10.01.2015 Views

determination of droplet surface temperature and drying kinetics of ...

determination of droplet surface temperature and drying kinetics of ...

determination of droplet surface temperature and drying kinetics of ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

DETERMINATION OF DROPLET SURFACE<br />

TEMPERATURE AND DRYING KINETICS<br />

OF PROTEIN SOLUTIONS<br />

USING AN ULTRASONIC LEVITATOR<br />

Der Naturwissenschaftlichen Fakultät<br />

der Friedrich-Alex<strong>and</strong>er-Universität Erlangen-Nürnberg<br />

zur<br />

Erlangung des Doktorgrades<br />

vorgelegt von<br />

Eva Cornelia Wulsten<br />

aus Delmenhorst


Als Dissertation genehmigt von der Naturwissenschaftlichen Fakultät<br />

der Universität Erlangen-Nürnberg<br />

Tag der mündlichen Prüfung:<br />

Vorsitzender der Promotionskommission:<br />

Erstberichterstatter:<br />

Zweitberichterstatter:<br />

Pr<strong>of</strong>. Dr. Eberhard Bänsch<br />

Pr<strong>of</strong>. Dr. Ge<strong>of</strong>frey Lee<br />

Pr<strong>of</strong>. Dr. Achim Göpferich


Für meine Eltern<br />

und Martin


Danksagung<br />

Die vorliegende Arbeit wurde von Oktober 2005 bis Februar 2009 am Lehrstuhl für<br />

Pharmazeutische Technologie der Friedrich-Alex<strong>and</strong>er-Universität Erlangen-Nürnberg<br />

angefertigt.<br />

Zuerst möchte ich Pr<strong>of</strong>. Dr. Ge<strong>of</strong>frey Lee für die Möglichkeit danken, in seinem<br />

Arbeitskreis an einem sehr interessanten und innovativen Thema arbeiten zu dürfen. Für<br />

seine Unterstützung und Diskussionsbereitschaft bei wissenschaftlichen Fragestellungen<br />

danke ich ihm genauso, wie für den Freiraum für eigene Ideen und selbständiges Arbeiten<br />

und das sehr angenehme Arbeitsklima am Lehrstuhl.<br />

Ferner danke ich der Deutschen Forschungsgemeinschaft für die finanzielle<br />

Unterstützung der Projekte (Le 626/8-1 und Le 626/8-2).<br />

Pr<strong>of</strong>. Dr. Achim Göpferich danke ich für die Übernahme des Zweitgutachtens.<br />

Ganz besonders danke ich für alle Unterstützung, die ich am Lehrstuhl für<br />

Pharmazeutische Technologie für meine Arbeit bekommen habe:<br />

• Dr. Stefan Seyferth für seine geduldige Hilfe bei Computerproblemen<br />

• Josef Hubert für seinen großartigen Einsatz beim Umbau des Levitators, außerdem<br />

für seine Kreativität und Hilfe bei technischen Fragestellungen<br />

• Luise Schedl für die rasche und geduldige Anfertigung unzähliger SEM-<br />

Aufnahmen der Partikel und Pulver<br />

• Christiane Blaha für die schnelle Beschaffung aller Substanzen und Materialien<br />

• Petra Neubarth für ihre Hilfe bei allen Verwaltungsangelegenheiten und die<br />

angenehme Zusammenarbeit bei der Organisation des Studentenpraktikums<br />

• Dr. Heiko Schiffter für die umfangreiche Einführung in mein Forschungsthema und<br />

seine Bereitschaft, fortwährend für meine Fragen zur Verfügung zu stehen<br />

• Dr. Alex<strong>and</strong>er Mauerer für die freundliche Aufnahme zu Beginn meiner<br />

Doktor<strong>and</strong>enzeit in seinem Labor und die umfangreiche Hilfe bei S<strong>of</strong>twarefragen<br />

• Dr. Joanna Manegold für die freundliche Aufnahme, die mir den Einstieg am<br />

Lehrstuhl leicht gemacht hat, und den fortwährenden freundschaftlichen Kontakt


• Dr. Andreas Ziegler und Dr. Peter Lassner für die Einführung in das Praktikum<br />

Liquida / Dermatika und für viele gute Gespräche im „Labor Harmonie“<br />

• Harald Pudritz, Anke Saß und Georg Straller für die gute Zusammenarbeit bei der<br />

Studentenbetreuung, für ihre Unterstützung in verschiedenen Bereichen meiner<br />

Arbeit und außerdem für die vielen schönen Abende in Erlangen, Nürnberg und<br />

Umgebung und bei den gemeinsam besuchten Weiterbildungsseminaren<br />

• den Pharmaziestudentinnen Silvia Meffert und Adriane Giraud für ihren Einsatz bei<br />

meiner Arbeit mit den Proteinen<br />

Auch bei meinen Kolleginnen und Kollegen während meiner Doktor<strong>and</strong>enzeit Jakob<br />

Beirowski, Dr. Jürgen Bögelein, Dr. Henning Gieseler, Silja von Graberg, Eva Meister,<br />

Simone Reismann, Susanne Rutzinger, Stefan Schneid, Sebastian Vonh<strong>of</strong>f und Dr.<br />

Henning Wegner möchte ich mich für ihre Unterstützung meiner Arbeit und das<br />

angenehme Arbeitsklima am Lehrstuhl bedanken.<br />

Allen meinen Freunden danke ich sehr für ihre Begleitung. Ganz besonderer Dank<br />

an Dr. Kerstin und Dirk Schöttler für das Korrekturlesen dieser Arbeit.<br />

Meinen Eltern gilt mein besonderer Dank für ihre liebevolle Unterstützung. Ohne<br />

ihre persönliche sowie finanzielle Unterstützung wäre mir dieser Weg nicht möglich<br />

gewesen. Meinem Bruder Jörg und Christine danke ich für die Unterstützung während<br />

meines Studiums und in meiner Doktor<strong>and</strong>enzeit - nicht nur bei diversen Umzügen - und<br />

für das Korrekturlesen dieser Arbeit.<br />

Ganz besonders danke ich meinem Mann Martin, der mir mit seiner Liebe und<br />

Unterstützung während der Doktor<strong>and</strong>enzeit eine sehr große Hilfe war. Angefangen bei<br />

seiner Bereitschaft, mit mir nach Nürnberg zu gehen, hat er mich mit sehr viel Motivation<br />

unterstützt. Für sein Interesse an meinen Versuchen, seine Hilfe bei S<strong>of</strong>twareproblemen<br />

und für das Korrekturlesen dieser Arbeit bin ich ihm sehr dankbar.


Parts <strong>of</strong> this thesis have already been presented or published:<br />

I. E. Wulsten <strong>and</strong> G. Lee (2008): “Drying behavior <strong>of</strong> levitated trehalose<br />

micro<strong>droplet</strong>s: <strong>surface</strong> <strong>temperature</strong> measurements <strong>and</strong> <strong>drying</strong> <strong>kinetics</strong>” 6 th World<br />

Meeting on Pharmaceutics, Biopharmaceutics <strong>and</strong> Pharmaceutical Technology in<br />

Barcelona (Spain)<br />

II.<br />

E. Wulsten <strong>and</strong> G. Lee (2008): “Surface <strong>temperature</strong> <strong>of</strong> acoustically levitated water<br />

micro<strong>droplet</strong>s measured using infra-red thermography” Chemical Engineering<br />

Science 2008, 63 (22), 5420-5424<br />

III.<br />

E. Wulsten <strong>and</strong> G. Lee (2008): “Residual activity <strong>of</strong> carbonic anhydrase depending<br />

on <strong>drying</strong> time using an ultrasonic levitator” DPhG Annual Meeting in Bonn<br />

(Germany)<br />

VI.<br />

E. Wulsten, F. Kiekens, F. van Dycke, J. Voorspoels <strong>and</strong> G. Lee: “Applications <strong>of</strong><br />

levitated single <strong>droplet</strong> <strong>drying</strong> in process <strong>and</strong> formulation development. Case study<br />

with itraconazole dried in binary organic solvent mixtures” International Journal <strong>of</strong><br />

Pharmaceutics 2009, in press


Contents I<br />

Table <strong>of</strong> contents<br />

1 Introduction ................................................................................................................ 1<br />

2 Single <strong>droplet</strong> <strong>drying</strong> .................................................................................................. 4<br />

2.1 A short history <strong>of</strong> single <strong>droplet</strong> <strong>drying</strong> ................................................................. 4<br />

2.2 Heat transfer ........................................................................................................... 6<br />

2.2.1 Conductive heat transfer .................................................................................. 6<br />

2.2.2 Convective heat transfer ................................................................................... 6<br />

2.2.3 Radiative heat transfer ...................................................................................... 7<br />

2.3 Diffusion ............................................................................................................... 11<br />

2.3.1 Fick’s first law <strong>of</strong> diffusion ............................................................................ 12<br />

2.3.2 Diffusion coefficients for gases ..................................................................... 12<br />

2.3.3 Convective mass transfer ............................................................................... 13<br />

2.4 Mass transfer theories ........................................................................................... 13<br />

2.5 Evaporation <strong>of</strong> pure solvent <strong>droplet</strong>s.................................................................... 14<br />

2.5.1 Evaporation <strong>of</strong> a single <strong>droplet</strong> ...................................................................... 14<br />

2.5.2 The d²-law ...................................................................................................... 15<br />

2.5.3 Diffusion-controlled evaporation <strong>of</strong> single <strong>droplet</strong>s ...................................... 16<br />

2.5.4 Evaporation <strong>of</strong> <strong>droplet</strong>s containing solvent mixtures ..................................... 16<br />

2.6 Evaporation <strong>of</strong> solution <strong>and</strong> suspension <strong>droplet</strong>s ................................................. 17<br />

2.6.1 Drying stages <strong>of</strong> solution <strong>and</strong> suspension <strong>droplet</strong>s ........................................ 17<br />

2.6.2 Particle formation ........................................................................................... 20<br />

2.6.3 Models for the <strong>drying</strong> <strong>of</strong> single <strong>droplet</strong>s containing solids ............................ 22<br />

3 Acoustic levitation .................................................................................................... 24<br />

3.1 History <strong>and</strong> application <strong>of</strong> acoustic levitation ...................................................... 24<br />

3.2 A short introduction to acoustics .......................................................................... 25<br />

3.3 Levitation forces ................................................................................................... 27<br />

3.3.1 St<strong>and</strong>ing acoustic wave .................................................................................. 27<br />

3.3.2 Interaction <strong>of</strong> ultrasonic wave <strong>and</strong> <strong>droplet</strong> ..................................................... 28<br />

3.4 Acoustic streaming ............................................................................................... 29<br />

3.4.1 Inner acoustic streaming ................................................................................ 29


Contents II<br />

3.4.2 Outer acoustic streaming ................................................................................ 30<br />

3.5 Vertical position <strong>of</strong> the levitated <strong>droplet</strong> .............................................................. 31<br />

3.6 Temperature <strong>of</strong> the levitation system ................................................................... 32<br />

3.7 Single <strong>droplet</strong> <strong>drying</strong> using an acoustic levitator ................................................. 32<br />

3.7.1 Pure solvent <strong>droplet</strong>s ...................................................................................... 32<br />

3.7.2 Solution <strong>and</strong> suspension <strong>droplet</strong>s ................................................................... 35<br />

4 Materials <strong>and</strong> Methods ............................................................................................ 37<br />

4.1 Materials ............................................................................................................... 37<br />

4.1.1 Proteins ........................................................................................................... 37<br />

4.1.1.1 Carbonic anhydrase .................................................................................. 37<br />

4.1.1.2 L-Lactic dehydrogenase ........................................................................... 39<br />

4.1.1.3 Trypsinogen <strong>and</strong> Trypsin ......................................................................... 40<br />

4.1.2 Itraconazole .................................................................................................... 41<br />

4.1.3 Excipients <strong>and</strong> reagents .................................................................................. 42<br />

4.2 Methods ................................................................................................................ 44<br />

4.2.1 Acoustic levitation ......................................................................................... 44<br />

4.2.2 Microdispensing system ................................................................................. 48<br />

4.2.3 Droplet size measurements ............................................................................ 51<br />

4.2.4 Droplet <strong>surface</strong> <strong>temperature</strong> measurements ................................................... 51<br />

4.2.5 Spray <strong>drying</strong> ................................................................................................... 52<br />

4.2.6 Scanning electron microscopy ....................................................................... 53<br />

4.2.7 Enzyme activity assay <strong>of</strong> carbonic anhydrase ................................................ 53<br />

4.2.8 Enzyme activity assay <strong>of</strong> L-lactic dehydrogenase ......................................... 54<br />

4.2.9 Enzyme activity assay <strong>of</strong> trypsinogen ............................................................ 55<br />

5 Results <strong>and</strong> Discussion ............................................................................................. 57<br />

5.1 Preliminary experiments using the IR-camera ..................................................... 57<br />

5.1.1 Heating-up <strong>of</strong> the levitation system ............................................................... 57<br />

5.1.2 Influence <strong>of</strong> <strong>droplet</strong> size on the <strong>surface</strong> <strong>temperature</strong> measurement ............... 58<br />

5.1.3 Determination <strong>of</strong> emissivities ........................................................................ 64<br />

5.2 Evaporation <strong>of</strong> pure solvent <strong>droplet</strong>s.................................................................... 67<br />

5.2.1 Evaporation <strong>of</strong> pure water <strong>droplet</strong>s ................................................................ 67


Contents III<br />

5.2.1.1 Influences <strong>of</strong> the initial <strong>droplet</strong> size on the evaporation process ............. 67<br />

5.2.1.2 Influence <strong>of</strong> the <strong>drying</strong> air stream on the evaporation process ................ 84<br />

5.2.2 Evaporation <strong>of</strong> pure organic solvent <strong>droplet</strong>s ................................................ 96<br />

5.2.3 Evaporation <strong>of</strong> <strong>droplet</strong>s containing solvent mixtures ................................... 103<br />

5.3 Evaporation <strong>of</strong> excipient solution <strong>droplet</strong>s ......................................................... 110<br />

5.3.1 Trehalose solution <strong>droplet</strong>s .......................................................................... 110<br />

5.3.2 Mannitol solution <strong>droplet</strong>s ........................................................................... 121<br />

5.3.3 Sucrose solution <strong>droplet</strong>s ............................................................................. 133<br />

5.3.4 Copolyvidone <strong>and</strong> HPMC solution <strong>droplet</strong>s ................................................ 136<br />

5.4 Itraconazole formulation experiments ................................................................ 141<br />

5.5 Drying <strong>of</strong> protein solutions ................................................................................ 148<br />

5.5.1 Bovine carbonic anhydrase (bCA) <strong>droplet</strong>s ................................................. 148<br />

5.5.1.1 bCA-trehalose 10 % (w/v) total solids formulations ............................. 151<br />

5.5.1.2 bCA 10 % (w/v) plus trehalose formulations ......................................... 157<br />

5.5.2 L-Lactic dehydrogenase (LDH) <strong>droplet</strong>s ..................................................... 161<br />

5.5.2.1 LDH-trehalose 10 % (w/v) total solids formulations ............................. 163<br />

5.5.2.2 LDH 10 % (w/v) plus trehalose formulations ........................................ 170<br />

5.5.3 Trypsinogen <strong>droplet</strong>s .................................................................................... 174<br />

6 Summary <strong>and</strong> Conclusions .................................................................................... 177<br />

7 Zusammenfassung .................................................................................................. 182<br />

8 References ............................................................................................................... 188


Abbreviations IV<br />

List <strong>of</strong> abbreviations<br />

Capital letters<br />

<br />

area<br />

coefficient (Equation 3.16, Table 5.3)<br />

, , liquid parameters (Equation 3.15, Table 3.2)<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

pressure amplitude at the source <strong>surface</strong><br />

effective pressure amplitude <strong>of</strong> the incident acoustic field<br />

Spalding transfer number<br />

binary diffusion coefficient<br />

binary diffusion coefficient for water<br />

horizontal diameter<br />

vertical diameter<br />

radial acoustic levitation force<br />

axial acoustic levitation force<br />

relative humidity<br />

latent heat <strong>of</strong> evaporation<br />

latent heat <strong>of</strong> evaporation for water<br />

molar flux <strong>of</strong> component A<br />

bulk modulus elasticity<br />

constant <strong>of</strong> the acoustic field<br />

sound pressure level<br />

distance <strong>of</strong> the reflector from the transducer<br />

molar mass<br />

molar mass <strong>of</strong> a gas<br />

molar mass <strong>of</strong> a liquid<br />

molar mass <strong>of</strong> water<br />

heat energy<br />

heat flow<br />

<br />

, <br />

heat flux in the liquid / surrounding gas (Figure 2.4)<br />

<br />

ideal gas constant


Abbreviations V<br />

Reynolds number<br />

<br />

<strong>surface</strong> area<br />

Schmidt number<br />

Sherwood number<br />

<br />

<strong>temperature</strong><br />

experimentally-measured <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong><br />

<strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> calculated according to Yarin et al. [1999]<br />

wet-bulb <strong>temperature</strong><br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

boiling point<br />

critical <strong>temperature</strong><br />

<strong>temperature</strong> <strong>of</strong> the <strong>drying</strong> air<br />

<strong>temperature</strong> <strong>of</strong> the bulk fluid<br />

<strong>surface</strong> <strong>temperature</strong><br />

<strong>temperature</strong> at infinity<br />

volume<br />

emittance<br />

emittance <strong>of</strong> a black body<br />

emittance <strong>of</strong> a black body over the wavelength<br />

mass fraction <strong>of</strong> the vapour<br />

mass fraction <strong>of</strong> the vapour at infinity<br />

characteristic acoustic impedance<br />

Small letters<br />

constants (Equation 3.12, Table 3.1)<br />

<br />

concentration<br />

, constants (Equation 2.5)<br />

<br />

<br />

<br />

<br />

<br />

<br />

<strong>droplet</strong> diameter (calculated from a <strong>surface</strong> equivalent sphere)<br />

frequency<br />

gravitational acceleration<br />

heat transfer coefficient<br />

thermal conductivity<br />

thermal conductivity <strong>of</strong> air


Abbreviations VI<br />

<br />

<br />

wave number<br />

mass flux <strong>of</strong> liquid<br />

mass flux <strong>of</strong> the <strong>droplet</strong> material (Figure 2.4)<br />

coefficient (Equation 3.16, Table 5.3)<br />

<br />

pressure<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

0<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

∆ <br />

reference effective sound pressure value<br />

partial vapour pressure at the <strong>surface</strong><br />

saturation vapour pressure<br />

partial vapour pressure at infinity<br />

sound pressure<br />

effective sound pressure<br />

heat flux<br />

<strong>droplet</strong> radius (calculated from a <strong>surface</strong> equivalent sphere)<br />

<strong>droplet</strong> radius at point <strong>of</strong> time t (calculated from a <strong>surface</strong> equivalent sphere)<br />

initial <strong>droplet</strong> radius (calculated from a <strong>surface</strong> equivalent sphere)<br />

measured horizontal <strong>droplet</strong> radius<br />

measured vertical <strong>droplet</strong> radius<br />

time<br />

air velocity<br />

particle velocity<br />

sound velocity<br />

length<br />

distance centre <strong>of</strong> <strong>droplet</strong> mass to next upper pressure node<br />

Small Greek letters<br />

<br />

absorptivity<br />

<br />

<br />

<br />

<br />

<br />

<br />

evaporation coefficient (calculated from <strong>droplet</strong> radius)<br />

adiabatic index<br />

thickness <strong>of</strong> layer / wall<br />

diffusion boundary layer<br />

acoustic boundary layer<br />

emissivity


Abbreviations VII<br />

<br />

λ<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

υ <br />

<br />

diffusivity <strong>of</strong> <strong>drying</strong> gas<br />

wavelength<br />

wavelength with max. emittance <strong>of</strong> a black body<br />

dynamic viscosity<br />

particle displacement<br />

reflectivity<br />

density<br />

density gas<br />

density liquid<br />

Stefan-Boltzmann constant<br />

transmissivity<br />

diffusion volume<br />

angular frequency<br />

Subscripts<br />

A, B components<br />

1, 2 numbering<br />

Expressions<br />

bSA bovine serum albumin<br />

bCA bovine carbonic anhydrase<br />

BAEE N α -benzoyl-L-arginine ethyl ester hydrochloride<br />

CEM controlled evaporation mixer<br />

CCD charged coupled device<br />

Diff diffusion model<br />

HF high frequency<br />

HPMC hydroxypropylmethylcellulose<br />

IR infrared<br />

LDH L-lactic dehydrogenase<br />

MCT mercury cadmium tellurium<br />

MWCO molecular weight cut <strong>of</strong>f<br />

p.a. pro analysi (for analysis)


Abbreviations VIII<br />

PCR<br />

PNPA<br />

RM<br />

SPL<br />

Treh<br />

UV<br />

VIS<br />

polymerase chain reaction<br />

p-nitrophenyl acetate<br />

Ranz-Marshall model<br />

sound pressure level<br />

trehalose<br />

ultraviolet<br />

visible


Introduction 1<br />

1 Introduction<br />

The process <strong>of</strong> <strong>drying</strong> is an important step in the production <strong>of</strong> various pharmaceutical<br />

dosage forms <strong>and</strong> has a wide influence on the quality <strong>and</strong> stability <strong>of</strong> the product. During<br />

the <strong>drying</strong> process liquids can be thermally removed from a product by evaporation. In this<br />

process heat transfer takes place via conduction, convection or radiation using various<br />

types <strong>of</strong> dryers like <strong>drying</strong> ovens, infrared tunnel dryers, fluid bed dryers or spray dryers.<br />

Spray <strong>drying</strong> has a special role in industrial <strong>drying</strong> technology, since it involves <strong>drying</strong> <strong>of</strong><br />

liquid feedstocks to powders with control over the final particle morphology <strong>and</strong> functional<br />

powder properties [Masters 2002]. As a result spray <strong>drying</strong> is an appropriate method to<br />

produce powders for various kinds <strong>of</strong> pharmaceutical applications.<br />

An increasing number <strong>of</strong> peptides <strong>and</strong> proteins is now available for the therapy <strong>of</strong><br />

diseases like cancer, immune mediated or metabolic diseases <strong>and</strong> for carrying out<br />

immunizations. One quarter <strong>of</strong> all drugs containing new active ingredients approved for<br />

Germany in 2007 were biopharmaceuticals, <strong>and</strong> there are currently more than 350<br />

biopharmaceuticals in clinical trials [BCG 2008]. Protein formulations are <strong>of</strong>ten developed<br />

for parenteral administration. Due to the fact that proteins are <strong>of</strong>ten instable in aqueous<br />

solution, it is not possible in most cases to achieve a sufficient shelf life stability. In this<br />

case the protein is stored in a dry form <strong>and</strong> dissolved directly before use. Spray <strong>drying</strong> is a<br />

potential alternative to lyophilisation for the production <strong>of</strong> parenteral powders for this use<br />

[Mumenthaler et al. 1994]. Protein loaded powders are furthermore used for needle-free<br />

powder injection. Vaccines for example can be efficiently delivered to the epidermis for<br />

epidermal powder immunization [Dean <strong>and</strong> Chen 2004]. Inhalation is an appropriate way<br />

to apply a protein not only for local treatment <strong>of</strong> respiratory diseases. An approved protein<br />

formulation for pulmonary delivery as an aerosol with local effect is recombinant human<br />

DNase (Pulmozyme ® ) for treatment <strong>of</strong> mucoviscidosis [Roche 2009]. For systemic<br />

delivery <strong>of</strong> a protein by absorption via the lung inhalation is also useful as it is an<br />

attractive, pain-free <strong>and</strong> self-administrable delivery method promoting a good compliance<br />

by the patient. The lung, where some 300 million alveoli constitute a capillarized area <strong>of</strong><br />

around 100 m² [Almer et al. 2002], <strong>of</strong>fers an enormous <strong>surface</strong> area with good blood<br />

supply <strong>and</strong> a thin absorption mucosal membrane in the distal lung <strong>of</strong> 0.1 - 0.2 μm [Yu <strong>and</strong><br />

Chien 1997] for high absorption <strong>of</strong> the protein. Spray <strong>drying</strong> is an efficient method to


Introduction 2<br />

produce powders loaded with various peptides <strong>and</strong> proteins for pulmonary delivery<br />

containing particles in the size <strong>of</strong> 0.5 - 5 μm for optimal deep deposition in the respiratory<br />

tract.<br />

Spray <strong>drying</strong> <strong>of</strong>fers various possibilities to influence the particle size <strong>and</strong> flow<br />

properties via optimization <strong>of</strong> the <strong>drying</strong> conditions or the formulation ingredients. In<br />

formulation development for a protein the risk <strong>of</strong> inactivation is <strong>of</strong>ten the main problem.<br />

Chemical <strong>and</strong> physical instability <strong>of</strong> the protein can occur, the latter refers to a change in<br />

the secondary, tertiary, or quaternary structure <strong>of</strong> a protein [Banga 1995]. Process stress<br />

like thermal stress, shear stress associated with atomization <strong>and</strong> especially air-water<br />

interfacial stress, has influence on protein stability during spray <strong>drying</strong> [Maa <strong>and</strong><br />

Prestrelski 2000]. Unfolding <strong>of</strong> the protein leads to adsorption, aggregation <strong>and</strong><br />

precipitation [Manning et al. 1989] <strong>and</strong> the activity <strong>of</strong> the protein decreases. This makes<br />

the addition <strong>of</strong> stabilizing excipients like non-reducing sugars, amino acids or surfactants<br />

in the spray <strong>drying</strong> formulation necessary [Adler <strong>and</strong> Lee 1999; Andya et al. 1999; Maury<br />

et al. 2005b].<br />

Single <strong>droplet</strong> <strong>drying</strong> experiments provide a possibility to analyze formulation<br />

aspects with a very small amount <strong>of</strong> substance <strong>and</strong> are therefore very cost-efficient. It is<br />

possible to investigate <strong>drying</strong> <strong>kinetics</strong> <strong>and</strong> particle formation during the whole <strong>drying</strong><br />

process <strong>of</strong> a sample <strong>droplet</strong> at different ambient conditions. During or after <strong>drying</strong> the<br />

sample can be removed from the acoustic field for further analysis. In an ultrasonic<br />

levitator a liquid or solid sample with a diameter <strong>of</strong> about 50 - 4000 μm [Kastner 2001] can<br />

be levitated in the pressure node <strong>of</strong> a st<strong>and</strong>ing acoustic wave without heat transfer by a<br />

thermal conducting holding device.<br />

The aim <strong>of</strong> this thesis is the analysis <strong>of</strong> the <strong>drying</strong> <strong>kinetics</strong>, <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> <strong>and</strong> particle formation <strong>of</strong> acoustically-levitated sample <strong>droplet</strong>s containing<br />

proteins to determine the <strong>drying</strong> behaviour <strong>of</strong> the solution. The influence <strong>of</strong> the <strong>drying</strong><br />

process on the proteins is analyzed by enzymatic activity measurement <strong>of</strong> the protein<br />

particles.<br />

In the first part <strong>of</strong> this work, an infrared camera is established in an existing<br />

levitation setup for simultaneous <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>and</strong> size measurements. Pure<br />

solvents <strong>and</strong> solvent mixtures are analyzed at different <strong>temperature</strong> <strong>and</strong> relative humidity<br />

conditions. For water <strong>droplet</strong>s the influence <strong>of</strong> different initial <strong>droplet</strong> diameter using a<br />

micro-dispensing system <strong>and</strong> the influence <strong>of</strong> <strong>drying</strong> air rate is investigated. The suitability


Introduction 3<br />

<strong>of</strong> <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> measurements to detect the critical point <strong>and</strong> to monitor the<br />

<strong>drying</strong> process in single <strong>droplet</strong> <strong>drying</strong> experiments is analyzed. In the second part<br />

solvent / excipient formulations at different ambient conditions <strong>and</strong> solid contents are<br />

examined. In the third part <strong>drying</strong> <strong>kinetics</strong> <strong>and</strong> <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>of</strong> protein<br />

solutions <strong>and</strong> protein / excipient formulations are examined <strong>and</strong> the enzymatic activity is<br />

measured at the end <strong>of</strong> the <strong>drying</strong> process. Results from spray <strong>drying</strong> experiments are<br />

compared to the results <strong>of</strong> the levitation experiments in terms <strong>of</strong> particle shape <strong>and</strong> protein<br />

activity. The time-dependent protein activity during the <strong>drying</strong> process is analyzed.


Single <strong>droplet</strong> <strong>drying</strong> 4<br />

2 Single <strong>droplet</strong> <strong>drying</strong><br />

2.1 A short history <strong>of</strong> single <strong>droplet</strong> <strong>drying</strong><br />

Since 1952 when Ranz <strong>and</strong> Marshall investigated the <strong>drying</strong> <strong>of</strong> a single <strong>droplet</strong> containing<br />

dissolved <strong>and</strong> suspended solids, single <strong>droplet</strong> experiments have been used in the study <strong>of</strong><br />

particle morphology. This method has been criticized as being unrealistic for spray <strong>drying</strong>,<br />

because the diameters are an order <strong>of</strong> magnitude larger than spray <strong>droplet</strong>s <strong>and</strong> the <strong>droplet</strong>s<br />

are not allowed to rotate freely while <strong>drying</strong>. Furthermore the <strong>droplet</strong> is suspended on a<br />

thermal conducting holding device, which is also a site for vapour bubble nucleation within<br />

the <strong>droplet</strong> <strong>and</strong> may cause deformation <strong>of</strong> the <strong>droplet</strong> in later <strong>drying</strong> process. The method,<br />

however, is still in use, because larger <strong>droplet</strong>s are easier to h<strong>and</strong>le <strong>and</strong> to observe [Lin <strong>and</strong><br />

Gentry 2003]. Table 2.1 gives a short review <strong>of</strong> single <strong>droplet</strong> <strong>drying</strong> experiments<br />

containing particle formation <strong>and</strong> morphology studies [Lin <strong>and</strong> Gentry 2003; Schiffter<br />

2006].<br />

Charlesworth <strong>and</strong> Marshall [1960] investigated the <strong>drying</strong> rates <strong>of</strong> inorganic salts.<br />

They analyzed the crust structure <strong>and</strong> appearance change in <strong>drying</strong> <strong>droplet</strong>s <strong>and</strong> also<br />

monitored <strong>droplet</strong> weight <strong>and</strong> <strong>temperature</strong> during <strong>drying</strong> process. Organic substances were<br />

analyzed by El-Sayed et al. [1990]. Walton <strong>and</strong> Mumford [1999] carried out several<br />

experiments with organic <strong>and</strong> inorganic substances as well as multicomponent mixtures.<br />

They identified three distinct categories <strong>of</strong> particle morphology: crystalline, skin forming<br />

<strong>and</strong> agglomerate. They found that although they are material-specific, each category is<br />

evidence <strong>of</strong> a characteristic <strong>drying</strong> behaviour dependent on feed concentration, the degree<br />

<strong>of</strong> feed aeration, <strong>and</strong> <strong>drying</strong> <strong>temperature</strong>. Lin <strong>and</strong> Gentry [2003] performed single <strong>droplet</strong><br />

<strong>drying</strong> studies with several inorganic salts <strong>and</strong> analyzed particle morphology. Low <strong>drying</strong><br />

<strong>temperature</strong> <strong>and</strong> a material with high latent heat <strong>of</strong> crystallization (for example<br />

endothermic crystallization) were found to be favourable for small, dense, <strong>and</strong> regularlyshaped<br />

particles. Materials that formed an elastic shell structure led to hollow particle<br />

formation, whereas materials with high solubility led to small, dense <strong>and</strong> irregularlyshaped<br />

particles. They found that a high initial solute concentration was favourable for the<br />

formation <strong>of</strong> large dense particles.


Single <strong>droplet</strong> <strong>drying</strong> 5<br />

Table 2.1: Selected single <strong>droplet</strong> <strong>drying</strong> studies [Lin <strong>and</strong> Gentry 2003; Schiffter<br />

2006]<br />

Investigators<br />

Suspending method<br />

Temperature<br />

Test material <strong>and</strong><br />

Year<br />

measurement<br />

initial drop size<br />

Ranz <strong>and</strong> Marshall<br />

1952<br />

Glass capillary<br />

80 μm<br />

Manganconstantan<br />

thermocouple<br />

Sodium chloride,<br />

ammonium nitrate<br />

0.6 - 11.0 mm<br />

Charlesworth <strong>and</strong><br />

Marshall<br />

1960<br />

340 μm glass<br />

filament, drop<br />

formed from a<br />

microburette<br />

Thermoelement<br />

formed from<br />

manganin <strong>and</strong><br />

constantan wired<br />

connected to a<br />

recording<br />

potentiometer<br />

Sodium sulphate,<br />

potassium sulphate,<br />

copper sulphate,<br />

ammonium nitrate<br />

calcium chloride,<br />

sodium acetate <strong>and</strong><br />

c<strong>of</strong>fee extract<br />

1.3 - 1.8 mm<br />

El-Sayed, Wallack<br />

<strong>and</strong> King<br />

1990<br />

Annealed glass with<br />

a glass bead on the<br />

tip<br />

0.12 mm Type-Ethermocouple<br />

Sucrose, maltodextrin,<br />

c<strong>of</strong>fee extract, skim<br />

milk<br />

0.38 - 1.5 mm<br />

Walton <strong>and</strong><br />

Mumford<br />

1999<br />

Rotating glass<br />

filament in a wind<br />

tunnel<br />

Thermocouple<br />

Different inorganic<br />

<strong>and</strong> organic salts,<br />

gelatine, semi-instant<br />

skimmed milk, codried<br />

egg <strong>and</strong><br />

skimmed milk<br />

powder, anionic<br />

detergents<br />

1.0 - 2.0 mm<br />

Lin <strong>and</strong> Gentry<br />

2003<br />

Capillary filament<br />

inside a drop holder<br />

Thermocouple<br />

Calcium acetate,<br />

sodium acetate,<br />

potassium carbonate,<br />

sodium chloride,<br />

ammonium chloride,<br />

lithium manganous<br />

nitrate, barium<br />

aliumino boro silicate<br />

1.2 - 1.3 mm


Single <strong>droplet</strong> <strong>drying</strong> 6<br />

2.2 Heat transfer<br />

A <strong>temperature</strong> gradient in a substance leads to heat flow from the part at higher<br />

<strong>temperature</strong> to that at lower <strong>temperature</strong>. This process takes place in solid materials as well<br />

as in liquids or gases [Eckert 1966]. Heat transfer can be divided in substance-bound heat<br />

transfer via conduction <strong>and</strong> convection, <strong>and</strong> substance-unbound heat transfer via thermal<br />

radiation.<br />

2.2.1 Conductive heat transfer<br />

Heat transfer via conduction occurs upon interaction <strong>of</strong> atoms <strong>and</strong> molecules, where<br />

energy is transferred from the more energetic to the less energetic molecules. In liquids <strong>and</strong><br />

gases the faster molecules <strong>of</strong> the warmer regions will be decelerated <strong>and</strong> the molecules <strong>of</strong><br />

the colder regions will be accelerated, which leads to <strong>temperature</strong> equalization over all<br />

parts <strong>of</strong> the liquid or gas. Heat transfer in solids that are poor conductors takes place via<br />

longitudinal oscillations <strong>of</strong> the atoms, whereas in metallic solids thermal conduction results<br />

from the motion <strong>of</strong> free electrons [Bosnjakovic 1997]. The basic equation used to express<br />

heat flow by conduction is Fourier’s law:<br />

Equation 2.1<br />

· · <br />

<br />

·<br />

with the heat energy , the thermal conductivity (specific for the material) , the heat<br />

transfer area , the <strong>temperature</strong> difference across the material , the material<br />

thickness , <strong>and</strong> the time . Considering the expressions for heat flow ⁄ <strong>and</strong> for<br />

heat flux ⁄ , Equation 2.1 becomes:<br />

Equation 2.2<br />

· <br />

<br />

·<br />

<br />

The material thickness is <strong>and</strong> the negative algebraic sign recognizes that heat flow takes<br />

place from warmer to colder regions.<br />

2.2.2 Convective heat transfer<br />

In liquids <strong>and</strong> gases thermal energy can be transported from one region to another by fluid<br />

(liquid or gas) flow. Parts <strong>of</strong> the fluid move between regions <strong>of</strong> different <strong>temperature</strong> in the


Single <strong>droplet</strong> <strong>drying</strong> 7<br />

substance <strong>and</strong> in this process they carry thermal energy. Convective heat transfer between<br />

a <strong>surface</strong> <strong>and</strong> a moving fluid at a different <strong>temperature</strong> is a combination <strong>of</strong> diffusion <strong>and</strong><br />

bulk motion <strong>of</strong> molecules. Near the <strong>surface</strong> the fluid velocity is low <strong>and</strong> diffusion<br />

dominates. Away from the <strong>surface</strong> bulk motion increases the influence <strong>and</strong> dominates<br />

[Engineering Toolbox 2009]. Convection is closely associated with fluid mechanics.<br />

Convective heat transfer takes the form <strong>of</strong> either natural (also known as free) or<br />

forced (also known as assisted) convection, depending on the forces used to create<br />

convection currents. In natural convection parts <strong>of</strong> the fluid heat up at the warmer <strong>surface</strong><br />

<strong>and</strong> the density <strong>of</strong> the liquid in this region decreases. The fluid will rise <strong>and</strong> be replaced by<br />

cooler fluid that will also heat <strong>and</strong> rise generating natural convection. If the currents are set<br />

in motion by the action <strong>of</strong> a mechanical device such as a pump or agitator, the flow is<br />

independent <strong>of</strong> density gradients <strong>and</strong> is called forced convection [Bosnjakovic 1997;<br />

McCabe et al. 2005].<br />

The relation describing heat transfer by convection is Newton’s law <strong>of</strong> cooling:<br />

Equation 2.3<br />

· <br />

which expresses how the heat flux is proportional to the <strong>temperature</strong> difference between<br />

the <strong>surface</strong> <strong>and</strong> the bulk fluid . The proportionality coefficient is the heat transfer<br />

coefficient , that is dependent on the type <strong>of</strong> media, its flow properties velocity or<br />

viscosity, <strong>and</strong> also other flow <strong>and</strong> <strong>temperature</strong> dependent properties [Engineering Toolbox<br />

2009]. Convective heat transfer is dependent on the intensity <strong>of</strong> bulk fluid motion, <strong>and</strong> thus<br />

fluid mechanics play a decisive role for convective heat calculations [Bosnjakovic 1997].<br />

2.2.3 Radiative heat transfer<br />

Heat transfer through radiation takes place in the form <strong>of</strong> electromagnetic waves emitted<br />

by an object <strong>and</strong> differs from conduction <strong>and</strong> convection, which both are substance-bound.<br />

The radiation emitted by a body results from the thermal agitation <strong>of</strong> its molecules <strong>and</strong> can<br />

spread even in vacuum due to its substance independence. In contrast to conductive <strong>and</strong><br />

convective heat transfer, thermal radiation is able to pass regions <strong>of</strong> lower <strong>temperature</strong> for<br />

heat transfer between objects [Bosnjakovic 1997]. When the electromagnetic waves reach<br />

an object, they will be transmitted, reflected or absorbed, whereby only the absorbed<br />

energy appears quantitatively as heat.


Single <strong>droplet</strong> <strong>drying</strong> 8<br />

Electromagnetic radiation covers a wide b<strong>and</strong> <strong>of</strong> wavelengths, from short cosmic rays<br />

having wavelengths <strong>of</strong> about 10 -11 cm to longwave broadcasting waves having lengths <strong>of</strong><br />

1000 m or more. The portion <strong>of</strong> the electromagnetic spectrum that is important for heat<br />

flow by radiation is in the wavelength range between 0.1 <strong>and</strong> 100 μm. At <strong>temperature</strong>s<br />

above about 500 °C heat radiation in the visible spectrum becomes substantial [McCabe et<br />

al. 2005]. The infrared spectrum is attached to the visible spectrum <strong>and</strong> includes<br />

wavelength from 780 nm up to 1 mm. For technical <strong>temperature</strong> measurement the<br />

wavelength range from 0.8 - 20 μm is used. It can be divided (based on the response <strong>of</strong><br />

various detectors) into short-wave infrared (SWIR) from 1 - 3 μm, mid-wave infrared<br />

(MWIR) from 3 - 5 μm, <strong>and</strong> long-wave infrared (LWIR) from 8 - 12 μm.<br />

A black body is an ideal emitter, which has the maximum attainable emissive<br />

power at any given <strong>temperature</strong> <strong>and</strong> absorbs all incoming radiation. The basic relationship<br />

for black body radiation is the Stefan-Boltzmann law, which states that the emittance (rate<br />

<strong>of</strong> radiant energy emission per unit area, equal to heat flux ) <strong>of</strong> a black body <br />

is proportional to the fourth power <strong>of</strong> the absolute <strong>temperature</strong>:<br />

Equation 2.4<br />

· <br />

using the Stefan-Boltzmann constant = 5.669·10 -12 W·cm -2·K -4 [McCabe et al. 2005;<br />

Schuster <strong>and</strong> Kolobrodov 2004].<br />

The distribution <strong>of</strong> energy in the spectrum <strong>of</strong> a black body is given by Planck’s law,<br />

where the emittance <strong>of</strong> a black body over the wavelength λ ⁄ λ is:<br />

Equation 2.5<br />

λ <br />

λ ·<br />

1<br />

⁄ · 1<br />

with the constants = 3.74·10 -16 W·m 2 <strong>and</strong> = 1.44·10 -2 K·m, the wavelength <strong>of</strong> radiation<br />

<strong>and</strong> the absolute <strong>temperature</strong> . Planck’s law at <strong>temperature</strong>s <strong>of</strong> 300, 800, 2500 <strong>and</strong><br />

5500 K is shown graphically in Figure 2.1 by curves <strong>of</strong> the same form without intersection.<br />

Note that the chart is double logarithmic <strong>and</strong> therefore the real proportions are not clearly<br />

visible. In the wavelength range <strong>of</strong> 8 - 14 μm for example the relation <strong>of</strong> the emittance <strong>of</strong><br />

the sun (5500 K curve) to that <strong>of</strong> a black body at 300 K is 470:1 [Schuster <strong>and</strong> Kolobrodov<br />

2004].


Single <strong>droplet</strong> <strong>drying</strong> 9<br />

Figure 2.1: Planck’s law according to Schuster <strong>and</strong> Kolobrodov [2004]<br />

The curves in Figure 2.1 also show that the wavelength at which a black body has its<br />

maximum emittance changes with the <strong>temperature</strong> <strong>of</strong> the body. The location <strong>of</strong> the<br />

maximum can be calculated by Wien’s displacement law:<br />

Equation 2.6<br />

λ <br />

2898 · μ<br />

<br />

The lower the <strong>temperature</strong> <strong>of</strong> the object, the higher is the wavelength for the maximum<br />

emittance [Schuster <strong>and</strong> Kolobrodov 2004].<br />

Black bodies do not exist in nature, though its characteristics are approximated by a<br />

hole in a box filled with highly absorptive material. For a real body the incident radiation is<br />

partly reflected, absorbed or transmitted [Engineering Toolbox 2009]. The fraction <strong>of</strong> the<br />

radiation falling on a body that is reflected is called the reflectivity , the fraction that is<br />

absorbed is called absorptivity <strong>and</strong> the fraction that is transmitted is called the<br />

transmissivity [McCabe et al. 2005]. All parts are wavelength-dependent, <strong>and</strong> the sum <strong>of</strong><br />

all fractions must be unity:


Single <strong>droplet</strong> <strong>drying</strong> 10<br />

Equation 2.7<br />

αλ λτλ 1<br />

According to Kirchh<strong>of</strong>f’s law for a body in thermal equilibrium the absorptivity is equal to<br />

the emissivity <strong>and</strong> the equation can be written as:<br />

Equation 2.8<br />

ελ λτλ 1<br />

Solids <strong>and</strong> liquids are usually not transparent for infrared radiation, so that τ 0 <strong>and</strong> the<br />

focus is on the emissivity <strong>and</strong> the reflectivity <strong>of</strong> the object.<br />

The Stefan-Boltzmann law for real bodies uses the emissivity <strong>of</strong> the object:<br />

Equation 2.9<br />

ε · · <br />

The emissivity is defined as the emittance <strong>of</strong> an object in comparison to the emittance <strong>of</strong><br />

a black body:<br />

Equation 2.10<br />

ε <br />

<br />

<br />

with ε 1 for a real body. A black body absorbs all radiation incident upon it, so that<br />

α ε 1. The emissivity <strong>of</strong> an object must be determined experimentally. It depends on<br />

the material, the wavelength, the <strong>temperature</strong> <strong>of</strong> the object, the <strong>surface</strong> texture, <strong>and</strong> the<br />

direction <strong>of</strong> the emitted beam. For a grey body the emissivity is unique for all wavelengths.<br />

If the emissivity is wavelength dependent, the object is called a selective radiator. Figure<br />

2.2 shows the idealized course <strong>of</strong> the emissivity. For a black body 1 for all<br />

wavelengths (curve 1). If 0 <strong>and</strong> 1, the substance is an ideal mirror, because it<br />

does not absorb radiation but reflects all <strong>of</strong> it (curve 2). Curve 2 is also characteristic for an<br />

ideal window in the case <strong>of</strong> 0 <strong>and</strong> 1, where all incident radiation is transmitted. A<br />

grey body has a constant emissivity at all wavelength const (curve 3), <strong>and</strong> for a<br />

selective radiator the emissivity depends on the wavelength <strong>of</strong> the measurement (curve 4).<br />

The emissivity for various substances can be found in the literature, keeping in<br />

mind that the emissivity is dependent on many parameters. For a few substances the<br />

emissivity is shown in Table 2.2.


Single <strong>droplet</strong> <strong>drying</strong> 11<br />

Figure 2.2: Emissivity depending on wavelength [Schuster <strong>and</strong> Kolobrodov 2004]<br />

Table 2.2: Emissivity <strong>of</strong> selected substances (20 °C, , angle <strong>of</strong><br />

radiation ± 20°) [Schuster <strong>and</strong> Kolobrodov 2004]<br />

Material<br />

Emissivity ε<br />

Iron (unmachined) 0.74<br />

Iron (rusty) 0.69<br />

Wooden board 0.96<br />

Water 0.96<br />

Snow 0.85<br />

Human skin 0.98<br />

2.3 Diffusion<br />

Diffusion is the movement, under the influence <strong>of</strong> a physical stimulus, <strong>of</strong> an individual<br />

component through a mixture. The most common cause <strong>of</strong> diffusion is a concentration<br />

gradient <strong>of</strong> the diffusing component, which tends to move the component in such a<br />

direction as to equalize concentrations <strong>and</strong> remove the gradient. In many mass-transfer<br />

operations the steady-state flux <strong>of</strong> the diffusing component takes place, where the gradient<br />

is maintained by constantly supplying the diffusing component to the high-concentration<br />

end <strong>of</strong> the gradient <strong>and</strong> removing it at the low-concentration end [McCabe et al. 2005].


