15.02.2013 Views

Single droplet drying of proteins and protein formulations

Single droplet drying of proteins and protein formulations

Single droplet drying of proteins and protein formulations

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.

SINGLE DROPLET DRYING OF<br />

PROTEINS AND PROTEIN FORMULATIONS<br />

VIA ACOUSTIC LEVITATION<br />

Den naturwissenschaftlichen Fakultäten<br />

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

zur<br />

Erlangung des Doktorgrades<br />

vorgelegt von<br />

Heiko Alex<strong>and</strong>er Schiffter<br />

aus Heilbronn-Neckargartach


Als Dissertation genehmigt von den<br />

naturwissenschaftlichen Fakultäten der Universität Erlangen-Nürnberg<br />

Tag der mündlichen Prüfung:<br />

Vorsitzender der Promotionskommision: Pr<strong>of</strong>. Dr. D.-P. Häder<br />

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

Zweitberichterstatter: Pr<strong>of</strong>. Dr. Anker Jensen


To my family<br />

in love <strong>and</strong> gratitude<br />

Schreib in den S<strong>and</strong>, die dich betrüben,<br />

Vergiß dann schnell, und schlafe drüber ein,<br />

Denn was du in den S<strong>and</strong> geschrieben,<br />

Das wird schon morgen nicht mehr sein.<br />

Schreib in den Stein, was du erfahren<br />

An Liebe, Seligkeit und Glück.<br />

Es gibt der Stein dir auch nach Jahren<br />

Dies als Erinnerung zurück.


ACKNOWLEDGEMENTS<br />

The present thesis has been acquired between August 2002 <strong>and</strong> December 2005 at the Department<br />

<strong>of</strong> Pharmaceutical Technology, Friedrich-Alex<strong>and</strong>er-University Nuremberg-Erlangen, Germany.<br />

First <strong>of</strong> all, Pr<strong>of</strong>essor Dr. Ge<strong>of</strong>frey Lee is gratefully acknowledged for <strong>of</strong>fering me the opportunity<br />

to work in this very interesting field <strong>of</strong> research, for many constructive discussions <strong>and</strong> support<br />

during this period, for promoting additional research projects, for the opportunity to develop a<br />

complete lecture in pharmacokinetics <strong>and</strong> for the friendly atmosphere within the department.<br />

Pr<strong>of</strong>essor Dr. Anker Jensen, Department <strong>of</strong> Chemical Engineering, Technical University <strong>of</strong><br />

Denmark, is thanked for kindly being co-referee for this thesis.<br />

Many thanks to Pr<strong>of</strong>essor Dr. Günther Brenn, Department <strong>of</strong> Chemical Engineering, Technical,<br />

University <strong>of</strong> Graz, Austria <strong>and</strong> Dr. Dirk Rensink, Adam Opel AG, Germany, for their help <strong>and</strong><br />

support with the acoustic levitation system <strong>and</strong> the knowledge they shared with me. They always<br />

found time to answer all <strong>of</strong> my questions in detail.<br />

Christian Schwartzbach <strong>and</strong> Sune Klint Anderson, Niro A/S, Denmark is thanked for gratefully<br />

supporting this project <strong>and</strong> for giving me new inspirations by discussing many critical details <strong>of</strong> the<br />

evaporation process. Also thank you very much for the invitation <strong>and</strong> the visit in Copenhagen.<br />

Thank you very much to Dr. Rudolf Tuckermann for helping me with the acoustic levitator,<br />

providing me with his thesis <strong>and</strong> giving me new inspirations when visiting my laboratory in<br />

Erlangen.<br />

Dr. Gerhard Simon <strong>and</strong> Dr. Stefan Seyferth is thanked for their support in problems with computer<br />

hardware <strong>and</strong> s<strong>of</strong>tware.<br />

Innerhalb des Instituts für pharmazeutische Technologie in Erlangen möchte ich mich zuerst bei<br />

Dr. Michael Maury für seine langjährige Freundschaft während des gemeinsamen Studiums in<br />

Heidelberg und der Promotion in Erlangen, für die vielen Diskussionen über Wissenschaft und<br />

Studentenpraktika, die Zeit im „Heidelberger Labor“ und die lustigen Abende Downtown<br />

bedanken. Vielen Dank an Dr. Henning Gieseler für die gemeinsame Zeit in Erlangen, die Hilfe und<br />

Ratschläge bei wissenschaftlichen Fragen, die kulinarischen Abende in Nürnberg und Umgebung<br />

und eine unvergesslich Woche auf dem AAPS-Kongress in Nashville. Danke auch an Alex<strong>and</strong>er<br />

Mauerer für die Hilfe bei jeder Art von S<strong>of</strong>t- und Hardwareproblemen, aber vor allem für die vielen


Stunden innerhalb und außerhalb der Universität, wie auch für die immer exakten<br />

Wegbeschreibungen zu den besten Locations der Region. Vielen Dank an Andreas Ziegler und<br />

Peter Lassner, dem selbsternannten „Labor Harmonie“, für die vielen <strong>of</strong>t lustigen, manchmal auch<br />

ernsthaften Gespräche während den diversen Erholungspausen und die Möglichkeit immer eine<br />

Anlaufstelle zu haben. Vielen Dank auch an Jürgen Bögelein und Henning Wegner für die vielen<br />

Next-door-Events während und außerhalb des Semesters und natürlich auch für die gemeinsame<br />

Solida-Betreuung. Danke an Dr. Stefan Seyferth, Dr. Christian Führling und Dr. Christian Rochelle<br />

für die gemeinsame Zeit im Institut, beim Tisch-Fußball oder bei diversen <strong>and</strong>eren Veranstaltungen.<br />

Bei Harald Pudritz möchte ich mich für die gemeinsame Zeit im Labor bedanken und natürlich für<br />

die vielen netten Geschichten und Sprüche.<br />

Euch allen: Ihr seid mir in der gemeinsamen Zeit in Erlangen sehr ans Herz gewachsen. Ich h<strong>of</strong>fe<br />

sehr wir bleiben ein Leben lang Freunde und werden uns noch <strong>of</strong>t wieder sehen.<br />

Danke auch an meine weiteren Kolleginnen und Kollegen Joanna Sawiec, Eva Meister,<br />

Silja von Graberg, Eva Schmidt, Anke Czerwinski, Dr. Doris Köpper und Dr. Marc Fitzner für die<br />

gemeinsamen Stunden im Praktikum und manche nette Unterhaltung.<br />

Ein ganz großes Dankeschön geht an Winfried Bauer und Josef Hubert, ohne deren h<strong>and</strong>werkliches<br />

Geschick die praktische Umsetzung des Levitationssystem nicht möglich gewesen wäre, an Luise<br />

Schedl für die Unmengen rasterelektonenmikroskopischer Bilder einzelner Partikel, an Petra<br />

Neubarth und Waltraut Klenk für alle administrativen Tätigkeiten und an Petra speziell für die<br />

Mithilfe beim Erstellen von Praktikumsplänen und der Pharmakokinetik-Klausur, an Christiane<br />

Blaha für die Bestellungen von Chemikalien, Glasgeräten und unzähligen Mikroliterspritzen und an<br />

Eberhardt Nürnberg für interessante Einblicke in die Erlangener Pharmazie lange vor unserer Zeit.<br />

Außerhalb der Universität möchte ich vor allem meiner Freundin Martina für ihre Liebe und für<br />

ihre Unterstützung meiner Arbeit in den letzten drei Jahren sowie natürlich auch für das<br />

Korrekturlesen danken.<br />

Zu guter letzt gilt mein größter Dank aber meiner Familie, meiner Mutter Gerlinde, meinem Bruder<br />

Jens, meiner Oma Erika, Werner, meinem Opa Hans und meiner Uroma Emsel. Den beiden letzten<br />

ist es leider nicht vergönnt dem Tag meiner Promotion auf Erden beizuwohnen. Ihr habt mich zu<br />

jeder Zeit meines Lebens unterstützt, gefördert und mir alle Türen und Tore geöffnet. Ohne Euch<br />

wäre all das hier nicht möglich gewesen. Keine Worte dieser Welt können ausdrücken wie sehr ich<br />

Euch für Eure Liebe und all Eure Unterstützung in meinem Leben dankbar bin.


PARTS OF THIS THESIS HAVE ALREADY BEEN PRESENTED OR PUBLISHED<br />

I. Heiko A. Schiffter, Ge<strong>of</strong>frey Lee (2004). „Protein spray-<strong>drying</strong>: single <strong>droplet</strong> <strong>drying</strong> kinetics<br />

via acoustical levitation.” 5 th World Meeting on Pharmaceutics, Biopharmaceutics <strong>and</strong><br />

Pharmaceutical Technology. Nuremberg (Germany).<br />

II. Heiko A. Schiffter, Ge<strong>of</strong>frey Lee (2005). „Protein spray-<strong>drying</strong>: influence <strong>of</strong> trehalose on the<br />

single <strong>droplet</strong> <strong>drying</strong> kinetics <strong>of</strong> solutions <strong>of</strong> bovine serum albumin in an acoustic levitator.”<br />

1 st European Congress <strong>of</strong> Life Science Technology. Nuremberg (Germany).<br />

III. Heiko A. Schiffter, Ge<strong>of</strong>frey Lee (2005). „Determination <strong>of</strong> single <strong>droplet</strong> <strong>drying</strong> kinetics <strong>of</strong><br />

<strong>protein</strong> solutions via acoustic levitation.” AAPS Annual Meeting. Nashville (USA)<br />

IV. Jacob Sloth, Soren Z. Kiil, Anker D. Jensen, Sune K. Anderson, Heiko A. Schiffter,<br />

Ge<strong>of</strong>frey Lee (2006). „Model based analysis <strong>of</strong> the <strong>drying</strong> <strong>of</strong> a single <strong>droplet</strong> in an ultrasonic<br />

levitator.” Chemical Engineering Science. Accepted <strong>and</strong> in press.


CONTENTS I<br />

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

1. General introduction ..........................................................................................1<br />

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

2.1 General heat transfer considerations......................................................................................... 9<br />

2.1.1 Heat transfer by conduction ...................................................................................................... 9<br />

Heat transfer through dry porous solid material ..................................................................... 10<br />

Heat transfer through humid porous solid material ................................................................ 11<br />

2.1.2 Convective heat transfer.......................................................................................................... 12<br />

2.1.3 Heat transfer by radiation........................................................................................................ 14<br />

2.2 Diffusion <strong>and</strong> mass transfer .................................................................................................... 16<br />

2.2.1 Diffusion ................................................................................................................................. 16<br />

Stefan-Maxwell Equation ....................................................................................................... 16<br />

Fick’s law <strong>of</strong> molecular diffusion ........................................................................................... 18<br />

Diffusion coefficient for gases................................................................................................ 18<br />

2.2.2 Convective mass transfer considerations ................................................................................ 19<br />

Equimolar diffusion in gases................................................................................................... 19<br />

General case for diffusion <strong>of</strong> gases plus convection............................................................... 20<br />

Unimolar diffusion.................................................................................................................. 21<br />

2.2.3 Film theory.............................................................................................................................. 22<br />

2.2.4 Boundary layer theory............................................................................................................. 24<br />

2.2.5 Penetration theory ................................................................................................................... 25<br />

2.2.6 Surface renewal theory............................................................................................................ 26<br />

2.3 Heat <strong>and</strong> mass transfer considerations for pure solvents ........................................................ 26<br />

2.3.1 The D 2 -law .............................................................................................................................. 26<br />

2.3.2 Abramzon <strong>and</strong> Sirignano’s model........................................................................................... 31<br />

2.3.3 Diffusion controlled evaporation <strong>of</strong> a single <strong>droplet</strong> .............................................................. 33<br />

2.3.4 Evaporation <strong>of</strong> single <strong>droplet</strong>s containing solvent mixtures................................................... 34<br />

2.3.4 Droplet-gas interactions .......................................................................................................... 35


II CONTENTS<br />

2.4 Evaporation <strong>of</strong> liquid <strong>droplet</strong>s containing dissolved solids .................................................... 36<br />

2.4.1 The <strong>drying</strong> stages <strong>of</strong> <strong>droplet</strong>s containing dissolved or suspended solids................................ 36<br />

The first <strong>drying</strong> stage .............................................................................................................. 37<br />

The critical moisture content................................................................................................... 38<br />

The second <strong>drying</strong> stage.......................................................................................................... 38<br />

2.4.2 The movement <strong>of</strong> moisture in solids....................................................................................... 41<br />

2.4.3 Modelling <strong>of</strong> <strong>drying</strong> <strong>of</strong> single <strong>droplet</strong>s containing dissolved or suspended solids ................. 42<br />

Droplets with dissolved solids ................................................................................................ 42<br />

Droplets with suspended inert solids ...................................................................................... 43<br />

2.5 Particle formation, size <strong>and</strong> morphology ................................................................................ 44<br />

2.5.1 Spray-dried particles ............................................................................................................... 44<br />

2.5.2 <strong>Single</strong> <strong>droplet</strong> <strong>drying</strong> experiments .......................................................................................... 46<br />

3. Acoustic levitation.............................................................................................49<br />

3.1 Basic principles <strong>of</strong> levitation................................................................................................... 49<br />

3.1.1 Application <strong>of</strong> acoustic levitation ........................................................................................... 49<br />

3.1.2 Fundamentals <strong>of</strong> acoustics ...................................................................................................... 51<br />

3.1.3 Forces <strong>of</strong> a st<strong>and</strong>ing acoustic wave......................................................................................... 53<br />

3.2. Influences on <strong>droplet</strong>s inside an acoustic levitator ................................................................. 61<br />

3.2.1 Interactions with the acoustic field ......................................................................................... 61<br />

3.2.2 Acoustic streaming.................................................................................................................. 61<br />

Influence <strong>of</strong> the inner acoustic streaming on mass transfer.................................................... 63<br />

