physical mixtures. The correlation was based on the average values of the amorphous lactose content. ......................................................................................................................... 57 Figure 4.6 FTIR spectra of whey powders with varying degrees of lactose crystallinity. The absorbance increases at 875 cm-1 and 900 cm-1 with increasing lactose crystallinity. The region in the square was used to determine the areas underneath the peaks, as shown in Figure 4.7. ................................................................................................................................ 58 Figure 4.7 FTIR spectra of whey powders normalized at 928 cm-1, with varying degrees of lactose crystallinity in the region from 860 cm-1 to 930 cm-1. ................................................. 59 Figure 4.8 The peak heights at 875 cm-1, 900 cm-1 and 915 cm-1 as function of the amorphous lactose content. A second order polynomial trendline has been fitted through the data points. The data points are the averages of five measurements. .......................................................... 60 Figure 4.9 A) Surface of spray-dried whey powder (amorphous) produced in a conventional spray-drying process at a high magnification. B:D) Morphological structure of processed whey powder in the fluidized bed. ........................................................................................... 63 Figure 4.10 Whey powder after three days at ambient condition, A and B) Spray-dried powder before and after test, respectively (A hard cake was formed (B)). C and D) Processed powder before and after test, respectively (still free-flowing (D)) (Run4). ............................. 64 Figure 5.1 Effect of the lactic acid concentration on the yield of the spray-dried products. .. 69 Figure 5.2 Effect of the lactic acid concentration on the moisture contents of the spray-dried products. ................................................................................................................................... 69 Figure 5.3 Moisture sorption behaviour of the spray-dried products from lactose for different lactic acid concentrations (the sorption test environment was at a temperature of 25oC and 75% relative humidity). ........................................................................................................... 73

Page 13: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

xi | P a g e

Figure 5.4 Sorption peak heights of the spray-dried products from lactose solutions for different lactic acid concentrations. ......................................................................................... 74 Figure 5.5 Sorption peak heights of the spray-dried products from lactose-WPI solutions for different lactic acid concentrations. ......................................................................................... 74 Figure 5.6 Thermal transitions during MDSC scans of the spray-dried products from lactose solutions for different lactic acid concentrations. .................................................................... 76 Figure 5.7 Effect of lactic acid concentration on the glass-transition temperatures, the crystallization temperatures, and the latent energies of crystallization for the spray-dried products from lactose solutions. .............................................................................................. 77 Figure 5.8 X-ray diffraction patterns of the spray-dried products from lactose solutions for different lactic acid concentrations. ......................................................................................... 79 Figure 5.9 X-ray diffraction patterns of the spray-dried products from lactose-WPI solutions for different lactic acid concentrations. ................................................................................... 80 Figure 6.1 Schematic diagram of the experimental setup and an illustration of the templating process...................................................................................................................................... 84 Figure 6.2 Effect of the ascorbic acid concentration on the glass-transition temperatures, the crystallization temperatures, and the latent energies of crystallization for the spray-dried products from lactose solutions. .............................................................................................. 93 Figure 6.3 Scanning electron micrographs of spray-dried powders from lactose/0.2% (w/w) WPI solutions with 1% w/w ascorbic acid, before (A) and after ethanol washing (B:D). ...... 99 Figure 6.4 Scanning electron micrographs of spray-dried powders from lactose solutions with 1% w/w ascorbic acid, after ethanol washing. ............................................................... 100 Figure 7.1 Effect of the WPI concentration on the BET surface area and the yield of the spray-dried products for mannitol (10% w/w) with 2% w/w citric acid (error bars indicate standard deviations). .............................................................................................................. 108 Figure 7.2 A and B) Scanning electron micrographs of spray-dried pure mannitol powder. C and D) spray-dried powder from mannitol with 0.5% (w/w) WPI and 2% (w/w) citric acid solution after ethanol washing. .............................................................................................. 109 Figure 7.3 Spray-dried powder from mannitol with 1% (w/w) Gum Arabic and 2% (w/w) citric acid solution after ethanol washing. ............................................................................. 110 Figure 7.4 FT-IR spectra of the spray-dried powder from mannitol (10% w/w) - citric acid (2% w/w) - WPI/Gum Arabic solutions before and after ethanol washing. A) 3% (w/w) WPI. B) 1% (w/w) WPI. C) 0.5% (w/w) WPI. D) 1% (w/w) Gum Arabic. The absorbance peaks at 1730 cm-1 show the citric acid content. ................................................................................. 112 Figure 8.1 Schematic illustration of the of the adsorption process. ...................................... 115 Figure 8.2 The concept of “Stickiness Barrier” for spray drying of sugar-rich materials (Imtiaz-Ul-Islam and Langrish, 2009). .................................................................................. 117 Figure 8.3 Pore size distributions for drug-loaded sample 5 in different acetaminophen solutions compared with sample 5 before loading. ................................................................ 122 Figure 8.4 Pore size distributions for drug-loaded sample 3 in different acetaminophen solutions compared with sample 3 before loading. ................................................................ 123 Figure 8.5 DSC thermograms of sample 3 loaded in 0.01 M and 0.5 M acetaminophen solutions compared with a DSC thermogram for a physical mixture of porous mannitol with 22 w/w % acetaminophen powder having a particle size of less than 63 µm. ....................... 124 Figure 8.6 X-ray pattern for drug-loaded sample 3 in 0.5 M acetaminophen solution compared with the physical mixture of porous mannitol and acetaminophen powder. ......... 127 Figure 8.7 DSC traces of different size fractions for drug-loaded sample 3 in 0.5 M acetaminophen solutions. ....................................................................................................... 129 Figure 8.8 DSC traces of different size fractions for drug-loaded sample 3 in 0.01 M acetaminophen solutions. ....................................................................................................... 130

Page 14: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

xii | P a g e

Figure 8.9 Scanning electron micrographs of spray-dried powders from sample 3, before (A) and after ethanol washing, compared with (B) drug-loaded sample 3 with a drug concentration of 0.5 M (C and D). ......................................................................................... 131 Figure 8.10 Dissolution profiles of gelatin capsules filled with both the drug-loaded sample 3 in 0.5 M acetaminophen, and the physical mixture of non-porous mannitol and acetaminophen powder (error bars indicate standard deviations). ......................................... 133 Figure 8.11 Comparison of tablet thickness as a function of compression pressure for tablets from the drug-loaded sample 3 in 0.5 M acetaminophen solution compared with the physical mixture of non-porous mannitol and acetaminophen powder (error bars indicate standard deviations). ............................................................................................................................. 134 Figure 8.12 Crushing strength of tablets compressed at different pressures for the tablets from the drug-loaded sample 3 in 0.5 M acetaminophen solution compared with the physical mixture of non-porous mannitol and acetaminophen powder (error bars indicate standard deviations). ............................................................................................................................. 135 Figure 9.1 DSC thermograms of mannitol drug-loaded in 0.16 M nifedipine compared with a DSC thermogram of a physical mixture of porous mannitol with 9 w/w% nifedipine powder................................................................................................................................................. 139 Figure 9.2 DSC thermograms of mannitol drug-loaded in 0.16 M indomethacin compared with a DSC thermogram of a physical mixture of porous mannitol with 13 w/w% indomethacin powder. ............................................................................................................ 140 Figure 9.3 X-ray pattern for drug-loaded mannitol in 0.16 M nifedipine solution compared with the physical mixture of porous mannitol and nifedipine powder. ................................. 142 Figure 9.4 X-ray pattern for drug-loaded mannitol in 0.16 M indomethacin solution compared with the physical mixture of porous mannitol and indomethacin powder. ........... 143 Figure 9.5 Scanning electron micrographs of spray-dried mannitol, before (A) and after ethanol washing (B) compared with for drug-loaded mannitol in 0.16 M indomethacin (C and D) and nifedipine (E and F) solutions. ................................................................................... 145 Figure 9.6 Dissolution profiles of gelatin capsules filled with drug-loaded mannitol in 0.16 M and 0.08 M nifedipine and the physical mixture of non-porous mannitol powder (error bars indicate standard deviations). ................................................................................................. 147 Figure 9.7 Dissolution profiles of gelatin capsules filled with drug-loaded mannitol in 0.16 M and 0.08 M indomethacin solutions and the physical mixture of non-porous mannitol indomethacin powder (error bars indicate standard deviations). ........................................... 148

Page 15: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

xiii | P a g e

LIST OF TABLES

Table 2.1 The typical compositions of sweet and acid whey. ................................................. 19 Table 3.1 List of materials used in the experiments. ............................................................... 30 Table 4.1 Summary of the experimental results for whey powder processed in the fluidized-bed dryer. ................................................................................................................................. 50 Table 4.2 XRD intensity values and the estimated degree of crystallinity at chosen 2θ values for the whey powders used here. .............................................................................................. 54 Table 4.3 Summary of the trendline equations fitted through the data points of Figure 4.8. . 60 Table 5.1 Summary of the experimental conditions used in spray drying for lactose-lactic acid solutions (lactose 10% w/w), inlet gas temperature of 180oC and outlet gas temperature of 108 ± 3oC. ............................................................................................................................ 70 Table 5.2 Summary of the experimental conditions used in spray drying for lactose-WPI-lactic acid solutions (lactose 15% w/w) + (WPI 5% w/w), inlet gas temperature of 180oC and outlet gas temperature of 112 ± 4oC. ....................................................................................... 71 Table 6.1 Summary of the experimental conditions used in spray drying for lactose- lactic acid solutions (lactose 10% w/w), inlet gas temperature of 180ºC and outlet gas temperature of 108 ± 3ºC (The data were mainly presented earlier in Table 5.1). ...................................... 86 Table 6.2 Summary of the experimental conditions used in spray drying for lactose-citric acid solutions (lactose 10% w/w), inlet gas temperature of 180ºC and outlet gas temperature of 113 ± 2ºC. ............................................................................................................................ 87 Table 6.3 Summary of the experimental conditions used in spray drying for lactose-ascorbic acid solutions (lactose 10% w/w), inlet gas temperature of 180ºC and outlet gas temperature 112 ± 2ºC. ................................................................................................................................ 88 Table 6.4 Summary of the experimental conditions used in spray drying for lactose (10% w/w) –WPI (0.2% w/w) -ascorbic acid solutions, inlet gas temperature of 180ºC and outlet gas temperature 112 ± 3ºC. ...................................................................................................... 89 Table 6.5 Summary of the experimental conditions used in spray drying for lactose (10 % w/w) - boric acid solutions, inlet gas temperature of 180ºC and outlet gas temperature of 98 ± 4ºC. ........................................................................................................................................... 89 Table 7.1 Summary of the experimental conditions used in spray drying for mannitol- citric acid - WPI solutions (mannitol 10% w/w), inlet gas temperature of 150oC and outlet gas temperature of 90 ± 4oC. ........................................................................................................ 105 Table 7.2 Summary of the experimental conditions used in spray drying for mannitol (10% w/w) - citric acid (2% w/w)- WPI solutions (% w/w), inlet gas temperature of 150oC and outlet gas temperature of 90 ± 3oC. ....................................................................................... 106 Table 7.3 Summary of the experimental conditions used in spray drying for mannitol- citric acid –Gum Arabic solutions (mannitol 10% w/w), inlet gas temperature of 150oC and outlet gas temperature of 89 ± 3oC. .................................................................................................. 106 Table 8.1 Summary of the experimental conditions used in spray drying for mannitol (10% w/w) - sucrose - WPI solutions, inlet gas temperature of 150oC and outlet gas temperature of 80 ± 4oC. ................................................................................................................................ 118 Table 8.2 Summary of the experimental results regarding drug loading, BET surface area, and pore volume for different mannitol samples loaded with acetaminophen solutions. ...... 121 Table 8.3 Content uniformity of different size fractions for the drug-loaded sample 3 in 0.01 M and 0.5 M acetaminophen solutions. ................................................................................. 129 Table 9.1 Summary of the experimental results for drug loading with mannitol samples (BET= 6.3 ± 0. 1 m2 g-1 and pore volume of 0.033 ± 0.002 ml g-1) loaded in drug solutions................................................................................................................................................. 137

Page 16: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

xiv | P a g e

Nomenclature

API Active Pharmaceutical Ingredient BET Brunauer, Emmett and Teller BJH Barrett-Joyner-Halenda DIAL Dairy Innovation Australia Limited DSC Differential Scanning Calorimetry FB Fluidized Bed FDA Food and Drug Administration FTIR Fourier Transform Infrared spectroscopy LSD Least Significant Difference MDSC Modulated Differential Scanning Calorimetry RSD Relative Standard Deviation SGF Simulated Gastric Fluid STD Standard Deviation WLF Williams–Landel–Ferry WP Whey Powder WPI Whey Protein Isolate XRD X-Ray Diffraction

Page 17: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

xv | P a g e

LIST OF PUBLICATIONS

The following publications and conference papers were published from content of the work

presented in this thesis at the time of submission of this thesis.

From work described in Chapter 4:

Significant results from chapter 4 were presented in the 5th Annual Student Research

Conference: Implementation of Crystallization in Fluidized Bed Processing. Faculty of

Engineering and IT, The University of Sydney. October, 2013 (Poster presentation).

From work described in Chapter 5:

Saffari, M. & Langrish, T. (2014), Effect of lactic acid on in-process crystallization of

lactose/protein powders during spray drying. Journal of Food Engineering, 137, 88-94.

Saffari, M. & Langrish, T.A.G. (2014), Effect of acidity on crystallization of lactose powders

during spray drying, 19th International Drying Symposium. Lyon, France (eds J. Andrieu, R.

Peczalski and S. Vessot), August 24-27, paper 130019 (Oral presentation).

From work described in Chapter 6:

Saffari, M., Ebrahimi, A. & Langrish, T.A.G. (2015), The Role of Acidity in Crystallization

of Lactose and Templating Approach for Highly-Porous Lactose Production, Journal of Food

Engineering, 164, 1-9.

From work described in Chapter 7:

Saffari, M., Ebrahimi, A. & Langrish, T.A.G. (2015), Highly-porous mannitol particle

production using a new templating approach, Food Research International, 67, 44-51.

Page 18: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

xvi | P a g e

From work described in Chapter 8:

Saffari, M., Ebrahimi, A. & Langrish, T.A.G. (2016), Nano-confinement of acetaminophen

into porous mannitol through adsorption method, Microporous & Mesoporous Materials,

227, 95-103.

From work described in Chapter 9:

Saffari, M., Ebrahimi, A. & Langrish, T.A.G. (2016), A novel formulation for solubility and

content uniformity enhancement of poorly water-soluble drugs using highly-porous mannitol,

European Journal of Pharmaceutical Sciences, 3, 52–61.

Page 19: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

1 | P a g e

CHAPTER 1 INTRODUCTION

1.1 Spray Drying

Spray drying is a low-cost and high-throughput continuous drying technique. It is among a

wide variety of drying operations that are applied to produce stable low moisture-content

products, which reduce costs for storage and transportation from their liquid or slurry states

in the food and pharmaceutical industries (Morgan and Vesey, 2009; Sollohub and Cal,

2010). The rapid removal of water from solutions by this technique, during drying, does not

allow sufficient time for full crystallization to occur for most materials and causes the

formation of amorphous, thermodynamically-unstable solids (Roos and Karel, 1991; Hogan

and O'Callaghan, 2010). This situation may lead to an increased rate of physicochemical

changes in dried products, such as sticking, collapse, caking, agglomeration and

crystallization (Bhandari and Howes, 1999). The stickiness phenomenon in spray drying, and

the caking phenomenon in storage, are associated with the presence of high sugar and food

acid concentrations in the product (Bhandari et al., 1997b), which can cause problems for

food powders and related substances.

1.2 Enhancements in Spray-Dried Powder Functionality

Processing of spray-dried powder into marketable attractive final products faces major

economical obstacles. The main constituents of dairy powders are carbohydrates, proteins,

water, and fat. Among the different ingredients of milk powders, amorphous lactose is one of

the most hygroscopic and unstable materials. Different operating conditions can be used to

obtain different extents of crystallinity in various processes to reduce the hygroscopic nature

of lactose in dairy powders by converting a portion of the lactose to the more stable

crystalline components, such as pre-crystallization and ‘‘in-process” crystallization steps

before and within spray drying, respectively and post process crystallization in fluidized-bed

drying, which is a controlled process to transform amorphous-lactose fraction to a crystalline

form rapidly as compared with the rate in storage (Bhandari et al., 1997b; Nijdam et al.,

2008; Hogan and O'Callaghan, 2010; Islam et al., 2010; Yazdanpanah and Langrish, 2011a).

Page 20: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

2 | P a g e

Due to this controlled process, more stable powder products can be produced, but crystalline

powders may have some undesirable properties, such as lower dissolution rates.

The creation of highly-porous crystalline powder with better functional (free flowing), and

storage (non-caking) properties would be helpful in the food industry and pharmaceutical

applications. Thus, a new spray-drying process has been developed with post-drying

treatment to create powder structures with significant porosities and high surface areas, while

maintaining the high stability of a highly crystalline powder.

1.3 Application of New Engineered Particles with High Porosities and

Controlled Characteristics

Pharmaceutical dosage forms that include tablets and capsules are some of the most popular

and commonly-employed methods of drug delivery, which currently account for over two-

thirds of the total number of medicines produced in the world (Zheng, 2009; Sahoo, 2012).

Their final formulation contains both active pharmaceutical ingredient (API) and excipients

added to aid the formulation and manufacture of the subsequent dosage form for

administration to patients (Rowe et al., 2003). Different types of sugars such as lactose,

mannitol, sorbitol, and dextrin and inorganic compounds, such as silica and calcium

carbonate are commonly used as excipients to provide inert or inactive ingredients (Rowe et

al., 2003; Millqvist-Fureby et al., 2014). However, this reason for using excipients has

undergone changes due to the needs for controlled properties of the final dosage form such as

bioavailability and stability and growth of novel forms of delivery (Rowe et al., 2009).

For instance, major challenges facing the design of oral dosage forms are their poor

bioavailability and uniformity of drug dosage (Garcia and Prescott, 2008; Huang and Sherry

Ku, 2010; Grigorov et al., 2013) and poor solubility is one of the main causes of low

bioavailability (Schreiner et al., 2005; Savjani et al., 2012). More than 40% of newly

developed drugs in the pharmaceutical industry are poorly water-soluble, which may cause a

range of medication control problems, such as insufficient dosing (Liu, 2008; Jain et al.,

2012). Formulation and dosage-form design problems of poorly water soluble drugs must be

addressed to better meet the needs of users and the targeted product requirements, such as

clinically acceptable performance.

Page 21: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

3 | P a g e

One potential approach to solve the abovementioned challenges for better control of the drug

release characteristics is loading or nano-confinement of drugs into porous excipients as hosts

(Salonen et al., 2005a; Qu et al., 2006; Xia and Chang, 2006; Prestidge et al., 2007).

Mesoporous silica is of specific interest as excipients for these drug-loading methods due to

its high specific area and large pore volume (Millqvist-Fureby et al., 2014).

Mesoporous silica is expensive and has low water solubility (Millqvist-Fureby et al., 2014)

and therefore other porous excipients, such as mannitol with its high water solubility (216 g/l;

Ohrem et al., 2014), may be potential carriers for controlled-release systems of poorly water-

soluble drugs despite the lower specific area and smaller pore volume compared with those of

mesoporous silica. It can be expected that preparing an excipient with higher surface areas

and larger pore volumes, such as mesoporous silica, would result in a higher drug loading, as

it provides more available pore volume in order to host the drug molecules. However, high

drug loadings are not always needed and, in the formulation and manufacture of low-dose

drug products, the amount of API in the solid-dosage form can be as low as 0.1 w/w %

(Zheng, 2009; Grigorov et al., 2013). Moreover, major challenges facing the design of the

oral dosage form, such as poor bioavailability and uniformity of drug dosage, are more

pronounced for low-dose drug products, which may result in undesirable variations in dosage

and (Cartilier and Moes, 1989; Garcia and Prescott, 2008; Zheng, 2009; Huang and Sherry

Ku, 2010). Therefore, providing porous carriers with enough specific surface area and

reasonable pore volumes for the targeted drug-loadings seems to be desirable. However, most

current techniques involve producing porous inorganic particles through templating process

and the key challenge is to extend this understanding of the process from inorganic materials

to organic ones.

There are several drug loading methods, such as incipient wetness impregnation (Mellaerts et

al., 2008; Van Speybroeck et al., 2009; Verraedt et al., 2010) and adsorption methods, in

which carrier particles are immersed in drug solution and drugs can be dispersed and

deposited at the surface and into the pore spaces of porous carriers (Qu et al., 2006; Xia and

Chang, 2006; Heikkilä et al., 2007; Hillerström et al., 2009).

This improvement can be ascribed to several factors. First, dissolution rate improvement due

to nanoconfinement of the drug molecules inside the internal void structures of nanoporous

excipients. This situation results in an increase in the solubility with a reduction in the

particle size of the drug, which maximizes the surface area of the compound that comes into

Page 22: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

4 | P a g e

contact with the dissolution medium as the carrier dissolves, often resulting in significant

increases in bioavailability. Second, excipients with high porosity can improve the drug

release rate into the dissolution medium by enabling faster release of the drug as a result of

better drug-excipient matrix permeability and penetration of solvent into the drug-excipient

matrix. Moreover, greater porosity of excipients facilities the diffusion of the API

components and increases the rate of drug adsorption. Last but not least, the adsorption

method is independent of the control strategy for both the particle size of the drug and

excipients in the finished dosage units, so the particle size of the final formulation does not

appear to have any effect on drug uniformity. The variability for the content uniformity and

delivered dose may then be low. Therefore, experimental design is necessary to create new

engineered frameworks with high porosity and controlled characteristics. These drug-loading

processes are gaining acceptance, as shown by the rapid expansion of this field in

pharmaceutical formulations, especially for controlled delivery and release applications

(Salonen et al., 2005a; Xia and Chang, 2006; Heikkilä et al., 2007; Mellaerts et al., 2007;

Prestidge et al., 2007; Mellaerts et al., 2008; Hillerström et al., 2009; Van Speybroeck et al.,

2009; Millqvist-Fureby et al., 2014).

1.4 Thesis Structure

Chapter 2 provides some background to the solid-state crystallization process for the spray-

dried form of dairy powders and also considers a controlled process using fluidized-bed

dryers for crystallization of dairy powders. This chapter also reviews the previous works on

templating techniques that have been applied to improve the porosity of spray-dried particles.

Chapter 3 reviews the applicable features regarding the characterization of products, such as

the particle size and shape, the moisture content, the glass-transition behavior and

amorphous/crystalline states, and other known issues concerning templating and drug-loading

processes, such as BET surface area analysis, tablet crushing tests, UV-spectrophotometry,

and in-vitro dissolution studies.

The effects of operating conditions on the crystallization properties of sweet whey powder in

a continuous multi-stage fluidized bed dryer to achieve the maximum degree of crystallinity

with reasonable overall processing times are described in Chapter 4. Powder was fed at

Page 23: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

5 | P a g e

different rates for each fluidization process at various sets of process temperatures,

humidities, and residence times. Gravimetric moisture sorption tests were performed on the

processed powders and compared with X-ray Powder Diffraction, FTIR, and Raman

spectroscopy analyses. The results show significant improvements in the crystallinity of the

lactose. Experimental results showed that there were significant changes on the degree of

lactose crystallinity between samples that were processed (crystallized) at different

humidities and temperatures. Increasing the rate of powder feeding did not have a significant

effect on the degrees of lactose crystallinity for the processed powders. The results of this

study are useful for guiding the processing conditions to produce and control crystallization

in carbohydrate-containing dairy powders during spray and fluidized-bed processing.

In Chapter 5, the effect of acidity on the final product properties from the spray drying of

lactose solutions has been investigated. Moreover, the crystallization kinetics of

lactose/protein solutions have been studied regarding their in-process crystallization

characteristics during spray drying, especially when lactic acid is added to the solution to

imitate the composition of a typical acid whey solution. Gravimetric moisture sorption tests

have been performed on the spray-dried powders processed under these conditions and

compared with X-ray powder diffraction and modulated differential scanning calorimetry

(MDSC) analyses to measure the degree of lactose crystallinity. It has been found that very

large changes in the degrees of lactose crystallinity for the final spray-dried product have

occurred when increasing the lactic acid concentration. The yields (or solids recoveries) from

spray drying have also been significantly decreased at higher concentrations of lactic acid.

The results of this study have implications in choosing the processing conditions to produce

and control crystallization in carbohydrate-containing dairy powders and also powders of

acid-rich foods, such as fruit juices, during spray and fluidized-bed processing.

A new production process has been designed to produce highly-porous powder, and this

process has been discussed in Chapters 6 and 7. This new templating process has been

successfully developed to create highly-porous lactose particles with high surface areas

through the spray drying of lactose solutions containing ethanol-soluble food-grade acids,

such as lactic acid, citric acid, boric acid, and ascorbic acid, as templating agents, and then

removing these acids by ethanol washing of the spray-dried powders. Chapter 6 also reports

the effect of using different concentrations for various food-grade acids, such as lactic acid,

citric acid, boric acid, and ascorbic acid, on the degree of crystallinity for spray-dried lactose

Page 24: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

6 | P a g e

powders and the porosity of the ethanol-washed powders. One innovative aspect of this

approach, which uses the insolubility of lactose in ethanol and the solubility of templating

acids in ethanol as a basis, is the potential to create highly-porous templates or frameworks of

lactose using relatively benign and non-toxic materials, such as food-grade acids. Another

novel contribution of this work is the use of the low-cost and high-throughput spray-drying

process, which is an improvement on the batch and semi-batch processes described so far.

Recently, the use of mannitol as an alternative to lactose in food products and pharmaceutical

formulations has significantly increased. Lactose is a reducing sugar, which is incompatible

with some active pharmaceutical ingredients, such as peptides and proteins, and also has

problems associated with lactose intolerance (Westermarck et al., 1998; Steckel and Bolzen,

2004; Littringer et al., 2012). Therefore, in Chapter 7, this new templating method has been

developed to create highly-porous mannitol through spray drying, using citric acid as the

templating agent. It has been demonstrated in Chapter 7 that textural properties, such as

surface area and pore volume of the resultant mannitol, can be tuned by varying the

concentrations of citric acid, whey protein isolate (WPI), and Gum Arabic.

A new adsorption method for solubility and content uniformity enhancements of poorly

water-soluble drugs was developed in Chapter 8. Highly-porous excipients with high surface

areas that were produced according to the templating method in Chapter 7 have been used in

an adsorption method to investigate drug dissolution. This work has been done by applying

an adsorption technique and modifying the current templating process described in Chapter 7.

Highly-porous mannitol particles with high surface areas have been successfully produced

through the spray drying of mannitol solutions containing ethanol-soluble food-grade acids,

such as citric acid as a templating agent, and then removing the citric acid by ethanol washing

of the spray-dried powders to create porous powders. Production of highly-porous

frameworks of mannitol having unique properties, such as a significant surface area of 9.1 ±

0.9 m2 g-1, and a total pore volume of 0.11± 0.03 ml g-1, which can be tuned by varying the

concentrations of citric acid and WPI. These features make this material suitable as a carrier

for adsorption method. However, this chapter has also investigated the possibility of using

carbohydrate sugars, such as sucrose, as non-acidic templates that are commonly used as

excipients in drug delivery for pharmaceutical applications. After ethanol washing, the

Page 25: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

7 | P a g e

resultant porous particles have been transferred to ethanol solutions to dissolve and carry

acetaminophen as a sample drug into the pores of porous mannitol.

Page 26: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

8 | P a g e

CHAPTER 2 LITERATURE REVIEW

2.1 Amorphous and Crystalline Powders

The rapid removal of water from solutions in spray drying does not allow sufficient time for

crystallization to occur for most materials and causes the formation of amorphous,

thermodynamically-unstable solids (Roos and Karel, 1991; Hogan and O'Callaghan, 2010).

Above the glass-transition temperature, an amorphous solid exists in a ''rubbery'' state, and

the molecular mobility of the matrix and the reactants are increased. This situation may lead

to an increased rate of physicochemical changes in dried products, such as sticking, collapse,

caking, agglomeration and crystallization (Bhandari and Howes, 1999), which are time,

temperature and moisture-dependent phenomena. The stickiness phenomenon in spray

drying, and the caking phenomenon in storage, are associated with the presence of high

concentrations of sugar in the product (Bhandari et al., 1997b).

2.2 Sticking and Caking Mechanism of Amorphous Lactose

Spray-dried milk powders, which are normally in amorphous states, have several advantages

compared with crystalline powders, such as greater dissolution rates and surface area

(Hancock and Parks, 2000; Salonen et al., 2005b; Babu and Nangia, 2011). However, they

are thermodynamically-unstable solids (Roos and Karel, 1991; Hogan et al., 2009; Hogan and

O'Callaghan, 2010), which can decrease the physical and chemical stability of the products

(Buckton and Darcy, 1995, 1996). They have a tendency to sorb moisture and form caked

powders that are not free flowing.

Stickiness and caking tendencies of hygroscopic, amorphous powders can be problems for

food powders and related substances. Stickiness has been defined as the situation when

particles stick to each other (cohesion) or to other surfaces (adhesion). A symptom is that

particles stick to one another or to the walls of the drying apparatus (Downton et al., 1982;

Wallack and King, 1988; Paterson et al., 2005). When the surface temperature of a sugar-

based powder particle exceeds Tg, a decrease in viscosity of the amorphous material from a

‘glassy’ to a ‘rubbery’ state allows the formation of liquid bridges between particles in

contact with each other or with equipment surfaces (Hogan and O'Callaghan, 2010). These

Page 27: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

9 | P a g e

liquid bridges may be of molten fat, amorphous sugar or a concentrated sugar solution

material. This situation can lead to various issues, such as lower product yield or recoveries,

and powder-handling difficulties and caking during storage. Over time, these liquid bridges

can crystallize, leading to the irreversible consolidation of the liquid bridges. Caking is

defined as the stage where crystallization has occurred and hence solid bridges form (Foster

et al., 2004, 2006). Caking may also occur as a result of recrystallization, either after fat

melting or after solubilization at crystal surfaces; surface wetting followed by moisture

equilibration or cooling; or electrostatic attraction between particles (Aguilera et al., 1995)

(Figure 2.1).

Figure 2.1 Sticking and caking mechanism of amorphous lactose.

Paterson et al. (2005) demonstrated that the sticking behaviour of amorphous lactose can be

characterised by the magnitude of the difference between the actual particle temperature and

the glass-transition temperature of the particle (T- Tg), regardless of which combination of

water activity and temperature are used to achieve this critical T- Tg level. They stated that

amorphous lactose, at the same value of T- Tg but at different temperatures and water

activities, showed similar relationships between the levels of cohesiveness for the powders

with time, and amorphous lactose became stickier much faster at higher values of T- Tg.

Crystallization

Caking

Sticking

High Humidity/ Temperature conditions

T≥ Glass-transition temperature (Tg)

Amorphous lactose

Rubber bridge

Solid bridge

Page 28: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

10 | P a g e

Above the glass-transition temperature, the molecular mobility increases, as shown by a

decrease in viscosity (Burnett et al., 2004). Water (moisture content) acts as a plasticizing, or

softening agent, in amorphous food materials, significantly lowering the overall glass-

transition temperature (Roos, 1995). The extent of Tg depression depends on the

concentration of the plasticizer and its interaction with the amorphous material (Roos, 1995).

Above the glass-transition point, amorphous materials of low molecular weight will

crystallize after a given time, and the water sorption capacity decreases, resulting in mass loss

since excess water is desorbed during crystallization (Roos and Karel, 1991; Burnett et al.,

2004).

The glass-transition temperature is affected by various factors, of which the composition of

the material, molecular weight and water (moisture) content are the most important. For the

calculation of the glass-transition temperature of the mixture, Gordon and Taylor (1952)

originally proposed an equation for binary polymer mixtures. Later, Roos (1993) found that

the Gordon–Taylor Equation ( 2.1) can be used to estimate the glass-transition temperature

(Tg) of carbohydrates at various water (moisture) contents:

2.1

where w1 and w2 are the weight fractions of the solute and water, respectively, Tg1 is the

glass-transition temperature of the solute, Tg2 is the glass-transition temperature of water (K),

and k is a curvature constant, which can be determined empirically. The weight fraction of

water, w2, is related to the moisture content, X, expressed on a dry basis, through the equation

w2 X⁄ (1+X).

2.3 Crystallization Kinetics

Crystallization from amorphous materials in the solid state occurs when molecules of the

amorphous solids rearrange themselves into an orderly structure (Jouppila and Roos, 1994a).

The transformation from solid amorphous to solid crystalline products has been investigated

by several researchers. They suggested that the crystallization rate of various amorphous

sugars can be estimated by using one of the following approaches: the Williams–Landel–

Ferry (WLF) equation (Williams et al., 1955), the Avrami equation or the activated-state

model (Das and Langrish, 2012). The difference between the glass-transition temperature (Tg)

Page 29: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

11 | P a g e

of the materials and the process temperature (T) has been widely accepted as a best indicator

to avoid this stickiness during drying (Bhandari et al., 1997b; Islam et al., 2010). According

to the Williams–Landel–Ferry (WLF) equation, a higher particle temperature and lower

particle glass-transition temperature increase the crystallization rate of the particles during the

spray-drying process.

2.3.1 Modelling the Kinetics of Solid-Phase Crystallization with the Williams–Landel–

Ferry Equation

The ratio (r) of the time for crystallization (θcr) at any temperature (T) to the time for

crystallization (θg) at the glass-transition temperature (Tg) can be correlated by the Williams–

Landel–Ferry (WLF) equation, Equation 2.2 (Williams et al., 1955). The WLF equation can

be expressed as a rate equation (Equation 2.3) for the crystallization process, and the

crystallization rate may be assumed to be inversely proportional to the crystallization time:

17.44

51.6

2.2

10^17.44

51.6

2.3

where, θcr is the time for crystallization (s), θg is the time for crystallization at the glass-

transition temperature (s), kcr is the rate of crystallization (s-1) at the particular local

conditions (T-Tg), and kg is the rate of crystallization at the glass-transition temperature (Tg).

The glass-transition temperature can be estimated from the Gordon–Taylor equation (1952).

The WLF equation suggests that the rate of crystallization is related to the difference between

the material temperature (T) and its glass-transition temperature (T-Tg), where the glass-

transition temperature depends on the moisture contents. The WLF equation is only valid for

temperatures up to 100ºC greater than the glass-transition temperature (Williams et al., 1955),

and the validity of the universal constants (17.44 and 51.6) in this equation has been debated

(Roos and Karel, 1991; Peleg, 1992). Roos and Karel (1991) found that these fixed constants

can be applied for the crystallization times of sucrose and lactose. However, Peleg (1992)

suggested that these constants depend on the substance investigated and the difference

Page 30: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

12 | P a g e

between T and Tg, particularly 20-30 K above the Tg. These constants are most likely to be

product specific (Langrish, 2008).

Another limitation of the WLF approach was reported by Roos and Karel (1992). They

showed that the kinetics for the crystallization process have been expressed in an integrated

form rather than as a function of reactant concentration. Therefore, this approach makes the

reaction appear to be a zero-order reaction regarding the concentration. Das and Langrish

(2012) reported that the WLF equation was inconsistent in predicting the rates of

crystallization at the glass-transition temperature and found that, since the WLF equation is a

temperature-shift equation, it does not take into account the fundamental effect of moisture

content on the crystallization rate. Therefore, with the above limitations, the WLF equation is

unlikely to predict high-temperature crystallization accurately in a quantitative manner.

2.3.2 Modelling the Kinetics of Solid-Phase Crystallization with the Avrami Equation

Another approach to modeling solid-phase crystallization is the Avrami equation (Avrami,

1940), which gives the degree of crystallinity as a function of time and not of reactant

concentration. Ibach and Kind (2007) determined the Avrami exponent (reaction order) over

a range of temperatures and humidities for lactose, whey, and whey-permeate powders.

The Avrami equation can be written in the form of Equation 2.4 to calculate the degree of

crystallinity from the XRD data, based on the extent of completion compared with

the total change in XRD peak intensity.

1 1 2.4

where θ is the degree of crystallinity, t is time, k is the rate constant, n is the Avrami

exponent, If is the maximum value of the intensity of the peaks or the maximum value of the

peak areas occurring at the maximum extent of sample crystallization, It is the intensity of the

peaks or the peak areas at time t, and I0 is the intensity of the peak or the peak areas for a

noncrystalline, amorphous sample. This approach can also be used with the moisture content

instead of the peak intensity (XRD), for water/ moisture sorption data (Langrish, 2008).

The Avrami equation can be rearranged as follows when fitting experimental data:

Page 31: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

13 | P a g e

1 2.5

A plot of 1 versus gives a straight line with a slope of n and an intercept of

. Values of k and n vary for different materials and conditions.

The applicability of the Avrami equation has been debated. It has been used as a correlation

for fitting data from isothermal crystallization experiments, and the equation does not

explicitly account for the effect of temperature (Langrish, 2008). Roos and Karel (1992)

tested the applicability of the Avrami equation by measuring the crystallization rate for

amorphous lactose using water-induced crystallization (crystallization at constant relative

humidity) and found that the equation did not fit the results well.

The physical process of crystallization consists of two main steps, which are nucleation and

crystal growth. Both the WLF and Avrami equations do not distinguish between the two

processes and instead lump them together (Das and Langrish, 2012). Langrish (2008) also

suggested that, due to the parameters in the Avrami equation being less universal compared

with those in the Williams–Landel–Ferry equation, the approach using the Avrami equation

currently appears to have less generality than using the Williams–Landel–Ferry equation.

2.4 The Controlled Crystallization Processes

The main constituents of dairy powders are carbohydrates, proteins, water, and fat. Among

the different ingredients of milk powders, amorphous lactose is a very hygroscopic and

unstable material. Stickiness and caking tendencies of hygroscopic, amorphous powders can

be problems for food powders and related substances. Stickiness occurs when the surface

temperature of the powder particle exceeds Tg, and a decrease in the viscosity of the

amorphous material from a ‘glassy’ to a ‘rubbery’ state allows the formation of liquid bridges

between particles that are in contact with each other or with equipment surfaces (Hogan and

O'Callaghan, 2010). Reducing stickiness in materials can be achieved through partial or

complete crystallization of the sticky components.