Single <strong>droplet</strong> <strong>drying</strong> 12<br />

2.3.1 Fick’s first law <strong>of</strong> diffusion<br />

For one-dimensional steady-state diffusion the following general equation, Fick’s first law<br />

<strong>of</strong> diffusion, for the molar flux <strong>of</strong> component A, (similar to heat flux ), is given by:<br />

Equation 2.11<br />

· <br />

<br />

using the binary diffusion coefficient <strong>and</strong> the concentration gradient ⁄ for a<br />

binary mixture [McCabe et al. 2005].<br />

2.3.2 Diffusion coefficients for gases<br />

Diffusivities are best estimated by experimental measurements, but such information is not<br />

always available for the system <strong>of</strong> interest. In this case the values can be estimated from<br />

published correlations. Binary gas-phase diffusion coefficients in cm²/s at a given<br />

<strong>temperature</strong> in °C <strong>and</strong> pressure in atm can be estimated by the method described by<br />

Fuller et al. in 1966:<br />

Equation 2.12<br />

1.00 · 10 . · ·1⁄<br />

1⁄<br />

<br />

·∑ ⁄ <br />

⁄ <br />

∑<br />

<br />

<br />

⁄<br />

<br />

<strong>and</strong> are the molar masses in g/mol <strong>of</strong> the substances <strong>and</strong> ∑ <strong>and</strong> ∑ are the<br />

summed atomic diffusion volumes given in Table 2.3 [Fuller et al. 1966].<br />

Table 2.3: Special atomic diffusion volumes [Fuller et al. 1966]<br />

Atom / substance<br />

Diffusion volume<br />

C 16.5<br />

H 1.98<br />

O 5.48<br />

N 2 17.9<br />

Air 20.1<br />

H 2 O 12.7


Single <strong>droplet</strong> <strong>drying</strong> 13<br />

2.3.3 Convective mass transfer<br />

Mass transfer from the <strong>surface</strong> to a moving fluid or between immiscible fluids depends on<br />

the material property <strong>of</strong> the components <strong>and</strong> the flow properties. As in heat transfer by<br />

convection, convective mass transfer in a current can be caused by density differences due<br />

to a gradient in <strong>temperature</strong> or concentration leading to free convection. It can also be<br />

caused by external influences by a pump for example leading to forced convection [Baehr<br />

<strong>and</strong> Stephan 2004]. Equimolar diffusion (both components move) <strong>and</strong> unimolar diffusion<br />

(where component A diffuses through a non-diffusing component B) have to be<br />

differentiated. Figure 2.3 gives a comparison <strong>of</strong> the concentration gradient for equimolar<br />

<strong>and</strong> unimolar diffusion. In Figure 2.3 (a) components A <strong>and</strong> B are diffusing at same molar<br />

rates in opposite directions. In Figure 2.3 (b) component A is diffusing <strong>and</strong> component B is<br />

stationary with respect to interface. For a broad summary <strong>of</strong> calculations for different<br />

convective mass transfer settings see Schiffter [2006], McCabe et al. [2005] or Baehr <strong>and</strong><br />

Stephan [2004].<br />

(a)<br />

Figure 2.3: Concentration gradient for (a) equimolar <strong>and</strong> (b) unimolar diffusion<br />

[McCabe et al. 2005]<br />

2.4 Mass transfer theories<br />

Several mass transfer theories are used to describe mass transfer <strong>and</strong> to calculate mass<br />

transfer coefficients. The film theory is based on a study by Lewis <strong>and</strong> Whitman in 1924.<br />

(b)


Single <strong>droplet</strong> <strong>drying</strong> 14<br />

Mass transfer from a <strong>surface</strong> to a moving fluid takes place in a very thin fluid film close to<br />

the <strong>surface</strong>. Concentrations <strong>and</strong> velocities are allowed to change in the direction<br />

perpendicular to the <strong>surface</strong> only, not with time or in the other two dimensions <strong>of</strong> space<br />

[Baehr <strong>and</strong> Stephan 2004]. The film theory is <strong>of</strong>ten used as a basis for complex problems<br />

<strong>of</strong> multicomponent diffusion or diffusion plus chemical reaction. The boundary layer<br />

theory is based on the assumption that mass transfer takes place in a thin boundary layer<br />

near a <strong>surface</strong> where the fluid is in laminar flow [McCabe et al. 2005]. In contrast to film<br />

theory, concentrations <strong>and</strong> velocities are allowed to change in three dimensions in space. In<br />

film theory <strong>and</strong> boundary layer theory a steady-state mass transfer is assumed. Both<br />

theories are not applicable if concentrations change with time [Baehr <strong>and</strong> Stephan 2004].<br />

The penetration theory makes use <strong>of</strong> the expression for the transient rate <strong>of</strong><br />

diffusion into a relatively thick mass <strong>of</strong> fluid with a constant concentration at the <strong>surface</strong>.<br />

The theory was first used by Higbie in 1935 for the situation <strong>of</strong> gas absorption in a liquid,<br />

showing that diffusing molecules will not reach the other side <strong>of</strong> a thin layer if the contact<br />

time is short [McCabe et al. 2005]. In 1951 Danckwerts developed the <strong>surface</strong> renewal<br />

theory as an extension <strong>of</strong> the penetration theory. Higbie stated that all elements <strong>of</strong> the fluid<br />

have the same contact time, whereas Danckwerts expected that fluid elements have<br />

different contact times so that elements <strong>of</strong> fluid at a transfer <strong>surface</strong> are r<strong>and</strong>omly replaced<br />

after some time with fresh fluid from the bulk stream [Baehr <strong>and</strong> Stephan 2004; McCabe et<br />

al. 2005].<br />

The two film theory, proposed by Whitman in 1923, recognizes that in many<br />

separation processes material must diffuse from one phase into another, <strong>and</strong> the rates <strong>of</strong><br />

diffusion in both phases affect the overall rate <strong>of</strong> mass transfer. In the two film theory<br />

equilibrium is assumed at the interface, <strong>and</strong> the resistances to mass transfer in the two<br />

phases are added to give an overall resistance [McCabe et al. 2005].<br />

2.5 Evaporation <strong>of</strong> pure solvent <strong>droplet</strong>s<br />

2.5.1 Evaporation <strong>of</strong> a single <strong>droplet</strong><br />

A single <strong>droplet</strong> with radius <strong>and</strong> <strong>temperature</strong> is brought into a gas phase with<br />

<strong>temperature</strong> <strong>and</strong> mass fraction <strong>of</strong> the vapour <strong>of</strong> the <strong>droplet</strong> material. Figure 2.4<br />

shows two possibilities for mass flux, heat flux <strong>and</strong> <strong>temperature</strong> pr<strong>of</strong>ile in the evaporating<br />

<strong>droplet</strong>. In the first example (Figure 2.4 (a)) the <strong>temperature</strong> <strong>of</strong> the <strong>droplet</strong> <strong>and</strong> the


Single <strong>droplet</strong> <strong>drying</strong> 15<br />

ambiance are initially the same. An evaporation process starts <strong>and</strong> a mass flux <strong>of</strong><br />

<strong>droplet</strong> material will leave the <strong>droplet</strong>. The latent heat <strong>of</strong> evaporation is taken from the<br />

inner energy <strong>of</strong> the <strong>droplet</strong> liquid, so that the <strong>temperature</strong> <strong>of</strong> the <strong>droplet</strong> will decrease <strong>and</strong> a<br />

<strong>temperature</strong> gradient will be established in the <strong>droplet</strong> <strong>and</strong> in the surrounding gas. The<br />

<strong>temperature</strong> inside the <strong>droplet</strong> is not constant. <br />

<strong>and</strong> <br />

represent the heat flux in the<br />

<strong>droplet</strong> <strong>and</strong> in the surrounding gas towards the <strong>droplet</strong> <strong>surface</strong>. In Figure 2.4 (b) the<br />

ambient <strong>temperature</strong> is higher than the boiling point <strong>of</strong> the <strong>droplet</strong> liquid. Here the latent<br />

heat <strong>of</strong> evaporation is delivered by the heat flux <br />

, which is directed towards the<br />

<strong>droplet</strong>. The <strong>droplet</strong> is heated until the heat that is delivered to the <strong>surface</strong> is completely<br />

needed for the latent heat <strong>of</strong> evaporation. Here the <strong>droplet</strong> stabilizes at a <strong>temperature</strong> below<br />

the boiling point [Frohn <strong>and</strong> Roth 2000].<br />

(a)<br />

(b)<br />

Figure 2.4: Mass flux, heat fluxes <strong>and</strong> radial <strong>temperature</strong> pr<strong>of</strong>iles in evaporating<br />

<strong>droplet</strong>s, (a) <strong>temperature</strong> <strong>of</strong> the <strong>droplet</strong> <strong>and</strong> the ambience are initially the same, (b)<br />

ambient <strong>temperature</strong> is higher than the boiling point <strong>of</strong> the <strong>droplet</strong> liquid [Frohn <strong>and</strong><br />

Roth 2000]<br />

2.5.2 The d²-law<br />

Droplet evaporation at the wet-bulb <strong>temperature</strong> shows a linear decrease <strong>of</strong> its squared<br />

radius with time as described by the d²-law:<br />

Equation 2.13 0 ·<br />

where 0 is the initial <strong>droplet</strong> radius, the time <strong>and</strong> is the evaporation coefficient<br />

given by:


Single <strong>droplet</strong> <strong>drying</strong> 16<br />

Equation 2.14<br />

2· · <br />

<br />

·ln1<br />

The density <strong>of</strong> the <strong>droplet</strong> liquid is , the density <strong>of</strong> the gas at the <strong>droplet</strong> <strong>surface</strong> is ,<br />

the diffusion coefficient is <strong>and</strong> the Spalding transfer number is [Frohn <strong>and</strong> Roth<br />

2000].<br />

2.5.3 Diffusion-controlled evaporation <strong>of</strong> single <strong>droplet</strong>s<br />

In the case that a <strong>droplet</strong> is surrounded by an atmosphere that has approximately the same<br />

<strong>temperature</strong> as the <strong>droplet</strong> liquid, <strong>and</strong> further this <strong>temperature</strong> is low in comparison with<br />

the boiling point <strong>of</strong> the <strong>droplet</strong> liquid, the vaporization process is dominated by diffusion<br />

processes in the vapour phase [Frohn <strong>and</strong> Roth 2000]. This situation can be described by:<br />

Equation 2.15<br />

0 2· · <br />

·<br />

· <br />

<br />

<br />

<br />

·<br />

with the molar mass <strong>of</strong> the liquid , the ideal gas constant <strong>and</strong> the partial vapour<br />

pressure at the <strong>droplet</strong> <strong>surface</strong> <strong>and</strong> at infinity . The evaporation coefficient is<br />

given by:<br />

Equation 2.16<br />

2· · <br />

·<br />

· <br />

<br />

<br />

<br />

<br />

2.5.4 Evaporation <strong>of</strong> <strong>droplet</strong>s containing solvent mixtures<br />

For the evaporation <strong>of</strong> a single <strong>droplet</strong> containing different solvents, the decrease in the<br />

radius squared with time is not linear but depends on the mass flux <strong>of</strong> the components. For<br />

a binary solvent mixture, for example, the component with the higher mass flux will<br />

predominate the evaporation in the first <strong>drying</strong> stage. As soon as most <strong>of</strong> the component<br />

has evaporated, the second component will predominate the evaporation process <strong>and</strong> the<br />

graph will show a slower decrease <strong>of</strong> the squared radius in the second stage [Yarin et al.<br />

2002]. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> changes during the <strong>drying</strong> process, when the<br />

second component starts to dominate evaporation [Tuckermann et al. 2005].


Single <strong>droplet</strong> <strong>drying</strong> 17<br />

2.6 Evaporation <strong>of</strong> solution <strong>and</strong> suspension <strong>droplet</strong>s<br />

2.6.1 Drying stages <strong>of</strong> solution <strong>and</strong> suspension <strong>droplet</strong>s<br />

The <strong>drying</strong> process <strong>of</strong> a solution or suspension <strong>droplet</strong> can be divided into two stages for<br />

non-hygroscopic materials <strong>and</strong> three stages for hygroscopic materials. When the <strong>droplet</strong><br />

comes in contact with the hot <strong>drying</strong> air, there is a very short phase for heating-up <strong>of</strong> the<br />

<strong>droplet</strong>. Then in the first <strong>drying</strong> stage, solvent evaporation takes place from the<br />

completely wetted <strong>droplet</strong> <strong>surface</strong> comparable to the evaporation <strong>of</strong> pure solvent <strong>droplet</strong>s.<br />

The limiting factor for evaporation is the mass transfer from the <strong>droplet</strong> <strong>surface</strong> to the<br />

<strong>drying</strong> air, <strong>and</strong> therefore the change in <strong>droplet</strong> diameter can be calculated using the d²-law<br />

as for pure solvent <strong>droplet</strong>s [Grassmann et al. 1997; Kastner 2001]. Diffusion <strong>of</strong> moisture<br />

in liquid form from inside the <strong>droplet</strong> maintains saturated <strong>surface</strong> conditions. Capillary <strong>and</strong><br />

diffusional mechanisms are involved. Their dominance depends upon the nature <strong>of</strong> the<br />

solid in the <strong>droplet</strong> as a solution or suspension. As long as the saturated <strong>surface</strong> conditions<br />

last, evaporation takes place at a constant rate. For this reason the first <strong>drying</strong> stage is also<br />

called the constant rate period. In this stage the evaporation rate is the highest during<br />

<strong>droplet</strong> <strong>drying</strong>. Figure 2.5 shows the evaporation rate depending on <strong>drying</strong> time. The first<br />

<strong>drying</strong> stage is represented by curve A-B. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> approximates<br />

the wet-bulb <strong>temperature</strong> <strong>and</strong> remains constant in the first <strong>drying</strong> stage. Solvent<br />

evaporation leads to a decrease in <strong>droplet</strong> diameter, <strong>and</strong> the mass fraction <strong>of</strong> the dissolved<br />

or suspended solid increases [Grassmann et al. 1997; Masters 2002].<br />

When the moisture content becomes too low to maintain saturated <strong>surface</strong><br />

conditions, the critical point (critical moisture content) is reached <strong>and</strong> a dry layer starts to<br />

form at the <strong>droplet</strong> <strong>surface</strong>. The <strong>surface</strong> is no longer completely wetted [Masters 2002].<br />

The critical point is the transition point from the first <strong>drying</strong> stage to the second <strong>drying</strong><br />

stage. The evaporation rate decreases <strong>and</strong> the evaporation curve shows a sharp inflexion<br />

point (Figure 2.5, point B). It is important to know when the critical point appears, because<br />

the evaporation principles between the first <strong>and</strong> the second <strong>drying</strong> stage differ [Grassmann<br />

et al. 1997].<br />

The second <strong>drying</strong> stage or falling rate period starts at the critical point. For a<br />

solution <strong>droplet</strong> the limiting factor for evaporation is the diffusion <strong>of</strong> moisture from inside<br />

the <strong>droplet</strong> through the dried <strong>surface</strong> layer, which presents a formidable resistance to mass<br />

transfer. The thickness <strong>of</strong> this layer increases with time as the moisture boundary retracts


Single <strong>droplet</strong> <strong>drying</strong> 18<br />

toward the centre <strong>of</strong> the <strong>droplet</strong> <strong>and</strong> the evaporation rate decreases (Figure 2.5, curve B-E<br />

for non-hygroscopic substances, curve B-C for hygroscopic substances). For a suspension<br />

<strong>droplet</strong> the limiting factor for evaporation is the capillary moisture flow in the pores<br />

forming in the particle [Kastner 2001; Masters 2002]. Due to decreasing evaporation rate,<br />

the <strong>droplet</strong> or particle <strong>surface</strong> <strong>temperature</strong> starts to increase after the critical point close to<br />

the ambient <strong>temperature</strong>.<br />

Figure 2.5: Evaporation rate with time (continuous line: hygroscopic substance,<br />

dashed line: non-hygroscopic substance) according to Grassmann et al. [1997]<br />

The <strong>drying</strong> mechanisms for solution <strong>and</strong> suspension <strong>droplet</strong>s differ in the second <strong>drying</strong><br />

stage [Masters 2002]. A solution <strong>droplet</strong> first evaporates at a constant rate, <strong>and</strong> the <strong>surface</strong><br />

<strong>temperature</strong> can be equated to that <strong>of</strong> a saturated solution. When the <strong>droplet</strong> moisture<br />

content falls to a critical value, solid phase appears at the <strong>surface</strong> <strong>and</strong> the constant rate<br />

ends. The rate <strong>of</strong> moisture migration to the <strong>surface</strong> decreases due to the increasing<br />

resistance to mass transfer through the solid phase. The <strong>droplet</strong> begins to heat up, because<br />

the rate <strong>of</strong> heat transfer exceeds that <strong>of</strong> mass transfer. If the heat transfer is high enough,<br />

sub-<strong>surface</strong> evaporation can occur <strong>and</strong> pressure can be built up in the <strong>droplet</strong>. If the crust is<br />

porous then the vapour can leak from the <strong>droplet</strong>. If it is a non-porous film then the <strong>droplet</strong><br />

may exp<strong>and</strong>, collapse, rupture or even disintegrate. The movement <strong>of</strong> volatiles in <strong>drying</strong> <strong>of</strong><br />

solution <strong>droplet</strong>s is first controlled by a liquid diffusional mechanism <strong>and</strong> then by a vapour<br />

diffusional mechanism. Each product shows different <strong>drying</strong> characteristics in the second


Single <strong>droplet</strong> <strong>drying</strong> 19<br />

<strong>drying</strong> stage, which are influenced by the solid <strong>surface</strong> layer. In a <strong>droplet</strong> with a porous<br />

<strong>surface</strong>, the vapour migrates easily through the crust <strong>and</strong> the <strong>drying</strong> rate is high. For a nonporous<br />

crust the <strong>drying</strong> rate will fall <strong>and</strong> evaporation time increases. A suspension <strong>droplet</strong><br />

evaporates also at a constant rate, as long as all suspended particles are completely wetted.<br />

Moisture loss brings the insoluble solids closer together, until capillary mechanisms are<br />

insufficient to maintain completely wet <strong>surface</strong>s. The falling rate period with particles in a<br />

funicular state continues until the pores are no longer filled with liquid <strong>and</strong> capillary flow<br />

ends. In the following pendular state conditions a vapour diffusion mechanism<br />

predominates until <strong>drying</strong> <strong>of</strong> the particle is completed [Masters 2002].<br />

For hygroscopic materials a third <strong>drying</strong> stage occurs <strong>and</strong> a second inflexion point<br />

in the evaporation curve appears (Figure 2.5, point C). The third <strong>drying</strong> stage begins as<br />

soon as the highest hygroscopic moisture content is reached. From here on the evaporation<br />

rate decreases faster than in the second <strong>drying</strong> stage (Figure 2.5, curve C-D) [Grassmann et<br />

al. 1997].<br />

The <strong>droplet</strong> temporal <strong>temperature</strong> curve together with the crust formation process<br />

<strong>of</strong> a spherical <strong>droplet</strong> <strong>of</strong> liquid containing solid is described in Figure 2.6 [Farid 2003].<br />

First the <strong>droplet</strong> is heated by the air until its <strong>surface</strong> reaches the air wet-bulb <strong>temperature</strong><br />

without substantial evaporation (Figure 2.6, curve A-B). In the following constant-rate<br />

period <strong>droplet</strong> shrinkage occurs, with the <strong>droplet</strong> <strong>temperature</strong> remaining at the wet-bulb<br />

<strong>temperature</strong> (Figure 2.6, curve B-C). The <strong>droplet</strong> does not heat up in this period <strong>and</strong> it can<br />

be assumed that all heat transferred to the <strong>droplet</strong> is used for evaporation. The shrinkage<br />

will continue until the solid starts crust formation. After the period <strong>of</strong> shrinkage,<br />

evaporation continues with the formation <strong>of</strong> a moving interface. This is assumed to move<br />

inward <strong>and</strong> separate the <strong>droplet</strong> into two regions, the dry crust <strong>and</strong> the inner wet core. The<br />

<strong>temperature</strong> <strong>of</strong> the interface is assumed to remain at the wet-bulb <strong>temperature</strong>, whereas the<br />

dry crust heats up. Farid [2003] calculated the average <strong>droplet</strong> <strong>temperature</strong> rising close to<br />

the <strong>temperature</strong> <strong>of</strong> the <strong>drying</strong> air (Figure 2.6, curve C-E). After <strong>drying</strong> is completed, the<br />

particle is heated up to the dry-bulb <strong>temperature</strong> <strong>of</strong> the air (Figure 2.6, curve E-F).


Single <strong>droplet</strong> <strong>drying</strong> 20<br />

Figure 2.6: Drying stages <strong>of</strong> a liquid <strong>droplet</strong> containing solids [Farid 2003]<br />

2.6.2 Particle formation<br />

During the <strong>drying</strong> process a <strong>droplet</strong> undergoes both size <strong>and</strong> shape changes, the extent <strong>of</strong><br />

which is specific to the given product. In several studies the influence <strong>of</strong> process <strong>and</strong><br />

formulation parameters in spray <strong>drying</strong> on the particle shape has been analyzed. Elversson<br />

et al. [2003] analyzed <strong>droplet</strong> <strong>and</strong> particle size relationship during spray <strong>drying</strong>. Increasing<br />

<strong>droplet</strong> size increased the end particle size, but the effect was also influenced by feed<br />

concentration whereby particles from low concentrated solutions were smaller than those<br />

from higher concentrations. In a further study they demonstrated that larger particles could


Single <strong>droplet</strong> <strong>drying</strong> 21<br />

be also obtained by decreasing the solubility <strong>of</strong> the solute. The apparent particle density<br />

was inversely correlated to the feed concentration [Elversson <strong>and</strong> Millqvist-F. 2005].<br />

Maa et al. [1997] investigated the shape <strong>and</strong> morphology <strong>of</strong> various spray dried<br />

protein powders as a function <strong>of</strong> spray <strong>drying</strong> conditions <strong>and</strong> also protein formulation.<br />

They observed a substantial change in particle shape from irregular to spherical with<br />

decreasing outlet <strong>temperature</strong>. In formulations with sugars, the spray dried particles<br />

exhibited a smooth <strong>surface</strong> when containing lactose. A rough <strong>surface</strong> appeared for<br />

formulations containing mannitol at more than 30 % <strong>of</strong> total solids due to recrystallization.<br />

The presence <strong>of</strong> surfactant resulted in noticeable smoother, more spherical particles. For<br />

bovine serum albumin (bSA) / trehalose / polysorbate 80 formulations Adler et al. [2000]<br />

confirmed that with rising amount <strong>of</strong> surfactant the <strong>surface</strong> shape becomes smooth,<br />

whereas bSA / trehalose particle are more wrinkled.<br />

Particle shapes formed during the spray <strong>drying</strong> process are given in Figure 2.7.<br />

After the atomized <strong>droplet</strong> contacts the hot <strong>drying</strong> air (phase 1) a dry <strong>surface</strong> forms <strong>and</strong> its<br />

properties control the ease <strong>of</strong> volatiles’ release <strong>and</strong> eventual particle shape <strong>and</strong> structure<br />

(phase 2). Phase 3 shows various shapes <strong>and</strong> structures <strong>of</strong> the dried particles, some tend to<br />

exp<strong>and</strong>, other collapse, form blow holes or agglomerate. Very thin-shelled particles can be<br />

so fragile that they disintegrate on mechanical h<strong>and</strong>ling [Masters 2002].<br />

Single <strong>droplet</strong> <strong>drying</strong> experiments are a useful tool to examine particle formation <strong>of</strong><br />

solution <strong>and</strong> suspension <strong>droplet</strong>s <strong>of</strong> inorganic <strong>and</strong> organic solids; some examples were<br />

already given in Table 2.1. Charlesworth <strong>and</strong> Marshall [1960] gave a summary <strong>of</strong> crust<br />

structure <strong>and</strong> consequential particle shape for suspended <strong>droplet</strong>s <strong>of</strong> an inorganic salt in<br />

aqueous solution. They found that the completion <strong>of</strong> the crust differed depending on the<br />

nature <strong>of</strong> the solute <strong>and</strong> its crust structure, <strong>and</strong> also the surrounding air <strong>temperature</strong> after<br />

the critical point. At a <strong>temperature</strong> <strong>of</strong> the surrounding air below the boiling point <strong>and</strong> with<br />

a rigid <strong>and</strong> porous crust structure, the particle shape does not change. Less porous crust<br />

structure can lead to fracture <strong>of</strong> the <strong>droplet</strong> <strong>and</strong> a crystal “fur” growing. Droplets with a<br />

pliable <strong>and</strong> impervious skin will dimple or wrinkle. If the <strong>drying</strong> air <strong>temperature</strong> is above<br />

the boiling point, the particle heats up after the critical point due to slowing <strong>of</strong> the mass<br />

transfer rate by the crust <strong>and</strong> vapour formed inside the particle. If the crust is rigid <strong>and</strong><br />

porous, liquid is forced through the pores <strong>and</strong> evaporates quickly. Droplets with a nonporous<br />

crust, which is not sufficiently permeable to the liquid, can fracture if the crust is<br />

rigid, or inflate if the crust is plastic [Charlesworth <strong>and</strong> Marshall 1960].


Single <strong>droplet</strong> <strong>drying</strong> 22<br />

Figure 2.7: Particle shapes formed during the spray <strong>drying</strong> process: 1: solid,<br />

spherical; 2: shrivelled, misshapen; 3: hollow, spherical; 4: cenospherical; 5:<br />

disintegrated [Masters 2002]<br />

In recent studies Kastner [2001] analyzed the <strong>drying</strong> <strong>of</strong> suspension <strong>droplet</strong>s in an ultrasonic<br />

levitator using glass particles suspended in a water <strong>droplet</strong> as a model. In later experiments<br />

he used a catalyst suspension. He developed a model for the computation <strong>of</strong> the <strong>drying</strong><br />

process. Schiffter <strong>and</strong> Lee [2007a; 2007b] investigated the <strong>drying</strong> behaviour <strong>of</strong><br />

formulations containing proteins, sugars <strong>and</strong> surfactants <strong>and</strong> compared the particle shape to<br />

spray dried powders, obtaining good agreement.<br />

2.6.3 Models for the <strong>drying</strong> <strong>of</strong> single <strong>droplet</strong>s containing solids<br />

Several models for heat <strong>and</strong> mass transfer during the <strong>drying</strong> <strong>of</strong> a single <strong>droplet</strong> containing<br />

solids have been developed <strong>and</strong> described in the literature. Especially the changes in heat<br />

<strong>and</strong> mass transfer in the second <strong>drying</strong> stage have to be taken into account, in contrast to<br />

the evaporation <strong>of</strong> a pure solvent <strong>droplet</strong>. Several single <strong>droplet</strong> <strong>drying</strong> models are<br />

summarised in Table 2.4.


Single <strong>droplet</strong> <strong>drying</strong> 23<br />

Table 2.4: Single <strong>droplet</strong> <strong>drying</strong> models [Adhikari et al. 2000; Farid 2003]<br />

Investigator<br />

Wijlhuizen, Kerkh<strong>of</strong> <strong>and</strong> Bruin<br />

[Wijlhuizen et al. 1979]<br />

Sano <strong>and</strong> Keey<br />

[Sano <strong>and</strong> Keey 1982]<br />

Nesic <strong>and</strong> Vodnic<br />

[Nesic <strong>and</strong> Vodnik 1991]<br />

Cheong, Jeffreys <strong>and</strong> Mumford<br />

[Cheong et al. 1986]<br />

Farid<br />

[Farid 2003]<br />

Model conditions<br />

Droplet is at a uniform <strong>temperature</strong>, water diffuses<br />

through the solid <strong>and</strong> evaporates at the <strong>surface</strong> <strong>of</strong> the<br />

<strong>droplet</strong><br />

Droplet is at a uniform <strong>temperature</strong>, model is based<br />

on the formation <strong>of</strong> crust with a receding crust-bulk<br />

interface, crust sensible heat was not included for<br />

<strong>droplet</strong> <strong>temperature</strong> calculations<br />

Model is based on a receding interface model with a<br />

<strong>droplet</strong> core <strong>temperature</strong> different from that <strong>of</strong> the<br />

<strong>surface</strong><br />

Moving boundary model considering a <strong>temperature</strong><br />

distribution inside the <strong>droplet</strong> <strong>and</strong> <strong>droplet</strong> shrinkage


Acoustic levitation 24<br />

3 Acoustic levitation<br />

3.1 History <strong>and</strong> application <strong>of</strong> acoustic levitation<br />

In an acoustic levitator a <strong>droplet</strong> or particle can be levitated in a pressure node <strong>of</strong> a<br />

st<strong>and</strong>ing acoustic wave generated by a piezoelectric crystal in the transducer. Acoustic<br />

levitation provides the opportunity to perform single <strong>droplet</strong> <strong>drying</strong> experiments without<br />

contact to a thermal-conducting holding device like a thermocouple or wire. The first<br />

theoretical background was described by King in 1934 <strong>and</strong> Gorkov in 1962. Further<br />

investigations were performed in the early 70s by the space agencies NASA <strong>and</strong> ESA for<br />

the development <strong>of</strong> a tool for “containerless processing” under microgravity conditions. In<br />

addition to acoustic levitation other levitation principles were also used, like electrostatic,<br />

electromagnetic <strong>and</strong> aerodynamic levitation. Today acoustic levitation is readily performed<br />

in terrestrial laboratories <strong>and</strong> is used for experiments in various fields <strong>of</strong> research [Lierke<br />

1996a; Lierke 2002].<br />

Lierke [1996a] described application fields for acoustic levitation experiments in<br />

exactly defined experimental conditions. These are controlled evaporation experiments,<br />

concentration <strong>of</strong> samples for trace analysis, single crystal growth, heat <strong>and</strong> mass transfer<br />

between single <strong>droplet</strong>s <strong>and</strong> their ambiance, measurements at isolated biological samples,<br />

or stability studies <strong>of</strong> suspensions <strong>and</strong> emulsions. A summary <strong>of</strong> experimental research<br />

using an acoustic levitation system for the application <strong>of</strong> density, <strong>surface</strong> tension <strong>and</strong><br />

viscosity measurements, micro <strong>and</strong> trace analysis, gas-solid reactions <strong>and</strong> <strong>determination</strong> <strong>of</strong><br />

enzyme <strong>kinetics</strong> for example is given by Schiffter [2006].<br />

Single <strong>droplet</strong> <strong>drying</strong> studies using an ultrasonic levitator were used to describe<br />

evaporation processes in several studies. The influence <strong>of</strong> the acoustic field <strong>and</strong> acoustic<br />

streaming on the evaporation process was examined [Burdukov et al. 1963]. Seaver et al.<br />

[1989] worked with an acoustic levitator in a free-jet wind tunnel <strong>and</strong> analyzed evaporation<br />

<strong>of</strong> liquid <strong>droplet</strong>s using different initial <strong>droplet</strong> sizes <strong>and</strong> the condensation <strong>of</strong> vapour on<br />

evaporating <strong>droplet</strong>s. In more recent studies Yarin et al. [1998] developed a method for<br />

calculation <strong>of</strong> the acoustic radiation pressure <strong>and</strong> gave a relationship between <strong>droplet</strong><br />

shape, <strong>droplet</strong> displacement <strong>and</strong> sound pressure in an acoustic levitator. In later studies<br />

they developed further calculations for the influence <strong>of</strong> the acoustic streaming on both<br />

<strong>droplet</strong> <strong>drying</strong> <strong>and</strong> <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> [Yarin et al. 1999]. The <strong>drying</strong> process <strong>of</strong>


Acoustic levitation 25<br />

solution <strong>and</strong> suspension <strong>droplet</strong>s was investigated by Kastner [2001] using an acoustic tube<br />

levitator, who also developed a model for the evaporation process <strong>of</strong> <strong>droplet</strong>s. The<br />

influence <strong>of</strong> <strong>droplet</strong> <strong>surface</strong> oscillations for different solvents, solvent mixtures <strong>and</strong><br />

suspensions on the evaporation was studied by Rensink [2004]. For <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> measurement Tuckermann et al. [2005] used an infrared camera <strong>and</strong> carried<br />

out experiments with different solvent <strong>droplet</strong>s. They also analyzed the formation <strong>of</strong><br />

octadecanol monolayers on a liquid <strong>droplet</strong> using the <strong>temperature</strong> change as a detection<br />

method [Tuckermann et al. 2007]. For pharmaceutical applications Schiffter <strong>and</strong> Lee<br />

[2007a; 2007b] analyzed excipients <strong>and</strong> model proteins in different formulations <strong>and</strong> under<br />

different experimental conditions. Drying <strong>kinetics</strong>, particle formation <strong>and</strong> protein activity<br />

in comparison to spray dried powders were examined.<br />

3.2 A short introduction to acoustics<br />

Acoustic waves are mechanical waves that spread in gases, liquids, <strong>and</strong> solids due to their<br />

elastic properties. Mechanical oscillations occur when molecules <strong>of</strong> the substance are<br />

moved out <strong>of</strong> their position <strong>of</strong> equilibrium by external forces. Due to elastic <strong>and</strong> inertial<br />

behaviour they start to oscillate periodically around their rest position. In acoustic waves<br />

the molecules do not move forwards, rather just the sound energy is transported. The<br />

appearance <strong>of</strong> sound depends on the existence <strong>of</strong> material, <strong>and</strong> therefore acoustic waves do<br />

not spread in a vacuum [Günther et al. 1978]. In liquids <strong>and</strong> gases acoustic waves occur as<br />

longitudinal waves. The human ear is able to hear frequencies from 16 Hz to 20 kHz.<br />

Ultrasonic waves have a frequency <strong>of</strong> > 20 kHz, <strong>and</strong> acoustic waves having a frequency <strong>of</strong><br />

< 16 Hz are called infrasound [Stroppe 2008].<br />

The sound velocity in a liquid can be calculated using the bulk modulus<br />

elasticity <strong>and</strong> the liquid density [Stroppe 2008]:<br />

Equation 3.1<br />

<br />

<br />

For an ideal gas the sound velocity is calculated using the adiabatic index , the ideal gas<br />

constant , the absolute <strong>temperature</strong> <strong>and</strong> the molar mass <strong>of</strong> the gas [Stroppe 2008]:<br />

Equation 3.2<br />

··


Acoustic levitation 26<br />

The particle velocity <strong>and</strong> the sound pressure are used to describe the sound field <strong>and</strong> to<br />

calculate the sound pressure level. The particle velocity is the alternate velocity <strong>of</strong> a<br />

particle in a medium that transmits a longitudinal wave <strong>of</strong> pressure. It is calculated for<br />

sinusoidal movement using the particle displacement <strong>and</strong> the angular frequency ω<br />

[Günther et al. 1978]:<br />

Equation 3.3<br />

··2··<br />

The sound pressure is the force <strong>of</strong> sound on a <strong>surface</strong> area perpendicular to the<br />

direction <strong>of</strong> sound. is an alternating quantity, that means it is a function <strong>of</strong> time<br />

which can be expressed as a peak value, arithmetic mean value, instantaneous value or as<br />

root-mean-square-value. In most cases the root-mean-square-value, the effective sound<br />

pressure , is used [Günther et al. 1978]. For sinusoidal waves the relation is<br />

[Hering et al. 2007]:<br />

Equation 3.4<br />

<br />

<br />

√2<br />

The relation <strong>of</strong> sound pressure <strong>and</strong> particle velocity in a plane wave is constant at any point<br />

<strong>and</strong> any time, so points with maximum particle velocity have also maximum sound<br />

pressure values. The proportionality factor is characteristic for the transmitting medium<br />

<strong>and</strong> is called the characteristic acoustic impedance . It is dependent on the density <strong>of</strong><br />

the medium <strong>and</strong> the sound velocity [Stroppe 2008]:<br />

Equation 3.5<br />

<br />

<br />

· <br />

Using this relation can be calculated by [Stroppe 2008]:<br />

Equation 3.6<br />

· ·


Acoustic levitation 27<br />

The sound pressure level (SPL or ) characterizes the acoustic field. Using the effective<br />

sound pressure values in the equation for it is defined as the logarithmic value <strong>of</strong> the<br />

sound pressure compared to a reference sound pressure [Hering et al. 2007]:<br />

Equation 3.7<br />

20·log <br />

<br />

The reference sound pressure value is = 2·10 -5 Pa (human auditory threshold) <strong>and</strong><br />

has the nondimensional unit decibel (dB) expressing the ratio <strong>of</strong> two values. A different<br />

definition for the SPL that is used for st<strong>and</strong>ing acoustic waves by Yarin et al. [1998] is<br />

introduced in Equation 3.24.<br />

A st<strong>and</strong>ing acoustic wave is generated by reflection <strong>of</strong> an acoustic wave. At the<br />

<strong>surface</strong> <strong>of</strong> reflection there is 0 <strong>and</strong> 0, whereas doubles. In the st<strong>and</strong>ing<br />

acoustic wave sound pressure <strong>and</strong> particle velocity nodes <strong>and</strong> antinodes are phase-shifted<br />

by 90° [Günther et al. 1978].<br />

3.3 Levitation forces<br />

3.3.1 St<strong>and</strong>ing acoustic wave<br />

In an acoustic levitator an acoustic wave generated by a piezoelectric crystal in the<br />

transducer is reflected creating a st<strong>and</strong>ing acoustic wave when the distance between<br />

transducer <strong>and</strong> reflector is an integral multiple <strong>of</strong> the half wavelength. In the st<strong>and</strong>ing<br />

acoustic wave pressure nodes <strong>and</strong> anti-nodes appear at fixed points separated by a distance<br />

<strong>of</strong> λ⁄ 2 (Figure 3.1). Small liquid or solid samples can be levitated in the vicinity <strong>of</strong> the<br />

pressure nodes. Under gravitational conditions the sphere is levitated just below the<br />

pressure node, dependent on its volume <strong>and</strong> density [Tuckermann 2002]. Calculations <strong>of</strong><br />

the position <strong>of</strong> the pressure nodes <strong>and</strong> anti-nodes <strong>and</strong> the levitation forces are given by<br />

Yarin et al. [1998].


Acoustic levitation 28<br />

Transducer<br />

Sound pressure<br />

Sample<br />

Sound particle<br />

velocity<br />

Reflector<br />

Figure 3.1: Sound pressure <strong>and</strong> sound particle velocity in a st<strong>and</strong>ing acoustic wave<br />

[Schiffter 2006; Tuckermann 2002]<br />

3.3.2 Interaction <strong>of</strong> ultrasonic wave <strong>and</strong> <strong>droplet</strong><br />

The levitation forces have influence on the <strong>droplet</strong><br />

shape, because axial <strong>and</strong> radial forces are unequal.<br />

Calculations <strong>of</strong> the forces are given by Lierke<br />

[1996b]. The axial forces are mainly responsible for<br />

compensating the gravitational forces, whereas the<br />

radial forces (Bernoulli forces) hold the sample<br />

horizontally in the pressure node, if there is radial<br />

displacement <strong>of</strong> the sample. For a deformable<br />

sample the axial <strong>and</strong> radial levitation forces lead to a<br />

deformation <strong>of</strong> the sample to the appearance <strong>of</strong> a<br />

spheroid (Figure 3.2). With increasing sound<br />

pressure the deformation increases; the sample can<br />

even disintegrate [Tuckermann 2002]. The ratio<br />

F r<br />

F z<br />

F z<br />

F r<br />

Figure 3.2: Radial (F r ) <strong>and</strong> axial<br />

(F z ) levitation forces according to<br />

Tuckermann [2002]


Acoustic levitation 29<br />

between the horizontal <strong>and</strong> vertical diameter is the aspect ratio <strong>of</strong> the <strong>droplet</strong>. It depends<br />

not only on the SPL, but also on the <strong>surface</strong> tension <strong>and</strong> the <strong>droplet</strong> size.<br />

If a sample is levitated in a pressure node, then some reflection <strong>of</strong> the ultrasonic<br />

wave occurs that is dependent on the sample size. As a result the SPL decreases. At the<br />

beginning <strong>of</strong> <strong>drying</strong> the <strong>droplet</strong> has a known volume <strong>and</strong> a definite SPL exists sufficient to<br />

levitate it. During the <strong>drying</strong> process <strong>droplet</strong> shrinkage leads to an increase in SPL due to a<br />

decreasing reflection area. The calculative link between <strong>droplet</strong> shape / <strong>droplet</strong> displacement<br />

<strong>and</strong> SPL is given by Yarin et al. [1998]. Kastner [2001] showed that the calculation<br />

model by Yarin corresponds with experimental values for the SPL. Rensink [2004]<br />

experimentally verified an increasing SPL with decreasing <strong>droplet</strong> volume for several pure<br />

solvents.<br />

3.4 Acoustic streaming<br />

3.4.1 Inner acoustic streaming<br />

The acoustic field induces streaming in the gas around a levitated <strong>droplet</strong>. This acoustical<br />

streaming can be divided into inner acoustic streaming that is located directly on the<br />

<strong>surface</strong> <strong>of</strong> the levitated <strong>droplet</strong> forming an acoustic boundary layer, <strong>and</strong> outer acoustic<br />

streaming which forms two toroidal vortices above <strong>and</strong> below the <strong>droplet</strong>. A sketch <strong>of</strong> the<br />

inner acoustic streaming field is given in Figure 3.3 taken from Yarin et al. [1999]. The<br />

four closed loops (AB, CD, EF, GH) inside the acoustic boundary layer represent the<br />

structure <strong>of</strong> the steady state flow with the acoustic <strong>and</strong> diffusion boundary layers <strong>and</strong><br />

. The acoustic boundary layer is very thin in comparison to the diffusion boundary<br />

layer. The inner acoustic streaming has to be taken into account for the <strong>droplet</strong> evaporation<br />

process, because it induces an additional convective influence that can increase mass<br />

transfer.


Acoustic levitation 30<br />

Figure 3.3: Inner acoustic streaming according to Yarin et al. [1999]<br />

3.4.2 Outer acoustic streaming<br />

Figure 3.4 shows the outer acoustical streaming field [Yarin et al. 1999]. The streaming <strong>of</strong><br />

the outer vortex on the left side above the <strong>droplet</strong> is left-h<strong>and</strong>ed <strong>and</strong> the one below the<br />

<strong>droplet</strong> is right-h<strong>and</strong>ed. The surrounding air streams therefore towards the <strong>droplet</strong><br />

equatorially <strong>and</strong> leaves the <strong>droplet</strong> at the poles [Kastner 2001]. In the outer acoustic<br />

streaming field vapour <strong>of</strong> the <strong>droplet</strong> liquid can accumulate. This has a wide influence on<br />

<strong>droplet</strong> evaporation because the evaporation rate can be slowed down. Using an axial<br />

ventilation airstream the vapour can, however, be blown out <strong>of</strong> the vortices, causing the<br />

concentration <strong>of</strong> vapour to decrease <strong>and</strong> the evaporation rate to increase. The ventilation<br />

airstream has to be adjusted so that the amount <strong>of</strong> vapour blown out <strong>of</strong> the vortices is equal<br />

to the amount <strong>of</strong> vapour accumulating in the vortices from the <strong>droplet</strong>. If the ventilation<br />

airstream is too strong, so that it has influence not only on the outer acoustic streaming but<br />

also on the <strong>droplet</strong> <strong>surface</strong>, a further convectional mass transfer is induced together with a<br />

rising evaporation rate [Rensink 2004].


Acoustic levitation 31<br />

Figure 3.4: Outer acoustic streaming according to Yarin et al. [1999]<br />

3.5 Vertical position <strong>of</strong> the levitated <strong>droplet</strong><br />

The position <strong>of</strong> the sample in an ultrasonic levitator under gravitational forces lies slightly<br />

below the pressure node due to mass <strong>of</strong> the <strong>droplet</strong>, as explained in 3.3. This displacement<br />

also depends on the SPL. A higher SPL as well as mass loss from the sample leads to a<br />

higher <strong>droplet</strong> position. During <strong>droplet</strong> evaporation there is mass loss <strong>of</strong> the <strong>droplet</strong>, <strong>and</strong><br />

this <strong>droplet</strong> shrinkage leads to an increase in the SPL. The <strong>droplet</strong> rises in the ultrasonic<br />

field towards the next upper pressure node during the <strong>drying</strong> process.<br />

Kastner [2001] performed calculations to analyze the influence <strong>of</strong> volume decrease<br />

(at constant density) <strong>and</strong> density decrease (at constant volume). A decrease in density has a<br />

much higher influence on the sample position than does the <strong>droplet</strong> volume [Kastner<br />

2001].


Acoustic levitation 32<br />

3.6 Temperature <strong>of</strong> the levitation system<br />

As soon as the levitator is put in operation the levitator transducer <strong>and</strong> the levitation<br />

chamber start to heat up due to oscillations <strong>of</strong> the transducer. Kastner found a <strong>temperature</strong><br />

increase <strong>of</strong> about 5 °C at an ambient <strong>temperature</strong> <strong>of</strong> 20 °C without additional ventilation in<br />

the levitation system [Kastner 2001]. For experiments at room <strong>temperature</strong> this<br />

<strong>temperature</strong> increase has to be taken into account, because the <strong>temperature</strong> increase<br />

influences the evaporation rate. Experiments at higher <strong>temperature</strong> need much external<br />

heat supply, so that here the <strong>temperature</strong> increase from the levitator can be neglected.<br />

3.7 Single <strong>droplet</strong> <strong>drying</strong> using an acoustic levitator<br />

3.7.1 Pure solvent <strong>droplet</strong>s<br />

For the calculation <strong>of</strong> <strong>droplet</strong> <strong>drying</strong> <strong>kinetics</strong> in an ultrasonic levitator the influence <strong>of</strong> the<br />

ultrasonic wave on the evaporation rate has to be included. The evaporation coefficient <br />

in the equation for diffusion-controlled evaporation <strong>of</strong> a pure solvent <strong>droplet</strong> (Equation<br />

2.16) is now modified to include the influence <strong>of</strong> inner acoustic streaming to mass transfer<br />

via the Sherwood number [Tuckermann et al. 2002]:<br />

Equation 3.8<br />

2· · <br />

·<br />

· <br />

<br />

<br />

<br />

· 2<br />

The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>of</strong> the levitated <strong>droplet</strong> can be calculated using the<br />

model described by Yarin et al. [1999]:<br />

Equation 3.9<br />

⁄<br />

<br />

<br />

· ⁄<br />

<br />

· · <br />

·<br />

· <br />

Equation 3.10<br />

1 2 · ⁄<br />

<br />

<br />

· · · <br />

·<br />

· ·<br />

<br />

<br />

is the diffusivity <strong>of</strong> the ambient gas ( 0.208 cm 2·s -1 for air), is the thermal<br />

conductivity <strong>of</strong> the ambient gas, is the latent heat <strong>of</strong> evaporation, is the relative<br />

humidity <strong>and</strong> <strong>and</strong> are the vapour pressure <strong>and</strong> <strong>temperature</strong> <strong>of</strong> the gas. The properties<br />

<strong>of</strong> the substance are dependent on the <strong>surface</strong> <strong>temperature</strong>, so that the equation has to be<br />

solved iteratively. For the evaporation <strong>of</strong> water <strong>droplet</strong>s <strong>and</strong> . The


Acoustic levitation 33<br />

thermal conductivity <strong>of</strong> air in cal·cm -1·°C -1·s -1 with in degrees Celsius is calculated<br />

by [Seaver et al. 1989]:<br />

Equation 3.11<br />

5.8·10 1.6·10 · <br />

For water vapour at atmospheric pressure the saturation vapour pressure in mbar is<br />

given by [Seaver et al. 1989]:<br />

Equation 3.12<br />

· · <br />

· · · · <br />

is taken in degrees Celsius <strong>and</strong> the values for the constants are given in Table<br />

3.1.<br />

Table 3.1: Parameters for the calculation <strong>of</strong> the saturation vapour pressure <strong>of</strong> water<br />

[Seaver et al. 1989]<br />

6.107799961<br />

4.436518521∙10 -1<br />

1.428945805∙10 -2<br />

2.650648731∙10 -4<br />

3.031240396∙10 -6<br />

2.034080948∙10 -8<br />

6.136820929∙10 -11<br />

The diffusion coefficient for water in cm 2·s -1 <strong>and</strong> the latent heat <strong>of</strong> evaporation for<br />

water in cal·g -1 for water vapour at atmospheric pressure are calculated by [Seaver et<br />

al. 1989]:<br />

Equation 3.13 0.211 · 273.15<br />

273.15 .


Acoustic levitation 34<br />

Equation 3.14<br />

595.3 · 0.483 ·1.2·10 <br />

For solution <strong>droplet</strong>s other than water the saturation vapour pressure in bar is<br />

calculated using the Antoine equation [Reid et al. 1987]:<br />

Equation 3.15 0.001333224 · <br />

<br />

Here is taken in degrees Kelvin <strong>and</strong> some values for the liquid parameters A-C are given<br />

in Table 3.2.<br />

Table 3.2: Liquid parameters for calculation <strong>of</strong> the saturation pressure <strong>of</strong> the vapour<br />

[Reid et al. 1987]<br />

Liquid A B C<br />

Acetone 16.6513 2940.46 -35.93<br />

2-Butanone 16.5986 3150.42 -36.65<br />

Dichloromethane 16.3029 2622.44 -41.70<br />

Ethanol 18.9919 3803.98 -41.68<br />

Ethyl acetate 16.1516 2790.50 -57.15<br />

Methanol 18.5875 3626.55 -34.29<br />

2-Propanol 18.6929 3640.20 -53.54<br />

Tetrahydr<strong>of</strong>uran 16.1069 2768.38 -46.90<br />

For solvent <strong>droplet</strong>s other than water the binary gas diffusion coefficient has to be<br />

calculated using the method by Fuller described in 2.3.2. The values for the latent heat <strong>of</strong><br />

evaporation can be found for each solvent in tables, where the values for are usually<br />

given for 25 °C or for the boiling point <strong>of</strong> the substance. An online-h<strong>and</strong>book containing<br />

values for numerous organic substances is “Yaws’ H<strong>and</strong>book <strong>of</strong> Thermodynamic <strong>and</strong><br />

Physical Properties <strong>of</strong> Chemical Compounds” [Yaws 2003]. To calculate in kJ·mol -1 for<br />

a special <strong>temperature</strong> in Kelvin the conversion is given by:


Acoustic levitation 35<br />

Equation 3.16<br />

· 1 <br />

<br />

<br />

For each substance the values for the coefficients , , <strong>and</strong> the molar mass for<br />

conversion <strong>of</strong> to kJ·kg -1 are also given in Yaws [2003]. For the calculation <strong>of</strong> <strong>droplet</strong>s<br />

containing binary solvent mixtures see Yarin et al. [2002].<br />

3.7.2 Solution <strong>and</strong> suspension <strong>droplet</strong>s<br />

The <strong>drying</strong> process <strong>of</strong> solution <strong>and</strong> suspension <strong>droplet</strong>s also has to be divided into two<br />

parts. In the first <strong>drying</strong> stage (constant rate) the <strong>droplet</strong>’s behaviour is like for pure solvent<br />

<strong>droplet</strong>s <strong>and</strong> can be calculated using the equations in 3.7.1. The vapour pressure lowering<br />

effect <strong>of</strong> the dissolved solids has to be taken into account. The mass flux in the first<br />

<strong>drying</strong> stage can be calculated directly from the volume decrease ∆ per time ∆ <strong>and</strong> the<br />

density <strong>of</strong> the solvent liquid [Kastner 2001]:<br />

Equation 3.17<br />

· ∆<br />

∆<br />

With the assumption <strong>of</strong> a volume decrease the negative algebraic sign leads to positive<br />

values for the mass flux. After the critical point the <strong>droplet</strong> volume remains constant, but<br />

mass <strong>and</strong> density <strong>of</strong> the levitated <strong>droplet</strong> decrease due to further evaporation through the<br />

crust from the liquid centre. The <strong>droplet</strong> rises in the ultrasonic field, <strong>and</strong> by using the<br />

distance to the pressure node the mass flux can be calculated [Kastner 2001]:<br />

Equation 3.18<br />

<br />

∆<br />

· sin2 · ·∆ sin2 · ·∆ <br />

The wave number is ⁄ <strong>and</strong> the distance from the mass centre <strong>of</strong> the <strong>droplet</strong> to its<br />

next upper pressure node is ∆ . The constant for the intensity <strong>of</strong> the ultrasonic field, ,<br />

is given by [Kastner 2001]:<br />

Equation 3.19<br />

·<br />

· <br />

· · Ω<br />

·


Acoustic levitation 36<br />

The function Ω with Ω · is given by [Kastner 2001]:<br />

Equation 3.20<br />

Ω 1<br />

Ω · <br />

<br />

· Ω <br />

<br />

2<br />

Ω · <br />

<br />

·Ω 3·1 <br />

<br />

<br />

<br />

<br />

1 1 ·<br />

Ω · <br />

<br />

<br />

<br />

<br />

· Ω · 2<br />

<br />

using for , <strong>and</strong> the spherical Bessel- <strong>and</strong> Neumann-functions Ω <strong>and</strong> Ω:<br />

Equation 3.21<br />

Ω Ω · Ω <br />

<br />

Ω · Ω<br />

Ω Ω · Ω <br />

<br />

Ω · Ω<br />

Ω Ω Ω ⁄<br />

<br />

For values at Ω 1 Lierke [1996a] gives a ready approximation:<br />

Equation 3.22 Ω 5 6 ·Ω· 3<br />

2Ω · sin2Ω 2Ω<br />

cos2Ω<br />

The effective pressure amplitude <strong>of</strong> the incident acoustic field is defined by Yarin et<br />

al. [1998] in relation to the pressure amplitude at the source <strong>surface</strong> using the distance<br />

<strong>of</strong> the reflector from the transducer , the angular frequency <strong>and</strong> the sound velocity :<br />

Equation 3.23<br />

<br />

<br />

cos · ⁄ <br />

The sound pressure level SPL is defined using the effective amplitude in dyn·cm -2 by<br />

Yarin et al. [1998] <strong>and</strong> for a given SPL can be calculated using this equation:<br />

Equation 3.24<br />

20 · log 74<br />

The SPL can be calculated using the aspect ratio ( ⁄ ) <strong>of</strong> the levitated <strong>droplet</strong> in a<br />

numerical solution according to Yarin et al. [1998], because <strong>droplet</strong> deformation is<br />

dependent on the SPL.