Influence <strong>of</strong> the outer acoustic streaming on mass transfer.................................................... 65<br />

3.2.3 Influence <strong>of</strong> <strong>droplet</strong> volume.................................................................................................... 67<br />

3.2.4 Vertical position <strong>of</strong> levitated <strong>droplet</strong> ...................................................................................... 68<br />

3.2.5 Influences <strong>of</strong> the ultrasonic transducer ................................................................................... 69<br />

3.3 <strong>Single</strong> <strong>droplet</strong> <strong>drying</strong> in an acoustic levitator ......................................................................... 70<br />

3.3.1 Pure solvent <strong>droplet</strong>s............................................................................................................... 70<br />

3.3.2 Droplets <strong>of</strong> binary liquid mixtures.......................................................................................... 71<br />

3.3.3 Solution <strong>and</strong> suspension <strong>droplet</strong>s............................................................................................ 72


CONTENTS III<br />

4. Materials <strong>and</strong> Methods ....................................................................................75<br />

4.1 Materials.................................................................................................................................. 75<br />

4.1.1 Proteins ................................................................................................................................... 75<br />

Bovine serum albumin ............................................................................................................ 75<br />

Catalase from bovine liver ...................................................................................................... 76<br />

4.1.2 Excipients <strong>and</strong> reagents........................................................................................................... 78<br />

4.1.3 Acoustic levitation system ...................................................................................................... 79<br />

Acoustic levitator .................................................................................................................... 79<br />

Controlled evaporation mixer ................................................................................................. 82<br />

CCD-camera <strong>and</strong> imaging s<strong>of</strong>tware........................................................................................ 84<br />

4.2 Methods................................................................................................................................... 86<br />

4.2.1 Acoustic levitation .................................................................................................................. 86<br />

Camera system <strong>and</strong> imaging s<strong>of</strong>tware .................................................................................... 86<br />

Acoustic levitator .................................................................................................................... 86<br />

Controlled evaporation mixer ................................................................................................. 86<br />

Levitation procedure ............................................................................................................... 87<br />

4.2.2 Maximum bubble pressure tensiometry.................................................................................. 87<br />

4.2.3 Ring tensiometry..................................................................................................................... 89<br />

4.2.4 Spray-<strong>drying</strong>............................................................................................................................ 90<br />

4.2.5 Karl-Fischer titration............................................................................................................... 92<br />

4.2.6 Enzyme activity assay <strong>of</strong> catalase........................................................................................... 92<br />

4.2.7 Kinetic viscosity...................................................................................................................... 93<br />

4.2.8 Liquid density ......................................................................................................................... 94<br />

4.2.9 Scanning electron microscopy ................................................................................................ 94<br />

5. Results <strong>and</strong> discussions ....................................................................................95<br />

5.1 Pre-liminary levitation tests .................................................................................................... 95<br />

5.1.1 Size measurement ................................................................................................................... 95<br />

5.1.2 Injectable volume.................................................................................................................... 97<br />

5.1.3 Heating <strong>of</strong> process chamber.................................................................................................... 99<br />

5.1.4 Parametric studies ................................................................................................................. 101


IV CONTENTS<br />

5.2 Evaporation <strong>of</strong> pure solvent <strong>droplet</strong>s .................................................................................... 103<br />

5.2.1 Data analysis ......................................................................................................................... 103<br />

Different initial <strong>droplet</strong> size .................................................................................................. 103<br />

Different <strong>droplet</strong> liquids........................................................................................................ 106<br />

Multicomponent <strong>droplet</strong>s ...................................................................................................... 107<br />

5.2.2 Evaporation <strong>of</strong> pure water <strong>droplet</strong>s....................................................................................... 109<br />

Calculation <strong>of</strong> limits <strong>of</strong> levitation range for water <strong>droplet</strong>s.................................................. 110<br />

Levitation <strong>of</strong> water <strong>droplet</strong>s .................................................................................................. 111<br />

5.2.3 Evaporation <strong>of</strong> pure ethanol <strong>droplet</strong>s.................................................................................... 124<br />

Calculation <strong>of</strong> limits <strong>of</strong> levitation range for ethanol <strong>droplet</strong>s ............................................... 124<br />

Levitation <strong>of</strong> pure ethanol <strong>droplet</strong>s ....................................................................................... 124<br />

5.2.4 Evaporation <strong>of</strong> ethanol-water mixtures................................................................................. 136<br />

5.3 Evaporation <strong>of</strong> solution <strong>droplet</strong>s........................................................................................... 138<br />

5.3.1 Evaporation <strong>of</strong> maltodextrin solution <strong>droplet</strong>s...................................................................... 138<br />

Properties <strong>of</strong> maltodextrin..................................................................................................... 138<br />

Data analysis <strong>of</strong> maltodextrin experiments........................................................................... 138<br />

5.3.2 Evaporation <strong>of</strong> trehalose solution <strong>droplet</strong>s............................................................................ 147<br />

Properties <strong>of</strong> trehalose........................................................................................................... 147<br />

Data analysis <strong>of</strong> trehalose experiments................................................................................. 149<br />

5.3.3 Evaporation <strong>of</strong> mannitol solution <strong>droplet</strong>s............................................................................ 159<br />

Properties <strong>of</strong> mannitol........................................................................................................... 159<br />

Data analysis <strong>of</strong> trehalose experiments................................................................................. 160<br />

5.3.4 Evaporation <strong>of</strong> bSA solution <strong>droplet</strong>s................................................................................... 170<br />

Properties <strong>of</strong> bSA.................................................................................................................. 170<br />

Data analysis <strong>of</strong> bSA solution experiments .......................................................................... 171<br />

Interfacial behaviour <strong>and</strong> surface excess <strong>of</strong> bSA.................................................................. 186<br />

Interfacial behaviour <strong>and</strong> surface excess <strong>of</strong> Pluronic F127 .................................................. 193<br />

Influence <strong>of</strong> trehalose on the interfacial behaviour <strong>of</strong> bSA .................................................. 195<br />

Determination <strong>of</strong> bSA monolayer......................................................................................... 199<br />

5.3.4 Evaporation <strong>of</strong> catalase solution <strong>droplet</strong>s ............................................................................. 202<br />

Properties <strong>of</strong> catalase ............................................................................................................ 202<br />

Data analysis <strong>of</strong> catalase solution experiments..................................................................... 204<br />

Interfacial behaviour <strong>and</strong> surface excess <strong>of</strong> catalase ............................................................ 213


CONTENTS V<br />

Influence <strong>of</strong> trehalose on the interfacial behaviour <strong>of</strong> catalase............................................. 216<br />

Comparison <strong>of</strong> surface tension <strong>and</strong> surface excess <strong>of</strong> bSA <strong>and</strong> catalase.............................. 218<br />

Enzyme activity <strong>of</strong> levitated catalase solution <strong>droplet</strong>s........................................................ 219<br />

6. Conclusions......................................................................................................222<br />

7. Zusammenfassung ..........................................................................................230<br />

8. Annex ...............................................................................................................239<br />

9. References........................................................................................................241


VI ABBREVIATIONS<br />

Capital letters<br />

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

A 0<br />

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

A0 e<br />

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

A0 em<br />

effective amplitude <strong>of</strong> the acoustic field at drop-out point<br />

A surface area<br />

B gas particle velocity<br />

C D<br />

drag coefficient<br />

D diffusion coefficient<br />

D AB<br />

binary diffusion coefficient<br />

E BM<br />

bulk modulus <strong>of</strong> elasticity<br />

F correction factor<br />

F L<br />

axial acoustic levitation force<br />

F r<br />

radial acoustic levitation force<br />

HC kinetic energy correction in viscometry<br />

I acoustic intensity<br />

J molar flux through a reference plane<br />

K acoustic<br />

constant <strong>of</strong> st<strong>and</strong>ing acoustic wave<br />

K SPL<br />

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

K viscometer<br />

constant <strong>of</strong> viscometer<br />

L dw, Lwd<br />

specific coefficients for a binary liquid mixture<br />

L I<br />

sound intensity level<br />

L p<br />

sound pressure level<br />

L R<br />

distance transducer <strong>and</strong> reflector<br />

L ν<br />

sound velocity level<br />

L W<br />

sound power level<br />

M molecular weigh<br />

N i<br />

flux <strong>of</strong> a component i<br />

Q heat flow/quantity


T temperature<br />

T M<br />

midpoint <strong>of</strong> thermal denaturation<br />

T S<br />

surface temperature<br />

T WB<br />

wet bulb temperature<br />

T ∞<br />

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

R gas constant<br />

U 0<br />

driving voltage <strong>of</strong> ultrasonic transducer<br />

ABBREVIATIONS VII<br />

U 0 m<br />

driving voltage <strong>of</strong> ultrasonic transducer at drop-out point<br />

V volume<br />

W acoustic power<br />

X moisture content<br />

Y i<br />

mass fraction <strong>of</strong> component i<br />

Z i<br />

mole fraction <strong>of</strong> component i<br />

Small letters<br />

a fraction <strong>of</strong> vertically arranged plates<br />

c concentration<br />

c m<br />

molar heat capacity<br />

c P<br />

heat capacity at constant pressure<br />

c s<br />

specific heat capacity, specific heat<br />

d diameter<br />

df dilution factor<br />

f frequency<br />

g gravitational acceleration<br />

h heat transfer coefficient<br />

h ci<br />

averaged heat transfer coefficient <strong>of</strong> the whole <strong>droplet</strong> surface <strong>of</strong> the liquid<br />

component i<br />

h v<br />

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

k mass transfer coefficient<br />

k′ modified mass transfer coefficient


VIII ABBREVIATIONS<br />

k 0<br />

wave number<br />

k r<br />

radial wave number<br />

k z<br />

axial wave number<br />

l length<br />

m mass<br />

m& vap<br />

mass flux <strong>of</strong> vapour per time<br />

n natural number<br />

p pressure<br />

p BM<br />

logarithmic mean value <strong>of</strong> partial pressures<br />

p′ i<br />

pressure perturbation <strong>of</strong> the incident acoustic wave<br />

p S<br />

saturation vapour pressure<br />

p′ S<br />

pressure perturbation <strong>of</strong> the scattered acoustic wave<br />

p ∞<br />

vapour pressure in <strong>drying</strong> air<br />

q& heat flux = transferred heat per time <strong>and</strong> area<br />

r radial coordinate<br />

r S<br />

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

s fraction rate <strong>of</strong> surface renewal<br />

t time<br />

u velocity<br />

u acoustic<br />

averaged velocity <strong>of</strong> the inner acoustic streaming<br />

w weight<br />

x fraction<br />

x D<br />

distance form centre <strong>of</strong> <strong>droplet</strong> mass to adjacent pressure node<br />

y mass fraction<br />

z length, horizontal coordinate<br />

Big Greek letters<br />

Γ surface excess concentration<br />

Π surface pressure<br />

Ω D<br />

collision integral


Small Greek letters<br />

α adiabatic index<br />

β coefficient <strong>of</strong> expansion<br />

ABBREVIATIONS IX<br />

β d<br />

evaporation coefficient calculated from the <strong>droplet</strong> diameter<br />

β r<br />

evaporation coefficient calculated from the <strong>droplet</strong> radius<br />

γ interfacial / surface tension<br />

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

δ IAS<br />

thickness <strong>of</strong> acoustic boundary layer<br />

δ diff<br />

thickness <strong>of</strong> diffusion boundary layer<br />

ε porosity<br />

θ perimeter angle<br />

ϕ relative humidity<br />

λ thermal conductivity (chapter 2)<br />

λ 0<br />

wavelength<br />

μ dynamic viscosity<br />

ν velocity<br />

ν 0<br />

velocity <strong>of</strong> sound<br />

ϑ temperature in °C<br />

κ gas<br />

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

ρ density<br />

σ surface tension<br />

σ Stefan-Boltzmann constant = 5,<br />

67 ⋅10<br />

SB<br />

σ AB<br />

effective collision diameter<br />

υ relaxation time in tensiometry<br />

ψ activity coefficient for binary mixtures<br />

ω emissivity (chapter 2)<br />

angular frequency (chapter 3)<br />

−8<br />

W⋅<br />

m<br />

ω SW<br />

angular frequency <strong>of</strong> st<strong>and</strong>ing acoustic wave<br />