Due to the hygroscopic nature of amorphous lactose, lactose-containing materials become

sticky when exposed to humid environments. Different operating conditions can be used to

obtain different extents of lactose crystallinity in various processes to reduce the hygroscopic

Page 32: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

14 | P a g e

nature of lactose in dairy powder by converting a portion of the lactose to the more stable

crystalline state. Examples include pre-crystallization and ‘‘in-process” crystallization steps

before and within spray drying, respectively, and post process crystallization in fluidized-bed

drying. The pre-crystallization process involves seeding a supersaturated solution with fine

lactose crystals to nucleate crystallization (Nijdam et al., 2008). The difference between the

glass-transition temperature (Tg) of the materials and the process temperature (T) has been

widely accepted as a good indicator to avoid this stickiness and to give specific advantages in

various processes, such as spray drying (Bhandari et al., 1997b; Islam et al., 2010) and post

processing in fluidized-bed units (Nijdam et al., 2008; Hogan and O'Callaghan, 2010;

Yazdanpanah and Langrish, 2011a).

2.4.1 Modified Spray-Drying Technique

Islam et al. (2010) proposed a modified spray-drying technique called “Humid Loop” to

spray dry lactose, which uses highly humid hot air to obtain highly crystalline lactose from

spray dryers. They suggested that, according to Williams–Landel–Ferry kinetics (WLF), a

higher particle temperature and lower glass-transition temperature should increase the

crystallization rate of the particles during the spray-drying process while still giving dry

powders (Figure 2.2). The results of their work showed that they could reach higher degrees

of lactose crystallinity compared with typical spray-dried lactose particles. Although the

effects of other components, such as proteins, on the crystallization rate, kinetics, crystal

shapes, and distributions have not been fully explained yet, this technique could also possibly

apply to other lactose-containing powders (Yazdanpanah and Langrish, 2011a).

Page 33: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

15 | P a g e

Figure 2.2 Effect of humidity on increasing the crystallinity of lactose in “Humid Loop”.

2.4.2 Crystallization in Fluidised-Bed Dryers/Crystallizers

Fluidized beds have found widespread applications for drying, mixing, granulation, coating,

heating and cooling due to their excellent mixing capabilities, which cause good heat transfer,

temperature uniformity and ease of process control (Rhodes, 2007). A very important

application of the fluidized bed is to the drying of solids. Fluidized-bed dryers can be used to

improve dairy powder crystallinity in food-processing industries (Cruz et al., 2005; Nijdam et

al., 2008; Yazdanpanah and Langrish, 2011a; 2011b; 2011c).

Nijdam et al. (2008) investigated the fluidization of partially- and fully-crystallized whey

powders at different relative humidities and air temperatures (Figure 2.3). They stated that

partially-crystallized whey powder can be fluidized in a vibrated bed at temperatures of 25ºC

to 40ºC above the glass-transition point of lactose, depending on the relative humidity of the

air, before the powder becomes too sticky to fluidize, while this temperature difference can

be increased up to 80ºC by fluidizing the fully-crystallized whey powder. They suggested

that, for partially-crystallized whey powder, the relative humidity can be increased by a

further 15% above the glass-transition curve found by Vuataz (2002). Fully-crystallized whey

powder can be fluidized up to a relative humidity of approximately two times that of

partially-crystallized whey powder. They also tried to fluidize a mixture of partially- and

fully-crystallized whey powders, with the aim of coating the sticky partially-crystallized

whey particles by non-sticky crystallized whey particles. They also assessed whether or not

an industrial whey-powder crystallization process is achievable, in which a portion of the

Higher humidity

Higher moisture contents

Decrease Tg

Increase

(T-Tg)

Enhancing the crystallization

rate

WLF equation

Page 34: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

16 | P a g e

whey powder already fully crystallized in a fluidized bed is recycled and mixed with fresh

partially-crystallized whey powder being fed into the fluidized bed. The result of this process

showed the highly agglomerated whey particles were formed during fluidization. In addition,

they could not reduce the time required to fully crystallize the amorphous-lactose fraction in

partially-crystallized whey powder sufficiently to be feasible in an industrial crystallization

process.

Figure 2.3 The upper limit on the relative humidity at which partially- and fully-crystallized whey

powders can be fluidized. In this figure, Vuataz [10] corresponds to the glass-transition curve found

by Vuataz (2002) for milk-based powders.

Hogan and O'Callaghan (2010) determined the relative stickiness behaviour for a range of

powders with different lactose/protein ratios, as a function of temperature and relative

humidity, using a fluidized bed. They found that increasing the proportion of protein

decreased the susceptibility of powders to sticking due to the combined influence of both Tg

and T- Tg. Their work also showed that the rate of transition of amorphous lactose from the

free-flowing, though the thermodynamically unstable, ‘glassy’ form, to the unstable,

‘rubbery’ (sticky) condition and subsequently to the stable crystalline state, was delayed by

proteins, due possibly to a competitive sorption mechanism.

Yazdanpanah and Langrish (2011a;c) advanced the above knowledge of stickiness by using

fluidized-bed dryers to crystallize lactose and milk powders with multiple stages of

processing at different temperatures and humidities to achieve different degrees of powder

crystallinity. They showed that the upper limit of fluidization depends on the percentage of

 

Page 35: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

17 | P a g e

crystallinity in the powders. The worst fluidization corresponds to the highest amorphicity in

the powders, and fully crystalline milk powder has very smooth fluidization at high

temperatures and humidities (Figure 2.4) (Yazdanpanah and Langrish, 2011c). Therefore in

their later study (Yazdanpanah and Langrish, 2011a), they showed that different fluidization

abilities may occur for different amounts of crystallinity in the powders. Hence in the

fluidization-crystallization process, the humidity and temperature can be increased as a

function of time while the crystallinity of powders is developing. In other words, the

processing time was suggested to be split into different modules to maximize the

crystallization rate by maintaining an adequately high differential temperature (T–Tg) in

different stages corresponding to different amounts of amorphicity in the powders (Figure

2.5). They found that, with proper spray-drying conditions followed by similar strategic

processes in fluidized-bed dryers, the overall processing time to reach the maximum degree

of lactose crystallinity could be decreased.

Figure 2.4 The upper limit of fluidization as related to the relative humidity of process air, for

different lactose base materials (Yazdanpanah and Langrish, 2011b).

Page 36: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

18 | P a g e

Figure 2.5 The upper limit on the relative humidity for which different degrees of crystallinity in

skim milk powder can be fluidized and the process pathways between different stages of

crystallization-drying. A) Fluidization of conventional spray-dried powders in three stages, B)

Fluidization of Humid Loop spray-dried powders in two stages (Yazdanpanah and Langrish, 2011a).

Processed milk powders show less moisture sorption and more lactose crystals, leading to an

improvement in the degree of amorphicity for spray-dried milk powders by post processing in

a series of laboratory-scale fluidized bed dryers.

Chapter 4 involves using continuous multistage fluidized beds, with the aim of reducing the

time required to crystallize the amorphous-lactose fraction sufficiently in sweet whey to be

feasible in an industrial-crystallization process, where each stage will have particular

temperatures and relative humidities. Based on the above findings, it is expected that the

process conditions (temperature and relative humidity) for each stage will depend on the feed

powder properties (glass-transition temperature, composition, degree of amorphicity) in terms

of the carbohydrate content for that stage. Multi-stage fluidized bed processing has been

developed to achieve greater control of lactose crystallinity than the control achievable

through operating a single- stage fluidized-bed dryer on its own.

2.5 Types and Composition of Whey

The main constituents of dairy powders are carbohydrates, proteins, water, and fat. Whey is a

multicomponent solution of various water-soluble milk constituents in water, which is the by-

 

Page 37: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

19 | P a g e

product remaining after the production of cheese or the removal of fat and casein (80% of the

proteins) from milk (De Wit, 2001; Jelen, 2009). There are two basic types of whey: sweet

and acid whey, where the acid whey is usually achieved by acidification to below pH 5.0

Table 2.1. While the main components of both sweet and acid wheys, after water, are lactose

(approximately 70–72% of the total solids), whey proteins (approximately 8–10%) and

minerals (approximately 12–15%), the main difference is found in the calcium and lactic acid

contents (Panesar et al., 2007; Jelen, 2009).

Table 2.1 The typical compositions of sweet and acid whey.

Components Sweet whey (g l-1) Acid whey (g l-1)

Total solids 63-70 63-70 Lactose 46-52 44-46 Protein 6-10 6-8 Calcium 0.4-0.6 1.2-1.6

Phosphate 1-3 2-4.5 Lactate 2 6.4

Chloride 1.1 1.1

(Source: Jelen, 2009; Panesar et al., 2007)

2.6 Effect of Acidity on Lactose Crystallinity

Many milk plants do not have a fully-developed strategy to recycle whey, and around half the

global production of whey is treated as a waste product (Salameh and Taylor, 2006). Due to

the presence of economically-important components, such as lactose and the whey proteins,

the disposal of whey solutions to drain is no longer acceptable. Whey and whey products, as

functional ingredients in food and pharmaceutical applications, and as nutrients in dietetic

and health foods, have received considerable attention (De Wit, 2001). Different types of

whey products contain some of the most valuable milk nutrients, but the processing of whey

into marketable and attractive final products faces major economical obstacles (Jelen, 2009).

It has been found that the spray drying of acid whey is not an easy task because powder sticks

to the dryer and cyclone walls due to the high content of lactic acid and the low pH (Modler

and Emmons, 1978; Salameh and Taylor, 2006; Bhandari, 2008). If the drying conditions and

the degree of lactose crystallization are similar, then differences in drying properties can be

linked to the lactic acid-lactate content of these products (Modler and Emmons, 1978).

Therefore, the rationale of this study has been to investigate the effects of lactic acid

Page 38: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

20 | P a g e

concentration on the degrees of powder crystallinity during the spray drying of lactose

solutions with various acid concentrations.

The main ingredients of acid whey are lactose, protein, and lactic acid. Due to the importance

of lactose in the food and pharmaceutical industries, there have been extensive studies on the

crystallization behavior of lactose (Roos and Karel, 1992; Jouppila et al., 1998). The effect of

lactic acid on the crystal growth of lactose needs to be studied due to its fundamental and

commercial importance in food processing. However, there is little known about the

crystallization behavior of lactose in the presence of lactic acid (Bhandari, 2008), and it has

been found that presence of protein at low concentrations (< 5 w/w %) significantly affects

the product yield and crystallinity (Haque and Roos, 2004). The crystallization kinetics of

lactose/protein solutions, especially when lactic acid is added to the solutions, for example

acid whey, have not been studied in terms of in-process crystallization during spray drying. It

is difficult to predict the crystallization behaviour of commercial acid whey due to the

presence of other impurities. Therefore, in Chapter 5, in one series of experiments, only solid-

phase crystallization of lactic acid and lactose was carried out. Then further experimentation

has been done on the crystallization behavior of powders produced during spray drying using

solutions of lactose/protein/lactic acid to imitate the composition of typical acid whey

solutions. Different methods have been used to assess the degree of crystallinity of the spray-

dried products for different concentrations of lactic acid.

2.7 Disadvantages of Crystalline Powders

As mentioned earlier, the spray-dried forms of dairy powders are normally amorphous and

hence relatively disordered in their molecular structures, which may have several advantages,

such as greater solubility and dissolution rates and surface areas (Hancock and Parks, 2000;

Salonen et al., 2005b; Babu and Nangia, 2011). However, they are thermodynamically-

unstable solids (Roos and Karel, 1991; Hogan and O'Callaghan, 2010), which may lead to an

increased rate of physicochemical changes in dried products, such as sticking, collapse,

caking, agglomeration and crystallization. To overcome these problems, a controlled process,

here a multistage fluidized bed, has been used where the transformation from the amorphous

state to the crystalline form has been done relatively rapidly compared with the rate in

storage, to reduce the time required to crystallize the components of amorphous lactose. Due

to the multistage fluidized-bed controlled process, more stable powder products can be

Page 39: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

21 | P a g e

produced, but crystalline powders may have some undesirable properties, such as lower

dissolution rates.

The creation of highly-porous crystalline powders with better functional (free flowing) and

storage (non-caking) properties, would be helpful for the food industry and pharmaceutical

applications. Thus, a new spray-drying process has been developed with post-drying

treatment to create powder structures with significant porosities and high surface areas, while

maintaining the high stability of a highly crystalline powder.

2.8 A New Templating Method to Produce Highly Crystalline Powder

with Significant Porosity and High Surface Area

The most common application of lactose in the pharmaceutical industry is its use as a diluent

in tablet processing with both direct compression and wet-granulation methods (Kibbe,

2000). In general, the compressibility of excipient powders, such as lactose and mannitol, has

been reported to depend on the powder properties, such as specific surface area, crystallinity,

polymorphism, particle size, and crystal habit (York, 1983).

Increasing the particle surface roughness increases the tablet crushing strength by promoting

the bonding mechanisms between solid surfaces in the compact agglomerate (Karehill et al.,

1990; Riepma et al., 1990). It has been suggested that the dissolution rates of particles with

large surface areas and porosities are relatively high, which results in fast tablet disintegration

and dissolution (Simões et al., 1996; Danesh et al., 2001; Parikh, 2011; Yamasaki et al.,

2011). It has been found that different crystal forms and polymorphs of theophylline have

different physicochemical properties, which influences their tableting behaviour (Suihko et

al., 2001). The same result has been reported for different polymorphs of mannitol (Debord et

al., 1987; Burger et al., 2000). McKenna and McCafferty (1982) reported that decreasing the

particle size of spray-dried lactose increases the tablet tensile strength and results in stronger

compacttions The effect of crystal habit on compressibility of powders has also been

extensively reviewed (Marshall and York, 1991; Garekani et al., 1999). Staniforth et al.

(1981) changed the crystal habit and porosity of mannitol and obtained a highly porous

mannitol for direct compression by using a special crystallization technique.

Page 40: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

22 | P a g e

Different techniques can be used to obtain porous spray-dried particles, but the dominant

method for producing porous inorganic particles by spray drying appears to be the use of

aqueous solutions containing the main (core) material with a templating agent that is removed

by heat treatment, including calcination (Lee et al., 2010; Nandiyanto et al., 2010; Oveisi et

al., 2010; Zhang et al., 2010; Balgis et al., 2011; Balgis et al., 2012; Majano et al., 2012;

Sachse et al., 2012; Emmanuelawati et al., 2013; Fiorilli et al., 2013; Jang et al., 2013; Melo

et al., 2013; Nandiyanto et al., 2013). The key challenge is to extend this understanding of the

powder-templating process from inorganic materials to organic ones, thereby generalizing

this understanding.

Nandiyanto et al. (2013) designed a template-driven self-assembly technique to prepare a

porous material with multisized pores. In this technique a spray-drying process was used for

solutions obtained by mixing silica (an inorganic material) and polystyrene spheres (organic

templates) with different sizes to produce porous silica with a range of pore sizes. In

materials with controllable pore sizes, an improvement in the material properties, compared

with a single-sized pore structure, was achievable. This material may also be very effective

for either selective adsorption or catalytic activity (Nandiyanto et al., 2010).

Zhang et al. (2010) reported a simple approach, using spray drying, to prepare macroporous

pure CoFe2O4 spinel microspheres. In this approach, the precursor slurry containing CoFeFe-

LDH, with or without sulfonated polystyrene microspheres as templates for constructing

macropores, was spray dried. Then the prepared CoFeFe-LDH microspheres were calcined in

air at 700 °C. They found that CoFe2O4 spinel microspheres maintain the original spherical

morphology of the precursor LDH microspheres during thermal decomposition. Layered

double hydroxides (LDHs), as excellent anion exchange materials, are also called anionic

clays. These materials have a wide variety of applications in catalysis, separation technology,

optics, medical science and nanocomposite material engineering (Khan et al., 2001;

Nalawade et al., 2009).

The use of polystyrene particles with various zeta potentials as templates to produce highly-

ordered porous agglomerates from inorganic colloids has been demonstrated by Lee et al.

(2010). Nanoparticles are important as the building blocks for the construction of periodic

and quasiperiodic crystal structures (Chen and Ozin, 2009). In this method, a suspension of

silica colloidal nanoparticles and polystyrene as organic template nanoparticles was spray

dried and then the template particles were removed by calcination at high temperatures of

Page 41: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

23 | P a g e

over 450oC from the product. It has been reported that the final structures of the particles

were highly-ordered porous and hollow nanostructures, the quality of which depends on the

initial concentration of particles, the particle size and the zeta potentials of the silica and

polystyrene particles.

Spray-dried mesoporous titania particles were prepared with a triblock copolymer as a

template by Oveisi et al. (2010). During the process, a suspension of precursor solution,

consisting of titanium tetraisopropoxide dissolved in hydrochloric acid with an ethanol

solution containing Pluronic F127 as a templating agent, was spray dried and then the

products were calcined to remove the surfactants and create the mesoporous particles. The

effect of various calcination temperatures on the mesostructures was also studied, and the

template was removed by calcination at different temperatures of 350oC, 400oC and 500oC. It

was found that the surface areas and the pore volumes were reduced by increasing the

calcination temperatures because of the distortion of the mesostructures, due to the grain

growth of the anatase phase and the transformation to the rutile phase during the calcination

process.

Porous NiO–ZrO2 powders were made by Balgis et al. (2011) using spray drying, with

polystyrene latex (PSL: 400 nm) as a template and the main materials being NiO powders (7

nm) and ZrO2 sols (1.2 nm). They made porous particles that had an average diameter of 4.5

mm, a surface area of 27 m2 g-1 and pore sizes of around 300 nm at a pH of 3.7 in the sprayed

solutions. They also found that, by adjusting the solution pH and the size of the template

particles, powders with a controlled morphology and pore sizes could be produced. The

particles were calcined at 900oC and 1200oC, and the particles shrunk by under 36%.

Majano et al. (2012) presented a one-step synthesis method for the formation of a

mesoporous matrix by spray drying a slurry consisting of a high mesoporosity surfactant-

templated silica matrix and a mesoporous zeolite. The mesoporous silica matrix produced by

the evaporation-induced assembly of the silica nanoparticles was combined, together with the

non-ionic surfactant from spray drying and the mesoporous zeolite obtained by alkaline

treatment, into an attrition-resistant containing sodium hydroxide and tetrapropylammonium

bromide. They reported that the final materials were multimodal and highly interconnected

porous networks, and they also found that the properties of the matrix could be tuned by

changing the spray-drying conditions and the slurry mixture.

Page 42: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

24 | P a g e

Balgis et al. (2012) designed a way to produce carbon-supported platinum (Pt/C) catalysts

with controlled morphologies (dense and hollow-porous) for high-performance PEM fuel cell

(PEMFCs) applications. Spray drying has been used to produce microspherical carbons

(modified catalyst support). The morphological control of carbon microspheres was done by

adding the polystyrene latex (PSL) as template particles prior to spray drying. PSL was

removed by heating the spray-dried particles to obtain hollow-porous microspheres. Pt/C

catalysts were obtained by impregnation of Pt nanoparticles on the surface of the carbon

particles. They found that the electrocatalytic activity of the modified Pt/C catalyst particles

was improved by comparison with a commercial Pt/C catalyst.

Silica-based microspheres were produced by Yurchenko et al. (2012) using

[(C2H5O)3Si]2C2H4 and (C2H5O)3Si(CH2)2P(O)(OC2H5)2 as functionalizing agents and

Pluronic 123 as a template through spray drying. The template was removed by boiling in

methanol. Silica microspheres produced by this templating method had a large specific

surface area of 747 m2 g-1.

Mesoporous silica–titania catalysts were made by Sachse et al. (2012) by spray drying an

aqueous solution containing an ethanolic suspension of siloxane oligomers with titanium

monomer precursor Ti(OiPr)2(acac)2 and α-chitin-nanorods as templating agents, then

calcining the obtained powders. They demonstrated that textural properties (surface area,

pore volume) of the synthesized materials could be tailored by varying the chitin volume

fraction during the synthesis and also by adjusting the Si/Ti molar ratios. Mesoporous silica–

titania catalyst may be applied to the oxidation of a variety of organic compounds, such as

sulfur-containing materials (Huybrechts et al., 1990; Reddy et al., 1992).

Melo et al. (2013) proposed a new approach to produce microbial-imprinted microspheres of

silica nanoparticles through the spray drying of mixed colloidal dispersions containing silica

and Escherichia coli as the templating agent, and then removing the templating materials by

calcination of the spray-dried powders at 300°C for 10 h in a muffle furnace. A high surface

area and a three-dimensional matrix were prepared by microbial imprinting, which can be

used to treat drinking water. The resulted materials showed effective filtration performance in

terms of filtering out microbial cells from aqueous media.

Nanoporous silica particles with controlled morphology, large pore size, and high pore

volume were made by Emmanuelawati et al. (2013) using spray drying, with commercial

Page 43: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

25 | P a g e

silica, Aerosil 200, as a precursor and lanthanum nitrate as a templating agent. The particles

were calcined in air at 550°C for 5 h. The resulted lanthanum oxide-functionalised

microspheres of silica showed high phosphate adsorption capacities of up to 2.317 mmol g-1.

In general, templating techniques have been extensively used to produce highly-porous

particles with multi-component mixtures of inorganic and colloidal materials (Lee et al.,

2010; Nandiyanto et al., 2010; Zhang et al., 2010; Balgis et al., 2011; Balgis et al., 2012;

Nandiyanto et al., 2013). The key challenge is to extend this understanding of the powder-

templating process from inorganic materials to organic ones, thereby generalizing this

understanding. In Chapter 6, a method has been developed to create highly-porous templates

or frameworks of lactose through the spray drying of lactose using food-grade acid as the

templating agent (both water soluble). Then the templating material has been removed by

ethanol washing of the spray-dried powders, simultaneously crystallizing the powders, giving

stable porous and crystalline powder particles. Novel contributions of the new process

reported in this work include the use of a food-grade, non-toxic and benign templating

material, citric acid, in a process involving spray drying, which is a low-cost continuous

drying technique relative to many other drying operations, such as freeze drying.

The effect of glass-transition temperature and strength of the acid (pH of the solution) on the

degree of crystallinity for lactose in spray-dried powders is critical, with implications for the

performance of the templating approach using acidic templates. Therefore, in this chapter the

effect of using different concentrations for various food-grade acids, such as lactic acid, citric

acid, boric acid, and ascorbic acid, on the degree of crystallinity for spray-dried lactose

powders and the porosity of the ethanol-washed powders has been investigated.

2.9 Mannitol as an Alternative to Lactose

Recently, the use of mannitol as an alternative to lactose in food products and pharmaceutical

formulations has significantly increased. Lactose is a reducing sugar, which is incompatible

with some active pharmaceutical ingredients, such as peptides and proteins, and also has

problems associated with lactose intolerance (Dangaran and Krochta, 2006; Dangaran et al.,

2006; Babu and Nangia, 2011; Littringer et al., 2012). The non-hygroscopic nature of

mannitol, and consequently its low moisture content, make it a desirable additive and

Page 44: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

26 | P a g e

excipient for moisture-sensitive ingredients (Babu and Nangia, 2011). Mannitol has also been

used as a primary material for protein stabilization in dried protein formulations (Bhandari et

al., 1997a; Bakaltcheva et al., 2007; Hulse et al., 2009).

Jayasundera et al. (2010) improved the tableting properties of mannitol by decreasing the

crystallinity of the particles through rapid solidification of the melted mannitol in a spray

dryer. However, the produced amorphous fused mannitol was rather unstable, and its

compressibility decreased as the crystallinity gradually increased during storage. Maas et al.

(2011) produced spray-dried mannitol particles with different surface roughnesses by spray

drying at different outlet temperatures. They found that increasing the outlet temperature of

the spray dryer increased the crystal size and particle surface roughness, which was attributed

to different crystallization mechanisms. In an attempt to increase the mannitol surface area,

Modler and Emmons (1978) found that exposing the mannitol to a high relative humidity

after spray drying significantly changed the morphology of the particles and increased their

surface area from 0.4 to 2.3 m2 g-1, which was associated with the moisture-induced

polymorphic transition of mannitol from the δ to the β form. The high surface-area mannitol

produced with this method has been shown to possess excellent compaction behaviour

(Drapier-Beche et al., 1997). Typical specific surface areas of mannitol powder have been

reported to be between 0.3 to 0.6 m2 g-1 (Babu and Nangia, 2011; Ho et al., 2012).

A templating approach has been applied here according to the method described in Chapter 6

to obtain highly-porous templates or frameworks of mannitol through the use of the low-cost

and high-throughput spray-drying process for mannitol using citric acid as the templating

agent. Then, the template material (citric acid) has been removed by ethanol washing of the

spray-dried powders. The core materials (mannitol), and the citric acid as the templating

agent (both water soluble), have been mixed and spray dried. The resulting powders have

been washed with ethanol to create porous powder particles by removing the citric acid. The

effects of surface active compounds, such as whey protein isolate (WPI) and Gum Arabic, on

the surface area and the pore volume of the resultant mannitol have also been investigated in

Chapter 7.

Page 45: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

27 | P a g e

2.10 Highly Porous Excipients as a Solution for Poor Solubility and Poor

Content Uniformity in Solid Dosage Form

Solid-dosage form drugs that include tablets and capsules are some of the most popular and

commonly-employed methods of drug delivery, which currently account for over two-thirds

of the total number of medicines produced in the world (Zheng, 2009; Sahoo, 2012). Their

final formulation contains an active pharmaceutical ingredient (API), its excipients and also

additives, which may be included in the formulations to enhance the physical appearance,

improve stability, and aid in disintegration after administration. However, major challenges

facing the design of the oral dosage form are poor bioavailability and uniformity of drug

dosage, especially in low-dose solid-drug products (Garcia and Prescott, 2008; Huang and

Sherry Ku, 2010; Grigorov et al., 2013). Among other essential qualities of a well-made

pharmacopeia, uniformity of drug dosage is a very important. The problems associated with

the mixing of an active pharmaceutical ingredient (API) with excipients are many and

complex, especially for low-dose (˂100µg) drugs during the manufacturing of solid dosage

forms, which may result in undesirable variations in dosage (Cartilier and Moes, 1989;

Garcia and Prescott, 2008; Zheng, 2009; Huang and Sherry Ku, 2010; Grigorov et al., 2013).

Therefore, it is critical that the drug be uniformly distributed in the final product. The

potential problems associated with the control of dose uniformity need to be addressed

carefully to ensure that the proper dosage of the drug is delivered to the patient. The content-

uniformity requirements are regulated throughout the world to ensure that the mixture of the

drug and its excipients is adequately uniform in the finished products (Garcia and Prescott,

2008). The Food and Drug Administration (FDA), in setting standards for the content

uniformity of pharmaceutical products, requires that the relative standard deviation (RSD) is

less than or equal to 6%, while the maximum acceptable deviation in the active substance

content of the finished products must not exceed ± 5% at the time of manufacture, according

to European Pharmacopoeia requirements (FDA, 2003; Directive, 2003/63/EC).

Particle size control of the API can enhance the uniformity of solid dosage forms; however,

this approach for drugs may not be able to maintain the level of homogeneity throughout

processing due to the possibility of blend desegregation, and consequently poor blending

uniformity as a result of differences in particle size, shape, or density of the materials being

blended (Garcia and Prescott, 2008; Zheng, 2009; Huang and Sherry Ku, 2010). In addition

Page 46: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

28 | P a g e

to blending uniformity, bioavailability, as one of the important quality parameters of drug

formulations, refers to the extent and rate at which an active drug reaches systemic

circulation. Poor solubility is one of the main causes of low bioavailability (Schreiner et al.,

2005; Savjani et al., 2012). More than 90% of small molecular-weight drugs are delivered to

the human body in crystalline form; however, 90% of crystalline drugs have low solubility in

water (Variankaval et al., 2008). Low aqueous solubility is the major problem encountered

with formulation development, and poorly soluble drugs often require high doses in order to

reach the desired bioavailability after oral administration, which may lead to increased side

effects. Formulation and dosage-form design must ensure that solubility requirements are met

for manufacturing as well as clinical applications.

There are various solubilization techniques to enhance the solubility of poorly water-soluble

drugs, such as solid dispersion, particle size reduction, complexation, the use of surfactants,

and novel excipients (Savjani et al., 2012; Ojha and Prabhakar, 2013; Dhillon et al., 2014).

Common strategies to increase the dissolution rate and hence the bioavailability of such drugs

include reduction of particle size (Leuner and Dressman, 2000) to generate effective surface

areas. As a particle becomes smaller, the surface area to volume ratio increases, and the larger

surface area allows greater interaction with the solvent, which causes an increase in solubility

(Savjani et al., 2012). Particle size reduction may also enhance the possibility of API

uniformity for solid dosage forms (Huang and Sherry Ku, 2010). Another potential approach

to solve the abovementioned challenges for better controllable release characteristics of a

drug is deposition or nano-confinement of drugs into porous excipients as hosts, in which a

precise amount of highly-concentrated drug solution can be dispersed in the pore spaces of a

porous carrier (Charnay et al., 2004; Beiner et al., 2007; Mellaerts et al., 2007; Mellaerts et

al., 2008; Rengarajan et al., 2008; Van Speybroeck et al., 2009; Ji et al., 2010; Verraedt et al.,

2010; Grigorov et al., 2013).

For successful adsorption of drugs, highly-porous excipients are needed, with high surface

areas being the most important property (Grigorov et al., 2013). However, it is unclear how

the effects of porous excipients are related to the solubility of final drug formulations.

Carriers with high porosity can improve the solubility by enabling faster release of the drug

as a result of better penetration of solvent into the drug-excipient matrix. Presenting a suitable

drug-loaded particle, which consists of dispersing APIs in the host pore space of highly-

porous excipients, results in particle size reduction of the drug powders to the size range of

Page 47: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

29 | P a g e

nanometres. Nanconfinement breaks up the intermolecular interactions of the API molecules

and separates them inside the internal void structures of nanoporous excipients. This situation

combines the benefits of an increase in the solubility with a reduction in the particle size of

the API and maximizes the surface area of the compound that comes into contact with the

dissolution medium as the carrier dissolves.

The reduction in the particle size of an active pharmaceutical ingredient (API) to the nano-

crystal size range may cause a significant increase in the dissolution rate of the API, often

resulting in significant increases in bioavailability. Greater porosity of excipients facilities the

diffusion of the API components and increases the rate of drug adsorption.

For this purpose, a method of adsorption has been developed using highly-porous excipients

with high surface areas produced according to the templating method described in Chapter 7.

These new engineered particles, furthermore, have been used as the key part of an adsorption

method to investigate drug dissolution. This work has been done by applying an adsoption

technique and modifying our current templating process, as described in the following

chapters (Chapters 8 and 9). Highly-porous mannitol particles with high surface areas have

been successfully produced through the spray drying of mannitol solutions containing

ethanol-soluble food-grade acids, such as citric acid as a templating agent, and then removing

the citric acid by ethanol washing of the spray-dried powders to create porous powders. The

resultant porous particles, after ethanol washing, have been transferred to ethanol solutions to

dissolve and carry acetaminophen as a sample drug into the pores of porous mannitol

(Chapter 8).

Chapter 9 has reported on the evaluation of the potential for this new adsorption technique to

improve the dissolution characteristics and bioavailability of poorly aqueous-soluble drugs,

such as indomethacin and nifedipine, using porous mannitol as the carrier.

Page 48: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

30 | P a g e

CHAPTER 3 MATERIALS, EQUIPMENT AND METHODS

This chapter reviews the most appropriate techniques regarding characterization of products,

such as the particle size and shape, the moisture content, the glass-transition behaviour and

amorphous/crystalline states, and other issues concerning templating and drug-loading

processes, such as BET surface area analysis, tablet crushing tests, UV-spectrophotometry,

and in-vitro dissolution studies.

3.1 Materials

The following materials have been used in the experiments, as shown in Table 3.1.

Table 3.1 List of materials used in the experiments.

Compound Chemical Formula Reagent Grade Supplier

pure α-lactose monohydrate crystals

C12H22O11.H2O analytical reagent Chem-Supply, Australia

sucrose C12H22O11 laboratory-grade reagent

Chem-Supply, Australia

lactic acid C3H6O3 88% w/w, laboratory reagent

Chem-Supply, Australia

D-Mannitol C6H14O6 analytical reagent Chem-Supply, Australia

citric acid as monohydrate

C6H8O7.H2O analytical reagent Chem-Supply, Australia

Gum Arabic from acacia trees

- analytical reagent Chem-Supply, Australia

ascorbic acid C6H8O6 analytical reagent Chem-Supply, Australia

boric acid H3BO3 laboratory reagent Ajax Finechem, Australia

sweet whey powder

- -

Dairy Innovation Australia Limited (DIAL)

potassium phosphate monobasic

KH2PO4 ACS reagent Chem-Supply, Australia

sodium hydroxide

NaOH ACS reagent Chem-Supply, Australia

Page 49: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

31 | P a g e

Compound Chemical Formula Reagent Grade Supplier

absolute ethanol 100% denatured

C2H5OH laboratory reagent Chem-Supply, Australia

methanol CH4O ACS spectrophotometric grade, ≥99.9%

Sigma-Aldrich

acetaminophen C8H9NO2 BioXtra, ≥99.0% Sigma-Aldrich

acetone C3H6O laboratory-grade reagent

Chem-Supply, Australia

indomethacin C19H16CINO4 ≥99%

Baoji Guokang Bio-Technology Co., China

nifedipine C17H18N2O6 ≥99%

Baoji Guokang Bio-Technology Co., China

sodium lauryl ether sulfate

(CH3(CH2)11(OCH2CH2)nOSO3Na) analytical reagent Chem-Supply, Australia

hydrochloric acid

HCl 32 %, laboratory-grade reagent

Chem-Supply, Australia

sodium chloride NaCl laboratory-grade reagent

Chem-Supply, Australia

pepsin laboratory-grade reagent

Chem-Supply, Australia

whey protein isolate

consisting of 92 g protein, 0.4 g fat (total) including 0.2 g saturated fat, 0.5 g carbohydrate, and sodium 0.6 g per 100 g

-

Balance, Vitaco Health Ltd, Auckland, New Zealand

3.2 Equipment

3.2.1 Spray Dryer

A mini spray dryer B-290, (Büchi AG, Switzerland) was used to dry the solutions at wide

ranges of inlet air temperatures (oC), a full air flow rate (38 m3 h-1), and different liquid pump

rates (mL min-1), nozzle air flow rate (L h-1) and feed concentrations (w/w) % for all the

experiments. A basic schematic diagram of the spray dryer is shown in Figure 3.1.

Page 50: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

32 | P a g e

Figure 3.1 Schematic diagram of the BüchiB-290 spray dryer.

3.2.2 Fluidized-Bed Dryer

A three-stage fluidized-bed drying (F.B.) system has been used for the experiments (Figure

3.1). The consecutive stages have been partially separated by fixed partitions. This

arrangement has given different sets of air temperatures and relative humidities for each

stage. Therefore, for continuous flow of solids through the stages, the particles have been

exposed to different process conditions to reach different levels of crystallinity. A pneumatic

linear vibrator (piston type, maximum pressure 6 bar) has been attached to the inlet side of

the bed to vibrate the fluidized bed and keep the solids well fluidized. The whole bed sits on

four springs located at the four corners of the bed. The springs and the supporting bed-frame

are securely attached to the base frame/structure of the unit. Each process stage has an air

blower (170 m3 h-1) and an air heater (containing customized heating coils, each coil 240 V

and 2500 W). The first stage-heater (warm-up zone) has two coils, the second stage-heater

(main processing zone) has three coils, and the third stage-heater (finishing zone) has a coil in

the heater. A ‘Humidifier’ is connected to the second stage. The humidifier has two inlets-one

for the hot air and the other for the steam, and two outlets-one for the humidified hot gas and

Air heater

Peristaltic pumpAtomizer

Air outlet fan

Drying chamber

Cyclone

Atomizer air

TinT Controller

Tout

Feed

Collection vessel

Nozzle cleaner

Page 51: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

33 | P a g e

the other for draining any condensate. This cylindrical column type unit has stainless steel

wool distributed on a support frame just below the inlets at the top of the column and a

similar distributor at the bottom of the column and just above the outlets. These arrangements

have been made to ensure better mixing of the hot gas and the steam without excessive

pressure drop. There is a ‘hood’, split into three sections by partitions, which are aligned with

the bed separators. At the top of the hood, three exit ducts are connected, to direct the gas to

the cyclones/exhaust duct. There are arrangements of perforated plates above the ‘Wind Box’

to ensure good distribution of air across the bed. The wind box is also partly filled with

stainless steel wool to distribute the gas within the confined space.

A set of dry- and wet-bulb temperature probes (RTD PT100) are placed just below (2~3 cm)

the fluid bed for each stage. The dry-bulb temperature can be set to a particular level using

the heater control system. Dry- and wet-bulb temperatures also give a relative humidity

measurement at these particular points. There are also three temperature probes on the

‘Hood’, close to the exit of each stage, to record the final/exit gas temperature of the each

stage. The air blowers can be adjusted for different flow rates using the variable speed drives

(Figures 3.2 and 3.3).

Page 52: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

34 | P a g e

Figure 3.2 Schematic diagram of the Pilot-Scale Fluidized Bed Dryer.

To 1st

Air distributor

Powder on bed

T1 DB T1 WB T2 DB T2 WB T3 DB T3 WB

T2 exit DB T3 exit DB

Powder Inlet

Powder Outlet

Air Fans 170m3/hr

Heaters each coil 240V and 2500W

Humidifier

Steam

To 2nd Stage

To 3rd Stage

To exhaust or recycle back to 1st stage-fan

T1 exit DB

Cyclone

Control board

Variable speed drives for fans

Heater Controllers

Fresh Air

Humid air recycles to the following stages

Data Record Ti WB

Ti exit DB

T3DB T2DB T1DB

Page 53: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

35 | P a g e

Figure 3.3 Pilot-Scale Fluidized Bed Dryer.

3.3 Powder Characterization

Determining the degree of crystallinity for lactose is important in understanding if the

product experiences changes during different processing or storage conditions. Different

methods have been utilized for the detection and qualification of the amorphous fractions or

components in lactose within dairy powders, including: Moisture sorption measurement, X-

ray diffraction (XRD), conventional differential scanning calorimetry (DSC), modulated

differential scanning calorimetry (MDSC), Infrared and Raman spectroscopy and microscopy

(Sperling, 1986; Miao and Roos, 2005; Lehto et al., 2006).

Most methods used in the characterization of the solid state are based on the detection of its

structural properties, such as the crystalline structure, which may be derived from X-ray

diffraction studies (XRD). XRD is a practical technique for identifying crystal forms and

following or characterizing rates of crystallization for components, especially when the

systems are complicated, such as systems with several crystallizing components (Miao and

Roos, 2005).

The amorphous states have disordered structures, and the physical state of amorphous

materials is related to molecular mobility (Roos, 1995). Molecular mobility is affected by the

Page 54: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

36 | P a g e

temperature and phase transitions. Various spectroscopic methods are available for the

examination of molecular mobility. Spectroscopic methods that have been used in the

determination of molecular mobility in food materials include Raman spectroscopy (Sperling,

1986) and Fourier transform infrared (FTIR) spectroscopy (Slade and Levine, 1995).