Materials <strong>and</strong> Methods 37<br />

4 Materials <strong>and</strong> Methods<br />

4.1 Materials<br />

4.1.1 Proteins<br />

Table 4.1 gives a summary <strong>of</strong> the proteins used in this work. Carbonic anhydrase from<br />

bovine erythrocytes, L-lactic dehydrogenase type II from rabbit muscle <strong>and</strong> trypsinogen<br />

from bovine pancreas were used as model proteins in the levitation experiments. Trypsin<br />

from bovine pancreas was used as a reagent in the enzyme assay <strong>of</strong> trypsinogen.<br />

Table 4.1: Enzymes with lot number <strong>and</strong> supplier used in this work<br />

Substance Lot Number Supplied by (Product Number)<br />

Carbonic anhydrase from bovine<br />

erythrocytes<br />

L-Lactic dehydrogenase solution<br />

Type II from rabbit muscle<br />

057K1277<br />

127K1564<br />

028K1625<br />

086K2026<br />

107K7405<br />

036K7400<br />

Sigma-Aldrich, Germany<br />

(C 3934)<br />

Sigma-Aldrich, Germany<br />

(L 2500)<br />

Trypsin from bovine pancreas 097K7025 Sigma-Aldrich, Germany<br />

(T 9201)<br />

Trypsinogen from bovine pancreas 077K7004 Sigma-Aldrich, Germany<br />

(T 1143)<br />

4.1.1.1 Carbonic anhydrase<br />

Carbonic anhydrase facilitates the hydration <strong>of</strong> carbon dioxide to hydrogen carbonate <strong>and</strong> a<br />

proton.<br />

CO 2<br />

+ H 2<br />

O<br />

carbonic anhydrase<br />

⎯⎯⎯⎯⎯⎯⎯⎯→ HCO - 3 + H+


Materials <strong>and</strong> Methods 38<br />

It is found in all animals <strong>and</strong> photosynthesizing organisms examined, as well as in some<br />

nonphotosynthetic bacteria. Carbon dioxide is the starting point for photosynthesis <strong>and</strong> the<br />

end point <strong>of</strong> substrate oxidation, therefore carbonic anhydrase is an ubiquitous enzyme<br />

[Cox <strong>and</strong> Phillips 2008]. Carbonic anhydrase is one <strong>of</strong> the fastest <strong>of</strong> all known enzymes.<br />

One enzyme is able to hydrolyse 10 6 molecules <strong>of</strong> CO 2 per second [Berg et al. 2007].<br />

Carbonic anhydrase can be found in different locations. In erythrocytes it is included in the<br />

transport <strong>of</strong> carbonic dioxide, in the kidneys in the reabsorption <strong>of</strong> hydrogen carbonate, in<br />

the stomach in the secretion <strong>of</strong> gastric acid, <strong>and</strong> in the eye in the production <strong>of</strong> aqueous<br />

humor. Carbonic anhydrase is a mostly spherical metalloenzyme containing zinc in the<br />

active site located in a cavity that almost reaches the centre <strong>of</strong> the molecule. The zinc ion is<br />

coordinated by three nitrogen atoms from histidin components in a tetrahedral geometry<br />

completed by water (or hydroxide) [Cox <strong>and</strong> Phillips 2008]. There are more than 120<br />

single-crystal X-ray structures <strong>of</strong> the native enzyme, its mutants, isozymes <strong>and</strong> various<br />

metal-substituted forms [Cronin et al. 2001]. Figure 4.1 shows the structure <strong>of</strong> human<br />

carbonic anhydrase.<br />

Figure 4.1: Structure <strong>of</strong> human carbonic anhydrase II <strong>and</strong> the active site with zinc<br />

according to [Berg et al. 2007; PDB 2009]<br />

In the reaction mechanism <strong>of</strong> carbonic anhydrase the enzyme uses the zinc-bond hydroxide<br />

ion to hydratase carbon dioxide as shown in Figure 4.2. First the water molecule is<br />

deprotonated (1) <strong>and</strong> carbon dioxide is bonded (2). Then a nucleophilic attack <strong>of</strong> the<br />

hydroxide ion to the carbon dioxide takes place <strong>and</strong> forms a hydrogen carbonate ion (3)<br />

that is set free <strong>and</strong> replaced by a new water molecule (4) [Berg et al. 2007]. In this thesis


Materials <strong>and</strong> Methods 39<br />

carbonic anhydrase from bovine erythrocytes (bCA) is used. The product is a mixture <strong>of</strong><br />

is<strong>of</strong>orms with the typically used molecular weight <strong>of</strong> 30 kDa [Sigma-Aldrich 2009].<br />

Figure 4.2: Reaction mechanism <strong>of</strong> carbonic anhydrase [Berg et al. 2007]<br />

4.1.1.2 L-Lactic dehydrogenase<br />

Lactic dehydrogenase is an enzyme in glycolysis that catalyzes the reduction <strong>of</strong> pyruvate to<br />

lactate. It is found in nearly all kinds <strong>of</strong> cells. Coenzyme is NADH+H + , which delivers the<br />

hydrogen for the reaction. L-Lactic dehydrogenase is specific for L-lactate.<br />

Pyruvate + NADH + H +<br />

lactic dehydrogenase<br />

⎯⎯⎯⎯⎯⎯⎯⎯→ lactate + NAD +<br />

Lactic dehydrogenase is composed <strong>of</strong> four subunits with a total molecular weight <strong>of</strong><br />

140 kDa. The subunits are either H- or M-subunits, which are similar in molecular weight,<br />

but differ in their amino acid composition [Sigma-Aldrich 2009]. The human polypeptide<br />

chains <strong>of</strong> the subunits differ by 25 % [Berg et al. 2007]. With this two types <strong>of</strong> subunits<br />

five isoenzymes, which are predominantly found in different organs, are built: H 4 (heart),<br />

MH 3 , M 2 H 2 , M 3 H <strong>and</strong> M 4 (skeletal muscles). Figure 4.3 shows the active site <strong>of</strong> the protein<br />

<strong>and</strong> the substrate <strong>and</strong> coenzyme binding. The reaction is stereospecific for the C4-<br />

hydrogens <strong>of</strong> NADH due to the positioning <strong>of</strong> the molecules [Cox <strong>and</strong> Phillips 2008]. In<br />

this thesis L-lactic dehydrogenase Type II from rabbit muscle (LDH) as crystalline<br />

suspension in 3.2 M (NH 4 ) 2 SO 4 solution pH 6.0 is used.


Materials <strong>and</strong> Methods 40<br />

Figure 4.3: The lactate dehydrogenase reaction [Cox <strong>and</strong> Phillips 2008]<br />

4.1.1.3 Trypsinogen <strong>and</strong> Trypsin<br />

Trypsinogen is an enzyme located in the pancreatic juice. It is the proenzyme <strong>of</strong> the<br />

pancreatic enzyme trypsin, which is a member <strong>of</strong> the serine protease family. The activation<br />

takes place after the proenzyme reaches the lumen <strong>of</strong> the small intestine by<br />

enteropeptidase. A hexapeptide is splitted <strong>of</strong>f the NH 2 -terminus, which leads to a change in<br />

conformation to reduce the distance from side chains in the active site <strong>of</strong> the protein. This<br />

way <strong>of</strong> using proenzymes prevents the pancreas from self-digestion. Trypsin itself is also<br />

able to activate trypsinogen to trypsin by autocatalysis. The physiological relevance <strong>of</strong><br />

trypsin is to cut peptide bonds by hydrolysis for the decomposition <strong>of</strong> amino acids in<br />

digestion [Karlson 1988; Sigma-Aldrich 2009]. Trypsinogen from bovine pancreas is used<br />

in this thesis. Bovine trypsinogen consists <strong>of</strong> a single polypeptide chain <strong>of</strong> 229 amino acids<br />

<strong>and</strong> is cross linked by six disulfide bridges (Figure 4.4). It has a molecular weight <strong>of</strong><br />

23.7 kDa [Karlson 1988; Sigma-Aldrich 2009].


Materials <strong>and</strong> Methods 41<br />

Figure 4.4: Structure <strong>of</strong> bovine trypsinogen [PDB 2009]<br />

4.1.2 Itraconazole<br />

Itraconazole (C 35 H 38 Cl 2 N 8 O 4 ) is a synthetic broad-spectrum triazole antifungal agent. It<br />

inhibits the synthesis <strong>of</strong> ergosterol, a vital cell membrane component in fungi. The molar<br />

mass is 705.63 g/mol [Sigma-Aldrich 2009]. The itraconazole used in this work was<br />

supplied by Janssen Pharmaceutica, Belgium (Lot Nr. 60000142). The chemical structure<br />

<strong>of</strong> itraconazole is given in Figure 4.5 <strong>and</strong> shows that the molecule is highly lipophilic.<br />

Itraconazole is applicable for oral <strong>and</strong> parenteral use [Mutschler et al. 2001].<br />

Figure 4.5: Chemical structure <strong>of</strong> itraconazole [Mutschler et al. 2001]


Materials <strong>and</strong> Methods 42<br />

4.1.3 Excipients <strong>and</strong> reagents<br />

The following tables give a summary <strong>of</strong> all substances used in this work as solvents (Table<br />

4.2), sugars (Table 4.3) or reagents (Table 4.4). The water for buffers <strong>and</strong> reagent solutions<br />

was prepared by double distillation in an all-glass apparatus <strong>and</strong> filtration using filter<br />

membranes with 0.2 μm pores.<br />

Table 4.2: Solvents with lot number <strong>and</strong> supplier used in this work<br />

Substance Lot Number Supplied by (Product Number)<br />

Acetone pure 03676547 Roth, Germany (CP40.1)<br />

2-Butanone p.a. 031K14530709 Merck, Germany (9708)<br />

Dichloromethane pure 15678452 Roth, Germany (CP45.1)<br />

Ethanol pure 19679127 Roth, Germany (P075.1)<br />

Ethyl acetate pure 03676549 Roth, Germany (CP42.1)<br />

Methanol pure 03676545 Roth, Germany (CP43.1)<br />

2-Propanol pure 39681217 Roth, Germany (CP41.1)<br />

Tetrahydr<strong>of</strong>uran p.a. 210K17393531 Merck, Germany (9731)<br />

Water p.a.<br />

OC683183<br />

HC803830<br />

Merck, Germany (1.16754.5000)<br />

Table 4.3: Sugars with lot number <strong>and</strong> supplier used in this work<br />

Substance Lot Number Supplied by (Product Number)<br />

D-Mannitol, SigmaUltra 066K00411 Sigma-Aldrich, Germany (M 9546)<br />

Sucrose (min. 99.5 %) 096K0026 Sigma-Aldrich, Germany (S 9378)<br />

D-(+)-Trehalose dihydrate<br />

124K3793<br />

018K3791-1<br />

Sigma-Aldrich, Germany (T 5251)


Materials <strong>and</strong> Methods 43<br />

Table 4.4: Reagents with lot number <strong>and</strong> supplier used in this work<br />

Substance Lot Number Supplied by (Product Number)<br />

N α -Benzoyl-L-arginine<br />

ethyl ester hydrochloride<br />

035K0767<br />

033K0991<br />

Sigma-Aldrich, Germany (B 4500)<br />

Calcium chloride dihydrate 31K1251 Sigma-Aldrich, Germany (C 3881)<br />

Copolyvidone<br />

(Kollidon® VA 64)<br />

59933416KO BASF, Germany (-)<br />

Hydrochloric acid, conc. 15786449 Roth, Germany (4625.1)<br />

Hydroxypropylmethylcellulose<br />

(Pharmacoat 615)<br />

20022509 Dow Chemicals, USA (-)<br />

β - Nicotinamide adenine<br />

dinucleotide, reduced<br />

disodium salt hydrate<br />

126K7025<br />

097K7016<br />

Sigma-Aldrich, Germany (N8129)<br />

p-Nitrophenyl acetate 1329140<br />

35007229<br />

Sigma-Aldrich, Germany (N 8130)<br />

Phosphoric acid, conc. K23985682 724 Merck, Germany (1.59382.0050)<br />

Potassium di-hydrogen<br />

phosphate<br />

di-Potassium hydrogen<br />

phosphate anhydrous<br />

Potassium hydroxide<br />

solution 1 M<br />

18887607 Roth, Germany (P018.1)<br />

07358012 Roth, Germany (T875.2)<br />

20420 Sigma-Aldrich, Germany (35113)<br />

Sodium pyruvate 117K0662 Sigma-Aldrich, Germany (P 2256)<br />

Trizma base 064K5405 Sigma-Aldrich, Germany (T 1503)<br />

Trizma HCl<br />

(reagent grade)<br />

086K5461<br />

057K5418<br />

Sigma-Aldrich, Germany (T 3253)


Materials <strong>and</strong> Methods 44<br />

4.2 Methods<br />

4.2.1 Acoustic levitation<br />

A 58 kHz levitator (tec5 AG, Oberursel, Germany) was used. According to tec5 AG the<br />

wavelength <strong>of</strong> the st<strong>and</strong>ing acoustic wave is 5.71 mm <strong>and</strong> the high frequency (HF)-power<br />

(continuously variable) can be adjusted from 0.65 to 5 Watt. The particle diameter range<br />

for levitation experiments is about 15 μm to 2.5 mm. The levitator can be used in a<br />

<strong>temperature</strong> range from 0 - 70 °C <strong>and</strong> a relative humidity (not condensating) <strong>of</strong> 10 - 90 %.<br />

In Figure 4.6 the transducer <strong>and</strong> reflector <strong>of</strong> the levitation system is shown. The transducer<br />

with the piezo-crystal is located at the top <strong>and</strong> connected to a power supply unit. The metal<br />

reflector at the base is replaceable. A<br />

planar reflector is an integral part <strong>of</strong><br />

the system on which the concave<br />

reflector can be placed as shown in<br />

the picture. All experiments in this<br />

Transducer<br />

work were performed using the<br />

concave reflector. The distance from<br />

transducer <strong>and</strong> reflector can be<br />

adjusted by a micrometer screw to a<br />

distance <strong>of</strong> a multiple <strong>of</strong> halfwavelengths.<br />

By two additional<br />

Reflector<br />

micrometer screws the horizontal<br />

position <strong>of</strong> the reflector can be<br />

adjusted. The hole in the reflector can<br />

be used for a <strong>drying</strong> air stream that is<br />

directed to the <strong>droplet</strong>. The<br />

dimensions <strong>of</strong> the transducer <strong>and</strong> the Figure 4.6: Transducer <strong>and</strong> reflector<br />

reflector are given in Figure 4.7.


Materials <strong>and</strong> Methods 45<br />

40.0 mm<br />

16.0 mm<br />

Transducer<br />

15.0 mm<br />

12.0 mm<br />

0.8 mm<br />

6.1 mm<br />

2.0 mm<br />

Concave<br />

reflector<br />

5.0 mm<br />

9.0 mm<br />

26.2 mm<br />

3.8 mm<br />

11.2 mm<br />

Planar<br />

reflector<br />

8.0 mm<br />

25.0 mm<br />

Figure 4.7: Dimensions <strong>of</strong> the transducer <strong>and</strong> the concave <strong>and</strong> planar (fixed) reflector


Materials <strong>and</strong> Methods 46<br />

The transducer <strong>and</strong> the reflector are<br />

located in an acrylic glass chamber.<br />

The chamber has a sample input<br />

window at the front that can be<br />

closed during measurement (Figure<br />

4.8 Nr. 1). Liquid samples are<br />

brought in the pressure node by a<br />

Hamilton 7105N 5 μl microliter<br />

syringe (Hamilton Company Europe,<br />

Bonaduz, Switzerl<strong>and</strong>) or by a<br />

micro-dispensing system described<br />

in 4.2.2. Solid samples are placed in<br />

by tweezers or spoon net. The spoon<br />

net is also used to remove the dried<br />

3<br />

particles out <strong>of</strong> the acoustic field. If<br />

liquid samples have to be removed,<br />

they are caught in 0.2 ml polymerase Figure 4.8: Inner acrylic glass box with<br />

levitator<br />

chain reaction (PCR) tubes<br />

(Eppendorf AG, Hamburg, Germany) where they can be dissolved for further analysis<br />

(Figure 4.9).<br />

1<br />

2<br />

Figure 4.9: Spoon net <strong>and</strong> PCR tube holder<br />

A Testo 605-H1 dew point hygrometer (Testo AG, Lenzkirch, Germany) is built in the<br />

chamber for humidity <strong>and</strong> <strong>temperature</strong> measurements (Figure 4.8 Nr. 2). A flexible tube<br />

(Figure 4.8 Nr. 3) connects the levitation chamber with the controlled evaporation mixer<br />

(CEM). The CEM-system consists <strong>of</strong> a flow meter for the liquid (L1-FAC-33-0, flow<br />

range 0.2 - 10.0 g/h H 2 O), a flow controller for the <strong>drying</strong> gas (F201C-FAC-33-V, flow


Materials <strong>and</strong> Methods 47<br />

range 0.04 - 2.0 ln/min N 2 ) <strong>and</strong> a <strong>temperature</strong> controlled evaporation unit (W202-330-P,<br />

range up to 200 °C) (Bronkhorst, Ruurlo, the Netherl<strong>and</strong>s, supplied by Wagner Mess- und<br />

Regeltechnik, Offenbach / Main, Germany). The feeding pressure was 1 bar <strong>and</strong> the system<br />

was used with water <strong>and</strong> dry air or nitrogen. The inner acrylic glass chamber is contained<br />

in an outer acrylic glass chamber which contains a Leister air heater Type 8D1 3000W<br />

(Leister, Kägiswill, Switzerl<strong>and</strong>) <strong>and</strong> has a front plate with two holes. The front plate has<br />

to be changed if the microdispensing system is used. The levitation system was<br />

equilibrated for 1 h (<strong>temperature</strong> <strong>and</strong> relative humidity) before starting an experiment. The<br />

conditions during the experiments were constantly monitored by the Testo dew point<br />

hygrometer. The setup <strong>of</strong> the levitation system is given in Figure 4.10 (without<br />

microdispensing system).<br />

Germanium<br />

window<br />

IR-Camera<br />

with<br />

macrolens<br />

External<br />

heating<br />

Controlled<br />

evaporation mixer<br />

CCD-Camera with<br />

macrolens<br />

Light source<br />

Transducer /<br />

reflector<br />

Inner <strong>and</strong> outer acrylic<br />

glass chamber<br />

Sample input<br />

Figure 4.10: Setup <strong>of</strong> the levitation system<br />

Figure 4.11 shows a picture <strong>of</strong> the complete setup <strong>of</strong> the levitation system including<br />

computer, control units <strong>and</strong> water supply: 1) CEM (three units), 2) charge coupled device<br />

(CCD) camera (left side <strong>of</strong> the chamber) <strong>and</strong> infrared (IR) camera (back side <strong>of</strong> the<br />

chamber), 3) light source, 4) inner <strong>and</strong> outer acrylic glass chamber with levitator, 5)<br />

external heating, 6) computer <strong>and</strong> control units for the levitator <strong>and</strong> CEM, 7) water supply.


Materials <strong>and</strong> Methods 48<br />

2<br />

1<br />

4<br />

3<br />

7<br />

6<br />

5<br />

Figure 4.11: Picture <strong>of</strong> the levitation setup<br />

4.2.2 Microdispensing system<br />

For contactless dispensing <strong>of</strong> very small <strong>droplet</strong><br />

sizes into the pressure nodes <strong>of</strong> the levitator a<br />

microdispensing system was used (microdrop<br />

Technologies GmbH, Norderstedt, Germany). A<br />

dispenser head (MD-K-130-012) was used with an<br />

inner nozzle diameter <strong>of</strong> 70 μm corresponding to a<br />

<strong>droplet</strong> volume <strong>of</strong> about 180 pl (Figure 4.12),<br />

dependent on the fluid, together with a control unit<br />

(MD-E-201-H). The dispenser head is suitable for<br />

liquids with a viscosity up to about 15 mPa·s. The<br />

<strong>droplet</strong>s have a velocity <strong>of</strong> about 2 m/s <strong>and</strong> the<br />

maximum <strong>droplet</strong> rate is 2000 Hz (depending on<br />

the fluid used).<br />

The dispenser head is based on piezodriven<br />

inkjet printing technology. The integrated<br />

piezo activator induces a shock-wave into the fluid<br />

contained in the head which causes a <strong>droplet</strong> to be<br />

emitted from the nozzle [microdrop 2009]. A<br />

series <strong>of</strong> <strong>droplet</strong>s is repeatedly shot into the<br />

Figure 4.12: Microdispensing<br />

system with rail (front view)


Materials <strong>and</strong> Methods 49<br />

pressure node <strong>of</strong> the st<strong>and</strong>ing acoustic wave <strong>of</strong> the levitator, so that the initial volume <strong>of</strong><br />

the levitated <strong>droplet</strong> is achieved by the number <strong>of</strong> single <strong>droplet</strong>s.<br />

The dispenser head has to be brought out <strong>of</strong> the levitation chamber, especially at<br />

experiments at high <strong>temperature</strong>, because <strong>of</strong> the <strong>temperature</strong> influence on the sample in the<br />

reservoir. For this reason a rail was constructed, on which the dispenser head <strong>and</strong> the<br />

reservoir are fixed on a mobile unit (Figure 4.12). The position <strong>of</strong> the rail can be adjusted<br />

in height <strong>and</strong> gradient to optimum position to the pressure node by screws fixed in the<br />

perpendicular bars (Figure 4.13). The total length <strong>of</strong> the rail minus stopper is 58 cm. The<br />

rail is fixed on a wooden plate that can be adjusted via four screws.<br />

Figure 4.13: Microdispensing system with rail (side view)<br />

For the fixation <strong>of</strong> the dispenser head the st<strong>and</strong>ard holder (MD-H-711-07) had to be<br />

modified (Figure 4.14) so that it can be moved through the small window into the inner<br />

levitation chamber close to the pressure node. The reservoir is outside the inner acrylic<br />

glass chamber while dispensing.<br />

Figure 4.15 shows the levitation setup for experiments using the microdispensing<br />

system. The rail is fixed directly on the front side <strong>of</strong> the outer acrylic glass box <strong>and</strong> an<br />

alternative front plate was used. There is one rectangular hole for the microdispensing head<br />

<strong>and</strong> one round hole for a h<strong>and</strong>. By several screws the rail can be adjusted for the best<br />

dispensing position.


Materials <strong>and</strong> Methods 50<br />

C<br />

A<br />

B<br />

Figure 4.14: Dispenser head holder: (A) dispenser head, (B) reservoir <strong>and</strong> (c) air filter<br />

Figure 4.15: Setup <strong>of</strong> the levitation system containing the microdispensing system


Materials <strong>and</strong> Methods 51<br />

4.2.3 Droplet size measurements<br />

For the <strong>droplet</strong> size measurements a JAI CV-M4 2/3’’ monochrome CCD-camera (JAI-<br />

AG, Copenhagen, Denmark) was used together with a Nikon Micro 60 mm objective 2.8<br />

diaphragm <strong>and</strong> a bellow. The CCD-camera is connected to a computer by a PcDIG LVDS<br />

frame grabber card. The pictures were taken using back-light illumination by a Schott KL<br />

1500 cold light source <strong>and</strong> recorded <strong>and</strong> analyzed by Image Pro Plus S<strong>of</strong>tware 4.51<br />

(MediaCybernetics, Bethesda, USA). The complete imaging system was supplied by Weiss<br />

Imaging <strong>and</strong> Solutions GmbH, Günding, Germany. The camera system was calibrated for<br />

each setting <strong>of</strong> the macro lens <strong>and</strong> bellows using a micrometer scale on a microscope slide<br />

that was placed in the levitator. For verification <strong>of</strong> the calibration a polypropylene sphere<br />

(Spherotech, Fulda, Germany) was levitated <strong>and</strong> the diameter was measured <strong>and</strong> compared<br />

to the given value <strong>of</strong> 1.85 ± 0.03 mm.<br />

During <strong>droplet</strong> <strong>drying</strong> pictures were taken in different intervals depending on the<br />

evaporation rate <strong>of</strong> the sample. The horizontal <strong>and</strong> vertical diameter <strong>and</strong> the aspect ratio<br />

were measured together with the position <strong>of</strong> the centre <strong>of</strong> mass directly in the picture <strong>and</strong><br />

saved in a data collector.<br />

4.2.4 Droplet <strong>surface</strong> <strong>temperature</strong> measurements<br />

Droplet <strong>surface</strong> <strong>temperature</strong> measurements were performed using the VARIOSCAN 3021<br />

ST high resolution camera (InfraTec, Dresden, Germany). The camera has a single<br />

mercury cadmium tellurium (MCT) detector for scanning measurements in the <strong>temperature</strong><br />

range from -40 to 1200 °C. A Stirling cooler is used as a cooling system for the detector.<br />

The camera has a thermal resolution <strong>of</strong> +/- 0.03 K <strong>and</strong> a spectral sensitivity in the range<br />

from 8 to 12 μm. The geometric resolution is 1.5 mrad <strong>and</strong> the field <strong>of</strong> view is 30° (H) x<br />

20° (V). Up to five infrared pictures can be taken per second dependent on the chosen<br />

camera settings with a sixfold electro optical zoom. The camera was used together with a<br />

trifold 25 μm macro lens [InfraTec 2009]. Since IR-measurements cannot be performed<br />

through acrylic glass, a round germanium window (100 mm x 5 mm) is located in the back<br />

wall <strong>of</strong> the acrylic glass box. From inside the levitation chamber the germanium window is<br />

covered by a thin plastic film for protection from the sample solvents. The IR-camera is<br />

connected to a computer via Ethernet <strong>and</strong> IRBIS®control s<strong>of</strong>tware (InfraTec, Dresden,<br />

Germany). For data analysis the IRBIS®pr<strong>of</strong>essional 2.2 s<strong>of</strong>tware (InfraTec, Dresden,


Materials <strong>and</strong> Methods 52<br />

Germany) was used. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> was measured at different time<br />

intervals during the <strong>drying</strong> process. The emissivity for the <strong>droplet</strong> substances was analyzed<br />

by comparison <strong>of</strong> the measured <strong>temperature</strong> by the IR-camera to the <strong>temperature</strong> <strong>of</strong> an<br />

Omega Type T thermocouple (OMEGA Engineering, Stamford, USA) with 0.13 mm<br />

diameter <strong>of</strong> the wire, on which the sample <strong>droplet</strong> was suspended according to<br />

Tuckermann [2002]. The <strong>temperature</strong> data were collected by an OM-CP-QUADTEMP<br />

data logger using the s<strong>of</strong>tware OMEGA engineering 2.00 (OMEGA Engineering,<br />

Stamford, USA).<br />

4.2.5 Spray <strong>drying</strong><br />

For spray <strong>drying</strong> experiments a<br />

Büchi Mini Spray Dryer B-191 was<br />

used (Büchi AG, Flawil,<br />

Switzerl<strong>and</strong>) using a high efficiency<br />

cyclone. The sample was collected<br />

directly in a 50 ml Sarstedt tube. The<br />

<strong>drying</strong> <strong>temperature</strong> was higher than<br />

in the levitation experiments, with<br />

130 °C inlet <strong>temperature</strong> <strong>and</strong> 80 °C<br />

outlet <strong>temperature</strong>. The atomization<br />

flow rate was 700 l/h <strong>and</strong> the sample<br />

volume was 0.4 - 5 ml, depending on<br />

the substance, with a liquid feed <strong>of</strong><br />

2 ml/min. The aspirator was set to<br />

Figure 4.16: Sketch <strong>of</strong> a Büchi Mini Spray<br />

90 % <strong>of</strong> the maximum rate <strong>of</strong><br />

Dryer B-191 [Büchi 1999]<br />

60 m³/h. Figure 4.16 shows a Büchi<br />

B-191 Mini Spray Dryer <strong>and</strong> the path <strong>of</strong> the <strong>drying</strong> air: 1) air inlet (air filter is connected<br />

upstream), 2) heating, 3) entry in <strong>drying</strong> chamber, 4) cyclone, in this work replaced by an<br />

improved cyclone for maximum powder recovery [Maury et al. 2005a], 5) aspirator, 6)<br />

<strong>temperature</strong> sensor air inlet, 7) <strong>temperature</strong> sensor air outlet, 8) product container, in this<br />

work replaced by a Sarstedt tube [Büchi 1999].


Materials <strong>and</strong> Methods 53<br />

4.2.6 Scanning electron microscopy<br />

Scanning electron microscopic pictures were taken by an Amray 1810T scanning electron<br />

microscope (Amray, Bedford, USA) for particle size <strong>and</strong> morphology analysis. The<br />

samples were sputtered with gold for 1.5 min (Hummer JR Technics, Munich, Germany)<br />

after fixing on aluminium sample stubs G301 (Plano GmbH, Wetzlar, Germany).<br />

4.2.7 Enzyme activity assay <strong>of</strong> carbonic anhydrase<br />

The enzymatic activity measurement <strong>of</strong> bovine carbonic anhydrase (bCA) is based on<br />

assay procedures by Sigma-Aldrich [2009] <strong>and</strong> by Verpoorte et al. [1967] <strong>and</strong> was used in<br />

a modified form. The principle <strong>of</strong> the assay is the reaction <strong>of</strong> p-nitrophenyl acetate (PNPA)<br />

<strong>and</strong> water to p-nitrophenol <strong>and</strong> acetate catalyzed by bCA:<br />

p - nitrophenyl acetate + H 2<br />

O<br />

bCA<br />

⎯⎯⎯→ p - nitrophenol + acetate + H +<br />

The samples were dissolved in 50 mM trizma buffer pH 7.5 (at 25 °C) <strong>and</strong> stored on ice.<br />

The buffer was prepared using trizma base, trizma hydrochloride <strong>and</strong> double-distilled<br />

water. The pH was adjusted to 7.5 at 25 °C using hydrochloric acid <strong>and</strong> potassium<br />

hydroxide solution. In the case <strong>of</strong> no inactivation, the final activity <strong>of</strong> the sample solution<br />

should be 100 - 200 units/ml. The levitation samples could be dissolved directly in the<br />

PCR tubes using 250 μl trizma buffer. Spray dried samples were dissolved in Sarstedt<br />

tubes. A reference sample using untreated bCA was also analyzed, the untreated sample<br />

solution being used as reference. A <strong>droplet</strong> was brought in the ultrasonic field for <strong>droplet</strong><br />

volume <strong>determination</strong> using the CCD-camera, immediately removed, diluted <strong>and</strong> stored on<br />

ice until measurement. A 3 mM PNPA-solution was prepared freshly daily by dissolving<br />

the PNPA in acetone <strong>and</strong> filling up the volume with double-distilled water. It was stored<br />

protected from light on ice. For the autozero the trizma buffer was used. The increase in<br />

absorption at 348 nm at 25 °C was measured by UV-spectroscopy using a Lambda 25<br />

UV/VIS-spectrometer connected to a PC with PerkinElmer UV WinLab V 5.0 s<strong>of</strong>tware<br />

(PerkinElmer LAS GmbH, Rodgau, Germany) in a 10 mm quartz cuvette. For sample<br />

measurements the cuvette contained 1.90 ml trizma buffer, 1.00 ml PNPA-solution <strong>and</strong><br />

0.10 ml <strong>of</strong> the enzyme solution. Blank measurements were performed using 2.00 ml trizma<br />

buffer <strong>and</strong> 1.00 ml PNPA-solution. The increase in absorption was recorded for 3 minutes<br />

after an initial 3 minutes’ equilibration time in the cuvette holder connected to a water bath


Materials <strong>and</strong> Methods 54<br />

Julabo Type MP-5 (Julabo Labortechnik GmbH, Seelbach, Germany) at 25 °C. The slopes<br />

<strong>of</strong> the sample measurements were compared to the slope <strong>of</strong> the reference sample (set to<br />

100 % activity) <strong>and</strong> the residual activity <strong>of</strong> the sample was thus calculated. All<br />

measurements could be performed only twice due to the small amount <strong>of</strong> sample solution<br />

in the PCR tubes.<br />

4.2.8 Enzyme activity assay <strong>of</strong> L-lactic dehydrogenase<br />

The L-lactic dehydrogenase (LDH) suspension was dialysed using Spectra/Por ® Dialysis<br />

Membrane with a molecular weight cut <strong>of</strong>f (MWCO) <strong>of</strong> 12 - 14,000 (Spectrum<br />

Laboratories, Rancho Domingues, USA) according to Adler [1999]. The dialysis tubing<br />

was placed in a hundredfold volume <strong>of</strong> a 100 mM potassium phosphate buffer pH 7.0 (at<br />

25 °C) on ice on a magnetic stirrer for 3 hours. The potassium phosphate buffer was<br />

prepared using potassium di-hydrogene phosphate <strong>and</strong> double-distilled water. The pH was<br />

adjusted using 1 M potassium hydroxide solution. After renewal <strong>of</strong> the dialysis medium the<br />

dialysis was continued for 14 hours. To achieve a high protein concentration the solution<br />

was concentrated using a centrifugal filter device Amicon Ultra-15 with a MWCO <strong>of</strong><br />

30,000 (Millipore Corporation, Billerica, USA) in the centrifugal rotor Minifuge RF<br />

(Heraeus Sepatech GmbH, Osterode, Germany). The concentration was first measured at<br />

280 nm after equilibration on 25 °C for 1 minute in a 10 mm quartz cuvette <strong>and</strong> then the<br />

sample was diluted to a suitable concentration.<br />

The enzymatic assay for LDH is based on the instruction by Sigma-Aldrich [2009]<br />

<strong>and</strong> the information given by Adler <strong>and</strong> Lee [1999]. LDH catalyzes the oxidation <strong>of</strong><br />

pyruvate <strong>and</strong> β-NADH to L-lactate <strong>and</strong> β-NAD + (see 4.1.1.2). The decrease <strong>of</strong> absorption<br />

results from the decomposition <strong>of</strong> β-NADH. The sample was dissolved in 100 mM<br />

potassium phosphate buffer pH 7.0 (at 25 °C) in PCR tubes or in Sarstedt tubes <strong>and</strong> diluted<br />

to an appropriate activity for the measurement with potassium phosphate buffer. In the case<br />

<strong>of</strong> no inactivation, the final activity <strong>of</strong> the solution should be about 1.3 units/ml. For<br />

autozero the potassium phosphate buffer was used. The decrease in absorption at 340 nm at<br />

25 °C is measured in a 10 mm quartz cuvette by UV-spectroscopy using a Lambda 25<br />

UV/VIS-spectrometer connected to a computer with PerkinElmer UV WinLab V 5.0<br />

s<strong>of</strong>tware (PerkinElmer LAS GmbH, Rodgau, Germany). For sample measurements 2.80 ml<br />

<strong>of</strong> a 0.13 mM β-NADH solution <strong>and</strong> 0.10 ml <strong>of</strong> a 69 mM sodium pyruvate solution were


Materials <strong>and</strong> Methods 55<br />

pipetted into the quartz cuvette. The 0.13 mM β-NADH solution <strong>and</strong> the 69 mM sodium<br />

pyryvate solution were prepared using potassium phosphate buffer as solvent. The cuvette<br />

was first equilibrated in the cuvette holder connected to a water bath (Julabo Labortechnik<br />

GmbH, Seelbach, Germany) at 25 °C for one minute. Then 0.10 ml <strong>of</strong> the sample solution<br />

was added <strong>and</strong> the decrease in absorption at 340 nm was measured for 2 minutes. Blank<br />

measurements were performed by using 0.10 ml potassium phosphate buffer instead <strong>of</strong> the<br />

protein solution. The slopes <strong>of</strong> the sample measurements were compared to the slope <strong>of</strong> a<br />

reference substance (set to 100 % activity, as described in 4.2.7) <strong>and</strong> the residual activity<br />

was calculated.<br />

4.2.9 Enzyme activity assay <strong>of</strong> trypsinogen<br />

In the enzyme activity assay <strong>of</strong> trypsinogen first the trypsinogen is activated to trypsin<br />

before the UV-detectable reaction takes place according to Sigma-Aldrich [2009] <strong>and</strong><br />

Faizyme Laboratories [2007].<br />

Trypsinogen<br />

trypsin<br />

⎯⎯⎯→ trypsin + hexapeptide<br />

Trypsin is able to catalyze the hydrolysis <strong>of</strong> esther <strong>and</strong> amide linkages <strong>of</strong> synthetic<br />

derivates <strong>of</strong> amino acids as in N α -benzoyl-L-arginine ethyl ester hydrochloride (BAEE). It<br />

is decomposed to N α -benzoyl-L-arginine <strong>and</strong> ethanol [Sigma-Aldrich 2009].<br />

BAEE + H 2<br />

O<br />

trypsin<br />

⎯⎯⎯⎯→ N α -benzoyl-L-arginine + ethanol<br />

For activation <strong>of</strong> trypsinogen the sample was dissolved in 1 mM hydrochloric acid.<br />

Levitation samples were dissolved in Eppendorf tubes <strong>and</strong> spray dried samples in Sarstedt<br />

tubes. The trypsinogen solution was added to an activation mixture containing 1 M CaCl 2 -<br />

solution, 400 mM trizma buffer pH 8.4 (at 25 °C) <strong>and</strong> 0.02 % (w/v) trypsin enzyme<br />

solution (1:19:1) <strong>and</strong> incubated at 4 - 8 °C for 24 hours. The CaCl 2 -solution <strong>and</strong> the trizma<br />

buffer were prepared using double-distilled water as a solvent. Trypsin was dissolved in<br />

1 M hydrochloric acid <strong>and</strong> freshly prepared daily. To stop the activation process the<br />

solution was diluted with 1 mM HCl. In the case <strong>of</strong> no inactivation, the final activity <strong>of</strong> the<br />

solution should be 500 units/ml. For autozero a potassium phosphate buffer was used. For<br />

activity measurement 3.00 ml <strong>of</strong> a 0.25 mM BAEE solution, prepared using a 67 mM<br />

potassium phosphate buffer pH 7.6 (at 25 °C) as solvent, <strong>and</strong> 0.20 ml <strong>of</strong> the sample


Materials <strong>and</strong> Methods 56<br />

solution were pipetted in a 10 mm quartz cuvette. The increase in absorption at 253 nm at<br />

25 °C was recorded for 1.5 minutes by using a Lambda 25 UV/VIS- spectrometer<br />

connected to a computer with PerkinElmer UV WinLab V 5.0 s<strong>of</strong>tware (PerkinElmer LAS<br />

GmbH, Rodgau, Germany). The slope <strong>of</strong> the sample measurement was compared to the<br />

slope <strong>of</strong> a reference sample set to 100 % activity (as described in 4.2.7) <strong>and</strong> the residual<br />

activity was calculated.


Results <strong>and</strong> Discussion 57<br />

5 Results <strong>and</strong> Discussion<br />

5.1 Preliminary experiments using the IR-camera<br />

5.1.1 Heating-up <strong>of</strong> the levitation system<br />

The transducer heats up due to its oscillations after switch-on <strong>of</strong> the levitator [Kastner<br />

2001]. In a preliminary experiment the infrared camera was used to depict this behaviour<br />

<strong>and</strong> to give a quantitative impression on the <strong>temperature</strong> increase for the levitation system.<br />

Figure 5.1 shows infrared pictures <strong>of</strong> the levitator (a) before power-on, (b) after 5 min <strong>and</strong><br />

(c) after 10 min from power-on at an ambient <strong>temperature</strong> <strong>of</strong> about 20 °C. The <strong>temperature</strong><br />

scale is set constant in all pictures <strong>and</strong> the increasing red colour shows higher <strong>temperature</strong>.<br />

(a)<br />

(b)<br />

(c)<br />

Figure 5.1: IR-pictures <strong>of</strong> the transducer (a) before power-on, (b) 5 minutes after<br />

power-on, (c) 10 minutes after power-on<br />

The heating-up <strong>of</strong> the transducer unit can be clearly seen. The <strong>temperature</strong> increase is<br />

about 2.6 °C after 10 minutes. The air <strong>temperature</strong> in the inner acrylic glass chamber also<br />

increases. Thus the consequence for the following experiments is that the minimum <strong>drying</strong><br />

air <strong>temperature</strong> must be set to 25 °C. Experiments at 20 °C are therefore not possible using


Results <strong>and</strong> Discussion 58<br />

this levitation set-up, because there is no cooling system <strong>and</strong> lower <strong>temperature</strong>s cannot be<br />

kept constant. For a constant <strong>temperature</strong> <strong>and</strong> relative humidity <strong>of</strong> the <strong>drying</strong> air it is<br />

necessary to equilibrate the chamber for one hour before an experiment is started. During<br />

the experiment the <strong>temperature</strong> <strong>and</strong> relative humidity are controlled continuously by a<br />

thermometer <strong>and</strong> a dew point hygrometer. For measurements at higher <strong>temperature</strong>, the<br />

<strong>temperature</strong> increase has little influence because <strong>of</strong> the higher external heat supply that can<br />

be adjusted to the correct <strong>temperature</strong> measured directly in the inner levitator chamber.<br />

The <strong>temperature</strong> distribution in the st<strong>and</strong>ing ultrasonic field <strong>of</strong> a 20 kHz levitator at<br />

21 °C, 43 % relative humidity <strong>and</strong> a SPL <strong>of</strong> about 168 dB was measured by Tuckermann<br />

[2002]. Tuckermann found that the maximum <strong>temperature</strong> in the levitation axis is located<br />

at the pressure antinodes where the particle velocity is maximal (2 - 3 °C <strong>temperature</strong><br />

increase). But also the other axial parts are heated up by a lower value <strong>of</strong> 1 - 1.5 °C. By use<br />

<strong>of</strong> different settings Tuckermann found that with rising SPL the <strong>temperature</strong> difference<br />

increases. At the maximum SPL <strong>of</strong> about 175 dB the <strong>temperature</strong> increase was about 8 °C<br />

in the pressure antinodes.<br />

5.1.2 Influence <strong>of</strong> <strong>droplet</strong> size on the <strong>surface</strong> <strong>temperature</strong> measurement<br />

Droplet <strong>drying</strong> experiments will be performed using different initial <strong>droplet</strong> sizes. Before<br />

starting the experiments the <strong>surface</strong> <strong>temperature</strong> measurement <strong>of</strong> <strong>droplet</strong>s <strong>and</strong> particles<br />

with different initial sizes was analyzed. At first the <strong>surface</strong> <strong>temperature</strong> <strong>of</strong> plastic spheres<br />

was analyzed at 25 °C <strong>and</strong> 40 °C at 1 % relative humidity using spheres having a diameter<br />

<strong>of</strong> 0.62 mm, 0.98 mm <strong>and</strong> 1.23 mm. Then the <strong>temperature</strong> pr<strong>of</strong>iles <strong>of</strong> water <strong>droplet</strong>s at<br />

three different initial <strong>droplet</strong> sizes <strong>of</strong> about 500 μm, 800 μm, <strong>and</strong> 1200 μm horizontal<br />

diameter are examined at 25 °C, 40 °C, <strong>and</strong> 60 °C at 1 % relative humidity. The<br />

<strong>temperature</strong> pr<strong>of</strong>ile was taken at the maximum diameter <strong>of</strong> the <strong>droplet</strong> or sphere. The size<br />

<strong>of</strong> the <strong>droplet</strong>s or spheres was measured using the CCD-camera.<br />

In comparison to the <strong>temperature</strong> <strong>of</strong> the ambient air the <strong>temperature</strong> pr<strong>of</strong>iles for the<br />

large sphere show a higher <strong>surface</strong> <strong>temperature</strong> for measurements at an ambient<br />

<strong>temperature</strong> <strong>of</strong> 25 °C (Figure 5.2 (a)) <strong>and</strong> lower <strong>surface</strong> <strong>temperature</strong> at 40 °C (Figure<br />

5.2 (b)). The smaller spheres showed similar behaviour. Note that there is no emissivity<br />

correction included for the material <strong>of</strong> these spheres, the focus lies on the pr<strong>of</strong>ile seen by


Results <strong>and</strong> Discussion 59<br />

the IR-camera <strong>and</strong> the area <strong>of</strong> minimum <strong>surface</strong> <strong>temperature</strong>. In the sphere pr<strong>of</strong>iles for<br />

clarity a circle shows the size <strong>of</strong> the sphere.<br />

(a)<br />

Figure 5.2: Surface <strong>temperature</strong> pr<strong>of</strong>ile <strong>of</strong> a levitated sphere (sphere diameter: 1.23<br />

mm) at (a) 25 °C <strong>and</strong> (b) 40 °C ambient <strong>temperature</strong> <strong>and</strong> 1% relative humidity<br />

The heating-up <strong>of</strong> the solid sphere placed in the ultrasonic levitator is shown in Figure<br />

5.3 (a). The maximum heating appears at the poles, the equatorial part is cooler. This effect<br />

was already described by Tuckermann [2002]. Tuckermann found that the heating <strong>of</strong> the<br />

poles is induced by the energy supplied by the ultrasonic field, whereas the cooling <strong>of</strong> the<br />

equatorial part is a consequence <strong>of</strong> the radial streaming air to the levitated sample due to<br />

acoustic convection. This phenomenon is not found for higher air <strong>temperature</strong>s, for<br />

example at 40 °C in Figure 5.3 (b).<br />

(b)<br />

(a)<br />

Figure 5.3: IR-pictures <strong>of</strong> a 1.23 mm plastic sphere at ambient conditions <strong>of</strong> (a) 25 °C<br />

<strong>and</strong> (b) 40 °C <strong>and</strong> 1% relative humidity<br />

(b)


Results <strong>and</strong> Discussion 60<br />

The <strong>temperature</strong> distribution is now overpowered by the high external heat supply from the<br />

ambient air. The <strong>temperature</strong> <strong>of</strong> the sphere appears cooler than the ambience because the<br />

air <strong>temperature</strong> cannot be measured by the IR-camera. This might also explain the decrease<br />

<strong>of</strong> the ambient <strong>temperature</strong> to the boundary <strong>of</strong> the <strong>temperature</strong> pr<strong>of</strong>ile.<br />

For water <strong>droplet</strong>s <strong>of</strong> all three initial sizes there is a decrease in the <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> due to the cooling effect <strong>of</strong> the evaporating solvent (Figure 5.4). This can be<br />

found for all <strong>drying</strong> air <strong>temperature</strong>s. The <strong>temperature</strong> <strong>of</strong> the water <strong>droplet</strong>s is assumed to<br />

be constant over the <strong>droplet</strong> <strong>surface</strong>, but the IR-camera does not show a uniform<br />

<strong>temperature</strong> at the <strong>droplet</strong> boundary to the ambience on the IR-picture. The <strong>temperature</strong> at<br />

the border <strong>of</strong> the <strong>droplet</strong> does not change immediately, but there is rather a transition<br />

region on the <strong>droplet</strong> <strong>surface</strong> where the <strong>temperature</strong> decreases. This effect was also seen<br />

for the plastics spheres at 40 °C.<br />

(a)<br />

(b)<br />

(c)<br />

Figure 5.4: IR-pictures <strong>of</strong> water <strong>droplet</strong>s at 25 °C <strong>and</strong> 1 % relative humidity, <strong>droplet</strong><br />

size: (a) 0.51 x 0.52 mm, (b) 0.77 x 0.97 mm <strong>and</strong> (c) 1.03 x 1.08 mm<br />

The IR-camera cannot be focused on the whole <strong>droplet</strong> <strong>surface</strong> area because <strong>of</strong> the<br />

sphericity <strong>of</strong> the <strong>droplet</strong>, <strong>and</strong> it also cannot be in a perpendicular direction <strong>of</strong> measurement<br />

to the <strong>droplet</strong> <strong>surface</strong> for all positions on the <strong>surface</strong>. Furthermore, the amount <strong>of</strong> measured


Results <strong>and</strong> Discussion 61<br />

background heat radiation, from the levitation chamber walls for example, by the camera<br />

on the <strong>droplet</strong> <strong>surface</strong> is higher at the border in comparison to the middle <strong>of</strong> the <strong>droplet</strong> or<br />

particle. This transition area is expected to have a higher influence on <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> measurement <strong>of</strong> small <strong>droplet</strong>s than for larger <strong>droplet</strong>s. In this case it would<br />

have to be taken into account that for small <strong>droplet</strong>s the measured <strong>surface</strong> <strong>temperature</strong><br />

(even for measurement in the middle <strong>of</strong> the <strong>droplet</strong> <strong>surface</strong>) may lead to higher values than<br />

the true <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> would be.<br />

The comparison <strong>of</strong> the <strong>temperature</strong> pr<strong>of</strong>iles <strong>of</strong> the 500 μm <strong>droplet</strong>s (Figure 5.5)<br />

shows that with increasing <strong>drying</strong> air <strong>temperature</strong> (T da ) the pr<strong>of</strong>ile <strong>of</strong> the <strong>droplet</strong> results in a<br />

smaller constant minimum <strong>temperature</strong> part.<br />

Figure 5.5: Surface <strong>temperature</strong> pr<strong>of</strong>ile <strong>of</strong> levitated water <strong>droplet</strong>s at <strong>droplet</strong><br />

diameters about 0.5 mm at 25 °C (<strong>droplet</strong> size 0.51 x 0.52 mm), 40 °C (<strong>droplet</strong> size<br />

0.57 x 0.58 mm) <strong>and</strong> 60 °C (<strong>droplet</strong> size 0.49 x 0.51 mm) <strong>and</strong> 1 % relative humidity<br />

In the comparison for different <strong>droplet</strong> sizes there is no visible dependence <strong>of</strong> the minimum<br />

<strong>droplet</strong> <strong>temperature</strong> on the <strong>droplet</strong> size (Figure 5.6, Figure 5.7, Figure 5.8). The minimum<br />

<strong>droplet</strong> <strong>temperature</strong>s change due to small differences in the <strong>drying</strong> air conditions.