ς friction coefficient between two components<br />

1,<br />

2<br />

ξ dimensionless radial coordinate<br />

−2<br />

⋅ K<br />

−4


X ABBREVIATIONS<br />

Mathematical sign<br />

∈ Lennard-Jones force<br />

Subscripts<br />

A , B components<br />

Expressions<br />

bSA bovine serum albumin<br />

CAT catalase from bovine liver<br />

CCD charged coupled device<br />

CD circular dichroism<br />

CEM controlled evaporation mixer<br />

ESCA electron spectroscopy for chemical analysis<br />

FTIR Fourier transformed infrared spectroscopy<br />

HFA hydr<strong>of</strong>luoroalkane<br />

LDH lactate dehydrogenase<br />

pMDI pressurized metered dose inhaler<br />

SD spray-<strong>drying</strong><br />

SDD single <strong>droplet</strong> <strong>drying</strong><br />

SEM scanning electron microscopy<br />

SIL sound intensity level<br />

SPL sound pressure level<br />

SPL effective<br />

effective sound pressure level<br />

SVL sound velocity level<br />

SWL sound power level


CHAPTER 1 GENERAL INTRODUCTION 1<br />

1 General Introduction<br />

Many manufacturing processes require one or more stages in which the unit operation “Drying” is<br />

carried out. Due to its influence in specifying the quality <strong>of</strong> a product, <strong>drying</strong> is one <strong>of</strong> the most<br />

important unit operations <strong>and</strong> has probably the widest application in industry. The <strong>drying</strong> is a<br />

cutting-<strong>of</strong>f process where liquid is removed from a solid in form <strong>of</strong> a solution, suspension, slurry,<br />

paste or a solid matrix. This usually involves the evaporation <strong>of</strong> the liquid <strong>and</strong> its subsequent<br />

diffusion away from the surface <strong>of</strong> the product. The evaporation can either result as a transfer <strong>of</strong><br />

heat to the liquid at temperatures ≥ than the boiling point <strong>of</strong> the liquid, or at room temperature<br />

which takes a much longer time for the <strong>drying</strong> process. In conductive <strong>drying</strong> the wet solid is, for<br />

example, put into a vessel <strong>and</strong> heated from the outside by placing onto a heated surface. For a heat<br />

supply by radiation the energy <strong>of</strong> infra-red irradiation is used. Convective <strong>drying</strong> uses the transfer <strong>of</strong><br />

heat from a bulk gas to the liquid. A large advantage <strong>of</strong> the latter is that the bulk gas can also be<br />

used for removal <strong>of</strong> the evaporated liquid. Using conductive or radiative heating a vent is installed<br />

additionally to the heat source to remove the vapour. Commercially a large number <strong>of</strong> dryer types is<br />

available using either conductive, radiative or convective heating. Table 1.1 gives an overview <strong>of</strong><br />

the different principles, characteristics <strong>and</strong> types <strong>of</strong> dryers. For the rate <strong>of</strong> evaporation not only the<br />

temperature <strong>and</strong> the way the heat is transferred to the liquid are important, but also the mechanism<br />

by which the liquid is bound in the solid matrix.<br />

Table 1.1: Overview <strong>of</strong> different methods <strong>of</strong> heat transfer, the <strong>drying</strong> principles <strong>and</strong> a sample <strong>of</strong> the corresponding<br />

commercial equipment.<br />

Nature <strong>of</strong> heat transfer Principle <strong>of</strong> heat transfer Apparatus (e.g.)<br />

Conductive<br />

Contact <strong>of</strong> the heating source <strong>and</strong> the<br />

product. Directly or via other<br />

materials<br />

Radiative Infra-red or microwave Irradiation<br />

Convective<br />

Heated bulk gas effacing the surface<br />

<strong>of</strong> the product<br />

- Plate dryer<br />

- Mixer dryer<br />

- Nutsche filter dryer<br />

- Infra-red tunnel dryer<br />

- Microwave vacuum dryer<br />

- Spray dryer<br />

- Fluidized bed dryer<br />

Spray-<strong>drying</strong>, as an example <strong>of</strong> <strong>drying</strong> using convective <strong>drying</strong> principles is the most widely<br />

used industrial process involving particle formation <strong>and</strong> <strong>drying</strong> [Niro 2005]. It is embedded in the<br />

techniques <strong>of</strong> suspended particle processing (SPP) that use liquid atomization to create <strong>droplet</strong>s<br />

which are dried to individual particles in the gaseous <strong>drying</strong> medium. Other SPP systems for


2 GENERAL INTRODUCTION<br />

example are spray-agglomeration, spray-granulation or spray-cooling. Spray-<strong>drying</strong> is a technique<br />

involving liquid spraying, <strong>droplet</strong> <strong>drying</strong>, particle formation, powder collection <strong>and</strong> h<strong>and</strong>ling<br />

[Masters 2002]. Figure 1.1 shows the flow diagram <strong>of</strong> a modern spray-<strong>drying</strong> plant. The first step is<br />

atomization by centrifugal, pressure or sonic nozzles, where the liquid feed is broken up into many<br />

<strong>of</strong> small individual <strong>droplet</strong>s forming a spray. This leads to an enormous increase in the specific<br />

surface area <strong>of</strong> the liquid resulting in very fast rates <strong>of</strong> heat <strong>and</strong> mass transfer in the subsequent<br />

<strong>drying</strong> process. During the second stage, hot air contact, moisture evaporates <strong>and</strong> particles are<br />

formed. In the third <strong>and</strong> last stage, an efficient particle collection is necessary, for example by a gassolid<br />

cyclone. As a continuous production method spray-<strong>drying</strong> is ideal to produce powders,<br />

granulates or agglomerate forms from a liquid feed. With relatively low device-related effort <strong>and</strong><br />

optimized energy consumption, high production rates <strong>of</strong> an end-product with specified requirements<br />

like particle size distribution, bulk <strong>and</strong> particle density, residual moisture content, dispersibility <strong>and</strong><br />

even flavour <strong>and</strong> aroma retention <strong>of</strong> volatile substances can be achieved.<br />

Figure 1.1: Basis flow diagram <strong>of</strong> a spray-dryer [Masters 2002].<br />

The popularity <strong>of</strong> spray-<strong>drying</strong> is shown in various application areas in different branches <strong>of</strong><br />

industry listed in Table 1.2 [Niro 2005]. The ability to control the above mentioned product<br />

characteristics by the <strong>drying</strong> procedure <strong>and</strong> to keep the <strong>drying</strong> temperature as low as possible, is the<br />

major advantage <strong>of</strong> the spray-<strong>drying</strong> technique over other <strong>drying</strong> methods. This is very important if<br />

the formulation to be dried contains a heat sensitive ingredient. In the food <strong>and</strong> dairy industry, for<br />

example, a large loss <strong>of</strong> a volatile flavour or aroma during the <strong>drying</strong> process would reduce the<br />

quality <strong>of</strong> the end-product.


CHAPTER 1 GENERAL INTRODUCTION 3<br />

Table 1.2: Industrial applications <strong>of</strong> spray-<strong>drying</strong> according to the application brochure published by GEA Niro A/S,<br />

Denmark in 2005<br />

Branch <strong>of</strong> Industry Industrial Application<br />

Pharmaceutical Industry<br />

Food <strong>and</strong> Dairy Industry<br />

Chemical Industry<br />

Polymer Industry<br />

Ceramic Industry<br />

- Analgesics<br />

- Antibiotics<br />

- Enzymes<br />

- Plasma <strong>and</strong> plasma substitutes<br />

- Baby food<br />

- Cheese <strong>and</strong> whey products<br />

- Coconut milk<br />

- C<strong>of</strong>fee <strong>and</strong> c<strong>of</strong>fee substitutes<br />

- C<strong>of</strong>fee whitener<br />

- Eggs<br />

- Flavours<br />

- Maltodextrines<br />

- Catalysts<br />

- Detergents<br />

- Dyestuffs<br />

- Fine (in)organic chemicals<br />

- Tannins<br />

- ABS<br />

- E-PVC<br />

- PMMA<br />

- UF/MF resins<br />

- Advanced ceramic <strong>formulations</strong><br />

- Carbides<br />

- Ferrites<br />

- Nitrides<br />

- Vaccines<br />

- Vitamins<br />

- Yeasts<br />

- Milk<br />

- Soup mixes<br />

- Soy-based food<br />

- Spices/herb extracts<br />

- Sugar-based food<br />

- Tea<br />

- Tomato<br />

- Vegetable <strong>protein</strong><br />

- Chelates<br />

- Fungicides<br />

- Herbicides<br />

- Insecticides<br />

- Oxides<br />

- Silicates<br />

- Steatites<br />

- Titanates<br />

In the pharmaceutical industry the application <strong>of</strong> <strong><strong>protein</strong>s</strong> <strong>and</strong> peptides as active ingredients<br />

has been rapidly increasing since the middle <strong>of</strong> the 1990s facilitated by improvements in modern<br />

biotechnology. Many are already available for therapy [Banga 1995]. The worldwide sale <strong>of</strong><br />

recombinant <strong>protein</strong> <strong>and</strong> peptide substances was 41.3 billion US$ in 2002 according to Frost &<br />

Sullivan, Banc Boston [ISB 2005]. The five top-selling products on the market are listed in<br />

Table 1.3. These peptides <strong>and</strong> <strong><strong>protein</strong>s</strong> are delivered by injection due to their poor bioavailability.<br />

Unfortunately, parenteral production is time- <strong>and</strong> cost-intensive <strong>and</strong> the injection itself needs skilled<br />

personal <strong>and</strong> is viewed sceptically by patients [Patton 1997]. Therefore, alternative ways <strong>of</strong> <strong>protein</strong><br />

delivery have been increasingly discussed <strong>and</strong> examined within the last decade. Possible <strong>and</strong><br />

favoured routes <strong>of</strong> administration can be pulmonary, transdermal <strong>and</strong> nasal delivery systems<br />

[Banga 1995].


4 GENERAL INTRODUCTION<br />

Table 1.3: Sales <strong>of</strong> the five top-selling recombinant drugs in 2004 according to the publication by the Information<br />

Secretary <strong>of</strong> Biotechnology 2005 [ISB 2005].<br />

Product Protein Effect / Therapeutic Use Marketed by<br />

Procrit /<br />

Eprex<br />

Rituxan<br />

(Mabthera)<br />

Remicade<br />

Enbrel<br />

Erythropoietin alpha<br />

Rituximab<br />

(Humanised MAb)<br />

Infliximab<br />

(Chimaeric MAb)<br />

Etanercept<br />

(Fusion <strong>protein</strong> <strong>of</strong><br />

antibody-Fc <strong>and</strong> p75-<br />

TNF receptor <strong>protein</strong>)<br />

Epogen Erythropoietin<br />

Stimulation <strong>of</strong> the production <strong>of</strong><br />

erythrocytes<br />

Worldwide sales 2004<br />

[US$m]<br />

J&J, Ortho Biotech 3589<br />

Leukaemia <strong>and</strong> Lymphoma Genentech, Roche 2989<br />

Rheumatoid arthritis,<br />

Morbus Crohn<br />

J&J 2891<br />

Rheumatoid arthritis Amgen 2580<br />

Stimulation <strong>of</strong> the production <strong>of</strong><br />

erythrocytes<br />

Amgen 2601<br />

Karil © is an example for nasal application containing salmon-calcitonin used in therapy to limit the<br />

risk <strong>of</strong> a vertebral fracture <strong>of</strong> women with post menopausal osteoporosis [Novartis 2005]. The<br />

bioavailability <strong>of</strong> the calcitonin compared with a parenteral injection is between 2 to 15% according<br />

to specific information supplied by Novartis Pharma. Possible reasons for the poor absorption <strong>of</strong> the<br />

<strong>protein</strong> via the nasal mucous membrane are inactivation by enzymes at the application area <strong>and</strong> an<br />

epithelial layer that acts as a permeation barrier [Banga 1995]. Considering these facts, the lungs<br />

where some 300 million alveoli constitute a capillarized area <strong>of</strong> approximately 100 m 2 are a<br />

promising alternative [Almer et al. 2002]. Already at the beginning <strong>of</strong> the 1990s the pulmonary<br />

absorption <strong>of</strong> leuprolide acetate, a nonapeptide with potent luteinising hormone releasing activity,<br />

was tested. Lung administration <strong>of</strong> a dose <strong>of</strong> 1.0mg to beagle dogs showed an absolute<br />

bioavailability <strong>of</strong> 40% [Adjei 1994]. Granulocyte colony stimulating factor (G-CSF) after<br />

intratracheal instillation formulated as an aerosol was also investigated. The bioavailability<br />

compared to parenteral application was 45.9% <strong>of</strong> the administered dose <strong>and</strong> 62.0% <strong>of</strong> the dose<br />

reaching the lung lobes [Niven et al. 1993]. Further <strong><strong>protein</strong>s</strong> <strong>of</strong> interest, not only for systemic<br />

delivery but also for the lungs as a local target are alpha-1 antitrypsin (therapy <strong>of</strong> emphysema),<br />

interleukin-1 receptor (therapy <strong>of</strong> asthma), cyclosporine (therapy <strong>of</strong> immunsuppression after<br />

transplantation), anti-cytomegalovirus antibody (therapy <strong>of</strong> cytomegalovirus) or the already<br />

approved <strong>and</strong> marketed recombinant human desoxyribonuclease (rhDNase) with the product name<br />

Pulmozyme © [Roche 2005].