The analysis of increasing intensities and areas of the peaks in X-ray diffraction (XRD)

patterns using X-ray diffractometry is an important technique in observing the properties of

crystalline solid materials, polymers, and food materials (Roos, 1995), which can be used to

obtain qualitative and quantitative information about the crystalline structure (Marsh and

Blanshard, 1988; Drapier-Beche et al., 1997; Jouppila et al., 1997; Corrigan et al., 2004;

Langrish and Wang, 2009; Islam et al., 2010; Yazdanpanah and Langrish, 2011b). In food

applications, X-ray diffraction patterns are extremely useful in the characterization of

crystalline states and crystal structure in such materials as starch and sugars, and in the

detection of crystallinity in partially amorphous food components. In this technique, the

powdered sample is exposed to a source of X-rays from a series of angles in a scanning

manner. The diffracted radiation can be detected using photographic film or an electronic

counter. True crystalline states will demonstrate a sharp peak of reflection from the lattice

structure, while amorphous food materials have no characteristic X-ray diffraction patterns

(Roos, 1995).

Water sorption in amorphous food polymers is known to be a time-dependent phenomenon

(Jouppila and Roos, 1994b). Amorphicity can be investigated by detecting the mass change

profile during the absorption of moisture and the subsequent desorption due to crystallization

taking place in the sample (changes in moisture content, percent (100× kg.kg−1 dry basis)).

The end of the crystallization process can be determined by the point where the sorption

curve reaches a plateau region (Lai and Schmidt, 1990; Jouppila and Roos, 1994b; Jouppila et

al., 1998). Amorphous materials absorb moisture more than their corresponding crystalline

counterparts, and after the moisture-induced recrystallization, the desorption of excess water

after crystallization of the amorphous components is observed (Lehto et al., 2006).

Changes in the physical state are observed from changes in the thermodynamic quantities,

which can be measured with a number of techniques, such as calorimetric ones (Roos, 1995).

Differential scanning calorimetry (DSC) is probably the most common technique for the

determination of phase transitions in inorganic, organic, polymeric and also food materials

(Roos and Karel, 1992; Arvanitoyannis and Blanshard, 1994; Roos, 1995; Kedward et al.,

Page 55: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

37 | P a g e

1998; Mazzobre et al., 2001; 2003). This technique can be used to observe phase transitions

and to determine glass-transition temperatures and the degree of crystallinity, for which the

observations of the melting and crystallization peaks are required. In the DSC method, the

samples are usually placed in pans that can be hermetically sealed, and the temperature

difference between the sample and the reference is recorded to derive the difference in the

energy supplied. Therefore, the method can be used to observe phase transitions and to

determine glass-transition temperatures (Roos, 1995). This technique also can be used to

determine the degree of crystallinity, for which the observations of the Tmelting

and Tcrystallization

peaks are required. The ratio of the heights or the areas under the peaks for the melting and

crystallization peaks give a value for the crystallinity of the sample.

Modulated differential scanning calorimetry (MDSC) allows separation of the total heat flow

signal into its thermodynamic (heat capacity) and kinetic components. MDSC offers

simultaneous improvements in sensitivity and resolution, and can separate overlapping events

that are difficult or impossible to do by standard DSC.

Microscopy and Infrared and Raman spectroscopy may also be applied to obtain information

on the crystallinity of amorphous systems (Akao et al., 2001; Mazzobre et al., 2003; Murphy

et al., 2005; Yazdanpanah and Langrish, 2013b). Microscopic methods, including optical

microscopy and electron microscopy, are particularly useful in the observation of changes

that may occur in the crystal size and crystallinity in foods during storage (Roos, 1995).

3.4 Measurements of Degree of Crystallization and Thermal Transitions

3.4.1 Moisture Sorption Tests

Time-dependent moisture sorption tests have been carried out for the powders from each

experiment. It has been suggested by Lehto et al. (2006) that the moisture adsorption peak

heights (changes in moisture content, percent (100× kg.kg−1 dry basis)) from the moisture-

sorption experiments can be used to assess the extent of lactose amorphicity by detecting the

mass change profile during the absorption of moisture and the subsequent desorption due to

the crystallization taking place in the sample. The end of the crystallization process can be

determined by the point where the sorption curve reaches a plateau region (Lai and Schmidt,

1990; Jouppila and Roos, 1994b; Jouppila et al., 1998). Amorphous materials absorb

moisture more than their corresponding crystalline counterparts, and after the moisture-

Page 56: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

38 | P a g e

induced recrystallization, the desorption of excess water after crystallization of the

amorphous components is observed (Lehto et al., 2006).

A mass of 1–2 g of the powder product (obtained just after spray drying the sample solution)

was placed on a Petri dish (borosilicate glass) of 10-cm-diameter. The dish was placed on an

analytical balance (± 0.0001 g, Mettler Toledo, AB 204-S, Switzerland), and the dish and the

balance were placed in a sealed container. In this container, electric light bulbs and a sodium-

chloride saturated-salt solution were used to control the temperature (24.5-25ºC) and the

relative humidity (70-75% RH), respectively, and the mass change as a function of storage

time was recorded by a computer once per minute over a period of 1–2 days until the mass of

the sample and dish stayed constant for more than six hours. The crystallization process has

been considered to be virtually complete when the moisture content remained constant for a

long time at a low level (Lehto et al., 2006). However, in the present study, the moisture

sorption curves have been normalized for the amount of lactose in the final spray-dried

powders (containing lactose and food-grade acids) since during the moisture sorption tests

only lactose crystallizes. Hence, the corresponding peak heights should be corrected on the

basis of the dry mass of lactose not the whole spray-dried powder (only for sample that the

lactose portions have been changed the final spray-dried powder).

3.4.2 X-Ray Diffraction (XRD)

A Siemens D5000 diffractometer with Cu target and maximum power of 2200 W has been

used to measure the overall powder crystallinity. Samples have been loaded into standard

plastic plates for measurements according to standard sample preparation procedure. The

operating conditions have included a range for scanning of 5–30º, 1 step/s was taken with a

0.02º step size, and the current and voltage have been 30 mA and 40 kV, respectively. An

analysis program, DIFFRAC Plus, Bruker analytical X-ray system, GmbH, has been used to

search for relevant peaks and to calculate the areas under the peaks when assessing the

relative degrees of lactose crystallinity for different samples.

3.4.3 Modulated Differential Scanning Calorimetry (MDSC)

The glass-transition and crystallization temperatures and also the heats of crystallization for

the powders have been determined by modulated differential scanning calorimetry (MDSC).

Page 57: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

39 | P a g e

Samples for MDSC measurements have been prepared according to standard procedures

using hermetically-sealed pans. Four to six milligrams of sample have been used in each

analysis. The samples have been heated from 5ºC to 250ºC using a ramp rate of 5ºC min-1

with a 1ºC modulated signal every 60 seconds using a modulated differential scanning

calorimeter (TA Instruments Q1000).

3.4.4 Raman Spectroscopy

Raman spectra were measured to assess the degree of surface crystallinity for the powders

using a Raman Station 400F (PerkinElmer, CA, USA). The samples were analysed using a

power level of 100% with a 785 nm laser and a five second exposure time, with five

exposures over a wavelengths range from 200 cm-1 to 1500 cm-1. The spectra were baseline

corrected using the software Spectrum v6.3.4.0164 and further analysed using Microsoft

Excel. Two spectral bands were chosen for the determination, one centred at 440 cm−1 for

amorphous lactose and another centred at 470 cm−1 for crystalline lactose. The band areas

were integrated between the spectral regions of 410 cm-1 to 490 cm-1 and 450 cm-1 to 490

cm−1. This spectra region has been chosen since this region has been used to estimate the

crystallinity of lactose by Niemelä et al. (2005) and Katainen et al. (2005).

3.4.5 Fourier Transform Infrared Spectroscopy (FTIR)

The FTIR absorbance spectra were used to distinguish the degree of surface crystallinity for

the lactose using a single bounce diamond ATR (Universal ATR) in a Nicolet 6700 FTIR

spectrometer (Thermo Fisher Scientific Inc.) controlled by OMNIC 8.2.387. The FTIR

spectra were collected at a resolution of 4 cm-1 with 32 scans over a wavelengths range from

600 cm-1 to 2000 cm-1. The graphs were base line corrected using the OMNIC 8.2.387

software and were further analysed using Microsoft Excel. Peaks have been analysed at 875

cm-1, 900 cm-1 and 1260 cm-1, since they have been reported to distinguish amorphous from

crystalline lactose (Listiohadi et al., 2009). The FT-IR absorbance spectra were also used to

measure the extent of citric acid removal. The spectrum of the pure citric acid solution has a

stretching band at 1,625 cm-1, attributed to the C=O in the dissociated carboxylic acid, while

it is 1,730 cm-1 when not dissociated (Socrates, 2001; Masoudpanah and Seyyed Ebrahimi,

2011).

Page 58: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

40 | P a g e

3.5 Characterization of Drug-Loaded Powders

3.5.1 Determination of Drug Loading and Blend Uniformity

An UV-spectrophotometric method has been developed according to an approach described

by Behera et al. (2012) using a solvent consisting of methanol and water (15:85, v/v) as a

diluent to determine the drug content in the final formulations. Six standard solutions of

acetaminophen in the range of 0–150 μg mL-1 were prepared using diluents, and UV

spectroscopic scanning (200–500 nm) was carried out by a UV Spectrophotometer (Cary 50,

Varian, USA) to obtain a calibration curve. At the maximum-absorption wavelength (λmax) of

275 nm, a linear response for acetaminophen was found with a correlation coefficient of

R2=0.999. However in the case of poorly water soluble drugs, indomethacin and nifedipine,

five standard solutions of drugs in the range of 0–50 μg mL-1 were prepared using a solvent

consisting of ethanol and water (50:50, v/v) as a diluent, and UV spectroscopic scanning

(200–500 nm) was carried out by a UV Spectrophotometer to obtain a calibration curve. At

the maximum absorbance wavelengths (λmax) of 335 nm and 320 nm, linear responses with a

correlation coefficient of R2=0.999 were found for nifedipine and indomethacin, respectively.

These calibration curves have been used to quantify the drug content in the final formulation.

The drug content of the drug-loaded mannitol was measured using the predetermined

standard curve. 50-150 mg samples of drug-loaded mannitol were carefully weighed,

dissolved in dilute solution and agitated, at which time the solutions were visually

transparent. From these solutions, 2 mL was taken and diluted again with diluents depending

on the estimated loading. The percentage relative standard deviation (% RSD) for the drug

content of drug-loaded mannitol has been calculated by taking at least five samples for each

drug concentration in each experiment, as the important acceptance criteria for the content-

uniformity dosage requirements of the pharmaceutical industry must be less than a 6% RSD

(FDA, 2003; Directive, 2003/63/EC).

3.5.2 Dissolution Experiments

In-vitro dissolution rate studies of drug-loaded mannitol in the form of capsules were

performed by Apparatus 2 (paddle assembly) according to the United States Pharmacopeia

(Dressman and Krämer, 2005) in a 1-L vessel with a paddle speed of 50 rpm.

Page 59: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

41 | P a g e

Size 0 gelatin capsules (Nutritional Health Supplements, Australia) were filled with each

sample using a capsule filling machine. At least three 300 mg samples of acetaminophen-

loaded mannitol were tested, while the temperature of the phosphate (900 mL) buffer at pH

5.8 (as the media) was maintained at 37°C by a water bath. During the test, samples were

taken to calculate the drug content of final formulation by a UV spectrophotometer at the

main absorbance wavelength of the corresponding drugs. The result was compared with the

release rate for the physical mixture of non-porous mannitol and drug powder in which the

ratios of drug to mannitol were similar to the weight ratios in the final formulation of drug-

loaded mannitol.

However, in the case of poorly water-soluble drugs, dissolution experiments were conducted

in a medium of simulated gastric fluid (SGF) with 0.5% SLS for indomethacin and release

medium of SGF with 0.35% SLS for nifedipine (Van Speybroeck et al., 2009). The

temperature of the release media (900 mL) was maintained at 37 ± 0.5 °C by a water bath. In

order to estimate the solubility of indomethacin and nifedipine in their release medium,

saturated solutions of both drugs were prepared. Excess amount of the drugs were added to

the release medium and then mixed using a magnetic stirrer for 48 h at the room temperature

of 25oC. The solutions were then filtered and analyzed by a UV Spectrophotometer to obtain

the absorbance wavelengths representing the fully-dissolved level of both drugs.

Drug release from the drug-loaded samples was compared to that of their corresponding

physical mixture of non-porous mannitol (as the carrier) and drug. The constant non-sink

condition of 0.01 mg drug/mL of release medium has been applied. Generally, sink

conditions for most dissolution studies are recommended (Liu et al., 2013). Sink conditions

normally require that the drug solubility be ten times the total concentration of drug in the

dissolution medium. However, such an approach may be inappropriate in dissolution studies

for a poorly water-soluble drug since it is not desirable to have incomplete drug dissolution or

over-saturated condition in the dissolution evaluation (Tang et al., 2001; Sirisuth et al., 2002;

Petralito et al., 2012). In addition, sink conditions or high amount of available drug increase

the dissolution rate and lead to rapid dissolution rates for crystalline drugs with different

particle sizes, causing difficulties in discriminating dissolution profiles between different

formulations (Liu et al., 2013). Since this study focused on the comparison of dissolution

profiles for nanometre size drug crystals in the formulation with physical mixture of non-

porous mannitol and crystalline drug (bulk drug), the non-sink condition for the dissolution

Page 60: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

42 | P a g e

study allow discriminating dissolution profiles with different drug particle size. This

condition was suggested by Van Speybroeck et al. (2009) in order to give better

differentiation of varous formulations and physical mixtures. Size 0 gelatin capsules

(Nutritional Health Supplements, Australia) were filled by different amounts of samples to

provide 0.01 mg drug/mL of release medium using the capsule filling machine. During the

test, 5 mL samples were taken, filtered over a 0.45 µm PTFE syringe filter, and analyzed by a

UV Spectrophotometer at the corresponding absorbance wavelengths of both nifedipine and

indomethacin. The absorption results at the appropriate wavelengths were then compared

with those of standard solutions (the absorbance wavelengths representing the fully-dissolved

level of drugs) to obtain the percentage of dissolved drugs.

3.5.3 Tableting Properties

Drug-loaded mannitol powders and physical mixtures of the non-porous mannitol with drug

powders were used for measuring the tableting properties. Tablets with an identical weight of

150 mg were prepared by manually introducing powders into a flat-faced punch of a

hydraulic press at different compaction loads in the range of 100 MPa to 400 MPa. Tablet

dimensions were measured using a caliper with an accuracy of 0.001 mm, and the tablet

thicknesses were plotted as a function of the applied compaction load. An Instron hardness

tester 5943 (Instron, Germany) with a strain rate of 0.2 mm/min and a load capacity of 100 N

was also used to determine the tablet hardness.

3.6 Particle Sizing and Morphology Studies

3.6.1 Scanning Electron Microscope (SEM)

The powders have been observed using a scanning electron microscope, to show the

structures of both the surface and the bulk material. Sample preparation has included placing

a small amount of the sample onto a carbon tape on an aluminium sample stab. A standard

thickness of 30 nm gold coating has been given for each sample to produce the conductive

surface (Emitech, K550X, Quorum Technologies, UK). A Hitachi S4500 FEG- scanning

electron microscope with an operating voltage of 5 kV has been used. A wide range of

magnifications has been used in the images, and the images have been captured using the

software ImageSlave v2.11.

Page 61: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

43 | P a g e

3.6.2 Particle Size Analysis

The particle size distribution was measured by the laser diffraction technique using a Malvern

Mastersizer 3000 (Malvern Instruments, UK) with a dry powder feeder unit (Aero S).

3.6.3 BET Surface Area Analysis

Surface area and pore volumes (or pore size distribution) were determined from nitrogen

adsorption and desorption isotherms, measured at the temperature of liquid N2 (77 K) using a

surface area analyzer (Quantachrome Autosorb-1). The surface area was calculated from the

Brunauer-Emmet-Teller (BET) equations, and the total pore volume was calculated according

to the Barrett-Joyner-Halenda (BJH) method. Outgassing at room temperature overnight has

been performed for all samples prior to all experiments.

3.7 Other Items of Methods and Equipments

3.7.1 Moisture Content Measurement

The free moisture content (MC, dry basis) has been measured by weighing the sample before

and after drying in a fan circulated oven (Labec, Australia) at 85ºC for 24 hours.

3.7.2 pH of Aqueous Solution

The pH of the solutions was measured with a pH electrode, InPro 3250 series (Mettler

Toledo, M 300, Switzerland).

3.7.3 Statistics

Data in this study have been presented as means ± STD (standard deviation) and were

obtained from at least three independent experiments. For statistical analysis, least significant

difference (LSD) tests have been performed first for comparisons between different groups,

and then one-way ANOVA tests have been performed to determine the overall significance

Page 62: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

44 | P a g e

of variations. The statistics program IBM SPSS Statistics 22 was employed for all statistical

analyses, and differences were considered significant at P < 0.05.

Page 63: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

45 | P a g e

CHAPTER 4 COMBINED DRYING AND

CRYSTALLIZATION IN A CONTINUOUS MULTI-STAGE

FLUIDIZED-BED DRYER

This chapter reports the investigation of the effects that operating conditions have on the

crystallization properties of whey powder in a continuous multi-stage fluidized-bed dryer to

achieve the maximum degree of lactose crystallinity within a reasonable overall processing

time. Powder was fed at different rates for each fluidization process at different sets of

process temperatures, humidities, and residence times using a continuous multistage fluidized

bed. The aim was to reduce the time required to crystallize the amorphous-lactose fraction

sufficiently to be feasible in an industrial-crystallization process, where each stage will have

particular temperatures and relative humidities (Figure 4.1). Based on the earlier findings in

dairy powder crystallization using fluidized-bed dryers (Cruz et al., 2005; Nijdam et al.,

2008; Yazdanpanah and Langrish, 2011a; 2011b; 2011c), it is expected that the process

conditions (temperature and relative humidity) for each stage should depend on the feed

powder properties (glass-transition temperature, composition, extent of amorphicity for the

lactose) for that stage. A multi-stage fluidized bed processing scheme has been developed to

achieve greater control of lactose crystallinity than the control achievable through operating a

single stage fluidized-bed dryer on its own.

Page 64: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

46 | P a g e

Figure 4.1 Multi-stage processing scheme for the fluidized-bed dryer.

4.1 Sample Preparation

4.1.1 Fresh Spray-Dried Whey Powder

Aqueous solutions of sweet whey powder (as supplied by Dairy Innovation Australia

Limited, DIAL) (35% w/w) were prepared. The solutions were spray dried using a Buchi-

B290 Mini Spray Dryer with an inlet temperature of 1500C, an aspirator setting of 100% (38

m3 h-1), a pump rate of 25% (8 mL min-1), and a nozzle air flow rate of 470 L h-1 (40 on the

nozzle rotameter scale). These operating conditions caused the outlet temperature to be 90 ±

20C. Freshly spray-dried milk powder was collected from a vessel at the bottom of a cyclone

and instantly fed to the fluidized bed.

4.1.2 Crystalline Whey Powders

In order to create highly-crystalline whey powders (WP), 100 g of fresh spray-dried whey

powder was placed in a Petri dish and stored in a sorption box with a saturated NaCl solution

(25°C and 75% relative humidity) for two weeks. Every day, the bulk material was gently

agitated to break up the caked materials and to mix the powders, exposing all particles to the

moisture in the air. Jouppila and Roos (1994b) and Langrish and Wang (2006) showed that

the amorphous lactose in milk powder is mostly crystallized under such conditions in two

weeks.

Stage 1 Stage 2 Stage 3

(T, RH%)-1 (T, RH%)-2 (T, RH%)-3

Amorphous powder from spray dryer

(MC,Tg) 1

Fluidized bed unit

(MC,Tg)2 (MC,Tg)3Crystallized

powder

B%Time, t1

Crystallinity, A%

C%Time, t2

D%Time, t3

Crystallinity: A < B < C < D

Page 65: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

47 | P a g e

4.2 Process Conditions

Table 4.1 represents all the experiments performed for this study. Amorphous powders were

produced by spray drying, and then the fluidized-bed dryer was used to process the powders

further. Many researchers have studied the sticky behavior of food powders. Yazdanpanah

and Langrish (2011a) found the upper limit of fluidization depends on the percentage of

lactose crystallinity in the powder (Figure 2.4). Preliminary trials were initially performed to

identify the combinations of temperatures, humidities, and residence times to achieve

maximum crystallinity of lactose in whey powders with reasonable processing time. It has

been found that processing condition of 45°C and 35% relative humidity in the first stage,

55°C and 45% relative humidity in the second stage, and 40°C and 40% relative humidity in

the third, allow fluidization to occur through minimizing the chances of bed collapse. These

process conditions also allow to maximize the crystallization rate by maintaining an

adequately high differential temperature (T–Tg) in different stages and to achieve a greater

degree of crystallinity in less time.

The powders were processed at 45°C and 35% relative humidity in the first stage and then

55°C and 45% relative humidity in the second stage and finally 40°C and 40% relative

humidity in the third and final stage. These safe process conditions of temperature and

humidity were below the upper limits on the relative humidity for which whey powder could

be fluidized and avoid cake formation. The residence time in each stage was ten minutes. In

the next step, different rates of powder feeding were used for each fluidization process in

order to assess the effect of feed rate on the crystallization properties of lactose in the

processed whey powder. Therefore, in all experiments, the air temperature, relative humidity,

and residence time were kept constant, and the only variable parameter was the feed rate. The

degrees of amorphicity for the lactose in the processed powders have been investigated by the

techniques described above. Samples were taken after the fluidization process and

immediately used for analytical tests (moisture content and gravimetric moisture analysis);

otherwise they were kept in sealed bags and in a refrigerator to do X-ray diffraction, FTIR,

and Raman spectroscopy on the following day.

Page 66: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

48 | P a g e

4.3 Results and Discussion

Experimental results showed that there were significant improvements in the degree of

crystallinity for the lactose in the spray-dried powder by post-processing in the multi-stage

fluidized bed at different humidities and temperatures with the same residence time, as shown

in Table 4.1. Gravimetric moisture sorption tests were performed on the degree of

crystallinity for the lactose in the spray-dried powders and the processed powders and

compared with X-ray powder diffraction, Fourier transform infrared spectroscopy, and

Raman spectroscopy analyses. The degrees of lactose crystallinity measured using all these

methods confirm that the lactose crystallinity of the processed powders has been increased.

However, there were significant differences in the estimated degree of crystallinity for lactose

by different crystallinity measurement techniques (Table 4.1). Fourier transform infrared

spectroscopy and Raman spectroscopy analyses have been reported to be able to measure the

degree of surface crystallinity of powders, while X-ray diffraction can be used to investigate

the bulk crystallinity in the powders. X-ray diffraction has been reported to have a sensitivity

range of 5% to 10% (Gombas et al., 2002). Therefore, the results of X-ray diffraction are

more precise than other techniques that have been used here, and it measures the bulk

crystallinity rather than surface one.

In scoping experiments of this study was on adjusting the process conditions. However, in the

next stage, project aimed to assess the feasibility of increasing the feed rate. Different rates of

powder feedings from 0.3 kg h-1 to 2.6 kg h-1 were investigated on the crystallization

properties of processed whey powder. The results showed that increasing the rate of powder

feeding did not have a significant and specific effect on the degrees of lactose crystallinity for

the processed powders (according to their moisture sorption peak heights). The heights of

powder bed prior to fluidization were varied from approximately 0.1 cm to 1 cm for 0.3 kg h-

1 to 2.6 kg h-1 powder feeding, respectively. The higher bed height may decrease the amount

of particle vibration and movement during fluidization and therefore higher chance of

agglomeration and subsequent bed de-fluidization. However, the lower bed depth increases

the possibility of fine particles being elutriated from the fluidized bed. Thus, the processed

powders fed with different rates of powder feeding are not expected to have different degrees

of lactose crystallinity since the residence time did not change during the fluidizations. The

results show that processed powder with higher degree of lactose crystallinity was achievable

for different throughputs at these sets of conditions in the fluidized-bed dryer, although no

Page 67: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

49 | P a g e

clear trend was found between the different rates of powder feeding and the degrees of

lactose crystallinity for the processed powders.

Page 68: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

50 | P a g e

Table 4.1 Summary of the experimental results for whey powder processed in the fluidized-bed dryer.

Run Feed rate Sample

Moisture sorption Crystallinity measurement (%)

Bulk

Crystallinity Surface crustallinity

Peak height,

(moisture

content change,

%, kg.kg-1 dry

basis)

Peak

height

change

, %

XRD Raman

FTIR

875

cm-1 900 cm-1

915

cm-1

Run1 50 g/10 min

(0.3 kg h-1)

Spray Dried

(35 w/w %) 4.4

50% ↑

FB -1st stage

FB -2nd stage

FB -3rd stage

Run2 80g /10 min

(0.5 kg h-1)

Spray Dried

(35 w/w %) 3.5

20% ↑

FB -1st stage

FB -2nd stage

FB -3rd stage 2.8

Run3 160 g/ 10min

(1 kg h-1)

Spray Dried

(35 w/w %) 3.3

20% ↑

FB -1st stage

FB -2nd stage

FB -3rd stage 2.7 12.4 39±0.5

Run4 250 g/10 min

(1.5 kg h-1)

Spray Dried

(35 w/w %) 3.5

15% ↑

FB -1st stage

FB -2nd stage

FB -3rd stage 3 10.5 12±1

Run5 440 g/10 min

(2.6 kg h-1)

Spray Dried

(35 w/w %) 3.4

18% ↑

FB -1st stage

FB -2nd stage

FB -3rd stage 2.8 13.0 14±4 32±2 31±0.2 11±1

Run6

440 g/10 min

(2.6 kg h-1)

Repeat of

Run 5

Spray Dried

(35 w/w %) 2.5

40% ↑

FB -1st stage

FB -2nd stage

FB -3rd stage 1.5 17.1 10±3 44±6 56±16 17±1

Page 69: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

51 | P a g e

4.3.1 Moisture Sorption Tests

Processed milk powders showed less moisture sorption and more lactose crystals, leading to a

reduction in the degree of lactose amorphicity for spray-dried milk powder by post-

processing in a series of laboratory-scale fluidized bed dryers. The net percentage changes in

mass (dry basis) as a function of time for whey powder, which has been processed in both

spray- and fluidized beds, have been displayed in Figure 4.2. The peak height (the distance

between the peak and the plateau of the crystallization process) for each powder characterised

the degree of lactose amorphicity for that material. Figure 4.2 has demonstrated the

significant decrease in amorphicity (less sorption) of lactose after processing the powders at

different temperatures and humidities.

The peak height analysis (Table 4.1) showed that the peak heights for spray-dried whey

powders at different runs were 2.5-4.4%. The peak height columns in Table 4.1 showed

progressive decreases in amorphicity of lactose and the lower amount of sorption for all runs

that were caused by different humidities, temperatures, and residence time within different

stages in the fluidized bed, as shown in Figure 4.2 and Table 4.1. By increasing the process

temperature and/or increasing the water activity by sorbing more moisture at a higher relative

humidity in the first two stages, T-Tg increased and the crystallization time decreased,

according to the Gordon–Taylor and Williams–Landel–Ferry equations (Gordon and Taylor,

1952, equation 2.1; Williams et al., 1955, equation 2.3). Therefore, as the results show, by

increasing the temperatures and relative humidities in different stages, the sorption peak

height decreased, suggesting that less amorphous lactose (the most hygroscopic component)

remained in the whey powder.

Page 70: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

52 | P a g e

Figure 4.2 Results of the moisture sorption behaviour of the spray-dried whey powder compared with

whey powder that was processed after being exposed to different conditions in the fluidized bed

(sorption test at 25°C and 75% relative humidity) (Run 1).

4.3.2 X-Ray Diffraction (XRD)

In this study, XRD analysis has been used to evaluate the formation of lactose crystals and

the improvement in the extent of crystallinity for different powders. The locations of the

peaks in the reference data from previous studies have been used to identify various types of

lactose crystals. The peaks for lactose are likely to be found at the following angles: 2θ =

12.5°, 16.4°, 20.0°, and 20.1° for α-lactose monohydrate; 10.5°, 20.9°, and 21.0° for

anhydrous β-lactose; 19.1°, 20.0°, and 20.1° for the mixture of anhydrous α:β with a molar

ratio of 5:3; and 19.5° for the mixture of anhydrous α:β with a molar ratio of 4:1 (Jouppila et

al., 1997; Haque and Roos, 2005).

To define the quantitative determination of crystallinity for whey powder, X-ray

diffractograms of the samples with different amorphous content were recorded. There are two

acceptable methods of X-ray diffraction analysis for quantitative studies; peak height

(intensities) measurement and measurement of the area under the peak. It has been suggested

by Gavish and Friedman (1973) that peak height analysis of X-ray diffraction data is

-4

-2

0

2

4

6

0 200 400 600 800 1000 1200

Moi

stu

re c

onte

nt c

han

ge

(%,

db

)

Time (min)

Spray Dried-WP

FB-1st stage

FB-2nd stage

Spray dried= 4.4%FB-1st Stage= 3.9%FB-2nd Stage= 2.2%

Page 71: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

53 | P a g e

preferable to peak area analysis since both showed a similar level of reproducibility, and the

peak height analysis is easier to perform especially when the curves are symmetrical.

On the diffractogram of the mixture, the intensity of the diffraction (counts/second) for the

individual components is proportional to the quantity of the components in the mixture.

Hence quantification is possible provided that the total amount of sample (and its

preparation) are kept constant from experiment to experiment (Gombas et al., 2002). In this

investigation, spray-dried powder was considered to be 0% crystalline, and physical mixtures

of amorphous and crystalline whey powders were prepared to give different crystalline

contents by mass.

The curves here have been observed to be symmetrical, so the heights of the curves can also

describe the intensity. Based on these references, the peak at 2θ= 20° was taken as the

characteristic peak to measure the intensity when crystalline lactose for assessing the relative

crystallinity of different whey powders (Table 4.2). The intensity of the characteristic peak at

2θ= 20.0° for α-lactose monohydrate with FB processed whey powder (Run 3) was found to

be 196 cps (Figure 4.3), while the peak height for spray-dried powder (considered to be 0%

crystalline), and powder mixtures of 10% crystallinity and 20% crystallinity were 88 cps, 151

cps, and 413 cps, respectively. Comparing the intensity values of the characteristic peak at

2θ= 20.0° for different runs with those for the 10% and 20% crystallinity mixtures of whey

powders suggested that the overall extent of crystallinity for processed whey powders has

been between 10% and 20% within the processes studied here (Run 3) (Figure 4.3).

In order to quantitatively determine the degree of lactose crystallinity for whey powder, the

intensities of the characteristic peaks at 2θ= 20.0° for 0% crystalline powder and mixtures

with 10%, and 20% lactose crystallinity have been fitted to a polynomial equation. A second-

order polynomial trendline (Y= 0.995x2 - 3.65x + 88) at 2θ= 20.0° appeared to fit the data

points closely (Figure 4.3). This function was then applied to estimate the degree of lactose

crystallinity for the spray-dried whey powders and the FB processed powders (Table 4.2).

Page 72: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

54 | P a g e

Table 4.2 XRD intensity values and the estimated degree of crystallinity at chosen 2θ values for the

whey powders used here.

Sample (degree of crystallinity) Peak at 2θ=

12° intensity,

cps

Peak at 2θ= 20°

intensity, cps

Degree of

crystallinity (%)

Spray dried whey powder (0%

Crystallinity)

55 88 0

Whey Powder with 10%

Crystallinity

116 151 10

Whey Powder with 20%

Crystallinity

179 413 20

FB-Processed whey powder Run 3 73 196 12.4

FB-Processed whey powder Run 4 59 159 10.5

FB-Processed whey powder Run 5 94 209 13.0

FB-Processed whey powder Run 6 176 318 17.1

Page 73: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

55 | P a g e

Figure 4.3 X-ray diffraction patterns for spray dried whey powder compared with FB processed whey

powder (Run 3), 10%, and 20% crystalline whey powders.

4.3.3 Estimating the Degree of Surface Crystallinity Using Raman and FTIR

Raman and Fourier Transform Infrared spectroscopy (FTIR) has been reported to be able to

distinguish between amorphous and crystalline lactose (Katainen et al., 2005; Niemelä et al.,

2005; Listiohadi et al., 2009; Islam et al., 2010; Yazdanpanah and Langrish, 2013a). Whey

powder consists mainly of lactose, and therefore Raman and FTIR were used to distinguish

the degree of crystallinity for the spray-dried and fluidized-bed whey powders.

Physical mixtures of whey powders, with an amorphous lactose content of 0-100% (w/w)

(0%, 10%, 20%, 30%, 40%, 50%, and 100%), were prepared to estimate the Raman band

ratio and the FTIR peak heights at 875 cm-1, 900 cm-1, and 915 cm-1 as a function of the

amorphous/crystalline lactose contents. The Raman measurements of the physical mixed

samples are shown in Figure 4.4. It can be seen that, with increasing lactose crystallinity, the

peak height at 476 cm-1 increases as well. The Raman band ratio as a function of the

amorphous content was estimated, and a polynomial second order trendline appeared to fit

the data points with a correlation of determination R2= 0.99, as shown in Figure 4.5. The

function was then used to estimate the degree of lactose crystallinity for the spray-dried whey

0

100

200

300

400

500

600

700

800

900

1000

1100

1200

10 12 14 16 18 20 22 24 26 28 30

Inte

nsit

y (c

ps)

Diffraction angle (2θ)

SPD

10% Crystallinity

FB Processed

20% Crystallinity

Page 74: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

56 | P a g e

powders and the fluidized powders, as shown in Table 4.1. The lactose in the spray-dried

whey powders have a lower degree of lactose crystallinity compared with the fluidized-bed

powders, which indicates that the fluidized-bed process increased the surface crystallinity of

lactose.

Figure 4.4 Raman spectra of whey powder with varying degrees of lactose crystallinity. The intensity

increases at 476 cm-1 with increasing lactose crystallinity. The band areas used for the calculations are

shown in Figure 4.5.

0

500

1000

1500

2000

2500

3000

3500

4000

4500

5000

360 380 400 420 440 460 480 500

Inte

nsi

ty

Raman shift [cm-1]

100% Crystallinity60% Crystallinity40% Crystallinity20% Crystallinity0% Crystallinity

Page 75: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

57 | P a g e

Figure 4.5 Ratio of the Raman band areas as a function of the amorphous lactose content for the

physical mixtures. The correlation was based on the average values of the amorphous lactose content.

The FTIR results for the physical mixtures show an increase in peak height with increasing

lactose crystallinity in the region from 860 cm-1 to 930 cm-1 and at a wavelength of 1260 cm-

1, as indicated in Figure 4.6, which seems to agree with Listiohadi et al. (2009). Further, the

region between 860 cm-1 and 930 cm-1 has been normalized for better comparison and further

analysis, as shown in Figure 4.6. The peak at 1260 cm-1 was not further considered, since it

corresponds to the vibration of the three amide groups in proteins, as indicated by Cai and

Singh (2000). This research focuses on the lactose crystallisation and the characteristic bands

in the region between 860 cm-1 and 930 cm-1 are probably due to configurational and

conformational transformations of the carbohydrate molecules, as mentioned by Listiohadi et

al. (2009).

The FTIR measurements of the physically mixed samples were used to find the peak heights

at wave numbers of 875 cm-1, 900 cm-1, and 915 cm-1, which were plotted as a function of the

amorphous lactose content, as shown in Figure 4.7. The equations of the trendlines fitted

through the data points of Figure 4.8, as well as the R2 values, have been summarised in

Table 4.3. A second order polynomial trendline fitted all the data points for all the different

peak heights well, with an R2 above 0.98.

y = -2E-05x2 - 0.0002x + 0.8445R² = 0.9961

0.6

0.65

0.7

0.75

0.8

0.85

0.9

0 20 40 60 80 100

Ram

an b

and

rat

io

Amorphous lactose content[%]

Page 76: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

58 | P a g e

Figure 4.6 FTIR spectra of whey powders with varying degrees of lactose crystallinity. The

absorbance increases at 875 cm-1 and 900 cm-1 with increasing lactose crystallinity. The region in the

square was used to determine the areas underneath the peaks, as shown in Figure 4.7.

0.00

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.10

800 1,000 1,200 1,400 1,600 1,800

Ab

sorb

ance

Wavelength [cm-1]

100% Crystallinity

50% Crystallinity

20% Crystallinity

0% Crystallinity

Region with significant peaks at 875, 900

and 915cm-1 1260 cm-1

Page 77: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

59 | P a g e

Figure 4.7 FTIR spectra of whey powders normalized at 928 cm-1, with varying degrees of lactose

crystallinity in the region from 860 cm-1 to 930 cm-1.

0.00

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

850 860 870 880 890 900 910 920 930 940

Ab

sorb

ance

Wavelength [cm-1]

100% Crystallinity

50% Crystallinity

20% Crystallinity

0% Crystallinity

875 cm-1900 cm-1

915 cm-1

Page 78: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

60 | P a g e

Figure 4.8 The peak heights at 875 cm-1, 900 cm-1 and 915 cm-1 as function of the amorphous lactose

content. A second order polynomial trendline has been fitted through the data points. The data points

are the averages of five measurements.

Table 4.3 Summary of the trendline equations fitted through the data points of Figure 4.8.

Peak Trendline equation

y=

R2

875 3E-06x2 - 0.0005x + 0.0292 0.99

900 4E-07x2 - 0.0001x + 0.0132 0.98

915 3E-07x2 - 9E-05x + 0.0068 0.99

The degrees of lactose crystallinity in the samples (as measured by FTIR) were estimated

using the peak heights at 875 cm-1, 900 cm-1, and 915 cm-1, since those peaks corresponded to

the characteristic bands for monosaccharides, due to conformational transformations of the

carbohydrate molecules as indicated by Listiohadi et al. (2009). In addition, the results have

been compared with the Raman method (Table 4.1). The results show that the degrees of

lactose crystallinity estimated at the peak height 915 cm-1 are similar to the degrees of lactose

crystallinity estimated with the Raman method. On the other hand, the degrees of lactose

0.00

0.01

0.02

0.03

0 20 40 60 80 100

Pea

k H

eigh

t

Amorphous lactose content [wt%]

▲ 875 cm-1

■ 900 cm-1

♦ 915 cm-1

Page 79: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

61 | P a g e

crystallinity estimated at the peaks of 875 cm-1 and 900 cm-1 are approximately three times

larger compared with Raman and the degree of lactose crystallinity estimated at the peak 915

cm-1. Overall, the Raman and FTIR results indicated that the fluidized-bed process increased

the surface lactose crystallinity of the whey powders, since the spray-dried powders (0%

lactose crystallinity) appeared to have a lower degree of lactose crystallinity compared with

the fluidized-bed powders.