Results <strong>and</strong> Discussion 62<br />

Figure 5.6: Surface <strong>temperature</strong> pr<strong>of</strong>iles <strong>of</strong> levitated water <strong>droplet</strong>s at 25 °C <strong>and</strong> 1 %<br />

relative humidity (<strong>droplet</strong> size: 0.51 x 0.52 mm, 0.77 x 0.79 mm <strong>and</strong> 1.03 x 1.08 mm)<br />

Figure 5.7: Surface <strong>temperature</strong> pr<strong>of</strong>iles <strong>of</strong> levitated water <strong>droplet</strong>s at 40 °C <strong>and</strong> 1 %<br />

relative humidity (<strong>droplet</strong> size: 0.57 x 0.58 mm, 0.83 x 0.84 mm <strong>and</strong> 1.13 x 1.03 mm)


Results <strong>and</strong> Discussion 63<br />

Figure 5.8: Surface <strong>temperature</strong> pr<strong>of</strong>iles <strong>of</strong> levitated water <strong>droplet</strong>s at 60 °C <strong>and</strong> 1 %<br />

relative humidity (<strong>droplet</strong> size: 0.49 x 0.51 mm, 0.75 x 0.90 mm <strong>and</strong> 1.35 x 1.13 mm)<br />

Note that the <strong>drying</strong> air <strong>temperature</strong> cannot be determined from the pr<strong>of</strong>ile, because the<br />

<strong>temperature</strong> <strong>of</strong> gas is not measureable for the IR-camera. Therefore the <strong>drying</strong> air<br />

<strong>temperature</strong> must be continuously measured by the thermometer in the dew point<br />

hygrometer during the experiments.<br />

There is no indication <strong>of</strong> an influence <strong>of</strong> the initial horizontal <strong>droplet</strong> diameter <strong>of</strong><br />

500, 800 <strong>and</strong> 1200 μm on the <strong>surface</strong> <strong>temperature</strong> measurement. For this reason for all<br />

three initial <strong>droplet</strong> sizes the minimum <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> can be used for<br />

analysis. For very small <strong>droplet</strong> sizes at the end <strong>of</strong> the <strong>drying</strong> process (about 200 - 300 μm)<br />

there is a large <strong>temperature</strong> increase measurable before the <strong>droplet</strong> disappears in the<br />

acoustic field (see data plots in 5.2). The <strong>temperature</strong> increase can be caused by<br />

measurement inaccuracy <strong>and</strong> by non-volatile impurities contained in the solvent or from<br />

the <strong>drying</strong> air [Tuckermann 2002]. For pure solvents the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> at the<br />

beginning <strong>of</strong> the measurement (after the <strong>surface</strong> <strong>temperature</strong> equilibration) is taken for<br />

comparison. For solution <strong>droplet</strong>s the <strong>droplet</strong> has a finite size after the critical point due to<br />

its solid content. Therefore the small size, where the increasing error due to incorrectly<br />

measured <strong>surface</strong> <strong>temperature</strong> has to be taken into account, will not be reached.


Results <strong>and</strong> Discussion 64<br />

5.1.3 Determination <strong>of</strong> emissivities<br />

For the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> measurements the emissivity ε <strong>of</strong> the measured<br />

substance has to be known. Values <strong>of</strong> ε for several substances are given in the literature.<br />

The emissivity has, however, to be determined experimentally because it depends on many<br />

factors like the wavelength range <strong>of</strong> the camera <strong>and</strong> the ambient <strong>temperature</strong>. The<br />

emissivities used for data analysis in this work were determined by measuring the<br />

<strong>temperature</strong> <strong>of</strong> a <strong>droplet</strong> suspended on a thermocouple by using both the IR-camera <strong>and</strong> the<br />

thermocouple. The <strong>temperature</strong> value <strong>of</strong> the thermocouple was used as a reference for the<br />

<strong>temperature</strong> measured by the IR-camera that was adjusted by variation <strong>of</strong> the emissivity. It<br />

is assumed that the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> measured by the IR-camera <strong>and</strong> the mean<br />

<strong>droplet</strong> <strong>temperature</strong> measured by the thermocouple correspond. This assumption is<br />

justified by the small size <strong>and</strong> the rapid internal convection <strong>of</strong> acoustically levitated drops<br />

[Yarin et al. 1999] which lead to rapid <strong>temperature</strong> equilibration [Tuckermann et al. 2005].<br />

The transmissivity <strong>of</strong> the germanium window was assumed to be 0.95 (according to<br />

the transmission spectrum <strong>of</strong> the window). Influences on the <strong>temperature</strong> measurement<br />

inside the levitation chamber were included in the value for the emissivity.<br />

This method was described by Tuckermann [2002]. Tuckermann determined<br />

experimentally emissivities for water <strong>and</strong> several organic solvents <strong>and</strong> found that during<br />

the <strong>drying</strong> process the <strong>temperature</strong> determined by the camera <strong>and</strong> the thermocouple<br />

diverges towards the end <strong>of</strong> measurement. Tuckermann attributed this to decreasing<br />

wetting <strong>of</strong> the thermocouple during the measurement. In this work the emissivity was<br />

determined using the <strong>temperature</strong> values as soon as a constant <strong>droplet</strong> <strong>temperature</strong><br />

appeared. The experiments were each performed three times. This method had to be<br />

exp<strong>and</strong>ed for solution <strong>droplet</strong>s because the <strong>surface</strong> characteristic <strong>of</strong> the <strong>droplet</strong> changes<br />

after the critical point. An example for the measurement <strong>of</strong> the <strong>temperature</strong> curve on a<br />

<strong>droplet</strong> containing mannitol 15 % (w/w) in water p.a. at 40 °C <strong>and</strong> 1 % relative humidity is<br />

given in Figure 5.9. Here the <strong>temperature</strong> curve measured by the thermocouple <strong>and</strong> the<br />

curve measured by the IR-camera using the best-fit emissivity in the first <strong>drying</strong> stage are<br />

compared. It is not possible to fit the whole curve with just one setting <strong>of</strong> the emissivity.<br />

For this reason the analysis <strong>of</strong> the <strong>temperature</strong> measurement has to be divided into two<br />

parts, before <strong>and</strong> after the critical point, each with a different setting for data analysis. A


Results <strong>and</strong> Discussion 65<br />

summary <strong>of</strong> the experimentally determined emissivities for pure water <strong>droplet</strong>s in the<br />

levitation system is given in Table 5.1.<br />

Figure 5.9: Droplet <strong>surface</strong> <strong>temperature</strong> curves <strong>of</strong> a mannitol 15 % (w/w) in water<br />

p.a. solution <strong>droplet</strong> suspended on a thermocouple, (continuous line: <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> measured by the IR-camera using best fitting emissivity for the <strong>droplet</strong>,<br />

dashed line: <strong>temperature</strong> measured by the thermocouple)<br />

The published value for the emissivity <strong>of</strong> water (20 °C) is 0.96 [Schuster <strong>and</strong> Kolobrodov<br />

2004]. Note that the values determined in the experiments include the properties <strong>of</strong> the<br />

levitation chamber <strong>and</strong> are determined by assuming a transmissivity <strong>of</strong> the germanium<br />

window <strong>of</strong> 0.95.<br />

Table 5.1: Summary <strong>of</strong> the emissivities ε for pure water <strong>droplet</strong>s<br />

1 % rel. humidity 20 % rel. humidity 40 % rel. humidity 60 % rel. humidity<br />

25 °C 0.96 0.99 1.00 1.00<br />

40 °C 0.98 1.00 1.00 1.00<br />

60 °C 0.96 1.00 1.00 1.00


Results <strong>and</strong> Discussion 66<br />

The emissivities were determined again using a 0.91 m/s <strong>drying</strong> air stream directed<br />

towards the <strong>droplet</strong> to make sure ventilation does not disturb the measurement. The result<br />

is that single values differ in comparison to the measurements without ventilation<br />

airstream, but the change is in both directions <strong>and</strong> stays in the range <strong>of</strong> 0.96 to 1.00. No<br />

clear dependency can be found for the influence <strong>of</strong> a direct <strong>drying</strong> air stream. Even in<br />

experiments with different <strong>drying</strong> air velocity settings <strong>of</strong> 0.75 m/s to 3.02 m/s, the<br />

emissivities remain in the range <strong>of</strong> 0.95 to 1.00.<br />

For organic solvents the emissivities were also examined. The problem was the<br />

<strong>temperature</strong> <strong>of</strong> 50 °C at which the evaporation experiments should be performed <strong>and</strong><br />

therefore the emissivities had to be analyzed. At a <strong>drying</strong> air <strong>temperature</strong> <strong>of</strong> 50 °C the<br />

<strong>temperature</strong> difference to the boiling point <strong>of</strong> the solvents (that lie between 39.0 °C <strong>and</strong><br />

82.4 °C) is low. This low <strong>temperature</strong> difference to the boiling point leads for some<br />

solvents to a very fast evaporation <strong>of</strong> the <strong>droplet</strong>s <strong>and</strong> a rapidly decreasing wetting <strong>of</strong> the<br />

thermocouple, so that the values for the emissivity were not reproducible. The published<br />

values for emissivities <strong>of</strong> organic solvents (especially hydrocarbons) are in a range <strong>of</strong> 0.9<br />

to 1.0 [Daif et al. 1999]. Therefore it was decided to use 0.90 as value for the emissivity for<br />

data analysis <strong>of</strong> the measurements <strong>of</strong> all organic solvents in this thesis, even though some<br />

experimentally determinate values were lower.<br />

The emissivities <strong>of</strong> solutions containing mannitol <strong>and</strong> trehalose 15 % (w/w) were<br />

also analyzed at different <strong>drying</strong> air <strong>temperature</strong> <strong>and</strong> relative humidity settings. Here the<br />

emissivities <strong>of</strong> the <strong>droplet</strong> <strong>and</strong> additionally <strong>of</strong> the particle forming after the critical point<br />

need to be determined. The measurement <strong>of</strong> the wet <strong>droplet</strong> leads to similar results as in<br />

the measurements for pure water. After the critical point the measurement started to be<br />

difficult. Mannitol did not form a particle, but rather started to crystallize at the wire <strong>of</strong> the<br />

thermocouple forming fur-like crystals. Trehalose climbed the wire <strong>and</strong> filled the space<br />

between the two thermocouple wires. The emissivities were determined by variation <strong>of</strong> the<br />

values for the transmissivity <strong>and</strong> the emissivity in the second <strong>drying</strong> stage. The values for<br />

mannitol <strong>and</strong> trehalose solutions differ slightly, so that for the protein solutions <strong>and</strong><br />

protein / sugar formulations the averaged values for emissivity <strong>and</strong> transmissivity are used<br />

without performing new experiments. For data analysis the course <strong>of</strong> the <strong>temperature</strong><br />

increase at the critical point is more important than the exact particle <strong>temperature</strong> that is<br />

assumed to be close to the <strong>temperature</strong> <strong>of</strong> the <strong>drying</strong> air. Therefore a small inaccuracy due


Results <strong>and</strong> Discussion 67<br />

to the emissivities in the second <strong>drying</strong> stage does not influence the interpretation <strong>of</strong> the<br />

data curves.<br />

For quantitative analysis note that there are inaccuracies due to the measuring error<br />

<strong>of</strong> the thermocouple <strong>and</strong> the thermometer <strong>of</strong> the dew point hygrometer. Together with<br />

<strong>temperature</strong> variation <strong>of</strong> the <strong>drying</strong> air during the measurement <strong>and</strong> the inaccuracy <strong>of</strong> the<br />

emissivities for data analysis, a variation in <strong>temperature</strong> <strong>of</strong> 1 - 2 °C has to be accepted.<br />

5.2 Evaporation <strong>of</strong> pure solvent <strong>droplet</strong>s<br />

5.2.1 Evaporation <strong>of</strong> pure water <strong>droplet</strong>s<br />

5.2.1.1 Influences <strong>of</strong> the initial <strong>droplet</strong> size on the evaporation process<br />

The <strong>drying</strong> behaviour <strong>of</strong> pure solvents was analyzed using the radius squared curves, the<br />

aspect ratio curves <strong>and</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curves. The horizontal <strong>and</strong> vertical<br />

<strong>droplet</strong> radii are unequal due to the flattening by the levitation forces <strong>and</strong> are converted to<br />

the radius <strong>of</strong> a <strong>surface</strong> equivalent sphere. The <strong>droplet</strong> is assumed to be an oblate spheroid,<br />

so that the <strong>surface</strong> area <strong>of</strong> the <strong>droplet</strong> can be calculated using the measured horizontal<br />

radius <strong>and</strong> vertical radius :<br />

Equation 5.1<br />

<br />

2·· <br />

<br />

·<br />

<br />

·ln1<br />

1 <br />

where is given by:<br />

Equation 5.2<br />

1 <br />

<br />

<br />

<br />

Via the formula for the <strong>surface</strong> <strong>of</strong> a sphere, the <strong>surface</strong> equivalent radius is calculated by<br />

solving the equation for the radius using the calculated <strong>droplet</strong> <strong>surface</strong> area value <strong>of</strong> the<br />

spheroid. To have comparable curves for different <strong>droplet</strong> sizes, the value <strong>of</strong> the radius<br />

squared <strong>and</strong> also for the <strong>drying</strong> time is divided by the initial radius squared <strong>of</strong> the <strong>droplet</strong>.<br />

This correction provides a comparison <strong>of</strong> the experimental data <strong>of</strong> variations in initial<br />

<strong>droplet</strong> size. If equal initial <strong>droplet</strong> sizes are required this correction is also needed,<br />

because even using a syringe without dead volume or the microdispensing system it is not<br />

possible to obtain absolutely identical initial <strong>droplet</strong> sizes. Using this correction the<br />

equation for the graph is given by Equation 2.13 [Frohn <strong>and</strong> Roth 2000]:


Results <strong>and</strong> Discussion 68<br />

Equation 5.3<br />

²<br />

0² 1 ·<br />

<br />

0²<br />

The evaporation coefficient given in Equation 2.16 is included in the equation by:<br />

Equation 5.4<br />

²<br />

12· ·<br />

· <br />

<br />

·<br />

0² · <br />

<br />

0²<br />

with the <strong>surface</strong>-equivalent <strong>droplet</strong> radius squared as a function <strong>of</strong> time ², the initial<br />

<strong>surface</strong> equivalent <strong>droplet</strong> radius squared 0 , the <strong>drying</strong> time , the binary diffusion<br />

coefficient calculated using the method by Fuller [1966], the molar mass , the<br />

density <strong>of</strong> water , the ideal gas constant , the partial vapour pressure at the <strong>droplet</strong><br />

<strong>surface</strong> <strong>and</strong> at infinity , the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> , <strong>and</strong> the <strong>drying</strong> air<br />

<strong>temperature</strong> . The curves show a linear decrease in the radius squared if the evaporation<br />

process can be described by the d²-law as a model for diffusional evaporation processes <strong>of</strong><br />

pure, spherical solvent <strong>droplet</strong>s at a fixed <strong>surface</strong> <strong>temperature</strong> <strong>and</strong> a fixed <strong>drying</strong> air<br />

<strong>temperature</strong> [Frohn <strong>and</strong> Roth 2000].<br />

The evaporation coefficient is directly given as the negative slope <strong>of</strong> the<br />

² ⁄ 0² versus ⁄ r0² plot <strong>and</strong> characterises the evaporation process. It is dependent<br />

on the thermodynamic properties <strong>of</strong> the <strong>droplet</strong> liquid <strong>and</strong> on conditions <strong>of</strong> the ambience<br />

like the <strong>temperature</strong>. The aspect ratio curve is given by the relation <strong>of</strong> the horizontal <strong>and</strong><br />

vertical diameter, measured in the shadow image taken by the CCD-camera using<br />

ImageProPlus s<strong>of</strong>tware. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curve is measured by the IRcamera.<br />

The minimum <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> values are corrected using the measured<br />

emissivity <strong>of</strong> the solvent at given conditions. For comparison <strong>of</strong> the <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong>s at different conditions the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> was taken from the<br />

linear part <strong>of</strong> the <strong>temperature</strong> curve after equilibration time <strong>and</strong> before the <strong>temperature</strong><br />

increases at the end <strong>of</strong> the measurement. All experiments were performed six times <strong>and</strong> the<br />

graphs presented are the averaged curves.<br />

Three different initial <strong>droplet</strong> sizes for water p.a. <strong>of</strong> approx. 500 μm, 800 μm <strong>and</strong><br />

1200 μm initial horizontal diameter were analyzed at a <strong>drying</strong> air <strong>temperature</strong> <strong>of</strong> 25 °C,<br />

40 °C <strong>and</strong> 60 °C <strong>and</strong> a relative humidity <strong>of</strong> 1 %, 20 %, 40 % <strong>and</strong> 60 % without a direct<br />

<strong>drying</strong> air stream towards the <strong>droplet</strong>. The water p.a. was filtered freshly before starting the<br />

experiments using a filtration unit with 0.2 μm pore diameter (Whatman®, FP 30/0.2,


Results <strong>and</strong> Discussion 69<br />

Schleicher & Schuell, Dassel, Germany). The <strong>droplet</strong>s were inserted in the acoustic field<br />

using the microdispenser system. Figure 5.10 shows the curves for 1200 μm water <strong>droplet</strong>s<br />

at 25 °C <strong>and</strong> 60 °C at 1 % relative humidity as an example for the curve progression.<br />

(a)<br />

(b)<br />

Figure 5.10: Drying behaviour <strong>of</strong> a 1200 μm water <strong>droplet</strong> at (a) 25 °C <strong>and</strong> (b) 60 °C<br />

at 1 % relative humidity


Results <strong>and</strong> Discussion 70<br />

The ² ⁄ 0² -graphs show a linear decrease <strong>of</strong> the radius squared. Only the last part <strong>of</strong><br />

the curve is a little flattened. For this reason the evaporation coefficient is calculated using<br />

the linear part <strong>of</strong> the curve. The aspect ratio is nearly constant at the beginning <strong>of</strong> the<br />

measurement <strong>and</strong> shows fluctuations at the end <strong>of</strong> the evaporation process due to <strong>droplet</strong><br />

oscillations. For higher <strong>temperature</strong>s a higher SPL is necessary to provide stable levitation<br />

<strong>of</strong> the <strong>droplet</strong> <strong>and</strong> to minimize oscillations that can influence the <strong>droplet</strong> size <strong>and</strong> <strong>surface</strong><br />

<strong>temperature</strong> measurement by the cameras. In this example the SPL at 25 °C was 161 dB<br />

<strong>and</strong> at 60 °C 165 dB. A higher SPL leads to further flattening <strong>of</strong> the <strong>droplet</strong> <strong>and</strong> therefore<br />

the aspect ratio is increased at 60 °C. After a short equilibration time the <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> curves show a constant <strong>temperature</strong> that is lower than the <strong>drying</strong> air<br />

<strong>temperature</strong> due to the evaporation <strong>of</strong> water. At the end <strong>of</strong> the measurement the<br />

<strong>temperature</strong> increases.<br />

Figure 5.11, Figure 5.12 <strong>and</strong> Figure 5.13 show the evaporation behaviour <strong>of</strong> water<br />

<strong>droplet</strong>s at 25 °C <strong>and</strong> 1 %, 20 %, 40 % <strong>and</strong> 60 % relative humidity in comparison <strong>of</strong> the<br />

different initial <strong>droplet</strong> sizes. With rising relative humidity the <strong>drying</strong> time increases <strong>and</strong><br />

the evaporation coefficient decreases. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> increases with<br />

increasing relative humidity for all initial <strong>droplet</strong> sizes. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong><br />

values correspond for all initial <strong>droplet</strong> sizes.<br />

Figure 5.11: Drying behaviour <strong>of</strong> water <strong>droplet</strong>s at 25 °C <strong>and</strong> 1 %, 20 %, 40 % <strong>and</strong><br />

60 % relative humidity with an initial <strong>droplet</strong> diameter <strong>of</strong> 500 μm (SPL = 160 dB)


Results <strong>and</strong> Discussion 71<br />

Figure 5.12: Drying behaviour <strong>of</strong> water <strong>droplet</strong>s at 25 °C <strong>and</strong> 1 %, 20 %, 40 % <strong>and</strong><br />

60 % relative humidity with an initial <strong>droplet</strong> diameter <strong>of</strong> 800 μm (SPL = 160 dB)<br />

Figure 5.13: Drying behaviour <strong>of</strong> water <strong>droplet</strong>s at 25 °C <strong>and</strong> 1 %, 20 %, 40 % <strong>and</strong><br />

60 % relative humidity with an initial <strong>droplet</strong> diameters <strong>of</strong> 1200 μm (SPL = 161 dB)<br />

The aspect ratio curves show that with the shrinking <strong>of</strong> the <strong>droplet</strong> more <strong>droplet</strong><br />

oscillations occur leading to an unsteady graph. For the aspect ratio no dependence on the<br />

relative humidity can be seen in these measurements. The aspect ratio increases with


Results <strong>and</strong> Discussion 72<br />

increasing initial <strong>droplet</strong> diameter, because <strong>droplet</strong>s having a higher mass need a stronger<br />

SPL for stable levitation that flattens the <strong>droplet</strong>. The initial SPL in the experiments was<br />

1 dB higher for the 1200 μm <strong>droplet</strong>s (161 dB) than for the smaller ones (160 dB).<br />

Figure 5.11 for the initial <strong>droplet</strong> size <strong>of</strong> 500 μm shows an aspect ratio value < 1<br />

leading to a prolate spheroid, whereas the <strong>droplet</strong> sizes <strong>of</strong> 1200 μm show mostly as<br />

assumed an oblate spheroid with an aspect ratio > 1. The <strong>droplet</strong>s having an initial <strong>droplet</strong><br />

size <strong>of</strong> 800 μm show an aspect ratio around 1. A conversion from an oblate to a prolate<br />

spheroid can sometimes also be seen for shrinking <strong>droplet</strong>s <strong>of</strong> higher initial diameters at<br />

the end <strong>of</strong> the evaporation process. In the case <strong>of</strong> an aspect ratio < 1 the calculation <strong>of</strong> the<br />

<strong>surface</strong> equivalent diameter has to be adapted using the formula for a prolate spheroid:<br />

Equation 5.5<br />

<br />

2·· <br />

2·· · <br />

<br />

·sin <br />

Equation 5.6<br />

1 <br />

<br />

<br />

<br />

The same evaporation behaviour in dependence <strong>of</strong> the relative humidity given for a <strong>drying</strong><br />

air <strong>temperature</strong> <strong>of</strong> 25 °C was found at higher <strong>drying</strong> air <strong>temperature</strong>s <strong>of</strong> 40 °C <strong>and</strong> 60 °C.<br />

The evaporation behaviour <strong>of</strong> water has previously been analyzed by Tuckermann<br />

[2002] <strong>and</strong> Schiffter <strong>and</strong> Lee [2007a]. Tuckermann published a measurement <strong>of</strong> a water<br />

<strong>droplet</strong> at 20 °C <strong>and</strong> 28 % relative humidity. For data analysis Tuckermann used the<br />

<strong>droplet</strong> <strong>surface</strong> area decrease which is also linear. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> was also<br />

monitored <strong>and</strong> was also almost constant at about 12.1 °C. The measurements for water<br />

were performed without variations <strong>of</strong> the <strong>drying</strong> air conditions. Schiffter <strong>and</strong> Lee<br />

performed measurements <strong>of</strong> evaporating water <strong>droplet</strong>s at different ambient conditions<br />

using a CCD-camera. They also found an almost linear decrease <strong>of</strong> the radius squared at<br />

different <strong>drying</strong> air <strong>temperature</strong>s <strong>and</strong> relative humidity.<br />

Figure 5.14 is a summary <strong>of</strong> the experimental evaporation coefficients (a)<br />

depending on the <strong>drying</strong> air <strong>temperature</strong> <strong>and</strong> (b) on the relative humidity for 1200 μm<br />

initial diameter <strong>droplet</strong>s. The values <strong>of</strong> the evaporation coefficients were compared to the<br />

experiments performed by Schiffter [2006]. The experimental values in Schiffter’s thesis<br />

show the same tendency as shown for the present experiments in Figure 5.14.


Results <strong>and</strong> Discussion 73<br />

(a)<br />

Figure 5.14: Evaporation coefficients <strong>of</strong> water <strong>droplet</strong>s with an initial horizontal<br />

<strong>droplet</strong> diameter <strong>of</strong> 1200 μm dependent on (a) the <strong>drying</strong> air <strong>temperature</strong> <strong>and</strong> (b) the<br />

relative humidity<br />

The evaporation coefficient increases with increasing <strong>drying</strong> air <strong>temperature</strong> <strong>and</strong><br />

decreasing relative humidity. The absolute values for 25 °C <strong>and</strong> 40 °C in the experiments<br />

correspond with the values by Schiffter, but the evaporation coefficients at 60 °C show<br />

lower values. This might be due to the way <strong>of</strong> bringing the <strong>droplet</strong> in the pressure node. By<br />

using the microdispensing system, the <strong>droplet</strong> has to be built up by several injections <strong>of</strong> a<br />

<strong>droplet</strong> stream into the pressure node. The <strong>droplet</strong>s can be redirected by streaming near the<br />

pressure node, so that this procedure needs more time <strong>and</strong> another setting <strong>of</strong> the SPL than<br />

with injection using a microsyringe. This may lead to <strong>temperature</strong> <strong>and</strong> relative humidity<br />

variations that are more noticeable at higher <strong>drying</strong> air <strong>temperature</strong>. Additionally the<br />

different SPL settings can lead to differences in the evaporation process.<br />

The predicted <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> according to Yarin et al. [1999]<br />

(T(Yarin)) was calculated according to Equation 3.9 for all experimental conditions. The<br />

<strong>surface</strong> <strong>temperature</strong> values were then used to calculate an evaporation coefficient.<br />

Additionally, the wet-bulb <strong>temperature</strong> <strong>of</strong> water (T(wb)) at the respective conditions was<br />

taken from a psychrometric calculator (www.linric.com/webpsysi.htm, Linric Company,<br />

Bedford, USA) <strong>and</strong> a psychrometric chart (Nautical dehumidifiers Inc., Huntington, USA).<br />

The evaporation coefficients using the wet-bulb <strong>temperature</strong> were also calculated for the<br />

comparison <strong>of</strong> the predicted (using T(Yarin) <strong>and</strong> T(wb)) <strong>and</strong> experimental <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong>s <strong>and</strong> evaporation coefficients. The comparison <strong>of</strong> the evaporation coefficients<br />

<strong>and</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s is given in Figure 5.15. The initial SPL was about<br />

(b)


Results <strong>and</strong> Discussion 74<br />

161 dB for experiments at 25 °C <strong>and</strong> 162 - 165 dB for experiments at 40 °C <strong>and</strong> 60 °C<br />

depending on the <strong>droplet</strong> size.<br />

(a)<br />

(b)<br />

(c)<br />

(d)<br />

(e)<br />

Figure 5.15: Comparison <strong>of</strong> the experimental <strong>and</strong> calculated values for the<br />

evaporation coefficients <strong>and</strong> for the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s at (a, b) 25 °C (c, d)<br />

40 °C <strong>and</strong> (e, f) 60 °C <strong>drying</strong> air <strong>temperature</strong> dependent on the relative humidity<br />

(f)


Results <strong>and</strong> Discussion 75<br />

Generally the evaporation coefficient increases with increasing <strong>drying</strong> air <strong>temperature</strong> <strong>and</strong><br />

decreasing relative humidity <strong>and</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> increases with increasing<br />

<strong>drying</strong> air <strong>temperature</strong> <strong>and</strong> increasing relative humidity. The comparisons <strong>of</strong> the<br />

evaporation coefficients show for all <strong>drying</strong> air <strong>temperature</strong>s a decrease with smaller initial<br />

<strong>droplet</strong> diameter that is most clearly seen at a relative humidity <strong>of</strong> 1 % (Figure 5.15). The<br />

calculated value using the wet-bulb <strong>temperature</strong> was lower than all measured evaporation<br />

coefficients. The evaporation coefficient calculated using the method described by Yarin et<br />

al. [1999] leads to lower values than using T(wb). With increasing relative humidity this<br />

order is constant for 25 °C, but the deviation between the values decrease. In the graphs for<br />

the 40 °C <strong>and</strong> the 60 °C experiments the values are closer together with rising relative<br />

humidity <strong>and</strong> at 60 °C <strong>and</strong> 40 % / 60 % relative humidity the calculated evaporation<br />

coefficients are now higher than the measured ones.<br />

The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> values also show a deviation between the<br />

experimental <strong>and</strong> the calculative values. The experimental <strong>surface</strong> <strong>temperature</strong> values are<br />

higher than the wet-bulb <strong>temperature</strong> <strong>and</strong> the calculated <strong>surface</strong> <strong>temperature</strong> using the<br />

method by Yarin et al. [1999]. In the <strong>surface</strong> <strong>temperature</strong> the experimental values show no<br />

substantial deviation for the three <strong>droplet</strong> sizes. As for the evaporation coefficient, with<br />

increasing relative humidity the theoretical <strong>and</strong> experimental values for the <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> become closer <strong>and</strong> are almost equal for high <strong>temperature</strong> <strong>and</strong> high relative<br />

humidity. As mentioned before, a <strong>temperature</strong> deviation <strong>of</strong> 1 - 2 °C may be due to<br />

measurement inaccuracy.<br />

Figure 5.16, Figure 5.17 <strong>and</strong> Figure 5.18 show this phenomenon in the<br />

experimental evaporation curves at 25 °C, 40 °C <strong>and</strong> 60 °C given at 1 % <strong>and</strong> 60 % relative<br />

humidity. Additionally, the calculated evaporation curves using the wet-bulb <strong>temperature</strong><br />

for the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>and</strong> a horizontal line depicting the wet-bulb <strong>temperature</strong><br />

are displayed. The diagrams for 25 °C clearly show that the wet-bulb <strong>temperature</strong> lies<br />

below the experimental <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s for low as well as for high relative<br />

humidity. At higher <strong>drying</strong> air <strong>temperature</strong> <strong>of</strong> 40 °C <strong>and</strong> 60 °C the wet-bulb <strong>temperature</strong> is<br />

at 1 % relative humidity lower than the experimental values, but at 60 % relative humidity<br />

the values are almost equal. The experimental ² ⁄ 0² -graphs for 25 °C <strong>drying</strong> air<br />

<strong>temperature</strong> have a steeper decrease than the calculated graph using the wet-bulb<br />

<strong>temperature</strong> values. For higher <strong>drying</strong> air <strong>temperature</strong> combined with high relative<br />

humidity this relation changes <strong>and</strong> the calculated evaporation coefficient is higher.


Results <strong>and</strong> Discussion 76<br />

(a)<br />

(b)<br />

Figure 5.16: Comparison <strong>of</strong> the experimental r(t)²/r(0)²-graphs, <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> curves <strong>and</strong> aspect ratio curves <strong>and</strong> the calculated evaporation curve<br />

using the wet-bulb <strong>temperature</strong> for water <strong>droplet</strong>s at (a) 25 °C <strong>and</strong> 1 % relative<br />

humidity <strong>and</strong> (b) 25 °C <strong>and</strong> 60 % relative humidity


Results <strong>and</strong> Discussion 77<br />

(a)<br />

(b)<br />

Figure 5.17: Comparison <strong>of</strong> the experimental r(t)²/r(0)²-graphs, <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> curves <strong>and</strong> aspect ratio curves <strong>and</strong> the calculated evaporation curve<br />

using the wet-bulb <strong>temperature</strong> for water <strong>droplet</strong>s at (a) 40 °C <strong>and</strong> 1 % relative<br />

humidity <strong>and</strong> (b) 40 °C <strong>and</strong> 60 % relative humidity


Results <strong>and</strong> Discussion 78<br />

(a)<br />

(b)<br />

Figure 5.18: Comparison <strong>of</strong> the experimental r(t)²/r(0)²-graphs, <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> curves <strong>and</strong> aspect ratio curves <strong>and</strong> the calculated evaporation curve<br />

using the wet-bulb <strong>temperature</strong> for water <strong>droplet</strong>s at (a) 60 °C <strong>and</strong> 1 % relative<br />

humidity <strong>and</strong> (b) 60 °C <strong>and</strong> 60 % relative humidity


Results <strong>and</strong> Discussion 79<br />

If the evaporation process <strong>of</strong> the three <strong>droplet</strong> sizes would correspond to the d²-law, then<br />

the curve progression should be independent <strong>of</strong> the initial <strong>droplet</strong> size. The reason for the<br />

increasing evaporation coefficient at a higher initial <strong>droplet</strong> radius can be found in the SPL<br />

used for levitation <strong>of</strong> the <strong>droplet</strong>. The aspect ratio is a parameter that shows the influence<br />

<strong>of</strong> the acoustic field. A higher SPL leads to a more flattened <strong>droplet</strong> <strong>and</strong> therefore to a<br />

higher aspect ratio. For stable levitation <strong>of</strong> the <strong>droplet</strong>s <strong>of</strong> 1200 μm initial diameter a<br />

higher SPL has to be used that can be seen at the higher aspect ratio > 1 in comparison to<br />

the smaller initial <strong>droplet</strong> sizes. The 800 μm <strong>and</strong> 500 μm <strong>droplet</strong>s have a lower aspect ratio<br />

<strong>of</strong> about 1 or even < 1. For the 25 °C <strong>and</strong> 40 °C experiments the values are similar, but in<br />

the 60 °C experiments a clear dependency on the <strong>droplet</strong> size can be seen. The higher SPL<br />

leads to a higher influence <strong>of</strong> primary acoustic streaming which enhances the evaporation<br />

process <strong>and</strong> leads to higher evaporation coefficients for <strong>droplet</strong>s having a higher initial<br />

<strong>droplet</strong> size. An additional effect is given with increasing <strong>drying</strong> air <strong>temperature</strong>, where a<br />

higher SPL has to be used to provide stable levitation for the experiments.<br />

The deviation <strong>of</strong> experimental <strong>and</strong> calculated graphs was already observed by<br />

Schiffter <strong>and</strong> Lee [2007a] for water <strong>and</strong> ethanol in comparison to diffusion controlled<br />

model. Yarin et al. [1999] gave as an explanation for the deviating evaporation coefficient<br />

the constitution <strong>of</strong> the inner <strong>and</strong> outer acoustic streaming which both increases <strong>and</strong><br />

decreases evaporation (see 3.4). The deviation in experimentally-measured <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> <strong>and</strong> the wet-bulb <strong>temperature</strong> or the <strong>surface</strong> <strong>temperature</strong> predicted by Yarin et<br />

al. [1999] can be explained by influences <strong>of</strong> the st<strong>and</strong>ing acoustic wave. This gap exp<strong>and</strong>s<br />

for increasing <strong>drying</strong> air <strong>temperature</strong> <strong>and</strong> decreasing relative humidity.<br />

Tuckermann [2002] found for a levitated water <strong>droplet</strong> at 20 °C <strong>and</strong> 28 % relative<br />

humidity at 167 dB a <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>of</strong> about 12.2 °C that is also above the<br />

wet-bulb <strong>temperature</strong> <strong>of</strong> 10.5 °C. In further experiments Tuckermann found for a levitated<br />

water <strong>droplet</strong> at 21 °C <strong>drying</strong> air <strong>temperature</strong>, a relative humidity <strong>of</strong> 35 - 40 % <strong>and</strong> a SPL<br />

<strong>of</strong> 160 dB in the pressure nodes a <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>of</strong> about 13 °C that is similar<br />

to the wet-bulb <strong>temperature</strong> <strong>of</strong> 12.3 - 13.1 °C [Tuckermann et al. 2005]. In the calculation<br />

model for the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>of</strong> acoustically levitated <strong>droplet</strong>s Yarin et al.<br />

[1999] take into account the influence <strong>of</strong> the inner acoustic streaming. The calculated<br />

values using this relation (3.7.1) are located further below the values <strong>of</strong> the wet-bulb<br />

<strong>temperature</strong> that is due to the additional convective driven evaporation induced by the<br />

inner acoustic streaming. The experiments presented in this chapter were not conducted


Results <strong>and</strong> Discussion 80<br />

with a <strong>drying</strong> air stream that was directed on the levitated <strong>droplet</strong>, so the outer acoustic<br />

streaming was not eliminated to fit the premises for the calculations by Yarin et al. [1999].<br />

The deviations between the experimental <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s <strong>and</strong> the values for<br />

the wet-bulb <strong>temperature</strong> are assumed to be due to the additional energy supply to the<br />

levitated <strong>droplet</strong> by the acoustic field. The <strong>temperature</strong> increase in the levitation chamber<br />

after power-on <strong>of</strong> the levitator (without CEM or external heating influence) due to the<br />

acoustic wave could clearly be seen <strong>and</strong> was in the range <strong>of</strong> up to 4 °C in this levitation<br />

set-up. The explanation <strong>of</strong> a heat input to the <strong>droplet</strong> by the acoustic wave seems therefore<br />

possible. This <strong>temperature</strong> increase leads to a faster evaporation <strong>of</strong> the water at higher<br />

<strong>temperature</strong> <strong>and</strong> therefore to the higher evaporation coefficients.<br />

To describe the deviation from the diffusion controlled evaporation process the<br />

Sherwood number is calculated using Equation 3.8 introduced in Equation 5.3 by:<br />

Equation 5.7<br />

²<br />

12· ·<br />

· <br />

<br />

·<br />

0² · 2 ·<br />

<br />

0²<br />

For either the experimental <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> or the wet-bulb <strong>temperature</strong> can<br />

be used. By fitting this equation to the experimental values for ² ⁄ 0² versus ⁄<br />

0²<br />

the values for the Sherwood number are calculated during the evaporation process (Figure<br />

5.19, Figure 5.20 <strong>and</strong> Figure 5.21). A value for the Sherwood number <strong>of</strong> 2 shows close<br />

agreement with the d²-law. Values > 2 are a sign for higher <strong>drying</strong> rates than predicted by<br />

the d²-law. The values calculated using the wet-bulb <strong>temperature</strong> are > 2 for most <strong>drying</strong><br />

air conditions <strong>and</strong> also higher than the values calculated using the experimental <strong>droplet</strong><br />

<strong>surface</strong> <strong>temperature</strong> which are close to 2. This shows good agreement using the<br />

experimental <strong>surface</strong> <strong>temperature</strong> values for the d²-law for experiments without direct<br />

<strong>drying</strong> air stream to the <strong>droplet</strong>. Regarding the d²-law, a higher <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong><br />

in the calculation will increase the binary diffusion coefficient <strong>and</strong> has also an influence on<br />

the saturation vapour pressure at T s . The increasing evaporation coefficients are in this case<br />

not sufficiently explained by the influence <strong>of</strong> inner acoustic streaming. The experimental<br />

<strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s lead to close agreement with the d²-law <strong>and</strong> the reason for the<br />

higher evaporation coefficients can be given by the higher <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>.


Results <strong>and</strong> Discussion 81<br />

(a)<br />

(b)<br />

Figure 5.19: Comparison <strong>of</strong> the Sherwood numbers <strong>of</strong> water <strong>droplet</strong>s calculated using<br />

the experimental <strong>surface</strong> <strong>temperature</strong> T(exp) <strong>and</strong> the wet-bulb <strong>temperature</strong> T(wb) at<br />

(a) 25 °C <strong>and</strong> 1 % relative humidity <strong>and</strong> (b) 25 °C <strong>and</strong> 60 % relative humidity


Results <strong>and</strong> Discussion 82<br />

(a)<br />

(b)<br />

Figure 5.20: Comparison <strong>of</strong> the Sherwood numbers <strong>of</strong> water <strong>droplet</strong>s calculated using<br />

the experimental <strong>surface</strong> <strong>temperature</strong> T(exp) <strong>and</strong> the wet-bulb <strong>temperature</strong> T(wb) at<br />

(a) 40 °C <strong>and</strong> 1 % relative humidity <strong>and</strong> (b) 40 °C <strong>and</strong> 60 % relative humidity


Results <strong>and</strong> Discussion 83<br />

(a)<br />

(b)<br />

Figure 5.21: Comparison <strong>of</strong> the Sherwood numbers <strong>of</strong> water <strong>droplet</strong>s calculated using<br />

the experimental <strong>surface</strong> <strong>temperature</strong> T(exp) <strong>and</strong> the wet-bulb <strong>temperature</strong> T(wb) at<br />

(a) 60 °C <strong>and</strong> 1 % relative humidity <strong>and</strong> (b) 60 °C <strong>and</strong> 40 % relative humidity


Results <strong>and</strong> Discussion 84<br />

The Sherwood number also shows that <strong>droplet</strong>s with a larger initial size have a higher<br />

divergence from the d²-law than the 500 μm <strong>and</strong> 800 μm <strong>droplet</strong>s. That shows a higher<br />

<strong>drying</strong> rate with a higher initial <strong>droplet</strong> size. Here the influence <strong>of</strong> acoustic streaming on<br />

<strong>droplet</strong> evaporation can be seen, because larger <strong>droplet</strong>s need a higher SPL for levitation<br />

that was already shown for the aspect ratio. With rising relative humidity the deviation <strong>of</strong><br />

the Sherwood number calculated using the different <strong>surface</strong> <strong>temperature</strong>s decreases.<br />

The deviation between the experimental <strong>and</strong> calculated values for the evaporation<br />

coefficient <strong>and</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> decreases for higher relative humidity <strong>of</strong><br />

the <strong>drying</strong> air. The initial SPL does not change noticeably, so this decrease cannot be<br />

sufficiently explained by less application <strong>of</strong> energy by the acoustic field to the levitated<br />

<strong>droplet</strong>. The decreasing deviation <strong>of</strong> the evaporation coefficients appears to be due to<br />

accumulation <strong>of</strong> water vapour in the outer acoustic streaming that decreases the<br />

evaporation process. This seems to be more important at higher relative humidity. But this<br />

does not explain the disappearance <strong>of</strong> the <strong>surface</strong> <strong>temperature</strong> deviation, because with<br />

increasing relative humidity around the <strong>droplet</strong> the <strong>surface</strong> <strong>temperature</strong> should increase. A<br />

possible explanation is that with increasing relative humidity the <strong>drying</strong> air <strong>temperature</strong><br />

close to the <strong>droplet</strong> decreases in comparison to the <strong>drying</strong> air <strong>temperature</strong> in the chamber<br />

due to the shielding effect <strong>of</strong> outer acoustic streaming. Especially for high <strong>drying</strong> air<br />

<strong>temperature</strong>s this effect could lead to a <strong>drying</strong> air <strong>temperature</strong> deviation. This idea could<br />

not be verified experimentally by <strong>temperature</strong> measurements in the levitation setup.<br />

5.2.1.2 Influence <strong>of</strong> the <strong>drying</strong> air stream on the evaporation process<br />

Evaporation experiments were performed using a <strong>drying</strong> air stream directed towards the<br />

<strong>droplet</strong> at different settings <strong>of</strong> the <strong>drying</strong> air velocity in order to minimize the influence <strong>of</strong><br />

outer acoustic streaming on the evaporation process <strong>of</strong> the <strong>droplet</strong>s. The effect on the<br />

<strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> was also analyzed. In these experiments the air streamed<br />

directly towards the <strong>droplet</strong>s via the hole in the reflector. To exclude the influence <strong>of</strong> the<br />

airstream on the emissivities used for the <strong>surface</strong> <strong>temperature</strong> analysis, the emissivities <strong>of</strong><br />

water at different conditions were determined again using different air stream settings. No<br />

dependency with the airstream velocity was detectable, the values are similar to the ones<br />

without airstream.<br />

In a first set <strong>of</strong> experiments the influence <strong>of</strong> different <strong>drying</strong> air velocities <strong>of</strong> 0.75,<br />

1.51, 2.26 <strong>and</strong> 3.02 m/s was analyzed at 25 °C <strong>and</strong> 1 % relative humidity <strong>and</strong> at 40 °C <strong>and</strong>


Results <strong>and</strong> Discussion 85<br />

20 % relative humidity. The <strong>droplet</strong>s <strong>of</strong> initial volume 1 μl were inserted into the ultrasonic<br />

field using the microsyringe. The SPL for experiments with <strong>and</strong> without ventilation<br />

airstream at 25 °C <strong>and</strong> 1 % relative humidity is 164.0 - 165.4 dB <strong>and</strong> at 40 °C <strong>and</strong> 20 %<br />

relative humidity 163.4 - 164.8 dB. Experiments at a <strong>drying</strong> air velocity <strong>of</strong> 3.02 m/s require<br />

a slightly higher SPL to avoid oscillations. The comparison <strong>of</strong> the /0²-graphs,<br />

<strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s <strong>and</strong> aspect ratios is given in Figure 5.22 <strong>and</strong> Figure 5.23. The<br />

evaporation coefficient increases as expected with increasing <strong>drying</strong> air velocity for both<br />

<strong>drying</strong> air conditions due to reduction <strong>of</strong> the slowing influence <strong>of</strong> outer acoustic streaming.<br />

The aspect ratio is not influenced by the <strong>drying</strong> air stream. For all airstream settings it is<br />

about 1.1 - 1.2. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> seems to increase for increasing <strong>drying</strong> air<br />

velocity at 40 °C.<br />

The <strong>temperature</strong> curves in Figure 5.23 show fluctuating lines, which means that the<br />

<strong>temperature</strong> measurement is subject to error by oscillations <strong>of</strong> the <strong>droplet</strong>s that increase<br />

with higher <strong>drying</strong> air velocity. The <strong>temperature</strong> fluctuations lead to measurement<br />

inaccuracies because <strong>of</strong> changes in the <strong>droplet</strong> distance to the IR-camera. The <strong>surface</strong><br />

<strong>temperature</strong> is slightly higher for increasing <strong>drying</strong> air velocity, but this variation is with<br />

1 - 2 °C close to the measurement inaccuracy.<br />

Figure 5.22: Drying behaviour <strong>of</strong> levitated water p.a. <strong>droplet</strong>s at 25 °C <strong>and</strong> 1 %<br />

relative humidity using different <strong>drying</strong> air stream settings


Results <strong>and</strong> Discussion 86<br />

Figure 5.23: Drying behaviour <strong>of</strong> levitated water p.a. <strong>droplet</strong>s at 40 °C <strong>and</strong> 20 %<br />

relative humidity using different <strong>drying</strong> air stream settings<br />

Rensink [2004] investigated the evaporation <strong>of</strong> water <strong>and</strong> organic solvent <strong>droplet</strong>s at<br />

different <strong>drying</strong> air velocities <strong>and</strong> found 1 m/s as the optimum airstream velocity in his<br />

experimental setup. At this setting the airflow was able to remove humidity accumulation<br />

in the outer acoustic streaming, whereas additional convective influence on the evaporation<br />

process was prevented. Schiffter [2006] found for his experimental setup the optimum<br />

<strong>drying</strong> air velocity <strong>of</strong> 0.82 m/s. In the following experiments the <strong>drying</strong> air velocity is set<br />

to 0.91 m/s at several <strong>drying</strong> air conditions. The <strong>droplet</strong>s <strong>of</strong> initial volume <strong>of</strong> 1 μl were<br />

inserted using the microsyringe. To ensure equal SPL settings the experiments without<br />

direct airstream were repeated using the microsyringe. The SPL for experiments with <strong>and</strong><br />

without a direct <strong>drying</strong> air stream is 163.6 dB at 25 °C, 164.5 dB at 40 °C <strong>and</strong> 166.6 dB at<br />

60 °C.<br />

Figure 5.24, Figure 5.25 <strong>and</strong> Figure 5.26 show the comparison <strong>of</strong> the evaporation<br />

behaviour with <strong>and</strong> without a <strong>drying</strong> air stream directed towards the <strong>droplet</strong> at different<br />

<strong>drying</strong> air conditions. The evaporation coefficient is as expected clearly higher for the<br />

measurements with a direct <strong>drying</strong> air stream at all <strong>drying</strong> air settings, <strong>and</strong> the <strong>drying</strong> time<br />

decreases. The aspect ratio does not show a dependency on the <strong>drying</strong> air stream, though<br />

<strong>droplet</strong> oscillations at high <strong>drying</strong> air velocities occur.