CHAPTER 1 GENERAL INTRODUCTION 5<br />

A major challenge in the development <strong>of</strong> stable peptide <strong>and</strong> <strong>protein</strong> <strong>formulations</strong> <strong>and</strong> dosage<br />

forms is to ensure their shelf life stability. Instability <strong>of</strong> peptides <strong>and</strong> <strong><strong>protein</strong>s</strong> can broadly be<br />

classified into two sections: physical <strong>and</strong> chemical instability. Physical instability refers to a change<br />

in secondary, tertiary or quaternary structure <strong>of</strong> the <strong>protein</strong> <strong>and</strong> includes denaturation, aggregation,<br />

precipitation <strong>and</strong> adsorption to surfaces [Banga 1995]. Due to the primary structure <strong>of</strong> the <strong><strong>protein</strong>s</strong><br />

consisting <strong>of</strong> various amino acids, there are multiple reactive sites for chemical reactions. Chemical<br />

instability involves covalent modification <strong>of</strong> the <strong>protein</strong> via bond formation or cleavage, including<br />

deamination, hydrolysis, oxidation, disulfide exchange, racemisation, β-elimination <strong>and</strong> Maillard-<br />

reaction [Banga 1995; Goolcharran et al. 2000].<br />

The challenge <strong>of</strong> deep lung delivery <strong>of</strong> costly <strong><strong>protein</strong>s</strong> <strong>and</strong> peptides has led to the development<br />

<strong>of</strong> dry powder <strong>formulations</strong> <strong>of</strong> <strong><strong>protein</strong>s</strong> <strong>and</strong> peptides to overcome the limitations <strong>of</strong> conventional<br />

aerosol delivery systems. These dry powder <strong>formulations</strong> have several advantages, including<br />

product <strong>and</strong> formulation stability, low susceptibility to microbiological growth <strong>and</strong> applicability to<br />

both soluble <strong>and</strong> insoluble drugs [Patton 1997]. The inhalable insulin Exubera © is marketed by<br />

Pfizer <strong>and</strong> developed in cooperation with Nektar [Nektar 2005]. To achieve an optimal deep lung<br />

delivery it is important to use the correct particle size <strong>of</strong> 1 to 5µm in diameter for best deposition<br />

efficiency <strong>and</strong> to ensure the stability <strong>of</strong> the peptide <strong>and</strong> <strong><strong>protein</strong>s</strong> in the production process <strong>of</strong> the<br />

microparticles as well as during their shelf life. To meet these requirements, spray-<strong>drying</strong> is an<br />

effective <strong>and</strong> efficient method to produce peptide or <strong>protein</strong> loaded powders suitable for pulmonary<br />

delivery. If the correct formulation <strong>and</strong> spray-<strong>drying</strong> conditions can be identified, a product with a<br />

high yield <strong>and</strong> the above mentioned required characteristics can be achieved [Maa et al. 1997] [Lee<br />

2002]. Since the <strong>drying</strong> time <strong>of</strong> the atomized liquid <strong>droplet</strong>s is < 1 s depending on their initial<br />

diameter [Stahl 1980] [Nürnberg 1980], spray-<strong>drying</strong> is considered as a gentle process for<br />

microparticle production out <strong>of</strong> peptide or <strong>protein</strong> solutions. Unfortunately, the spray-<strong>drying</strong> <strong>of</strong><br />

aqueous solutions <strong>of</strong> pure <strong><strong>protein</strong>s</strong> without any excipients can lead to unfolding, aggregation or<br />

inactivation. Therefore the <strong>protein</strong> must be stabilized by formulation with one or more substances,<br />

for example glassy carriers such as sucrose to improve the process as well as the storage stability.<br />

Proteins such as recombinant human granulocyte-colony stimulating factor (rhG-CSF) [Niven et al.<br />

1994], recombinant human growth hormone rhGH [Maa et al. 1998 a], trypsinogen [Tzannis 1999],<br />

lactic dehydrogenase (LDH) [Adler 1999], recombinant humanized anti-IgE monoclonal antibody<br />

[Andya et al. 1999] <strong>and</strong> human immunoglobulin G (IgG) [Maury et al. 2005 a] have been<br />

successfully stabilized in this way. Unfortunately, these kinds <strong>of</strong> <strong><strong>protein</strong>s</strong> are usually very expensive<br />

<strong>and</strong> available only in small quantities for formulation studies. 500mg lyophilized powder <strong>of</strong> a


6 GENERAL INTRODUCTION<br />

human IgG with a <strong>protein</strong> content <strong>of</strong> approximately 80% in reagent grade cost almost 1300 €<br />

[Sigma 2004]. The currently favoured equipment in both industrial <strong>and</strong> university research is<br />

therefore a small laboratory-scale spray dryer such as the Büchi Mini Spray-Dryer B-290 TM<br />

[Büchi 2005] or the Niro Lab Spray Dryer SD Micro TM [Niro 2005]. Newer developments in<br />

cyclone technology for powder collection can maximize the yield during the <strong>drying</strong> process <strong>and</strong><br />

reduce the utilized peptide or <strong>protein</strong> amount [Maury et al. 2005 b].<br />

Research on the formulation <strong>of</strong> peptide or <strong>protein</strong> solutions for spray <strong>drying</strong> is not only<br />

concerned with the stabilisation <strong>and</strong> yield optimisation, but also with prediction <strong>of</strong> particle<br />

morphology <strong>and</strong> <strong>drying</strong> kinetics. Though some investigations <strong>of</strong> variables has already been done<br />

[Maa et al. 1997] [Adler et al. 2000] [Elverson et al. 2003] it is still a challenge to examine <strong>drying</strong><br />

kinetics <strong>and</strong> predict particle morphology from <strong>formulations</strong>.<br />

The application <strong>and</strong> properties <strong>of</strong> a dry powder containing a pharmaceutical peptide or <strong>protein</strong><br />

are determined by market forces. A fast <strong>and</strong> cost-efficient adaptation <strong>of</strong> the spray-<strong>drying</strong> process is<br />

therefore one <strong>of</strong> the main components to success. This leads to search for a technique to investigate<br />

<strong>and</strong> to optimize spray-<strong>drying</strong> processes taking the targeted product properties into account. Because<br />

<strong>of</strong> difficulties associated with investigating spray <strong>drying</strong> in situ, research has tended to be divided<br />

into four main areas:<br />

1. Atomizing studies (Correlation <strong>of</strong> <strong>droplet</strong> size distribution with feed characteristics, atomizer<br />

design, pressure or rotational speed, throughput or shearing stress for peptides <strong>and</strong> <strong><strong>protein</strong>s</strong>);<br />

2. Studies <strong>of</strong> gas flow patterns <strong>and</strong> residence time;<br />

3. <strong>Single</strong> <strong>droplet</strong> <strong>drying</strong> studies;<br />

4. Mathematical models, with data from point 1 to 3, to simulate dryer performance <strong>and</strong> to predict<br />

<strong>droplet</strong> <strong>drying</strong> rates <strong>and</strong> <strong>drying</strong> behaviour.<br />

The main interests have been studies <strong>of</strong> atomization <strong>and</strong> gas flow patterns. However, within the last<br />

decade the investigation <strong>of</strong> single <strong>droplet</strong> <strong>drying</strong> has increased rapidly. To underst<strong>and</strong> more about<br />

the behaviour <strong>of</strong> a <strong>protein</strong> during the spray-<strong>drying</strong> process, its stability or instability, its particle<br />

formation <strong>and</strong> particle morphology, precise knowledge <strong>of</strong> the heat <strong>and</strong> mass transfer processes<br />

during the <strong>drying</strong> period is necessary. The <strong>drying</strong> period is one <strong>of</strong> the most important sub-processes,<br />

because residual water content, <strong>protein</strong> activity <strong>and</strong> particle morphology are determined herein.<br />

Inside the spray-<strong>drying</strong> tower precise examinations are not possible. All these considerations show<br />

the need for a tool <strong>and</strong> a method for single <strong>droplet</strong> <strong>drying</strong> to investigate the <strong>drying</strong> behaviour <strong>and</strong><br />

kinetics <strong>of</strong> peptide <strong>and</strong> <strong><strong>protein</strong>s</strong> <strong>formulations</strong> during spray-<strong>drying</strong>.


CHAPTER 1 GENERAL INTRODUCTION 7<br />

In the literature different setups for single <strong>droplet</strong> <strong>drying</strong> experiments are described. They can<br />

be divided into techniques with contact [Walton 1994] <strong>and</strong> techniques without contact to the<br />

evaporating <strong>droplet</strong>. The former have the disadvantage <strong>of</strong> interferences in heat transfer, change in<br />

surface <strong>and</strong> volume <strong>of</strong> the <strong>droplet</strong> <strong>and</strong> influence on particle formation. In contrast, the contact-less<br />

techniques, also termed as levitation techniques or containerless-processing, cause less interference<br />

to the sample <strong>droplet</strong>. Table 1.4 shows an overview over some levitation techniques, their<br />

underlying physical principle <strong>and</strong> their characteristics [Lierke 1995]. Acoustic ultrasonic levitation<br />

has the large advantage that liquid <strong>droplet</strong>s <strong>of</strong> pure solvents, suspensions or solutions with a size<br />

range from 50µm to 4000µm can be levitated very easily [Kastner 2001]. In combination with<br />

embedded air flow or with aerodynamic principles in a hybrid levitator, experiments under different<br />

flow velocities are possible for comparison with spray-<strong>drying</strong> procedure.<br />

Table 1.4: Summary <strong>of</strong> the different levitation principles for containerless processing [Lierke 1995]<br />

Levitator Principle <strong>of</strong> levitation Characteristics<br />

Aerodynamic levitator<br />

Electrostatic levitator<br />

Acoustic levitator<br />

Acoustic-aerodynamic<br />

hybrid levitator<br />

Acoustic-electrostatic<br />

hybrid levitator<br />

Levitation in a upwards air-flow <strong>of</strong> a<br />

wind tunnel low in turbulence or in a<br />

upwards spreading air-flow<br />

Levitation <strong>of</strong> electrically charged<br />

<strong>droplet</strong>s in an electric field<br />

(“Millican”-experiment)<br />

Levitation in the pressure nodes <strong>of</strong> a<br />

st<strong>and</strong>ing ultrasonic wave between a<br />

transducer <strong>and</strong> a reflector<br />

Levitation as a combination <strong>of</strong><br />

acoustic <strong>and</strong> aerodynamic principles<br />

Levitation as a combination <strong>of</strong><br />

acoustic <strong>and</strong> electrostatic principles<br />

- Precise adjustments necessary due to<br />

unstable balance <strong>of</strong> forces determining flow<br />

velocity <strong>and</strong> <strong>droplet</strong> deformation<br />

- Similar to the aerodynamic levitator, precise<br />

adjustments for every <strong>droplet</strong> size necessary<br />

- Levitation without air flow<br />

- for every <strong>droplet</strong> diameter is a large range<br />

for stable levitation adjustment<br />

- Levitation in still air or with air flow<br />

possible<br />

- Utilization <strong>of</strong> stable levitation area <strong>and</strong> the<br />

possibilities <strong>of</strong> high flow velocity<br />

- Unique variation possibilities <strong>of</strong> engineering<br />

operation.<br />

The aim <strong>of</strong> this work in this thesis was to set up an acoustic levitation system to examine<br />

pharmaceutical spray-<strong>drying</strong> <strong>formulations</strong> <strong>of</strong> peptides <strong>and</strong> <strong><strong>protein</strong>s</strong>. The work can be divided into<br />

three different sections:<br />

1. Set up the acoustic levitation system for single <strong>droplet</strong> <strong>drying</strong> experiments under varying <strong>drying</strong><br />

conditions;<br />

2. Investigation <strong>of</strong> the evaporation behaviour <strong>of</strong> single <strong>droplet</strong>s with pure solvents or mixtures <strong>of</strong><br />

solvent, <strong>and</strong> verification <strong>of</strong> the results obtained with simple gas-phase models <strong>and</strong> data from the<br />

literature;


8 GENERAL INTRODUCTION<br />

3. Investigation <strong>of</strong> the evaporation behaviour <strong>of</strong> single solution <strong>droplet</strong>s containing sugars,<br />

peptides or <strong><strong>protein</strong>s</strong> <strong>and</strong> complex <strong>protein</strong> <strong>formulations</strong>. Comparison <strong>and</strong> verification <strong>of</strong> the<br />

results with data form spray-<strong>drying</strong> experiments <strong>and</strong> the literature.<br />

The first part <strong>of</strong> this work presents the theoretical background <strong>of</strong> <strong>drying</strong> <strong>and</strong> acoustic levitation.<br />

At the beginning the basic physical <strong>and</strong> thermodynamic principles <strong>of</strong> single <strong>droplet</strong> <strong>drying</strong> are<br />

discussed. Also a brief summary <strong>of</strong> published models for the evaporation <strong>of</strong> single <strong>droplet</strong>s is given.<br />

Then the historical background <strong>and</strong> the fundamentals <strong>of</strong> acoustic levitation as a containerlessprocessing<br />

technique are focused on. Existing techniques for the determination <strong>of</strong> single <strong>droplet</strong><br />

<strong>drying</strong> kinetics <strong>of</strong> suspension <strong>droplet</strong>s [Kastner et al. 2001] [Groenewold et al. 2002] inside an<br />

acoustic tube levitator are discussed.<br />

The second part <strong>of</strong> this work deals with the evaporation <strong>of</strong> single <strong>droplet</strong>s <strong>of</strong> pure solvents <strong>and</strong><br />

mixtures <strong>of</strong> solvents. The influences <strong>of</strong> temperature, humidity, air flow <strong>and</strong> sound pressure level<br />

(SPL) on the evaporation kinetics inside the levitation system are examined. The part concludes<br />

with the set-up <strong>of</strong> a simple model describing the evaporation process <strong>of</strong> solvent <strong>droplet</strong>s inside the<br />

acoustic levitator.<br />

In the third part the <strong>drying</strong> <strong>of</strong> single solution <strong>droplet</strong>s in the ultrasonic levitator is shown. Again,<br />

the influence <strong>of</strong> process parameters on the <strong>drying</strong> kinetics <strong>and</strong> on the formation <strong>of</strong> solid particles<br />

<strong>and</strong> their morphology is dealt with. Different sugar <strong>and</strong> <strong>protein</strong> solutions are considered <strong>and</strong> the<br />

influence <strong>of</strong> the ingredients on the <strong>drying</strong> kinetics <strong>and</strong> particle morphology is discussed. Additional<br />

these results are compared to experimental data <strong>of</strong> laboratory spray-dryers <strong>and</strong> from the literature.


CHAPTER 2 SINGLE DROPLET DRYING 9<br />

2 SINGLE DROPLET DRYING<br />

2.1 General heat transfer considerations<br />

2.1.1 Heat transfer by conduction<br />

Where there exists a temperature gradient within a body, heat energy, Q , will flow from the region<br />

<strong>of</strong> high temperature, T1, to the region <strong>of</strong> low temperature, T2. If the temperatures on both sides <strong>of</strong> a<br />

flat homogeneous isotropic barrier with the surface area, A, <strong>and</strong> the thickness, δ, are constant with<br />

time, the heat transfer is described by Fourier’s Law.<br />

T1<br />

− T2<br />

Equation 2.1 Q = λ ⋅ A⋅<br />

⋅ t<br />

δ<br />

Using the expressions Q & = Q / t <strong>and</strong> q & = Q&<br />

/ A for the heat flux, Equation 2.1 becomes<br />

Equation 2.2<br />

T1<br />

− T2<br />

∂T<br />

q&<br />

= −λ<br />

⋅ = -λ<br />

⋅<br />

δ ∂z<br />

The coefficient λ in the above equations is called heat conductivity <strong>and</strong> is specific for the wall<br />

material. The negative algebraic sign in Equation 2.2 includes the heat transfer with decreasing heat<br />

gradient ∂ T / ∂z<br />

[Bosnjakovic 1997].<br />

If a material or substance is inhomogeneous <strong>and</strong> anisotropic, for example a porous solid,<br />

heat transfer by conduction has to be discussed differently by using a segment model (Figure 2.1).<br />

The segments consist <strong>of</strong> different parallel plates that are arranged vertically <strong>and</strong> horizontally to the<br />

direction <strong>of</strong> heat flow. The distance between the plates can be varied to set the porosity, ε , to a<br />

certain value [Kneule 1975].