The spray-dried powders showed a larger sorption peak height compared with the fluidized-

bed processed powders, which indicates that the spray-dried lactose in the powders are more

amorphous lactose compared with the fluidized-bed powders. The trend from the moisture

sorption test appears to have been confirmed by the FTIR and Raman results, where the

spray-dried whey powders have a lower surface lactose crystallinity compared with the

fluidized-bed powders.

4.3.4 Particle Size Analysis

The particle size distribution was measured for spray-dried and processed whey powders.

Each test was carried out two or three times, and the average mean particle size was reported.

Particle size analysis of processed materials has shown some particle agglomeration for the

processed powders compared with the spray-dried whey powders. The agglomeration and

size enlargement was observed for all types of processed powders due to processing at high

humidity conditions that caused particles to stick together during the process. The particles

have been enlarged from the size of 7 ± 1 μm (D[3,2], Sauter mean diameter (SMD) for

amorphous whey powder) to 59 ± 14 μm (D[3,2] for a processed powder). The agglomeration

and particle enlargement was not necessarily the result of crystallization of lactose in

powders; however agglomeration was not avoidable in this technique due to processing by

high-humidity air and at a temperature above the glass-transition temperature of the powders.

The agglomerated powders might result in better dissolution than the spray-dried powder

since they could sink and disperse easier in water despite the higher degree of lactose

crystallinity compared with that of spray-dried powders. The bulk densities (ρb) of the

powders, which was the mass of the powders divided by the total volumes they occupied for

the spray-dried and processed whey powders, were 665 ± 5 kg m-3 and 620 ± 10 kg m-3,

respectively.

Page 80: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

62 | P a g e

4.3.5 Surface Morphology

Microscopy has also been applied to obtain information on the lactose crystallinity of

powders. SEM micrographs show some morphological changes for whey powders after

processing in the fluidized beds at different humidities and temperatures. Crystals appeared to

be formed on the surface and inside the particles. Figure 4.9 shows the whey powder before

and after processing (crystallization) in a fluidized-bed dryer. The surface of the spray-dried

whey powder (produced by conventional spray drying) appeared to be amorphous (Figure 4.9

A), while the processed whey powder had a heavily textured appearances, suggesting that

part of the particle (the lactose) is crystalline (Figure 4.9 B:D).

The processed powders remained free-flowing at room temperature and humidity for at least

three days (Figure 4.10 C, D), while the raw powder has created a solid cake and become

stuck in the Petri dish (Figure 4.10 A, B). The room conditions during the three-day test were

19–25°C and 55–65% relative humidity. Cake formation and agglomeration are other

physical properties that are strongly affected by the surface structure of the particles, so the

crystalline surfaces developed by this experiment prevented bridge formation between

particles due to more hygroscopic amorphous lactose. The crystal texture on the surface

appeared to preserve the particle core from deteriorative changes and also appeared to

enhance the powder's physical properties.

Page 81: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

63 | P a g e

Figure 4.9 A) Surface of spray-dried whey powder (amorphous) produced in a conventional spray-

drying process at a high magnification. B:D) Morphological structure of processed whey powder in

the fluidized bed.

A B

C D

Page 82: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

64 | P a g e

Figure 4.10 Whey powder after three days at ambient condition, A and B) Spray-dried powder before and

after test, respectively (A hard cake was formed (B)). C and D) Processed powder before and after test,

respectively (still free-flowing (D)) (Run4).

4.4 Conclusions

This chapter has reported the effects of operating conditions on the crystallization behaviour of

whey powders. Powder was fed at different rates to the continuous multistage fluidized bed,

where each stage had particular temperatures and relative humidities. Gravimetric moisture

sorption tests were performed on the processed powders to assess the degree of lactose

crystallinity and compared using X-ray Powder Diffraction, FTIR, and Raman spectroscopy

analyses. The results showed significant improvements in the degree of lactose crystallinity for

the powders. Experimental results showed that there were significant changes in lactose

crystallinity between samples that were processed (crystallized) at different humidities and

Page 83: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

65 | P a g e

temperatures, and increasing the rate of powder feeding did not have a significant effect on the

degrees of lactose crystallinity for the processed powders. The results of this study help when

choosing the processing conditions to produce and control lactose crystallization in

carbohydrate-containing dairy powders during spray and fluidized-bed processing.

Page 84: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

66 | P a g e

CHAPTER 5 EFFECT OF ACIDITY ON IN-PROCESS

CRYSTALLIZATION OF LACTOSE DURING SPRAY

DRYING1

Spray drying of acid whey is a significant problem for the dairy industry due to its high content

of lactic acid. The main ingredients of acid whey are lactose, protein, and lactic acid. Due to the

importance of lactose in the food and pharmaceutical industries, there have been extensive

studies on the crystallization behavior of lactose (Roos and Karel, 1992; Jouppila et al., 1998).

Studying the effect of lactic acid on the crystal growth of lactose is necessary due to its

fundamental and commercial importance in food processing. However, there is little known

about the lactose crystallization behavior of lactose in the presence of lactic acid. Bhandari

(2008) found that lactic acid significantly influenced the crystallization of lactose during drying

by decreasing glass-transition temperature of lactose and the depression of the crystallization

temperature. He also found that adding lactic acid to the lactose solutions decreased the yield the

spray-drying process. It has been found that the presence of proteins at low concentrations (< 5

w/w %) significantly affects the product yield and crystallinity (Haque and Roos, 2004). The

crystallization kinetics of lactose/protein solutions, especially when lactic acid is added to the

solution, for example acid whey, have not been studied in terms of in-process crystallization

during spray drying. It is difficult to predict the performance of the crystallization for

commercial acid whey due to the presence of other impurities. Therefore in one series of

experiments in this paper, only solid-phase crystallization of lactic acid and lactose was carried

out. Then further experimentation has been done on the crystallization behavior of powders

produced during spray drying using solutions of lactose/protein/lactic acid to imitate the

composition of typical acid whey solutions. Different methods have been used to assess the

degree of crystallinity for the spray-dried products produced with different concentrations of

lactic acid.

1 The content of this chapter has been published as the following paper: Saffari, M. & Langrish, T. (2014), Effect of lactic acid on in-process crystallization of lactose/protein powders during spray drying. Journal of Food Engineering, 137, 88-94.

Page 85: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

67 | P a g e

5.1 Materials and Methods

5.1.1 Sample Preparation

One series of experiments was carried out by varying the composition of the solution with

additions of whey protein isolate from 1 to 10% (w/w), whilst maintaining the level of lactose

concentration in solution at 10% (w/w). Further experimentation has been done by adding lactose

and whey protein isolate to create solutions with 15% (w/w) lactose and 5% (w/w) WPI (lactose

lactose/WPI ratios of 75:25 (w/w)), and then varying the composition of the solution with

additions of lactic acid from 1% to 20% (w/w). Each solution was made up to a total weight of

100 g. The composition of the solution was used at that ratio, in order to imitate the typical

composition of whey solution(s) regarding the amounts of sugar and protein (Panesar et al.,

2007; Jelen, 2009). All solutions were magnetically stirred at the room temperature of 25oC for

at least 30 minutes, so a clear solution was obtained without any visible crystals being present.

The clear solutions were then spray dried. The pH of the solutions was measured with a pH

electrode, InPro 3250 series (Mettler Toledo, M 300, Switzerland).

5.1.2 Operating Conditions for Spray Drying

The solutions were spray dried using a Buchi-B290 Mini Spray Dryer with an inlet gas

temperature of 180oC, an aspirator setting of 100% (38 m3 h-1), a pump rate of 25% (8 mL min-1),

and a nozzle air flow rate of 470 L h-1 (40 on the nozzle rotameter scale). These operating

conditions caused the outlet gas temperatures to be 108 ± 3oC and 112 ± 4oC for lactose and

lactose/WPI solutions, respectively. All experiments were performed in triplicate. Freshly spray-

dried powder was collected from a collection vessel at the bottom of a cyclone and was

immediately used for analytical tests (moisture content, MDSC analysis and gravimetric

moisture analysis); otherwise the powders were kept in sealed bags and in a refrigerator for X-

ray diffraction tests on the following day.

Page 86: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

68 | P a g e

5.2 Results and Discussion

5.2.1 Yield from Spray Drying

It can be seen, in Tables 5.1 and 5.2 and Figure 5.1, that adding lactic acid to the lactose and to

the lactose/WPI solutions decreased the yield (or recovery) of the spray-drying process, which

agreed well with the result of Bhandari (2008). This decrease in the yield can be attributed to

stickiness during the spray-drying process. A key explanation for the steady downward trend in

the yield can be seen in the work of Hanus and Langrish (2007). In their study, it was found that

direct deposition of very wet particles on the dryer walls is a very significant phenomenon in

small-scale spray dryers, which may be linked with the limited drying time that is possible in the

equipment and corresponding increase in moisture content of product when increasing the lactic

acid concentration (Figure 5.2). Another reason (not mutually exclusive) for this yield behavior

might also be due to the powders having lower glass-transition temperatures (Jouppila and Roos,

1994a; Roos, 1995; Bhandari and Howes, 1999). This situation means that, as the lactic acid

concentrations increase, the glass-transition temperatures decrease, making the particles stickier

(Table 5.1). The statistical analysis of the yield data showed that even adding 1% (w/w) of lactic

acid to the lactose solution significantly decreased the yield of the process to 60%. However,

with increasing the lactic acid content to more than 5% (w/w), producing dry powder was

impossible (Table 5.1). However, adding 2% (w/w) lactic acid to the lactose/WPI solution did

not have a significant effect on the yield of the process, whereas increasing the amount of lactic

acid concentration to 15% (w/w) significantly decreased the yield of the process to 26% (Table

5.2). The higher yield of the spray-drying process for lactose/WPI solutions than lactose

solutions with the same concentration of lactic acid can be explained by the decrease in wall

deposition during spray drying due to the surface coverage of proteins, which remain non-sticky

due to their higher glass-transition temperatures (Adhikari et al., 2009; Wang and Langrish,

2010).

Page 87: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

69 | P a g e

Figure 5.1 Effect of the lactic acid concentration on the yield of the spray-dried products.

Figure 5.2 Effect of the lactic acid concentration on the moisture contents of the spray-dried products.

10

20

30

40

50

60

70

80

90

0 2 4 6 8 10 12 14 16

Yie

ld, %

Lactic acid concentration (% w/w)

lactose-Lactic acid solutions

lactose-WPI-lactic acid solutions

0

1

2

3

4

5

6

7

8

9

0 2 4 6 8 10 12 14 16

Moi

stu

re C

onte

nt,

%

Lactic acid concentration (% w/w)

lactose-Lactic acid solutions

lactose-WPI-lactic acid solutions

Page 88: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

70 | P a g e

Table 5.1 Summary of the experimental conditions used in spray drying for lactose-lactic acid solutions (lactose 10% w/w), inlet gas temperature

of 180oC and outlet gas temperature of 108 ± 3oC.

Analysis Lactic acid concentration (%, w/w)

0 1 3 5 10

Spray-dried powder

Yield (%, w/w) 71±2a 60±2b 44±1c 22±3d *NA

Moisture content,% (dry basis) 1.3±0.1a 2.7±0.5b 7.4±0.8c 8.1±0.5c *NA

Solution pH 6.7±0.1a 2.53±0.01b 2.24±0.01c 2.11±0.03d 1.91±0.01e

Sorption test

Peak height, (moisture content change, %, kg.kg-1 dry basis of

lactose) 6±1a 5.5±0.4a 4±1b 0.7±0.3c *NA

Modulated differential scanning calorimetry

Glass transition temperature (oC)

82±8a 64±2b 40±10c 20±10d *NA

Crystallization temperature (oC)

153±9a 142±7a 90±10b 69±6c *NA

Latent energy of Crystallization (J/g)

89±2a 60±10b 46±4b 30±4c *NA

X-ray diffraction Intensity (cps) 140±10a 270±20b 440±20c 570±50d *NA

All values are mean ±standard deviation of at least two replicate analyses. For each property, means in a row followed by different letters (a-e) are significantly different at p= 0.05. *NA: not available

Page 89: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

71 | P a g e

Table 5.2 Summary of the experimental conditions used in spray drying for lactose-WPI-lactic acid solutions (lactose 15% w/w) + (WPI 5%

w/w), inlet gas temperature of 180oC and outlet gas temperature of 112 ± 4oC.

Analysis Lactic acid concentration (%, w/w)

0 1 2 7 10 15 20

Spray-dried powder

Yield (%, w/w) 81±5a 77±3a 75±6a 58±3b 42±1c 26±7d *NA

Moisture content,% (dry basis)

1.2±0.2a 1.2±0.3a 3.5±0.2b 5.1±0.3c 5.9±0.6d 6.5±0.1e *NA

Solution pH 6.7±0.1a 3.7±0.3b 3.2±0.2c 2.5±0.1d 2.3±0.1e 2.1±0.1f 1.98±0.08g

Sorption test

Peak height, (moisture content change, %, kg.kg-1

dry basis of lactose) 6±1a 5.6±0.4b 4.8±0.3b 4.4±0.6b 4.5±0.5b 4.3±0.4b *NA

X-ray diffraction Intensity (cps) 80±10a 94±4a 150±20b 160±20b 300±20c 380±70d *NA

All values are mean ±standard deviation of at least two replicate analyses. For each property, means in a row followed by different letters (a-g) are significantly different at p= 0.05. *NA: not available

Page 90: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

72 | P a g e

5.2.2 Assessment of the Particle Crystallinity by Moisture Sorption Tests

Figure 5.3 illustrates the moisture sorption behavior (moisture content change, %, kg.kg-1 dry

basis of lactose) for spray-dried products of lactose solutions at different lactic acid

concentrations, which shows a gradual improvement in lactose crystallinity through less moisture

sorption, representing lower degrees of lactose amorphicity in the powders. There were

significant changes between different spray-dried lactose powders with different lactic acid

concentrations. The first curve (highest peak) is the moisture sorption by powders from the spray

drying of pure lactose solutions (0%, w/w), which were mostly amorphous. The highest peak

shows the greatest amorphicity of lactose compared with the other spray-dried powders. The

other curves (for example, Figure 5.3, curve 4) show a significant reduction in moisture sorption

by spray-dried powder from lactose solutions with 5% (w/w) lactic acid.

Increasing the lactic acid concentration in the lactose solutions significantly decreases the degree

of lactose amorphicity in the spray-dried powders, which leads to lower peak heights. The

sorption peak heights for lactose can be altered from 6% ± 1 % to about 0.7% ± 0.3%. The

statistical analysis of the peak height data showed that increasing the lactic acid concentration

from 1% to 5% (w/w) significantly decreased the peak heights for spray-dried powder (Figure

5.4).

Page 91: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

73 | P a g e

Figure 5.3 Moisture sorption behaviour of the spray-dried products from lactose for different lactic acid

concentrations (the sorption test environment was at a temperature of 25oC and 75% relative humidity).

The net percentage changes in mass (dry basis) as a function of time for spray-dried lactose/WPI

mixtures with different concentrations of lactic acids are displayed in Figure 5.5, which

demonstrates the significant decrease in lactose amorphicity (less sorption) when increasing the

lactic acid concentrations in the lactose/WPI solutions. The statistical analysis of the moisture

sorption behaviour also showed that adding lactic acid to the lactose/WPI solution had a

significant effect on the peak heights of the spray-dried powders, where increasing the amount of

lactic acid concentration from 1% to 15% (w/w) significantly decreased the peak heights from

4.6% ± 0.3% to 2.9% ± 0.6% (Table 5.2).

The sorption peak heights for lactose can be altered from 6% ± 1% to about 0.7% ± 0.3%,

compared with 6% ± 1% to 4.3% ± 0.4% for the spray-dried lactose/WPI mixtures. The presence

of proteins during the spray drying of the lactose/WPI mixtures is likely to reduce the extent of

crystallization for lactose during spray drying.

-4

-2

0

2

4

6

8

0 100 200 300 400 500 600

Moi

stu

re c

onte

nt

chan

ge (

%)

Time (min)

0 (%, w/w)

1 (%, w/w)

3 (%, w/w)

5 (%, w/w)

Page 92: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

74 | P a g e

Figure 5.4 Sorption peak heights of the spray-dried products from lactose solutions for different lactic

acid concentrations.

Figure 5.5 Sorption peak heights of the spray-dried products from lactose-WPI solutions for different

lactic acid concentrations.

5.2.3 Further Assessment of the Particle Crystallinity by MDSC

The MDSC results of spray-dried powders from lactose solutions with different lactic acid

concentrations showed very clear phase transitions during the thermal analysis, while those for

lactose/WPI solutions were complicated to analyze due to multiple components being present in

0

1

2

3

4

5

6

7

8

0 1 2 3 4 5

Sor

pti

on p

eak

hei

ght,

(m

oist

ure

co

nte

nt

chan

ge, %

, kg.

kg-

1 d

ry b

asis

of

lact

ose)

Lactic acid concentration (% w/w)

2

3

4

5

6

7

8

0 2 4 6 8 10 12 14 16

Sor

pti

on p

eak

hei

ght,

(m

oist

ure

co

nte

nt

chan

ge, %

, kg.

kg-

1 d

ry b

asis

of

lact

ose)

Lactic acid concentration (% w/w)

Page 93: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

75 | P a g e

the particles (Figure 5.6). The presence of protein, fat, additives, and mineral contents in whey

isolate powder made it difficult to distinguish each transition in the thermogram separately

(Vuataz, 2002), because the transitions of each component may have overlapped with those of

lactose.

The MDSC results (Figure 5.7) suggest that increasing the lactic acid concentration in the lactose

solutions decreased the glass transition and crystallization temperatures of the spray-dried

powders. There is a downward trend in the latent energy of crystallization for the spray-dried

products with increasing acid concentrations, which indicates a decrease in the amorphous

content of the spray-dried powders. The downward trends of glass-transition temperatures with

increasing lactic acid concentrations are similar to the results of Bhandari (2008). The glass-

transition temperature of pure lactose was 82 ± 8oC, and the crystallization peak appeared at 153

± 9oC, with the exothermic heat of crystallization being 89 ±2 J/g. The corresponding figures for

spray-dried powders from lactose solutions with 5% (w/w) lactic acid concentrations

significantly decreased to 20 ± 10oC, 30 ± 4oC, and 69 ± 6 J/g, respectively (Table 5.1).

The statistical analysis of the MDSC results showed that increasing the lactic acid concentration

from 1% to 5% (w/w) significantly decreased all of these values (Table 5.1) (the glass-transition

temperature, the crystallization temperature, and the latent heat of crystallization).

Page 94: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

76 | P a g e

Figure 5.6 Thermal transitions during MDSC scans of the spray-dried products from lactose solutions for

different lactic acid concentrations.

The results from all these methods agree well with the results from the moisture sorption tests

and suggest that the crystallinity of the lactose powders increased with greater lactic acid

concentrations. The results suggested that, when the lactic acid concentrations increased, less

amorphous lactose (the most hygroscopic component) remained in the spray-dried powder,

which could be due to the greater temperature difference (T-Tg) and consequently the decrease in

the crystallization time, according to the Gordon–Taylor and Williams–Landel–Ferry equations

( 2.1, Gordon and Taylor, 1952; equation 2.3, Williams et al., 1955).

-1

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

0.8

0 50 100 150 200 250

Hea

t F

low

(w

/g)

Temperature (°C)

0% (w/w)1% (w/w)3%(w/w)5%(w/w)

Endothermic transition associated with Melting

High exothermic peak, hence higher heat of crystallization indicates more amorphous lactose

Glass-transition temperature

Page 95: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

77 | P a g e

Figure 5.7 Effect of lactic acid concentration on the glass-transition temperatures, the crystallization

temperatures, and the latent energies of crystallization for the spray-dried products from lactose solutions.

5.2.4 X-Ray Diffraction (XRD)

In this study, XRD analysis has been used to evaluate the formation of lactose crystals and the

improvement in the extent of crystallinity for different powders. The different types of lactose

crystals were identified by the location of peaks in the reference data of previous studies. The

most noted representative peaks for crystalline carbohydrates have been assigned at 2θ= 12.5°,

16.4°, 20.0°, and 20.1° for α-lactose monohydrate; 10.5°, 20.9°, and 21.0° for anhydrous β-

lactose; 19.1°, and 21.1° for the mixture of anhydrous α:β with a molar ratio of 5:3; and 19.5° for

mixture of anhydrous α:β with a molar ratio of 4:1 (Simpson et al., 1982; Drapier-Beche et al.,

1997; Jouppila et al., 1997; Haque and Roos, 2005). When the curves are symmetrical, the height

of the curves can also describe the intensity. In this investigation, based on these references, the

peak at 2θ= 20.1° was taken as the characteristic peak to measure its intensity for finding the

relative crystallinity of different powders (Tables 5.1 and 5.2). The diffraction patterns for spray-

dried lactose with different lactic acid concentrations show clear sharp peaks associated with α-

lactose monohydrate at high intensities (2θ= 12.5°, 20.0°, and 20.1°). There is no distinguishable

peak for anhydrous β-Lactose, and the peak intensities for the mixture of anhydrous α:β lactose

0

20

40

60

80

100

120

140

160

0

20

40

60

80

100

0 1 2 3 4 5

Gla

ss-t

ran

siti

on t

emp

erat

ure

,0 CC

ryst

alli

zati

on t

emp

erat

ure

, 0 C

Hea

t of

cry

stal

liza

tion

(J/

g)

Lactic acid concentration (%w/w)

Heat of crystallization

Crystallization temperature

Glass-transition temperature

Page 96: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

78 | P a g e

with different molar ratios at 19.1°and 19.5° are relatively low. Table 5.1 and Figure 5.8 show

some improvement in the degree of crystallinity for lactose powders with increasing lactic acid

concentrations. The intensity of the characteristic peak at 2θ= 20.1° for the spray-dried lactose

without lactic acid was found to be 140±10 cps while those for the spray-dried lactose with 1%

(w/w), 3% (w/w), and 5% (w/w) lactic acid concentrations were 270 ± 20 cps, 440 ± 20 cps, and

570 ± 50 cps, respectively. Simply comparing the intensity values of the characteristic peak at

2θ= 20.1° for different lactic acid concentrations suggests that the crystallinity of the lactose

increased by increasing the lactic acid concentrations in lactose solutions. The statistical analysis

of the XRD results showed that adding lactic acid to the lactose solutions had a significant effect

by increasing the crystallinity of lactose (Table 5.1).

For lactose/WPI feed-solutions, Table 5.2 and Figure 5.9 show that adding lactic acid increased

the crystallinity of lactose powders to some extent. This increase was considerable when more

than 5% (w/w) lactic acid was added to the solution. The diffraction patterns are similar to those

of spray-dried lactose, which show clear sharp peaks associated with α-lactose monohydrate at

high intensity (2θ= 20.0° and 20.1°) and indicate that most of the lactose crystals are in the α-

monohydrate form. There is no distinguishable peak for anhydrous β-lactose, and the peak

intensities for the mixture of anhydrous α:β anomers with different molar ratios at 19.1° and

19.5° are relatively low. The statistical analysis of the XRD results also showed that adding

lactic acid to the lactose/WPI solution had a considerable effect on increasing the intensity values

of spray-dried powder, where increasing the lactic acid concentration from 1% to 15% (w/w)

significantly increased the intensity values from 80 ± 10 cps to 380 ± 70 cps (Table 5.2).

Page 97: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

79 | P a g e

Figure 5.8 X-ray diffraction patterns of the spray-dried products from lactose solutions for different lactic

acid concentrations.

0

200

400

600

800

1000

1200

10 12 14 16 18 20 22 24

Inte

nsi

ty, c

ps

Angle, 2θ

5%(w/w)

3%(w/w)

1%(w/w)

0%(w/w)

Page 98: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

80 | P a g e

Figure 5.9 X-ray diffraction patterns of the spray-dried products from lactose-WPI solutions for different

lactic acid concentrations.

5.3 Conclusions

This chapter describes experiments in which lactic acid, in various proportions, was added to

lactose and lactose/WPI solutions to investigate the effect of acidity on lactose crystallization.

Higher concentrations of acid appear to increase the extent of lactose crystallization during spray

drying. The sorption peak heights for lactose were altered from 6% ± 1% to about 0.7% ± 0.3%

for lactose/lactic acid mixtures and from 6% ± 1% to 4.3% ± 0.4% for spray-dried

lactose/WPI/lactic acid mixtures. However, the yield from spray drying lactose/lactic acid was

significantly decreased at higher concentrations of acid, and the yield for lactose/WPI/lactic acid

solutions was higher than lactose solutions with the same concentration of lactic acid, as

expected, due to the surface coverage of proteins when using WPI. The degrees of crystallinity

measured using MDSC and XRD methods agreed well with the results of the moisture sorption

tests and confirm that the crystallinity of the lactose increased with increasing lactic acid

0

100

200

300

400

500

600

700

800

900

1000

10 12 14 16 18 20 22 24 26

Inte

nsi

ty, c

ps

Angle, 2θ

15%(w/w)10%(w/w)7%(w/w)2%(w/w)1%(w/w)0%(w/w)

Page 99: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

81 | P a g e

concentration. This increase in crystallinity may be due to increasing the T-Tg difference (with

the low Tg of lactic acid) and consequently decreasing the crystallization time, according to the

Gordon–Taylor and Williams–Landel–Ferry equations ( 2.1 and 2.3, respectively). The results of

this study have implications for choosing processing conditions to control crystallization in

carbohydrate-containing dairy powders and also powders of acid-rich foods, such as fruit juices,

during spray and fluidized-bed processing.

Page 100: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

82 | P a g e

CHAPTER 6 A NEW TEMPLATING APPROACH FOR HIGH-

POROSITY SPRAY-DRIED PARTICLE PRODUCTION2

In this chapter, a new production process is used to create highly-porous powder templates

through the spray drying of lactose solutions containing ethanol-soluble food-grade acids, such

as lactic acid, citric acid, boric acid, and ascorbic acid, as templating agents and then removing

these acids by ethanol washing of the spray-dried powders. This chapter has also investigated the

effect of different concentrations for various templating acids on the porosity of the final

products from a new templating approach, and the extents of in-process crystallization for the

lactose/templating acids powders have been studied during spray drying.

It has been found in Chapter 5 that significant increases in the degrees of crystallinity for the

final spray-dried products of lactose solutions have occurred at increasing lactic acid

concentrations, and the yields from spray drying have also been significantly decreased at higher

concentrations of lactic acid. This increase in crystallinity may be due to increasing the T-Tg

difference (with the low Tg of lactic acid) and consequently decreasing the crystallization time,

according to the Gordon–Taylor and Williams–Landel–Ferry equations. The increase in the rate

of lactose crystallization at low pH may also be connected with the increase in the mutarotation

rate and the rate of orientation of lactose molecules into the crystals (Twieg and Nickerson,

1968; Nickerson and Moore, 1974; Jenness, 1988; Singh et al., 1991).

The effect of glass-transition temperature and strength of the acid (pH of the solution) on the

degree of crystallinity for lactose in spray-dried powders is critical, with implications for the

performance of the templating approach using acidic templates. Therefore, the aim of this

research is to investigate the effect of using different concentrations for various food-grade acids,

such as lactic acid, citric acid, boric acid, and ascorbic acid, on the degree of crystallinity for

spray-dried lactose powders and the porosity of the ethanol-washed powders. The results of this 2 The content of this chapter has been published as the following paper: Saffari, M., Ebrahimi, A. & Langrish, T.A.G. (2015b), The Role of Acidity in Crystallization of Lactose and Templating Approach for Highly-Porous Lactose Production, Journal of Food Engineering, 164, 1-9.

Page 101: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

83 | P a g e

study have implications in choosing the best processing conditions to control the extent of

lactose crystallization for producing spray-dried powders with high concentrations of surface

asperities and large specific surface areas in order to increase the dissolution rates of food

products, such as food infusions, and also in flavoring or encapsulating processes.

6.1 Materials and Methods

6.1.1 Sample Preparation

In order to investigate and report the effects of different templating acid concentrations on the

BET surface areas of lactose, experiments were carried out by varying the composition of the

solutions with different amounts of templating acids (citric acid, ascorbic acid, boric acid, and

lactic acid), whilst maintaining the level of lactose concentration in solution at 10% (w/w). Each

solution was made up to a total weight of 100 g. A magnetic stirrer was used to enhance the

dissolution rate of lactose at the room temperature of 25oC for at least 30 minutes, so that clear

solutions were obtained without any visible crystals being present. The clear solutions were then

spray dried. The pH of the solutions was measured with a pH electrode. Further experimentation

has been done by adding whey protein isolate (0.2%, w/w) to 10% (w/w) lactose solutions to

increase the yield of the process (Islam et al., 2013) and also by controlling the degree of

crystallinity and varying the compositions of the solutions with different additions of templating

acids.

6.1.2 Operating Conditions for Spray Drying

A BüchiB-290 spray dryer has been used in the experiments. The inlet air temperature has been

180oC, the main air flow rate through the dryer has been 38 m3 h-1 (aspirator setting of 100%),

the pump rate has been 8 mL min-1 (25% of the maximum rate), and the nozzle air flow rate has

been 470 L h-1 (40 on the nozzle rotameter scale). Freshly spray-dried powder was collected

from a collection vessel at the bottom of a cyclone and was immediately used for analytical tests

(moisture content, modulated differential scanning calorimetry (MDSC), and gravimetric

moisture analysis); otherwise, the powders were kept in sealed bags and in a refrigerator for X-

Page 102: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

84 | P a g e

ray diffraction and scanning electron microscope (SEM) tests on the following day. Part of the

powder collected from the vessel at the bottom of the cyclone has been washed with ethanol for

24 h at the room temperature of 25oC to remove the templating acid, filtered under vacuum, and

dried for one hour in a laboratory oven at 60oC. Crushing, grinding, and sieving the dried powder

mass was used to produce the final powders. For each concentration, experiments have been

done three times, and at least two replicate analyses have been done on samples. Data in this

study have been presented as means ± STD (standard deviation) and were obtained from at least

three independent experiments. Figure 6.1 illustrates the overall process.

Droplet Templating agent

Water evaporation

Template removed with ethanol washing

Core material

Templating agent

Crystalline templated porous particle

Amorphous spray-dried particle

Core material solution with different concentrations of

templating agent

Cyclone

Air outlet fanDrying chamber

Air heater Gas flow

Production collection

vessel

Figure 6.1 Schematic diagram of the experimental setup and an illustration of the templating process.

Page 103: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

85 | P a g e

6.2 Results and Discussion

6.2.1 Yield from Spray Drying

Experimental results showed that the addition of templating acids to the lactose solutions

decreased the yield (or solids recovery) from the spray-drying process. At higher concentrations

of templating acids, it has also been found that the moisture contents of the particles are higher,

as shown in Tables 6.1-6.5. These trends in yield and moisture content have also been observed

in Chapter 5. The main reason for this yield behavior is likely to be due to having lower glass-

transition temperatures in the powders and also greater product moisture contents at higher

templating-acid concentrations (Jouppila and Roos, 1994a; Roos, 1995; Bhandari and Howes,

1999). This situation means that, as both the templating-acid concentration and the product

moisture contents increase, the glass-transition temperatures decrease, making the particles

stickier (Tables 6.1-6.5). Another reason for the steady downward trend in the yield can be

attributed to direct deposition of very wet particles on the dryer walls, which is a very significant

phenomenon in small-scale spray dryers (Hanus and Langrish, 2007).

It can be seen in Tables 6.1-6.5 that even adding 1% (w/w) of lactic acid to the lactose solution

significantly decreased the yield of the process from 71% ± 2% to 60% ± 2%, and the yield was

reduced to 51% ± 2% by adding only 1% (w/w) citric acid to the lactose solution. Likewise, in

the case of using ascorbic acid as a templating acid, increasing the ascorbic acid concentration to

1% (w/w) significantly decreased the yield of the process to 35% ± 4%. By increasing the

ascorbic acid content to more than 2% (w/w), producing dry powder was impossible (Table 6.3).

The use of boric acid at concentrations of about 3% (w/w) did not have a significant effect on the

yield of the process, suggesting that boric acid is the best templating material for powder

recovery. However, increasing the boric acid concentration to 7% (w/w) decreased the yield of

the process to 56% ± 5% (Table 6.5).

In order to increase the yield of the spray-drying process, 0.2% (w/w) whey protein isolate has

been added to 10% (w/w) lactose with different concentrations of ascorbic acid. The yield of the

spray-drying process can be increased from 35% ± 4% to about 87% ± 2% for the spray-dried

lactose/WPI mixture with 1% (w/w) ascorbic acid (Table 6.4). The effect of WPI in creating high

yields with potentially stickier core materials (with more citric acid) may be due to the reduction

Page 104: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

86 | P a g e

in wall deposition caused by the WPI coating the particle surfaces, with the WPI having higher

glass-transition temperatures and hence being less sticky than lactose, ascorbic acid, or their

mixtures (Adhikari et al., 2004; Wang and Langrish, 2010; Islam et al., 2013).

Table 6.1 Summary of the experimental conditions used in spray drying for lactose- lactic acid solutions

(lactose 10% w/w), inlet gas temperature of 180ºC and outlet gas temperature of 108 ± 3ºC (The data

were mainly presented earlier in Table 5.1).

Analysis Lactic acid concentration (%, w/w)

0 1 3

Spray-dried powder

Yield (%, w/w) 71±2 60±2 44±1

Moisture content

(% dry basis) 1.3±0.1 2.7±0.5 7.4±0.8

Solution pH 6.69±0.09 2.53±0.01 2.24±0.01

Sorption test

Peak height

(moisture content change, %, kg.kg-1 dry basis of

lactose)

6±1 5.5±0.4 4±1

Modulated differential scanning calorimetry

Glass-transition temperature (oC)

82±8 64±2 40±10

Crystallization temperature (oC)

153±9 142±7 90±10

Latent energy of crystallization (J/g)

89±2 60±10 46±4

X-ray diffraction Intensity (cps) 140±10 270±20 440±20

BET of ethanol-washed powder

Surface area (m2 g-1) 0.3±0.2 14.9±0.9 9.5±0.8

All values are means ±standard deviations of at least three replicate analyses.

Page 105: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

87 | P a g e

Table 6.2 Summary of the experimental conditions used in spray drying for lactose-citric acid solutions

(lactose 10% w/w), inlet gas temperature of 180ºC and outlet gas temperature of 113 ± 2ºC.

Analysis Citric acid concentration (%, w/w)

0 1 3

Spray-dried powder

Yield (%, w/w) 71±2 51±2 16±4

Moisture content

(% dry basis) 1.3±0.1 1.7±0.2 2.7±0.1

Solution pH 6.69±0.09 2.4±0.1 2.12±0.02

Sorption test

Peak height

(moisture content change, %, kg.kg-1 dry basis of

lactose)

6±1 5.4±0.5 2.9±0.6

X-ray diffraction Intensity (cps) 140±10 290±20 410±20

BET of ethanol-washed powder

Surface area (m2 g-1) 0.3±0.2 20.8±0.9 12.3±0.8

All values are means ±standard deviations of at least three replicate analyses.

Page 106: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

88 | P a g e

Table 6.3 Summary of the experimental conditions used in spray drying for lactose-ascorbic acid

solutions (lactose 10% w/w), inlet gas temperature of 180ºC and outlet gas temperature 112 ± 2ºC.

Analysis Ascorbic acid concentration (%,

w/w) 1 2

Spray-dried powder

Yield (%, w/w) 35±4 10±2

Moisture content

(% dry basis) 1.8±0.2 1.9±0.1

Solution pH 2.83±0.04 2.62±0.01

Sorption test

Peak height (moisture content change, %, kg.kg-1

dry basis of lactose) 5.9±0.9 4.8±0.5

Modulated differential scanning calorimetry

Glass-transition temperature (oC)

54±4 45±2

Crystallization temperature (oC)

100±6 90±3

Latent energy of crystallization (J/g)

39±5 33±3

X-ray diffraction Intensity (cps) 240±20 470±30

BET of ethanol-washed powder Surface area (m2 g-1) 10.5±0.6 8.3±0.4

All values are means ± standard deviations of at least three replicate analyses.

Page 107: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

89 | P a g e

Table 6.4 Summary of the experimental conditions used in spray drying for lactose (10% w/w) –WPI

(0.2% w/w) -ascorbic acid solutions, inlet gas temperature of 180ºC and outlet gas temperature 112 ± 3ºC.

Analysis Ascorbic acid concentration (%, w/w)

1 2 3

Spray-dried powder

Yield (%, w/w) 87±2 83±1 76±1

Moisture content (% dry basis) 1.5±0.2 1.8±0.1 2.1±0.2

Solution pH 2.93±0.05 2.84±0.02 2.81±0.03

Sorption test

Peak height (moisture content change, %, kg.kg-1

dry basis of lactose)

14±1 14±1 14±1

BET of ethanol-washed powder Surface area (m2 g-1) 21.5±0.5 19.3±0.9 18.2±0.7

All values are means ± standard deviations of at least three replicate analyses.

Table 6.5 Summary of the experimental conditions used in spray drying for lactose (10 % w/w) - boric

acid solutions, inlet gas temperature of 180ºC and outlet gas temperature of 98 ± 4ºC.

Analysis Boric acid concentration (%, w/w)

1 2 3 5 7

Spray-dried powder

Yield (%, w/w) 75±4 75±1 73±2 67±2 56±5

Moisture content (% dry basis) 4.0±0.1 4.2±0.1 5.2±0.3 6.8±0.1 7.2±0.3

Solution pH 5.34±0.08 4.93±0.02 4.47±0.01 3.91±0.05 3.46±0.03

Sorption test

Peak height (moisture content change, %, kg.kg-1

dry basis of lactose) 7±1 7±1 7±1 6±1 6±1

BET Surface area (m2 g-1) 17.2±0.5 18.3±0.4 18.5±0.8 11.1±0.4 9.9±0.6

All values are means ± standard deviations of at least three replicate analyses.

Page 108: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

90 | P a g e

6.2.2 Assessment of the Degree of Crystallinity for the Lactose in Spray-Dried Powders

6.2.2.1 Moisture Sorption Characteristics

Time-dependent moisture sorption tests (moisture content change, %, kg.kg-1 dry basis of

lactose) showed significant changes in the crystallinity of lactose in the spray-dried products

from lactose solutions with different templating-acid concentrations. The moisture sorption

behavior of spray-dried powders demonstrated a gradual increase in the apparent degree of

lactose crystallinity through less moisture sorption, when the acid concentration was increased.