Results <strong>and</strong> Discussion 87<br />

(a)<br />

(b)<br />

Figure 5.24: Evaporation <strong>of</strong> water <strong>droplet</strong>s using 0.0 <strong>and</strong> 0.91 m/s <strong>drying</strong> air velocity<br />

at (a) 25 °C <strong>and</strong> 1 % relative humidity <strong>and</strong> (b) 25 °C <strong>and</strong> 40 % relative humidity


Results <strong>and</strong> Discussion 88<br />

(a)<br />

(b)<br />

Figure 5.25: Evaporation <strong>of</strong> water <strong>droplet</strong>s using 0.0 <strong>and</strong> 0.91 m/s <strong>drying</strong> air velocity<br />

at (a) 40 °C <strong>and</strong> 1 % relative humidity <strong>and</strong> (b) 40 °C <strong>and</strong> 40 % relative humidity


Results <strong>and</strong> Discussion 89<br />

(a)<br />

(b)<br />

Figure 5.26: Evaporation <strong>of</strong> water <strong>droplet</strong>s using 0.0 <strong>and</strong> 0.91 m/s <strong>drying</strong> air velocity<br />

at (a) 60 °C <strong>and</strong> 1 % relative humidity <strong>and</strong> (b) 60 °C <strong>and</strong> 40 % relative humidity<br />

The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> for the measurements without a direct air stream in the<br />

experiments at a <strong>drying</strong> air <strong>temperature</strong> <strong>of</strong> 25 °C is higher than the <strong>surface</strong> <strong>temperature</strong><br />

with a direct air stream to the <strong>droplet</strong>. With increasing <strong>drying</strong> air velocity the <strong>droplet</strong><br />

<strong>surface</strong> is cooled by the faster evaporation, so that the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>


Results <strong>and</strong> Discussion 90<br />

decreases. At a <strong>drying</strong> air <strong>temperature</strong> <strong>of</strong> 40 °C the <strong>surface</strong> <strong>temperature</strong>s are similar,<br />

whereas at 60 °C the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> with a direct <strong>drying</strong> air stream is higher.<br />

This might be due to the faster supply <strong>of</strong> fresh hot <strong>drying</strong> air to the <strong>droplet</strong> at higher <strong>drying</strong><br />

air velocity that overcomes the effect <strong>of</strong> increasing evaporation.<br />

Figure 5.27 shows a comparison <strong>of</strong> the evaporation coefficients <strong>and</strong> the <strong>droplet</strong><br />

<strong>surface</strong> <strong>temperature</strong>s <strong>of</strong> the experiments with <strong>and</strong> without a direct <strong>drying</strong> air stream<br />

towards the <strong>droplet</strong>. Additionally the calculated evaporation coefficients (using T(wb)) <strong>and</strong><br />

the wet-bulb <strong>temperature</strong>s are depicted. The evaporation coefficients for 0.91 m/s <strong>drying</strong><br />

air velocity lie at all ambient conditions above the values for the measurements without a<br />

direct <strong>drying</strong> air stream <strong>and</strong> also above the values calculated using the wet-bulb<br />

<strong>temperature</strong> in the diffusion model. This is due to the reduction <strong>of</strong> the influence <strong>of</strong> the<br />

outer acoustic streaming via the <strong>drying</strong> air stream. The deviation becomes less with rising<br />

relative humidity. In the experiments without additional ventilation, the increase in<br />

accumulation <strong>of</strong> water vapour in the outer vortices might have less influence on the<br />

evaporation process at high relative humidity, because there is a higher <strong>drying</strong> air humidity<br />

in the whole levitation chamber.<br />

The relation <strong>of</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s changes with increasing <strong>drying</strong> air<br />

<strong>temperature</strong> as already shown in Figure 5.24, Figure 5.25 <strong>and</strong> Figure 5.26. At low <strong>drying</strong><br />

air <strong>temperature</strong>s the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> is less higher in the experiments without<br />

ventilation than in the experiments with 0.91 m/s <strong>drying</strong> air velocity. At 60 °C <strong>drying</strong> air<br />

<strong>temperature</strong> this relation changes <strong>and</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>of</strong> the measurements<br />

using a direct <strong>drying</strong> air stream is higher. The clearest deviation is given in the experiments<br />

at 60 °C, where the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> using a direct <strong>drying</strong> air stream is about<br />

5 °C higher than in the experiments without ventilation. A <strong>temperature</strong> deviation <strong>of</strong> about<br />

1 °C can be due to measurement errors.<br />

The increase <strong>of</strong> the evaporation coefficient with increasing <strong>drying</strong> air velocity<br />

corresponds with the results by Schiffter [2006]. The absolute values are slightly higher<br />

than the values given here due to the higher SPL at Schiffter’s experiments. Kastner [2001]<br />

also measured faster evaporation using a direct <strong>drying</strong> air stream.


Results <strong>and</strong> Discussion 91<br />

(a)<br />

(b)<br />

(c)<br />

(d)<br />

(e)<br />

Figure 5.27: Experimental values for the evaporation coefficient at (a) 25 °C (c) 40 °C<br />

<strong>and</strong> (e) 60 °C <strong>and</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> at (b) 25 °C (d) 40 °C <strong>and</strong> (f) 60 °C<br />

for 0.0 <strong>and</strong> 0.91 m/s <strong>drying</strong> air velocity in comparison to the calculated evaporation<br />

coefficient (using the wet-bulb <strong>temperature</strong>) <strong>and</strong> T(wb)<br />

(f)


Results <strong>and</strong> Discussion 92<br />

To quantify the deviation from diffusion controlled evaporation in the experiments the<br />

Sherwood number is calculated. It is calculated using the diffusion model (Diff). For the<br />

experiments with a direct <strong>drying</strong> air stream the Ranz-Marshall (RM) model for forced<br />

convection is also used [Ranz <strong>and</strong> Marshall 1952a; Ranz <strong>and</strong> Marshall 1952b]:<br />

Equation 5.8<br />

2 0.6 · ⁄ · ⁄<br />

<br />

The Reynolds number is calculated by [Baehr <strong>and</strong> Stephan 2004]:<br />

Equation 5.9<br />

·· <br />

<br />

using the <strong>droplet</strong> diameter (<strong>of</strong> a <strong>surface</strong> equivalent sphere) , the density <strong>of</strong> the air , the<br />

air velocity <strong>and</strong> the dynamic viscosity <strong>of</strong> the air . The Schmidt number is<br />

calculated using the dynamic viscosity <strong>of</strong> air , the binary diffusion coefficient <strong>and</strong> the<br />

density <strong>of</strong> the air by [Baehr <strong>and</strong> Stephan 2004]:<br />

Equation 5.10<br />

<br />

<br />

·<br />

Schiffter [2006] performed experiments using various <strong>drying</strong> air velocities <strong>and</strong> found that<br />

the Ranz-Marshall model is accurate for an air velocity <strong>of</strong> 1 m/s or more. The air velocity<br />

setting <strong>of</strong> 0.91 m/s in the present experiments is close to this setting, so that the comparison<br />

using the Ranz-Marshall model is assumed to be applicable in this case.<br />

The Sherwood number curves are given in Figure 5.28, Figure 5.29 <strong>and</strong> Figure<br />

5.30. The experiments without a direct <strong>drying</strong> air stream using the microsyringe show a<br />

Sherwood number at about 2, as well as in the experiments using the microdispenser head<br />

in 5.2.1.1. The experiments with a direct <strong>drying</strong> air stream have initial Sherwood numbers<br />

<strong>of</strong> about 4 - 6 calculated using the diffusion model, <strong>and</strong> about 6 calculated using the Ranz-<br />

Marshall model for forced convection. This shows the increasing evaporation rate due to<br />

the <strong>drying</strong> air stream by reducing vapour accumulation around the <strong>droplet</strong> in the outer<br />

acoustic streaming. The calculation using the diffusion model shows decreasing Sherwood<br />

numbers with increasing <strong>temperature</strong> <strong>and</strong> relative humidity. This means that the<br />

evaporation approximates diffusional rates. For measurements with <strong>drying</strong> air stream<br />

directed to the <strong>droplet</strong> the Ranz-Marshall model for forced convection is more applicable.


Results <strong>and</strong> Discussion 93<br />

(a)<br />

(b)<br />

Figure 5.28: Comparison <strong>of</strong> the Sherwood numbers <strong>of</strong> water <strong>droplet</strong>s calculated using<br />

the experimental <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> at (a) 25 °C <strong>and</strong> 1 % relative humidity<br />

<strong>and</strong> (b) 25 °C <strong>and</strong> 40 % relative humidity (Diff = diffusion model; RM = Ranz-<br />

Marshall model)


Results <strong>and</strong> Discussion 94<br />

(a)<br />

(b)<br />

Figure 5.29: Comparison <strong>of</strong> the Sherwood number <strong>of</strong> water <strong>droplet</strong>s calculated using<br />

the experimental <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> at (a) 40 °C <strong>and</strong> 1 % relative humidity<br />

<strong>and</strong> (b) 40 °C <strong>and</strong> 40 % relative humidity (Diff = diffusion model; RM = Ranz-<br />

Marshall model)


Results <strong>and</strong> Discussion 95<br />

(a)<br />

(b)<br />

Figure 5.30: Comparison <strong>of</strong> the Sherwood number <strong>of</strong> water <strong>droplet</strong>s calculated using<br />

the experimental <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> at (a) 60 °C <strong>and</strong> 1 % relative humidity<br />

<strong>and</strong> (b) 60 °C <strong>and</strong> 40 % relative humidity (Diff = diffusion model; RM = Ranz-<br />

Marshall model)


Results <strong>and</strong> Discussion 96<br />

5.2.2 Evaporation <strong>of</strong> pure organic solvent <strong>droplet</strong>s<br />

The experiments were performed at just one single <strong>drying</strong> air <strong>temperature</strong>, 50 °C, <strong>and</strong> 1 %<br />

relative humidity. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> for each organic solvent was<br />

additionally calculated using the method by Yarin et al. [1999]. The material properties<br />

<strong>and</strong> the coefficients for the formulas are given in Table 5.2 <strong>and</strong> Table 5.3. The calculation<br />

steps were already described in 3.7.1.<br />

The <strong>droplet</strong>s <strong>of</strong> 1 μl initial volume were inserted into the acoustic field using the<br />

microsyringe. Due to the fast evaporation <strong>of</strong> acetone, 2-butanone, ethyl acetate <strong>and</strong><br />

tetrahydr<strong>of</strong>uran the initial <strong>droplet</strong> volume for these solvents was set to 2 μl to have a<br />

sufficient test duration. The SPL was adjusted to the lowest possible setting <strong>and</strong> was<br />

between 162.8 - 165.5 dB dependent on the organic solvent. All experiments were<br />

performed without a direct <strong>drying</strong> air stream.<br />

Table 5.2: Properties <strong>of</strong> the organic solvents [Engineering Toolbox 2009; Fuller et al.<br />

1966; Yaws 2003]<br />

Substance name Molecular<br />

Molecular<br />

Boiling<br />

Density at<br />

Diffusion<br />

formula<br />

weight<br />

point T b<br />

20 °C<br />

volume<br />

[g/mol]<br />

[K]<br />

[kg/m³]<br />

(Fuller)<br />

Acetone C 3 H 6 O 58.08 329.44 791 66.86<br />

2-Butanone C 4 H 8 O 72.11 352.79 805 87.32<br />

Dichloromethane CH 2 Cl 2 84.93 312.90 1326 59.46<br />

Ethanol C 2 H 6 O 46.07 351.44 789 50.36<br />

Ethyl acetate C 4 H 8 O 2 88.11 350.21 901 92.80<br />

Methanol CH 4 O 32.04 337.85 791 29.90<br />

2-Propanol C 3 H 8 O 60.10 355.41 785 70.82<br />

Tetrahydr<strong>of</strong>uran C 4 H 8 O 72.11 338.00 888 67.12<br />

Water H 2 O 18.02 373.15 998 12.70


Results <strong>and</strong> Discussion 97<br />

Table 5.3: Material properties <strong>and</strong> coefficients for the calculation <strong>of</strong> the enthalpy <strong>of</strong><br />

vaporization [Yaws 2003]<br />

Substance name Enthalpy <strong>of</strong><br />

vaporization at<br />

T b [kJ/mol]<br />

Coefficients for the formula:<br />

· ⁄ <br />

A T c [K] n<br />

Acetone 29.79 49.24 508.2 0.481<br />

2-Butanone 31.22 50.65 535.5 0.450<br />

Dichloromethane 28.38 41.91 510.0 0.410<br />

Ethanol 39.40 43.12 516.3 0.079<br />

Ethyl acetate 32.23 49.35 523.3 0.385<br />

Methanol 35.14 52.72 512.6 0.377<br />

2-Propanol 39.87 58.98 508.3 0.326<br />

Tetrahydr<strong>of</strong>uran 30.26 44.44 540.2 0.391<br />

Experiments with pure dichloromethane could not be performed, because the <strong>droplet</strong>s<br />

explosively evaporated in the acoustic field, even when the levitator settings (variations <strong>of</strong><br />

the distance between transducer <strong>and</strong> reflector or different SPL settings) were changed. The<br />

boiling point <strong>of</strong> the substance is 39.8 °C, lower than the <strong>drying</strong> air <strong>temperature</strong>. In general<br />

it is possible to analyze the evaporation behaviour <strong>of</strong> <strong>droplet</strong>s at a <strong>temperature</strong> higher than<br />

their boiling point due to the cooling effect <strong>of</strong> the evaporation. Dichloromethane was<br />

analyzed as a component in solvent mixtures.<br />

The evaporation behaviour for each organic solvent averaged from three<br />

measurements is given in Figure 5.31 (acetone <strong>and</strong> 2-butanone), Figure 5.32 (ethanol <strong>and</strong><br />

ethyl acetate), Figure 5.33 (methanol <strong>and</strong> 2-propanol) <strong>and</strong> Figure 5.34 (tetrahydr<strong>of</strong>uran <strong>and</strong><br />

water). The aspect ratio is higher than for pure water <strong>droplet</strong>s <strong>and</strong> decreases towards<br />

sphericity at the end <strong>of</strong> the measurement. This is due to the higher <strong>surface</strong> tension <strong>of</strong> water<br />

in comparison to other solvents that leads to more spherical <strong>droplet</strong>s for water. The<br />

measured <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> is, as in the experiments with water, higher than the<br />

calculated values according to Yarin et al. [1999].


Results <strong>and</strong> Discussion 98<br />

This deviation is assumed to be caused by the energy supply <strong>of</strong> the acoustic field. The<br />

increase in <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> at the end <strong>of</strong> the measurement can also be seen. The<br />

measured evaporation coefficient is much higher than the calculated one using the<br />

diffusion model due to the influence <strong>of</strong> inner acoustic streaming that increases the<br />

evaporation rate. This behaviour was also seen for pure water <strong>droplet</strong>s.<br />

Experiments for the analysis <strong>of</strong> the evaporation behaviour <strong>of</strong> organic solvents using<br />

a IR-camera were already performed by Tuckermann et al. [2002; 2005]. They used a<br />

different <strong>drying</strong> air <strong>temperature</strong> <strong>of</strong> 21 °C <strong>and</strong> 35 - 40 % relative humidity. The organic<br />

solvents were alkanes <strong>and</strong> alkanols, but also dichloromethane was examined successfully<br />

at the lower <strong>temperature</strong>. The <strong>surface</strong> area decrease <strong>and</strong> the <strong>surface</strong> <strong>temperature</strong> were<br />

monitored. Most <strong>of</strong> the organic solvents show a linear decrease in <strong>surface</strong> area as for the<br />

radius in the present experiments at higher <strong>drying</strong> air <strong>temperature</strong>. However, some solvents<br />

show a break in the <strong>surface</strong> area decrease. Tuckermann et al. [2002; 2005] explained this<br />

effect by condensation <strong>of</strong> water vapour on the cold <strong>droplet</strong> <strong>surface</strong> at the experimental<br />

conditions <strong>of</strong> 35 - 40 % relative humidity. This leads to an increase in the water fraction in<br />

the <strong>droplet</strong>. The water now evaporates predominately in the second part <strong>of</strong> the curves with<br />

a slower evaporation rate. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curves confirm this. After a<br />

short <strong>surface</strong> <strong>temperature</strong> decrease in the equilibration time <strong>of</strong> the <strong>droplet</strong>, the <strong>droplet</strong><br />

<strong>surface</strong> <strong>temperature</strong> has an almost constant value <strong>and</strong> increases only at the end <strong>of</strong> the<br />

measurement.<br />

In the case <strong>of</strong> water condensation on the <strong>droplet</strong> the <strong>temperature</strong> increases, when<br />

water predominately starts to evaporate <strong>and</strong> is constant for the water evaporation time. The<br />

equilibration time at the beginning <strong>of</strong> the measurement described by Tuckermann cannot<br />

be seen in the experiments at 50 °C. It might take place in the time <strong>of</strong> the adjustment <strong>of</strong> the<br />

acoustic field before the measurement could be started. Water condensation may be absent<br />

at a relative humidity <strong>of</strong> 1 % in the 50 °C experiments. Tuckermann et al. [2002; 2005] did<br />

not present experiments at higher <strong>drying</strong> air <strong>temperature</strong>s or lower relative humidity,<br />

therefore the absolute values cannot be compared.


Results <strong>and</strong> Discussion 99<br />

(a)<br />

(b)<br />

Figure 5.31: Evaporation behaviour <strong>of</strong> (a) acetone <strong>and</strong> (b) 2-butanone <strong>droplet</strong>s,<br />

<strong>drying</strong> air conditions: 50 °C <strong>and</strong> 1 % relative humidity


Results <strong>and</strong> Discussion 100<br />

(a)<br />

(b)<br />

Figure 5.32: Evaporation behaviour <strong>of</strong> (a) ethanol <strong>and</strong> (b) ethyl acetate <strong>droplet</strong>s,<br />

<strong>drying</strong> air conditions: 50 °C <strong>and</strong> 1 % relative humidity


Results <strong>and</strong> Discussion 101<br />

(a)<br />

(b)<br />

Figure 5.33: Evaporation behaviour <strong>of</strong> (a) methanol <strong>and</strong> (b) 2-propanol <strong>droplet</strong>s,<br />

<strong>drying</strong> air conditions: 50 °C <strong>and</strong> 1 % relative humidity


Results <strong>and</strong> Discussion 102<br />

(a)<br />

(b)<br />

Figure 5.34: Drying behaviour <strong>of</strong> (a) tetrahydr<strong>of</strong>uran <strong>and</strong> (b) water <strong>droplet</strong>s, <strong>drying</strong><br />

air conditions: 50 °C <strong>and</strong> 1 % relative humidity


Results <strong>and</strong> Discussion 103<br />

Figure 5.35 shows the comparison <strong>of</strong> the evaporation coefficients <strong>and</strong> the experimental <strong>and</strong><br />

predicted <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> versus the boiling point <strong>of</strong> the substances. The<br />

evaporation coefficient has no clear dependence on the boiling point, but the tendency is<br />

that for a higher boiling point the evaporation coefficient decreases. This can be due to<br />

different influence <strong>of</strong> inner acoustic streaming <strong>and</strong> to <strong>droplet</strong> oscillations. The solvents<br />

needed different adjustments <strong>of</strong> the SPL to provide less oscillations <strong>and</strong> an acceptable<br />

sphericity. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> increases clearly with increasing boiling point<br />

<strong>and</strong> the calculated <strong>temperature</strong> values are always lower than the experimental <strong>droplet</strong><br />

<strong>surface</strong> <strong>temperature</strong>s.<br />

(a)<br />

Figure 5.35: (a) Evaporation coefficient <strong>and</strong> (b) <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>of</strong> the<br />

organic solvents ordered by the boiling point (squares: experimental values, triangles:<br />

calculated values), <strong>drying</strong> air conditions: 50 °C <strong>and</strong> 1 % relative humidity<br />

5.2.3 Evaporation <strong>of</strong> <strong>droplet</strong>s containing solvent mixtures<br />

The evaporation <strong>of</strong> solvent mixtures was analyzed for the combination <strong>of</strong><br />

ethanol / dichloromethane (Figure 5.36), ethanol / ethyl acetate (Figure 5.37) <strong>and</strong><br />

ethanol / tetrahydr<strong>of</strong>uran (Figure 5.38) at 50 °C, 1 % relative humidity <strong>and</strong> an initial<br />

<strong>droplet</strong> volume <strong>of</strong> 2 μl. The weight proportions <strong>of</strong> ethanol-second solvent <strong>of</strong> 10:90, 25:75,<br />

50:50, 75:25 <strong>and</strong> 90:10 were analyzed at a SPL <strong>of</strong> about 165.8 dB for<br />

ethanol / dichloromethane <strong>and</strong> 163.3 dB for ethanol / ethyl acetate or tetrahydr<strong>of</strong>uran. For<br />

all solvent mixtures with a high ethanol content <strong>droplet</strong> oscillations occur that can be seen<br />

in the /0²-graphs.<br />

(b)


Results <strong>and</strong> Discussion 104<br />

(a)<br />

(b)<br />

Figure 5.36: (a) r(t)²/r(0)² <strong>and</strong> (b) <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curves <strong>of</strong> solvent<br />

mixture <strong>droplet</strong>s composed by ethanol <strong>and</strong> dichloromethane, <strong>drying</strong> air conditions:<br />

50 °C <strong>and</strong> 1 % relative humidity


Results <strong>and</strong> Discussion 105<br />

(a)<br />

(b)<br />

Figure 5.37: (a) r(t)²/r(0)² <strong>and</strong> (b) <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curves <strong>of</strong> solvent<br />

mixture <strong>droplet</strong>s composed by ethanol <strong>and</strong> ethyl acetate, <strong>drying</strong> air condition: 50 °C<br />

<strong>and</strong> 1 % relative humidity


Results <strong>and</strong> Discussion 106<br />

(a)<br />

(b)<br />

Figure 5.38: (a) r(t)²/r(0)² <strong>and</strong> (b) <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curves <strong>of</strong> solvent<br />

mixture <strong>droplet</strong>s composed by ethanol <strong>and</strong> tetrahydr<strong>of</strong>uran, <strong>drying</strong> air conditions:<br />

50 °C <strong>and</strong> 1 % relative humidity


Results <strong>and</strong> Discussion 107<br />

Dichloromethane has a very low boiling point in comparison to ethanol leading to very fast<br />

evaporation. For very small amounts <strong>of</strong> dichloromethane its evaporation cannot be seen in<br />

the graph, because the solvent has evaporated before the SPL could be finally adjusted <strong>and</strong><br />

the measurement started. For a higher dichloromethane content a break in the curve where<br />

the ethanol starts to dominate the evaporation process can clearly be seen.<br />

The graphs for ethanol / dichloromethane do not show a sequence dependent on the<br />

mass content <strong>of</strong> the solvents. This might be due to the very fast evaporation <strong>of</strong> low<br />

amounts <strong>of</strong> dichloromethane that cannot be exactly determined. The trend is that the<br />

evaporation is faster at a higher dichloromethane content. The <strong>temperature</strong> curves for high<br />

ethanol content do not show an increase at the beginning <strong>of</strong> the measurement due to<br />

dichloromethane evaporation. This confirms the theory <strong>of</strong> measurement inaccuracy by fast<br />

evaporation <strong>of</strong> small amounts <strong>of</strong> dichloromethane. For higher dichloromethane content a<br />

decrease in the initial <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> occurs as expected.<br />

Ethyl acetate <strong>and</strong> ethanol neither show clear breaks in the /0²-curves nor a<br />

clear dependency <strong>of</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> on the solvent ratio. A dependency <strong>of</strong><br />

the /0²-graph on the mixture ratio <strong>of</strong> this solvents can be seen. For higher ethanol<br />

content the evaporation slows down, though ethyl acetate has only a slightly lower boiling<br />

point. The curves for the ethanol / tetrahydr<strong>of</strong>uran mixtures also show breaks in the<br />

/0²-graph with slower evaporation for higher ethanol content. Here the increase in<br />

<strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> with increasing ethanol content can clearly be seen.<br />

The aspect ratio curves in Figure 5.39, Figure 5.40 <strong>and</strong> Figure 5.41 show clearly the<br />

increasing <strong>droplet</strong> oscillations with increasing ethanol content for all solvent mixtures.<br />

Most flattened <strong>droplet</strong>s occur for the ethanol / dichloromethane mixtures with an initial<br />

aspect ratio <strong>of</strong> 1.5 - 2.0 (without considering the oscillations). Ethanol in combination with<br />

ethyl acetate <strong>and</strong> tetrahydr<strong>of</strong>uran <strong>droplet</strong>s are more spherical having an initial aspect ratio<br />

<strong>of</strong> about 1.5. For all mixture samples the aspect ratio decreases towards the end <strong>of</strong> the<br />

measurement.<br />

Schiffter [2006] performed experiments for ethanol / water mixtures at 40 °C <strong>and</strong><br />

60 °C. For the evaporation coefficient he found an increase with decreasing water content.<br />

Calculation <strong>of</strong> the evaporation coefficient in the first <strong>and</strong> last part <strong>of</strong> the measurement<br />

shows that the terminal evaporation coefficient corresponds for all mixture ratios well with<br />

the value for pure water. At the beginning the evaporation coefficient is higher with higher<br />

ethanol content.


Results <strong>and</strong> Discussion 108<br />

Figure 5.39: Aspect ratio curves <strong>of</strong> solvent mixture <strong>droplet</strong>s composed by ethanol <strong>and</strong><br />

dichloromethane, <strong>drying</strong> air conditions: 50 °C <strong>and</strong> 1 % relative humidity<br />

Figure 5.40: Aspect ratio curves <strong>of</strong> solvent mixture <strong>droplet</strong>s composed by ethanol <strong>and</strong><br />

ethyl acetate, <strong>drying</strong> air conditions: 50 °C <strong>and</strong> 1 % relative humidity


Results <strong>and</strong> Discussion 109<br />

Figure 5.41: Aspect ratio curves <strong>of</strong> solvent mixture <strong>droplet</strong>s composed by ethanol <strong>and</strong><br />

tetrahydr<strong>of</strong>uran, <strong>drying</strong> air conditions: 50 °C <strong>and</strong> 1 % relative humidity<br />

This phenomenon can be also seen in the present experiments, where the steeper curves<br />

flatten to the end <strong>of</strong> the measurement. The relation for ethanol is inverse, because ethanol<br />

is the solvent having the highest boiling point in comparison to the experiments with water<br />

by Schiffter [2006].<br />

Tuckermann et al. [2005] analyzed ideal <strong>and</strong> azeotropic solvent mixtures. For<br />

water / methanol as an almost ideal mixture they found a decreasing <strong>surface</strong> <strong>temperature</strong> <strong>of</strong><br />

the evaporation <strong>droplet</strong> with increasing methanol content. Water / propanol mixtures show<br />

a <strong>temperature</strong> minimum for middle mixture ratios <strong>and</strong> maximum <strong>surface</strong> <strong>temperature</strong> for<br />

the pure solvents. This is due to the azeotropic mixture <strong>and</strong> corresponds closely with the<br />

boiling points <strong>of</strong> the solvent mixture ratios. The mixture ethanol / dichloromethane has also<br />

azeotropic behaviour. The azeotropic mixture ratio is 0.09:0.91. This cannot be determined<br />

by the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> in the experiments at 50 °C, because the pure<br />

dichloromethane that should consequently have a higher <strong>surface</strong> <strong>temperature</strong> as the<br />

ethanol / dichloromethane 10:90 mixture was not measureable under this condition.


Results <strong>and</strong> Discussion 110<br />

5.3 Evaporation <strong>of</strong> excipient solution <strong>droplet</strong>s<br />

5.3.1 Trehalose solution <strong>droplet</strong>s<br />

Trehalose (α,α-trehalose) is a disaccharide formed by two 1→1 linked α,α units <strong>of</strong><br />

glucopyranose [Gil et al. 1996]. The bond is via the reducing groups, so trehalose itself is a<br />

non-reducing sugar. It is a natural substance that can be found in some insects, plants <strong>and</strong><br />

microorganisms [Richards et al. 2002]. In periods <strong>of</strong> desiccation trehalose has a protective<br />

effect on the cells. Trehalose has the molecular formula C 12 H 22 O 11 <strong>and</strong> a molecular weight<br />

<strong>of</strong> 342.31 g/mol. It is commonly used as a dihydrate with a molecular weight <strong>of</strong><br />

378.33 g/mol. The melting point <strong>of</strong> trehalose dihydrate is 97 °C, where the water<br />

evaporates <strong>and</strong> the trehalose solidifies again at about 130 °C. The anhydrous trehalose<br />

melts at 203 °C [Richards et al. 2002]. The maximum solubility in water at 20 °C is 68.9 g<br />

in 100 g water [Higashiyama 2002]. Trehalose is used as an excipient in pharmaceutical<br />

spray <strong>drying</strong> formulation due to its protective effect on proteins in aqueous solution <strong>and</strong> in<br />

the dry product [Adler <strong>and</strong> Lee 1999]. In this chapter the <strong>drying</strong> behaviour <strong>of</strong> pure<br />

trehalose solutions is analyzed before subsequently in 5.5 the protective effect on the<br />

protein formulation is analyzed.<br />

Trehalose has an amorphous modification that affects its <strong>drying</strong> behaviour. Walton<br />

<strong>and</strong> Mumford [1999] defined in their study three distinct morphological types for powder<br />

samples: agglomerate, skin-forming <strong>and</strong> crystalline structure. If classified in this way, then<br />

trehalose is an example for skin-forming behaviour. Solutions <strong>of</strong> 15 % (w/w) trehalose in<br />

water p.a. were prepared <strong>and</strong> filtered using filters <strong>of</strong> 0.2 μm pores (Schleicher <strong>and</strong> Schuell,<br />

Dassel, Germany) before use. The <strong>drying</strong> conditions were 25 °C, 40 °C, 60 °C <strong>and</strong> 1 %,<br />

20 %, 40 % relative humidity <strong>and</strong> the experiments were performed without a direct <strong>drying</strong><br />

air stream. The starting <strong>droplet</strong>s had a volume <strong>of</strong> 1.5 μl <strong>and</strong> the initial SPL was between<br />

164.4 - 166.4 dB. All experiments were performed six times <strong>and</strong> the average curves were<br />

calculated. As in the water experiments, the /0²-curve, the aspect ratio <strong>and</strong> the<br />

<strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> were used for analysis <strong>of</strong> the <strong>drying</strong> behaviour <strong>and</strong> here also<br />

for the detection <strong>of</strong> the critical point.<br />

The graphs for measurement at 25 °C <strong>and</strong> 1 % relative humidity are given in Figure<br />

5.42 as an example. The /0²-graph shows a linear decrease <strong>of</strong> the radius squared<br />

until the critical point appears where a crust forms. The transition is not a sharp break, but


Results <strong>and</strong> Discussion 111<br />

changes over a long time forming a bend in the curve. The initial aspect ratio is about 1.2<br />

<strong>and</strong> almost constant in this phase. At the critical point the <strong>droplet</strong> starts to flatten <strong>and</strong><br />

slowly the aspect ratio increases to about 2.0 at the end <strong>of</strong> the <strong>drying</strong> process. Therefore the<br />

aspect ratio curve is useful to determine the critical point in the <strong>drying</strong> process. The <strong>droplet</strong><br />

<strong>surface</strong> <strong>temperature</strong> decreases at the beginning <strong>of</strong> measurement to a constant value in the<br />

first <strong>drying</strong> stage. At the critical point it starts to increase <strong>and</strong> approaches the <strong>drying</strong> air<br />

<strong>temperature</strong>. Also at low <strong>drying</strong> air <strong>temperature</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curve is<br />

rounded <strong>and</strong> shows no sharp transition. This is due to the amorphous character <strong>of</strong> trehalose.<br />

When the water evaporates it starts to form a film on the <strong>surface</strong> <strong>of</strong> the <strong>droplet</strong> in a rubberlike<br />

state which is deformable. Afterwards the solid glassy state is reached. This behaviour<br />

avoids sharp transition points in the curves.<br />

Figure 5.42: Drying behaviour <strong>of</strong> a trehalose 15 % (w/w) solution <strong>droplet</strong> at 25 °C<br />

<strong>and</strong> 1 % relative humidity<br />

The comparison <strong>of</strong> the curves for different <strong>drying</strong> air conditions in Figure 5.43, Figure 5.44<br />

<strong>and</strong> Figure 5.45 shows that this behaviour can be found for all settings. The <strong>drying</strong> time<br />

increases for increasing relative humidity due to a decrease in the water vapour pressure<br />

gradient between the <strong>droplet</strong> <strong>surface</strong> <strong>and</strong> the surrounding air. The critical point appears<br />

later in the /0²-graph as well as for the aspect ratio <strong>and</strong> the <strong>droplet</strong> <strong>surface</strong>


Results <strong>and</strong> Discussion 112<br />

<strong>temperature</strong> curves. The final particle size is similar under all <strong>drying</strong> air conditions <strong>and</strong><br />

also the aspect ratio shows no clear dependence on the <strong>drying</strong> air conditions.<br />

Figure 5.43: Evaporation behaviour <strong>of</strong> trehalose 15 % (w/w) solution <strong>droplet</strong>s at<br />

25 °C <strong>drying</strong> air <strong>temperature</strong><br />

Figure 5.44: Evaporation behaviour <strong>of</strong> trehalose 15 % (w/w) solution <strong>droplet</strong>s at<br />

40 °C <strong>drying</strong> air <strong>temperature</strong>


Results <strong>and</strong> Discussion 113<br />

Figure 5.45: Evaporation behaviour <strong>of</strong> trehalose 15 % (w/w) solution <strong>droplet</strong>s at<br />

60 °C <strong>drying</strong> air <strong>temperature</strong><br />

Schiffter [2006] <strong>and</strong> Schiffter <strong>and</strong> Lee [2007b] performed experiments using trehalose<br />

solutions without <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> measurement. They found similar results for<br />

the <strong>drying</strong> curves, but in the example for a solution <strong>of</strong> 100 mg/ml <strong>and</strong> 200 mg/ml trehalose<br />

content at 60 °C <strong>and</strong> 0.1 % / 5 % relative humidity the aspect ratio increased only to 1.5.<br />

Figure 5.46 gives a comparison <strong>of</strong> the evaporation coefficients <strong>and</strong> the <strong>droplet</strong><br />

<strong>surface</strong> <strong>temperature</strong>s in the first <strong>drying</strong> stage for all <strong>drying</strong> air conditions. As expected the<br />

evaporation coefficient increases with increasing <strong>drying</strong> air <strong>temperature</strong> <strong>and</strong> decreasing<br />

relative humidity. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> increases with increasing <strong>drying</strong> air<br />

<strong>temperature</strong> <strong>and</strong> increasing relative humidity.<br />

In comparison to water (5.2.1.1) the values for the evaporation coefficient <strong>and</strong> the<br />

<strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> are similar. The dissolved solid should lead to a depression <strong>of</strong><br />

the water vapour pressure <strong>and</strong> the <strong>droplet</strong> <strong>temperature</strong> should increase in comparison to the<br />

pure solvent [Masters 1991]. The evaporation coefficients for water at 25 °C are even<br />

lower than the values for the trehalose solution, but at higher <strong>drying</strong> air <strong>temperature</strong>s the<br />

values approximate <strong>and</strong> no evident deviation due to the vapour pressure lowering effect <strong>of</strong><br />

trehalose can be seen.


Results <strong>and</strong> Discussion 114<br />

(a)<br />

Figure 5.46: (a) Evaporation coefficients <strong>and</strong> (b) <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s <strong>of</strong> a<br />

15 % (w/w) trehalose solution <strong>droplet</strong> at different <strong>drying</strong> air conditions<br />

The evaporation rate for solution <strong>droplet</strong>s was calculated using the equations given in<br />

3.7.2. In the first <strong>drying</strong> stage the <strong>droplet</strong> shrinkage was used for the calculations, whereas<br />

in the second <strong>drying</strong> stage the vertical position <strong>of</strong> the mass centre <strong>of</strong> the <strong>droplet</strong> / particle to<br />

the next upper pressure node was measured. The problem here is that small <strong>droplet</strong><br />

oscillations have a large influence on the vertical <strong>droplet</strong> position measurement. This leads<br />

to fluctuations, so that at the end <strong>of</strong> the second <strong>drying</strong> stage slightly negative evaporation<br />

rates are calculated.<br />

In spite <strong>of</strong> these inaccuracies, the <strong>drying</strong> stages can be clearly seen in the graphs<br />

given in Figure 5.47, Figure 5.48 <strong>and</strong> Figure 5.49 for different <strong>drying</strong> air <strong>temperature</strong> <strong>and</strong><br />

relative humidity. The first part <strong>of</strong> the curves shows a linear evaporation rate that increases<br />

somewhat towards the critical point. At the critical point the evaporation starts to decrease<br />

sharply because <strong>of</strong> the reduced evaporation due to formation <strong>of</strong> a <strong>surface</strong> layer on the<br />

<strong>droplet</strong>. The fluctuations at the end <strong>of</strong> the measurement are caused by slight oscillations <strong>of</strong><br />

the particle around its vertical position. The trehalose 15 % (w/w) solution <strong>droplet</strong>s show<br />

the highest initial evaporation rate at <strong>drying</strong> air conditions <strong>of</strong> 60 °C <strong>and</strong> 1 % relative<br />

humidity. The initial evaporation rate decreases for increasing relative humidity <strong>and</strong><br />

decreasing <strong>drying</strong> air <strong>temperature</strong>, as expected. Therefore, the lowest initial evaporation<br />

rate can be seen in the experiments at <strong>drying</strong> air conditions <strong>of</strong> 25 °C <strong>and</strong> 40 % relative<br />

humidity.<br />

(b)


Results <strong>and</strong> Discussion 115<br />

Figure 5.47: Evaporation rate <strong>of</strong> trehalose 15 % (w/w) solution <strong>droplet</strong>s at 1 %<br />

relative humidity<br />

Figure 5.48: Evaporation rate <strong>of</strong> trehalose 15 % (w/w) solution <strong>droplet</strong>s at 20 %<br />

relative humidity


Results <strong>and</strong> Discussion 116<br />

Figure 5.49: Evaporation rate <strong>of</strong> trehalose 15 % (w/w) solution <strong>droplet</strong>s at 40 %<br />

relative humidity<br />

The dried particles were removed from the acoustic field using the spoon net, stored in<br />

Eppendorf tubes, <strong>and</strong> analyzed by scanning electron microscopy. The SEM-pictures are<br />

given in Figure 5.50. The particles for all <strong>drying</strong> air conditions are flattened, corresponding<br />

to the aspect ratio curves. The <strong>surface</strong> is smooth <strong>and</strong> on the top side folded <strong>and</strong> wrinkled.<br />

This corresponds with the <strong>surface</strong> <strong>of</strong> spray dried trehalose particles which also have<br />

a smooth <strong>surface</strong>. The spray dried particles are spherical in contrast to the levitated ones. In<br />

spray <strong>drying</strong> the <strong>droplet</strong>s are allowed to rotate freely while <strong>drying</strong>. The levitated samples<br />

are flattened due to the influence <strong>of</strong> the levitation forces that do not occur in the spray<br />

<strong>drying</strong> process. Additionally in the levitation experiments the top side starts to collapse<br />

downwards in the second <strong>drying</strong> stage.<br />

This behaviour was also shown for particles produced from trehalose solutions <strong>of</strong><br />

100 mg/ml <strong>and</strong> 200 mg/ml at 60 °C <strong>and</strong> 5 % relative humidity [Schiffter 2006; Schiffter<br />

<strong>and</strong> Lee 2007b]. In the present experiments for the 15 % (w/w) solution <strong>droplet</strong>s no<br />

influence <strong>of</strong> the <strong>drying</strong> air conditions on the <strong>droplet</strong> shape or <strong>surface</strong> smoothness can be<br />

observed.


Results <strong>and</strong> Discussion 117<br />

(a) 25 °C, 20 % rel. humidity (100x)<br />

(b) 25 °C, 20 % rel. humidity (1000x)<br />

(c) 40 °C, 1 % rel. humidity (100x)<br />

(d) 40 °C, 1 % rel. humidity (1000x)<br />

(e) 60 °C, 20 % rel. humidity (100x)<br />

(f) spray dried at T in = 130 °C (3000x)<br />

Figure 5.50: SEM-pictures <strong>of</strong> trehalose particles, (a-e) levitated <strong>and</strong> (f) spray dried<br />

using a 15 % (w/w) trehalose solution at different <strong>drying</strong> air conditions


Results <strong>and</strong> Discussion 118<br />

In additional experiments the influence <strong>of</strong> the trehalose content <strong>of</strong> the solution was<br />

analyzed. The trehalose content was changed from 0 % to 40 % (w/w) <strong>and</strong> the solution<br />

<strong>droplet</strong>s were dried at 60 °C <strong>and</strong> 1 % relative humidity. The starting <strong>droplet</strong> volume was<br />

1.5 μl <strong>and</strong> the initial SPL was between 165.9 - 166.7 dB. Figure 5.51 <strong>and</strong> Figure 5.52 give<br />

the comparison <strong>of</strong> the resulting curves. The ²/0² -curve shows an earlier critical<br />

point with increasing trehalose content, <strong>and</strong> the particles are larger. The <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> curves correspond with the critical point <strong>of</strong> the radius squared curves.<br />

Figure 5.51: Drying behaviour <strong>of</strong> trehalose solution <strong>droplet</strong>s at 60 °C <strong>and</strong> 1 %<br />

relative humidity dependent on the trehalose content (r(t)²/r(0)² <strong>and</strong> <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong>)<br />

The <strong>temperature</strong> increase is sharper for a lower trehalose content. Droplets containing<br />

30 % <strong>and</strong> 40 % trehalose start to form a <strong>surface</strong> layer immediately after the <strong>droplet</strong> is<br />

brought into the acoustic field. Their <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> increase is slower in<br />

comparison to the less concentrated <strong>droplet</strong>s at their critical point. The more concentrated<br />

<strong>droplet</strong>s did not dry, though their <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>and</strong> aspect ratio did not<br />

change further. They stuck to the tube wall so that the SEM pictures cannot show the<br />

original shape. This may be due to fast formation <strong>of</strong> a <strong>surface</strong> layer that encloses moisture<br />

inside the particle even for long <strong>drying</strong> times. The aspect ratio (Figure 5.52) has a tendency<br />

to increase with increasing trehalose content. There are also more <strong>droplet</strong> oscillations at the<br />

critical point especially with the more concentrated samples.