10 SINGLE DROPLET DRYING<br />

Figure 2.1: (a-c) Different segment models to describe the heat conduction inside a porous solid material; (I) Segment<br />

model with lengthwise plates, parallel to the heat flow; (II) Segment model with plates vertical to the heat flow [Kneule<br />

1975].<br />

Heat transfer through dry porous solid materials<br />

For dry porous solids, whose interstitials are filled with air, two different cases are possible. In<br />

case I with lengthwise plates (Figure 2.1 (I)), parallel to the heat flow, the transferred heat is given<br />

by<br />

Equation 2.3<br />

&<br />

∂T<br />

A⋅<br />

λ ⋅<br />

∂z<br />

Qtot I = I<br />

λ I is the average thermal conductivity including the porous structure <strong>of</strong> solid<br />

Equation 2.4 λI = ε ⋅ λgas<br />

+ ( 1−<br />

ε ) ⋅ λsolid<br />

Case II (Figure 2.1 (II)) describes the arrangement with a vertical heat flow, which can be treated<br />

similar to a serial connection in electrical engineering<br />

Equation 2.5<br />

&<br />

∂T<br />

A⋅<br />

λ ⋅<br />

∂z<br />

Qtot I = II<br />

With the thermal conductivity λ II<br />

Equation 2.6<br />

λ<br />

II<br />

=<br />

1<br />

ε 1−<br />

ε<br />

+<br />

λ λ<br />

gas<br />

solid


CHAPTER 2 SINGLE DROPLET DRYING 11<br />

To characterize the effective heat conductivity λ eff <strong>of</strong> porous substances a combination <strong>of</strong> both<br />

cases is necessary. The parameter a describes the fraction <strong>of</strong> vertically arranged plates, ( − a)<br />

fraction <strong>of</strong> lengthwise arrange plates [Bosnjakovic 1997; Kneule 1975].<br />

Equation 2.7<br />

λ<br />

1<br />

=<br />

1−<br />

+<br />

λ λ<br />

eff a a<br />

Heat transfer through humid porous solid materials<br />

I<br />

II<br />

1 the<br />

To calculate the effective thermal conductivity <strong>of</strong> humid materials or substances two new cases<br />

have to be taken into account. Case A is applied if the pores <strong>of</strong> the solid substance are totally filled<br />

with water. According to the considerations for dry porous materials, the effective thermal<br />

conductivity is calculated by [Kneule 1975]<br />

Equation 2.8 λA I = ε ⋅ λliquid<br />

+ ( 1−<br />

ε ) ⋅ λsolid<br />

Equation 2.9<br />

λ<br />

A II<br />

=<br />

ε<br />

λ<br />

liquid<br />

1<br />

1−<br />

ε<br />

+<br />

λ<br />

solid<br />

In case B only parts <strong>of</strong> the pores are filled with a gaseous medium. In contrast, the walls <strong>of</strong> the<br />

pores are wetted by the liquid. If there are temperature differences within the porous solid, diffusion<br />

<strong>of</strong> vapour due to a vapour gradient occurs, with evaporation at the warmer spots <strong>and</strong> condensation<br />

at the colder spots [Kneule 1975; Walton 1994]. The total heat transferred inside an ideal material<br />

with parallel pores can be calculated using the thermal conductivity <strong>of</strong> the dry gas, λ gas , <strong>and</strong> the<br />

diffusion related thermal conductivity, λ diff<br />

A<br />

T A T<br />

Equation 2.10 Q& ∂<br />

∂<br />

tot = Q&<br />

gas + Q&<br />

diff = − ⋅ ( λgas<br />

+ λdiff<br />

) ⋅ = − ⋅ λtot<br />

⋅<br />

ε<br />

∂z<br />

ε ∂z<br />

DAB<br />

p ∂pvapour<br />

Equation 2.11 λdiff =<br />

⋅ ⋅ ⋅ Δhv<br />

R ⋅T<br />

p − p ∂T<br />

D<br />

vapour


12 SINGLE DROPLET DRYING<br />

2.1.2 Convective heat transfer<br />

Heat energy transferred between a surface <strong>and</strong> a moving fluid at different temperatures is known as<br />

convection. It is a combination <strong>of</strong> diffusion <strong>and</strong> bulk motion <strong>of</strong> molecules. Near the surface the fluid<br />

velocity is low <strong>and</strong> diffusion dominates. Away from the surface, bulk motion increases influence<br />

<strong>and</strong> therefore dominates. Convective heat transfer may take the form <strong>of</strong> either<br />

- natural (free) convection<br />

- forced (assisted) convection<br />

The essential ingredients for the analysis <strong>of</strong> convective heat transfer are given by Newton’s Law <strong>of</strong><br />

Cooling (Equation 2.12). The rate <strong>of</strong> surface cooling <strong>of</strong> a solid material, immersed in a colder fluid,<br />

is proportional to the difference between temperature <strong>of</strong> the surface <strong>and</strong> the temperature <strong>of</strong> the<br />

cooling fluid [Earle 2004].<br />

Equation 2.12 q& = h ⋅ ( T − T )<br />

A<br />

S<br />

The coefficient h is called the heat transfer coefficient (synonymous film coefficient or film<br />

conductance) <strong>and</strong> can be regarded as the conductance <strong>of</strong> a hypothetical surface film <strong>of</strong> the<br />

thickness, δ , <strong>of</strong> the fluid<br />

Equation 2.13<br />

λ fluid<br />

h =<br />

δ<br />

Heat transfer by convection is more difficult to analyse than heat transfer by conduction, because no<br />

single property <strong>of</strong> the heat transfer medium, such as thermal conductivity, can be defined to<br />

describe the mechanism. Heat transfer by convection varies from situation to situation upon the<br />

fluid flow properties <strong>and</strong> conditions, <strong>and</strong> it is frequently coupled with the mode <strong>of</strong> fluid flow. In<br />

practise, analyses <strong>of</strong> heat transfer <strong>and</strong> heat transfer coefficients are treated empirically by direct<br />

observation [Engineersedge 2005]. Factors that can affect the heat transfer coefficient are<br />

- fluid velocity;<br />

- fluid properties (e.g. fluid density, fluid viscosity);<br />

- heat flux, q& ;<br />

- surface characteristics (e.g. shape, roughness);<br />

- type <strong>of</strong> flow (e.g. single-phase, two-phase).<br />

Of particular interest, not only for <strong>drying</strong> operations, is the heat transfer between a fluid <strong>and</strong> a<br />

definite surface. For the magnitude <strong>of</strong> the heat transfer the properties <strong>of</strong> the fluid layer on the<br />

surface are most important. For this reason L. Pr<strong>and</strong>tl set up the boundary-layer theory in 1904 as a


CHAPTER 2 SINGLE DROPLET DRYING 13<br />

part <strong>of</strong> fluid dynamics essential for heat <strong>and</strong> mass transfer calculations. To calculate the heat<br />

transfer coefficient it has to be connected to the temperature gradient. Directly on the surface (z→0)<br />

the gas velocity can be set zero <strong>and</strong> the heat is only transferred by conduction. Instead <strong>of</strong> Newton’s<br />

law <strong>of</strong> cooling, Fourier’s law is valid [Schlichting 2000].<br />

⎛ ∂T<br />

⎞<br />

Equation 2.14 q& = −λ<br />

⋅ ⎜ ⎟<br />

⎝ ∂z<br />

⎠<br />

Here − λ represents the thermal conductivity <strong>of</strong> the fluid at surface temperature. The heat flux q& is<br />

determined via the slope <strong>of</strong> the tangent at z = 0 <strong>of</strong> the temperature gradient in the fluid.<br />

Equation 2.15<br />

⎛ ∂T<br />

⎞<br />

⎜ ⎟<br />

z<br />

h<br />

⎝ ∂<br />

= −λ<br />

⋅<br />

⎠<br />

T − T<br />

F<br />

S<br />

To calculate the heat transfer coefficient, precise knowledge <strong>of</strong> the temperature gradient inside the<br />

fluid influenced by the flow conditions is necessary. Therefore, not only thermodynamical but also<br />

fluid mechanic principles are important for the survey <strong>of</strong> the heat transfer coefficient<br />

[Bosnjakovic 1997; Earle 2004].<br />

Natural convection is caused by buoyancy forces due to density differences caused by<br />

temperature variations in the fluid. At heating the density change in the boundary layer will cause<br />

the fluid to rise <strong>and</strong> be replaced by cooler fluid that also will heat <strong>and</strong> rise. This continues<br />

phenomena can be seen, for example, in the hot air rising <strong>of</strong>f the surface <strong>of</strong> a radiator<br />

[EngineeringToolBox 2005]. Natural convection rates depend up on the physical properties <strong>of</strong> the<br />

fluid, density ρ , dynamic viscosity μ , thermal conductivity λ , specific heat at constant<br />

pressure c P <strong>and</strong> coefficients <strong>of</strong> thermal expansion β which for gases is = 1/<br />

T by Charles’ Law.<br />

Other factors that affect the convective heat transfer are some linear dimensions <strong>of</strong> the system like<br />

diameter d or length, temperature gradient Δ T <strong>and</strong> the gravitational acceleration g since it is<br />

density differences acted upon by gravity that create circulation [Earle 2004]. It has been shown<br />

experimentally that heat transfer under natural or free convection can be described in terms <strong>of</strong> these<br />

factors grouped in dimensionless numbers [Baehr 1998, Walton 1994]:<br />

- Nusselt number<br />

h ⋅ d<br />

Nu<br />

=<br />

λ


14 SINGLE DROPLET DRYING<br />

- Pr<strong>and</strong>tl number<br />

cP ⋅ µ<br />

Pr =<br />

λ<br />

3 2<br />

d ⋅ ρ ⋅ g ⋅ β ⋅ ΔT<br />

- Grash<strong>of</strong> number Gr =<br />

2<br />

µ<br />

If we assume that these ratios can be related by a simple function, we can write the most general<br />

equation for natural convection [Baehr 1998; Earle 2004; Kneule 1975].<br />

Equation 2.16 ( ) n<br />

m<br />

Nu = K ⋅ (Pr) ⋅ Gr<br />

Experimental work has evaluated K, m <strong>and</strong> n for different technical set-ups under various<br />

conditions. The magnitude <strong>of</strong> the constants in single <strong>droplet</strong> <strong>drying</strong> under natural convection will be<br />

discussed in chapter 2.3.<br />

Forced convection occurs when a fluid flow is induced by an external flow, such as a pump,<br />

fan or mixer [EngineeringToolBox 2005]. The fluid is constantly replaced <strong>and</strong> the rates <strong>of</strong> heat<br />

transfer are therefore higher than for natural convection. In case <strong>of</strong> low fluid flow velocities, where<br />

rates <strong>of</strong> natural convection are comparable to those <strong>of</strong> forced convection, the Grash<strong>of</strong> number is still<br />

significant. But normally the influence <strong>of</strong> natural circulation due to coefficients <strong>of</strong> thermal<br />

expansion <strong>and</strong> gravitational acceleration is replaced by dependence on circulation velocities factors<br />

grouped in another dimensionless number, the Reynolds number [Earle 2004].<br />

- Reynolds number<br />

⋅ ρ ⋅υ<br />

=<br />

μ<br />

d<br />

Re<br />

As for heat transfer for natural convection, a general equation for the heat transfer under forced<br />

convection can be derived [Baehr 1998; Kneule 1975].<br />

Equation 2.17 ( ) n<br />

m<br />

Nu = K ⋅ (Pr) ⋅ Re<br />

The constants K, m <strong>and</strong> n relevant in single <strong>droplet</strong> <strong>drying</strong> under forced convection have to be found<br />

empirically <strong>and</strong> will be discussed in chapter 2.3.<br />

2.1.3 Heat transfer by radiation<br />

Radiant heat transfer is the transfer <strong>of</strong> heat by electromagnetic radiation that arises due to the<br />

temperature <strong>of</strong> a body. The energy is carried by photons <strong>of</strong> light in the infrared <strong>and</strong> visible portions


CHAPTER 2 SINGLE DROPLET DRYING 15<br />

<strong>of</strong> the electromagnetic spectrum in a wavelength range <strong>of</strong> 0.1 to 100 microns [Engineersedge 2005].<br />

Radiation operates independently <strong>of</strong> the medium itself through which it occurs <strong>and</strong> depends upon<br />

the relative temperature, geometric arrangements <strong>and</strong> surface structures like area, reflectivity <strong>and</strong><br />

emissivity <strong>of</strong> the materials emitting or absorbing heat [Earle 2004]. The basic formula for radiant<br />

heat transfer is described by the Stefan-Boltzmann Law<br />

Equation 2.18<br />

4<br />

q& = σ SB ⋅T<br />

T is the absolute temperature in Kelvin, A the total radiating surface area <strong>and</strong> σ SB the Stefan-<br />

Boltzmann constant<br />

−8<br />

−2<br />

−4<br />

σ SB = 5,<br />

67 ⋅10<br />

W⋅<br />

m ⋅ K [Tipler 2000]. This law describes the radiation<br />

energy <strong>of</strong> a perfect radiator, a so-called black body. A black body gives the maximum amount <strong>of</strong><br />

emitted radiation possible at its particular temperature. Equation 2.18 overestimates the energy<br />

emitted by real surfaces at a temperature T , but it was found, that many emit a constant fraction <strong>of</strong><br />

the radiation from a black body [Earle 2004; Haas 2002]. For these real bodies Equation 2.18 can be<br />

rewritten<br />

Equation 2.19<br />

&<br />

4<br />

q = ω ⋅σ<br />

SB ⋅T<br />

The constant ω is called the emissivity <strong>of</strong> a particular body <strong>and</strong> is a number between zero <strong>and</strong> one.<br />