Adding 1% (w/w) lactic acid to the lactose solution decreased the degree of lactose amorphicity

for the spray-dried powders, by lowering the sorption peak heights from 6% ± 1% to about 5.5%

± 0.4% (Table 6.1). Further increasing the lactic acid concentrations to about 3% (w/w)

decreased the sorption peak height to around 4% ± 1%, suggesting higher degrees of lactose

crystallinity in the spray-dried powders. Similarly, when using citric acid and ascorbic acid as the

templating acid, a steady increase in the degree of lactose crystallinity through less moisture

sorption (lower peak heights) has been observed. Table 6.2 shows that spray drying of lactose

solution with 3% (w/w) citric acid resulted in a significant reduction in peak height to about

2.9% ± 0.6%. The highest ascorbic acid concentration used in this study was 2 % (w/w). This

concentration decreased the peak heights from 6% ± 1% to about 4.8% ± 0.5% (Table 6.3).

However, the addition of ascorbic acid to the lactose/WPI solutions did not significantly change

the degrees of lactose crystallinity in the spray-dried powders. Using ascorbic acid at

concentrations ranging from 1% to 3% (w/w) gave almost the same peak heights, of about 14%.

The higher peak heights for lactose/WPI solutions (compared with lactose solutions at the same

concentration of ascorbic acid) were expected, as it has been found that the presence of proteins

at low concentrations significantly decreases the lactose product crystallinity (Jouppila and Roos,

1994a).

Increasing the boric acid concentrations from 1% to 7% (w/w) also did not have a significant

effect on lactose crystallinity, as the peak heights only changed from 7% ± 1% to about 6% ± 1%

(Table 6.5). Lower degrees of lactose crystallinity for the spray-dried powders from the

lactose/boric acid solutions may be explained by the lower acidity of boric acid, as suggested by

the greater pHs of the solutions. An explanation for the greater lactose crystallinity at higher

Page 109: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

91 | P a g e

templating-acid concentrations may be may be connected with the increase in mutarotation rate

and the rate of orientation of lactose molecules into the crystals at lower pHs (Twieg and

Nickerson, 1968; Nickerson and Moore, 1974; Jenness, 1988; Singh et al., 1991).

6.2.2.2 Modulated Differential Scanning Calorimetry

Modulated Differential scanning calorimetry (MDSC) has been used to observe phase transitions

and to determine glass-transition temperatures and the degree of lactose crystallinity for spray-

dried powders from lactose solutions with different templating-acid concentrations. The MDSC

results for spray-dried powders with different ascorbic acid concentrations showed very clear

phase transitions during the thermal analysis (Figure 6.2), while those for lactose/WPI solutions

were complicated to analyse due to multiple components being present in the particles. The

presence of protein, fat, additives, and mineral contents in whey isolate powder made it difficult

to distinguish each transition in the thermogram separately (Vuataz, 2002), because the

transitions of each component may have overlapped with those of lactose.

The MDSC results presented in Figure 6.2 suggested that increasing the ascorbic acid

concentrations in the lactose solutions decreased the glass-transition and crystallization

temperatures for the lactose-ascorbic acid mixtures in the spray-dried powders. There was a

downward trend in the latent energy of crystallization for the spray-dried products, which

indicated a decrease in the amorphous content of the spray-dried powders as the acid

concentration increased. The glass-transition temperature of pure lactose was 82 ± 8oC, and the

crystallization peak appeared at 153 ± 9oC with an exothermic heat of crystallization of 89 ± 2

J/g. The corresponding figures for spray-dried powders from lactose solutions with 2% (w/w)

ascorbic acid concentrations significantly decreased to 45 ± 2oC, 90 ± 3oC, and 33 ± 3 J/g,

respectively (Tables 6.1 and 6.2).

The MDSC results of spray-dried powder of lactose solutions with different lactic acid

concentrations also showed clear phase transitions during the thermal analysis. The glass-

transition temperature, the crystallization temperature, and the exothermic heat of crystallization

for the spray-dried lactose with 1% (w/w) lactic acid concentrations were found to be 64 ± 2oC,

Page 110: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

92 | P a g e

142 ± 7oC, and 57 ± 13 J/g, respectively while those for the spray-dried lactose with 3% (w/w)

lactic acid concentrations decreased to 40 ± 10oC, 90 ± 10oC, and 46 ± 4 J/g, respectively (Table

6.1).

The results from the MDSC technique agreed well with the results from the moisture sorption

tests and confirmed that the crystallinity of the lactose powders increased with greater

ascorbic/lactic acids contents. During the spray-drying process, several phase transformations

may occur within the material. Depending on the process intensity and material physical

properties, the dispersed liquid turns into a solid with some crystalline regions and many

amorphous ones (governed by drying kinetics and liquid-phase crystallization) due to the

evaporation of the solvent, and the predominantly amorphous regions may also transform to

crystalline ones (governed by Williams– Landel–Ferry kinetics) (Bhandari et al., 1997b; Islam et

al., 2010). The results suggested that using ascorbic/lactic acids, with lower glass-transition

temperatures than that for lactose, decreased the glass-transition temperature of the spray-dried

mixture according to the Gordon–Taylor equation ( 2.1). Lower glass-transition temperatures

increase the temperature difference (T-Tg), which in turn decreases the crystallization time as

suggested by the Williams–Landel–Ferry equation ( 2.3), leading to higher degrees of

crystallinity for the spray-dried particles.

Page 111: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

93 | P a g e

Figure 6.2 Effect of the ascorbic acid concentration on the glass-transition temperatures, the

crystallization temperatures, and the latent energies of crystallization for the spray-dried products from

lactose solutions.

6.2.2.3 X-Ray Diffraction (XRD)

In this investigation, based on the references (Jouppila et al., 1997; Haque and Roos, 2005), the

peak at 2θ= 20.1° was taken as the characteristic peak to measure its intensity for finding the

relative lactose crystallinity of different powders (Tables 6.1-6.3). The diffraction patterns for

spray-dried lactose with different lactic acid/citric acid concentrations show clear sharp peaks

associated with α-lactose monohydrate at high intensities (2θ= 12.5°, 20.0°, and 20.1°). There is

no distinguishable peak for anhydrous β-lactose, and the peak intensities for the mixture of

anhydrous α:β with different molar ratios at 19.1°and 19.5° are relatively low. There were some

improvements in the degrees of crystallinity for lactose powders with increasing lactic acid

concentrations (Table 6.1). The intensity of the characteristic peak at 2θ= 20.1° for the spray-

dried pure lactose was found to be 140 ± 10 cps, while the intensities for the spray-dried lactose

with 1% (w/w) and 3% (w/w) lactic acid concentrations were 270 ±20 cps and 440 ± 20 cps,

respectively. Simply comparing the intensity values of the characteristic peak at 2θ= 20.1° for

0

20

40

60

80

100

120

140

160

0

20

40

60

80

100

0 1 2 3

Gla

ss-t

ran

siti

on t

emp

erat

ure

,0 CC

ryst

alli

zati

on t

emp

erat

ure

, 0 C

Hea

t of

cry

stal

liza

tion

(J/

g)

Ascorbic acid concentration (%w/w)

Heat of crystallization

Crystallization temperature

Glass-transition temperature

Page 112: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

94 | P a g e

different lactic acid concentrations suggests that the lactose crystallinity of the powders increased

by raising the lactic acid concentrations in lactose solutions.

Likewise, XRD analyses have shown distinct peaks associated with α-lactose monohydrate at

high intensities by increasing the citric acid concentrations in lactose solutions. These results

support the results of MDSC and sorption tests in confirming the increase in the degree of lactose

crystallinity when the lactic/citric acids contents were raised (Table 6.1). The intensities of the

characteristic peak at 2θ= 20.1° for the lactose in the spray-dried powders with 1% (w/w) and

3% (w/w) citric acid concentrations were 290 ± 20 cps and 410 ± 20 cps, respectively. Similarly,

in the case of using ascorbic acid as a templating acid, there were distinct peaks associated with

α-lactose monohydrate at high intensities by increasing the ascorbic acid concentrations in

lactose solutions. The intensity of the characteristic peak at 2θ= 20.1° for the spray-dried lactose

with 1% (w/w) ascorbic acid was found to be 240 ± 20 cps, while that for the spray-dried lactose

with 2% (w/w) ascorbic acid concentrations was 470 ± 30 cps. However, addition of ascorbic

acid to the lactose/WPI solutions did not significantly change the degrees of lactose crystallinity

in the spray-dried powders, as there were no distinguishable peaks associated with different types

of lactose crystals. These results agreed well with the results of the moisture sorption tests and

suggested that the resulted spray-dried powders from lactose/WPI solutions with different

concentrations of ascorbic acid were amorphous. These results are also consistent with the effect

of adding WPI on the product crystallinity, since it has been found that the presence of proteins

at low concentrations significantly affects the product crystallinity (Jouppila and Roos, 1994a;

Haque and Roos, 2006) (Table 6.3). There were also no significant changes in peak intensities

for different types of lactose crystals in the spray-dried powders from lactose solutions with

different boric acid concentrations, which means that the relative lactose crystallinity of different

powders has not been changed by increasing the boric acid content. The evidence supports this

suggestion, with the sorption peak heights being almost the same when the concentration of boric

acid was increased (Table 6.3).

The degrees of crystallinity measured using MDSC and XRD methods have agreed well with the

results of the moisture sorption tests and confirm that the crystallinity of the lactose increased

when the templating-acid concentration was raised. The results of the crystallinity tests and the

moisture contents of the final products support the suggestion that the crystallization rate is

Page 113: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

95 | P a g e

related to the difference between the material temperature and the glass-transition temperature,

which has been observed by other researchers (Bhandari and Howes, 1999; Islam et al., 2010).

This situation means that, as the templating-acid concentrations and the product moisture

contents increase, the glass-transition temperatures decrease. Therefore, the difference between

the glass-transition temperatures (Tg) of the materials and the process temperatures (T) increases,

thereby enhancing the crystallization rate, according to the Williams–Landel–Ferry equation (Eq.

2.2).

The acidity or the pH of the lactose solutions also had a significant impact on the lactose

crystallinity. The pHs of the lactose solutions containing lactic acid, citric acid, and ascorbic acid

(almost constant under different acid concentrations) were much lower than the ones for

lactose/boric acid solutions (Tables 6.1-6.3). Hence, the higher degrees of lactose crystallinity for

the spray-dried powders with lactic/citric/ascorbic acids, as suggested by the moisture sorption,

MDSC, and XRD techniques, were expected. Higher acidic strengths of lactic acid, citric acid,

and ascorbic acid compared with boric acid decreased the pH of the solution, which increased the

mutarotation rate and then the crystallization rate of lactose (Twieg and Nickerson, 1968;

Nickerson and Moore, 1974; Jenness, 1988; Singh et al., 1991).

6.2.2.4 BET Specific Surface Area

The BET surface areas of the processed lactose particles using templating methods with different

concentrations of templating acids have been shown in Tables 6.1-6.3. There has been a

significant improvement in the surface area by using lactic acid as a templating material, from

0.3 ± 0.2 m2 g-1 (for conventional spray-dried lactose) to 14.9 ± 0.9 m2 g-1 at a lactic acid

concentration of 1 w/w %. Above this concentration, there was a considerable reduction in the

surface area to 9.5 ± 0.8 m2 g-1 (3 w/w %). It can also be observed, in Table 6.1, that adding 1%

w/w citric acid as a templating material significantly increased the BET surface area of lactose to

20.8 ± 0.9 m2 g-1. However, adding further citric acid content (3 w/w %) decreased the BET

surface area of lactose to 12.3 ± 0.8 m2 g-1.

Page 114: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

96 | P a g e

The BET surface areas of the lactose particles, using ascorbic acid as a templating acid, have

been given in Table 6.3. The highly-porous lactose powders with surface areas of 10.5 ± 0.6 m2

g-1, a total pore volume of 0.096 ± 0.008 m2 g-1 and a mean pore diameter of 31.4 ± 0.2 Å have

been achieved using 1% w/w ascorbic acid, and adding more ascorbic acid (2% w/w) decreased

the BET surface area to 8.3 ± 0.4 m2 g-1. However, when increasing the ascorbic acid content to

more than 2% (w/w), producing dry powder was impossible. As mentioned earlier, in order to

increase the yield of the spray-drying process and also reducing the degree of crystallinity, 0.2%

(w/w) whey protein isolate (WPI) has been added to 10% (w/w) lactose with different

concentrations of ascorbic acid. WPI addition significantly improved the BET surface area of the

ethanol-washed particles from 10.5 ± 0.6 m2 g-1 to 21.5 ± 0.5 m2 g-1 using 1% w/w ascorbic acid.

The BET surface area of lactose (0.2% w/w WPI) decreased from 21.5 ± 0.5 m2 g-1 for 1% w/w

ascorbic acid to about 18.2 ± 0.7 m2 g-1 with 3% w/w ascorbic acid. The total pore volumes and

mean pore diameters were also affected at different ascorbic acid concentrations, such that the

templated lactose particles (after ascorbic acid removal) showed a reduction in pore volume from

0.32 ± 0.05 ml g-1 to 0.27 ± 0.04 ml g-1 (with a change in ascorbic acid concentration from 1% to

3 % w/w) and the mean pore diameters decreased from 14.8 ± 0.5 Å to 28.3 ± 0.9 Å.

It can also be observed, in Table 6.5, that the BET surface area of lactose can be altered from

17.2 ± 0.5 m2 g-1 for the spray-dried lactose with 1% w/w boric acid as a templating material to

about 9.9±0.6 m2 g-1 for 7% w/w boric acid. It can be seen that adding more than 2% w/w boric

acid did not have a significant effect on the surface area of the ethanol-washed powders. Again,

with increasing the templating-acid concentration, the porosity of lactose significantly decreased,

with the particle BET surface area decreasing from 18.5 ± 0.8 m2 g-1 for 3% w/w boric acid to

9.9 ± 0.6 m2 g-1 for 7% w/w boric acid.

These results have supported the link between the extent of lactose crystallization in the spray-

dried powders, concentrations of the templating acids, and the BET surface areas of the

processed powder, as will now be explained. Crystalline solids are known to have distinctive and

organized internal structures, called crystal lattices, while the atoms and molecules in amorphous

solids are arranged randomly with no regular patterns (Averill and Eldredge, 2007). During the

crystallization process, molecules of each component are more likely to join their own crystal

structure (Jones and Mirkin, 2013), and the impurities (templating materials) tend to be rejected

Page 115: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

97 | P a g e

during crystal growth (Bhandari and Howes, 1999; Lorenz and Beckmann, 2013). Therefore,

each growing crystal tends to consist of only one type of molecule.

Lower degrees of crystallinity for spray-dried lactose powders mean that the lactose crystal

structures have become less regular, which in turn means that the number of templating-acid

molecules incorporated in the whole lactose structure (amorphous and crystal) has increased.

Washing with ethanol removes the templating-acid molecules, either scattered in the amorphous

portion of the lactose particles or accumulated on the surface of the crystals, rather than inside

the crystal lattice. Therefore, the porosities and the BET surface areas of the powders increased

as the amorphous lactose content increased, which may be connected with the increase in the

amount of incorporated templating-acid molecules, due to the decrease in the degree of lactose

crystallinity for the spray-dried powders. This explanation implies lower surface areas at higher

templating-acid concentrations, as seen experimentally here (Table 6.1-6.3).

The presence of proteins during the spray drying of the lactose/WPI mixtures is likely to reduce

the rate and extent of crystallization for lactose during spray drying. The presence of WPI

increased the glass-transition temperatures of the surfaces for the particles, making it difficult for

the lactose in the particles to crystallize completely in the process and resulting in higher levels

of amorphicity in the particles (Jouppila and Roos, 1994a; Haque and Roos, 2006). However,

this work aimed to produce pure porous lactose, which means that using less surfactant to reach

this goal has been preferred. Hence higher concentrations of WPI have not been desired or tested

for this product development, since increasing the WPI amounts have had detrimental effects on

the product properties.

6.2.2.5 SEM Micrographs

Scanning electron microscopy has also been used to study the morphology of the powders.

Figure 6.3 shows the SEM micrographs of the spray-dried powders from lactose/0.2% (w/w)

WPI solutions with 1% w/w ascorbic acid before and after ethanol washing, where the surfaces

of the spray dried powder appeared to be non-porous and amorphous (Figure 6.3A), while

ethanol-washed powders show porosity on the surface (Figure 6.3B) and also inside the particles

(Figure 6.3C and D). The final processed particles are highly-interconnected porous networks

Page 116: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

98 | P a g e

showing porosity in the bulk as well as on their surfaces, suggesting the relatively uniform

distribution of the pores throughout the particles. However, the SEM images of the ethanol-

washed particles from lactose solution with 1 w/w % ascorbic acid (Figure 6.4) show some dense

and non-porous block crystal areas, suggesting low porosities, as also indicated by the BET

measurements. Overall, the results are consistent with the measurements of the BET surface

areas for the samples.

The morphologies of the particles support the critical role of amorphicity in the spray-dried

powders for achieving a well-dispersed lactose/template mixture to reach high and uniform

porosities in the processed lactose particles. Adding WPI to the lactose/templating acid solutions

decreased the degree of lactose crystallinity in the spray-dried particles, leading to homogenous

dispersion of the templating acid into the lactose structure to produce highly porous particles

upon removal of acid through ethanol washing (Figure 6.3A and B).

The measurements of the particles size gave a mean particle size of about 19 ± 2 μm (D[3,2]) for

the final processed powder, while that for the spray-dried powder prior to ethanol washing was 5

± 1 μm (D[3,2]). The D[4,3] values for these two powders were 122 ± 7 μm and 43 ± 3 μm,

respectively. The large processed particles have been produced by grinding the dried

agglomerate/filtrate obtained after ethanol washing, which consists of small ethanol-

washed/porous particles, suggesting that the high porosity and relatively large processed particles

are actually agglomerates of smaller porous particles.

Page 117: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

99 | P a g e

Figure 6.3 Scanning electron micrographs of spray-dried powders from lactose/0.2% (w/w) WPI

solutions with 1% w/w ascorbic acid, before (A) and after ethanol washing (B:D).

A B

C D

Page 118: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

100 | P a g e

Figure 6.4 Scanning electron micrographs of spray-dried powders from lactose solutions with

1% w/w ascorbic acid, after ethanol washing.

A B

C D

Page 119: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

101 | P a g e

6.3 Conclusions

This chapter has investigated the effects of different types of templating acids at various

concentrations on the porosity of lactose through doing experiments in which templating acids,

in various proportions, were added to lactose solutions and then spray dried. The optimum

concentrations of templating acids have been found. The optimum concentration of lactic acid,

citric acid, and ascorbic acid were found to be 1 w/w % with the BET surface area of 14.9 ± 0.9

m2 g-1, 20.8 ± 0.9 m2 g-1, and 10.5 ± 0.6 m2 g-1, respectively while that for the spray-dried lactose

with boric acid as a templating agent was 3 w/w % with the BET surface area of 18.5 ± 0.8 m2 g-

1. Higher concentrations of different templating acids appear to decrease the surface area of the

ethanol-washed lactose particles due to the increase in the extent of lactose crystallinity as the

acid concentrations increased. The results of moisture sorption tests, MDSC, and XRD tests

confirmed that the crystallinity of the lactose in the spray-dried powders increased with higher

templating-acid concentrations. This increase in lactose crystallinity may be due to increasing the

T-Tg difference, consequently decreasing the crystallization time, according to the Williams–

Landel–Ferry equation, or increasing the acidity of the solution, which increases the mutarotation

rate and then the crystallization rate of lactose. It has also been found that the increasing

templating acid concentration decreases the yield of the spray-drying process due to stickiness.

The decrease in yields is associated with the decrease in the glass-transition temperatures for the

amorphous spray-dried intermediate products, since the templating agent has a very low glass-

transition temperature. The results of this study can be of specific interest for controlled drug

deliveries and release applications (Millqvist-Fureby et al., 2014) by choosing the processing

conditions to control the extent of crystallization to produce highly-porous particles with the

wide range of porosities.

Page 120: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

102 | P a g e

CHAPTER 7 FORMATION OF HIGH POROSITY MANNITOL

PARTICLES BY A NEW TEMPLATING PROCESS3

In the previous chapter (Chapter 6), a new method has been successfully developed to create

highly-porous powder templates through spray drying, using food-grade acid as the templating

agent and then removing the template material by ethanol washing of the spray-dried powders.

As mentioned earlier, the use of mannitol as an alternative to lactose in food products and

pharmaceutical formulations has significantly increased. The non-hygroscopic nature of

mannitol, and consequently its low moisture content, make it a desirable additive and excipient

for moisture sensitive ingredients (Babu and Nangia, 2011). In this chapter, a templating

approach has been applied according to the method described in Chapter 6, to create highly-

porous templates or frameworks of mannitol through the spray drying of mannitol using citric

acid as the templating agent and then removing the template material (citric acid) by ethanol

washing of the spray-dried powders. It is also demonstrated that textural properties, such as the

surface area and the pore volume of the resultant mannitol, can be altered by varying the

concentrations of citric acid, WPI, and Gum Arabic.

7.1 Materials and Methods

7.1.1 Sample Preparation

In order to investigate and report the effects of citric acid, WPI, and Gum Arabic concentrations

on the BET surface areas of mannitol, experiments were carried out by varying the composition

of the solutions with additions of citric acid, WPI, and Gum Arabic, whilst maintaining the level

of mannitol concentration in solution at 10% (w/w). Each solution was made up to a total weight

of 100 g. All solutions were magnetically stirred at room temperature of 25oC for at least 30

3 The content of this chapter has been published as the following paper: Saffari, M., Ebrahimi, A. & Langrish, T.A.G. (2015a), Highly-porous mannitol particle production using a new templating approach, Food Research International, 67, 44-51.

Page 121: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

103 | P a g e

minutes, so a clear solution was obtained without any visible crystals being present. The clear

solutions were then spray dried. The pH of the solutions was measured with a pH electrode.

7.1.2 Operating Conditions for Spray Drying

The solutions have been spray dried using a Buchi-B290 Mini Spray Dryer with an inlet gas

temperature of 150oC, an aspirator setting of 100% (38 m3 h-1), a pump rate of 25% (8 mLmin-1),

and a nozzle air flow rate of 470 L h-1 (40 on the nozzle rotameter scale). Freshly spray-dried

powder has been collected from a collection vessel at the bottom of a cyclone, and a portion of

that powder has been treated with ethanol for 24 h to remove the citric acid, vacuum filtered, and

oven dried at 60°C for one hour, as shown in Figure 6.1. The final powders have been produced

by crushing and grinding the dried cake. The freshly spray-dried and ethanol washed powders

were immediately used for analytical tests (moisture content, gravimetric moisture analysis,

Fourier transform infrared spectroscopy (FTIR), and BET); otherwise the powders were kept in

sealed bags and in a refrigerator for scanning electron microscope (SEM) tests on the following

day. The free moisture content (dry basis) was measured by weighing the sample before and after

drying in a fan circulated oven (Labec, Australia) at 85ºC for 24 hours.

7.2 Results and Discussion

7.2.1 Yield from Spray Drying

It can be seen in Table 7.1 that adding citric acid to the mannitol decreased the yield (or

recovery) of the spray-drying process, which can be attributed to greater stickiness during the

spray-drying process. The yield for mannitol was decreased from 70% ± 2% to about 8% ± 1%

by increasing the citric acid concentration to 2% (w/w). Likewise, in the case of using 0.5%

(w/w) Gum Arabic as an additive, increasing the citric acid concentration to 2% (w/w)

significantly decreased the yield of the process down to 15% (Table 7.3). A key explanation for

the steady downward trend in the yield can be seen in the work of Bhandari et al. (1997b), due to

stickiness. There is a significant increase in the moisture content of the product when increasing

the citric acid concentration (Table 7.1). The increase in moisture content means an increase in

Page 122: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

104 | P a g e

stickiness due to the effect of moisture content in depressing the glass-transition and sticky-point

temperatures (Bhandari et al., 1997b). The presence of citric acid must also decrease the glass-

transition temperature and hence increase stickiness at any moisture content. In order to increase

the yield of spray-drying process, 0.2% (w/w) whey protein isolate has been added to 10% (w/w)

mannitol with different concentrations of citric acid. The yield of the spray-drying process can be

altered from 19% ± 5% to about 84% ± 3% for the spray-dried mannitol/WPI mixture with 1%

(w/w) citric acid, compared with 8% ± 1% to 33% ± 8% with 1% (w/w) citric acid (Table 7.1).

The higher yields of the spray-drying process for mannitol with WPI solutions compared with

mannitol solutions at the same concentrations of citric acid can be explained by the decrease in

wall deposition during spray drying due to the effective encapsulating properties of proteins,

which remain non-sticky due to their higher glass-transition temperatures (Adhikari et al., 2004;

Wang and Langrish, 2010; Islam et al., 2013).

Further experimentation has been done to investigate the effect of WPI and Gum Arabic on both

BET surface area and the yield of the process by adding whey protein isolate and Gum Arabic to

10% (w/w) mannitol, and then varying the composition of the solution with additions of citric

acid. The results suggest that only 0.5% (w/w) WPI concentration was enough to give high

spray-drying yields and that further increases in the WPI content in the mannitol/WPI mixtures

did not change the process yield significantly (Table 7.2). In the case of using Gum Arabic as a

surfactant, increasing the concentration of Gum Arabic did not improve the yield of the process;

increasing the amount of Gum Arabic concentration to 3% (w/w) decreased the yield of the

process down to 8%. However, the lower yields of the spray-drying process for mannitol with

Gum Arabic solutions than mannitol/WPI solutions with the same concentration of citric acid

can be attributed to the lower surface activity of Gum Arabic than WPI. This result suggests that

more Gum Arabic stays in the particle structure as an incorporated material (Table 7.3).

Page 123: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

105 | P a g e

Table 7.1 Summary of the experimental conditions used in spray drying for mannitol- citric acid - WPI

solutions (mannitol 10% w/w), inlet gas temperature of 150oC and outlet gas temperature of 90 ± 4oC.

Analysis Citric acid: WPI concentration (%, w/w)

0:0 1:0 1:0.2 2:0 2:0.2

Spray-dried powder

Yield (%, w/w) 70±2 19±5 84±3 8±1 33±8

Moisture content,% (dry basis) 0.15±0.1 1.9±0.2 2.1±0.3 3.4±0.5 2.3±0.1

Solution pH 7.3±0.1 2.41±0.04 2.45±0.03 2.23±0.02 2.26±0.02

BET Surface area(m2 g-1) 0.71±0.05 1.91±0.04 3.20±0.90 1.60±0.09 5.65±0.60

All values are mean ± standard deviation of at least two replicate analyses.

Page 124: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

106 | P a g e

Table 7.2 Summary of the experimental conditions used in spray drying for mannitol (10% w/w) - citric acid (2% w/w)- WPI solutions (% w/w),

inlet gas temperature of 150oC and outlet gas temperature of 90 ± 3oC.

Analysis WPI concentration (%, w/w)

0 0.2 0.5 1 2 3 4

Spray-dried powder Yield (%, w/w) 8±1 33±8 64±5 60±5 72±6 66±12 68±8

Moisture content,% (dry basis) 3.4±0.5 2.3±0.1 2.3±0.1 2.1±0.2 1.7±0.1 1.5±0.1 1.3±0.1

Solution pH 2.23±0.03 2.26±0.02 2.30±0.01 2.42±0.01 2.64±0.04 2.75±0.03 2.82±0.02

BET Surface area(m2 g-1) 1.6±0.1 5.6±0.6 8.0±0.4 3.4±0.1 2.7±0.1 3.5±0.1 4.5±0.1

All values are mean ± standard deviation of at least two replicate analyses.

Table 7.3 Summary of the experimental conditions used in spray drying for mannitol- citric acid –Gum Arabic solutions (mannitol 10% w/w),

inlet gas temperature of 150oC and outlet gas temperature of 89 ± 3oC.

Analysis Citric acid: Gum Arabic concentrations (%, w/w)

0.5:0.5 1:0.5 2:0.5 2:1 2:2 2:3

Spray-dried powder Yield (%, w/w) 51±3 33±2 15±3 19±6 22±4 8±2

Moisture content,% (dry basis) 0.8±0.1 1.4±0.1 2.3±0.2 3.8±0.4 5.8±0.5 NA

Solution pH 2.65±0.09 2.46±0.02 2.30±0.02 2.38±0.01 2.48±0.01 2.57±0.02

BET Surface area(m2 g-1) 1.8±0.1 1.8±0.1 5.1±0.2 7.3±0.3 2.1±0.3 NA

All values are mean ± standard deviation of at least two replicate analyses. *NA: not available

Page 125: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

107 | P a g e

7.2.2 BET Specific Surface Area

It can be observed from Table 7.1 that adding 1% w/w citric acid as a templating material

significantly increased the BET surface area of mannitol from 0.71 ± 0.05 m2 g-1 to 1.91 ±

0.04 m2 g-1. Increasing the citric acid content (2% w/w) did not have a significant effect on

the surface area of the ethanol-washed powders. However, adding 0.2% (w/w) WPI

significantly increased the particle porosity. The BET surface area of mannitol can be altered

from 1.91 ± 0.04 m2 g-1 to about 3.20 ± 0.90 m2 g-1 for 1% w/w citric acid, compared with

1.60 ± 0.09 m2 g-1 to 5.60 ± 0.60 m2 g-1 for the spray-dried mannitol/WPI mixture with 2%

w/w citric acid.

This result suggested that there is a link between the spray-dried powder’s mannitol

crystallinity and the porosity of the ethanol-washed powers, as explained in Section 6.2.2.4.

The results are similar to those of spray-dried mannitol with 0.5% w/w Gum Arabic, which

showed that adding 2% w/w citric acid significantly increased the BET surface area of

mannitol to 5.1 ± 0.2 m2 g-1 compared with 1.8 ± 0.1 m2 g-1 for the spray-dried mannitol/Gum

Arabic mixture with 0.5% w/w citric acid (Table 7.3).

The effects of WPI and Gum Arabic concentration on the porosity and BET surface area of

the mannitol particles have been shown in Tables 7.2 and 7.3. Due to the higher BET surface

area of the mannitol/0.2% (w/w) WPI mixture with 2% w/w citric acid, this concentration of

templating material was kept fixed in subsequent experiments. It can be observed that

increasing the WPI and Gum Arabic concentrations changed the surface area of mannitol. It

can also be seen in Table 7.2 and Figure 7.1 that adding WPI to the mannitol/citric acid

solutions significantly increased the BET surface of mannitol up to 8.0 ± 0.4 m2 g-1.

However, higher concentrations of WPI, more than 0.5% w/w, have a detrimental effect on

the particle porosity. The results also suggest that only 0.5% (w/w) WPI concentration was

enough to give high spray-drying yields and that further increases in the WPI content in the

mannitol/WPI mixtures did not change the process yield significantly (Figure 7.1). When

using Gum Arabic as a surfactant, increasing the Gum Arabic content to 1% (w/w) made it

possible to produce mannitol with a BET surface area of 7.3 ± 0.3 m2 g-1, while increasing the

amount of Gum Arabic concentration to 2% (w/w) significantly decreased the particle

porosity down to 2.1 ± 0.3 m2 g-1 (Table 7.3). These decreases in the BET surface areas with

increasing WPI or Gum Arabic concentrations can be explained by the results of FTIR

spectra in Section 7.2.4.

Page 126: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

108 | P a g e

Figure 7.1 Effect of the WPI concentration on the BET surface area and the yield of the spray-dried

products for mannitol (10% w/w) with 2% w/w citric acid (error bars indicate standard deviations).

7.2.3 SEM Micrographs

Scanning electron microscopy has also been used to study the morphology of the powders.

The SEM micrographs of the spray-dried pure mannitol and spray-dried powders from

mannitol with WPI/Gum Arabic solutions having different concentrations of citric acid are

shown in Figures 7.2 and 7.2. These micrographs show some morphological changes for the

mannitol powders, before and after citric acid removal (ethanol washing). More porosity

appeared to be present on the surface and inside of the particles after citric acid removal.

Figures 7.2 A and B show the spray-dried pure mannitol, where the surfaces of the spray

dried powder appeared to be non-porous and amorphous, while the ethanol-washed powders

have a more porous structure, suggesting the formation of a porous network (Figures 7.2 C

and D and 7.3). The results are consistent with the measurements of the BET surface areas for

the samples.

Particle size analysis of powders has shown that the particle size of the processed powder

after grinding is about 6.8 ± 0.5 μm (D[3,2]), while the particle size of the spray-dried

lactose/WPI powder before ethanol washing was in the range of 6.6 ± 0.3 μm (D[3,2]). The

D[4,3] values for these two powders are 12.8±0.2 and 108±6 μm, respectively. The large

processed particles have been produced by grinding the dried agglomerate/filtrate obtained

0

1

2

3

4

5

6

7

8

9

0

10

20

30

40

50

60

70

80

90

0 1 2 3 4

BE

T s

urf

ace

area

(m2 g-1

)

Yie

ld, %

WPI concentration (%w/w)

Yield, %

BET surface area(m2 g-1)

Page 127: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

109 | P a g e

after ethanol washing, which consists of small ethanol-washed/porous particles. Hence, it is

not unreasonable to assume that the high porosity and relatively large processed particles are

actually agglomerates of smaller porous particles.

Figure 7.2 A and B) Scanning electron micrographs of spray-dried pure mannitol powder. C and D)

spray-dried powder from mannitol with 0.5% (w/w) WPI and 2% (w/w) citric acid solution after

ethanol washing.

A B

C D

Page 128: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

110 | P a g e

Figure 7.3 Spray-dried powder from mannitol with 1% (w/w) Gum Arabic and 2% (w/w) citric acid

solution after ethanol washing.

A B

C D

E F

Page 129: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

111 | P a g e

7.2.4 Fourier Transform Infrared Spectroscopy (FTIR)

The FTIR spectra for the spray-dried powders from solutions with 2% (w/w) citric acid, at

different WPI and Gum Arabic concentrations, are presented in Figure 7.4, before and after

ethanol washing. The citric acid absorbance peaks at 1730 cm-1 for the powders with 0.5% w/w

of WPI before and after ethanol washing show almost complete citric acid removal from the

mannitol structure. However, the difference between the absorbance peaks before and after

ethanol washing decreases as the WPI concentration increases, suggesting that full citric acid

removal has not been achieved for higher contents of WPI. The decrease in the BET surface area

from 8.0 ± 0.4 m2 g-1 (WPI: 0.5% w/w) to 3.4 ± 0.1 m2 g-1 (WPI: 1% w/w) and 3.5 ± 0.1 m2 g-1

(WPI: 3% w/w) can be explained by the poor templating removal from the main mannitol

structure as a result of an increase in the WPI concentration. This behaviour can be attributed to

the accumulation of the WPI on the particle surface during spray drying, as suggested in the

literature (Izutsu et al., 2004; Islam et al., 2013). Increasing the WPI concentration increases the

thickness of the protein layer covering the particle surface, thereby decreasing the availability of

the citric acid molecules for the ethanol molecules to achieve high removal performance. The

complete disappearance of the citric acid absorbance peak has also been shown in Figure 7.4 D

for the ethanol-washed powder from solutions with 1% of Gum Arabic. A high BET surface area

of 7.3 ± 0.3 m2 g-1 has been measured for this powder, which can be attributed to the high citric

acid removal observed by FTIR spectra. Comparing Figure 7.4 B and D, it can be observed that

using equal amounts of WPI and Gum Arabic (1% w/w) can lead to different templating removal

performances. Gum Arabic is known as an emulsifying agent rather than a surface-active

material. The less surface-active properties of Gum Arabic compared with WPI means that Gum

Arabic remains incorporated in the particle structure.

Page 130: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

112 | P a g e

Figure 7.4 FT-IR spectra of the spray-dried powder from mannitol (10% w/w) - citric acid (2% w/w) -

WPI/Gum Arabic solutions before and after ethanol washing. A) 3% (w/w) WPI. B) 1% (w/w) WPI. C)

0.5% (w/w) WPI. D) 1% (w/w) Gum Arabic. The absorbance peaks at 1730 cm-1 show the citric acid

content.

0

0.1

0.2

0.3

0.4

0.5

700 900 1100 1300 1500 1700 1900

Ab

sorb

ance

Wavelength [cm-1]

3% w/w WPI-Before ethanol washing

3% w/w WPI-After ethanol washing

Citric acid peak

0

0.1

0.2

0.3

700 900 1100 1300 1500 1700 1900

Ab

sorb

ance

Wavelength [cm-1]

1% w/w WPI-Before ethanol washing

1% w/w WPI-After ethanol washing

Citric acid peak

0

0.1

0.2

0.3

0.4

0.5

700 900 1100 1300 1500 1700 1900

Ab

sorb

ance

Wavelength [cm-1]

0.5% w/w WPI-Before ethanol washing

0.5% w/w WPI-After ethanol washing

Citric acid peak

0

0.1

0.2

0.3

0.4

700 900 1100 1300 1500 1700 1900

Ab

sorb

ance

Wavelength [cm-1]

1% w/w Gum Arabic-Before ethanol washing

1% w/w Gum Arabic-After ethanol washing

Citric acid peak

Page 131: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

113 | P a g e

7.3 Conclusions

Highly porous mannitol was prepared through a new production route by the spray drying of

solutions containing mannitol and citric acid as the templating agent using a Buchi-B290 mini

spray dryer with an inlet temperature of 150oC. It has been demonstrated that textural properties,

such as the surface area and the pore volume of the resultant mannitol, can be tuned by varying

the concentrations of citric acid, WPI, and Gum Arabic. The resulting powder structure is a

mannitol network with significant porosity and a high surface area of 8.0 ± 0.4 m2 g-1 and a total

pore volume of 0.066 ml g-1 for mannitol solution with citric acid (2% w/w) and 0.5% (w/w)

WPI. This optimum preparation method has been used with the aim of producing the highest

possible surface area of the ethanol-washed mannitol particles for drug delivery purpose as

discussed in next chapter. The scanning electron microscope micrographs were consistent with

the BET surface areas. It has been found that higher concentrations of WPI, more than 0.5%

w/w, have a detrimental effect on the porosity of the spray-dried powders from mannitol/WPI

solutions with 2% (w/w) citric acid, while the concentration of Gum Arabic causing decreased

porosity was 1% (w/w) due to the decrease in the performance of citric acid removal through

ethanol washing. The results of this study have implications in the pharmaceutical industry for

producing excipient powders with better properties, such as greater specific surface areas.