Results <strong>and</strong> Discussion 119<br />

Figure 5.52: Drying behaviour <strong>of</strong> trehalose solution <strong>droplet</strong>s at 60 °C <strong>and</strong> 1 %<br />

relative humidity dependent on the trehalose content (aspect ratio)<br />

The comparison <strong>of</strong> the evaporation coefficients in the first <strong>drying</strong> stage (Figure 5.53)<br />

shows an initial increase for increasing solids content until the trehalose content produces<br />

reduced evaporation due to rapid crust formation at 30 % <strong>and</strong> 40 % (w/w) trehalose<br />

content.<br />

Figure 5.53: Evaporation coefficients <strong>and</strong> <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s dependent on<br />

the trehalose content <strong>of</strong> the solutions, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 1 % relative<br />

humidity


Results <strong>and</strong> Discussion 120<br />

These values have to be interpreted with circumspection because <strong>of</strong> <strong>droplet</strong> oscillations<br />

even at the beginning <strong>of</strong> the <strong>drying</strong> process. The initial <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> does<br />

not change much. There is a tendency to a higher initial <strong>surface</strong> <strong>temperature</strong> at 30 % <strong>and</strong><br />

40 % trehalose content, but the difference <strong>of</strong> 1 °C is in the range <strong>of</strong> the measurement<br />

inaccuracy. Masters [1991] predicts a vapour pressure lowering effect <strong>of</strong> the dissolved<br />

solids. In experiments with calcium chloride solutions <strong>of</strong> different solids content an<br />

increase in salt concentration led to a decrease in the evaporation rate [Charlesworth <strong>and</strong><br />

Marshall 1960]. Unfortunately, this relation cannot be shown for the complete<br />

concentration range in these experiments.<br />

The SEM-pictures for trehalose particles from 5 %, 10 %, 15 % <strong>and</strong> 20 % (w/w)<br />

trehalose content are given in Figure 5.54.<br />

(a) 5 % (w/w) trehalose (100x)<br />

(b) 10 % (w/w) trehalose (100x)<br />

(c) 15 % (w/w) trehalose (100x)<br />

(d) 20 % (w/w) trehalose (100x)<br />

Figure 5.54: SEM-pictures <strong>of</strong> trehalose particles dried using solutions <strong>of</strong> different<br />

trehalose content, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 1 % relative humidity


Results <strong>and</strong> Discussion 121<br />

For all initial trehalose concentrations the <strong>surface</strong> <strong>of</strong> the particle is smooth <strong>and</strong> the particle<br />

flattened, they just differ in size due to their trehalose content. Schiffter [2006] performed<br />

experiments at different trehalose content. Schiffter achieved similar evaporation rate<br />

curves, which also show no clear first <strong>drying</strong> stage at high trehalose content (400 mg/ml).<br />

With increasing trehalose content the evaporation rate in the first <strong>drying</strong> stage decreases<br />

due to the amount <strong>of</strong> dissolved molecules <strong>and</strong> the related water vapour pressure depression<br />

[Schiffter 2006]. Schiffter also calculated the evaporation coefficients for different<br />

trehalose content <strong>and</strong> found a continuous decrease with increasing trehalose content. This<br />

effect cannot clearly be seen in the evaporation coefficients <strong>of</strong> the present experiments.<br />

However, here the evaporation coefficients increase slightly for low trehalose content, but<br />

decrease for high trehalose content as expected. In summary, the trehalose experiments<br />

show that for the <strong>determination</strong> <strong>of</strong> the critical point in the evaporation experiments it is<br />

useful to measure the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> in addition to the <strong>droplet</strong> radius <strong>and</strong> the<br />

aspect ratio. The results regarding particle shape <strong>and</strong> <strong>surface</strong> structure are comparable to<br />

spray <strong>drying</strong> experiments.<br />

5.3.2 Mannitol solution <strong>droplet</strong>s<br />

Mannitol is a sugar alcohol with the molecular formula C 6 H 14 O 6 <strong>and</strong> a molecular weight <strong>of</strong><br />

182.17 g/mol. The melting point <strong>of</strong> D-mannitol is 167 - 170 °C <strong>and</strong> its maximum solubility<br />

in water at 20 °C is 182 g/l [Sigma-Aldrich 2009]. D-Mannitol is a common excipient in<br />

freeze <strong>and</strong> spray <strong>drying</strong>. Its chief advantage is chemical stability [Yu et al. 1998]. D-<br />

mannitol has a strong tendency to crystallize, but also exists in a fully or partially<br />

amorphous state in certain formulations [Yu et al. 1998]. It is used in this work as a model<br />

for the <strong>drying</strong> behaviour <strong>of</strong> a crystallizing substance according to the morphological types<br />

given by Walton <strong>and</strong> Mumford [1999].<br />

A mannitol 15 % (w/w) solution in water p.a. was prepared <strong>and</strong> filtered using filters<br />

<strong>of</strong> 0.2 μm pores. The <strong>drying</strong> air conditions were 25 °C, 40 °C, 60 °C <strong>and</strong> 1 %, 20 % <strong>and</strong><br />

40 % relative humidity without a direct <strong>drying</strong> air stream. The <strong>droplet</strong>s were inserted in the<br />

pressure node using a microsyringe <strong>and</strong> had an initial <strong>droplet</strong> volume <strong>of</strong> 1.5 μl. The SPL<br />

was set to the lowest level that gives sufficient <strong>droplet</strong> stability. In the experiments it was<br />

between 162.2 - 167.9 dB depending on the <strong>drying</strong> air conditions. The experiments were<br />

performed six times <strong>and</strong> the average curves were calculated. Figure 5.55 gives an example


Results <strong>and</strong> Discussion 122<br />

for the curve progression <strong>of</strong> <strong>drying</strong> mannitol solution <strong>droplet</strong>s at 25 °C <strong>and</strong> 1 % relative<br />

humidity. In contrast to the trehalose experiments the transition from the first in the second<br />

<strong>drying</strong> stage at the critical point is very steep <strong>and</strong> clearly visible in the ²/0² -graph<br />

<strong>and</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curve. The aspect ratio increases after the critical point<br />

to 1.3. This increase is not as high as for the trehalose solution <strong>droplet</strong>s that have a final<br />

aspect ratio <strong>of</strong> up to 2.0. The mannitol particles are therefore more spherical than the<br />

trehalose particles.<br />

Figure 5.55: Drying behaviour <strong>of</strong> a mannitol 15 % (w/w) solution <strong>droplet</strong> at 25 °C<br />

<strong>and</strong> 1 % relative humidity<br />

A comparison <strong>of</strong> the curves for different <strong>drying</strong> air conditions in Figure 5.56, Figure 5.57<br />

<strong>and</strong> Figure 5.58 shows that with increasing <strong>drying</strong> air <strong>temperature</strong> the transition to the<br />

second <strong>drying</strong> stage at the critical point becomes sharper for the <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> <strong>and</strong> the /0²-curves. The aspect ratio is constant after the critical point<br />

or decreases at high <strong>drying</strong> air <strong>temperature</strong>. Due to the crystallization <strong>of</strong> mannitol the<br />

particle formation is faster than for amorphous substances like trehalose leading to the<br />

sharp breaks seen in the mannitol curves. The <strong>determination</strong> <strong>of</strong> the critical point in the<br />

<strong>drying</strong> curves is therefore easier for the crystallizing substance with sharp transition than<br />

for trehalose with longer transition times. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> is especially at


Results <strong>and</strong> Discussion 123<br />

higher <strong>drying</strong> air <strong>temperature</strong>s a useful tool for the <strong>determination</strong> <strong>of</strong> the critical point due<br />

to the fast increase in <strong>temperature</strong>.<br />

Figure 5.56: Drying behaviour <strong>of</strong> mannitol 15 % (w/w) solution <strong>droplet</strong>s at 25 °C<br />

<strong>drying</strong> air <strong>temperature</strong><br />

Figure 5.57: Drying behaviour <strong>of</strong> mannitol 15 % (w/w) solution <strong>droplet</strong>s at 40° <strong>drying</strong><br />

air <strong>temperature</strong>


Results <strong>and</strong> Discussion 124<br />

Figure 5.58: Drying behaviour <strong>of</strong> mannitol 15 % (w/w) solution <strong>droplet</strong>s at 60 °C<br />

<strong>drying</strong> air <strong>temperature</strong><br />

The values for the evaporation coefficients <strong>and</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s are given<br />

in Figure 5.59. The evaporation coefficients at 40 °C / 60 °C <strong>and</strong> low relative humidity are<br />

higher for mannitol in comparison to trehalose. At higher relative humidity <strong>and</strong> for the<br />

25 °C-measurements the values correspond to the trehalose experiments. The <strong>droplet</strong><br />

<strong>surface</strong> <strong>temperature</strong> is similar for <strong>droplet</strong>s <strong>of</strong> both solutions.<br />

(a)<br />

Figure 5.59: (a) Evaporation coefficients <strong>and</strong> (b) <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s <strong>of</strong><br />

mannitol 15 % (w/w) solution <strong>droplet</strong>s at different <strong>drying</strong> air conditions<br />

(b)


Results <strong>and</strong> Discussion 125<br />

Due to the lower molecular weight <strong>of</strong> mannitol the 15 % (w/w) solution contains more<br />

molecules than a trehalose 15 % (w/w) solution. This should lead to higher vapour pressure<br />

depression for mannitol <strong>and</strong> lower evaporation coefficients. However, the results do not<br />

correspond with this assumption, especially for high <strong>drying</strong> air <strong>temperature</strong>s. Schiffter<br />

[2006] performed mannitol experiments with a 100 mg/ml mannitol solution. The results<br />

were higher evaporation rates for mannitol than for trehalose or maltodextrin solutions.<br />

The evaporation rate curves <strong>of</strong> mannitol 15 % (w/w) solutions at different <strong>drying</strong><br />

air conditions are given in Figure 5.60, Figure 5.61 <strong>and</strong> Figure 5.62. As for the trehalose<br />

experiments the <strong>drying</strong> rate increases with increasing <strong>drying</strong> air <strong>temperature</strong> <strong>and</strong><br />

decreasing relative humidity. The problem that occurs in the calculation <strong>of</strong> the evaporation<br />

rate was the precise measurement <strong>of</strong> the vertical <strong>droplet</strong> position. At the critical point the<br />

<strong>droplet</strong>s <strong>of</strong>ten started to oscillate <strong>and</strong> the <strong>determination</strong> <strong>of</strong> the vertical position <strong>of</strong> the<br />

<strong>droplet</strong> or particle was imprecise or even impossible. The oscillations lead to fluctuations<br />

in the evaporation rate curves, especially at the critical point <strong>and</strong> in the second <strong>drying</strong><br />

stage. This has to be considered for curve interpretation.<br />

Figure 5.60: Evaporation rate <strong>of</strong> mannitol 15 % (w/w) solution <strong>droplet</strong>s at 1 %<br />

relative humidity


Results <strong>and</strong> Discussion 126<br />

Figure 5.61: Evaporation rate <strong>of</strong> mannitol 15 % (w/w) solution <strong>droplet</strong>s at 20 %<br />

relative humidity<br />

Figure 5.62: Evaporation rate <strong>of</strong> mannitol 15 % (w/w) solution <strong>droplet</strong>s at 40 %<br />

relative humidity


Results <strong>and</strong> Discussion 127<br />

The initial evaporation rates are similar for the mannitol <strong>and</strong> the trehalose experiments.<br />

The values <strong>of</strong> the initial evaporation rates are in the range <strong>of</strong> 0.2 - 1.4 μg/s/mm² depending<br />

on the <strong>drying</strong> air conditions. The experiments at 40 °C <strong>and</strong> 1 % relative humidity are an<br />

exception. Here the evaporation rate <strong>of</strong> the mannitol solution is slightly higher. The curve<br />

progression for mannitol solutions at the critical point cannot be compared to the trehalose<br />

curves, because <strong>droplet</strong> oscillations deform the curves. Especially at 40 % relative<br />

humidity the graphs show strong fluctuations.<br />

Schiffter [2006] found a sharp break at the critical point <strong>of</strong> the evaporation rate<br />

curves for solutions <strong>of</strong> 100 mg/ml mannitol, but he also had a problem with oscillating<br />

<strong>droplet</strong>s at the critical point. The sharp breaks in the evaporation rates from his<br />

experiments correspond with the /0²-graphs <strong>and</strong> <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong><br />

curves <strong>of</strong> the present measurements.<br />

The particle shape was analyzed by scanning electron microscopy <strong>of</strong> the dried<br />

particle. Several examples for different <strong>drying</strong> air conditions are presented in Figure 5.63.<br />

The particles dried at low <strong>temperature</strong> are more spherical than the trehalose particles,<br />

leading to a lower aspect ratio. The <strong>surface</strong> is rough due to the crystallizing properties <strong>of</strong><br />

mannitol, <strong>and</strong> the particles have “blow holes”. At higher <strong>temperature</strong> <strong>and</strong> relative humidity<br />

the particles are also mostly spherical, but both rough <strong>and</strong> smoothed particle <strong>surface</strong>s can<br />

be found. In some experiments at 60 °C <strong>and</strong> 40 % relative humidity the <strong>droplet</strong>s did not<br />

form a closed spherical particle, but rather a “bowl” that is open at the top. The spray dried<br />

particles <strong>of</strong> a mannitol 15 % (w/w) solution are more spherical <strong>and</strong> seem to have a<br />

smoother <strong>surface</strong> than most <strong>of</strong> the levitated particles. Blow holes can also be seen in the<br />

powder particles.<br />

The SEM-pictures correspond with the pictures given by Schiffter [2006] <strong>and</strong><br />

Schiffter <strong>and</strong> Lee [2007b] for a lower concentrated solution <strong>of</strong> 100 mg/ml mannitol. The<br />

<strong>surface</strong> shape is rough for low <strong>drying</strong> air <strong>temperature</strong> <strong>and</strong> smoother at higher <strong>drying</strong> air<br />

<strong>temperature</strong>. The particle shape is also dependent on the relative humidity <strong>of</strong> the <strong>drying</strong> air.<br />

At low relative humidity the particles were smoother <strong>and</strong> rougher for high relative<br />

humidity. In the present experiments a clear dependency <strong>of</strong> the <strong>surface</strong> shape on the<br />

relative humidity <strong>of</strong> the <strong>drying</strong> air could not be observed.


Results <strong>and</strong> Discussion 128<br />

(a) 25°C, 1% rel. humidity (75x)<br />

(b) 25°C, 1% rel. humidity (750x)<br />

(c) 40 °C, 20 % rel. humidity (75x)<br />

(d) 60 °C, 40 % rel. humidity (100x)<br />

(e) 60 °C, 40 % rel. humidity (100x)<br />

(f) spray dried T in = 130 °C (3000x)<br />

Figure 5.63: SEM-pictures <strong>of</strong> (a-e) levitated <strong>and</strong> (f) spray dried mannitol 15 % (w/w)<br />

solution particles at different <strong>drying</strong> air conditions


Results <strong>and</strong> Discussion 129<br />

Experiments at different mannitol concentrations were performed. Due to the low<br />

maximum solubility in water the concentration range <strong>of</strong> 0 - 15 % (w/w) mannitol content<br />

was chosen. The inserted <strong>droplet</strong> volume was 1.0 μl <strong>and</strong> the initial SPL was between<br />

163.1 - 167.6 dB depending on the mannitol content. The <strong>drying</strong> air <strong>temperature</strong> was 60 °C<br />

<strong>and</strong> 1 % relative humidity. The curves for the radius squared <strong>and</strong> the <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> are given in Figure 5.64. For all mannitol contents the break at the critical<br />

point is very sharp <strong>and</strong> corresponds for the /0²-graphs <strong>and</strong> the <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> curves. Due to the higher solids content the critical point appears earlier in the<br />

curves for higher concentration. The initial <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> is almost identical<br />

at 30 °C for all mannitol concentrations.<br />

Figure 5.64: Drying behaviour <strong>of</strong> mannitol solution <strong>droplet</strong>s at 60 °C <strong>and</strong> 1 % relative<br />

humidity dependent on the solids content (r(t)²/r(0)² <strong>and</strong> <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>)<br />

The aspect ratio curves (Figure 5.65) stay constant or even decrease at the critical point, so<br />

that the aspect ratio graphs are not always useful for the detection <strong>of</strong> the critical point. A<br />

comparison <strong>of</strong> the evaporation rates (Figure 5.66) for higher mannitol content shows a<br />

decrease in the initial evaporation rate. This corresponds with the assumption <strong>of</strong> a slower<br />

evaporation due to an increasing number <strong>of</strong> dissolved molecules. The critical point appears<br />

earlier in the evaporation curves for the higher concentrated samples.


Results <strong>and</strong> Discussion 130<br />

Figure 5.65: Drying behaviour <strong>of</strong> mannitol solution <strong>droplet</strong>s at 60 °C <strong>and</strong> 1 % relative<br />

humidity dependent on the solids content (aspect ratio)<br />

Figure 5.66: Drying behaviour <strong>of</strong> mannitol solution <strong>droplet</strong>s at 60 °C <strong>and</strong> 1 % relative<br />

humidity dependent on the solids content (evaporation rate)<br />

The unusual curves <strong>of</strong> the 5 % <strong>and</strong> especially for the 1 % (w/w) mannitol solution <strong>droplet</strong>s<br />

can be explained by <strong>droplet</strong> oscillations. For these concentrations oscillations occured even<br />

at the beginning <strong>of</strong> the measurement leading to measurement inaccuracies.


Results <strong>and</strong> Discussion 131<br />

The SEM-pictures <strong>of</strong> the particles produced using solutions <strong>of</strong> different mannitol content<br />

are given in Figure 5.67 <strong>and</strong> Figure 5.68. The particles at low mannitol content have a<br />

smooth <strong>surface</strong> <strong>and</strong> a relatively large hole in the middle <strong>of</strong> the particle. The picture at<br />

larger magnification shows a star-like pattern on the <strong>surface</strong> that shows crystalline<br />

influence during the <strong>drying</strong> process.<br />

The particles are irregular, more so than the trehalose particles. At higher mannitol<br />

content the particles start to become rougher on the <strong>surface</strong>. In the present experiments this<br />

behaviour can be found at a concentration ≥ 12.5 % (w/w). Schiffter [2006] found rough<br />

particles for a 100 mg/ml mannitol solution at 60 °C <strong>and</strong> 20 % relative humidity.<br />

(a) 1 % (w/w) mannitol (100x)<br />

(b) 5 % (w/w) mannitol (100x)<br />

(c) 5 % (w/w) mannitol (1500x)<br />

(d) 7.5 % (w/w) mannitol (100x)<br />

Figure 5.67: SEM-pictures <strong>of</strong> mannitol particles dried using solutions <strong>of</strong> different<br />

mannitol content, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 1 % relative humidity


Results <strong>and</strong> Discussion 132<br />

(a) 10 % (w/w) mannitol (100x)<br />

(b) 12.5 % (w/w) mannitol (100x)<br />

(c) 15 % (w/w) mannitol (75x)<br />

(d) 15 % (w/w) mannitol (1000x)<br />

Figure 5.68: SEM-pictures <strong>of</strong> mannitol particles dried using solutions <strong>of</strong> different<br />

mannitol content, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 1 % relative humidity<br />

The comparison <strong>of</strong> spray dried samples <strong>of</strong> 5 % <strong>and</strong> 15 % (w/w) mannitol content solutions<br />

(T in = 130 °C) is given in Figure 5.69. The SEM-pictures show little difference in particle<br />

shape for the mannitol solutions, but the particles from the lower concentrated solution are<br />

smaller. The spray dried particles show blow holes as seen with the levitated samples <strong>and</strong><br />

seem to have a smooth <strong>surface</strong>. The <strong>surface</strong> structure cannot clearly be seen, even in the<br />

picture at the highest magnification from the 15 % (w/w) mannitol solution particles in<br />

Figure 5.63 (f). Here at a magnification <strong>of</strong> 3000x the <strong>surface</strong> also seems to be smooth, <strong>and</strong><br />

the blow holes can be seen more clearly. For the spray dried mannitol particles no clear<br />

dependence <strong>of</strong> the <strong>surface</strong> structure on the mannitol content, as found for the levitated<br />

mannitol particles, could be observed in the present experiments.


Results <strong>and</strong> Discussion 133<br />

(a) 5 % (w/w) mannitol (2000x)<br />

(b) 15 % (w/w) mannitol (2000x)<br />

Figure 5.69: SEM-pictures <strong>of</strong> spray dried mannitol using solutions <strong>of</strong> different<br />

mannitol content, T in = 130 °C<br />

5.3.3 Sucrose solution <strong>droplet</strong>s<br />

Sucrose was analyzed as a substance with high solubility in water in order to achieve a<br />

high solids content in the sample solutions. The solubility <strong>of</strong> sucrose in water is ca.<br />

200 g/100 ml at 20 °C <strong>and</strong> therefore higher than the solubility <strong>of</strong> trehalose or mannitol.<br />

Sucrose is a disaccharide <strong>of</strong> D-fructose <strong>and</strong> D-glucose connected at their reducing groups.<br />

Therefore sucrose is a non-reducing sugar with molecular formula C 12 H 22 O 11 <strong>and</strong><br />

molecular weight 342.30 g/mol. The melting point is 185 - 187 °C [Sigma-Aldrich 2009].<br />

Solutions <strong>of</strong> 20 % - 60 % (w/w) sucrose content in water p.a. were prepared <strong>and</strong><br />

filtered as before. The <strong>droplet</strong>s <strong>of</strong> 1.5 μl were inserted into the pressure node using the<br />

microsyringe. The initial SPL was between 164.5 dB <strong>and</strong> 165.0 dB. The <strong>drying</strong> air<br />

conditions were 60 °C <strong>and</strong> 1 % relative humidity without a direct <strong>drying</strong> air stream to the<br />

<strong>droplet</strong> <strong>and</strong> the experiments for each concentration were performed three times. The<br />

comparison <strong>of</strong> the /0 -curves, the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curves <strong>and</strong> the<br />

aspect ratio curves is given in Figure 5.70 <strong>and</strong> Figure 5.71. The curves show typical<br />

amorphous behaviour without a sharp transition at the critical point. This behaviour has<br />

already been explained for the trehalose solutions. For high concentrated sucrose solution<br />

<strong>droplet</strong>s the curve transition at the critical point is highly rounded. At sucrose<br />

concentrations ≥ 40 % (w/w) the first <strong>drying</strong> stage <strong>and</strong> the critical point is not clearly<br />

visible in the /0 -curves or in the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curves.


Results <strong>and</strong> Discussion 134<br />

Figure 5.70: Drying behaviour <strong>of</strong> sucrose solution <strong>droplet</strong>s at 60 °C <strong>and</strong> 1 % relative<br />

humidity dependent on the sucrose content (r(t)²/r(0)² <strong>and</strong> <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong>)<br />

Figure 5.71: Drying behaviour <strong>of</strong> sucrose solution <strong>droplet</strong>s at 60 °C <strong>and</strong> 1 % relative<br />

humidity dependent on the sucrose content (aspect ratio curves)<br />

The evaporation coefficients (Figure 5.72) are almost equal, except at sucrose contents <strong>of</strong><br />

50 % <strong>and</strong> 60 % (w/w). Here the <strong>droplet</strong>s formed a <strong>surface</strong> film immediately after insertion


Results <strong>and</strong> Discussion 135<br />

in the acoustic field leading to reduced evaporation. The initial <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong><br />

is also equal for low concentrations, but increases for 50 % <strong>and</strong> 60 % (w/w) sucrose<br />

content due to the fast formation <strong>of</strong> the <strong>surface</strong> film <strong>and</strong> therefore reduced evaporation<br />

cooling <strong>of</strong> the <strong>droplet</strong>. In the radius squared <strong>and</strong> <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curves the<br />

critical point appears later with decreasing sucrose content, as expected. The aspect ratio<br />

curves show that the particles, especially the low concentration samples, become more<br />

spherical during the <strong>drying</strong> time.<br />

Figure 5.72: Evaporation coefficients <strong>and</strong> <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s <strong>of</strong> sucrose<br />

solution <strong>droplet</strong>s at 60 °C <strong>and</strong> 1 % relative humidity dependent on the sucrose<br />

content<br />

The problem in the <strong>drying</strong> <strong>of</strong> sucrose solution <strong>droplet</strong>s was that for a high sucrose content<br />

the particles did not dry completely. The particles <strong>of</strong> low concentration have a smooth<br />

<strong>surface</strong> <strong>and</strong> were solid, whereas the high concentrated samples at 50 % <strong>and</strong> 60 % (w/w)<br />

sucrose content stayed sticky <strong>and</strong> deformable even after long <strong>drying</strong> times. This can also<br />

be seen in the SEM-pictures in Figure 5.73, where the 20 % (w/w) sucrose solution particle<br />

shows a smooth <strong>surface</strong>. The 30 % (w/w) sucrose solution particle has a popcorn-like<br />

shape because it burst during the preparation for the SEM analysis due to the residual<br />

water content. This happened to all particles produced from higher concentrated sample<br />

<strong>droplet</strong>s <strong>and</strong> is a good indication <strong>of</strong> not completely dry particles.


Results <strong>and</strong> Discussion 136<br />

(a) 20 % (w/w) sucrose (50x)<br />

(b) 30 % (w/w) sucrose (35x)<br />

Figure 5.73: SEM-pictures <strong>of</strong> sucrose particles dried using solutions <strong>of</strong> different<br />

sucrose content, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 1 % relative humidity<br />

Spray <strong>drying</strong> <strong>of</strong> pure sucrose solutions is difficult [Adler 1999], because the <strong>droplet</strong>s stick<br />

in the chamber <strong>and</strong> cyclone walls forming a glass-like layer. The glass transition<br />

<strong>temperature</strong> <strong>of</strong> pure sucrose is 75 °C [Hancock <strong>and</strong> Zografi 1997]. Considering the residual<br />

water content <strong>of</strong> the powder, the <strong>temperature</strong> <strong>of</strong> the cyclone wall is higher than the glass<br />

transition <strong>temperature</strong> <strong>of</strong> the powder [Adler 1999]. For this reason the behaviour <strong>of</strong> high<br />

concentrated sucrose solutions in the levitation experiments is not surprising. For high<br />

water content in the particles it is possible that the glass transition <strong>temperature</strong> is lower<br />

than the <strong>drying</strong> air conditions in the levitator. Even for the very wide concentration range<br />

<strong>of</strong> the sucrose solutions it was not possible to detect a difference in the initial evaporation<br />

rate or the initial <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> (except for immediate film building at high<br />

concentrations <strong>of</strong> 50 % <strong>and</strong> 60 % (w/w)). The experiments show again how important the<br />

additional measurement <strong>of</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> for the detection <strong>of</strong> the critical<br />

point is.<br />

5.3.4 Copolyvidone <strong>and</strong> HPMC solution <strong>droplet</strong>s<br />

Preliminary experiments for the itraconazole measurements in Chapter 5.4 examine the<br />

excipients copolyvidone <strong>and</strong> hydroxypropylmethylcellulose (HPMC) dissolved in a solvent<br />

mixture containing different ratios <strong>of</strong> ethanol <strong>and</strong> dichloromethane. The <strong>drying</strong><br />

experiments were performed using nitrogen as <strong>drying</strong> gas at a <strong>drying</strong> gas <strong>temperature</strong> <strong>of</strong><br />

50 °C <strong>and</strong> 0.5 % relative humidity without a <strong>drying</strong> air stream directed towards the <strong>droplet</strong>.<br />

The <strong>droplet</strong>s with an initial volume <strong>of</strong> 2 μl were inserted in the pressure node using the


Results <strong>and</strong> Discussion 137<br />

microsyringe. The SPL was set to the lowest possible value <strong>of</strong> 160.0 dB (copolyvidone)<br />

<strong>and</strong> 160.4 dB (HPMC). The solvent ratios <strong>of</strong> dichloromethane-ethanol were 58.8:41.2<br />

(start formulation), 70:30, 60:40, 50:50, 40:60 <strong>and</strong> 30:70. The polymer content for the pure<br />

excipient solution was 4.9 % (w/w) for each polymer.<br />

The <strong>drying</strong> behaviour <strong>of</strong> copolyvidone <strong>and</strong> HPMC in the start formulations is given<br />

in Figure 5.74 <strong>and</strong> Figure 5.75. The experiments were performed for a longer time than<br />

depicted, <strong>and</strong> the figure is a part <strong>of</strong> the <strong>drying</strong> behaviour in the first <strong>drying</strong> stage <strong>and</strong> the<br />

critical point. The copolyvidone solution <strong>droplet</strong>s show the common curve progression for<br />

an amorphous substance that was already found for trehalose <strong>and</strong> sucrose. The <strong>droplet</strong>s<br />

form a crust at the critical point that leads to an immediate decrease in evaporation,<br />

therefore the critical point is clearly visible in all three curves. In contrast to the<br />

copolyvidone solution the HPMC solution does not show a clear first <strong>drying</strong> stage in the<br />

<strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>and</strong> the aspect ratio curves. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong><br />

increases immediately after the <strong>droplet</strong> was placed into the pressure node. The aspect ratio<br />

shows fluctuations at the beginning <strong>of</strong> the measurement <strong>and</strong> then it increases slightly. The<br />

radius squared plot changes its slope, but the critical point cannot clearly be seen due to<br />

oscillations at the beginning <strong>of</strong> the measurement. For both excipient solutions the known<br />

<strong>drying</strong> behaviour <strong>of</strong> the pure binary solvent mixture cannot be seen in the graphs. The<br />

initial <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> is higher than the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>of</strong> the<br />

binary solvent mixture. Here the influence <strong>of</strong> the fast dichloromethane evaporation that<br />

leads to cooling to very low <strong>temperature</strong> <strong>of</strong> the <strong>droplet</strong> <strong>surface</strong> at the beginning <strong>of</strong> the<br />

measurement is strongly reduced by addition <strong>of</strong> the polymer. The <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> <strong>of</strong> the polymer solutions is more similar to the ethanol <strong>surface</strong> <strong>temperature</strong>.<br />

When placing the sample in the pressure node it could be observed for the HPMC<br />

solutions that immediately after the solution leaves the syringe a film appears on the<br />

<strong>droplet</strong> <strong>surface</strong>. This leads to the early increase in <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>and</strong> the<br />

differing <strong>drying</strong> behaviour in comparison to the copolyvidone solutions. In the first <strong>drying</strong><br />

stage strong <strong>droplet</strong> oscillations occur until the <strong>surface</strong> film is sufficiently solidified. This<br />

leads to strong measurement inaccuracies in the /0²-graph that make it impossible<br />

to determine the evaporation coefficients for the HPMC formulations. Neither different<br />

SPL settings nor different levitator adjustments were able to stop the oscillations <strong>and</strong> give<br />

more precise results.


Results <strong>and</strong> Discussion 138<br />

Figure 5.74: Drying behaviour <strong>of</strong> a copolyvidone start formulation <strong>droplet</strong> at 50 °C<br />

<strong>and</strong> 0.5 % relative humidity<br />

Figure 5.75: Drying behaviour <strong>of</strong> a HPMC start formulation <strong>droplet</strong> at 50 °C <strong>and</strong><br />

0.5 % relative humidity


Results <strong>and</strong> Discussion 139<br />

The comparison <strong>of</strong> the formulations at different solvent ratios is given in Figure 5.76 for<br />

copolyvidone <strong>and</strong> Figure 5.77 for HPMC. The copolyvidone <strong>temperature</strong> graphs show a<br />

tendency to a lower initial <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> for higher dichloromethane content<br />

due to its faster evaporation <strong>and</strong> therefore stronger cooling <strong>of</strong> the <strong>droplet</strong>. The critical point<br />

that can clearly be seen in the measured curves appears later for increasing ethanol content<br />

also because <strong>of</strong> the slower evaporation <strong>of</strong> ethanol in comparison to dichloromethane.<br />

The critical points in the radius squared curves correspond with the <strong>temperature</strong><br />

increases as well as with the aspect ratio increases. The graphs for the HPMC solution<br />

<strong>droplet</strong>s are not analyzable due to the strong <strong>droplet</strong> oscillations, but the curves show that<br />

the “critical point” must be early in the <strong>drying</strong> process in comparison to the copolyvidone<br />

solutions. Note that the end <strong>of</strong> the depicted graphs occurs after a third <strong>of</strong> the measurement,<br />

the first <strong>drying</strong> stage <strong>and</strong> the critical point <strong>of</strong> the different formulations are shown in detail.<br />

Also for the HPMC solutions the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> increases up to about 50 °C<br />

by the end <strong>of</strong> the measurement.<br />

Figure 5.76: Comparison <strong>of</strong> the <strong>drying</strong> behaviour <strong>of</strong> the copolyvidone formulations<br />

having different solvent ratios, <strong>drying</strong> gas conditions: 50 °C <strong>and</strong> 0.5 % relative<br />

humidity


Results <strong>and</strong> Discussion 140<br />

Figure 5.77: Comparison <strong>of</strong> the <strong>drying</strong> behaviour <strong>of</strong> the HPMC formulations having<br />

different solvent ratios, <strong>drying</strong> gas conditions: 50 °C <strong>and</strong> 0.5 % relative humidity<br />

The SEM-pictures in Figure 5.78 show the different particle shapes for both substances in<br />

the start formulation. The <strong>surface</strong> <strong>of</strong> the copolyvidone as well as for the HPMC particle is<br />

smooth as expected for amorphous substances. The particle shape is more crinkled for the<br />

HPMC particles than for the copolyvidone particles. The <strong>droplet</strong> movement for the HPMC<br />

solutions due to the crust collapsing leading to crinkled particles could be the reason for<br />

the strong <strong>droplet</strong> oscillations at the beginning <strong>of</strong> the measurements. The oscillations stop<br />

when the final particle size is reached.<br />

(a) copolyvidone start (100x)<br />

(b) HPMC start (100x)<br />

Figure 5.78: SEM-pictures <strong>of</strong> the start formulation particles containing copolyvidone<br />

or HPMC as polymer, <strong>drying</strong> gas conditions: 50 °C <strong>and</strong> 0.5 % relative humidity


Results <strong>and</strong> Discussion 141<br />

The particle shape is not dependent on the solvent ratio as the pictures for the start<br />

formulations <strong>and</strong> the 30:70 or 40:60 formulations for the polymers in Figure 5.79 show.<br />

SEM-pictures for the other solvent ratios (not shown) confirm this, they all look similar to<br />

the related start formulations.<br />

(a) copolyvidone 40:60 (100x)<br />

(b) copolyvidone 40:60 (1000x)<br />

(c) HPMC 30:70 (75x)<br />

(d) HPMC 30:70 (1000x)<br />

Figure 5.79: SEM-pictures <strong>of</strong> particles <strong>of</strong> different polymer formulations, <strong>drying</strong> gas<br />

conditions: 50 °C <strong>and</strong> 0.5 % relative humidity<br />

5.4 Itraconazole formulation experiments<br />

Itraconazole formulations containing as a polymer excipient either copolyvidone or HPMC<br />

were analyzed at different ratios <strong>of</strong> the solvents dichloromethane <strong>and</strong> ethanol. The aim <strong>of</strong><br />

this single <strong>droplet</strong> <strong>drying</strong> experiment is to find a dependence <strong>of</strong> the kind <strong>of</strong> polymer<br />

excipient in the formulation <strong>and</strong> the solvent ratio on the final particle shape. Previous spray<br />

<strong>drying</strong> experiments <strong>of</strong> different itraconazole formulations had shown different powder


Results <strong>and</strong> Discussion 142<br />

appearances. The information from the levitation experiments can be integrated in the<br />

development <strong>of</strong> an optimum formulation for spray <strong>drying</strong> <strong>of</strong> itraconazole.<br />

The formulations contained 3.2 % itraconazole, 4.8 % copolyvidone or HPMC <strong>and</strong><br />

92 % solvent mixture <strong>of</strong> different solvent ratio <strong>of</strong> dichloromethane <strong>and</strong> ethanol. The<br />

solvent ratios were 58.8:41.2 (start formulation), 70:30, 60:40, 50:50, 40:60 <strong>and</strong> 30:70, the<br />

same as for the pure excipient solutions. The <strong>drying</strong> gas was nitrogen <strong>and</strong> the experiments<br />

were performed at a <strong>drying</strong> air <strong>temperature</strong> <strong>of</strong> 50 °C <strong>and</strong> 0.5 % relative humidity without a<br />

direct <strong>drying</strong> air stream. The initial <strong>droplet</strong> volume was 2 μl, <strong>and</strong> the initial SPL, as set to<br />

the lowest setting for stable levitation, was 159.5 dB. The comparison <strong>of</strong> the <strong>drying</strong> curves<br />

for both start formulations is given in Figure 5.80 <strong>and</strong> Figure 5.81. The<br />

itraconazole / copolyvidone formulation shows similar behaviour to the pure copolyvidone<br />

formulation. The critical point can clearly be seen in the graphs for the radius squared, the<br />

<strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>and</strong> the aspect ratio. In contrast to the excipient formulation,<br />

the critical point appears earlier in the curves due to the higher solid content achieved by<br />

itraconazole addition. The particle stays more spherical, so that the aspect ratio curve is<br />

almost constant around 1.1, in contrast to the pure copolyvidone solutions.<br />

Figure 5.80: Drying behaviour <strong>of</strong> itraconazole / copolyvidone start formulation<br />

<strong>droplet</strong>s, <strong>drying</strong> gas conditions: 50 °C <strong>and</strong> 0.5 % relative humidity


Results <strong>and</strong> Discussion 143<br />

Figure 5.81: Drying behaviour <strong>of</strong> itraconazole / HPMC start formulation <strong>droplet</strong>s,<br />

<strong>drying</strong> gas conditions: 50 °C <strong>and</strong> 0.5 % relative humidity<br />

The itraconazole / HPMC formulation is also similar to its basic pure excipient<br />

formulation. Again the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> increases directly after the particle was<br />

inserted into the acoustic field. The radius squared <strong>and</strong> aspect ratio curves show <strong>droplet</strong><br />

oscillations, but much less than for the pure excipient solution. The critical point is not<br />

clearly detectable in the curves. The aspect ratio increases from 1.1 to about 1.3, in contrast<br />

to the formulation containing copolyvidone. Note that the particles were dried for a longer<br />

time than depicted.<br />

Figure 5.82 <strong>and</strong> Figure 5.83 show the comparison <strong>of</strong> the formulations having<br />

different solvent ratio. The itraconazole / copolyvidone formulations show that the critical<br />

point in the curves appears later for increasing amount <strong>of</strong> ethanol in the solvent mixture.<br />

This is due to its slower evaporation as already shown for the pure excipient solutions. The<br />

30:70 formulation needs nearly twice the <strong>drying</strong> time to the critical point than the 70:30<br />

solution. The solubility <strong>of</strong> itraconazole could have an opposite effect, because itraconazole<br />

is less soluble in ethanol (0.2 g/l) than in dichloromethane (250 g/l). But this influence<br />

cannot be seen in the <strong>drying</strong> time to the critical point. The critical point can be clearly seen<br />

in all measured curves for different solvent ratios. The curve progression is similar to the<br />

pure copolyvidone solutions, showing that the evaporation behaviour <strong>of</strong> the solution is<br />

dependent on the polymer <strong>and</strong> not on the itraconazole.


Results <strong>and</strong> Discussion 144<br />

Figure 5.82: Comparison <strong>of</strong> the <strong>drying</strong> behaviour <strong>of</strong> itraconazole / copolyvidone<br />

formulations, <strong>drying</strong> gas conditions: 50 °C <strong>and</strong> 0.5 % relative humidity at different<br />

solvent ratios<br />

Figure 5.83: Comparison <strong>of</strong> the <strong>drying</strong> behaviour <strong>of</strong> itraconazole / HPMC<br />

formulations, <strong>drying</strong> gas conditions: 50 °C <strong>and</strong> 0.5 % relative humidity at different<br />

solvent ratios


Results <strong>and</strong> Discussion 145<br />

Figure 5.84 gives the comparison <strong>of</strong> the evaporation coefficients depending on the ethanol<br />

content for the pure copolyvidone <strong>and</strong> the itraconazole / copolyvidone formulations. The<br />

absolute values for both experimental series are similar, but the pure copolyvidone values<br />

show more fluctuations. For both series the evaporation coefficient decreases clearly with<br />

higher ethanol content. The itraconazole / HPMC solutions show less oscillation in<br />

comparison to the pure HPMC solutions. Here the curves can be better compared, but still<br />

the calculation <strong>of</strong> the evaporation coefficients is imprecise. For none <strong>of</strong> the measured<br />

curves the critical point is detectable. The radius squared curves, the <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> <strong>and</strong> the aspect ratio have similar curve progressions for all solvent ratios. The<br />

absence <strong>of</strong> a clear critical point can again be explained by an immediate <strong>surface</strong> film<br />

forming on insertion <strong>of</strong> the solution <strong>droplet</strong> in the acoustic field.<br />

Figure 5.84: Evaporation coefficients <strong>of</strong> copolyvidone <strong>and</strong> itraconazole / copolyvidone<br />

formulations depending on the ethanol content, <strong>drying</strong> gas conditions: 50 °C <strong>and</strong><br />

0.5 % relative humidity<br />

The SEM-pictures given in Figure 5.85 show the particle shapes <strong>of</strong> the start formulation<br />

particles either containing copolyvidone or HPMC. The particle shape departs completely<br />

from that <strong>of</strong> the pure excipient solutions. The itraconazole / copolyvidone particle is almost<br />

spherical but much more crinkled on the <strong>surface</strong>. The itraconazole / HPMC particle is also<br />

more spherical than the pure excipient formulation, but has a less crinkled <strong>surface</strong>. For the<br />

HPMC formulation the addition <strong>of</strong> itraconazole leads to smoothing <strong>of</strong> the <strong>surface</strong>, whereas


Results <strong>and</strong> Discussion 146<br />

for the copolyvidone formulation the particle becomes more crinkled. For a higher ethanol<br />

ratio the particles are more crinkled than in the start formulations for both <strong>of</strong> the<br />

itraconazole / polymer formulations as shown Figure 5.86.<br />

(a) itraconazole / copolyvid. start (100x)<br />

(b) itraconazole / copolyvid. start (1000x)<br />

(c) itraconazole / HPMC start (100x)<br />

(d) itraconazole / HPMC start (1000x)<br />

Figure 5.85: SEM-pictures <strong>of</strong> start formulation particles <strong>of</strong> itraconazole with<br />

copolyvidone <strong>and</strong> HPMC, <strong>drying</strong> gas conditions: 50 °C <strong>and</strong> 0.5 % relative humidity<br />

The itraconazole / copolyvidone particle at a solvent ratio <strong>of</strong> 30:70 is more spherical than<br />

the particle produced from the HPMC formulation. The difference in the particle shape for<br />

the formulation containing HPMC is noticeable, where for high ethanol content the <strong>surface</strong><br />

structure changes from smooth to crinkled for the 40:60 <strong>and</strong> the 30:70 formulation. This<br />

can be due to the sample that changes for high ethanol content in both formulations to a<br />

suspension because the itraconazole is not sufficiently soluble in the solvent mixture due to<br />

the high ethanol ratio. In the case <strong>of</strong> the formulation containing HPMC <strong>and</strong> itraconazole


Results <strong>and</strong> Discussion 147<br />

the dispersed particles in the suspension may to lead to the crinkled <strong>surface</strong> in comparison<br />

to the solution formulations with higher dichloromethane content.<br />

(a) itraconazole / copolyvid. 30:70 (100x)<br />

(b) itraconazole / copolyvid. 30:70 (1000x)<br />

(c) itraconazole / HPMC 30:70 (100x)<br />

(d) itraconazole / HPMC 30:70 (1000x)<br />

(e) itraconazole / HPMC 50:50 (100x)<br />

(f) itraconazole / HPMC 50:50 (1000x)<br />

Figure 5.86: SEM-pictures <strong>of</strong> itraconazole / HPMC <strong>and</strong> itraconazole / copolyvidone<br />

formulation particles, <strong>drying</strong> gas conditions: 50 °C <strong>and</strong> 0.5 % relative humidity


Results <strong>and</strong> Discussion 148<br />

As a result the itraconazole formulations containing copolyvidone or HPMC show a<br />

completely different <strong>drying</strong> behaviour. For the copolyvidone formulation the dependence<br />

<strong>of</strong> the <strong>drying</strong> time to the critical point on the solvent mixture is clearly visible. The <strong>drying</strong><br />

rate is dependent on the polymer <strong>and</strong> not on the itraconazole, shown by the similar curve<br />

progression for both the pure copolyvidone <strong>and</strong> the itraconazole / copolyvidone graphs. In<br />

contrast the final particle shape differs between the pure excipient <strong>and</strong> the<br />

itraconazole / excipient formulations <strong>and</strong> also between the solutions <strong>and</strong> suspensions for<br />

itraconazole / HPMC. Here the itraconazole has a strong influence on the particle <strong>surface</strong><br />

appearance.<br />

5.5 Drying <strong>of</strong> protein solutions<br />

5.5.1 Bovine carbonic anhydrase (bCA) <strong>droplet</strong>s<br />

The aim is to analyze the <strong>drying</strong> behaviour, the particle formation <strong>and</strong> the final particle<br />

shape <strong>of</strong> different bCA solution formulations. Any protein stabilization in the <strong>drying</strong><br />

process due to trehalose addition is analyzed via measurement <strong>of</strong> the residual activity <strong>of</strong><br />

bCA. The results <strong>of</strong> formulations with different trehalose content are compared using<br />

levitated particles as well as spray dried powders. In further experiments the <strong>kinetics</strong> <strong>of</strong><br />

protein inactivation in the <strong>drying</strong> process are examined by removing the <strong>droplet</strong> or particle<br />

from the acoustic field at different times. The characterisation <strong>of</strong> bCA is given in 4.1.1.1.<br />

Trehalose is able to stabilize proteins in solution, during the <strong>drying</strong> process <strong>and</strong> in<br />

the dry product. The three main hypotheses for the protection mechanisms <strong>of</strong> trehalose are<br />

water-replacement, water-layer (or preferential exclusion), <strong>and</strong> mechanical-entrapment (or<br />

glass immobilisation) [Lins et al. 2004]. The water-replacement hypothesis attributes the<br />

stabilisation <strong>of</strong> the protein to direct interaction via hydrogen bonds between the protein <strong>and</strong><br />

the hydroxyl-groups <strong>of</strong> trehalose [Allison et al. 1999; Arakawa et al. 2001]. The waterlayer<br />

hypothesis assumes that a layer <strong>of</strong> water molecules is trapped close to the protein<br />

<strong>surface</strong> due to the preferential exclusion <strong>of</strong> trehalose from the vicinity <strong>of</strong> the protein<br />

[Belton <strong>and</strong> Gil 1994]. Trehalose is an amorphous substance that builds up a glass state<br />

during the <strong>drying</strong> process that reduces protein unfolding due to mechanical entrapment<br />

[Hagen et al. 1995; Lins et al. 2004].<br />

The bCA <strong>and</strong> trehalose were dissolved in 50 mM trizma-buffer pH 7.5 <strong>and</strong> stored on ice<br />

until the measurement begins. The experiments were performed first with solutions <strong>of</strong> bCA


Results <strong>and</strong> Discussion 149<br />

<strong>and</strong> trehalose total content <strong>of</strong> 10 % (w/v). Further experiments used a bCA content <strong>of</strong><br />

10 % (w/v) plus trehalose to provide either a constant total solids content or a constant<br />

protein content. The experiments were performed at <strong>drying</strong> air conditions <strong>of</strong> 60 °C <strong>and</strong><br />

10 % relative humidity without a direct <strong>drying</strong> air stream. The initial <strong>droplet</strong> volume was<br />

1.5 μl, the initial SPL was about 167.0 dB, <strong>and</strong> the experiments were performed six times<br />

each. Experiments with bCA <strong>and</strong> also other tested proteins are much more difficult than<br />

the excipient experiments using sugars or polymers. The <strong>droplet</strong>s start to oscillate strongly<br />

<strong>and</strong> at / after the critical point the particles <strong>of</strong>ten break <strong>and</strong> fall out <strong>of</strong> the acoustic field so<br />

that the experiment has to be repeated. The measurement <strong>of</strong> the distance to the next upper<br />

pressure node for the calculation <strong>of</strong> the evaporation rate is therefore not possible due to<br />

strong oscillations.<br />

Figure 5.87 shows the <strong>drying</strong> behaviour <strong>of</strong> a 10 % (w/v) bCA solution. Both the<br />

/0²-curve <strong>and</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curve show the critical point in the<br />

<strong>drying</strong> process as a curve, as already seen with amorphous substances. The aspect ratio is<br />

about 1.4, decreases before the critical point, <strong>and</strong> rises again to 1.4 for the final particle.<br />

The aspect ratio is here not useful for clear detection <strong>of</strong> the critical point.<br />

Figure 5.87: Drying behaviour <strong>of</strong> a <strong>droplet</strong> containing bCA 10 % (w/v) in 50 mM<br />

trizma buffer pH 7.5, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 % relative humidity


Results <strong>and</strong> Discussion 150<br />

The SEM-picture <strong>of</strong> a 10 % (w/v) bCA solution particle is given in Figure 5.88, as well as<br />

a SEM-picture <strong>of</strong> spray dried powder particles using a 10 % (w/v) bCA solution. In both<br />

cases the bCA particles have a smooth <strong>surface</strong> <strong>and</strong> a hollow interior. The pure levitated<br />

protein particles are very fragile <strong>and</strong> <strong>of</strong>ten break after the critical point or during storage in<br />

Eppendorf-tubes before SEM-pictures can be taken. Even a small amount <strong>of</strong> trehalose in<br />

the formulation (20 % <strong>of</strong> total solids) leads to much better levitation behaviour <strong>and</strong> storage<br />

stability <strong>of</strong> the particles. Schiffter <strong>and</strong> Lee [2007b] found also hollow particles for pure<br />

catalase as levitated particles <strong>and</strong> as a spray dried formulation. This particle shape seems to<br />

be common for pure or buffered aqueous protein solutions <strong>and</strong> will also be shown for pure<br />

LDH particles in 5.5.2.<br />

(a) 60 °C, 10 % rel. humidity (100x)<br />

(b) spray dried T in = 130 °C (3000x)<br />

Figure 5.88: SEM-pictures <strong>of</strong> (a) a levitated particle <strong>and</strong> (b) the spray dried powder<br />

<strong>of</strong> bCA 10 % (w/v) in 50 mM trizma buffer pH 7.5<br />

A trehalose solution 10 % (w/v) in 50 mM trizma buffer pH 7.5 was also dried in the<br />

levitator at 60°C <strong>and</strong> 10% relative humidity <strong>and</strong> in the spray dryer at T in = 130 °C. The<br />

SEM-pictures <strong>of</strong> a levitated particle <strong>and</strong> <strong>of</strong> the spray dried powder are given in Figure 5.89.<br />

Both the levitated <strong>and</strong> the spray dried particles show similar particle shape compared to the<br />

experiments using trehalose in pure water in 5.3.1. The levitated particle is (as for the<br />

trehalose in pure water) not completely spherical <strong>and</strong> has a smooth <strong>surface</strong>, but it is not as<br />

much flattened as the trehalose in water particle. The spray dried particles are almost<br />

spherical.


Results <strong>and</strong> Discussion 151<br />

(a) 60 °C, 10 % rel. humidity (100x)<br />

(b) spray dried T in = 130 °C (3000x)<br />

Figure 5.89: SEM-pictures <strong>of</strong> (a) a levitated particle <strong>and</strong> (b) the spray dried powder<br />

<strong>of</strong> trehalose 10 % (w/v) in 50 mM trizma buffer pH 7.5<br />

5.5.1.1 bCA-trehalose 10 % (w/v) total solids formulations<br />

In the first set <strong>of</strong> experiments trehalose was added to the bCA solutions achieving a total<br />

solids’ content <strong>of</strong> 10 % (w/v) (except buffer salts). The bCA-trehalose ratios used are 8:2,<br />

6:4, 4:6 <strong>and</strong> 2:8. Figure 5.90 shows the comparison <strong>of</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>and</strong><br />

the /0²-curves for the different formulations. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> is<br />

little influenced by the amount <strong>of</strong> trehalose in the mixtures. For higher trehalose content<br />

the <strong>temperature</strong> increase is slightly shifted to later <strong>drying</strong> times. The /0²-curves<br />

show a similar slope at the beginning. The critical point tends to appear also later for<br />

higher trehalose content. The aspect ratio <strong>of</strong> the <strong>droplet</strong> is about 1.3 - 1.4 at the beginning<br />

<strong>of</strong> the measurement <strong>and</strong> increases after the critical point as shown in Figure 5.91. For high<br />

bCA content the increase cannot clearly be seen, but with increasing trehalose content the<br />

aspect ratio increases up to about 2.4 for the 2:8 formulation. Yet pure trehalose<br />

10 % (w/v) solution <strong>droplet</strong>s do not show such an increase in the aspect ratio after the<br />

critical point.<br />

The dried particles were removed from the acoustic field to determine both enzyme<br />

activity assay <strong>and</strong> to take SEM-pictures. The total <strong>drying</strong> time was set to 563 s.<br />

Additionally, the protein formulations were spray dried at 130 °C inlet <strong>temperature</strong> to<br />

produce powders for comparison <strong>of</strong> the residual activity <strong>of</strong> bCA <strong>and</strong> <strong>of</strong> the particle shape.<br />

Figure 5.92 shows the comparison <strong>of</strong> the residual activity <strong>of</strong> the bCA formulations after the<br />

<strong>drying</strong> process in the levitator or spray dryer.