It depends not only on the properties <strong>of</strong> the material but also on the composition <strong>of</strong> the surface area,<br />

for example roughness. Bodies obeying this equation are termed grey bodies. If radiation between<br />

two surfaces occurs the energy transferred depends again upon the above mentioned factors <strong>of</strong> the<br />

two surfaces. For two parallel surfaces, facing each other <strong>and</strong> neglecting edge effects, each must<br />

intercept the total energy emitted by the other, either absorbing or reflecting it [Earle 2004]. Then<br />

the net heat transfer is calculated by<br />

4 4<br />

Equation 2.20 q& = Ε ⋅σ<br />

SB ⋅ ( T − T )<br />

with the temperaturesT 1 <strong>and</strong>T 2 <strong>of</strong> the two bodies <strong>and</strong> with<br />

1 1 1<br />

Equation 2.21 = + −1<br />

E<br />

ω ω<br />

1<br />

2<br />

1<br />

2


16 SINGLE DROPLET DRYING<br />

In case <strong>of</strong> a small body in surroundings that are at uniform temperature, the net heat transfer is<br />

given by<br />

4 4<br />

Equation 2.22 q& = ω ⋅σ<br />

SB ⋅ ( T − T )<br />

1<br />

T 1 is the temperature <strong>of</strong> the body with emissivity 1 ω <strong>and</strong> 2<br />

the surroundings [Earle 2004; Engineersedge 2005].<br />

2.2 Diffusion <strong>and</strong> Mass transfer<br />

2.2.1 Diffusion<br />

1<br />

2<br />

T is the uniform absolute temperature <strong>of</strong><br />

Diffusion is the movement under the influence <strong>of</strong> a physical<br />

stimulus <strong>of</strong> an individual component through a mixture. The<br />

most common driving force <strong>of</strong> diffusion is a concentration<br />

gradient <strong>of</strong> the diffusing component. Molecular diffusion<br />

caused by temperature is called thermal diffusion <strong>and</strong> in case <strong>of</strong><br />

an external field like a pressure gradient, forced diffusion. A<br />

concentration gradient tends to move the diffusing component<br />

into the direction as to equalize the concentrations <strong>and</strong> to<br />

destroy the gradient. If the gradient is maintained by constantly Figure 2.2: Schematic diagram <strong>of</strong><br />

molecular diffusion. A r<strong>and</strong>om path that<br />

supplying component at high concentration side <strong>and</strong> removing molecule A might take in diffusion<br />

through B molecule from point (1) to (2)<br />

it a low concentration side, a steady-state flux is achieved. This is shown [Geankoplis 1993]<br />

is characteristic for many mass transfer operations [McCabe et al. 2005]. In evaporation it is<br />

important to have a look at the diffusion inside a solution or suspension <strong>droplet</strong> as well as at the<br />

vapour diffusion through the liquid-gas interface into the gas phase. In Figure 2.2 the molecular<br />

diffusion process is shown schematically.<br />

Stefan-Maxwell Equation<br />

A simple experimental setup with two glass bulbs connected by a capillary will help to underst<strong>and</strong><br />

the diffusion phenomena. The bulb at the left side contains hydrogen <strong>and</strong> the right side nitrogen.<br />

The two glass bulbs are connected via a capillary. The whole system is at ambient pressure <strong>and</strong><br />

temperature. At the beginning a sharp front between the two gases can be seen in the middle <strong>of</strong> the<br />

capillary. After some hours the mol fraction x <strong>of</strong> hydrogen <strong>and</strong> nitrogen inside the left bulb


CHAPTER 2 SINGLE DROPLET DRYING 17<br />

approaches 0.5. Figure 2.3 shows the concentration equalization inside the capillary at the point z<br />

[Geankoplis 1993].<br />

Figure 2.3: Two glass bulbs connected via a capillary to explain the diffusion phenomena. The points z <strong>and</strong> z + Δz<br />

are the reference points for the derivation <strong>of</strong> the Maxwell-Stefan equation [Krishna <strong>and</strong> Wesselingh 2000]<br />

Consider the mixture between two nearby points z <strong>and</strong> z + Δz<br />

in the capillary. The velocities u A<br />

<strong>and</strong> u B (average diffusive velocities) with which the two components are moving through each<br />

other cause friction between the two species. If the momentum balance <strong>of</strong> hydrogen (component A )<br />

in the slice between z <strong>and</strong> z + Δz<br />

is considered, this contains the following terms [Krishna 2000]:<br />

- a force due to the partial pressure at z : A z A p ⋅<br />

- a force due to the partial pressure at z + dz : A z dz A p − ⋅ +<br />

- the friction force exerted by the two components proportional to their partial pressures:<br />

A<br />

B<br />

( u u )<br />

∝ p ⋅ p ⋅ −<br />

B<br />

A<br />

In steady-state considerations, taking the limit for Δz → 0 <strong>and</strong> using the ideal gas law<br />

= p / RT , the three remaining terms lead to<br />

cA A<br />

( )<br />

RT dpA<br />

Equation 2.23 − ∝ R ⋅T<br />

⋅ pB<br />

⋅ ( uA<br />

− uB<br />

)<br />

p dz<br />

A<br />

Krishna <strong>and</strong> Wesselingh [Krishna 2000] describe the left-h<strong>and</strong> side <strong>of</strong> Equation 2.23 as the driving<br />

force F 1 on component A , whereas the right h<strong>and</strong> side st<strong>and</strong>s for the friction force exerted by<br />

component B on component A . Using xB ∝ pB<br />

, Equation 2.23 can be rewritten as<br />

( u u )<br />

ς A,<br />

B xB ⋅ A − B , where A, B<br />

ς is the friction coefficient between species A <strong>and</strong> B . The complete<br />

equation with driving force, friction force <strong>and</strong> generalized to any component i surrounded by other<br />

components j is given by


18 SINGLE DROPLET DRYING<br />

F i = ∑ i,<br />

j<br />

i≠<br />

j<br />

Equation 2.24 ς x ⋅ ( u − u )<br />

j<br />

i<br />

j<br />

This equation is called the Maxwell-Stefan Equation <strong>and</strong> is much more general than the following<br />

Fick’s law <strong>of</strong> diffusion. It does yield Fick’s equation as a limiting case for simple but important<br />

diffusion problems [Krishna 2000]. Simplifying mathematics the approximated difference form <strong>of</strong><br />

the composition driving force <strong>of</strong> the Maxwell-Stefan equation for the potential gradient can also be<br />

written in the form <strong>of</strong> [Krishna 2000].<br />

Equation 2.25<br />

RT Δxi<br />

F1<br />

= − ⋅<br />

Δz<br />

x<br />

Fick’s law for molecular diffusion<br />

i<br />

The general Fick’s first law equation for one-dimensional steady-state diffusion for the molar flux<br />

J A <strong>of</strong> component A through a reference plane in the binary mixture <strong>of</strong> A <strong>and</strong> B can be written as<br />

Equation 2.26<br />

J<br />

A<br />

= −D<br />

AB<br />

dc<br />

⋅<br />

dz<br />

A<br />

The diffusion flux J A is assumed to be proportional to the concentration gradient dcA / dz . The<br />

diffusivity <strong>of</strong> component A in its mixture with component B is denoted by the binary diffusion<br />

coefficient D AB [Geankoplis 1993; McCabe et al. 2005].<br />

In case <strong>of</strong> unsteady-state diffusion with no chemical reactions, only a two component<br />

system, a constant binary diffusion coefficient <strong>and</strong> a one-dimensional mass transfer, Fick’s second<br />

law can be taken into account [McCabe et al. 2005]<br />

2<br />

A d c<br />

Equation 2.27 = DAB<br />

⋅ 2<br />

dt dz<br />

Diffusion coefficients for gases<br />

dc A<br />

Diffusivities can be determined by a number <strong>of</strong> different experimental methods, but <strong>of</strong>ten the<br />

desired values are not available for the system <strong>of</strong> interest [Geankoplis 1993]. A second possibility is<br />

the estimation from published data or correlations. For diffusion in gases a third approach based on<br />

modern kinetic gas theory allows the prediction <strong>of</strong> binary diffusion coefficients dependent on the<br />

different sizes <strong>and</strong> velocities <strong>of</strong> the molecules <strong>and</strong> the mutual interactions as they approach one


CHAPTER 2 SINGLE DROPLET DRYING 19<br />

another. For a pair <strong>of</strong> non-polar molecules a reasonable approximation to the forces is the Lennard-<br />

Jones function [Geankoplis 1993; McCabe et al. 2005]<br />

Equation 2.28<br />

D<br />

AB<br />

=<br />

0.<br />

001858 ⋅T<br />

1.<br />

5<br />

[ ( M + M ) / ( M ⋅ M ) ]<br />

⋅<br />

p ⋅<br />

A<br />

σ AB<br />

2<br />

σ AB is the effective collision diameter in Å <strong>and</strong> D<br />

AB<br />

A<br />

B<br />

Ω<br />

D<br />

B<br />

A<br />

B<br />

0.<br />

5<br />

Ω the collision integral f ( k T ∈ )<br />

⋅ / with<br />

∈ = ∈ ∈ <strong>and</strong> ∈= the Lennard-Jones force constant for common gases. Equation 2.28 is<br />

relatively complicated to use <strong>and</strong> <strong>of</strong>ten some <strong>of</strong> the contents such asσ AB are not available or<br />

difficult to estimate. Hence, the semi-empirical more convenient method <strong>of</strong> Fuller et al. [1966],<br />

described in Annex A is <strong>of</strong>ten used.<br />

2.2.2 Convective mass transfer considerations<br />

The mass transfer from a surface to a streaming fluid depends on the properties <strong>of</strong> the participating<br />

components <strong>and</strong> on the flow properties <strong>of</strong> the fluid. Similar to the possible conditions for convective<br />

heat transfer, an induced fluid flow leads to mass transfer by forced convection, whereas mass<br />

transfer due to density differences caused by temperature or concentration gradient is called mass<br />

transfer by natural convection [Walton 1994]. Considering mass transfer it has to be differentiated<br />

between two-component mass transfer (equimolar diffusion) <strong>and</strong> one-component mass transfer<br />

(one-way or synonymous unimolar diffusion) [McCabe et al. 2005].<br />

Equimolar diffusion in gases<br />

For equimolar diffusion the mass <strong>of</strong> component A transferred can be calculated using Fick’s first<br />

law according to Equation 2.26 [Baehr 1998].<br />

Equation 2.29<br />

( c − c ) = k ⋅ ( c − )<br />

dm dc D<br />

= c<br />

A⋅<br />

dt dz δ<br />

A AB<br />

J A −J<br />

B = = −D<br />

AB ⋅ = ⋅ Surface ∞<br />

Surface<br />

= , the partial pressure at the surface, p surface , <strong>and</strong> at the edge <strong>of</strong> the layer, p ∞ ,<br />

With c p / ( RT )<br />

Equation 2.29 can be rearranged with the convective mass transfer coefficient k to<br />

k<br />

Equation 2.30 J A<br />

= ⋅ ( pSurface<br />

− p∞<br />

)<br />

R ⋅T<br />

∞<br />

AB


20 SINGLE DROPLET DRYING<br />

General case for diffusion <strong>of</strong> gases plus convection<br />

For a scientific consideration <strong>of</strong> the evaporation process it is necessary to look at the diffusion <strong>of</strong><br />

gases, when the whole fluid is moving in bulk or convective flow. According to the derivation by<br />

Geankoplis, the rate at which moles <strong>of</strong> gas A passes a fixed point to the right, will be taken as<br />

positive flux, J A .<br />

Equation 2.31 J A = ν Ad ⋅ cA<br />

The molar average velocity <strong>of</strong> the whole fluid relative to a stationary point is ν M . The diffusion<br />

velocity ν Ad <strong>of</strong> component A is now measured relatively to the moving fluid. To a stationary<br />

observer, A is moving faster than the bulk <strong>of</strong> the phase, since its diffusion velocity ν Ad is added to<br />

the fluid velocity ν M [Geankoplis 1993]<br />

Equation 2.32 ν A = ν Ad + ν M<br />

Multiplying Equation 2.32 with c A , the first term A A c ⋅ ν can be represented by the flux N A <strong>of</strong> A<br />

relative to a stationary point. The second term is given by Equation 2.31 as J A . The third term is<br />

the convective flux <strong>of</strong> A relative to a stationary point.<br />

Equation 2.33 N A = J A + ν M ⋅ cA<br />

If N is the total flux <strong>of</strong> the whole stream relative to a stationary point, then<br />

Equation 2.34 N = ν M ⋅ cA<br />

= N A + N B<br />

Solving <strong>of</strong> Equation 2.34 for ν M <strong>and</strong> substituting into Equation 2.33 leads to<br />

cA<br />

Equation 2.35 N A = J A + ⋅ ( N A + N B )<br />

c<br />

Since J A is described by Fick’s first law (Equation 2.26) the final general equation for diffusion<br />

plus convection to use when the flux N A is regarded, relative to a stationary point, is given by<br />