Page 132: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

114 | P a g e

CHAPTER 8 HIGHLY POROUS EXCIPIENTS AS A

SOLUTION FOR POOR SOLUBILITY AND POOR CONTENT

UNIFORMITY IN SOLID DOSAGE FORMS4

Poor blending uniformity of drug dosage and the solubility behavior of a drug are key challenges

facing the design of solid dosage forms. A new adsorption method has been developed

successfully in this chapter, which includes the production of highly-porous mannitol particles

with high surface areas according to the method described in Chapter 7, through the spray drying

of mannitol solutions containing ethanol-soluble sugar or food-grade acids as templating agents,

and then transferring the resulted porous particles after ethanol washing to ethanol solutions

containing the drug component.

8.1 Materials and Methods

8.1.1 Sample Preparation

Mannitol solutions with citric acid (2% w/w) and WPI (0.5% w/w) were prepared, with the aim

being to produce the highest possible surface area of the ethanol-washed mannitol particles, as

described in Chapter 7. In order to investigate the effects of sucrose concentrations as a

templating sugar on the BET surface areas of ethanol-washed mannitol, experiments were

carried out by varying the composition of the solutions with different amounts of sucrose. In

order to increase the yield of the spray-drying process, 0.5% (w/w) whey protein isolate has been

added to 10 % (w/w) mannitol with different concentrations of sucrose.

8.1.2 Porous Mannitol Production and Drug Loading Process

Freshly spray-dried powder has been collected from a collection vessel at the bottom of a

cyclone and has been washed with ethanol for 48 h at the room temperature of 25oC to remove

the templating agents (citric acid or sucrose) and then filtered under vacuum. The resultant pastes

4 The content of this chapter has been published as the following paper: Saffari, M., Ebrahimi, A. & Langrish, T.A.G. (2016a), Nano-confinement of acetaminophen into porous mannitol through adsorption method, Microporous & Mesoporous Materials, 227, 95-103.

Page 133: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

115 | P a g e

have been immersed in the ethanol solutions with four different concentrations of acetaminophen

(0.01 M, 0.05 M, 0.2 M, and 0.5 M) as a model drug in order to load the drug onto highly-porous

mannitol for 12 h at the room temperature of 25oC. The final-processed powder has been

obtained after vacuum filtering and oven drying at 60°C for one hour to remove any residual

ethanol. Sieving has then been carried out on the portion of dried filtrate with a set of sieve sizes

ranging between 63 µm and 300 µm, and the final powder has been used for analytical tests.

Figure 8.1 illustrates the overall process. In order to investigate the effect of excipient porosity

on the efficiency of the adsorption approach, porous mannitol with a range of porosities has been

produced through the templating process by varying the concentrations of templating agents and

WPI. This material has been used for the adsorption process. In this study, for each drug

concentration, an experiment has been repeated three times. Data in this study have been

presented as means ± STD (standard deviation). The drug content uniformity of final powder, the

percentage relative standard deviation (RSD), has been obtained from one set of independent

experiments for different drug concentrations.

Droplet

Water evaporation

Spray drying

Template removed with ethanol washing

Core material (mannitol) Templating agent (sucrose)

Porous particle

Adsorption

Loading of porous mannitol withdrug in

acetone Drug-loaded mannitol

Core material (mannitol)

Loaded drugAmorphous spray-dried

particle

Figure 8.1 Schematic illustration of the of the adsorption process.

8.2 Results and Discussion

8.2.1 Highly-Porous Mannitol Particles Using Sucrose as a Templating Agent

As has been shown in Table 8.1, by adding 0.5% (w/w) WPI, the yield of the spray-drying

process could be altered from 50% ± 6% to about 87% ± 5% for the spray-dried mannitol/WPI

mixture with 1% (w/w) sucrose, compared with 64% ± 2% to 86% ± 7% with 2% (w/w) sucrose

Page 134: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

116 | P a g e

(Table 8.1). However, when increasing the sucrose content to more than 2% (w/w), producing

dry powder was impossible. The higher yields of the spray-drying process for mannitol with WPI

solutions than mannitol solutions with the same concentrations of sucrose can be explained by

the decrease in wall deposition during spray drying due to the effective encapsulating properties

of proteins, which remained non-sticky due to their higher glass-transition temperatures

(Adhikari et al., 2004; Islam et al., 2013).

Normally, a higher glass-transition temperature leads to less wall deposition, due to a lower Tp-Tg

giving less stickiness. If the temperature difference (Tp-Tg) is less than 20°C, then the particles

are likely to be non-sticky (Bhandari et al., 1997b), and spray-drying operations would be

considered to be “normal”( zone 1, Figure 8.2). Temperature differences of 20°C to 50 °C (Tp-Tg)

are generally considered to cause stickiness and operability problems in spray drying (zone 2,

Figure 8.2). This zone has been described as the “Stickiness Barrier” (Imtiaz-Ul-Islam and

Langrish, 2009). Imtiaz-Ul-Islam and Langrish (2009) also found that at temperature difference

(Tp-Tg ) above 50°C, the yields increase significantly due to a high enough Tp-Tg to give a

sufficient crystallization rate, leading to lower wall deposition (zone 3, Figure 8.2).

The virtually-zero yield with 2% (w/w) sucrose might be due to the higher glass-transition

temperature for the mixture of sucrose (62ºC; Roos, 1995) and mannitol (10.7ºC; Telang et al.,

2003) compared with the glass-transition temperature of mannitol with 1% (w/w) sucrose. A

lower glass-transition temperature, gives greater crystallization in spray drying, leading to lower

wall deposition, due to a high enough Tp-Tg to give a high enough crystallization rate according

to the Gordon–Taylor and Williams–Landel–Ferry equations (Figure 8.2, zone 3) ( 2.1, Gordon

and Taylor, 1952; equation 2.3, Williams et al., 1955).

Page 135: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

117 | P a g e

Figure 8.2 The concept of “Stickiness Barrier” for spray drying of sugar-rich materials (Imtiaz-Ul-Islam

and Langrish, 2009).

The BET surface areas showed that using sucrose as the templating agent at concentrations up to

2% (w/w) significantly increased the specific surface areas of the mannitol particles. However,

adding 0.5% (w/w) WPI caused an increase in the BET surface area of particles compared with

mannitol powders having the same concentrations of sucrose (samples 1 and 3 compared with

samples 2 and 4, respectively). The presence of WPI increased the glass-transition temperatures

of the particle surfaces, making it difficult for the particles to crystallize completely in the

process, resulting in higher levels of amorphicity in the particles and consequently higher

porosities and BET surface areas for the ethanol-washed particles (Table 8.1).

Yield

Tp-Tg (0C)

Material is sticky, because Tp-Tg is not high enough for crystallization to occur

The Stickiness Barrier (Zone 2)

Tp-Tg ˂ 200C

Tp-Tg is high enough for crystallization to occur (Zone 3)

200C ˂Tp-Tg ˂ 500C Tp-Tg ˃ 500C

Tp-Tg ˂ 200C is known as non-sticky zone (Zone 1)

Page 136: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

118 | P a g e

Table 8.1 Summary of the experimental conditions used in spray drying for mannitol (10% w/w) -

sucrose - WPI solutions, inlet gas temperature of 150oC and outlet gas temperature of 80 ± 4oC.

Analysis Sucrose: WPI concentration (%, w/w)

Citric acid (2% w/w)- WPI (0.5% w/w)

Sample 1 Sample 2 Sample 3 Sample 4 Sample5 1:0 1:0.5 2:0 2:0.5

Yield of spray drying (%, w/w) 50±6 87±5 64±2 86±7 64±5

BET surface area (m2 g-1) 3.5±0.2 3.9±0.1 6.1±0.1 7.5±0.2 10.5±0.5

Pore volume (ml g-1) 0.013±0.002 0.021±0.002 0.037±0.003 0.052±0.007 0.11±0.05

All values are means ± standard deviations of at least three independent experiments.

8.2.2 Effect of Porous Mannitol on Drug Loading

There appears to be a relationship between the extent of drug loading in the adsorption process

with different concentrations of drug solutions, and the BET surface areas of the processed

powders before drug loading, as will now be explained. It can be seen in Table 8.2 that the

current adsorption method for the whole range of drug loadings (low to high) has been

successful. With increasing the acetaminophen solution concentration, the drug contents for

samples 1, 3 and 5 have been increased.

The drug content of sample 1 with the BET surface area of 3.5 ± 0.2 m2 g-1 gradually increased

from 0.47 ± 0.02 w/w % for 0.01 M acetaminophen solutions to 16.2 ± 0.9 w/w % drug loading

for 0.5 M acetaminophen solutions. The results of the drug loading for sample 3 with higher

surface areas and pore volumes of 6.1 ± 0.1 m2 g-1 and 0.037 ± 0.003 ml g-1, respectively,

significantly increased from 0.62 ± 0.03 w/w % to 21.7 ± 0.8 w/w % drug loading when the

concentration of acetaminophen was increased from 0.01 M to 0.5 M (Table 8.2).

It can be expected that preparing an excipient with a higher BET surface area, such as lactose,

which can be produced according to the method described in Chapter 6, would result in a higher

drug loading as it (lactose) provides more available pore volume in order to host the drug

molecules. For instance, porous lactose powders produced using ascorbic acid as a templating

agent have surface areas of 21.5 ± 0.5 m2 g-1, which have a pore volumes of 0.27 ± 0.04 ml g-1.

In addition, a probable outcome of using higher concentrations of drug solutions (˃0.5 M) might

Page 137: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

119 | P a g e

be higher drug contents, since the driving force for the adsorption process is the drug

concentration. Hence for the very low drug loadings (˂0.1 w/w %), drug loading of porous

mannitol with lower concentrations of drug solutions (˂0.01 M) would appear to be achievable.

The question then arises as to why sample 5, even when having a higher surface area and more

available pore volume of 10.5 ± 0.5 m2 g-1 and 0.11 ± 0.05 ml g-1, respectively, does not have a

better performance for the adsorption process. This sample resulted in a final formulation with a

lower drug content with the same acetaminophen solution concentrations compared with sample

3, which has a lower BET surface area and pore volume. A similar result has been found for the

drug-loaded formulations from samples 2 and 4 compared with samples 1 and 3, where samples

1 and 3 have lower surface areas and pore volumes and have higher drug loadings. However, it

has been confirmed that, for all the porous mannitol samples without WPI, higher BET surface

areas led to higher drug contents, and the same trend has been found for the porous mannitol

containing WPI. WPI has some advantages, such as improving the yield and resulting in lower

crystallinity for spray-dried powders, which helps to produce final porous particles with higher

surface areas. Nevertheless, WPI accumulates on the particle surfaces during spray drying, as

suggested in the literature (Izutsu et al., 2004; Islam et al., 2013). However, the low binding of

acetaminophen to protein (Milligan et al., 1994; Zhou et al., 1997) and the barrier properties of

WPI for moisture and gas permeation (Cinelli et al., 2014) might be reasons for less

sedimentation of drug molecules into the voids of porous mannitol.

The pore size distributions of samples 5 and 3 before and after drug loading in different

concentrations of acetaminophen solutions have been presented in Figures 8.3 and 8.4,

respectively. The pore size distributions for drug-loaded samples 3 and 5 demonstrated a gradual

decrease in the total pore volume through lower pore volumes for the pore diameters between

20-200nm when the drug concentration was increased. The pore size distributions of sample 3

also showed shifting of the pore size distribution to a range of narrower pores at about 4-6 nm as

a result of drug sedimentation inside the wider pores.

Comparing the BET surface areas and pore volume values of the porous mannitol before and

after drug loading with different acetaminophen concentrations suggests that there is still

Page 138: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

120 | P a g e

significant residual porosity after drug loading, which indicates that acetaminophen molecules

have been precipitated onto the voids in the porous mannitol (Grigorov et al., 2013).

Page 139: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

121 | P a g e

Table 8.2 Summary of the experimental results regarding drug loading, BET surface area, and pore volume for different mannitol samples loaded

with acetaminophen solutions.

Drug loading, % w (API) /w (mannitol)

Acetaminophen solution

concentration (M)

Sample 1 Sample 3 Sample 5

Drug content, w/w %

BET (m2 g-1)

after drug

loading

Pore volume after loading

(ml g-1)

Drug content, w/w %

BET (m2 g-1)

after drug

loading

Pore volume after loading

(ml g-1)

Drug content, w/w %

BET (m2 g-1) after

drug loading

Pore volume

after loading (ml

g-1)0.01 0.47±0.02 3.5±0.1 0.013±0.001 0.62±0.03 3.9±0.1 0.013±0.004 0.56±0.02 10.5±0.1 0.10±0.01

0.05 1.95±0.08 3.5±0.1 0.013±0.002 3.2±0.07 3.7±0.2 0.013±0.005 2.45±0.04 8.5±0.2 0.09±0.01

0.2 8.1±0.4 3.0±0.2 0.01±0.002 10.4±0.6 3.4±0.1 0.011±0.04 7.3±0.4 6.9±0.2 0.06±0.02

0.5 16.2±0.9 2.5±0.4 0.01±0.0 21.7±0.8 3.0±0.1 0.01±0.005 14.2±0.5 6.2±0.3 0.05±0.01

Acetaminophen solution

concentration (M)

Sample 2 Sample 4

Drug content, w/w %

BET (m2 g-1)

after drug

loading

Pore volume after loading

(ml g-1)

Drug content, w/w %

BET (m2 g-1)

after drug

loading

Pore volume after loading

(ml g-1)

0.5 8.8 ±0.2 3.6±0.2 0.012±0.01 12.8±0.6 5.5±0.3 0.05±0.01

All values are means ± standard deviations of at least three independent experiments (Sample numbers are the same as in Table 8.1).

Page 140: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

122 | P a g e

Figure 8.3 Pore size distributions for drug-loaded sample 5 in different acetaminophen solutions

compared with sample 5 before loading.

0.00

0.02

0.04

0.06

0.08

0.10

0.12

2 20 200

Por

e vo

lum

e, c

m3 /

g

Pore diameter (nm)

Sample 5 before drug loading

Sample 5 loaded in 0.01M drug

Sample 5 loaded in 0.05M drug

Sample 5 loaded in 0.2M drug

Sample 5 loaded in 0.5M drug

Page 141: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

123 | P a g e

Figure 8.4 Pore size distributions for drug-loaded sample 3 in different acetaminophen solutions

compared with sample 3 before loading.

8.2.3 Physical State of Loaded Drug

Differential scanning calorimetry provides information about the physical state and the

crystalline character of a loaded drug by analyzing corresponding thermal transitions, which are

the processes of transitions (for example) between crystalline components and amorphous ones

(Westesen and Bunjes, 1995; Jenning et al., 2000; Xue et al., 2001; Feng and Kamal, 2004;

Beiner et al., 2007; Doktorovová et al., 2010; de Souza et al., 2012). Physical mixtures of porous

mannitol and acetaminophen powder were prepared with a size range of less than 63 µm (22 w/w

% drug loading), in which the ratios of drug to mannitol were similar to the weight ratios in the

final formulations of drug-loaded mannitol in 0.5 M acetaminophen solutions. These materials

were prepared in order to assess the melting point depression and also to evaluate the crystalline

character of the loaded drug in the final formulations of sample 3 with 0.5 M and 0.01 M

acetaminophen solutions. DSC thermograms suggested that the drug molecules were present in

0.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

0.016

0.018

0.020

2 20 200

Por

e vo

lum

e, c

m3 /

g

Pore diameter (nm)

Sample 3 before drug loading

Sample 3 loaded in 0.01M drug

Sample 3 loaded in 0.05M drug

Sample 3 loaded in 0.2M drug

Sample 3 loaded in 0.5M drug

Page 142: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

124 | P a g e

their crystalline forms in the final formulations since there is no crystallization peak (Figure 8.5).

A higher degree of drug crystallinity helps the processing and storage of the drug, such as better

physical and chemical stability, resulting in a stable product, so 90% of drug molecules are

delivered in the crystalline form (Variankaval et al., 2008).

Figure 8.5 DSC thermograms of sample 3 loaded in 0.01 M and 0.5 M acetaminophen solutions

compared with a DSC thermogram for a physical mixture of porous mannitol with 22 w/w %

acetaminophen powder having a particle size of less than 63 µm.

Melting curves from differential scanning calorimetry also can used to compare crystal size

distributions (Feng and Kamal, 2004; de Souza et al., 2012). The result of this MDSC analysis

suggested that a less ordered arrangement of drug crystals was dispersed in the drug-loaded

sample 3 with two different drug loadings concentrations of 0.5 M and 0.01 M, as shown by the

onset shifts for melting temperatures. The thermodynamic melting temperature is the temperature

where a perfect structure (a large crystal with no defects) would melt. Melting of crystals with

smaller sizes (high surface area) and poor quality (large number of defects) is expected to happen

-8.00

-6.00

-4.00

-2.00

0.00

2.00

4.00

6.00

8.00

10.00

12.00

25 50 75 100 125 150 175 200 225 250

Hea

t F

low

(J/

g)

Temperature(°C)

Sample 3 loaded in 0.01M acetaminophen solution, unsieved

Physical mixture of porous mannitol with 22 wt % acetaminophen powder less than 63 µmSample 3 loaded in 0.5M acetaminophen solution, unsieved

Page 143: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

125 | P a g e

at lower temperatures (Westesen and Bunjes, 1995; Xue et al., 2001), and the depression in the

melting point increased with decreasing crystal size. The melting peaks in differential

thermograms for sample 3 with two different loadings from 0.5 M and 0.01 M acetaminophen

solutions also exhibited a broad melting peak compared with a single sharp endothermic peak for

the physical mixture of porous mannitol with 22 w/w % drug loading. The broad peak mainly

occurred because of the wide crystal size distribution (Feng and Kamal, 2004; de Souza et al.,

2012). The onset shifts of the melting peaks for both sample 3 materials loaded with 0.5 M and

0.01 M acetaminophen solutions were 146.5 ± 0.5oC, with the peak temperatures of 172.2 ±

0.8oC and 167.5 ± 0.3oC, respectively. These peaks should be compared with the thermogram for

the physical mixture of porous mannitol with acetaminophen powder (22 w/w % drug loading).

This situation indicates that melting commences at 153.5 ± 0.6oC, and goes through a maximum

at 172 ± 1oC, following the established phase behavior for acetaminophen (Giordano et al., 2002;

Qi et al., 2008). There is an onset shift by almost 7°C for both drug-loaded samples compared

with the physical mixture of porous mannitol and acetaminophen powder, which agrees with the

observation that acetaminophen confinement into small pores shows different melting behavior

than in the bulk (Beiner et al., 2007). The depression of the peak melting point is more marked

for sample 3 with 0.01 M acetaminophen solutions. The reasons for the higher shift in the

melting peak to lower temperatures can be seen in the work of Beiner et al. (2007). In their

study, it was found that, for acetaminophen confined to nanopores with the largest pore diameter

of 103 nm, the melting point value was only slightly below the bulk melting point, while shifts to

lower values in the melting point increased as the pore diameter decreases.

The X-ray powder diffraction analysis (XRD) was also used to compare the crystal structure of

acetaminophen present in the drug-loaded sample 3 having a drug loading concentration of 0.5

M with the sample of porous mannitol and acetaminophen powder (22 w/w % drug loading)

prepared by powder mixing and recrystallization methods (Figure 8.6). The XRD pattern for the

physical mixture of acetaminophen with mannitol is simply a superposition of each component

with the peaks from both drugs and carriers. Figure 8.6 shows the X-ray diffraction pattern of

sample 3 loaded with a drug concentration of 0.5 M, compared with the sample of porous

mannitol with 22 w/w % acetaminophen powder, which exhibits significant line broadening in

all corresponding peak positions for acetaminophen reported in the literature (Jordan, 1993;

Nichols and Frampton, 1998). A decrease was found in the areas of the peaks in the X-ray

Page 144: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

126 | P a g e

pattern of the drug-loaded sample, compared with the areas from the physical mixtures of porous

mannitol and acetaminophen powder that contained the original crystalline acetaminophen (with

a particle size of less than 63 µm).

Less intense and highly-diffused peaks of drug in the diffractograms of the drug-loaded sample

indicate that the drug is found in the reduced ordering of the crystal lattice in the final

formulation (Ghodke et al., 2010; Jain et al., 2012). In addition, according to the Scherrer

equation (Scherrer, 1918), which relates the width of a powder diffraction peak to the average

dimensions to estimate the size of the crystallites, the peak width is inversely proportional to the

crystallite size. As the crystallite size gets smaller, the peak gets broader. In other words, as the

number of lattice repeats increases, the full width becomes narrower at the half maximum height

(FWHM), which is the width of the diffraction peak at a height half-way between the

background level and the peak maximum in the XRD (Burton et al., 2009). The results from the

XRD method agreed well with the results from the MDSC test and confirm that the smaller and

less ordered arrangement of drug crystals was dispersed and confined in the pores of mannitol

compared with the physical mixture of porous mannitol and acetaminophen powder (Beiner et

al., 2007).

Page 145: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

127 | P a g e

Figure 8.6 X-ray pattern for drug-loaded sample 3 in 0.5 M acetaminophen solution compared with the

physical mixture of porous mannitol and acetaminophen powder.

8.2.4 Blend Uniformity

In order to evaluate the drug loading efficiency, the relative standard deviation (% RSD) for the

drug content of final powder has been calculated by taking at least five samples for each drug

loading concentrations of 0.01 M and 0.5 M for one set of experiments, as the important

acceptance criteria (must be less than 6% RSD) for the content-uniformity dosage requirements

of the pharmaceutical industry. The blend uniformity was around 4% RSD or less, which

indicates a highly uniform blend, even for low-dose drug products. A content uniformity test and

a BET surface area analysis were also conducted across the particle-size distribution of the final

formulation (Table 8.3). Low variability was found for the content uniformity and delivered dose

across the particle-size distribution in sample 3 loaded in 0.01 M and 0.5 M acetaminophen.

There were also no significant changes in the BET surface areas for the different size ranges of

the drug-loaded sample 3 in 0.5 M acetaminophen, which means that the uniformity of pore sizes

0

2000

4000

6000

8000

10000

12000

14000

16000

18000

8 10 12 14 16 18 20 22 24 26 28 30

Inte

nsi

ty, c

ps

Angle, 2

Sample 3 loaded in 0.5M acetaminophen

Physical mixture of porous mannirol with 22 wt% acetaminophen powder (˂ 63 µm)

Page 146: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

128 | P a g e

after drug loading over different size ranges was very good. These results suggest that the

particle size of the final formulation does not have any significant effect on the drug content and

uniformity of pore sizes.

The evidence supports this suggestion, with the MDSC scans of all size fractions for drug-loaded

mannitol in 0.5 M and 0.01 M drug solutions being almost the same (Table 8.3). As shown in

Figures 8.6 and 8.7 for all size ranges of each drug concentration, there were similar endothermic

peaks, representing good drug dispersion throughout the mannitol.

As mentioned earlier, pharmaceutical manufacturers are required to ensure the consistency of

dosage units. However, the content uniformity of the finished products, the homogeneity of

blend and the poor blending uniformity of finished products, especially for low-dose drug

product (high RSD), are typical problems due to a combination of factors. These problems

include insufficient blending, segregation, and the large particle size of the drug substance when

using blending or mixing unit operations. Control of the particle size for the drug substance may

be a solution to ensure the content uniformity of the finished products, but the particle size and

size distribution of excipients also have a significant impact on blending homogeneity, powder

segregation, and flowability. This situation can result in unacceptable content uniformity and

control of excipients for product quality (Zheng, 2009). These results here suggest that the

present adsorption method is independent of the control strategy for both the particle size of the

drug and excipients in the finished dosage units, so here the particle size of the final formulation

does not appear to have any effect on drug uniformity.

Page 147: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

129 | P a g e

Table 8.3 Content uniformity of different size fractions for the drug-loaded sample 3 in 0.01 M and 0.5 M

acetaminophen solutions.

Acetaminophen solution

concentration (M)

Size fraction, µm

Below 63 63-106 106-150 150-212 212-300 Over 300

0.01 Drug

content, w/w %

0.60±0.02 0.60±0.01 0.59±0.01 0.61±0.01 0.61±0.00 0.60±0.02

0.5

Drug content, w/w %

20.8±0.2 20.9±0.4 21.0±0.1 21.0±0.3 20.7±0.2 20.9±0.3

BET(m2 g-1) after drug loading

2.9±0.2 2.9±0.1 3.0±0.1 3.2±0.3 3.1±0.2 3.0±0.3

All values are means ± standard deviations of at least two replicate analyses for one set of independent experiment.

Figure 8.7 DSC traces of different size fractions for drug-loaded sample 3 in 0.5 M acetaminophen

solutions.

-6

-5

-4

-3

-2

-1

0

1

25 50 75 100 125 150 175 200 225

Hea

t F

low

(W

/g)

Temperature (°C)

Over 300 µmBetween 212-300µmBetween 150-212µmBetween 106-150µmBetween 63-106µmLess than 63 µm

Page 148: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

130 | P a g e

Figure 8.8 DSC traces of different size fractions for drug-loaded sample 3 in 0.01 M acetaminophen

solutions.

8.2.5 Morphological Characterization

Scanning electron microscopy has also been used to study the morphology of the powders.

Figure 8.9 shows the SEM micrographs of the spray-dried powders (sample 3) before and after

ethanol washing compared with drug-loaded sample 3 that had a drug concentration of 0.5 M.

The surfaces of the spray-dried powders appeared to be non-porous and amorphous (Figure 8.9

A), while ethanol-washed powders showed porosity on the surface and also inside the particles

(Figure 8.9 B).

The final drug-loaded particles were also highly-interconnected porous networks showing

porosity in the bulk as well as on their surfaces, suggesting a relatively uniform distribution of

the pores throughout the particles after drug loading (Figures 8.9 C and D) and significant

residual porosity after drug loading, as also indicated by the BET measurements. The

morphologies of the particles support the measurements of the BET surface areas for the samples

and indicate that acetaminophen molecules have been adsorbed onto the voids of porous

mannitol, leading to homogenous dispersion of the drug into the mannitol structure. The

measurements of the particle sizes showed a mean particle size of about 17 ± 3 μm (D[3,2]) for

-6

-5

-4

-3

-2

-1

0

1

25 50 75 100 125 150 175 200 225

Hea

t F

low

(W

/g)

Temperature (°C)

Over 300 µmBetween 212-300µmBetween 150-212µmBetween 106-150µmBetween 63-106µmLess than 63 µm

Page 149: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

131 | P a g e

the final drug-loaded powder, while the mean size for the spray-dried powder prior to ethanol

washing was 7 ± 1 μm (D[3,2]), suggesting that the high porosity and relatively large processed

particles were actually agglomerates of smaller porous particles. The D[4,3] values for these two

powders were 112 ± 5 μm and 46 ± 2 μm, respectively.

Figure 8.9 Scanning electron micrographs of spray-dried powders from sample 3, before (A) and after

ethanol washing, compared with (B) drug-loaded sample 3 with a drug concentration of 0.5 M (C and D).

A B

C D

Page 150: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

132 | P a g e

8.2.6 Dissolution Profiles

Figure 8.10 shows in-vitro dissolution profiles of gelatin capsules filled with drug-loaded sample

3 in 0.5 M acetaminophen compared with the physical mixture of non-porous mannitol and

acetaminophen powders. The drug release rate for the drug-loaded sample 3 was slightly faster

than the release rate for the physical mixture of non-porous mannitol and acetaminophen powder.

The amount of the drug released within the first five minutes of the experiment for the drug-

loaded sample 3 was 80 %, which meets the USP requirement of 80% release in 20 minutes or

less (FDA, 2003; Grigorov et al., 2013). One of the factors that has led to the enhanced

dissolution rates of drug-loaded sample 3 is mannitol with high porosity. Porous mannitol may

improve the solubility by enabling faster release of the drug as a result of better penetration of

solvent into the acetaminophen-mannitol matrix and the large surface area of the compound that

comes into contact with the dissolution medium. In addition, deposition of acetaminophen in the

pore space of highly-porous mannitol results in particle size reduction of the drug powders to the

size range of nanometres, which may cause a significant increase in the dissolution rate of the

acetaminophen particles. The in-vitro release results agreed well with the results of MDSC, XRD

techniques and BET surface areas, and these results suggested that nano-confinement of drugs

into the porous excipients has occurred through adsorption method.

Page 151: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

133 | P a g e

Figure 8.10 Dissolution profiles of gelatin capsules filled with both the drug-loaded sample 3 in 0.5 M

acetaminophen, and the physical mixture of non-porous mannitol and acetaminophen powder (error bars

indicate standard deviations).

8.2.7 Tableting Properties of the Products

The tableting characteristics of drug-loaded sample 3 in 0.5 M acetaminophen solutions were

compared with the physical mixture of non-porous mannitol and acetaminophen powder by

measuring the thickness and crushing strength of tablets at different compaction loads. The tablet

thicknesses as a function of the applied compaction load clearly showed that, when increasing

the compression force, the tablet thickness decreased for both the drug-loaded sample 3 in 0.5 M

acetaminophen solution and for the physical mixture of non-porous mannitol and acetaminophen

powder (Figure 8.11). However, the thickness of the tablet for drug-loaded sample 3 is lower

than the tablet thickness for the physical mixture of non-porous mannitol and acetaminophen

powder at the same compression forces. This result may be explained by the higher porosity and

the lower density of drug-loaded sample 3 compared with the physical mixture of non-porous

mannitol and acetaminophen powder, which results in a higher thickness for the final tablets.

0

20

40

60

80

100

0 1 2 3 4 5 6 7 8 9 10

Dru

g re

leas

e (%

)

Time (min)

Drug-loaded sample 3 in 0.5M acetaminophen solution

Physical mixture of non-porous mannitol and acetaminophen powder

Page 152: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

134 | P a g e

Figure 8.11 Comparison of tablet thickness as a function of compression pressure for tablets from the

drug-loaded sample 3 in 0.5 M acetaminophen solution compared with the physical mixture of non-

porous mannitol and acetaminophen powder (error bars indicate standard deviations).

Figure 8.12 shows the crushing strength of the tablets for the drug-loaded sample 3 in 0.5 M

acetaminophen solutions compared with the physical mixture of non-porous mannitol and

acetaminophen powder at different compression pressures. It can be seen that, with increasing

compaction load, the tablet strength increased significantly. The differences between tablet

strengths became less pronounced with increasing the compaction load. This result may be

explained by the higher porosity of drug-loaded sample 3 compared with the physical mixture of

non-porous mannitol and acetaminophen powder, suggesting that powder-tableting properties

and the binding capacities of the powders are related to the particle porosity and surface area. It

has been found that particles with higher porosities and surface areas lead to more coherent

tablets and a subsequent increase in tablet strength, due to greater fragmentation and a decrease

in the initial particle size during compression (Vromans et al., 1987a; Vromans et al., 1987b;

Zuurman et al., 1994; Caillard et al., 2012).

1.5

1.7

1.9

2.1

2.3

2.5

2.7

50 150 250 350 450

Tab

let

thic

kn

ess,

mm

Compression pressure, MPa

Drug-loaded sample 3 in 0.5 M acetaminophen solution

Physical mixture of non-porous mannitol and acetaminophen powder

Page 153: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

135 | P a g e

Figure 8.12 Crushing strength of tablets compressed at different pressures for the tablets from the drug-

loaded sample 3 in 0.5 M acetaminophen solution compared with the physical mixture of non-porous

mannitol and acetaminophen powder (error bars indicate standard deviations).

8.3 Conclusions

Together with the uniformity of drug dosage, especially in low-dose solid drug products, the

solubility behavior of a drug is a key challenge facing the design of solid dosage forms. A new

adsorption method was developed to successfully produce highly-porous excipients with nano-

confinement drugs dispersed in the pore spaces of a porous carrier. The drug content of porous

mannitol with the surface area and pore volume of 6.1 ± 0.1 m2 g-1 and 0.037 ± 0.003 m2 g-1,

respectively, significantly increased from 0.62 ± 0.03 w/w % to 21.7 ± 0.8 w/w % when the

concentration of acetaminophen was increased from 0.01 M to 0.5 M. This method has increased

the dissolution rate of the API due to both the nano-confinement of drugs into porous excipients

and the better penetration of the dissolution medium into the pores of the excipients, giving 80%

release of the drug within five minutes.

0

20

40

60

80

100

120

50 150 250 350 450

Cru

shin

g st

ren

gth

, N

Compression pressure, MPa

Drug-loaded sample 3 in 0.5M acetaminophen solution

Physical mixture of non-porous mannitol and acetaminophen powder

Page 154: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

136 | P a g e

CHAPTER 9 A NOVEL FORMULATION FOR SOLUBILITY

AND CONTENT UNIFORMITY OF POORLY WATER-

SOLUBLE DRUGS5

This chapter investigates the enhancement of the dissolution profile of poorly-water soluble

drugs by the adsorption method described in Chapter 8. Indomethacin and nifedipine as model

drugs were used in this technique to incorporate the drug molecules onto this porous structure.

9.1 Porous Mannitol Production and Drug Loading Process

Mannitol (10% w/w)-sucrose (2% w/w) solutions were prepared according to the method

explained in Chapter 8 and then spray dried using a Büchi mini spray dryer B-290. Freshly

spray-dried powder has been collected from a collection vessel at the bottom of a cyclone and

two grams of powder has been washed with 80 mL of ethanol for 48 h at the room temperature

of 25oC to remove the templating agent (sucrose) and then filtered under vacuum. The resultant

pastes were immersed in acetone solutions with two different concentrations of indomethacin

and nifedipine (0.08 M and 0.16 M) as model drugs in order to load the drug onto highly-porous

mannitol for 12 h at the room temperature of 25oC. The carrier: drug ratios were 1:2 (w/w) and

1:4 (w/w) for 0.08 M and 0.16 M drug solutions, respectively. The final processed powder were

obtained after vacuum filtering and oven drying at 60°C for one hour to remove any residual

acetone. There was no change in the mass for drug-loaded powder after one hour of oven drying,

suggested that no solvent residues were left in the final material. The final powder has been used

for analytical tests. In this study, for each drug concentration, an experiment has been done three

times, and data have been presented as means ± STD (standard deviation). The drug content

uniformity of final powder, % RSD, was obtained from one set of independent experiment for

each drug and different drug concentrations.

5 The content of this chapter has been published as the following paper: Saffari, M., Ebrahimi, A. & Langrish, T.A.G. (2016b), A novel formulation for solubility and content uniformity enhancement of poorly water-soluble drugs using highly-porous mannitol, European Journal of Pharmaceutical Sciences, 3, 52–61.

Page 155: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

137 | P a g e

9.2 Results and Discussion

9.2.1 Analysis of Drug Loading

The experimental result of drug loading for porous mannitol with a surface area and pore volume

of 6.3 ± 0.1 m2 g-1 and 0.033 ± 0.002 ml g-1, respectively, suggested that increasing the

concentrations of drug solution results in higher drug contents for both nifedipine and

indomethacin (Table 9.1). Hence different drug loadings with adsorption method of porous

mannitol over a range of concentrations for drug solutions seem to be achievable, since the

driving force for the adsorption process is the difference between the drug concentration in the

porous mannitol and drug solution (Hillerström et al., 2009).

Table 9.1 Summary of the experimental results for drug loading with mannitol samples (BET=

6.3 ± 0. 1 m2 g-1 and pore volume of 0.033 ± 0.002 ml g-1) loaded in drug solutions.

Drug

Drug solution

concentration

(M)

Drug content,

w/w %

BET(m2 g-1)

after drug

loading

Pore volume

after loading

(ml/g)

Drug

content

uniformity,

% RSD

Nifedipine

0.08 3.2±0.1 3.8±0.1 0.026±0.002 2.8

0.16 9.1±0.3 2.5±0.2 0.019±0.006 3.2

Indomethacin

0.08 4.1±0.2 3.6±0.1 0.024±0.004 3.9

0.16 12.6±0.4 1.9±0.4 0.015±0.006 3.0

All values are means ± standard deviations of at least three independent experiments except the data for drug content uniformity, which was obtained from one independent experiment.

The results suggested that the current adsorption method has been successful for the whole range

of drug loadings. The drug content for nifedipine gradually increased from 3.2 ± 0.1 w/w % for

0.08 M drug solution to 4.1 ± 0.2 w/w % drug loading for 0.16 M drug solution. However the

final drug content for nifedipine adsorption is slightly lower than the drug content for

indomethacin. Indomethacin gave a better performance for the adsorption process and resulted in

Page 156: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

138 | P a g e

a final formulation with higher drug content with the same drug solution concentrations

compared with nifedipine, with 4.1 ± 0.2 w/w % and 12.6 ± 0.4 w/w % for 0.08 M and 0.16 M

concentrations of indomethacin, respectively, which might be due to the fact that in adsorption

method depending on the chemical nature of the drug, different drug- loading are achieved

(Salonen et al., 2005a; Millqvist-Fureby et al., 2014).

The results of drug loading for porous mannitol before and after drug adsorption in different

concentrations of drug solutions are presented in Table 9.1 and demonstrated a gradual decrease

in the total pore volume as a result of drug sedimentation inside the pores. However, there is still

significant residual porosity for drug-loaded mannitol, which indicates that drug molecules have

been precipitated onto the voids of porous mannitol. Porous mannitol with the pore volume of

0.033±0.002 mL g-1 seems not to have enough space to host drug molecules for the resulted

drug-loadings. In general, in adsorption method drug molecules are loaded at the surface and in

the internal pores of porous carriers (Prestidge et al., 2007). However, the extent of drug loading

is more likely to be a function of the degree of surface area available rather than the internal pore

volume (Prestidge et al., 2007). In the present work, in addition to the internal pores, porous

mannitol provides a large surface area for drug to be deposited on. Besides, the agglomeration of

small porous mannitol during the drug-loading stage creates secondary external voids (Boissiere

et al., 2011), which provides more available space in order to host the drug molecules.

9.2.2 Physical State of the Loaded Drugs

The physical state of the drugs in the final formulations were investigated by DSC thermograms

and X-ray patterns of the drug-loaded samples (Westesen and Bunjes, 1995; Jenning et al., 2000;

Xue et al., 2001; Feng and Kamal, 2004; Beiner et al., 2007; Doktorovová et al., 2010; de Souza

et al., 2012). Physical mixtures of porous mannitol and drug powders (in which the ratios of drug

to mannitol were similar to the weight ratios in the final formulation of drug-loaded mannitol)

were prepared in order to assess the melting-point depression and also to evaluate the crystalline

character of the loaded drug in the final formulation of different samples. DSC thermograms

suggested that the drug molecules loaded into the porous mannitol host were in their crystalline

forms in the final formulations, since there was no crystallization peak (Figures 9.1 and 9.2).