Results <strong>and</strong> Discussion 152<br />

Figure 5.90: Comparison <strong>of</strong> the <strong>drying</strong> behaviour <strong>of</strong> bCA-trehalose 10 % (w/v) total<br />

solids formulation <strong>droplet</strong>s, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 % relative humidity<br />

(r(t)²/r(0)² <strong>and</strong> <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>)<br />

Figure 5.91: Comparison <strong>of</strong> the <strong>drying</strong> behaviour <strong>of</strong> bCA-trehalose 10 % (w/v) total<br />

solids formulation <strong>droplet</strong>s, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 % relative humidity<br />

(aspect ratio)


Results <strong>and</strong> Discussion 153<br />

For a trehalose portion <strong>of</strong> ≥ 40 % <strong>of</strong> the total solids the residual activity for both the<br />

levitated <strong>and</strong> the spray dried samples is about 100 %. With ≤ 20 % trehalose a lower<br />

residual activity for the levitated samples <strong>and</strong> the pure bCA 10 % (w/v) solution is clearly<br />

shown. In spite <strong>of</strong> the scatter in the measured activity values this figure shows the<br />

stabilizing effect <strong>of</strong> even small amounts <strong>of</strong> trehalose on bCA.<br />

The similarity between the levitation <strong>and</strong> spray <strong>drying</strong> experiments is remarkable.<br />

During spray <strong>drying</strong> the <strong>droplet</strong>s are exposed to a much higher <strong>drying</strong> air <strong>temperature</strong>.<br />

Additionally the <strong>droplet</strong> <strong>surface</strong> area, where proteins can adsorb <strong>and</strong> unfold [Adler <strong>and</strong> Lee<br />

1999], is much higher in spray <strong>drying</strong> (related to the same sample volume) due to the<br />

smaller <strong>droplet</strong> size. This was assumed to lead to a higher protein deactivation during the<br />

spray <strong>drying</strong> process. However, the results compared to the levitation experiments are<br />

similar, possibly because <strong>of</strong> the longer <strong>drying</strong> time in the levitator. An influence <strong>of</strong> the<br />

ultrasonic field was excluded by Weis <strong>and</strong> Nardozzi [2005], who found similar enzyme<br />

<strong>kinetics</strong> in a bulk protein solution <strong>and</strong> in levitated protein <strong>droplet</strong>s. Residual activity<br />

experiments using catalase <strong>and</strong> trehalose were performed by Schiffter [2006]. Schiffter<br />

achieved sufficient protein stability at a trehalose content <strong>of</strong> 20 % <strong>of</strong> total solids.<br />

Figure 5.92: Residual activity <strong>of</strong> bCA depending on the trehalose content <strong>of</strong> the<br />

formulations in comparison <strong>of</strong> levitated <strong>and</strong> spray dried samples (10 % (w/v) total<br />

solids)


Results <strong>and</strong> Discussion 154<br />

The SEM-pictures <strong>of</strong> the levitated particles <strong>and</strong> the spray dried powders are given in Figure<br />

5.93 <strong>and</strong> Figure 5.94. All <strong>of</strong> the levitated particles show a smooth <strong>and</strong> less crinkled <strong>surface</strong>.<br />

The 8:2 sample has, as already shown for the pure bCA particles, a hole at the top <strong>and</strong><br />

shows its hollow interior. For higher trehalose content the holes in the particles disappear<br />

<strong>and</strong> the particles are more flattened. This relation was already shown in the aspect ratio<br />

curves in Figure 5.91. However, the pure trehalose particle (Figure 5.89 (a)) is more<br />

spherical than the particles having high trehalose content. The spray dried powders show<br />

crinkled particles for all bCA formulations. Particles with a high amount <strong>of</strong> bCA show<br />

fewer but deeper dimples than the particles with high trehalose content. In comparison, the<br />

pure trehalose powder particles are nearly spherical (Figure 5.89 (b)). The dimples seen in<br />

the spray dried powder particles cannot directly be found in the levitated particles.<br />

(a) bCA-trehalose 10 % 8:2 (100x)<br />

(b) bCA-trehalose 10 % 8:2 (3000x)<br />

(c) bCA-trehalose 10 % 6:4 (100x)<br />

(d) bCA-trehalose 10 % 6:4 (3000x)<br />

Figure 5.93: SEM-pictures <strong>of</strong> the (a, c) levitated <strong>and</strong> (b, d) spray dried bCA-trehalose<br />

10 % (w/v) total solids formulations, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 % relative<br />

humidity


Results <strong>and</strong> Discussion 155<br />

(a) bCA-trehalose 10 % 4:6 (100x)<br />

(b) bCA-trehalose 10 % 4:6 (3000x)<br />

(c) bCA-trehalose 10 % 2:8 (100x)<br />

(d) bCA-trehalose 10 % 2:8 (3000x)<br />

Figure 5.94: SEM-pictures <strong>of</strong> the (a, c) levitated <strong>and</strong> (b, d) spray dried bCA-trehalose<br />

10 % (w/v) total solids formulations, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 % relative<br />

humidity<br />

A similarity between the levitated particle samples <strong>and</strong> the spray dried powder is the<br />

inclination <strong>of</strong> the particle top side that can be seen in the 6:4 <strong>and</strong> 2:8 formulation particles.<br />

Spray <strong>drying</strong> experiments with catalase [Schiffter 2006] led to similar results for the<br />

change <strong>of</strong> the particle shape. With increasing trehalose content the particle <strong>surface</strong> is more<br />

crinkled.<br />

The period in which protein inactivation takes place was analyzed by experiments<br />

with different <strong>drying</strong> times between 0 - 563 s. Before the critical point the <strong>droplet</strong>s were<br />

collected directly into the PCR tubes, whereas after the critical point the particles were<br />

removed using a spoon net. The experiments were performed for all bCA-trehalose<br />

formulations at 6 different points <strong>of</strong> time <strong>and</strong> the results are given in Figure 5.95. The 6:4,<br />

4:6 <strong>and</strong> 2:8 formulations do not show any substantial decrease in the residual protein


Results <strong>and</strong> Discussion 156<br />

activity during the <strong>drying</strong> process. The 8:2 formulation <strong>and</strong> pure bCA 10 % (w/v) show a<br />

decrease in residual activity towards the end <strong>of</strong> the measurement. The critical points for<br />

these two solutions appear at about 265 s <strong>drying</strong> time. The residual activity results show<br />

that inactivation takes place predominantly after the critical point.<br />

Protein inactivation during the <strong>drying</strong> process may take place mostly via <strong>surface</strong><br />

adsorption <strong>of</strong> the proteins to the <strong>droplet</strong> <strong>surface</strong> <strong>and</strong> also the influence <strong>of</strong> <strong>temperature</strong> [Maa<br />

<strong>and</strong> Hsu 1997; Mumenthaler et al. 1994]. Surface adsorption <strong>of</strong> the protein will take place<br />

before the critical point, but the holes <strong>and</strong> the hollow interior <strong>of</strong> the particles having high<br />

bCA content show an increase in the <strong>surface</strong> area at crust formation. Therefore a larger<br />

area for <strong>surface</strong> adsorption <strong>of</strong> the protein is available after the critical point. The<br />

<strong>temperature</strong> <strong>of</strong> the <strong>droplet</strong> or particle increases after the critical point leading to higher<br />

thermal stress on the protein especially when it is still able to unfold in the liquid parts <strong>of</strong><br />

the <strong>droplet</strong> / particle. The results show therefore that single <strong>droplet</strong> <strong>drying</strong> experiments are<br />

a helpful method to examine the <strong>drying</strong> process <strong>and</strong> its influence on residual protein<br />

activity.<br />

Figure 5.95: Residual activity <strong>of</strong> bCA depending on the <strong>drying</strong> time for the bCAtrehalose<br />

10 % (w/v) total solids formulations, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 %<br />

relative humidity


Results <strong>and</strong> Discussion 157<br />

5.5.1.2 bCA 10 % (w/v) plus trehalose formulations<br />

The protein stabilizing experiments were repeated using a constant bCA 10 % (w/v)<br />

content plus trehalose in the ratios <strong>of</strong> 10:10, 10:5 <strong>and</strong> 10:3.3 with a 50 mM trizma buffer<br />

pH 7.5 as solvent. The advantage <strong>of</strong> these experiments is a constant bCA content in the<br />

protein activity assays. The <strong>drying</strong> behaviour <strong>of</strong> this formulations is compared in Figure<br />

5.96 <strong>and</strong> Figure 5.97. The /0²-curves for all mixtures have critical points that lie<br />

close together, though the solution <strong>droplet</strong>s have different solids content. In the <strong>droplet</strong><br />

<strong>temperature</strong> curves the critical point can clearly be seen. These curves show an earlier<br />

<strong>temperature</strong> increase for a higher trehalose content <strong>and</strong> therefore for a higher solids<br />

content, as expected. For formulations with high trehalose content the <strong>temperature</strong> increase<br />

is prolonged. However, pure trehalose <strong>droplet</strong>s have a faster <strong>temperature</strong> increase in<br />

comparison to the 10:5 <strong>and</strong> 10:10 mixtures. The formulations having higher trehalose<br />

content have a higher aspect ratio than the pure protein or trehalose <strong>droplet</strong>s, but no clear<br />

dependency can be seen.<br />

Figure 5.96: Comparison <strong>of</strong> the <strong>drying</strong> behaviour <strong>of</strong> bCA 10 % (w/v) plus trehalose<br />

formulation <strong>droplet</strong>s, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 % relative humidity<br />

(r(t)²/r(0)² <strong>and</strong> <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>)<br />

The residual protein activity <strong>of</strong> the formulation particles does not substantially change with<br />

increasing trehalose content as shown in Figure 5.98. The residual activity <strong>of</strong> the pure bCA<br />

particles is much lower than that <strong>of</strong> the stabilized formulations. The results for the


Results <strong>and</strong> Discussion 158<br />

levitation <strong>and</strong> the spray <strong>drying</strong> experiments are again similar to the bCA-trehalose 10 %<br />

total solids formulations.<br />

Figure 5.97: Comparison <strong>of</strong> the <strong>drying</strong> behaviour <strong>of</strong> bCA 10 % (w/v) plus trehalose<br />

formulation <strong>droplet</strong>s dried at 60 °C <strong>and</strong> 10 % relative humidity (aspect ratio)<br />

Figure 5.98: Residual activity <strong>of</strong> bCA depending on the trehalose content in<br />

comparison <strong>of</strong> levitated <strong>and</strong> spray dried samples (bCA 10 % (w/v) plus trehalose)


Results <strong>and</strong> Discussion 159<br />

The results for the residual protein activity measurements <strong>of</strong> bCA at different <strong>drying</strong> times<br />

are given in Figure 5.99. Here none <strong>of</strong> the bCA-trehalose formulations shows a decrease in<br />

the residual protein activity, because <strong>of</strong> the relative high trehalose content. Even the 10:3.3<br />

formulation shows sufficient stabilization <strong>of</strong> the protein. This confirms the results for the<br />

bCA-trehalose 10 % (w/v) total solids formulations, where the minimum trehalose ratio<br />

necessary for stabilization is between 8:2 <strong>and</strong> 6:4.<br />

Figure 5.99: Residual activity <strong>of</strong> bCA depending on the <strong>drying</strong> time for the bCA<br />

10 % (w/v) plus trehalose formulations, <strong>drying</strong> air conditions: 60 °C, 10 % relative<br />

humidity<br />

Figure 5.100 gives the comparison <strong>of</strong> the SEM-pictures <strong>of</strong> levitated particles <strong>and</strong> spray<br />

dried powders for the bCA 10 % (w/v) plus trehalose formulations. The levitated particles<br />

<strong>of</strong> all three formulations show a smooth <strong>surface</strong> <strong>and</strong> have an inclination <strong>of</strong> the top side that<br />

leads to a hole with increasing bCA content. The partially fractured particle <strong>of</strong> the 10:3.3<br />

formulation shows clearly the hole <strong>and</strong> the hollow interior <strong>of</strong> the particle. The spray dried<br />

powders show deeper dimples for low trehalose content in the 10:3.3 formulation than for<br />

high trehalose content in the 10:10 formulation. With increasing trehalose content the<br />

powder particles are more crinkled.


Results <strong>and</strong> Discussion 160<br />

(a) bCA 10 % plus trehalose 10:3.3 (100x)<br />

(b) bCA 10 % plus trehalose 10:3.3(3000x)<br />

(c) bCA 10 % plus trehalose 10:5 (100x)<br />

(d) bCA 10 % plus trehalose 10:5 (3000x)<br />

(e) bCA 10 % plus trehalose 10:10 (100x)<br />

(f) bCA 10 % plus trehalose 10:10 (3000x)<br />

Figure 5.100: SEM-pictures <strong>of</strong> the (a, c, e) levitated <strong>and</strong> (b, d, f) spray dried bCA<br />

10 % (w/v) plus trehalose formulations, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 %<br />

relative humidity


Results <strong>and</strong> Discussion 161<br />

5.5.2 L-Lactic dehydrogenase (LDH) <strong>droplet</strong>s<br />

L-Lactic dehydrogenase from rabbit muscle is chosen as a second model protein to achieve<br />

a higher activity loss in the stabilization experiments. LDH is assumed to be more sensitive<br />

to thermal stress [Adler <strong>and</strong> Lee 1999]. Porcine LDH was used in spray <strong>drying</strong><br />

experiments by Adler <strong>and</strong> Lee [1999] using trehalose as a carrier to analyze powder<br />

particle shape, the residual protein activity after the <strong>drying</strong> process <strong>and</strong> the storage<br />

stability. They found an opposing effect <strong>of</strong> the inlet <strong>temperature</strong> that led for a low<br />

<strong>temperature</strong> setting to high residual protein activity after the <strong>drying</strong> process, but the higher<br />

water content <strong>of</strong> the spray dried powder produced less storage stability.<br />

In the present experiments LDH was (after dialysis <strong>and</strong> concentration <strong>of</strong> the<br />

original LDH-suspension) diluted with a 100 mM potassium phosphate buffer pH 7.0.<br />

Trehalose was added for the two test series <strong>of</strong> LDH-trehalose 10 % (w/v) total solids<br />

(except buffer salts) <strong>and</strong> 10 % (w/v) LDH plus trehalose. The experiments were performed<br />

at a <strong>drying</strong> air <strong>temperature</strong> <strong>of</strong> 60 °C <strong>and</strong> a relative humidity <strong>of</strong> 10 % without a direct <strong>drying</strong><br />

air stream. The initial <strong>droplet</strong> volume was 1.5 μl <strong>and</strong> the initial SPL was about 167.0 dB.<br />

The <strong>drying</strong> behaviour measured via the ²/0²-curve, <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>and</strong><br />

aspect ratio, <strong>and</strong> the residual activity <strong>of</strong> the protein for the levitated <strong>and</strong> spray dried<br />

formulations was analyzed. Strong oscillations occurred <strong>and</strong> especially the pure LDH<br />

10 % (w/v) solution particles break after the critical point or fell out <strong>of</strong> the pressure node.<br />

For this reason the LDH measurements were difficult to perform.<br />

The <strong>drying</strong> behaviour <strong>of</strong> LDH 10 % (w/v) in 100 mM potassium phosphate buffer<br />

pH 7.0 is presented in Figure 5.101. The ²/0²-curves <strong>and</strong> the <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> curves show a bend at the critical point that indicates an amorphous state <strong>of</strong><br />

the protein. The aspect ratio increases in comparison to the bCA 10 % (w/v in 50 mM<br />

trizma buffer pH 7.5) <strong>droplet</strong>s at the critical point from 1.3 up to 1.8. Particle oscillations<br />

can be seen in the graph, reflecting the levitation problems for the pure LDH solution<br />

<strong>droplet</strong>s. As mentioned before, after the critical point the particles started to oscillate <strong>and</strong><br />

<strong>of</strong>ten fell out <strong>of</strong> the pressure node. Fragile particles break, making it difficult to achieve a<br />

sufficient <strong>drying</strong> time for the LDH 10 % (w/v) particles. As already shown for the bCA<br />

particles, the addition <strong>of</strong> small amounts <strong>of</strong> trehalose led to better <strong>drying</strong> behaviour in the<br />

levitator.


Results <strong>and</strong> Discussion 162<br />

Figure 5.101: Drying behaviour <strong>of</strong> a LDH 10 % (w/v) in 100 mM phosphate buffer<br />

pH 7.0 <strong>droplet</strong>, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 % relative humidity<br />

The SEM-picture <strong>of</strong> the levitated LDH 10 % (w/v) particle in Figure 5.102 shows the<br />

fragility <strong>of</strong> the sample. Even if the particle could be removed intact after <strong>drying</strong> was<br />

completed, it breaks during the preparation for the SEM analysis. But the broken particle<br />

still shows that pure LDH 10 % (w/v) solution <strong>droplet</strong>s lead to hollow particles with a hole<br />

at the top side. The spray dried powder <strong>of</strong> the same formulation (inlet <strong>temperature</strong> 130 °C)<br />

shows particles with dimples <strong>and</strong> holes that correspond with the shape <strong>of</strong> the levitated<br />

particle.<br />

(a) 60 °C, 10 % rel. humidity (100x)<br />

(b) spray dried T in = 130 °C (3000x)<br />

Figure 5.102: SEM-pictures <strong>of</strong> (a) a levitated particle <strong>and</strong> (b) the spray dried powder<br />

<strong>of</strong> LDH 10 % (w/v) in 100 mM phosphate buffer pH 7.0


Results <strong>and</strong> Discussion 163<br />

Trehalose 10 % (w/v) dissolved in 100 mM potassium phosphate buffer pH 7.0 shows a<br />

very strange particle shape <strong>of</strong> the levitator dried <strong>droplet</strong> shown in Figure 5.103. The <strong>droplet</strong><br />

formed a <strong>surface</strong> film immediately after it was brought into the pressure node, <strong>and</strong> starts to<br />

flatten. The resulting particles look like a disc. This appearance is due to the potassium<br />

phosphate buffer. Trehalose dissolved in 50 mM trizma buffer pH 7.5 or in pure water<br />

leads in the levitation experiments to much less flattened particles. The strange particle<br />

shape <strong>of</strong> the levitated samples is not seen in the spray dried powder. The spray dried<br />

particles are nearly spherical.<br />

(a) 60 °C, 10 % rel. humidity (50x)<br />

(b) spray dried T in = 130 °C (3000x)<br />

Figure 5.103: SEM-pictures <strong>of</strong> (a) a levitated particle <strong>and</strong> (b) the spray dried powder<br />

<strong>of</strong> trehalose 10 % (w/v) in 100 mM phosphate buffer pH 7.0<br />

5.5.2.1 LDH-trehalose 10 % (w/v) total solids formulations<br />

In the first set <strong>of</strong> experiments trehalose was added to the LDH solution using 100 mM<br />

potassium phosphate buffer pH 7.0 as solvent to a solids content <strong>of</strong> 10 % (w/v) (except<br />

buffer salts). The LDH-trehalose ratios examined are 8:2, 6:4, 4:6 <strong>and</strong> 2:8 as for the bCA<br />

solution experiments. Figure 5.104, Figure 5.105 <strong>and</strong> Figure 5.106 give the <strong>drying</strong><br />

behaviour <strong>of</strong> the <strong>droplet</strong>s in comparison <strong>of</strong> the different formulations. The strange curve<br />

progression <strong>of</strong> pure 10 % (w/v) trehalose <strong>droplet</strong>s is remarkable. The <strong>drying</strong> behaviour <strong>of</strong><br />

this solution is explained by sudden film formation after the <strong>droplet</strong> was inserted in the<br />

pressure node that leads to a reduction <strong>of</strong> solvent evaporation <strong>and</strong> therefore to an increase<br />

in the initial <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>. The increasing aspect ratio is due to the extreme<br />

flattening <strong>of</strong> the <strong>droplet</strong> during the <strong>drying</strong> process that was already shown in the SEM-


Results <strong>and</strong> Discussion 164<br />

picture <strong>of</strong> the pure trehalose particle (Figure 5.103 (a)). The /0²-curves do not<br />

show a clear dependency on the trehalose content <strong>of</strong> the formulations. Only the 2:8<br />

formulation has an clearly earlier critical point in comparison to formulations containing a<br />

lower amount <strong>of</strong> trehalose. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curves show that with<br />

increasing amount <strong>of</strong> trehalose in the formulation the <strong>temperature</strong> increase at the critical<br />

point is shifted to earlier <strong>drying</strong> times. The <strong>temperature</strong> increase is slower in comparison to<br />

formulations with higher protein content. Here the formation <strong>of</strong> a <strong>surface</strong> film similar to<br />

the pure trehalose solution <strong>droplet</strong>s can be seen as the reason for the earlier <strong>surface</strong><br />

<strong>temperature</strong> increase. The aspect ratio for the 8:2 formulation surprisingly does not change<br />

at the critical point, whereas the 6:4 formulations shows an increase in the aspect ratio to<br />

about 2.0. At high trehalose content the aspect ratio curves for the 4:6 <strong>and</strong> 2:8 formulations<br />

show clearly the strong flattening <strong>of</strong> the <strong>droplet</strong> / particle with increasing trehalose content<br />

leading to disc-shaped particles. The irregular flat particle shape can lead to oscillations<br />

<strong>and</strong> the particles can fall out <strong>of</strong>f the pressure node, so that sufficiently long <strong>drying</strong> times<br />

can hardly be reached.<br />

Figure 5.104: Drying behaviour <strong>of</strong> LDH-trehalose 10 % (w/v) total solids formulation<br />

<strong>droplet</strong>s, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 % relative humidity (r(t)²/r(0)²)


Results <strong>and</strong> Discussion 165<br />

Figure 5.105: Drying behaviour <strong>of</strong> LDH-trehalose 10 % (w/v) total solids formulation<br />

<strong>droplet</strong>s, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 % relative humidity (<strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong>)<br />

Figure 5.106: Drying behaviour <strong>of</strong> LDH-trehalose 10 % (w/v) total solids formulation<br />

<strong>droplet</strong>s dried at 60 °C <strong>and</strong> 10 % relative humidity (aspect ratio)


Results <strong>and</strong> Discussion 166<br />

In the SEM-pictures <strong>of</strong> the levitated LDH formulation particles (Figure 5.107 <strong>and</strong> Figure<br />

5.108) the flattening at increasing trehalose content can clearly be seen. The LDHtrehalose<br />

8:2 particle (Figure 5.107) has a similar particle shape to the pure LDH particle<br />

(Figure 5.102 (a)) with the hole at the top side <strong>and</strong> the hollow interior. As soon as trehalose<br />

is added to the formulation, the particles do not break anymore. Even a small amount <strong>of</strong><br />

trehalose has a strong influence on the particle stability in the 8:2 formulation. The 6:4<br />

formulation particle is obviously more flattened than the 8:2 particle <strong>and</strong> shows an<br />

inclination <strong>of</strong> the particles top side. The spray dried powders for both formulations show<br />

strongly dimpled particles, most for the 6:4 formulation. Some particles, especially in the<br />

8:2 powder formulation, seem to have a hollow interior.<br />

(a) LDH-trehalose 8:2 (100x)<br />

(b) LDH-trehalose 8:2 (3000x)<br />

(c) LDH-trehalose 6:4 (100x)<br />

(d) LDH-trehalose 6:4 (3000x)<br />

Figure 5.107: SEM-pictures <strong>of</strong> the (a, c) levitated particles <strong>and</strong> (b, d) spray dried<br />

powders <strong>of</strong> LDH-trehalose 10 % (w/v) total solids formulations, <strong>drying</strong> air conditions:<br />

60 °C <strong>and</strong> 10 %relative humidity


Results <strong>and</strong> Discussion 167<br />

The levitated particles with a higher trehalose content <strong>of</strong> 4:6 <strong>and</strong> 2:8 (Figure 5.108) are<br />

clearly more flattened <strong>and</strong> have a mostly smooth <strong>surface</strong>. The flatness may lead to the<br />

oscillations that have already been seen in the aspect ratio curves. The spray dried powders<br />

show shallower dimpled but more wrinkled particles for higher trehalose content. This<br />

relation could also be seen for the bCA-trehalose formulations before.<br />

SEM-pictures <strong>of</strong> LDH / trehalose formulations using porcine LDH were previously<br />

given by Adler <strong>and</strong> Lee [1999] for spray <strong>drying</strong> solutions containing 10 % (w/w) total<br />

solids. The trehalose ratio was very high at 99.7 % <strong>and</strong> 95 % <strong>of</strong> the total solids. The spray<br />

dried powder particles are for higher trehalose content more wrinkled than the nearly pure<br />

trehalose powder particles containing only 0.3 % LDH. This confirms the results from the<br />

present experiments.<br />

(a) LDH-trehalose 4:6 (100x)<br />

(b) LDH-trehalose 4:6 (3000x)<br />

(c) LDH-trehalose 2:8 (100x)<br />

(d) LDH-trehalose 2:8 (3000x)<br />

Figure 5.108: SEM-pictures <strong>of</strong> the (a, c) levitated particles <strong>and</strong> (b, d) spray dried<br />

powders <strong>of</strong> LDH-trehalose 10 % (w/v) total solids formulations, <strong>drying</strong> air conditions:<br />

60 °C <strong>and</strong> 10 % relative humidity


Results <strong>and</strong> Discussion 168<br />

The <strong>surface</strong> structure <strong>of</strong> the levitated particles dependent on the trehalose ratio in the<br />

formulation is given in the detailed SEM-pictures in Figure 5.109. The 8:2 formulation<br />

particle has a smooth, but cracked <strong>surface</strong>. The same mainly smooth but less cracked<br />

<strong>surface</strong> is shown for the 6:4 formulation particle, whereas the 4:6 particle has no cracks<br />

<strong>and</strong> is more wrinkled. The 2:8 formulation particle shows the smoothest <strong>surface</strong> <strong>of</strong> all<br />

mixture ratios. The <strong>surface</strong> structure <strong>of</strong> the particles reflects the increasing amount <strong>of</strong><br />

trehalose in the formulations. Pure trehalose particles have a highly smooth <strong>surface</strong> <strong>and</strong> are<br />

stable in the levitation process, whereas the pure LDH particles show a high fragility. The<br />

fragility <strong>of</strong> LDH can be seen in the cracks <strong>of</strong> the particle <strong>surface</strong> that decrease with<br />

increasing trehalose content in the formulation. The particle-stabilizing influence <strong>of</strong><br />

trehalose is responsible for less particle breakage after the critical point in comparison to<br />

pure LDH particles.<br />

(a) LDH-trehalose 8:2 (1000x)<br />

(b) LDH-trehalose 6:4 (1000x)<br />

(c) LDH-trehalose 4:6 (1000x)<br />

(d) LDH-trehalose 2:8 (1000x)<br />

Figure 5.109: Surface structure <strong>of</strong> levitated LDH-trehalose 10 % (w/v) total solids<br />

particles, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 % relative humidity


Results <strong>and</strong> Discussion 169<br />

The residual activity <strong>of</strong> LDH in the different levitated <strong>and</strong> spray dried LDH-trehalose 10 %<br />

total solids formulations was also analyzed. The results for the spray dried powder<br />

formulations are given in Figure 5.110. A stabilizing effect <strong>of</strong> trehalose addition on the<br />

residual LDH activity can be seen for most formulations, but a dependency on the amount<br />

<strong>of</strong> trehalose in the mixture is ambiguous. The levitated LDH formulation particles were<br />

also analyzed, but the results show large discrepancy <strong>of</strong> the measured activity values. The<br />

measured residual activity <strong>of</strong> the fresh sample solution was reproducible, but the measured<br />

activity values <strong>of</strong> the levitated particles deviated greatly.<br />

Adler <strong>and</strong> Lee [1999] performed LDH activity studies using a spray dried powder<br />

produced from a solution containing 0.3 % porcine LDH stabilized by 99.7 % trehalose<br />

(10 % total solids). Using spray <strong>drying</strong> conditions <strong>of</strong> T in = 90 °C, T out = 60 °C led to an<br />

activity loss <strong>of</strong> 11 % for the protein. At a higher inlet <strong>temperature</strong> <strong>of</strong> 150 °C the activity<br />

loss increases to about 25 %. This is a much higher activity loss than found for the<br />

stabilized formulations containing a higher ratio <strong>of</strong> LDH in the present experiments, but it<br />

has to be considered that in the present experiments LDH from rabbit muscle was used.<br />

Figure 5.110: Residual activity <strong>of</strong> the spray dried LDH-trehalose 10 % (w/v) total<br />

solids formulations depending on the trehalose content


Results <strong>and</strong> Discussion 170<br />

5.5.2.2 LDH 10 % (w/v) plus trehalose formulations<br />

The <strong>drying</strong> behaviour described in the ²/0²-curves, the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong><br />

curves <strong>and</strong> the aspect ratio curves for the LDH 10 % (w/v) plus trehalose solutions using<br />

100 mM potassium phosphate buffer pH 7.0 with the LDH-trehalose ratios <strong>of</strong> 10:10, 10:20<br />

<strong>and</strong> 10:30 is given in Figure 5.111, Figure 5.112 <strong>and</strong> Figure 5.113. As expected, with<br />

increasing total solids <strong>of</strong> the formulations the critical point appears earlier in the ²/<br />

0²-curves <strong>and</strong> the final particle is larger. In the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curves the<br />

critical point appears consequently also at earlier <strong>drying</strong> times for higher solids content <strong>and</strong><br />

the <strong>temperature</strong> increases substantially slower. The aspect ratio increase <strong>of</strong> the <strong>droplet</strong>s is<br />

not as high as for the LDH-trehalose 10 % total solids formulations, the highest value is<br />

only about 2.5. For the 4:6 <strong>and</strong> 2:8 <strong>droplet</strong>s (10 % total solids formulations) the <strong>droplet</strong> is<br />

more flattened in comparison with the LDH-trehalose ratio 10:30. For a higher solids<br />

content the flattening <strong>of</strong> the particle is therefore reduced. The levitated particles produced<br />

from the high concentrated solutions are <strong>of</strong>ten not completely dry at the end <strong>of</strong> the<br />

measurement <strong>and</strong> stick at the tube wall before the SEM-pictures could be taken. This leads<br />

subsequently to particle deformation, so that the visible particle shape is not necessarily<br />

representative <strong>of</strong> the particle shape directly after the <strong>drying</strong> process.<br />

Figure 5.111: Drying behaviour <strong>of</strong> LDH 10 % (w/v) plus trehalose formulation<br />

<strong>droplet</strong>s, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 % relative humidity (r(t)²/r(0)²)


Results <strong>and</strong> Discussion 171<br />

Figure 5.112: Drying behaviour <strong>of</strong> LDH 10 % (w/v) plus trehalose formulation<br />

<strong>droplet</strong>s, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 % relative humidity (<strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong>)<br />

Figure 5.113: Drying behaviour <strong>of</strong> LDH 10 % (w/v) plus trehalose formulation<br />

<strong>droplet</strong>s, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 % relative humidity (aspect ratio)


Results <strong>and</strong> Discussion 172<br />

The SEM-pictures <strong>of</strong> the spray dried powders for each formulation are given in Figure<br />

5.114. The 10:10 formulation leads to dimpled powder particles. With increasing trehalose<br />

content the dimples are reduced in the 10:20 formulation, <strong>and</strong> the 10:30 formulation shows<br />

spherical but fused particles. This could be due to the higher water content in the spray<br />

dried powder produced from the high concentrated solution. A possible explanation is also<br />

the sticky nature <strong>of</strong> trehalose [Maa et al. 1997].<br />

(a) LDH plus trehalose 10:10 (3000x)<br />

(b) LDH plus trehalose 10:20 (3000x)<br />

(c) LDH plus trehalose 10:30 (3000x)<br />

Figure 5.114: Spray dried powders <strong>of</strong> LDH 10 % (w/v) plus trehalose formulations,<br />

T in = 130 °C<br />

For the levitated particles the activity values again show discrepancies so that the results<br />

are not useful for interpretation <strong>of</strong> the stabilizing effect <strong>of</strong> trehalose. The residual activity<br />

<strong>of</strong> the spray dried samples is around 80 % <strong>and</strong> there is no increase <strong>of</strong> activity due to<br />

trehalose addition visible (Figure 5.115). No further increase in the amount <strong>of</strong> trehalose in<br />

the formulation was analyzed because <strong>of</strong> the influence <strong>of</strong> the trehalose content on the


Results <strong>and</strong> Discussion 173<br />

<strong>drying</strong> process leading to stickiness <strong>of</strong> the particles (shown for the 10:30 formulation). In<br />

the 10 % (w/v) formulations an increase in the trehalose content would lead to smaller<br />

amounts <strong>of</strong> LDH <strong>and</strong> therefore lower activity <strong>of</strong> the sample particle <strong>and</strong> still higher<br />

measurement inaccuracies. For this reason no further experiments using LDH formulations<br />

were performed.<br />

Figure 5.115: Residual activity <strong>of</strong> spray dried LDH 10 % (w/v) plus trehalose<br />

formulations depending on the trehalose content<br />

The protein activity measurement at different <strong>drying</strong> times was performed for the pure<br />

10 % (w/v) LDH in 100 mM potassium phosphate buffer solution. The results are given as<br />

an example in Figure 5.116 to illustrate the behaviour that occurs. An unambiguous<br />

decrease in the LDH activity is not seen. The activity <strong>of</strong> the <strong>droplet</strong>s removed well before<br />

the critical point (about 250 s) are reproducible <strong>and</strong> show no reduction <strong>of</strong> the LDH activity.<br />

However, for particles removed close to the critical point the st<strong>and</strong>ard deviations <strong>of</strong> the<br />

measured LDH activity values become very large. Additionally, as mentioned before after<br />

the critical point the particles break or fall out <strong>of</strong> the pressure node, so that it was<br />

impossible to analyze the whole <strong>drying</strong> process. Trehalose addition (8:2 formulation, 10 %<br />

total solids) leads to longer <strong>drying</strong> times, but not to an improvement <strong>of</strong> the reproducibility


Results <strong>and</strong> Discussion 174<br />

<strong>of</strong> the LDH activity values at <strong>and</strong> after the critical point. LDH is therefore not suitable for<br />

the analysis <strong>of</strong> protein inactivation during the whole <strong>drying</strong> process with this technique.<br />

Figure 5.116: Residual activity <strong>of</strong> LDH 10 % (w/v) solution <strong>droplet</strong>s dependent on the<br />

<strong>drying</strong> time, <strong>drying</strong> air conditions: 60 °C <strong>and</strong> 10 % relative humidity<br />

5.5.3 Trypsinogen <strong>droplet</strong>s<br />

Trypsinogen was investigated for its suitability in protein stabilization experiments.<br />

Tzannis <strong>and</strong> Prestrelski [1999] used a 2 % (w/v) trypsinogen solution in 1 M HCl for<br />

protein stabilization studies with sucrose as a stabilizing excipient in spray <strong>drying</strong><br />

formulations (T in = 120 °C, T out = 85 °C). They found a reduction <strong>of</strong> the residual protein<br />

activity in the spray dried pure trypsinogen powder to about 85 % related to the untreated<br />

solution. Even small amounts <strong>of</strong> sucrose stabilized the protein. The highest residual<br />

activity <strong>of</strong> 99.9 % was found for the 1:1 mixture. This activity reduction is promising for<br />

the projected levitation experiments.<br />

The <strong>drying</strong> behaviour <strong>of</strong> the 10 % (w/v) trypsinogen solution <strong>droplet</strong> using water<br />

p.a. as solvent is given in Figure 5.117. The <strong>drying</strong> air conditions were 60 °C <strong>and</strong> 20 %<br />

relative humidity without a direct <strong>drying</strong> air stream. Lower humidity settings do not lead to<br />

sufficiently long <strong>drying</strong> times due to <strong>droplet</strong> oscillations. The ²/0²-curve <strong>and</strong> the<br />

<strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curve both show the curve progression for an amorphous


Results <strong>and</strong> Discussion 175<br />

substance. The ²/0²-curve shows a bend at the critical point <strong>and</strong> a slow <strong>droplet</strong><br />

<strong>surface</strong> <strong>temperature</strong> increase can be seen. The aspect ratio decreases during the<br />

measurement from 1.3 to about 1.1. A slight sudden displacement <strong>of</strong> the curve to lower<br />

values appears at the critical point.<br />

Figure 5.117: Drying behaviour <strong>of</strong> a trypsinogen 10 % (w/v) solution <strong>droplet</strong>, <strong>drying</strong><br />

air conditions: 60 °C <strong>and</strong> 20 % relative humidity<br />

The SEM-pictures <strong>of</strong> the pure trypsinogen particles are given in Figure 5.118. The<br />

levitated particles <strong>and</strong> the spray dried powder particles show a smooth <strong>surface</strong>. The<br />

partially fractured levitated particle has a hollow interior. A hole as seen for the pure bCA<br />

<strong>and</strong> LDH solution particles is not visible in the pictures. The spray dried powder consists<br />

<strong>of</strong> hollow particles with a hole, as already seen for the 10 % (w/v) carbonic anhydrase <strong>and</strong><br />

LDH powders.<br />

The measured residual activity <strong>of</strong> a 10 % (w/v) trypsinogen solution in water p.a.<br />

both for the levitated particles dried at 60 °C <strong>and</strong> 20 % relative humidity without a direct<br />

<strong>drying</strong> air stream <strong>and</strong> the spray dried powder (T in = 130 °C, T out = 80 °C) was, however,<br />

100 %. For this reason trypsinogen is also not useful for protein stabilization experiments<br />

in the levitator. No explanation for this behaviour or that <strong>of</strong> LDH was sought further in this<br />

work.


Results <strong>and</strong> Discussion 176<br />

(a) 60 °C, 20 % rel. humidity (100x)<br />

(b) 60 °C, 20 % rel. humidity (100x)<br />

(c) spray dried T in = 130 °C (3000x)<br />

Figure 5.118: SEM-pictures <strong>of</strong> the (a, b) levitated <strong>and</strong> (c) spray dried trypsinogen<br />

10 % (w/v) solution particles


Summary <strong>and</strong> Conclusions 177<br />

6 Summary <strong>and</strong> Conclusions<br />

This thesis deals with the analysis <strong>of</strong> the evaporation / <strong>drying</strong> behaviour <strong>and</strong> the particle<br />

formation <strong>of</strong> pure solvents, excipient solutions <strong>and</strong> solutions containing an active<br />

pharmaceutical ingredient or model protein. The analytical method used for this purpose is<br />

single <strong>droplet</strong> <strong>drying</strong> using an ultrasonic levitator. During the whole <strong>drying</strong> process the<br />

levitated <strong>droplet</strong>s are monitored by a CCD-camera for <strong>droplet</strong> size <strong>and</strong> position<br />

measurement, <strong>and</strong> an IR-camera for observation <strong>of</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>. The<br />

dried particles are removed for either SEM-pictures or for residual activity measurements<br />

<strong>of</strong> protein. The first part <strong>of</strong> this thesis discusses the evaporation behaviour <strong>of</strong> pure solvent<br />

<strong>droplet</strong>s. In the second part experiments using excipient solutions are presented. Solutions<br />

containing the active pharmaceutical ingredient itraconazole or the model proteins bovine<br />

carbonic anhydrase (bCA), L-lactic dehydrogenase from rabbit muscle (LDH) or bovine<br />

trypsinogen are discussed in the third part <strong>of</strong> this thesis.<br />

In the first part <strong>of</strong> this thesis water, the most common solvent for protein<br />

formulations, was examined. To investigate the influence <strong>of</strong> different <strong>droplet</strong> sizes on<br />

evaporation, <strong>droplet</strong>s with an initial diameter <strong>of</strong> 500 μm, 800 μm <strong>and</strong> 1200 μm were<br />

analyzed at different <strong>drying</strong> air conditions (<strong>temperature</strong> <strong>and</strong> relative humidity) without a<br />

direct airstream to the <strong>droplet</strong>. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> is constant for the most part<br />

<strong>of</strong> the measurement. The initial diameter <strong>of</strong> the <strong>droplet</strong>s shows no influence on their<br />

<strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>. The experimental <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s (taken from the<br />

constant <strong>temperature</strong> part) were compared to the wet-bulb <strong>temperature</strong> taken from a<br />

psychrometric chart for the respective ambient conditions. Additionally the values are<br />

compared to the theoretical <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> taking into account the ultrasonic<br />

field according to Yarin et al. [1999]. The experimental values are higher than the wet-bulb<br />

<strong>temperature</strong> <strong>and</strong> the theoretical <strong>surface</strong> <strong>temperature</strong> calculated according to Yarin et al.<br />

[1999]. The deviation <strong>of</strong> the <strong>surface</strong> <strong>temperature</strong>s can be explained by the energy supply<br />

from the acoustic wave to the <strong>droplet</strong>. This leads to a heating-up <strong>of</strong> the <strong>droplet</strong> <strong>and</strong><br />

therefore to higher <strong>surface</strong> <strong>temperature</strong>.<br />

For all initial <strong>droplet</strong> sizes an almost linear decrease <strong>of</strong> the ²/0²-graph can<br />

be seen. The experimental evaporation coefficients (negative slope <strong>of</strong> the ²/0²graph)<br />

are higher for larger initial diameters, because evaporation is accelerated by the


Summary <strong>and</strong> Conclusions 178<br />

higher SPL setting that is necessary for stable levitation <strong>of</strong> larger <strong>droplet</strong>s. The<br />

experimental values <strong>of</strong> the evaporation coefficient are higher than the calculated values<br />

using the wet-bulb <strong>temperature</strong> or the <strong>surface</strong> <strong>temperature</strong> according to Yarin et al. [1999].<br />

Using the experimental <strong>surface</strong> <strong>temperature</strong> Sherwood numbers lie in the range <strong>of</strong> 2,<br />

indicating correspondence <strong>of</strong> the evaporation process with the d²-law. Therefore, the<br />

accelerated evaporation cannot be sufficiently explained only by the influence <strong>of</strong> the inner<br />

acoustic streaming. It is also dependent on the energy supply by the acoustic wave.<br />

By using a <strong>drying</strong> airstream directed to the <strong>droplet</strong> the evaporation can be<br />

accelerated <strong>and</strong> therefore the evaporation coefficient increases. In experiments at an<br />

ambient <strong>temperature</strong> <strong>of</strong> 25 °C the measured <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> is lower using a<br />

<strong>drying</strong> airstream than without one. Due to the reduction in the amount <strong>of</strong> water vapour in<br />

the outer acoustic stream there is an increased cooling <strong>of</strong> the <strong>droplet</strong>. In contrast, at an<br />

ambient <strong>temperature</strong> <strong>of</strong> 60 °C the values using a direct <strong>drying</strong> airstream are higher. The<br />

higher supply <strong>of</strong> hot <strong>drying</strong> air to the <strong>droplet</strong> via direct airstream compared to<br />

measurements without airstream appears to predominate.<br />

The organic solvents acetone, 2-butanone, dichloromethane, ethanol, ethyl acetate,<br />

2-propanol <strong>and</strong> tetrahydr<strong>of</strong>uran were also analyzed <strong>and</strong> show different evaporation<br />

coefficients <strong>and</strong> <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s dependent on their liquid properties. In<br />

comparison to the calculated values there is also a deviation, with higher values for the<br />

experimental evaporation coefficients <strong>and</strong> <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s. Some solvent<br />

mixtures show a break in the ²/0²-graphs indicating the fast evaporation <strong>of</strong> the<br />

more volatile solvent in the mixture at the beginning <strong>of</strong> the measurement. As soon as the<br />

more volatile fraction has almost completely evaporated, a <strong>surface</strong> <strong>temperature</strong> increase<br />

takes place. Other solvent mixtures show an almost linear decrease in the ²/0²graph.<br />

In this case the evaporation is faster with increasing amount <strong>of</strong> the more volatile<br />

solvent in the mixture <strong>and</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> is lower.<br />

In the second part <strong>of</strong> this thesis excipient solutions were examined using trehalose<br />

<strong>and</strong> sucrose as amorphous substances <strong>and</strong> mannitol as a crystalline substance. The curve<br />

progression <strong>of</strong> the ²/0²-graphs, <strong>of</strong> the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong>s <strong>and</strong> the aspect<br />

ratios were compared. In the first <strong>drying</strong> stage solution <strong>droplet</strong>s show a linear decrease in<br />

the ²/0²-graph. After the critical point the <strong>droplet</strong> size stays almost constant. The<br />

<strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> increases at the critical point, because the evaporation <strong>of</strong> the<br />

solvent is slowed down due to crust formation. Depending on the sugar properties,


Summary <strong>and</strong> Conclusions 179<br />

amorphous or crystalline, the curve progression at the critical point differs. An amorphous<br />

substance like trehalose shows a bend in the ²/0²-graph <strong>and</strong> the <strong>droplet</strong> <strong>surface</strong><br />

<strong>temperature</strong> increase is slow. The aspect ratio changes before <strong>and</strong> after the critical point,<br />

because the substance does not form a stable particle but it is still deformable. In contrast,<br />

the crystalline substance mannitol shows a sharp break in the ²/0²-graph <strong>and</strong> an<br />

immediate <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> increase. The aspect ratio is almost constant in the<br />

second <strong>drying</strong> stage. This <strong>drying</strong> behaviour <strong>of</strong> mannitol is due to the rapid formation <strong>of</strong> a<br />

solid crystalline crust at the critical point. The present experiments show that <strong>droplet</strong><br />

<strong>surface</strong> <strong>temperature</strong> measurement is, in addition to the <strong>droplet</strong> size measurement, a helpful<br />

analytical method to detect the critical point in the <strong>drying</strong> process <strong>of</strong> solution <strong>droplet</strong>s.<br />

The polymer excipients copolyvidone (Kollidon® VA 64) <strong>and</strong> HPMC (Pharmacoat<br />

615) dissolved in various dichloromethane-ethanol-mixtures were analyzed. The<br />

experimental results show a completely different <strong>drying</strong> behaviour <strong>of</strong> the two substances.<br />

The ²/0²-graph <strong>of</strong> the copolyvidone solution shows a break at the critical point <strong>and</strong><br />

the <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> increases as expected. In contrast, the HPMC solution<br />

<strong>droplet</strong>s form a <strong>surface</strong> film immediately after the <strong>droplet</strong>s were inserted into the acoustic<br />

field. This leads to a continual <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> increase <strong>and</strong> to a large bend in<br />

the ²/0²-graph, so that a critical point for the HPMC solutions is not clearly<br />

detectable. With increasing amount <strong>of</strong> the less volatile solvent ethanol in the<br />

dichloromethane-ethanol-mixture there is a shift <strong>of</strong> the critical point to later <strong>drying</strong> times<br />

for the copolyvidone solutions. The HPMC solutions do not show substantial changes in<br />

the curve progression.<br />

In the third part <strong>of</strong> this thesis solutions <strong>of</strong> itraconazole as an active pharmaceutical<br />

ingredient <strong>and</strong> solutions <strong>of</strong> the model proteins bCA, LDH <strong>and</strong> trypsinogen together with<br />

excipients were examined. Itraconazole was analyzed using either copolyvidone or HPMC<br />

as a polymer excipient. The <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> curve, the ²/0²-graph <strong>and</strong><br />

the change <strong>of</strong> the aspect ratio during the measurement are not influenced by the<br />

itraconazole addition, but depend on the added polymer. The final particle shape is,<br />

however, predominated by itraconazole. Particles containing itraconazole show a rougher<br />

<strong>surface</strong> in comparison to the pure polymer particles. The analysis <strong>of</strong> the two<br />

itraconazole / polymer formulation series in the levitator illustrates a potential application<br />

<strong>of</strong> single <strong>droplet</strong> <strong>drying</strong> for the optimization <strong>of</strong> formulations.


Summary <strong>and</strong> Conclusions 180<br />

The <strong>drying</strong> behaviour <strong>of</strong> the pure 10 % (w/v) bCA solution as well as <strong>of</strong> different bCAtrehalose<br />

formulations shows the characteristic curve progression for amorphous<br />

substances. SEM-pictures show that with increasing trehalose content the particles have a<br />

more smoothed <strong>surface</strong>, but are more flattened. Levitated pure carbonic anhydrase particles<br />

have a hole at the top side <strong>and</strong> are hollow, as also observed with spray dried powder<br />

particles. The residual protein activity increases for both the levitated particles <strong>and</strong> the<br />

spray dried powder samples with trehalose addition. A trehalose addition <strong>of</strong> more than<br />

20 % <strong>of</strong> the total solids does not lead to higher protein protection. In further experiments<br />

the bCA-trehalose formulations were tested for protein stability during the <strong>drying</strong> process.<br />

At different time points in the <strong>drying</strong> process <strong>droplet</strong>s or particles were removed out <strong>of</strong> the<br />

levitator <strong>and</strong> the residual activity <strong>of</strong> the protein was measured. For the pure 10 % (w/v)<br />

bCA solution a decrease in the residual activity after the critical point can be seen. The<br />

stabilized formulations show almost constant activity during the <strong>drying</strong> process as<br />

expected.<br />

LDH was chosen as second model protein, because it should show higher losses <strong>of</strong><br />

activity during <strong>drying</strong>. However, the activity measurements <strong>of</strong> the levitated particles show<br />

wide deviations <strong>of</strong> the measured values <strong>and</strong> cannot be interpreted. The spray dried samples<br />

<strong>of</strong> the formulation show a stabilization effect even for a low trehalose ratio <strong>of</strong> 20 % <strong>of</strong> the<br />

total solids. SEM-pictures <strong>of</strong> the levitated particles show an extreme flattening with<br />

increasing trehalose content. The particles produced from the pure 10 % (w/v) trehalose<br />

solution <strong>droplet</strong>s are also very flattened <strong>and</strong> disc-shaped. This behaviour is a result <strong>of</strong> using<br />

the 100 mM potassium phosphate buffer as solvent, because trehalose dissolved in pure<br />

water or in 50 mM trizma buffer lead to predominantly spherical particles in the levitator.<br />

The flat particles started to oscillate making measurements difficult. Pure LDH particles<br />

have a hole at the top side <strong>and</strong> are hollow, as also seen with the pure bCA particles.<br />

Trypsinogen was analyzed as the third protein, but no activity decrease due to the<br />

<strong>drying</strong> process could be observed. The <strong>drying</strong> behaviour <strong>of</strong> the pure 10 % (w/v)<br />

trypsinogen solution <strong>droplet</strong>s is similar to the other two proteins.<br />

Single <strong>droplet</strong> <strong>drying</strong> experiments including <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong><br />

measurements are a valuable tool for the analysis <strong>of</strong> the evaporation behaviour <strong>of</strong> solvents<br />

<strong>and</strong> <strong>of</strong> the <strong>drying</strong> process <strong>of</strong> pharmaceutical solution formulations for spray <strong>drying</strong>. The<br />

results for particle shape <strong>and</strong> residual protein activity from spray dried powders <strong>and</strong><br />

levitated particles correspond well. Residual protein activity measurements are possible,


Summary <strong>and</strong> Conclusions 181<br />

because the <strong>droplet</strong> or particle can be removed out <strong>of</strong> the acoustic field at any time. For this<br />

reason formulations containing expensive pharmaceutical proteins can be optimized by<br />

using very small amounts <strong>of</strong> substance. The <strong>kinetics</strong> <strong>of</strong> protein inactivation during the<br />

<strong>drying</strong> process needs, however, further analysis using a more <strong>temperature</strong>-sensitive model<br />

protein. Furthermore, surfactants should be analyzed as excipients in the formulations, also<br />

in combination with sugars like trehalose, for the protein stabilizing effect due to their<br />

accumulation at the <strong>droplet</strong> <strong>surface</strong>. Via integration <strong>of</strong> a cooling system for the piezocrystal<br />

in the transducer, experiments could be performed at higher <strong>drying</strong> air<br />

<strong>temperature</strong>s.