∂cA<br />

cA<br />

Equation 2.36 N A =<br />

−D<br />

AB ⋅ + ⋅ ( N A + N B )<br />

∂z<br />

c


CHAPTER 2 SINGLE DROPLET DRYING 21<br />

Unimolar diffusion (A diffusing through stagnant, non-diffusing B)<br />

It <strong>of</strong>ten occurs that only component A diffuses<br />

through a stagnant, non-diffusing component B .<br />

The rate <strong>of</strong> mass transfer for a given concentration<br />

gradient is then greater than if component B<br />

would diffuse in the opposite direction. A typical<br />

example is the evaporation <strong>of</strong> liquids into a gas.<br />

The humid surface acts like a one-way permeable<br />

layer. It is permeable for the water vapour Figure 2.4: Diffusion <strong>of</strong> component A through a<br />

stagnant, non-diffusing B [Geankoplis 1993]<br />

diffusing into the surrounding air but impermeable<br />

for a diffusion <strong>of</strong> air into the water [Baehr 1998]. This process can be explained by the Maxwell-<br />

Stefan Equation 2.24 as well as Equation 2.36 describing the diffusion <strong>of</strong> gases plus convection. If<br />

only component A is diffusing in stagnant, non-diffusing B , then Equation 2.36 can be transferred<br />

using = 0<br />

N <strong>and</strong> the ideal gas law c = p / ( RT )<br />

B<br />

DAB<br />

dpA<br />

p A<br />

Equation 2.37 N A = − ⋅ + ⋅ N A<br />

RT dz p<br />

Rearrangement <strong>and</strong> integration according to Figure 2.4 <strong>of</strong> Equation 2.37 leads to<br />

Equation 2.38<br />

N<br />

A<br />

DAB<br />

= ⋅<br />

RT<br />

p<br />

p −<br />

⋅ ln<br />

−<br />

p<br />

A2<br />

( z2<br />

− z1)<br />

p p A1<br />

Equation 2.38 is <strong>of</strong>ten rewritten in another form using the log mean value <strong>of</strong> the inert component B<br />

as follows. Since p p A1<br />

+ pB1<br />

= pA2<br />

+ pB<br />

2<br />

logarithmic mean value is<br />

Equation 2.39<br />

p p + p<br />

p = p + p the<br />

= <strong>and</strong> B1<br />

= A1<br />

as well as B2<br />

A2<br />

p<br />

BM<br />

pB<br />

2 − pB1<br />

pA1<br />

− p A2<br />

=<br />

=<br />

ln A<br />

( p / p ) ln[<br />

( p − p ) / ( p − p ) ]<br />

B2<br />

B1<br />

Substituting Equation 2.39 in Equation 2.38 leads to [Baehr 1998; Geankoplis 1993]<br />

Equation 2.40<br />

D<br />

A2<br />

AB<br />

N A =<br />

⋅ ⋅ p −<br />

R ⋅T<br />

⋅ 2 1 BM<br />

p<br />

A1<br />

A2<br />

( z − z ) p<br />

The flux <strong>of</strong> component A for a given concentration difference is therefore greater for one-way<br />

diffusion than for equimolar diffusion, since the term p BM is always less than p . This can also be<br />

p<br />

1


22 SINGLE DROPLET DRYING<br />

seen from the approximated driving force <strong>of</strong> the Maxwell-Stefan Equation 2.25 [Krishna 2000]. The<br />

term p / pBM<br />

is a kind <strong>of</strong> correction for unimolar diffusion <strong>and</strong> is called Stefan-flow [Baehr 1998].<br />

The concentration gradient for one-way diffusion is not linear (Figure 2.5). It is steeper at low<br />

values <strong>of</strong> p A . There is no transfer <strong>of</strong> B towards the interface in spite <strong>of</strong> the large concentration<br />

gradient. The explanation is that B tends to diffuse towards the region <strong>of</strong> lower concentration, but<br />

the diffusion flux is just matched by the convective flow carrying B in the opposite direction<br />

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

Figure 2.5: Concentration gradient for equimolar <strong>and</strong> unimolar diffusion. On the left side component A <strong>and</strong> B diffusing<br />

at the same molar rates in opposite directions; on the right side component A diffusion, component B stationary with<br />

respect to interface [McCabe et al. 2005].<br />

Rearrangement <strong>of</strong> Equation 2.40 similar to Equation 2.29 leads to<br />

DAB<br />

p 1<br />

k′<br />

Equation 2.41 N A = ⋅ ⋅ ⋅ ( pA1<br />

− pA2<br />

) = ⋅ ( pA1<br />

− pA2<br />

)<br />

δ p R ⋅T<br />

R ⋅T<br />

with k = ( D / ) ⋅ ( p / p )<br />

AB<br />

BM<br />

BM<br />

′ δ [Baehr 1998, Kneule 1975, McCabe et al. 2005].<br />

2.2.3 Film theory<br />

The basic concept <strong>of</strong> the film theory considers the resistance to diffusion equivalent to that in a<br />

stagnant film <strong>of</strong> a certain thickness δ [McCabe et al. 2005]. This one-dimensional model was<br />

developed to examine the dependence <strong>of</strong> momentum, heat <strong>and</strong> mass transfer on the flow properties<br />

<strong>of</strong> a fluid [Kastner 2001]. Figure 2.6 illustrates the main principles <strong>of</strong> the film theory for a better<br />

underst<strong>and</strong>ing. Component A is transferred from a solid or liquid surface to a streaming fluid. The


CHAPTER 2 SINGLE DROPLET DRYING 23<br />

concentration <strong>of</strong> A at the surface is c A surface <strong>and</strong> inside the<br />

fluid c A ∞ . Concentration <strong>and</strong> diffusion velocity can only<br />

change along the y-axis <strong>and</strong> not along the other axis or<br />

with time [Baehr 1998]. Depending on equimolar or oneway<br />

diffusion, the concentration pr<strong>of</strong>ile <strong>of</strong> A within the<br />

layer can be seen in Figure 2.6 <strong>and</strong> calculated as shown in chapter 2.2.2.<br />

The two film theory was proposed by Lewis <strong>and</strong> Whitman in 1924 (Figure 2.7). It assumes<br />

that a stagnant film layer in each phase δ liquid <strong>and</strong> δ gas represents the resistance to mass transfer <strong>and</strong><br />

that the interface concentrations A surface<br />

surfactants at the interface) [Musonge 2005].<br />

p <strong>and</strong> c A surface are in equilibrium (exception: occurrence <strong>of</strong><br />

Figure 2.7: Concentration pr<strong>of</strong>ile inside the liquid film <strong>and</strong> the gas film according to the concept <strong>of</strong> the two-film theory<br />

by Lewis <strong>and</strong> Whitman in idealized <strong>and</strong> more realistic form [McCabe et al. 2005]<br />

The more complex aspect than considering heat transfer is the discontinuity at the interface,<br />

which occurs because the concentration or mole fraction <strong>of</strong> the diffusing component is hardly ever<br />

the same on opposite sides <strong>of</strong> the interface. The resistance to mass transfer in the two phases is<br />

added to get an overall mass transfer coefficient. If the thickness <strong>of</strong> the film layer is known, it is<br />

possible to determine the overall mass transfer coefficient <strong>of</strong> a particular system using k = D / δ .<br />

The derivation <strong>and</strong> the equation to calculate the film thickness δ <strong>and</strong> the resulting mass transfer<br />

coefficient k can be seen at Bird [1960] [Kastner 2001].<br />

Figure 2.6: Concentration pr<strong>of</strong>ile within the<br />

layer <strong>of</strong> thickness δ at a flat plate<br />

[Baehr 1998].<br />

AB


24 SINGLE DROPLET DRYING<br />

2.2.4 Boundary-layer theory<br />

The boundary layer theory bases upon the assumption that mass transfer takes place in a thin<br />

boundary layer near the surface where the fluid is in laminar flow. In contrast to the film theory,<br />

concentrations <strong>and</strong> velocities are allowed to change not only along the y-axis but also along the<br />

other axis [Baehr 1998; Schlichting 2000]. If the velocity gradient in the boundary layer is linear<br />

<strong>and</strong> the velocity at the surface is zero, the equations for diffusion <strong>and</strong> flow can be solved to give the<br />

concentration gradient <strong>and</strong> the mass transfer coefficient [McCabe et al. 2005]. For flows over a flat<br />

plate or around a cylinder or sphere, the velocity pr<strong>of</strong>iles are also linear at the surface, but the<br />

gradients decrease as the velocity approaches that <strong>of</strong> the main stream at the outer edge <strong>of</strong> the<br />

boundary layer (Figure 2.8) [Hampe et al. 2001].<br />

The analogy in heat <strong>and</strong> mass transfer for boundary-layer considerations permits<br />

correlations for heat transfer to be used for mass transfer, too [McCabe et al. 2005]. Therefore, the<br />

value <strong>of</strong> the convective mass transfer coefficient depends on the same assumptions than the<br />

convective heat transfer coefficient under natural or free convection <strong>and</strong> can be described in terms<br />

<strong>of</strong> these factors grouped in dimensionless numbers [Walton 1994]:<br />

- Sherwood number<br />

- Schmidt number<br />

h ⋅ d<br />

Sh =<br />

D<br />

Sc =<br />

AB<br />

DAB<br />

μ<br />

⋅ ρ<br />

3 2<br />

d ⋅ ρ ⋅ g ⋅ β ⋅ ΔT<br />

- Grash<strong>of</strong> number Gr =<br />

2<br />

µ<br />

- Reynolds number<br />

d ⋅ ρ ⋅υ<br />

Re =<br />

µ<br />

If we assume that these ratios can again be related by a simple function we can write the most<br />

general equation for mass transfer under natural convection<br />

Equation 2.42 ( ) n<br />

m<br />

Sh = K ⋅ (Sc) ⋅ Gr<br />

<strong>and</strong> for mass transfer under forced convection<br />

Equation 2.43 ( ) n<br />

m<br />

Sh =<br />

K ⋅ (Sc) ⋅ Re


CHAPTER 2 SINGLE DROPLET DRYING 25<br />

Similar to the equation for convective heat transfer for natural convection, the constants K, m <strong>and</strong> n<br />

relevant in single <strong>droplet</strong> <strong>drying</strong> under natural or forced convection have to be found empirically<br />

<strong>and</strong> will be discussed in chapter 2.3.<br />

Figure 2.8: Development <strong>of</strong> laminar boundary-layer conditions at a flat solid surface [Hampe et al. 2001].<br />

2.2.5 Penetration theory<br />

Both, the film theory <strong>and</strong> the boundary layer theory assume<br />

steady-state mass transfer [Baehr 1998]. The penetration theory in<br />

contrast makes use <strong>of</strong> the expression for transient rate <strong>of</strong> diffusion<br />

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

surface. The change in concentration with distance <strong>and</strong> time is<br />

usually governed by Fick’s second diffusion law (Equation 2.27)<br />

[McCabe et al. 2005]. The penetration model was first developed<br />

by Higbie in 1935 who examined gas absorption in a liquid,<br />

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

thin layer if the contact time is short. He stated that liquid elements<br />

move from the bulk liquid to the interface. These elements are<br />

loaded transient at the interface <strong>and</strong> penetrate back into the bulk<br />

liquid afterwards (Figure 2.9). According to him, the mass transfer<br />

coefficient is inversely proportional to the square root <strong>of</strong> the contact time t contact<br />

[Baehr 1998; Hampe et al. 2001; McCabe et al. 2005].<br />

Equation 2.44<br />

k = ⋅<br />

π<br />

2<br />

D<br />

t<br />

AB<br />

contact<br />

Figure 2.9: In-stationary loading <strong>of</strong><br />

liquid elements according to the<br />

concept <strong>of</strong> the penetration theory by<br />

Higbie [Hampe et al. 2001]


26 SINGLE DROPLET DRYING<br />

2.2.6 Surface renewal theory<br />

Danckwerts developed an alternative form <strong>of</strong> the penetration<br />

theory (Figure 2.10). Because it was unlikely that all elements<br />

<strong>of</strong> the fluid had the same interfacial residence time, he<br />

considered the case where the elements <strong>of</strong> fluid at the transfer<br />

surface are r<strong>and</strong>omly replaced by fresh fluid, <strong>and</strong> the mass<br />

transfer coefficient is given by k DAB<br />

s ⋅ = where s is the<br />

fraction rate <strong>of</strong> surface renewal in seconds -1 [Hampe et al.<br />

2001; McCabe et al. 2005].<br />

2.3 Heat <strong>and</strong> mass transfer considerations for pure solvent <strong>droplet</strong>s<br />

2.3.1 The D 2 -law<br />

Figure 2.10: Model <strong>of</strong> the surface<br />

renewal theory with a r<strong>and</strong>omly<br />

replacement <strong>of</strong> the liquid elements<br />

according to Danckwerts [Hampe et al.<br />

2001]<br />

Two methods are usually used to analyse the mass transfer from an evaporating <strong>droplet</strong> through the<br />

gas phase <strong>of</strong> the surrounding environment. One is to solve the conservation equations for a<br />

motionless <strong>droplet</strong> in an infinite stagnant medium <strong>and</strong> to employ an empirical correction factor to<br />

account for natural or forced convection around the <strong>droplet</strong> [Faeth 1977; Yao et al. 2003]. This<br />

method is termed the “classic method” <strong>and</strong> will be discussed in this chapter. The other method is to<br />

employ the film theory which involves analysing the thickness <strong>of</strong> a layer for heat transfer <strong>and</strong> for<br />

mass transfer [Abramzon 1989; Bird 1960; Sirignano 1999]. The values <strong>of</strong> the layers are<br />

determined empirically <strong>and</strong> heat <strong>and</strong> mass transfer on film thickness are neglected. This last<br />

suggestion leads to two models with almost identical equations for heat <strong>and</strong> mass transfer rates<br />

[Yao et al. 2003]. Chapter 2.4.2 will address to the basic equations <strong>of</strong> the model by Abramzon <strong>and</strong><br />

Sirignano based on the film theory.<br />

A single <strong>droplet</strong> with the radius r S (diameter d S ) <strong>and</strong> the temperature T S is brought into an<br />

environment with the temperature T ∞ <strong>and</strong> the mass fraction <strong>of</strong> vapour y A ∞ <strong>of</strong> the <strong>droplet</strong> material.<br />