Page 157: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

139 | P a g e

This might be due to the pore size distributions of porous mannitol, which seems to be large

enough to allow the formation of crystalline form of drugs rather than retaining drugs in its

amorphous state due to the limited pore sizes (Salonen et al., 2005a). Although the amorphous

form of drugs is known to have an advantage of higher dissolution rates than the crystalline form

(Hancock and Parks, 2000), but this crystallinity is very important when considering long-term

time stability and storage of the drug (Hecq et al., 2005). More than 90% of drug molecules are

delivered in the crystalline form due to better physical and chemical stability, which results in a

stable final formulation (Variankaval et al., 2008).

Figure 9.1 DSC thermograms of mannitol drug-loaded in 0.16 M nifedipine compared with a DSC

thermogram of a physical mixture of porous mannitol with 9 w/w% nifedipine powder.

-7

-6

-5

-4

-3

-2

-1

0

20 40 60 80 100 120 140 160 180 200 220 240

Hea

t F

low

(W

/g)

Temperature (°C)

Mannitol loaded in 0.16 M nifedipine solution

Physical mixture of porous mannitol with 9 w/w % nifedipine powder

Page 158: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

140 | P a g e

Figure 9.2 DSC thermograms of mannitol drug-loaded in 0.16 M indomethacin compared with a DSC

thermogram of a physical mixture of porous mannitol with 13 w/w% indomethacin powder.

The results of the MDSC analysis show that the characteristic endothermic peak, corresponding

to drug melting, was broadened and shifted towards the lower temperature, with a reduced

intensity in the drug-loaded samples for both drugs, compared with the peaks for the physical

mixture of porous mannitol and corresponding drug powder. This result could be attributed to a

less ordered arrangement of drug crystals dispersed in the drug-loaded samples (Feng and

Kamal, 2004; Modi and Tayade, 2006; de Souza et al., 2012) (Figures 9.3 and 9.4). The

depression in the melting point of a perfectly structure drug (a large crystal with no defects)

increases with a decrease in the size and the quality of the crystal (Westesen and Bunjes, 1995;

Xue et al., 2001). Differential thermograms at the melting peaks for all drug-loaded samples also

exhibit a broad melting peak compared with a single sharp endothermic peak for the physical

mixture of porous mannitol with the same drug loadings because of the broad crystal size

distribution (Feng and Kamal, 2004; de Souza et al., 2012).

The onset shifts of the melting peaks for porous mannitol loaded with 0.16 M nifedipine and

indomethacin solutions were 150.5 ± 0.6oC and 148.6 ± 0.4oC, respectively, with the peak

-8

-7

-6

-5

-4

-3

-2

-1

0

1

20 40 60 80 100 120 140 160 180 200 220 240

Hea

t F

low

(W

/g)

Temperature (°C)

Mannitol loaded in 0.16 M indomethacin solution

Physical mixture of porous mannitol with 13 w/w% indomethacin powder

Page 159: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

141 | P a g e

temperatures of 164.8 ± 0.2oC and 164.4 ± 0.3oC, respectively. These results should be compared

with the thermogram for the physical mixture of porous mannitol and drug with the same drug

contents. This situation indicates that for nifedipine and indomethacin melting commences at

165.1 ± 0.4oC and 157.5 ± 0.4oC, respectively, and goes through the maximum at 167.1 ± 0.3oC

and 168.7 ± 0.4oC, respectively.

There are the onset shifts of about 15 and 9°C for drug-loaded mannitol with nifedipine and

indomethacin solutions, respectively, compared with the physical mixture of porous mannitol

and the corresponding drug powder, which agreed with the observation that drug confinement

into small pores shows different melting behavior than in the bulk (Hecq et al., 2005; Beiner et

al., 2007).

X-ray powder diffraction analysis (XRD) has also been used to compare the crystal structure of

drug present in the drug-loaded samples with a drug loading concentration of 0.16 M with the

physically-mixed sample of porous mannitol and drug powder (same drug loading) (Figures 9.3

and 9.4). The X-Ray diffraction pattern for the physical mixture of drugs with mannitol is simply

a superposition of each component with the peaks from both drugs. Figures 9.3 and 9.4 show the

X-ray diffraction patterns of mannitol loaded with the 0.16 M drug solution for nifedipine and

indomethacin, respectively, compared with the sample of porous mannitol having the same drug

loading, which showed significant line broadening in all corresponding peak positions for the

drugs reported in the literature (Vippagunta et al., 2002; Otsuka et al., 2003; Padrela et al., 2012).

A decrease in the areas of the peaks in the X-ray pattern of the drug-loaded sample, compared

with the areas from the physical mixture of porous mannitol and drug powder that contained the

original crystalline drug, indicated the presence of significant disorder in the crystalline material

present in the final formulation. Less intense and highly-diffused peaks of drug in the

diffractograms of solid dispersions indicate that the drug is found in the reduced ordering of the

crystal lattice in these dispersions (Ghodke et al., 2010; Jain et al., 2012). The results from the

XRD method agreed well with the results from the MDSC test and support the suggestion that

the smaller and less ordered arrangement of drug crystals was dispersed and confined in the

pores of mannitol (Beiner et al., 2007).

Page 160: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

142 | P a g e

Figure 9.3 X-ray pattern for drug-loaded mannitol in 0.16 M nifedipine solution compared with the

physical mixture of porous mannitol and nifedipine powder.

0

2000

4000

6000

8000

10000

12000

14000

16000

10 15 20 25 30

Inte

nsi

ty, c

ps

Angle, 2θ

Physical mixture of porous mannirol with 9 w/w% nifedipine powder

Drug-loaded mannitol in 0.16 M nifedipine solution

Page 161: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

143 | P a g e

Figure 9.4 X-ray pattern for drug-loaded mannitol in 0.16 M indomethacin solution compared with the

physical mixture of porous mannitol and indomethacin powder.

9.2.3 Drug Content Uniformity

In order to evaluate the drug loading efficiency, the relative standard deviation (% RSD) for the

drug content of the final powder has been calculated by taking at least five samples for each drug

loading concentration of 0.8 M and 0.16 M for one set of experiments. The blend uniformity was

less than 4% RSD for drug-loaded mannitol in both nifedipine and indomethacin with two

solution concentrations of 0.08 M and 0.16 M. This result indicates highly uniform blends, even

for low-dose drug products, as the content uniformity and the homogeneity of blend for

pharmaceutical manufacturers, especially for low-dose drug products, are typical problems due

to a combination of factors (Zheng, 2009) (Table 9.1). These results also suggest that the present

adsorption method is independent of the control strategies for the particle size and shape of both

the drug and excipients in the finished dosage units (Wong and Pilpel, 1990; Vikas Anand

Saharan, 2008).

0

2000

4000

6000

8000

10000

12000

10 15 20 25 30

Inte

nsi

ty, c

ps

Angle, 2θ

Physical mixture of porous mannirol with 13 w/w% indomethacin powder

Drug-loaded mannitol in 0.16 M indomethacin solution

Page 162: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

144 | P a g e

9.2.4 Microscopic Characterization

A scanning electron microscope was also used to study the morphology of the powders and to

observe the powders in terms of the surface and bulk structures. Figure 9.5 shows the SEM

micrographs of the spray-dried mannitol before and after ethanol washing compared with drug-

loaded mannitol from a drug concentration of 0.16 M. These micrographs show some

morphological changes for the mannitol powders, before (Figure 9.5 A) and after sucrose

removal by ethanol washing (Figure 9.5 B). The surfaces of the spray-dried powders appeared to

be non-porous, while more porosity appeared to be formed on the surface and also inside the

ethanol-washed particles. The final drug-loaded particles were also highly-interconnected porous

networks showing porosity in the bulk as well as on their surfaces, suggesting the relatively

uniform distribution of the pores throughout the particles even after drug loading (Figures 9.5

C:F). Particle size analysis of the powders has shown that the particle size of the final drug-

loaded powder after grinding is about 98 ± 7 μm (D[4,3]), while the particle size of the spray-

dried mannitol/sucrose powder before ethanol washing was 32 ± 4 μm (D[4,3]), suggesting that

the high porosity and relatively large processed particles are actually agglomerates of smaller

porous particles. The results are consistent with the measurements of the BET surface areas for

the samples, suggesting high residual porosity after drug loading. Therefore, from all the results

so far, it can be suggested that drug molecules have been precipitated onto the voids of porous

mannitol, leading to homogenous dispersion of the drug into the carrier structure despite the

initial differences in particle size and shape of porous mannitol before drug loading.

Page 163: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

145 | P a g e

Figure 9.5 Scanning electron micrographs of spray-dried mannitol, before (A) and after ethanol washing

(B) compared with for drug-loaded mannitol in 0.16 M indomethacin (C and D) and nifedipine (E and F)

solutions.

B

DC

A

EF

Page 164: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

146 | P a g e

9.2.5 Dissolution Profiles

Figures 9.6 and 9.7 the in vitro dissolution profiles of gelatin capsules filled with the drug-loaded

samples compared with the corresponding physical mixture of non-porous mannitol and drug.

Similar drug-release profiles have been found for the drug-loaded samples with different drug

loadings. Therefore, dissolution profiles of porous mannitol loaded in different drug

concentration of 0.16 M and 0.08 M has been depicted in Figures 9.6 and 9.7 as a single release

profile with error bars indicating standard deviations.

The dissolution rate of both drugs was significantly faster for the drug-loaded mannitol than the

release rate for the physical mixture of non-porous mannitol and drug powder. More than 80 %

of the drug was released within the first fifteen minutes of the experiment for the drug-loaded

mannitol for both nifedipine and indomethacin, which meets the USP requirement of 80%

release in 20 minutes or less (FDA, 2003). This improvement in the drug dissolution rate can be

ascribed to several factors. Having a less ordered arrangement of drug crystals that are dispersed

and confined into the voids of the carrier, which results in particle size reduction of drug powder

to the nanometre size range, compared with the physical mixture of mannitol and drug powder, is

one main factor leading to the enhanced dissolution of drug-loaded porous mannitol. Moreover,

the concomitant increase in wettability and dispersibility of drug into the drug-excipient matrix

(porous mannitol) may be another reason for the solubility improvement and the faster release of

the drug. Mannitol with high porosity has given better wettability and penetration of the

dissolution medium into the drug-excipient matrix, has maximized the surface area of the

compound that comes into contact with the dissolution medium. The in vitro release results have

agreed well with the results of MDSC, XRD techniques, and BET surface areas and have

suggested that nano-confinement of drugs into porous excipients has occurred through

adsorption of drug into highly-porous excipients.

Page 165: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

147 | P a g e

Figure 9.6 Dissolution profiles of gelatin capsules filled with drug-loaded mannitol in 0.16 M and 0.08 M

nifedipine and the physical mixture of non-porous mannitol powder (error bars indicate standard

deviations).

0

20

40

60

80

100

0 10 20 30 40 50 60

Rel

ease

(%

)

Time (min)

Drug-loaded mannitol in nifedipine

Physical mixture of non-porous mannitol with nifedipine powder

Page 166: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

148 | P a g e

Figure 9.7 Dissolution profiles of gelatin capsules filled with drug-loaded mannitol in 0.16 M and 0.08 M

indomethacin solutions and the physical mixture of non-porous mannitol indomethacin powder (error bars

indicate standard deviations).

9.3 Conclusions

It was found that indomethacin has a better performance for the adsorption process and resulted

in a final formulation with a higher drug content with the same drug solution concentrations,

compared with nifedipine. The blend uniformity was less than 4% RSD for drug-loaded mannitol

in both nifedipine and indomethacin, which indicates highly uniform blends, even for low-dose

drug products. This method significantly enhanced the dissolution rate for both poorly water-

soluble drugs due to the nano-confinement of drugs into the porous excipients and the high water

solubility of mannitol, where 80% of the drug was released within the first fifteen minutes of the

experiment for the drug-loaded samples.

0

20

40

60

80

100

0 10 20 30 40 50 60

Rel

ease

(%

)

Time (min)

Drug-loaded mannitol indomethacin

Physical mixture of non-porous mannitol with indomethacin powder

Page 167: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

149 | P a g e

CHAPTER 10 CONCLUSIONS AND RECOMMENDATIONS

10.1 Significant Outcomes of This Study

10.1.1 Crystallization in Fluidized-Bed Dryers/Crystallizer

One of the main foci of the present work was to investigate different processes, such as ‘‘in-

process” crystallization within spray drying and post-process crystallization in fluidized-bed

drying, which is a controlled process to transform the amorphous-lactose fraction to a crystalline

one. The experimental work consisted of studying the effects of operating conditions on the

crystallization properties of sweet-whey powder in a continuous multi-stage fluidized bed dryer

to achieve the maximum degree of lactose crystallinity with a reasonable overall processing time.

Powder was fed at different rates for each fluidization process at different sets of process

temperatures, humidities, and residence times. Experimental results showed that there were

significant reductions in the amorphicity of lactose for the processed (crystallized) powders at

different humidities and temperatures. The overall extent of lactose crystallinity for processed

whey powders has been found to be between 10% and 20% within the processing time of 30

minutes. Increasing the rate of powder feeding (from 0.3 kg h-1 to 2.6 kg h-1) did not have a

significant effect on the degrees of lactose crystallinity for the processed powders.

10.1.2 Effect of Acidity on Lactose Crystallinity

Experimental studies have also been performed to better understand the choice of processing

conditions to produce and control lactose crystallization in acid-rich carbohydrate-powders, such

as acid whey, during spray drying. Spray drying of acid whey is a significant problem for the

dairy industry due to its high content of lactic acid. It has been found that significant increases in

the degrees of lactose crystallinity for the final spray-dried products of lactose solutions have

occurred at increasing lactic acid concentrations, and the yields from spray drying have also been

significantly decreased at higher concentrations of lactic acid. The sorption peak heights for

lactose were altered from 6% ± 1% to about 0.3% ± 0.2% for lactose/lactic acid mixtures with

increasing the lactic acid concentration from 1 to 5% (w/w), and adding 5% (w/w) of lactic acid

Page 168: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

150 | P a g e

to the lactose solution significantly decreased the yield of the process by up to 22%. This

increase in crystallinity may be due to increasing the T-Tg difference (with the low Tg of lactic

acid) and consequently decreasing the crystallization time, according to the Gordon–Taylor and

Williams–Landel–Ferry equations. The increase in the rate of lactose crystallization at low pHs

may also be connected with the increase in the mutarotation rate and the rate of orientation of

lactose molecules into the crystals.

10.1.3 Producing Highly Crystalline Lactose with Significant Porosity and High Surface Area

Although crystalline powders have several advantages compared with spray-dried powders, such

as better stability and longer shelf-life, they may have some undesirable properties, such as lower

dissolution rates. Therefore, highly-porous crystalline powders with better functional (free

flowing), storage (non-caking) properties, have been developed using a new technique.

The experimental work consisted of producing highly-porous powder templates through the

spray drying of lactose solutions containing ethanol-soluble food-grade acids, such as lactic acid,

citric acid, boric acid, and ascorbic acid, as templating agents and then removing these acids by

ethanol washing of the spray-dried powders. The results showed that textural properties, such as

the surface area of the resultant lactose, changed considerably by varying the concentrations of

different templating acids, due to changes in the degrees of lactose crystallinity for the final

spray-dried products. For instance, in the case of using lactic acid as a templating acid, a change

in lactic acid concentration from 1 to 3 w/w % showed a reduction in the surface areas from 14.9

± 0.9 m2 g-1 to about 9.5 ± 0.8 m2 g-1 since the sorption peak heights decreased from 5.5% ± 0.4%

to about 4% ± 1%, suggesting higher degrees of lactose crystallinity in the spray-dried powders.

It has also been found so far that the pH of the feed solution significantly affected the BET

surface area and porosity of the processed lactose powder, where these outputs have also been

linked to the crystallinity of the spray-dried powder before ethanol washing. For example,

increasing the boric acid concentrations from 1 to 3% (w/w) did not have a significant effect on

lactose crystallinity compared with lactic acid, as the peak heights remained constant (7% ± 1%).

The BET surface area of lactose altered from 17.2 ± 0.5 m2 g-1 for the spray-dried lactose with

Page 169: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

151 | P a g e

1% w/w boric acid to about 18.5 ± 0.8 m2 g-1 for 3% w/w boric acid. Higher BET surface areas

of the lactose particles, using boric acid as a templating acid compared with lactic acid, may be

explained by the lower acidity of boric acid, as suggested by the greater pHs of the solutions and

therefore lower degrees of lactose crystallinity for the spray-dried powders from the lactose/boric

acid solutions. The pH for the lactose/boric acid solutions changed from 5.34 ± 0.08 to 4.47 ±

0.01 with increasing boric concentrations from 1 to 3 w/w %, while pHs measurements for the

lactose/lactic acid solutions were 2.53 ± 0.01 and 2.24 ± 0.01, respectively.

Increasing the acid concentration, and consequently decreasing the pH of the solution, increased

the crystallinity of the spray-dried lactose/WPI powders, giving a decrease in the BET surface

area. Increasing the number of templating molecules incorporated in the main structure is

expected to increase the porosity of the templated particle. It has also been stated that acidity can

significantly decrease the yield of the spray-drying process due to an increase in stickiness,

resulting from a decrease in the glass-transition temperature of the spray-dried powder. For

instance, in the case of using lactic acid as a templating acid a change in lactic acid concentration

from 1 to 3 w/w % the yield of the spray-drying process altered from 71% ± 2% to about 44% ±

1%. Likewise, increasing citric acid content to 3 w/w% decreased the yield of the process

significantly to 16% ± 4%.

10.1.4 Producing Highly Porous Mannitol with High Surface Area

Highly porous mannitol was also prepared through this templating route through the spray drying

of solutions containing mannitol and citric acid as the templating agent. It is demonstrated that

textural properties, such as surface area and pore volume of the resultant mannitol, can be tuned

by varying the concentrations of citric acid, WPI, and Gum Arabic. The resulting powder

structure is a mannitol network with significant porosity and high surface area of 8.0 ± 0.4 m2 g-1

and total pore volume of 0.075 ml g-1. It has been found that adding 1% w/w citric acid as a

templating material significantly increased the BET surface area of mannitol from 0.7 ±0.1 m2g-1

to 1.91 ± 0.01 m2 g-1. However, adding WPI 0.2% (w/w) significantly increased the particle

surface area to about 3.2 ± 0.9 m2 g-1 for 1% w/w citric acid, compared with 5.1 ± 1.1 m2 g-1 for

the spray-dried mannitol/WPI mixture with 2% w/w citric acid (0.2% w/w WPI). The results

Page 170: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

152 | P a g e

were similar to those of spray-dried mannitol with 0.5% w/w Gum Arabic, which showed that

adding 2% w/w citric acid significantly increased the BET surface area of mannitol to 5.1 ± 0.2

m2 g-1, compared with 1.8 ± 0.1 m2 g-1 for the spray-dried mannitol/Gum Arabic mixture with

0.5% w/w citric acid. It has been found that using emulsifiers such as gum Arabic as an additive

could improve the dispersity of the templating agents in the main structure during the

solidification step (particle formation) of the spray-drying process.

10.1.5 Effect of acidity on a new templating approach for high-porosity spray-dried particle production

It can be observed in Tables 6.2 and 7.1 that adding 1% w/w citric acid as a templating material

significantly increased the BET surface area of lactose to 20.8 ± 0.9 m2 g-1. However, adding

further citric acid content (3 w/w%) decreased the BET surface area of lactose to 12.3±0.8 m2g-1.

The BET surface areas of the mannitol particles, using citric acid as a templating acid, can be

altered from 1.91 ± 0.01 m2 g-1 for 1% w/w citric acid, compared with 1.6 ± 0.1 m2 g-1 for the

spray-dried mannitol with 2% w/w citric acid. These results showed that the surface areas both of

the resultant lactose and mannitol (core materials), decreased considerably by increasing the

concentrations of templating acid. This result suggested that there is a link between crystallinity

of core material in the spray-dried powders and the porosity of the ethanol-washed powders, as

will now be explained. Crystalline solids are known to have distinctive and organized internal

structures, called crystal lattices, while the atoms and molecules in amorphous solids are

arranged randomly with no regular patterns (Averill and Eldredge, 2007). During the

crystallization process, molecules of each component are more likely to join their own crystal

structure (Jones and Mirkin, 2013). Therefore, each growing crystal tends to consist of only one

type of molecule.

Lower degrees of core material crystallinity for spray-dried powders mean that the crystal

structures have become less developed, which in turn means that the number of citric acid

molecules incorporated in the whole core material structure (amorphous and crystal) has

increased. Washing with ethanol removes the citric acid molecules, either scattered in the

amorphous portion of the core material or accumulated on the surface of the crystals, rather than

inside the crystal lattice. Therefore, the porosity and BET surface areas of the powders have

Page 171: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

153 | P a g e

increased as the amorphous content has increased, which may be connected with the increase in

the amount of incorporated citric acid molecules due to the decrease in the degree of crystallinity

for the spray-dried powders.

The degrees of crystallinity measured using moisture sorption tests and XRD methods confirmed

that the crystallinity of the lactose increased when the templating-acid concentration was raised.

Spray drying of lactose solution with 3% (w/w) citric acid, resulted in a significant reduction in

peak height to about 2.9% ± 0.6% compared with 5.4% ± 0.5% for 1 % (w/w) citric acid. The

intensities of the characteristic peak at 2θ= 20.1° for the lactose in the spray-dried powders with

1% (w/w) and 3% (w/w) citric acid concentrations were 290 ± 20 cps and 410 ± 20 cps,

respectively.

However, these results have shown that increasing the porosity of naturally-crystalline materials,

such as mannitol may be much more difficult than lactose. The WLF equation suggests that the

rate of crystallization is related to the difference between the material temperature (T) and its

glass-transition temperature (T-Tg) which has been observed by other researchers (Bhandari and

Howes, 1999; Islam et al., 2010). This situation means that the lower BET surface area for

mannitol compared with lactose might be due to the lower glass-transition temperature for the

mixture of mannitol (10.7ºC; Telang et al., 2003) with citric acid compared with the glass-

transition temperature of lactose (101ºC; Roos and Karel, 1991) with the same citric acid

concentration. This situation means that, as the glass-transition temperatures decrease, the

difference between the glass-transition temperatures (Tg) of the materials and the process

temperatures (T) increases, thereby enhancing the crystallization rate, according to the Williams–

Landel–Ferry equation.

The results from experiments on the effects of WPI concentrations on the BET surface areas of

mannitol have agreed well with the above-mentioned suggestion, since adding WPI 0.2% (w/w)

significantly increased the particle porosity. The BET surface area of mannitol increased from

1.91 ± 0.01 to about 3.2 ± 0.9 m2 g-1 for 1% w/w citric acid, compared with 1.6 ± 0.1 to 5.1 ± 1.1

m2 g-1 for the spray-dried mannitol/WPI mixture with 2% w/w citric acid (0.2% w/w WPI). It

also have been found that adding more WPI (0.5% w/w WPI ) to the mannitol/citric acid

solutions significantly increased the BET surface of mannitol up to 8.0 ± 0.4 m2 g-1 (Table 7.2).

Page 172: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

154 | P a g e

This result is also consistent with the effect of adding WPI on the product crystallinity, since it

has been found that the presence of proteins at low concentrations significantly affects the

product crystallinity (Jouppila and Roos, 1994a; Haque and Roos, 2006).

10.1.6 Solubility and Content Uniformity Enhancement of Poorly Water-Soluble Drugs Using Highly-Porous Excipients

In the next step, a new adsorption method has been developed to investigate enhancement of the

dissolution rate of poorly-water soluble drugs and better blending uniformity of drug dosage

using mannitol particles with high porosity and high surface area. The engineered particles were

used as a porous excipient in an adsorption technique to incorporate the drug molecules onto this

porous structure. Indomethacin, nifedipine, and acetaminophen as model drugs were studied. The

results of this study showed that adsorption of the drug in the pore-space of highly-porous

mannitol resulted in better blending uniformity of drug dosage, as well as faster drug release rate

due to the nano-confinement of drugs into the porous excipients and the better penetration of the

solvent into the drug-porous mannitol matrix. Porous mannitol particles with a surface area and

pore volume of 6.3 ± 0.1 m2g-1 and 0.036 ± 0.002 mlg-1, respectively, were drug loaded with two

different concentrations of indomethacin and nifedipine. This adsorption method for the whole

range of drug loadings (low to high) was successful and, with increasing the drug solution

concentrations, the drug contents for porous mannitol were increased. The results of drug loading

for nifedipine showed an increase from 3.2 ± 0.1 w/w % for a 0.08 M drug solution to 9.1±0.3

w/w % drug loading for a 0.16 M drug solution, while indomethacin had slightly better

performance for the adsorption process, with 4.1 ± 0.2 w/w % and 12.6 ± 0.4 w/w % for 0.08 M

and 0.16 M concentrations of indomethacin, respectively, in the final formulation. Low

variability in the drug content of all drug-loaded samples was also found for the finished

products, even for low-dose drug products. This result indicated highly-uniform blends with a

percentage relative standard deviation of less than 4% for drug-loaded mannitol in both

nifedipine and indomethacin. This method gave a significant enhancement in the dissolution rate

for both drugs due to nano-confinement of drugs into porous excipients and high solubility of

porous mannitol, with 80% drug release within the first fifteen minutes for the drug-loaded

samples.

Page 173: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

155 | P a g e

10.2 Recommendations and Future Work

Further areas of interest include studying the possible application of engineered particles with

highly-porous structures in food infusions, flavoring, and encapsulating processes. The current

wet-adsorption method can be used to enhance the release rate of many other poorly water-

soluble drugs using different types of highly-porous excipients. The current templating method

might be a potential process for high-porosity particle production of other food and non-food

materials, such as calcium and magnesium salts, sodium or potassium chloride, and sodium

carbonate. However, the templating research results have shown that increasing the crystallinity

of spray-dried powder decreases the porosity of the processed powder. Hence, it can be assumed

that increasing the porosity of naturally-crystalline materials, such as sodium carbonate, may be

much more difficult. This additional difficulty with materials that crystallize readily during spray

drying can be seen in the results here when using mannitol (readily crystallizes during spray

drying) and lactose (crystallizes less readily). Controlling the degree of crystallinity and crystal

size of these materials by applying different additives might be helpful, which can be

investigated.

The micrographs taken in this thesis showed some morphological changes for both the lactose

and mannitol powders, before and after citric acid removal (ethanol washing). More porosity

appeared to be present on the surface and inside of the particles after removing the templating

agent, while the surfaces of the spray-dried powder appeared to be non-porous and amorphous.

However, a broad pore size distribution in the range of 2-200 nm has been found for both porous

mannitol and lactose particles, suggesting that the templating agents were not evenly distributed

within the core materials during the spray-drying process. By contrast, mesoporous silica, as a

common carrier for controlled drug delivery, has a narrow pore size distribution (PSD) in the

range of mesopores (2-50 nm), resulting from a homogenous distribution of templates during the

synthesis procedure (Millqvist-Fureby et al., 2014). In order to achieve desired PSD and to

provide a high degree of control over the pore size distributions, different approaches can be

taken in future studies. Study of a multicomponent particle-drying model, such as the work by

Wang and Langrish (2009), can be developed to describe the component migration during spray

drying and distribution of components inside the spray-dried particles. This model takes the

diffusion processes of water and dissolved solids, and the effects of the solubilities of the solids

Page 174: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

156 | P a g e

into consideration to simulate the spray-drying process. Another possible study is the use of

organic templating agents, such as polyvinylalcohol (PVA) with different particle sizes through a

similar template-driven self-assembly technique to produce porous particles with tuneable and

narrow pore size distributions.

Page 175: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

157 | P a g e

REFERENCES

Adhikari, B., Howes, T., Bhandari, B. R. & Langrish, T. A. G. 2009. Effect of Addition of Proteins on the Production of Amorphous Sucrose Powder through Spray Drying. Journal of Food Engineering, 94 (2), 144-153.

Adhikari, B., Howes, T., Bhandari, B. R. & Troung, V. 2004. Effect of Addition of Maltodextrin on Drying Kinetics and Stickiness of Sugar and Acid-Rich Foods During Convective Drying: Experiments and Modelling. Journal of Food Engineering, 62 (1), 53-68.

Aguilera, J. M., del, V. J. M. & Karel, M. 1995. Caking Phenomena in Amorphous Food Powders. Trends in Food Science & Technology, 6 (5), 149-55.

Akao, K.-i., Okubo, Y., Asakawa, N., Inoue, Y. & Sakurai, M. 2001. Infrared Spectroscopic Study on the Properties of the Anhydrous Form II of Trehalose. Implications for the Functional Mechanism of Trehalose as a Biostabilizer. Carbohydrate Research, 334 (3), 233-241.

Arvanitoyannis, I. & Blanshard, J. M. V. 1994. Rates of Crystallization of Dried Lactose-Sucrose Mixtures. Journal of Food Science, 59 (1), 197-205.

Averill, B. & Eldredge, P., 2007. Chemistry: Principles, Patterns, and Applications, San Francisco, Pearson Benjamin Cummings, 1070-1072.

Avrami, M. 1940. Kinetics of Phase Change. II. Transformation-Time Relations for Random Distribution of Nuclei. The Journal of Chemical Physics, 8 (2), 212-24.

Babu, N. J. & Nangia, A. 2011. Solubility Advantage of Amorphous Drugs and Pharmaceutical Cocrystals. Crystal Growth & Design, 11 (7), 2662-2679.

Bakaltcheva, I., O'Sullivan, A. M., Hmel, P. & Ogbu, H. 2007. Freeze-Dried Whole Plasma: Evaluating Sucrose, Trehalose, Sorbitol, Mannitol and Glycine as Stabilizers. Thrombosis Research, 120 (1), 105-116.

Balgis, R., Anilkumar, G. M., Sago, S., Ogi, T. & Okuyama, K. 2012. Nanostructured Design of Electrocatalyst Support Materials for High-Performance PEM Fuel Cell Application. Journal of Power Sources, 203, 26-33.

Balgis, R., Iskandar, F., Ogi, T., Purwanto, A. & Okuyama, K. 2011. Synthesis of Uniformly Porous NiO/ZrO2 Particles. Materials Research Bulletin, 46 (5), 708-715.

Behera, S., Ghanty, S., Ahmad, F., Santra, S. & Banerjee, S. 2012. UV-Visible Spectrophotometric Method Development and Validation of Assay of Paracetamol Tablet Formulation. Journal of Analytical Bioanalytical Techniques, 3 (6), 1000151/1-1000151/6.

Beiner, M., Rengarajan, G. T., Pankaj, S., Enke, D. & Steinhart, M. 2007. Manipulating the Crystalline State of Pharmaceuticals by Nanoconfinement. Nano Letters, 7 (5), 1381-1385.

Bhandari, B. 2008. Crystallisation of Amorphous Lactose in the Presence Lactic Acid Factors Affecting Crystallization of Amorphous Solids. International Union of Food Science & Technology. Shanghai, China.

Bhandari, B. R., Datta, N., Crooks, R., Howes, T. & Rigby, S. 1997a. A Semi-Empirical Approach to Optimise the Amount of Drying Aids Required to Spray Dry Sugar-Rich Foods. Drying Technology, 15 (10), 2509-2525.

Bhandari, B. R., Datta, N. & Howes, T. 1997b. Problems Associated with Spray Drying of Sugar-Rich Foods. Drying Technology, 15 (2), 671-684.

Page 176: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

158 | P a g e

Bhandari, B. R. & Howes, T. 1999. Implication of Glass Transition for the Drying and Stability of Dried Foods. Journal of Food Engineering, 40 (1), 71-79.

Boissiere, C., Grosso, D., Chaumonnot, A., Nicole, L. & Sanchez, C. 2011. Aerosol Route to Functional Nanostructured Inorganic and Hybrid Porous Materials. Advanced Materials, 23 (5), 599-623.

Buckton, G. & Darcy, P. 1995. The Influence of Additives on the Recrystallisation of Amorphous Spray Dried Lactose. International Journal of Pharmaceutics, 121 (1), 81-87.

Buckton, G. & Darcy, P. 1996. Water Mobility in Amorphous Lactose Below and Close to the Glass Transition Temperature. International Journal of Pharmaceutics, 136 (1), 141-146.

Burger, A., Henck, J. O., Hetz, S., Rollinger, J. M., Weissnicht, A. A. & Stöttner, H. 2000. Energy/Temperature Diagram and Compression Behavior of the Polymorphs of D-Mannitol. Journal of Pharmaceutical Sciences, 89 (4), 457-468.

Burnett, D. J., Thielmann, F. & Booth, J. 2004. Determining the Critical Relative Humidity for Moisture-Induced Phase Transitions. International Journal of Pharmaceutics, 287 (1-2), 123-133.

Burton, A. W., Ong, K., Rea, T. & Chan, I. Y. 2009. On the Estimation of Average Crystallite Size of Zeolites from the Scherrer Equation: A Critical Evaluation of Its Application to Zeolites with One-Dimensional Pore Systems. Microporous and Mesoporous Materials, 117 (1–2), 75-90.

Cai, S. & Singh, B. R. 2000. Determination of the Secondary Structure of Proteins from Amide I and Amide III Infrared Bands using Partial Least-Square Method. ACS Symposium Series, 750 (Journal Article), 117-129.

Caillard, R., Guillet-Nicolas, R., Kleitz, F. & Subirade, M. 2012. Tabletability of Whey Protein Isolates. International Dairy Journal, 27, 92-98.

Cartilier, L. H. & Moes, A. J. 1989. Effect of Drug Agglomerates Upon the Kinetics of Mixing of Low-Dosage Cohesive Powder Mixtures. Drug Development and Industrial Pharmacy, 15 (12), 1911-31.

Charnay, C., Bégu, S., Tourné-Péteilh, C., Nicole, L., Lerner, D. A. & Devoisselle, J. M. 2004. Inclusion of Ibuprofen in Mesoporous Templated Silica: Drug Loading and Release Property. European Journal of Pharmaceutics and Biopharmaceutics, 57 (3), 533-540.

Chen, J. I. L. & Ozin, G. A. 2009. Heterogeneous Photocatalysis with Inverse Titania Opals: Probing Structural and Photonic Effects. Journal of Materials Chemistry, 19 (18), 2675-2678.

Cinelli, P., Schmid, M., Bugnicourt, E., Wildner, J., Bazzichi, A., Anguillesi, I. & Lazzeri, A. 2014. Whey Protein Layer Applied on Biodegradable Packaging Film to Improve Barrier Properties while Maintaining Biodegradability. Polymer Degradation and Stability, 108, 151-157.

Corrigan, D. O., Corrigan, O. I. & Healy, A. M. 2004. Predicting the Physical State of Spray Dried Composites: Salbutamol Sulphate/Lactose and Salbutamol Sulphate/Polyethylene Glycol Co-Spray Dried Systems. International Journal of Pharmaceutics, 273 (1), 171-182.

Cruz, M. A. A., Passos, M. L. & Ferreira, W. R. 2005. Final Drying of Whole Milk Powder in Vibrated-Fluidized Beds. Drying Technology, 23 (9-11), 2021-2037.

Page 177: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

159 | P a g e

Danesh, A., Connell, S. D., Davies, M. C., Roberts, C. J., Tendler, S. J. B., Williams, P. M. & Wilkins, M. J. 2001. An In Situ Dissolution Study of Aspirin Crystal Planes (100) and (001) by Atomic Force Microscopy. Pharmaceutical Research, 18 (3), 299-303.

Dangaran, K. L. & Krochta, J. M. 2006. Kinetics of Sucrose Crystallization in Whey Protein Films. Journal of Agricultural and Food Chemistry, 54 (19), 7152-7158.

Dangaran, K. L., Renner‐Nantz, J. & Krochta, J. M. 2006. Whey Protein-Sucrose Coating Gloss and Integrity Stabilization by Crystallization Inhibitors. Journal of Food Science, 71 (3), E152-E157.

Das, D. & Langrish, T. A. G. 2012. An Activated-State Model for the Prediction of Solid-Phase Crystallization Growth Kinetics in Dried Lactose Particles. Journal of Food Engineering, 109 (4), 691-700.

de Souza, A., Andreani, T., Nunes, F., Cassimiro, D., de Almeida, A., Ribeiro, C., Sarmento, V., Gremião, M., Silva, A. & Souto, E. 2012. Loading of Praziquantel in the Crystal Lattice of Solid Lipid Nanoparticles. Journal of Thermal Analysis and Calorimetry, 108 (1), 353-360.

De Wit, J. N., 2001. Lecturer's Handbook on Whey and Whey Products, Belgium, European Whey Products Association, 67-68.

Debord, B., Lefebvre, C., Guyot-Hermann, A. M., Hubert, J., Bouché, R. & Cuyot, J. C. 1987. Study of Different Crystalline forms of Mannitol: Comparative Behaviour under Compression. Drug Develpment and Industrial Pharmacy, 13 (9-11), 1533-1546.

Dhillon, B., Goyal, N. K., Malviya, R. & Sharma, P. K. 2014. Poorly Water Soluble Drugs: Change in Solubility for Improved Dissolution Characteristics a Review. Global Journal of Pharmacology, 8 (1), 26-35.

Directive, C. 2003/63/EC. Specifications and Control Tests on the Finished Product. Official Journal of the European Union.

Doktorovová, S., Araújo, J., Garcia, M. L., Rakovský, E. & Souto, E. B. 2010. Formulating Fluticasone Propionate in Novel PEG-Containing Nanostructured Lipid Carriers (PEG-NLC). Colloids and Surfaces B: Biointerfaces, 75 (2), 538-542.

Downton, G. E., Flores-Luna, J. L. & King, C. J. 1982. Mechanism of Stickiness in Hygroscopic, Amorphous Powders. Industrial & Engineering Chemistry Fundamentals, 21 (4), 447-51.

Drapier-Beche, N., Fanni, J., Parmentier, M. & Vilasi, M. 1997. Evaluation of Lactose Crystalline Forms by Nondestructive Analysis. Journal of Dairy Science, 80 (3), 457-463.

Dressman, J. B. & Krämer, J., 2005. Pharmaceutical Dissolution Testing, Boca Raton, Taylor & Francis.

Emmanuelawati, I., Yang, J., Zhang, J., Zhang, H., Zhou, L. & Yu, C. 2013. Low-Cost and Large-Scale Synthesis of Functional Porous Materials for Phosphate Removal with High Performance. Nanoscale, 5 (13), 6173-6180.

FDA 2003. Guidance for Industry: Powder Blends and Finished Dosage Units – Stratified in-Process Dosage Unit Sampling and Assessment.