Zusammenfassung 182<br />

7 Zusammenfassung<br />

Diese Arbeit beschäftigt sich mit der Untersuchung des Verdampfungs- sowie<br />

Trocknungsverhaltens und der Partikelbildung von reinen Lösungsmitteln, von<br />

Hilfsst<strong>of</strong>flösungen und von Lösungen, die einen Wirkst<strong>of</strong>f oder ein Modellprotein<br />

enthalten. Die für diesen Zweck verwendete analytische Methode ist die Trocknung von<br />

Einzeltropfen in einem Ultraschall-Levitator. Während des gesamten Trocknungsprozesses<br />

werden die schwebenden Tropfen durch eine CCD-Kamera für die Messung der<br />

Tropfengröße und Position, und durch eine IR-Kamera für die Beobachtung der<br />

Temperatur auf der Tropfenoberfläche überwacht. Die getrockneten Partikel werden für<br />

SEM-Aufnahmen oder für Messungen der Restaktivität bei Proteinen entnommen. Im<br />

ersten Teil dieser Arbeit wird das Verdampfungsverhalten von reinen Lösungsmitteltropfen<br />

beh<strong>and</strong>elt. Im zweiten Teil werden Versuche mit Hilfsst<strong>of</strong>flösungen beschrieben.<br />

Lösungen, die den Wirkst<strong>of</strong>f Itraconazol oder die Modellproteine Carboanhydrase vom<br />

Rind (bCA), L-Lactat-Dehydrogenase aus Kaninchenmuskel (LDH) oder Trypsinogen<br />

vom Rind enthalten, werden im dritten Teil dieser Arbeit beh<strong>and</strong>elt.<br />

Wasser, das am weitesten verbreitete Lösungsmittel für Proteinformulierungen,<br />

wurde zuerst untersucht. Um den Einfluss verschiedener Tropfengrößen auf die<br />

Verdampfung zu ermitteln, wurden Tropfen mit einem initialen Durchmesser von 500 μm,<br />

800 μm und 1200 μm bei verschiedenen Trocknungsluftbedingungen (Temperatur und<br />

relative Luftfeuchte) ohne direkten Luftstrom auf den Tropfen untersucht. Die<br />

Oberflächentemperatur des Tropfens ist über den größten Teil der Messung konstant. Ein<br />

Einfluss des initialen Durchmessers der Tropfen auf ihre Oberflächentemperatur konnte<br />

nicht festgestellt werden. Die experimentell ermittelten Oberflächen<strong>temperature</strong>n der<br />

Tropfen (aus dem konstanten Temperaturbereich) wurden mit der Kühlgrenztemperatur<br />

verglichen, die mit Hilfe eines psychrometrischen Diagramms für die jeweiligen<br />

Umgebungsbedingungen ermittelt wurde. Zusätzlich erfolgte ein Vergleich mit der<br />

theoretischen Oberflächentemperatur der Tropfen unter Berücksichtigung des Einflusses<br />

durch das Ultraschallfeld, berechnet nach Yarin et al. [1999]. Die experimentell ermittelten<br />

Temperaturwerte sind größer als die Kühlgrenztemperatur und als die theoretische<br />

Oberflächentemperatur berechnet nach Yarin et al. [1999]. Die Differenz der Oberflächen<strong>temperature</strong>n<br />

lässt sich durch den Energieeintrag aus der akustischen Welle in den Tropfen


Zusammenfassung 183<br />

erklären. Dieser führt zu einem Aufheizen des Tropfens und folglich zu einer höheren<br />

Oberflächentemperatur.<br />

Bei allen initialen Durchmessern der Tropfen ist eine nahezu lineare Abnahme des<br />

²/0²-Graphen zu erkennen. Die experimentellen Verdampfungskoeffizienten<br />

(negative Steigung des ²/0²-Graphen) sind bei größeren initialen Durchmessern<br />

höher, da die Verdampfung durch eine stärkere SPL-Einstellung, die für ein stabiles<br />

Schweben größerer Tropfen erforderlich ist, beschleunigt wird. Im Vergleich sind die<br />

experimentell ermittelten Werte des Verdampfungskoeffizienten höher als die berechneten<br />

Werte unter Verwendung der Kühlgrenztemperatur oder der Oberflächentemperatur<br />

berechnet nach Yarin et al. [1999]. Bei Verwendung der experimentell ermittelten<br />

Oberflächentemperatur liegen die Sherwood-Zahlen im Bereich von 2, was auf eine<br />

Übereinstimmung des Verdampfungsprozesses mit dem d²-Gesetz hinweist. Die<br />

beschleunigte Verdampfung kann daher mit dem Einfluss der inneren akustischen<br />

Strömung allein nicht zufriedenstellend erklärt werden. Sie ist auch abhängig von dem<br />

Energieeintrag durch die akustische Welle.<br />

Unter Verwendung eines auf den Tropfen gerichteten Trocknungsluftstrahls kann<br />

die Verdampfung beschleunigt werden und folglich erhöht sich der Verdampfungskoeffizient.<br />

In Versuchen bei 25 °C Umgebungstemperatur ist die gemessene<br />

Oberflächentemperatur der Tropfen bei Verwendung eines Trocknungsluftstrahls geringer<br />

als ohne. Aufgrund der Reduzierung des Wasserdampfanteils in der äußeren akustischen<br />

Strömung kommt es zu einer verstärkten Kühlung des Tropfens. Bei 60 °C<br />

Umgebungstemperatur sind die Werte mit direktem Trocknungsluftstrom dagegen höher.<br />

Die größere Zufuhr an heißer Trocknungsluft an den Tropfen durch den direkten Luftstrom<br />

im Vergleich zu Messungen ohne Luftstrom scheint hier dominierend zu sein.<br />

Die organischen Lösungsmittel Aceton, 2-Butanon, Dichlormethan, Ethanol,<br />

Ethylacetat, 2-Propanol und THF wurden ebenfalls untersucht und zeigen verschiedene<br />

Verdampfungskoeffizienten und Oberflächen<strong>temperature</strong>n der Tropfen abhängig von ihren<br />

Flüssigkeitseigenschaften. Im Vergleich zu den berechneten Werten gibt es auch hier eine<br />

Abweichung, die Werte für die experimentellen Verdampfungskoeffizienten und die<br />

Oberflächen<strong>temperature</strong>n der Tropfen sind höher. Einige Lösungsmittelmischungen zeigen<br />

Kurven in den ²/0²-Graphen, was auf eine schnelle Verdampfung des leichter<br />

flüchtigen Lösungsmittels in der Mischung zu Anfang der Messung hinweist. Sobald der<br />

leichter flüchtige Anteil fast vollständig verdampft ist, findet ein Anstieg der


Zusammenfassung 184<br />

Oberflächentemperatur statt. Andere Lösungsmittelmischungen zeigen eine fast lineare<br />

Abnahme des ²/0²-Graphen. In diesem Fall ist die Verdampfung mit steigendem<br />

Anteil des leichter flüchtigen Lösungsmittels in der Mischung schneller und die<br />

Oberflächentemperatur der Tropfen ist geringer.<br />

Im zweiten Teil dieser Arbeit wurden Hilfsst<strong>of</strong>flösungen unter Verwendung von<br />

Trehalose und Saccharose als amorphe Substanzen und Mannitol als kristalline Substanz<br />

untersucht. Die Kurvenverläufe der ²/0²-Graphen, der Oberflächen<strong>temperature</strong>n<br />

und der Achsenverhältnisse der Tropfen wurden verglichen. In der ersten Trocknungsphase<br />

zeigen Lösungstropfen eine lineare Abnahme des ²/0²-Graphen. Nach<br />

Überschreiten des kritischen Punktes bleibt die Tropfengröße nahezu konstant. Die<br />

Oberflächentemperatur der Tropfen steigt am kritischen Punkt an, da die Verdampfung des<br />

Lösungsmittels durch die Krustenbildung verlangsamt wird. Abhängig von den<br />

Eigenschaften des Zuckers, amorph oder kristallin, unterscheidet sich der Kurvenverlauf<br />

am kritischen Punkt. Eine amorphe Substanz wie Trehalose zeigt eine Kurve im ²/<br />

0²-Graphen und der Anstieg der Oberflächentemperatur der Tropfen ist langsam. Das<br />

Achsenverhältnis ändert sich vor und nach dem kritischen Punkt, da die Substanz nicht<br />

s<strong>of</strong>ort einen festen Partikel bildet, sondern dieser immer noch formbar bleibt. Im<br />

Gegensatz dazu zeigt die kristalline Substanz Mannitol einen scharfen Knick im ²/<br />

0²-Graphen und einen s<strong>of</strong>ortigen Temperaturanstieg der Tropfenoberfläche. Das<br />

Achsenverhältnis ist nahezu konstant in der zweiten Trocknungsphase. Dieses<br />

Trocknungsverhalten von Mannitol ist bedingt durch den s<strong>of</strong>ortigen Aufbau einer festen<br />

kristallinen Kruste am kritischen Punkt. Die vorliegenden Versuche zeigen, dass die<br />

Messung der Tropfenoberflächentemperatur, zusätzlich zu der Tropfengrößenbestimmung,<br />

ein hilfreiches analytisches Verfahren ist, um den kritischen Punkt im Trocknungsprozess<br />

von Lösungstropfen zu bestimmen.<br />

Die polymeren Hilfsst<strong>of</strong>fe Copolyvidon (Kollidon® VA 64) und HPMC<br />

(Pharmacoat 615) wurden, gelöst in unterschiedlichen Dichlormethan-Ethanol-<br />

Mischungen, analysiert. Die Versuchsergebnisse zeigen ein ganz unterschiedliches<br />

Trocknungsverhalten der zwei Substanzen. Der ²/0²-Graph der<br />

Copolyvidonlösung weist einen Knick am kritischen Punkt auf und die<br />

Oberflächentemperatur der Tropfen steigt wie erwartet an. Im Gegensatz dazu bilden die<br />

HPMC–Lösungstropfen unmittelbar nach Einbringen des Tropfens in das Ultraschallfeld


Zusammenfassung 185<br />

einen Oberflächenfilm aus. Dieser führt zu einem kontinuierlichen Anstieg der<br />

Oberflächentemperatur des Tropfens und zu einer weiten Kurve im ²/0²-Graphen,<br />

so dass der kritische Punkt für die HPMC-Lösungen nicht klar bestimmbar ist. Bei den<br />

Copolyvidonlösungen erfolgt mit ansteigendem Anteil des geringer flüchtigen<br />

Lösungsmittels Ethanol in der Dichlormethan-Ethanol-Mischung eine Verlagerung des<br />

kritischen Punktes zu späteren Trocknungszeitpunkten. Die HPMC-Lösungen zeigen keine<br />

wesentlichen Änderungen im Kurvenverlauf.<br />

Im dritten Teil dieser Arbeit wurden Lösungen mit Itraconazol als Wirkst<strong>of</strong>f und<br />

Lösungen mit den Modellproteinen bCA, LDH und Trypsinogen zusammen mit<br />

Hilfsst<strong>of</strong>fen untersucht. Itraconazol wurde unter Verwendung von Copolyvidon oder<br />

HPMC als polymerem Hilfsst<strong>of</strong>f untersucht. Der Oberflächentemperaturverlauf der<br />

Tropfen, der ²/0²-Graph und die Entwicklung des Achsenverhältnisses während<br />

der Messung werden nicht durch den Itraconazolzusatz beeinflusst, sondern sind abhängig<br />

vom eingesetzten Polymer. Das Aussehen des fertigen Partikels wird allerdings von<br />

Itraconazol bestimmt. Partikel, die Itraconazol enthalten, zeigen eine rauere Oberfläche im<br />

Vergleich zu reinen Polymerpartikeln. Die Untersuchungen der zwei Itraconazol/Polymer-<br />

Formulierungsserien im Levitator veranschaulichen eine mögliche Anwendung der<br />

Einzeltropfentrocknung bei der Optimierung von Formulierungen.<br />

Das Trocknungsverhalten der reinen 10 %igen (m/V) bCA-Lösung und das<br />

verschiedener bCA-Trehalose-Formulierungen zeigt den für amorphe Substanzen<br />

typischen Kurvenverlauf. Die SEM-Aufnahmen zeigen, dass die Partikel mit ansteigendem<br />

Trehalosegehalt eine glattere Oberfläche haben, insgesamt jedoch flacher sind. Levitierte<br />

reine Carboanhydrasepartikel haben ein Loch an der Oberseite und sind hohl, ebenso wie<br />

bei den sprühgetrockneten Pulverpartikeln beobachtet. Die Restaktivität des Proteins steigt<br />

sowohl für die levitierten Partikel als auch für die sprühgetrockneten Pulverproben mit<br />

Trehalosezusatz an. Ein Zusatz an Trehalose von mehr als 20 % des Gesamtfestst<strong>of</strong>fanteils<br />

führt jedoch zu keiner verbesserten Proteinstabilisierung. Die bCA-Trehalose-<br />

Formulierungen wurden in weiteren Versuchen auf die Proteinstabilität während des<br />

Trocknungsvorgangs hin untersucht. Zu verschiedenen Zeitpunkten des<br />

Trocknungsprozesses wurden Tropfen oder Partikel aus dem Levitator entnommen und die<br />

Restaktivität des Proteins bestimmt. Bei der reinen 10 %igen (m/V) bCA-Lösung ist eine<br />

Abnahme der Restaktivität nach dem kritischen Punkt zu sehen. Die stabilisierten


Zusammenfassung 186<br />

Formulierungen zeigen wie erwartet eine nahezu konstante Aktivität während des<br />

Trocknungsprozesses.<br />

LDH wurde als zweites Modellprotein ausgewählt, da es höhere Aktivitätsverluste<br />

während der Trocknung zeigen sollte. Die Aktivitätsmessungen der levitierten Partikel<br />

zeigen jedoch große Abweichungen der Messwerte und können nicht interpretiert werden.<br />

Die sprühgetrockneten Proben der Formulierungen zeigen einen Stabilisierungseffekt<br />

schon bei einem geringen Trehaloseanteil von 20 % des Gesamtfestst<strong>of</strong>fanteils. Die SEM-<br />

Aufnahmen der levitierten Partikel zeigen eine extreme Abflachung mit ansteigendem<br />

Trehalosegehalt. Die aus Tropfen der 10 %igen (m/V) Trehaloselösung hergestellten<br />

Partikel sind ebenfalls sehr flach und „disc“-förmig. Dieses Verhalten ist durch die<br />

Verwendung des 100 mM Kaliumphosphat-Puffers als Lösungspuffer hervorgerufen, denn<br />

Trehalose gelöst in reinem Wasser oder in 50 mM Trizma-Puffer führt zu vorwiegend<br />

abgerundeten Partikeln im Levitator. Die flachen Partikel fingen an zu oszillieren, was die<br />

Durchführung der Messungen erschwerte. Reine LDH-Partikel haben ein Loch an der<br />

Oberseite und sind ebenso wie die reinen bCA-Partikel hohl.<br />

Trypsinogen wurde als drittes Protein untersucht, allerdings konnte keine<br />

Aktivitätsabnahme aufgrund des Trocknungsprozesses festgestellt werden. Das<br />

Trocknungsverhalten von Tropfen der reinen 10 %igen (m/V) Trypsinogenlösung ist<br />

vergleichbar mit den zwei <strong>and</strong>eren Proteinen.<br />

Trocknungsexperimente an Einzeltropfen, die Messungen der<br />

Oberflächentemperatur einschließen, sind ein nützliches Instrument bei der Analyse des<br />

Verdampfungsverhaltens von Lösungsmitteln und des Trocknungsprozesses von<br />

pharmazeutischen Lösungsformulierungen für die Sprühtrocknung. Die Ergebnisse<br />

bezüglich des Aussehens der Partikel und der Restaktivität der Proteine von den<br />

sprühgetrockneten Pulvern und den levitierten Partikeln stimmen gut überein. Messungen<br />

der Restaktivität eines Proteins sind möglich, da die Tropfen oder Partikel jederzeit aus<br />

dem akustischen Feld entnommen werden können. Daher können Formulierungen, die<br />

teure pharmazeutische Proteine enthalten, unter Verwendung sehr kleiner Substanzmengen<br />

optimiert werden. Die Kinetik der Inaktivierung des Proteins während des<br />

Trocknungsprozesses erfordert hingegen weitere Untersuchungen mit einem noch<br />

temperatursensibleren Modellprotein. Des Weiteren sollten Tenside als Hilfsst<strong>of</strong>fe in den<br />

Formulierungen, auch in Kombination mit Zuckern wie Trehalose, auf die<br />

proteinstabilisierende Wirkung aufgrund ihrer Anreicherung an der Tropfenoberfläche


Zusammenfassung 187<br />

untersucht werden. Durch den Einbau einer Kühlung für den Piezokristall im Transducer<br />

könnten die Versuche bei höheren Temperaturen der Trocknungsluft durchgeführt werden.


References 188<br />

8 References<br />

[Adhikari et al. 2000]<br />

[Adler 1999]<br />

[Adler <strong>and</strong> Lee 1999]<br />

[Adler et al. 2000]<br />

[Allison et al. 1999]<br />

Adhikari, B., Howes, T., Bh<strong>and</strong>ari, B. R. <strong>and</strong> Truong,<br />

V. (2000). Experimental studies <strong>and</strong> <strong>kinetics</strong> <strong>of</strong> single<br />

drop <strong>drying</strong> <strong>and</strong> their relevance in <strong>drying</strong> <strong>of</strong> sugarrich<br />

foods: A review. International Journal <strong>of</strong> Food<br />

Properties 3 (3): 323-351.<br />

Adler, M. (1999). Sprüheinbettung von Proteinen in<br />

Gerüstbildner: Stabilität und Oberflächenanalyse.<br />

Division <strong>of</strong> Pharmaceutics. Friedrich-Alex<strong>and</strong>er-<br />

University, Erlangen, Germany. Ph.D. Thesis.<br />

Adler, M. <strong>and</strong> Lee, G. (1999). Stability <strong>and</strong> <strong>surface</strong><br />

activity <strong>of</strong> lactate dehydrogenase in spray-dried<br />

trehalose. Journal <strong>of</strong> Pharmaceutical Sciences 88 (2):<br />

199-208.<br />

Adler, M., Unger, M. <strong>and</strong> Lee, G. (2000). Surface<br />

composition <strong>of</strong> spray-dried particles <strong>of</strong> bovine serum<br />

albumin/trehalose/surfactant. Pharmaceutical<br />

Research 17 (7): 863-870.<br />

Allison, S. D., Chang, B., R<strong>and</strong>olph, T. W. <strong>and</strong><br />

Carpenter, J. F. (1999). Hydrogen bonding between<br />

sugar <strong>and</strong> protein is responsible for inhibition <strong>of</strong><br />

dehydration-induced protein unfolding. Archives <strong>of</strong><br />

Biochemistry <strong>and</strong> Biophysics 365 (2): 289-298.<br />

[Almer et al. 2002] Almer, L.-O., Wollmer, P., Jonson, B. <strong>and</strong> Almer, A. T.<br />

(2002). Insulin inhalation with absorption enhancer<br />

at meal-times results in almost normal postpr<strong>and</strong>ial<br />

insulin pr<strong>of</strong>iles. Clinical Physiology <strong>and</strong> Functional<br />

Imaging 22 (3): 218-221.<br />

[Andya et al. 1999]<br />

Andya, J. D., Maa, Y.-F., Costantino, H. R., Nguyen, P.-<br />

A., Dasovich, N., Sweeney, T. D., Hsu, C. C. <strong>and</strong> Shire,<br />

S. J. (1999). The effect <strong>of</strong> formulation excipients on<br />

protein stability <strong>and</strong> aerosol performance <strong>of</strong> spraydried<br />

powders <strong>of</strong> a recombinant humanized anti-IgE<br />

monoclonal antibody. Pharmaceutical Research 16 (3):<br />

350-358.


References 189<br />

[Arakawa et al. 2001]<br />

[Baehr <strong>and</strong> Stephan 2004]<br />

[Banga 1995]<br />

Arakawa, T., Prestrelski, S. J., Kenney, W. C. <strong>and</strong><br />

Carpenter, J. F. (2001). Factors affecting short-term<br />

<strong>and</strong> long-term stabilities <strong>of</strong> proteins. Advanced Drug<br />

Delivery Reviews 46 (1-3): 307-326.<br />

Baehr, H. D. <strong>and</strong> Stephan, K. (2004). Wärme- und<br />

St<strong>of</strong>fübertragung. Springer Verlag, Berlin, Heidelberg,<br />

Germany.<br />

Banga, A. K. (1995). Therapeutic Peptides <strong>and</strong><br />

Proteins: Formulation, Processing, <strong>and</strong> Delivery<br />

Systems. Technomic Publishing Company, Lancaster,<br />

USA.<br />

[BCG 2008] BCG (2008). Medizinische Biotechnologie in<br />

Deutschl<strong>and</strong> 2008. The Boston Consulting Group<br />

GmbH, München, Germany.<br />

[Belton <strong>and</strong> Gil 1994]<br />

Belton, P. S. <strong>and</strong> Gil, A. M. (1994). IR <strong>and</strong> Raman<br />

spectroscopic studies <strong>of</strong> the interaction <strong>of</strong> trehalose<br />

with hen egg white lysozyme. Biopolymers 34 (7):<br />

957-961.<br />

[Berg et al. 2007] Berg, J. M., Stryer, L. <strong>and</strong> Tymoczko, J. L. (2007).<br />

Stryer Biochemie. Elsevier GmbH, Spektrum<br />

Akademischer Verlag, Heidelberg, München, Germany.<br />

[Bosnjakovic 1997] Bosnjakovic, F. (1997). Technische Thermodynamik -<br />

Teil 2. Dr. Dietrich Steinkopff Verlag GmbH & Co.<br />

KG, Darmstadt, Germany.<br />

[Büchi 1999] Büchi (1999). Büchi Mini Spray Dryer B-191<br />

Manual. Büchi Labortechnik AG, Flawil, Switzerl<strong>and</strong>.<br />

[Burdukov et al. 1963] Burdukov, A. P., Elchin, V. I. <strong>and</strong> Nakoryakov, V. E.<br />

(1963). Some mass transfer problems in an<br />

ultrasonic field. Kinetika Goreniya Iskop. Topliv,<br />

Akad. Nauk SSSR, Sibirsk. Otd., Khim.-Met. Inst.: 97-<br />

109.<br />

[Charlesworth <strong>and</strong> Marshall 1960] Charlesworth, D. H. <strong>and</strong> Marshall, W. R., Jr. (1960).<br />

Evaporation from drops containing dissolved solids.<br />

AIChE Journal 6: 9-23.


References 190<br />

[Cheong et al. 1986] Cheong, H. W., Jeffreys, G. V. <strong>and</strong> Mumford, C. J.<br />

(1986). A receding interface model for the <strong>drying</strong> <strong>of</strong><br />

slurry <strong>droplet</strong>s. AIChE Journal 32 (8): 1334-1346.<br />

[Cox <strong>and</strong> Phillips 2008]<br />

Cox, M. M. <strong>and</strong> Phillips, G. N. J. (2008). H<strong>and</strong>book <strong>of</strong><br />

Proteins: Structure, Functions <strong>and</strong> Methods. John<br />

Wiley & Sons, Chichester, Engl<strong>and</strong>.<br />

[Cronin et al. 2001] Cronin, L., Foxon, S. P., Lusby, P. J. <strong>and</strong> Walton, P. H.<br />

(2001). Syntheses <strong>and</strong> structures <strong>of</strong> M(L)(X)BPh4<br />

complexes {M = Co(II), Zn(II); L = cis-1,3,5-tris[3-<br />

(2-furyl)prop-2-enylideneamino]cyclohexane, X =<br />

OAc, NO3}: structural models <strong>of</strong> the active site <strong>of</strong><br />

carbonic anhydrase. JBIC, Journal <strong>of</strong> Biological<br />

Inorganic Chemistry 6 (4): 367-377.<br />

[Daif et al. 1999] Daif, A., Bouaziz, M., Bresson, J. <strong>and</strong> Grisenti, M.<br />

(1999). Surface <strong>temperature</strong> <strong>of</strong> hydrocarbon <strong>droplet</strong><br />

in evaporation. Journal <strong>of</strong> Thermophysics <strong>and</strong> Heat<br />

Transfer 13 (4): 553-555.<br />

[Dean <strong>and</strong> Chen 2004]<br />

[Eckert 1966]<br />

Dean, H. J. <strong>and</strong> Chen, D. (2004). Epidermal powder<br />

immunization against influenza. Vaccine 23 (5): 681-<br />

686.<br />

Eckert, R. G. (1966). Einführung in den Wärme- und<br />

St<strong>of</strong>faustausch. Springer Verlag, Berlin, Heidelberg,<br />

Germany.<br />

[El-Sayed et al. 1990] El-Sayed, T. M., Wallack, D. A. <strong>and</strong> King, C. J. (1990).<br />

Changes in particle morphology during <strong>drying</strong> <strong>of</strong><br />

drops <strong>of</strong> carbohydrate solutions <strong>and</strong> food liquids. 1.<br />

Effect <strong>of</strong> composition <strong>and</strong> <strong>drying</strong> conditions.<br />

Industrial & Engineering Chemistry Research 29 (12):<br />

2346-2354.<br />

[Elversson et al. 2003]<br />

Elversson, J., Millqvist-Fureby, A., Alderborn, G. <strong>and</strong><br />

El<strong>of</strong>sson, U. (2003). Droplet <strong>and</strong> particle size<br />

relationship <strong>and</strong> shell thickness <strong>of</strong> inhalable lactose<br />

particles during spray <strong>drying</strong>. Journal <strong>of</strong><br />

Pharmaceutical Sciences 92 (4): 900-910.<br />

[Elversson <strong>and</strong> Millqvist-F. 2005] Elversson, J. <strong>and</strong> Millqvist-F., A. (2005). Particle size<br />

<strong>and</strong> density in spray <strong>drying</strong>-effects <strong>of</strong> carbohydrate<br />

properties. Journal <strong>of</strong> Pharmaceutical Sciences 94 (9):<br />

2049-2060.


References 191<br />

[Engineering Toolbox 2009] The Engineering Toolbox (2009).<br />

http://www.engineeringtoolbox.com<br />

[Farid 2003]<br />

[Frohn <strong>and</strong> Roth 2000]<br />

Farid, M. (2003). A new approach to modelling <strong>of</strong><br />

single <strong>droplet</strong> <strong>drying</strong>. Chemical Engineering Science<br />

58 (13): 2985-2993.<br />

Frohn, A. <strong>and</strong> Roth, N. (2000). Dynamics <strong>of</strong> <strong>droplet</strong>s.<br />

Springer Verlag, Berlin, Heidelberg, Germany.<br />

[Fuller et al. 1966] Fuller, E. N., Schettler, P. D. <strong>and</strong> Giddings, J. C. (1966).<br />

A new method for prediction <strong>of</strong> binary gas-phase<br />

diffusion coefficients. Journal <strong>of</strong> Industrial <strong>and</strong><br />

Engineering Chemistry (Washington, D. C.) 58 (5): 18-<br />

27.<br />

[Gil et al. 1996] Gil, A. M., Belton, P. S. <strong>and</strong> Felix, V. (1996).<br />

Spectroscopic studies <strong>of</strong> solid a-a trehalose.<br />

Spectrochimica Acta, Part A: Molecular <strong>and</strong><br />

Biomolecular Spectroscopy 52A (12): 1649-1659.<br />

[Grassmann et al. 1997] Grassmann, P., Widmer, F. <strong>and</strong> Sinn, H. (1997).<br />

Einführung in die thermische Verfahrenstechnik.<br />

Walter de Gruyter, Berlin, Germany.<br />

[Günther et al. 1978] Günther, B. C., Hansen, K.-H. <strong>and</strong> Veit, I. (1978).<br />

Technische Akustik. Lexika-Verlag, Grafenau,<br />

Germany.<br />

[Hagen et al. 1995] Hagen, S. J., H<strong>of</strong>richter, J. <strong>and</strong> Eaton, W. A. (1995).<br />

Protein reaction <strong>kinetics</strong> in a room-<strong>temperature</strong><br />

glass. Science (Washington, D. C.) 269 (5226): 959-<br />

962.<br />

[Hancock <strong>and</strong> Zografi 1997]<br />

[Hering et al. 2007]<br />

Hancock, B. C. <strong>and</strong> Zografi, G. (1997). Characteristics<br />

<strong>and</strong> significance <strong>of</strong> the amorphous state in<br />

pharmaceutical systems. Journal <strong>of</strong> Pharmaceutical<br />

Sciences 86 (1): 1-12.<br />

Hering, E., Martin, R. <strong>and</strong> Stohrer, M. (2007). Physik<br />

für Ingenieure. Springer Verlag, Berlin, Heidelberg,<br />

Germany.<br />

[Higashiyama 2002] Higashiyama, T. (2002). Novel functions <strong>and</strong><br />

applications <strong>of</strong> trehalose. Pure <strong>and</strong> Applied Chemistry<br />

74 (7): 1263-1269.


References 192<br />

[InfraTec 2009]<br />

[Karlson 1988]<br />

[Kastner 2001]<br />

[Lierke 1996a]<br />

InfraTec GmbH (2009). http://www.InfraTec.de<br />

Karlson, P. (1988). Kurzes Lehrbuch der Biochemie<br />

für Mediziner und Naturwissenschaftler. Georg<br />

Thieme Verlag, Stuttgart, Germany.<br />

Kastner, O. (2001). Theoretische und experimentelle<br />

Untersuchungen zum St<strong>of</strong>fübergang von<br />

Einzeltropfen in einem akustischen Rohrlevitator.<br />

Institute <strong>of</strong> Fluid Mechanics. Friedrich-Alex<strong>and</strong>er-<br />

University, Erlangen, Germany. Ph.D. Thesis.<br />

Lierke, E. G. (1996a). Akustische Positionierung - Ein<br />

umfassender Überblick über Grundlagen und<br />

Anwendungen. Acta Acustica 82: 220-237.<br />

[Lierke 1996b] Lierke, E. G. (1996b). Kontrollierte<br />

Massenänderungen von Tropfen in einem<br />

akustischen Stehwellen-Positionierer. Forschung im<br />

Ingenieurwesen 62 (1/2): 21-31.<br />

[Lierke 2002]<br />

[Lin <strong>and</strong> Gentry 2003]<br />

Lierke, E. G. (2002). Deformation <strong>and</strong> displacement<br />

<strong>of</strong> liquid drops in an optimized acoustic st<strong>and</strong>ing<br />

wave levitator. Acta Acustica united with Acustica 88<br />

(2): 206-217.<br />

Lin, J.-C. <strong>and</strong> Gentry, J. W. (2003). Spray <strong>drying</strong> drop<br />

morphology: experimental study. Aerosol Science <strong>and</strong><br />

Technology 37 (1): 15-32.<br />

[Lins et al. 2004] Lins, R. D., Pereira, C. S. <strong>and</strong> Huenenberger, P. H.<br />

(2004). Trehalose-protein interaction in aqueous<br />

solution. Proteins: Structure, Function, <strong>and</strong><br />

Bioinformatics 55 (1): 177-186.<br />

[Maa et al. 1997]<br />

Maa, Y.-F., Costantino, H. R., Nguyen, P.-A. <strong>and</strong> Hsu,<br />

C. C. (1997). The effect <strong>of</strong> operating <strong>and</strong> formulation<br />

variables on the morphology <strong>of</strong> spray-dried protein<br />

particles. Pharmaceutical Development <strong>and</strong> Technology<br />

2 (3): 213-223.<br />

[Maa <strong>and</strong> Hsu 1997] Maa, Y.-F. <strong>and</strong> Hsu, C. C. (1997). Protein<br />

denaturation by combined effect <strong>of</strong> shear <strong>and</strong> airliquid<br />

interface. Biotechnology <strong>and</strong> Bioengineering 54<br />

(6): 503-512.


References 193<br />

[Maa <strong>and</strong> Prestrelski 2000] Maa, Y.-F. <strong>and</strong> Prestrelski, S. J. (2000).<br />

Biopharmaceutical powders: particle formation <strong>and</strong><br />

formulation considerations. Current Pharmaceutical<br />

Biotechnology 1 (3): 283-302.<br />

[Manning et al. 1989] Manning, M. C., Patel, K. <strong>and</strong> Borchardt, R. T. (1989).<br />

Stability <strong>of</strong> protein pharmaceuticals. Pharmaceutical<br />

Research 6 (11): 903-918.<br />

[Masters 1991] Masters, K. (1991). Spray Drying H<strong>and</strong>book.<br />

Longman Scientific & Technical, Harlow, Engl<strong>and</strong>.<br />

[Masters 2002]<br />

Masters, K. (2002). Spray Drying in Practice.<br />

SprayDryConsult International ApS, Charlottenlund,<br />

Denmark.<br />

[Maury et al. 2005a] Maury, M., Murphy, K., Kumar, S., Shi, L. <strong>and</strong> Lee, G.<br />

(2005a). Effects <strong>of</strong> process variables on the powder<br />

yield <strong>of</strong> spray-dried trehalose on a laboratory spraydryer.<br />

European Journal <strong>of</strong> Pharmaceutics <strong>and</strong><br />

Biopharmaceutics 59 (3): 565-573.<br />

[Maury et al. 2005b]<br />

Maury, M., Murphy, K., Kumar, S., Mauerer, A. <strong>and</strong><br />

Lee, G. (2005b). Spray-<strong>drying</strong> <strong>of</strong> proteins: effects <strong>of</strong><br />

sorbitol <strong>and</strong> trehalose on aggregation <strong>and</strong> FT-IR<br />

amide I spectrum <strong>of</strong> an immunoglobulin G. European<br />

Journal <strong>of</strong> Pharmaceutics <strong>and</strong> Biopharmaceutics 59 (2):<br />

251-261.<br />

[McCabe et al. 2005] McCabe, W. L., Smith, J. C. <strong>and</strong> Harriott, P. (2005).<br />

Unit operations <strong>of</strong> chemical engineering. McGraw-<br />

Hill Book Company, New York, USA.<br />

[microdrop 2009] microdrop Technologies GmbH (2009).<br />

http://www.microdrop.de<br />

[Mumenthaler et al. 1994] Mumenthaler, M., Hsu, C. C. <strong>and</strong> Pearlman, R. (1994).<br />

Feasibility study on spray-<strong>drying</strong> protein<br />

pharmaceuticals: recombinant human growth<br />

hormone <strong>and</strong> tissue-type plasminogen activator.<br />

Pharmaceutical Research 11 (1): 12-20.<br />

[Mutschler et al. 2001]<br />

Mutschler, E., Geisslinger, G., Kroemer, H. K. <strong>and</strong><br />

Schäfer-Korting, M. (2001). Mutschler<br />

Arzneimittelwirkungen.<br />

Wissenschaftliche<br />

Verlagsgesellschaft mbH, Stuttgart, Germany.


References 194<br />

[Nesic <strong>and</strong> Vodnik 1991]<br />

Nesic, S. <strong>and</strong> Vodnik, J. (1991). Kinetics <strong>of</strong> <strong>droplet</strong><br />

evaporation. Chemical Engineering Science 46 (2):<br />

527-537.<br />

[PDB 2009] RCSB Protein Data Bank (2009).<br />

www.pdb.org/pdb/home/home.do<br />

[Ranz <strong>and</strong> Marshall 1952a]<br />

[Ranz <strong>and</strong> Marshall 1952b]<br />

Ranz, W. E. <strong>and</strong> Marshall, W. R., Jr. (1952a).<br />

Evaporation from drops. I. Chemical Engineering<br />

Progress 48: 141-146.<br />

Ranz, W. E. <strong>and</strong> Marshall, W. R., Jr. (1952b).<br />

Evaporation from drops. II. Chemical Engineering<br />

Progress 48: 173-180.<br />

[Reid et al. 1987] Reid, R. C., Prausnitz, J. M. <strong>and</strong> Poling, B. E. (1987).<br />

The properties <strong>of</strong> liquids <strong>and</strong> gases. McGraw-Hill<br />

Book Company, New York, USA.<br />

[Rensink 2004]<br />

[Richards et al. 2002]<br />

Rensink, D. (2004). Verdunstung akustisch levitierter<br />

schwingender Tropfen aus homogenen und<br />

heterogenen Medien. Institute <strong>of</strong> Fluid Mechanics.<br />

Friedrich-Alex<strong>and</strong>er-University, Erlangen, Germany.<br />

Ph.D. Thesis.<br />

Richards, A. B., Krakowka, S., Dexter, L. B., Schmid,<br />

H., Wolterbeek, A. P. M., Waalkens-Berendsen, D. H.,<br />

Shigoyuki, A. <strong>and</strong> Kurimoto, M. (2002). Trehalose: a<br />

review <strong>of</strong> properties, history <strong>of</strong> use <strong>and</strong> human<br />

tolerance, <strong>and</strong> results <strong>of</strong> multiple safety studies. Food<br />

<strong>and</strong> Chemical Toxicology 40 (7): 871-898.<br />

[Roche 2009] Roche Deutschl<strong>and</strong> Holding GmbH (2009).<br />

http://www.roche.de<br />

[Sano <strong>and</strong> Keey 1982]<br />

[Schiffter 2006]<br />

Sano, Y. <strong>and</strong> Keey, R. B. (1982). The <strong>drying</strong> <strong>of</strong> a<br />

spherical particle containing colloidal material into a<br />

hollow sphere. Chemical Engineering Science 37 (6):<br />

881-889.<br />

Schiffter, H. (2006). Single <strong>droplet</strong> <strong>drying</strong> <strong>of</strong> proteins<br />

<strong>and</strong> protein formulations via acoustic levitation.<br />

Division <strong>of</strong> Pharmaceutics. Friedrich-Alex<strong>and</strong>er-<br />

University, Erlangen, Germany. Ph.D. Thesis.


References 195<br />

[Schiffter <strong>and</strong> Lee 2007a]<br />

[Schiffter <strong>and</strong> Lee 2007b]<br />

Schiffter, H. <strong>and</strong> Lee, G. (2007a). Single-<strong>droplet</strong><br />

evaporation <strong>kinetics</strong> <strong>and</strong> particle formation in an<br />

acoustic levitator. Part 1: evaporation <strong>of</strong> water<br />

micro<strong>droplet</strong>s assessed using boundary-layer <strong>and</strong><br />

acoustic levitation theories. Journal <strong>of</strong> Pharmaceutical<br />

Sciences 96 (9): 2274-2283.<br />

Schiffter, H. <strong>and</strong> Lee, G. (2007b). Single-<strong>droplet</strong><br />

evaporation <strong>kinetics</strong> <strong>and</strong> particle formation in an<br />

acoustic levitator. Part 2: <strong>drying</strong> <strong>kinetics</strong> <strong>and</strong><br />

particle formation from micro<strong>droplet</strong>s <strong>of</strong> aqueous<br />

mannitol, trehalose, or catalase. Journal <strong>of</strong><br />

Pharmaceutical Sciences 96 (9): 2284-2295.<br />

[Schuster <strong>and</strong> Kolobrodov 2004] Schuster, N. <strong>and</strong> Kolobrodov, V. G. (2004).<br />

Infrarotthermographie. WILEY-VCH Verlag GmbH<br />

& Co. KGaA, Weinheim, Germany.<br />

[Seaver et al. 1989] Seaver, M., Galloway, A. <strong>and</strong> Manuccia, T. J. (1989).<br />

Acoustic levitation in a free-jet wind tunnel. Review<br />

<strong>of</strong> Scientific Instruments 60 (11): 3452-3459.<br />

[Sigma-Aldrich 2009]<br />

[Stroppe 2008]<br />

Sigma-Aldrich (2009). http://www.sigmaaldrich.com<br />

Stroppe, H. (2008). Physik für Studierende der<br />

Natur- und Ingenieurwissenschaften. Fachbuchverlag<br />

Leipzig im Carl Hanser Verlag, München, Germany.<br />

[Tuckermann et al. 2002] Tuckermann, R., Bauerecker, S. <strong>and</strong> Neidhart, B.<br />

(2002). Evaporation rates <strong>of</strong> alkanes <strong>and</strong> alkanols<br />

from acoustically levitated drops. Analytical <strong>and</strong><br />

Bioanalytical Chemistry 372 (1): 122-127.<br />

[Tuckermann 2002]<br />

Tuckermann, R. (2002). Gase, Aerosole, Tropfen und<br />

Partikel in stehenden Ultraschallfeldern: Eine<br />

Untersuchung zur Anreicherung schwerer Gase,<br />

Verdampfung levitierter Tropfen, Kristall- und<br />

Partikelbildung. Faculty <strong>of</strong> Science. Technical<br />

University Carolo-Wilhelmina, Braunschweig,<br />

Germany. Ph.D. Thesis.<br />

[Tuckermann et al. 2005] Tuckermann, R., Bauerecker, S. <strong>and</strong> Cammenga, H. K.<br />

(2005). IR-thermography <strong>of</strong> evaporating acoustically<br />

levitated drops. International Journal <strong>of</strong> Thermophysics<br />

26 (5): 1583-1594.


References 196<br />

[Tuckermann et al. 2007] Tuckermann, R., Bauerecker, S. <strong>and</strong> Cammenga, H. K.<br />

(2007). Generation <strong>and</strong> characterization <strong>of</strong> <strong>surface</strong><br />

layers on acoustically levitated drops. Journal <strong>of</strong><br />

Colloid <strong>and</strong> Interface Science 310 (2): 559-569.<br />

[Tzannis <strong>and</strong> Prestrelski 1999]<br />

Tzannis, S. T. <strong>and</strong> Prestrelski, S. J. (1999). Activitystability<br />

considerations <strong>of</strong> trypsinogen during spray<br />

<strong>drying</strong>: Effects <strong>of</strong> sucrose. Journal <strong>of</strong> Pharmaceutical<br />

Sciences 88 (3): 351-359.<br />

[Verpoorte et al. 1967] Verpoorte, J. A., Mehta, S. <strong>and</strong> Edsall, J. T. (1967).<br />

Esterase activities <strong>of</strong> human carbonic anhydrases B<br />

<strong>and</strong> C. The Journal <strong>of</strong> Biological Chemistry 242 (18):<br />

4221-4229.<br />

[Walton <strong>and</strong> Mumford 1999]<br />

[Weis <strong>and</strong> Nardozzi 2005]<br />

Walton, D. E. <strong>and</strong> Mumford, C. J. (1999). The<br />

morphology <strong>of</strong> spray-dried particles. The effect <strong>of</strong><br />

process variables upon the morphology <strong>of</strong> spraydried<br />

particles. Chemical Engineering Research <strong>and</strong><br />

Design 77 (A5): 442-460.<br />

Weis, D. D. <strong>and</strong> Nardozzi, J. D. (2005). Enzyme<br />

<strong>kinetics</strong> in acoustically levitated <strong>droplet</strong>s <strong>of</strong><br />

supercooled water: A novel approach to<br />

cryoenzymology. Analytical Chemistry 77 (8): 2558-<br />

2563.<br />

[Wijlhuizen et al. 1979] Wijlhuizen, A. E., Kerkh<strong>of</strong>, P. J. A. M. <strong>and</strong> Bruin, S.<br />

(1979). Theoretical study <strong>of</strong> the inactivation <strong>of</strong><br />

phosphatase during spray <strong>drying</strong> <strong>of</strong> skim-milk.<br />

Chemical Engineering Science 34 (5): 651-660.<br />

[Yarin et al. 1998] Yarin, A. L., Pfaffenlehner, M. <strong>and</strong> Tropea, C. (1998).<br />

On the acoustic levitation <strong>of</strong> <strong>droplet</strong>s. Journal <strong>of</strong> Fluid<br />

Mechanics 356: 65-91.<br />

[Yarin et al. 1999]<br />

Yarin, A. L., Brenn, G., Kastner, O., Rensink, D. <strong>and</strong><br />

Tropea, C. (1999). Evaporation <strong>of</strong> acoustically<br />

levitated <strong>droplet</strong>s. Journal <strong>of</strong> Fluid Mechanics 399:<br />

151-204.<br />

[Yarin et al. 2002] Yarin, A. L., Brenn, G. <strong>and</strong> Rensink, D. (2002).<br />

Evaporation <strong>of</strong> acoustically levitated <strong>droplet</strong>s <strong>of</strong><br />

binary liquid mixtures. International Journal <strong>of</strong> Heat<br />

<strong>and</strong> Fluid Flow 23 (4): 471-486.


References 197<br />

[Yaws 2003] Yaws, C. L. (2003). Yaws' H<strong>and</strong>book <strong>of</strong><br />

Thermodynamic <strong>and</strong> Physical Properties <strong>of</strong><br />

Chemical Compounds. Knovel, www.knovel.com.<br />

[Yu <strong>and</strong> Chien 1997]<br />

Yu, J. <strong>and</strong> Chien, Y. W. (1997). Pulmonary drug<br />

delivery: physiologic <strong>and</strong> mechanistic aspects.<br />

Critical Reviews in Therapeutic Drug Carrier Systems<br />

14 (4): 395-453.<br />

[Yu et al. 1998] Yu, L., Mishra, D. S. <strong>and</strong> Rigsbee, D. R. (1998).<br />

Determination <strong>of</strong> the glass properties <strong>of</strong> D-mannitol<br />

using sorbitol as an impurity. Journal <strong>of</strong><br />

Pharmaceutical Sciences 87 (6): 774-777.


Lebenslauf<br />

Persönliche Daten<br />

Name:<br />

Eva Cornelia Wulsten, geb. Schmidt<br />

Geburtsdatum/-ort: 11.04.1979 in Delmenhorst<br />

Anschrift: Saßnitzer Straße 19<br />

90425 Nürnberg<br />

Familienst<strong>and</strong>:<br />

verheiratet<br />

Schulbildung<br />

1985 - 1989 Grundschule G<strong>and</strong>erkesee-Bookholzberg<br />

1989 - 1991 Orientierungsstufe G<strong>and</strong>erkesee-Bookholzberg<br />

1991 - 1998 Gymnasium G<strong>and</strong>erkesee<br />

Abschluss mit Abitur am 22.06.1998<br />

Studium<br />

10/1998 - 11/2002 Studium der Pharmazie an der Technischen Universität<br />

Carolo-Wilhelmina zu Braunschweig<br />

1. Abschnitt der Pharm. Prüfung am 25.08.2000<br />

2. Abschnitt der Pharm. Prüfung am 11.11.2002<br />

3. Abschnitt der Pharm. Prüfung am 08.01.2004<br />

Approbation als Apothekerin am 21.01.2004<br />

Praktika<br />

03/99 Famulatur in der Stern-Apotheke Bookholzberg<br />

08/99 Famulatur in der Zentralapotheke des Zentralkrankenhauses<br />

St.-Jürgen-Straße in Bremen<br />

12/2002 - 05/2003 Pharmaziepraktikum bei der Lilly Forschung GmbH in<br />

Hamburg, Abteilung Pharmazeutische Entwicklung<br />

(Analytik)<br />

06/2003 - 11/2003 Pharmaziepraktikum in der Apotheke und Sanitätshaus<br />

Gamsen, Gifhorn-Gamsen


Berufliche Tätigkeiten<br />

01/2004 - 06/2005 Apothekerin in der Apotheke und Sanitätshaus Gamsen in<br />

Gifhorn-Gamsen<br />

07/2005 - 09/2005 Apothekerin in der Apotheke im Hansa-Carré in Bremen<br />

10/2005 - 02/2009 Wissenschaftliche Mitarbeiterin und Doktor<strong>and</strong>in am<br />

Lehrstuhl für Pharmazeutische Technologie der Friedrich-<br />

Alex<strong>and</strong>er-Universität Erlangen-Nürnberg<br />

Anfertigung der Doktorarbeit zum Thema „Determination<br />

<strong>of</strong> <strong>droplet</strong> <strong>surface</strong> <strong>temperature</strong> <strong>and</strong> <strong>drying</strong> <strong>kinetics</strong> <strong>of</strong><br />

protein solutions using an ultrasonic levitator“<br />

05/2006 - 08/2008 zeitweise Nebentätigkeit in Apotheken in Nürnberg und<br />

Übernahme von Apothekennotdiensten<br />

Weiterbildung<br />

seit 02/2006<br />

Weiterbildung zum Fachapotheker für Pharmazeutische<br />

Technologie<br />

Nürnberg, 20.02.2009

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!