Feng, L. & Kamal, M. R. 2004. Distributions of Crystal Size from DSC Melting Traces for Polyethylenes. The Canadian Journal of Chemical Engineering, 82 (6), 1239-1251.

Fiorilli, S., Tallia, F., Pontiroli, L., Vitale-Brovarone, C. & Onida, B. 2013. Spray-Dried Mesoporous Silica Spheres Functionalized with Carboxylic Groups. Materials Letters, 108, 118-121.

Page 178: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

160 | P a g e

Foster, K. D., Bronlund, J. E. & Paterson, A. H. J. 2004. The Contribution of Milk Fat Towards the Caking of Dairy Powders. International Dairy Journal, 15 (1), 85-91.

Foster, K. D., Bronlund, J. E. & Paterson, A. H. J. 2006. Glass Transition Related Cohesion of Amorphous Sugar Powders. Journal of Food Engineering, 77 (4), 997-1006.

Garcia, T. P. & Prescott, J. K. 2008. Blending and Blend Uniformity. In: Thomas P. Garcia & Prescott, J. K. (eds.) Pharmaceutical Dosage Forms: Tablets. 111-174.

Garekani, H. A., Ford, J. L., Rubinstein, M. H. & Rajabi-Siahboomi, A. R. 1999. Formation and Compression Characteristics of Prismatic Polyhedral and Thin Plate-Like Crystals of Paracetamol. International Journal of Pharmaceutics, 187 (1), 77-89.

Gavish, E. & Friedman, G. M. 1973. Quantitative Analysis of Calcite and Mg-Calcite by X-Ray Diffraction: Effect of Grinding on Peak Height and Peak Area. Sedimentology, 20 (3), 437-444.

Ghodke, D. S., Chaulang, G. M., Patil, K. S., Nakhat, P. D., Yeole, P. G., Naikwade, N. S. & Magdum, C. S. 2010. Solid State Characterization of Domperidone: Hydroxypropyl-β-Cyclodextrin Inclusion Complex. Indian Journal of Pharmaceutical Sciences, 72 (2), 245-249.

Giordano, F., Rossi, A., Bettini, R., Savioli, A., Gazzaniga, A. & Novák, C. 2002. Thermal Behavior of Paracetamol-Polymeric Excipients Mixtures. Journal of Thermal Analysis and Calorimetry, 68 (2), 575-590.

Gombas, A., Szabo-Revesz, P., Kata, M., Regdon, G., Jr. & Eros, I. 2002. Quantitative Determination of Crystallinity of α-Lactose Monohydrate by DSC. Journal of Thermal Analysis and Calorimetry, 68 (2), 503-510.

Gordon, M. & Taylor, J. S. 1952. Ideal Copolymers and the Second-Order Transitions of Synthetic Rubbers. I. Noncrystalline Copolymers. Journal of Applied Chemistry, 2, 493-500.

Grigorov, P. I., Glasser, B. J. & Muzzio, F. J. 2013. Formulation and Manufacture of Pharmaceuticals by Fluidized-Bed Impregnation of Active Pharmaceutical Ingredients onto Porous Carriers. AIChE Journal, 59 (12), 4538-4552.

Hancock, B. C. & Parks, M. 2000. What Is the True Solubility Advantage for Amorphous Pharmaceuticals? Pharmaceutical research, 17 (4), 397-404.

Hanus, M. J. & Langrish, T. A. G. 2007. Re-Entrainment of Wall Deposits from a Laboratory-Scale Spray Dryer. Asia-Pacific Journal of Chemical Engineering, 2 (2), 90-107.

Haque, M. K. & Roos, Y. H. 2004. Water Plasticization and Crystallization of Lactose in Spray-dried Lactose/Protein Mixtures. Journal of Food Science, 69 (1), FEP23-FEP29.

Haque, M. K. & Roos, Y. H. 2005. Crystallization and X-ray Diffraction of Spray-Dried and Freeze-Dried Amorphous Lactose. Carbohydrate Research, 340 (2), 293-301.

Haque, M. K. & Roos, Y. H. 2006. Differences in the Physical State and Thermal Behavior of Spray-Dried and Freeze-Dried Lactose and Lactose/Protein Mixtures. Innovative Food Science and Emerging Technologies, 7 (1), 62-73.

Hecq, J., Deleers, M., Fanara, D., Vranckx, H. & Amighi, K. 2005. Preparation and Characterization of Nanocrystals for Solubility and Dissolution Rate Enhancement of Nifedipine. International Journal of Pharmaceutics, 299 (1–2), 167-177.

Heikkilä, T., Salonen, J., Tuura, J., Hamdy, M. S., Mul, G., Kumar, N., Salmi, T., Murzin, D. Y., Laitinen, L., Kaukonen, A. M., Hirvonen, J. & Lehto, V. P. 2007. Mesoporous Silica Material TUD-1 as a Drug Delivery System. International Journal of Pharmaceutics, 331 (1), 133-138.

Page 179: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

161 | P a g e

Hillerström, A., Van Stam, J., Andersson, M., Karlstads, u., Avdelningen för kemi och biomedicinsk, v. & Fakulteten för teknik- och, n. 2009. Ibuprofen Loading into Mesostructured Silica Using Liquid Carbon Dioxide as a Solvent. Green Chemistry, 11 (5), 662-667.

Ho, R., Naderi, M., Heng, J. Y. Y., Williams, D. R., Thielmann, F., Bouza, P., Keith, A. R., Thiele, G. & Burnett, D. J. 2012. Effect of Milling on Particle Shape and Surface Energy Heterogeneity of Needle-Shaped Crystals. Pharmaceutical Research, 29 (10), 2806-2816.

Hogan, S. A. & O'Callaghan, D. J. 2010. Influence of Milk Proteins on the Development of Lactose-Induced Stickiness in Dairy Powders. International Dairy Journal, 20 (3), 212-221.

Hogan, S. A., O'Callaghan, D. J. & Bloore, C. G. 2009. Application of fluidised bed stickiness apparatus to dairy powder production. Milchwissenschaft, 64 (Copyright (C) 2012 American Chemical Society (ACS). All Rights Reserved.), 308-312.

Huang, C.-Y. & Sherry Ku, M. 2010. Prediction of Drug Particle Size and Content Uniformity in Low-Dose Solid Dosage Forms. International Journal of Pharmaceutics, 383 (1–2), 70-80.

Hulse, W. L., Forbes, R. T., Bonner, M. C. & Getrost, M. 2009. Influence of Protein on Mannitol Polymorphic form Produced During Co-Spray Drying. International journal of pharmaceutics, 382 (1-2), 67-72.

Huybrechts, D. R. C., Bruycker, L. D. & Jacobs, P. A. 1990. Oxyfunctionalization of Alkanes with Hydrogen Peroxide on Titanium Silicalite. Nature, 345 (6272), 240-242.

Ibach, A. & Kind, M. 2007. Crystallization Kinetics of Amorphous Lactose, Whey-Permeate and Whey Powders. Carbohydrate Research, 342 (10), 1357-1365.

Imtiaz-Ul-Islam, M. & Langrish, T. A. G. 2009. Comparing the Crystallization of Sucrose and Lactose in Spray Dryers. Food and Bioproducts Processing, 87 (2), 87-95.

Islam, M. I.-U., Edrisi, M. & Langrish, T. 2013. Improving Process Yield by Adding WPI to Lactose During Crystallization and Spray Drying Under High-Humidity Conditions. Drying Technology, 31 (4), 393-404.

Islam, M. I. U., Langrish, T. A. G. & Chiou, D. 2010. Particle Crystallization During Spray Drying in Humid Air. Journal of Food Engineering, 99 (1), 55-62.

Izutsu, K.-i., Ocheda, S. O., Aoyagi, N. & Kojima, S. 2004. Effects of Sodium Tetraborate and Boric Acid on Nonisothermal Mannitol Crystallization in Frozen Solutions and Freeze-Dried Solids. International Journal of Pharmaceutics, 273 (1-2), 85-93.

Jain, S., Sandhu, P., Gurjar, M. & Malvi, R. 2012. Solubility Enhancement by Solvent Deposition Technique: An Overview. Asian Journal of Pharmaceutical and Clinical Research, 5 (4), 15-19.

Jang, H. R., Oh, H. J., Kim, J. H. & Jung, K. Y. 2013. Synthesis of Mesoporous Spherical Silica via Spray Pyrolysis: Pore Size Control and Evaluation of Performance in Paclitaxel Pre-Purification. Microporous and Mesoporous Materials, 165, 219-227.

Jayasundera, M., Adhikari, B. P., Adhikari, R. & Aldred, P. 2010. The Effect of Food-Grade Low-Molecular-Weight Surfactants and Sodium Caseinate on Spray Drying of Sugar-Rich Foods. Food Biophysics, 5 (2), 128-137.

Jelen, P. 2009. Dried Whey, Whey Proteins, Lactose and Lactose Derivative Products. Dairy Powders and Concentrated Products. Wiley-Blackwell, 255-267.

Jenness, R. 1988. Composition of Milk. In: Wong, N. P., Jenness, P., Keeney, M. & Marth, E. H. (eds.) Fundamentals of Dairy Chemistry. Third Edition ed.: Springer US, pp. 1-38.

Page 180: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

162 | P a g e

Jenning, V., Thünemann, A. F. & Gohla, S. H. 2000. Characterisation of a Novel Solid Lipid Nanoparticle Carrier System Based on Binary Mixtures of Liquid and Solid Lipids. International Journal of Pharmaceutics, 199 (2), 167-177.

Ji, C., Barrett, A., Poole-Warren, L. A., Foster, N. R. & Dehghani, F. 2010. The Development of a Dense Gas Solvent Exchange Process for the Impregnation of Pharmaceuticals into Porous Chitosan. International Journal of Pharmaceutics, 391 (1–2), 187-196.

Jones, M. R. & Mirkin, C. A. 2013. Bypassing the Limitations of Classical Chemical Purification with DNA-Programmable Nanoparticle Recrystallization. Angewandte Chemie International Edition, 52 (10), 2886-2891.

Jordan, D. D. 1993. Optical Crystallographic Characteristics of Some USP Drugs. Journal of Pharmaceutical Sciences, 82 (12), 1269-1271.

Jouppila, K., Kansikas, J. & Roos, Y. H. 1997. Glass Transition, Water Plasticization, and Lactose Crystallization in Skim Milk Powder. Journal of Dairy Science, 80 (12), 3152-3160.

Jouppila, K., Kansikas, J. & Roos, Y. H. 1998. Crystallization and X-Ray Diffraction of Crystals Formed in Water-Plasticized Amorphous Lactose. Biotechnology Progress, 14 (2), 347-350.

Jouppila, K. & Roos, Y. H. 1994a. Glass Transitions and Crystallization in Milk Powders. Journal of Dairy Science, 77 (10), 2907-2915.

Jouppila, K. & Roos, Y. H. 1994b. Water Sorption and Time-Dependent Phenomena of Milk Powders. Journal of Dairy Science, 77 (7), 1798-1808.

Karehill, P. G., Glazer, M. & Nyström, C. 1990. Studies on Direct Compression of Tablets. XXIII. The Importance of Surface Roughness for the Compactability of some Directly Compressible Materials with Different Bonding and Volume Reduction Properties. International Journal of Pharmaceutics, 64 (1), 35-43.

Katainen, E., Niemelä, P., Harjunen, P., Suhonen, J. & Järvinen, K. 2005. Evaluation of the amorphous content of lactose by solution calorimetry and Raman spectroscopy. Talanta, 68 (1), 1-5.

Kedward, C. J., Macnaughtan, W., Blanshard, J. M. V. & Mitchell, J. R. 1998. Crystallization Kinetics of Lactose and Sucrose Based on Isothermal Differential Scanning Calorimetry. Journal of Food Science, 63 (2), 192-197.

Khan, A. I., Lei, L., Norquist, A. J. & O'Hare, D. 2001. Intercalation and Controlled Release of Pharmaceutically Active Compounds from a Layered Double Hydroxide. Chemical Communications, 10.1039/b106465g(22), 2342-2343.

Kibbe, A. H., 2000. Handbook of pharmaceutical excipients, American Pharmaceutical Association.

Lai, H.-M. & Schmidt, S. J. 1990. Lactose Crystallization in Skim Milk Powder Observed by Hydrodynamic Equilibria, Scanning Electron Microscopy and 2H Nuclear Magnetic Resonance. Journal of Food Science, 55 (4), 994-999.

Langrish, T. & Wang, E. H.-L. 2006. Crystallization of Powders of Spray-Dried Lactose, Skim Milk and Lactose-Salt Mixtures. International Journal of Food Engineering, 2 (4), 8-15.

Langrish, T. A. G. 2008. Assessing the Rate of Solid-Phase Crystallization for Lactose: The Effect of the Difference between Material and Glass-Transition Temperatures. Food Research International, 41 (6), 630-636.

Langrish, T. A. G. & Wang, S. 2009. Crystallization Rates for Amorphous Sucrose and Lactose Powders from Spray Drying: A Comparison. Drying Technology, 27 (4), 606-614.

Page 181: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

163 | P a g e

Lee, S. Y., Gradon, L., Janeczko, S., Iskandar, F. & Okuyama, K. 2010. Formation of Highly Ordered Nanostructures by Drying Micrometer Colloidal Droplets. ACS Nano, 4 (8), 4717-4724.

Lehto, V.-P., Tenho, M., Vähä-Heikkilä, K., Harjunen, P., Päällysaho, M., Välisaari, J., Niemelä, P. & Järvinen, K. 2006. The Comparison of Seven Different Methods to Quantify the Amorphous Content of Spray Dried Lactose. Powder Technology, 167 (2), 85-93.

Leuner, C. & Dressman, J. 2000. Improving Drug Solubility for Oral Delivery Using Solid Dispersions. European Journal of Pharmaceutics and Biopharmaceutics, 50 (1), 47-60.

Listiohadi, Y., Hourigan, J., Sleigh, R. & Steele, R. 2009. Thermal Analysis of Amorphous Lactose and α-Lactose Monohydrate. Dairy Science & Technology, 89 (1), 43-67.

Littringer, E. M., Mescher, A., Eckhard, S., Schrottner, H., Langes, C., Fries, M., Griesser, U., Walzel, P. & Urbanetz, N. A. 2012. Spray Drying of Mannitol as a Drug Carrier-The Impact of Process Parameters on Product Properties. Drying Technology, 30 (1), 114-124.

Liu, P., De Wulf, O., Laru, J., Heikkilä, T., van Veen, B., Kiesvaara, J., Hirvonen, J., Peltonen, L. & Laaksonen, T. 2013. Dissolution Studies of Poorly Soluble Drug Nanosuspensions in Non-sink Conditions. AAPS PharmSciTech, 14 (2), 748-756.

Liu, R., 2008. Water-Insoluble Drug Formulation, CRC Press, page 1. Lorenz, H. & Beckmann, W. 2013. Purification by Crystallization. 129-148. Maas, S. G., Schaldach, G., Littringer, E. M., Mescher, A., Griesser, U. J., Braun, D. E., Walzel,

P. E. & Urbanetz, N. A. 2011. The Impact of Spray Drying Outlet Temperature on the Particle Morphology of Mannitol. Powder Technology, 213 (1), 27-35.

Majano, G., Restuccia, A., Santiago, M. & Perez-Ramirez, J. 2012. Assembly of a Hierarchical Zeolite-Silica Composite by Spray Drying. CrystEngComm, 14 (18), 5985-5991.

Marsh, R. D. L. & Blanshard, J. M. V. 1988. The Application of Polymer Crystal Growth Theory to the Kinetics of Formation of the B-Amylose Polymorph in a 50% Wheat-Starch Gel. Carbohydrate Polymers, 9 (4), 301-17.

Marshall, P. V. & York, P. 1991. The Compaction Properties of Nitrofurantoin Samples Crystallised from Different Solvents. International Journal of Pharmaceutics, 67 (1), 59-65.

Masoudpanah, S. M. & Seyyed Ebrahimi, S. A. 2011. Fe/Sr Ratio and Calcination Temperature Effects on Processing of Nanostructured Strontium Hexaferrite Thin Films by a Sol–Gel Method. Research on Chemical Intermediates, 37 (2), 259-266.

Mazzobre, M. a. F., Soto, G., Aguilera, J. M. & Buera, M. a. P. 2001. Crystallization Kinetics of Lactose in Sytems Co-Lyofilized with Trehalose. Analysis by Differential Scanning Calorimetry. Food Research International, 34 (10), 903-911.

Mazzobre, M. F., Aguilera, J. M. & Buera, M. P. 2003. Microscopy and Calorimetry as Complementary Techniques to Analyze Sugar Crystallization from Amorphous Systems. Carbohydrate Research, 338 (6), 541-548.

McKenna, A. & McCafferty, D. F. 1982. Effect of Particle Size on the Compaction Mechanism and Tensile Strength of Tablets. Journal of Pharmacy and Pharmacology, 34 (6), 347-351.

Mellaerts, R., Aerts, C. A., Humbeeck, J. V., Augustijns, P., den Mooter, G. V. & Martens, J. A. 2007. Enhanced Release of Itraconazole from Ordered Mesoporous SBA-15 Silica Materials. Chemical Communications, 10.1039/b616746b(13), 1375-1377.

Page 182: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

164 | P a g e

Mellaerts, R., Jammaer, J. A. G., Van Speybroeck, M., Chen, H., Humbeeck, J. V., Augustijns, P., Van den Mooter, G. & Martens, J. A. 2008. Physical State of Poorly Water Soluble Therapeutic Molecules Loaded into SBA-15 Ordered Mesoporous Silica Carriers: A Case Study with Itraconazole and Ibuprofen. Langmuir, 24 (16), 8651-8659.

Melo, J. S., Sen, D., Mazumder, S. & D'Souza, S. F. 2013. Spray Drying As a Novel Technique for Obtaining Microbial Imprinted Microspheres and Its Application in Filtration. Soft Matter, 9 (3), 805-810.

Miao, S. & Roos, Y. H. 2005. Crystallization Kinetics and X‐ray Diffraction of Crystals Formed in Amorphous Lactose, Trehalose, and Lactose/Trehalose Mixtures. Journal of Food Science, 70 (5), E350-E358.

Milligan, T. P., Morris, H. C., Hammond, P. M. & Price, C. P. 1994. Studies on Paracetamol Binding to Serum Proteins. Annals of Clinical Biochemistry, 31, 492-6.

Millqvist-Fureby, A., Larsson, A., Järn, M., Speets, E., Ahniyaz, A., Macakova, L. & Elofsson, U. 2014. Chapter 25 - Mesoporous Solid Carrier Particles in Controlled Delivery and Release. In: Gaonkar, A. G., Vasisht, N., Khare, A. R. & Sobel, R. (eds.) Microencapsulation in the Food Industry. San Diego: Academic Press, 299-319.

Modi, A. & Tayade, P. 2006. Enhancement of Dissolution Profile by Solid Dispersion (Kneading) Technique. AAPS PharmSciTech, 7 (3), E87-E92.

Modler, H. W. & Emmons, D. B. 1978. Calcium as an Adjuvant for Spray-Drying Acid Whey. Journal of Dairy Science, 61 (3), 294.

Morgan, C. & Vesey, G., 2009. Encyclopedia of Microbiology, Middletown, Academic Press. Murphy, B. M., Prescott, S. W. & Larson, I. 2005. Measurement of Lactose Crystallinity Using

Raman Spectroscopy. Journal of Pharmaceutical and Biomedical Analysis, 38 (1), 186-190.

Nalawade, P., Aware, B., Kadam, V. J. & Hirlekar, R. S. 2009. Layered Double Hydroxides: A Review. Journal of Scientific & Industrial Research, 68 (4), 267-272.

Nandiyanto, A. B. D., Arutanti, O., Ogi, T., Iskandar, F., Kim, T. O. & Okuyama, K. 2013. Synthesis of Spherical Macroporous WO3 Particles and Their High Photocatalytic Performance. Chemical Engineering Science, 101, 523-532.

Nandiyanto, A. B. D., Hagura, N., Iskandar, F. & Okuyama, K. 2010. Design of a Highly Ordered and Uniform Porous Structure with Multisized Pores in Film and Particle Forms Using a Template-Driven Self-Assembly Technique. Acta Materialia, 58 (1), 282-289.

Nickerson, T. A. & Moore, E. E. 1974. Factors Influencing Lactose Crystallization. Journal of Dairy Science, 57 (11), 1315-1319.

Nichols, G. & Frampton, C. S. 1998. Physicochemical Characterization of the Orthorhombic Polymorph of Paracetamol Crystallized from Solution. Journal of Pharmaceutical Sciences, 87 (6), 684-693.

Niemelä, P., Päällysaho, M., Harjunen, P., Koivisto, M., Lehto, V.-P., Suhonen, J. & Järvinen, K. 2005. Quantitative analysis of amorphous content of lactose using CCD-Raman spectroscopy. Journal of Pharmaceutical and Biomedical Analysis, 37 (5), 907-911.

Nijdam, J., Ibach, A. & Kind, M. 2008. Fluidisation of Whey Powders above the Glass-Transition Temperature. Powder Technology, 187 (1), 53-61.

Ohrem, H. L., Schornick, E., Kalivoda, A. & Ognibene, R. 2014. Why Is Mannitol Becoming More and More Popular as a Pharmaceutical Excipient in Solid Dosage Forms? Pharmaceutical Development and Technology, 19 (3), 257-262.

Page 183: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

165 | P a g e

Ojha, N. & Prabhakar, B. 2013. Advances in Solubility Enhancement Techniques. International Journal of Pharmaceutical Sciences Review and Research, 21 (2), 351-358.

Otsuka, M., Kato, F., Matsuda, Y. & Ozaki, Y. 2003. Comparative Determination of Polymorphs of Indomethacin in Powders and Tablets by Chemometrical Near-Infrared Spectroscopy and X-Ray Powder Diffractometry. AAPS PharmSciTech, 4 (Copyright (C) 2015 American Chemical Society (ACS). All Rights Reserved.), 147-158.

Oveisi, H., Suzuki, N., Beitollahi, A. & Yamauchi, Y. 2010. Aerosol-Assisted Fabrication of Mesoporous Titania Spheres with Crystallized Anatase Structures and Investigation of Their Photocatalitic Properties. Journal of Sol-Gel Science and Technology, 56 (2), 212-218.

Padrela, L., de Azevedo, E. G. & Velaga, S. P. 2012. Powder X-Ray Diffraction Method for the Quantification of Cocrystals in the Crystallization Mixture. Drug Development and Industrial Pharmacy, 38 (8), 923-929.

Panesar, P. S., Kennedy, J. F., Gandhi, D. N. & Bunko, K. 2007. Bioutilisation of Whey for Lactic Acid Production. Food Chemistry, 105 (1), 1-14.

Parikh, D. M. 2011. Recent Advances in Particle Engineering for Pharmaceutical Applications. American Pharmaceutical Review, 14 (2), 98-103.

Paterson, A. H. J., Brooks, G. F., Bronlund, J. E. & Foster, K. D. 2005. Development of Stickiness in Amorphous Lactose at Constant T−Tg Levels. International Dairy Journal, 15 (5), 513-519.

Peleg, M. 1992. On the Use of the WLF Model in Polymers and Foods. Critical Reviews in Food Science and Nutrition, 32 (1), 59-66.

Prestidge, C. A., Barnes, T. J., Lau, C.-H., Barnett, C., Loni, A. & Canham, L. 2007. Mesoporous Silicon: A Platform for the Delivery of Therapeutics. Expert Opinion on Drug Delivery, 4 (2), 101-110.

Qi, S., Avalle, P., Saklatvala, R. & Craig, D. Q. M. 2008. An Investigation into the Effects of Thermal History on the Crystallisation Behaviour of Amorphous Paracetamol. European Journal of Pharmaceutics and Biopharmaceutics, 69 (1), 364-371.

Qu, F., Zhu, G., Lin, H., Zhang, W., Sun, J., Li, S. & Qiu, S. 2006. A Controlled Release of Ibuprofen by Systematically Tailoring the Morphology of Mesoporous Silica Materials. Journal of Solid State Chemistry, 179 (7), 2027-2035.

Reddy, R. S., Reddy, J. S., Kumar, R. & Kumar, P. 1992. Sulfoxidation of Thioethers using Titanium Silicate Molecular Sieve Catalysts. Journal of the Chemical Society, Chemical Communications, 10.1039/c39920000084(2), 84-85.

Rengarajan, G. T., Enke, D., Steinhart, M. & Beiner, M. 2008. Stabilization of the Amorphous State of Pharmaceuticals in Nanopores. Journal of Materials Chemistry, 18, 2537-2539.

Rhodes, M., 2007. Introduction to Particle Technology (2nd Edition), John Wiley & Sons, Ltd, 169-s191.

Riepma, K. A., Lerk, C. F., de Boer, A. H., Bolhuis, G. K. & Kussendrager, K. D. 1990. Consolidation and Compaction of Powder Mixtures. I. Binary Mixtures of Same Particle Size Fractions of Different Types of Crystalline Lactose. International Journal of Pharmaceutics, 66 (1), 47-52.

Roos, Y. 1993. Melting and Glass Transitions of Low Molecular Weight Carbohydrates. Carbohydrate Research, 238 (C1-C10), 39-48.

Roos, Y. & Karel, M. 1991. Plasticizing Effect of Water on Thermal Behavior and Crystallization of Amorphous Food Models. Journal of Food Science, 56 (1), 38-43.

Page 184: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

166 | P a g e

Roos, Y. & Karel, M. 1992. Crystallization of Amorphous Lactose. Journal of Food Science, 57 (3), 775-777.

Roos, Y. H., 1995. Phase Transitions in Foods, Elsevier Science. Rowe, R. C., Sheskey, P. J., Quinn, M. E. & American Pharmacists, A., 2009. Handbook of

pharmaceutical excipients, London; Washington, DC; Greyslake, IL, Pharmaceutical Press.

Rowe, R. C., Sheskey, P. J. & Weller, P. J., 2003. Handbook of pharmaceutical excipients, Washington, DC; London, Pharmaceutical Press.

Sachse, A., Hulea, V., Kostov, K. L., Marcotte, N., Boltoeva, M. Y., Belamie, E. & Alonso, B. 2012. Efficient Mesoporous Silica-Titania Catalysts from Colloidal Self-Assembly. Chemical Communications 48 (86), 10648-10650.

Saffari, M., Ebrahimi, A. & Langrish, T. 2015a. Highly-Porous Mannitol Particle Production Using a New Templating Approach. Food Research International, 67 (0), 44-51.

Saffari, M., Ebrahimi, A. & Langrish, T. 2015b. The Role of Acidity in Crystallization of Lactose and Templating Approach for Highly-Porous Lactose Production. Journal of Food Engineering, 164 (C), 1-9.

Saffari, M. & Langrish, T. 2014. Effect of Lactic Acid In-Process Crystallization of Lactose/Protein Powders During Spray Drying. Journal of Food Engineering, 137, 88-94.

Sahoo, P. K. 2012. Tablets. Pharmacy Pharmaceutical Technology NISCAIR. Salameh, A. K. & Taylor, L. S. 2006. Role of Deliquescence Lowering in Enhancing Chemical

Reactivity in Physical Mixtures. The Journal of Physical Chemistry. B, 110 (20), 10190-10196.

Salonen, J., Laitinen, L., Kaukonen, A. M., Tuura, J., Björkqvist, M., Heikkilä, T., Vähä-Heikkilä, K., Hirvonen, J. & Lehto, V. P. 2005a. Mesoporous Silicon Microparticles for Oral Drug Delivery: Loading and Release of Five Model Drugs. Journal of Controlled Release, 108 (2), 362-374.

Salonen, J., Laitinen, L., Kaukonen, A. M., Tuura, J., Björkqvist, M., Heikkilä, T., Vähä-Heikkilä, K., Hirvonen, J. & Lehto, V. P. 2005b. Mesoporous Silicon Microparticles for Oral Drug Delivery: Loading and Release of Five Model Drugs. Journal of Controlled Release, 108 (2-3), 362-374.

Savjani, K. T., Gajjar, A. K. & Savjani, J. K. 2012. Drug Solubility: Importance and Enhancement Techniques. ISRN Pharmaceutics, 2012, 195727.

Scherrer, P. 1918. Bestimmung der Grösse und der inneren Struktur von Kolloidteilchen mittels Röntgenstrahlen. Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen, Mathematisch-Physikalische Klasse, 26, 98-100.

Schreiner, T., Schaefer, U. F. & Loth, H. 2005. Immediate Drug Release from Solid Oral Dosage Forms. Journal of Pharmaceutical Sciences, 94 (1), 120-133.

Simões, S., Sousa, A. & Figueiredo, M. 1996. Dissolution Rate Studies of Pharmaceutical Multisized Powders — A Practical Approach Using the Coulter Method. International Journal of Pharmaceutics, 127 (2), 283-291.

Simpson, T. D., Parrish, F. W. & Nelson, M. L. 1982. Crystalline Forms of Lactose Produced in Acidic Alcoholic Media. Journal of Food Science, 47 (6), 1948-51, 1954.

Singh, R. K., Nielsen, S. S. & Chambers, J. V. 1991. Lactose Crystallization from Whey Permeate in an Ethanol-Water Mixture. Journal of Food Science, 56 (4), 935-937.

Page 185: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

167 | P a g e

Sirisuth, N., Augsburger, L. L. & Eddington, N. D. 2002. Development and Validation of a Non-Linear IVIVC Model for a Diltiazem Extended Release Formulation. Biopharmaceutics & Drug Disposition, 23 (1), 1-8.

Slade, L. & Levine, H. 1995. Glass Transitions and Water-Food Structure Interactions. Advances in food and nutrition research, 38 (2), 103-269.

Socrates, G., 2001. Infrared and Raman Characteristic Group Frequencies: Tables and Charts, Chichester, Wiley, pp. 331-332.

Sollohub, K. & Cal, K. 2010. Spray Drying Technique: II. Current Applications in Pharmaceutical Technology. Journal of Pharmaceutical Sciences, 99 (2), 587-597.

Sperling, L. H., 1986. Introduction to Physical Polymer Science, New York, Wiley. Staniforth, J. N., Rees, J. E., Kayes, J. B., Priest, R. C. & Cotterill, N. J. 1981. The Design of a

Direct Compression Tableting Excipient. Drug Development and Industrial Pharmacy, 7 (2), 179-190.

Steckel, H. & Bolzen, N. 2004. Alternative Sugars as Potential Carriers for Dry Powder Inhalations. International Journal of Pharmaceutics, 270 (1-2), 297-306.

Petralito., S., Zanardi, I., Memoli, A., Annesini, C., Milucci, V., & Travagli, V. 2012. Apparent Solubility and Dissolution Profile at Non-Sink Conditions as Quality Improvement Tools. In: Basnet, P. (ed.) Promising Pharmaceuticals. InTech, 84, Ch. 5, 83-100. Rijeka (Croatia)].

Suihko, E., Lehto, V. P., Ketolainen, J., Laine, E. & Paronen, P. 2001. Dynamic Solid-State and Tableting Properties of Four Theophylline Forms. International Journal of Pharmaceutics, 217 (1-2), 225-236.

Tang, L., Khan, S. U. & Muhammad, N. A. 2001. Evaluation and Selection of Bio-relevant Dissolution Media for a Poorly Water-Soluble New Chemical Entity. Pharmaceutical development and technology, 6 (4), 531-540.

Telang, C., Suryanarayanan, R. & Yu, L. 2003. Crystallization of D-Mannitol in Binary Mixtures with NaCl: Phase Diagram and Polymorphism. Pharmaceutical Research, 20 (12), 1939-1945.

Twieg, W. C. & Nickerson, T. A. 1968. Kinetics of Lactose Crystallization. Journal of Dairy Science, 51 (11), 1720-1724.

Van Speybroeck, M., Barillaro, V., Thi, T. D., Mellaerts, R., Martens, J., Van Humbeeck, J., Vermant, J., Annaert, P., Van Den Mooter, G. & Augustijns, P. 2009. Ordered Mesoporous Silica Material SBA-15: A Broad-Spectrum Formulation Platform for Poorly Soluble Drugs. Journal of Pharmaceutical Sciences, 98 (8), 2648-2658.

Variankaval, N., Cote, A. S. & Doherty, M. F. 2008. From Form to Function: Crystallization of Active Pharmaceutical Ingredients. AIChE Journal, 54 (7), 1682-1688.

Verraedt, E., Pendela, M., Adams, E., Hoogmartens, J. & Martens, J. A. 2010. Controlled Release of Chlorhexidine from Amorphous Microporous Silica. Journal of Controlled Release, 142, 47-52.

Vikas Anand Saharan, V. K., Mahesh Kataria, Vandana Kharb, Pratim Kumar Choudhury 2008. Ordered mixing: mechanism, process and applications in pharmaceutical formulations. AJPS, 3 (6), 240-259.

Vippagunta, S. R., Maul, K. A., Tallavajhala, S. & Grant, D. J. W. 2002. Solid-State Characterization of Nifedipine Solid Dispersions. International Journal of Pharmaceutics, 236 (1–2), 111-123.

Page 186: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

168 | P a g e

Vromans, H., Bolhuis, G. K., Lerk, C. F. & Kussendrager, K. D. 1987a. Studies of Tableting Properties of Lactose. IX. The Relationship Between Particle Structure and Compactibility of Crystalline Lactose. International Journal of Pharmaceutics, 39 (3), 207-212.

Vromans, H., Bolhuis, G. K., Lerk, C. F., van de Biggelaar, H. & Bosch, H. 1987b. Studies of Tableting Properties of Lactose. VII. The Effect of Variations in Primary Particle Size and Percentage of Amorphous Lactose in Spray Dried Lactose Products. International Journal of Pharmaceutics, 35 (1–2), 29-37.

Vuataz, G. 2002. The Phase Diagram of Milk: A New Tool for Optimising the Drying Process. Lait, 82 (4), 485-500.

Wallack, D. A. & King, C. J. 1988. Sticking and Agglomeration of Hygroscopic, Amorphous Carbohydrate and Food Powders. Biotechnology Progress, 4 (1), 31-5.

Wang, S. & Langrish, T. A. G. 2009. A distributed parameter model for particles in the spray drying process. Advanced Powder Technology, 20 (3), 220-226.

Wang, S. & Langrish, T. A. G. 2010. The Use of Surface Active Compounds as Additives in Spray Drying. Drying Technology, 28 (3), 341-348.

Westermarck, S., Juppo, A. M., Kervinen, L. & Yliruusi, J. 1998. Pore Structure and Surface Area of Mannitol Powder, Granules and Tablets Determined with Mercury Porosimetry and Nitrogen Adsorption. European Journal of Pharmaceutics and Biopharmaceutics, 46 (1), 61-68.

Westesen, K. & Bunjes, H. 1995. Do Nanoparticles Prepared from Lipids Solid at Room Temperature Always Possess a Solid Lipid Matrix? International Journal of Pharmaceutics, 115 (1), 129-131.

Williams, M. L., Landel, R. F. & Ferry, J. D. 1955. The Temperature Dependence of Relaxation Mechanisms in Amorphous Polymers and Other Glass-Forming Liquids. Journal of the American Chemical Society, 77 (14), 3701-7.

Wong, L. W. & Pilpel, N. 1990. Effect of Particle Shape on the Mixing of Powders. Journal of Pharmacy and Pharmacology, 42 (1), 1-6.

Xia, W. & Chang, J. 2006. Well-Ordered Mesoporous Bioactive Glasses (MBG): A Promising Bioactive Drug Delivery System. Journal of Controlled Release, 110 (3), 522-530.

Xue, Y., Zhao, Q. & Luan, C. 2001. The Thermodynamic Relations between the Melting Point and the Size of Crystals. Journal of Colloid And Interface Science, 243 (2), 388-390.

Yamasaki, K., Kwok, P. C. L., Fukushige, K., Prud'homme, R. K. & Chan, H.-K. 2011. Enhanced Dissolution of Inhalable Cyclosporine Nano-Matrix Particles with Mannitol as Matrix Former. International Journal of Pharmaceutics, 420 (1), 34-42.

Yazdanpanah, N. & Langrish, T. A. G. 2011a. Crystallization and Drying of Milk Powder in a Multiple-Stage Fluidized Bed Dryer. Drying Technology, 29 (9), 1046-1057.

Yazdanpanah, N. & Langrish, T. A. G. 2011b. Egg-Shell Like Structure in Dried Milk Powders. Food Research International, 44 (1), 39-45.

Yazdanpanah, N. & Langrish, T. A. G. 2011c. Fast Crystallization of Lactose and Milk Powder in Fluidized Bed Dryer/Crystallizer. Dairy Science & Technology, 91 (3), 323-340.

Yazdanpanah, N. & Langrish, T. A. G. 2013a. Comparative Study of Deteriorative Changes in the Ageing of Milk Powder. Journal of Food Engineering, 114 (1), 14-21.

Yazdanpanah, N. & Langrish, T. A. G. 2013b. Comparative Study of Eeteriorative Changes in the Ageing of Milk Powder. Journal of Food Engineering, 114 (1), 14-21.

Page 187: Copyright and use of this thesis · iii | Page lactose) to 21.5 ± 0.5 m2 g-1.The typical specific surface areas of mannitol powders have been reported to be 0.3 m2 g-1 (Babu and

169 | P a g e

York, P. 1983. Solid-State Properties of Powders in the Formulation and Processing of Solid Dosage forms. International Journal of Pharmaceutics, 14 (1), 1-28.

Yurchenko, G. R., Matkovskii, A. K., Mel’nik, I. V., Dudarko, O. A., Stolyarchuk, N. V., Zub, Y. L. & Alonso, B. 2012. Adsorption Properties of Silica-Based Sorbents Containing Phosphonic Acid Residues. Colloid Journal, 74 (3), 386-390.

Zhang, F. Z., Xie, Y. R., Xu, S. L., Zhao, X. F. & Lei, X. D. 2010. Facile Fabrication and Magnetic Properties of Macroporous Spinel Microspheres from Layered Double Hydroxide Microsphere Precursor. Chemistry Letters, 39 (6), 588-590.

Zheng, J., 2009. Formulation and Analytical Development for Low-Dose Oral Drug Products, John Wiley & Sons, Inc.

Zhou, L., Erickson, R. R. & Holtzman, J. L. 1997. Studies Comparing the Kinetics of Cysteine Conjugation and Protein Binding of Acetaminophen by Hepatic Microsomes from Male Mice. Biochimica et Biophysica Acta (BBA) - General Subjects, 1335 (1–2), 153-160.

Zuurman, K., Riepma, K. A., Bolhuis, G. K., Vromans, H. & Lerk, C. F. 1994. The Relationship Between Bulk Density and Compactibility of Lactose Granulations. International Journal of Pharmaceutics, 102 (1), 1-9.