In case the temperature <strong>of</strong> the ambience is much higher than the initial <strong>droplet</strong> temperature, the


CHAPTER 2 SINGLE DROPLET DRYING 27<br />

resulting temperature-time curve can be seen in Figure 2.11. First, there is little mass diffusion from<br />

the <strong>droplet</strong> early in the process due to a low solvent concentration at the liquid surface. Then the<br />

<strong>droplet</strong> heats up like any other cold body placed into a hot environment. As the liquid temperature<br />

rises, the rate <strong>of</strong> mass transfer increases as a result <strong>of</strong> higher solvent vapour concentration at the<br />

<strong>droplet</strong> surface [Faeth 1977]. According to Frohn <strong>and</strong> Roth the temperatures are not uniform within<br />

the <strong>droplet</strong>. There is a maximum liquid temperature at the surface [Frohn 2000]. The increased mass<br />

transfer leads to an increasing portion <strong>of</strong> energy reaching the drop surface that supplies the latent<br />

heat <strong>of</strong> vaporization <strong>of</strong> the evaporating solvent. Also the outward flow <strong>of</strong> vapour in the boundary<br />

layer reduces the rate <strong>of</strong> heat transfer to the <strong>droplet</strong>. This effect slows down the rate <strong>of</strong> increase <strong>of</strong><br />

the liquid surface temperature <strong>and</strong> later in the process a stage is reached where all the heat utilized<br />

to the surface is needed for the latent heat <strong>of</strong> vaporization <strong>and</strong> the <strong>droplet</strong> stabilizes at the wet bulb<br />

temperature [Faeth 1977].<br />

Figure 2.11: Representation <strong>of</strong> the mass flux <strong>of</strong> vapour m& <strong>and</strong> the heat fluxes in the gas phase vap<br />

Qgas & <strong>and</strong> in the liquid<br />

Qliq & <strong>of</strong> an evaporation <strong>droplet</strong>. In the diagram the mass fraction <strong>of</strong> vapour Y vap = y <strong>and</strong> the radial distribution <strong>of</strong> the<br />

vap<br />

temperature are shown. The pr<strong>of</strong>ile on the left side is observed, if the temperature <strong>of</strong> the ambience is much higher than<br />

the initial <strong>droplet</strong> temperature. The pr<strong>of</strong>ile on the right side is achieved, if the ambient <strong>and</strong> initial <strong>droplet</strong> temperature<br />

are equal [Frohn 2000].<br />

A <strong>droplet</strong> evaporating at its wet bulb temperature will have a linear decrease <strong>of</strong> its squared diameter<br />

or radius with time [Faeth 1977; Law 1982]:<br />

2<br />

dd S<br />

Equation 2.45 = −β<br />

d<br />

dt<br />

dr<br />

dt<br />

or<br />

2<br />

S = −β<br />

r<br />

β d <strong>and</strong> β r are the evaporation coefficients, which are a function <strong>of</strong> fluid properties <strong>and</strong> the ambient<br />

conditions. This relationship is called “d 2 -Law”. It is the simplest model describing the


28 SINGLE DROPLET DRYING<br />

vaporisation <strong>of</strong> a <strong>droplet</strong> [Law 1982]. The major assumptions built in this theory are discussed by<br />

Law [1982] <strong>and</strong> can be seen in Table 2.1. Considering that liquid phase heat <strong>and</strong> mass transport<br />

processes are completely neglected, the d 2 -law is essentially a gas-phase model.<br />

Table 2.1: Major assumptions built into the d 2 -law theory according to Law [1982].<br />

Assumption Explanation<br />

Spherical symmetry<br />

Forced <strong>and</strong> natural convection are neglected. This reduces the analysis to one<br />

dimension.<br />

No spray effects Isolated one immersed <strong>droplet</strong> in an infinite environment<br />

Diffusion being rate-controlling<br />

Isobaric process<br />

Constant gas-phase properties Specific heats <strong>and</strong> thermal conductivities are constant during the evaporation<br />

process<br />

Gas-phase quasi-steadiness Due to significant density disparity between liquid <strong>and</strong> gas. The liquid possesses<br />

great inertia resulting in a much slower change <strong>of</strong> e.g. regression rate, species<br />

concentrations <strong>and</strong> temperature than those <strong>of</strong> the gas phase transport processes.<br />

<strong>Single</strong> solvent species Thus it is unnecessary to analyse liquid phase mass transport<br />

Constant <strong>and</strong> uniform <strong>droplet</strong><br />

temperature<br />

Saturation vapour pressure at<br />

surface<br />

No Soret, no Dufour <strong>and</strong> no<br />

radiation effects<br />

No <strong>droplet</strong> heating<br />

Phase-change process between liquid <strong>and</strong> vapour occurs at rates much faster than<br />

those for gas-phase transport. Therefore gasification at the surface is at equilibrium<br />

producing fuel vapour at its saturation vapour pressure at the wet bulb temperature.<br />

No Soret: no mass flow because <strong>of</strong> a temperature gradient<br />

No Dufour: no heat flow because <strong>of</strong> a concentration gradient.<br />

No effects due to radiative heat transfer.<br />

The basic gas phase equations for the conservation <strong>of</strong> mass, vapour <strong>and</strong> energy are given by Bird<br />

[1960], Faeth [1977] or Law [1982] as well as Frohn <strong>and</strong> Roth [Frohn 2000]:<br />

Conservation <strong>of</strong> mass:<br />

d<br />

2<br />

Equation 2.46 ( ρ<br />

gas ⋅ r ⋅ν<br />

r ) = 0<br />

dr


CHAPTER 2 SINGLE DROPLET DRYING 29<br />

Conservation <strong>of</strong> species (vapour):<br />

d ⎡ 2 ⎛<br />

dyi<br />

⎞⎤<br />

Equation 2.47 ⎢r<br />

⋅ ⎜ ρ gas ⋅ν<br />

r ⋅ yi<br />

− ρ gas ⋅ DAB<br />

⋅ ⎟⎥<br />

= 0<br />

dr ⎣ ⎝<br />

dr ⎠⎦<br />

Conservation <strong>of</strong> energy:<br />

d ⎡ 2 ⎛<br />

dT ⎞⎤<br />

Equation 2.48 ⎢r<br />

⋅ ⎜ ρ gas ⋅ν<br />

r ⋅ cP<br />

⋅ ( T − T∞<br />

) − λgas<br />

⋅ ⎟⎥<br />

= 0<br />

dr ⎣ ⎝<br />

dr ⎠⎦<br />

In these equations r represents the radial coordinate, ρ gas the density <strong>of</strong> the gas <strong>and</strong> ν r the radial<br />

velocity <strong>of</strong> the gas surrounding the <strong>droplet</strong>. y i is the mass fraction <strong>of</strong> vapour, c P the specific heat<br />

<strong>of</strong> the gas <strong>and</strong> λ gas the heat conductivity <strong>of</strong> the ambience. Since only the solvent has net mass<br />

transfer, D AB represents the binary diffusion coefficient <strong>of</strong> the solvent with respect to the gas phase<br />

species. The analysis considers only two species, solvent <strong>and</strong> ambient gas. By definition ∑ y i = 1<br />

<strong>and</strong> only one conservation equation must be solved [Faeth 1977; Law 1982]. Integration <strong>of</strong><br />

Equation 2.46 yields<br />

m&<br />

2<br />

vap<br />

Equation 2.49 r ⋅ ρliquid<br />

⋅ν<br />

r = = constant<br />

4π<br />

With boundary conditions<br />

r = r<br />

S<br />

:<br />

r = ∞ :<br />

T = T ,<br />

S<br />

T = T ,<br />

∞<br />

y<br />

y<br />

vap<br />

vap<br />

= y<br />

= y<br />

vap S<br />

vap ∞<br />

the solution <strong>of</strong> Equation 2.49 leads to the following<br />

Equation 2.50<br />

m&<br />

vap<br />

= 4π ⋅ ρ<br />

gas S<br />

⋅ D<br />

AB<br />

⎛ yvap<br />

S − y<br />

⋅ r ⋅ ⎜ S ln 1−<br />

⎜<br />

⎝ 1−<br />

yvap<br />

Where r S is the <strong>droplet</strong> radius <strong>and</strong> ρ gas S the density <strong>of</strong> the gas at the <strong>droplet</strong> surface temperature.<br />

The so-called Spalding transfer number B M is a convenient driving potential for the definition <strong>of</strong> a<br />

mass transfer coefficient for the diffusion <strong>of</strong> gas through a stagnant gas [Faeth 1977].<br />

vap ∞<br />

S<br />

⎞<br />

⎟<br />


30 SINGLE DROPLET DRYING<br />

Equation 2.51<br />

B M<br />

=<br />

y<br />

vap S<br />

− y<br />

1 − y<br />

vap S<br />

vap ∞<br />

The conservation <strong>of</strong> mass for a liquid <strong>droplet</strong> yields<br />

Equation 2.52<br />

dm<br />

dt<br />

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

d ⎛ 4 3 ⎞<br />

2 drS<br />

( ρliquid<br />

⋅V<strong>droplet</strong><br />

) = ⎜ π ⋅ rS<br />

⋅ ρliquid<br />

⎟ = 4π<br />

⋅ rS<br />

⋅ liquid ⋅ = −m&<br />

vap<br />

d<br />

= ρ<br />

dt<br />

dt ⎝ 3<br />

⎠<br />

Employing Equation 2.51 <strong>and</strong> Equation 2.52 into Equation 2.50 the radius-time curse is given by<br />

2 2<br />

Equation 2.53 = r − β ⋅ t<br />

where S , 0<br />

rS S,<br />

0 r<br />

r is the initial <strong>droplet</strong> radius <strong>and</strong> β r is calculated by<br />

2 ⋅ ρ gas ⋅ DAB<br />

Equation 2.54 β r =<br />

⋅ ln(<br />

1−<br />

BM<br />

)<br />

ρ<br />

liquid<br />

The evaporation model set up in Equation 2.53 <strong>and</strong> Equation 2.54 is the “classical” way <strong>of</strong><br />

predicting the radius-time course <strong>of</strong> an evaporation <strong>droplet</strong> following the assumptions <strong>of</strong> the d 2 -law<br />

mentioned above. To employ an empirical correction to account for forced or natural convection<br />

many experimental approaches where carried out within the 20 th century. Only two <strong>of</strong> them are<br />

introduced here, because <strong>of</strong> their most basic character. In 1927 Froessling examined the evaporation<br />

<strong>of</strong> pure solvent <strong>droplet</strong>s <strong>of</strong> nitrobenzene, aniline <strong>and</strong> water under air flow conditions at Reynolds<br />

numbers from 2 to 800 [Froessling 1927]. The <strong>droplet</strong>s with diameters <strong>of</strong> 0.2 mm to 1.8 mm were<br />

suspended from a thin glass filament or thermocouple <strong>and</strong> photographed in planned intervals during<br />

the evaporation experiments. In correlation to his theoretical considerations he found:<br />

Equation 2.55<br />

Sh = 2 + 0.<br />

552 ⋅ Re ⋅<br />

1/ 2 1/<br />

3<br />

Sc<br />

Ranz <strong>and</strong> Marshall constructed a special dryer in form <strong>of</strong> a hydrophobic glass filament,<br />

thermocouple <strong>and</strong> capillary burette to investigate the evaporation kinetics <strong>of</strong> <strong>droplet</strong>s under natural<br />

<strong>and</strong> forced convection [Ranz 1952 a; Ranz 1952 b]. They examined the evaporation <strong>of</strong> <strong>droplet</strong>s <strong>of</strong><br />

0.6mm to 1.1 mm <strong>of</strong> pure solvents (water, benzene, aniline), solutions (Water/NH4NO3;<br />

Water/NaCl) <strong>and</strong> suspensions at temperatures from 20°C to 220°C. The evaporation rate was<br />

determined by monitoring the decrease <strong>of</strong> the <strong>droplet</strong> diameter with time or by measuring the<br />

dt


CHAPTER 2 SINGLE DROPLET DRYING 31<br />

decrease in liquid level within the burette while keeping the <strong>droplet</strong> diameter constant. The<br />

correlation factor found for natural convection is given by<br />

Equation 2.56<br />

<strong>and</strong> for forced convection by<br />

Equation 2.57<br />

Nu = 2 + 0.<br />

6 ⋅ Pr ⋅<br />

Sh = 2 + 0.<br />

6 ⋅Sc<br />

⋅<br />

Nu = 2 + 0.<br />

6 ⋅ Re ⋅<br />

Sh = 2 + 0.<br />

6 ⋅ Re ⋅<br />

0. 33 0.<br />

25<br />

Gr<br />

0. 33 0.<br />

25<br />

Gr<br />

1/ 2 1/<br />

3<br />

Pr<br />

1/ 2 1/<br />

3<br />

Sc<br />

To describe the mass transfer <strong>and</strong> the evaporation rate under natural <strong>and</strong> forced convection for a<br />

pure solvent <strong>droplet</strong>, Equation 2.54 is changed to<br />

2 ⋅ϕ<br />

gas ⋅ DAB<br />

Sh<br />

Equation 2.58 β r =<br />

⋅ ln(<br />

1−<br />

BM<br />

) ⋅<br />

ρ<br />

2<br />

liquid<br />

2.3.2 Abramzon <strong>and</strong> Sirignano’s model<br />

An approach based on the film theory was carried out by Abramzon <strong>and</strong> Sirignano [Abramzon<br />

1989; Sirignano 1999]. The film theory assumes that the resistance to heat <strong>and</strong> mass transfer<br />

between a surface <strong>and</strong> a gas flow is modelled by introducing the concept <strong>of</strong> film thickness: δ T <strong>and</strong><br />

δ M [Yao et al. 2003]. The presence <strong>of</strong> the Stefan flow will influence the values <strong>of</strong> δ T <strong>and</strong> δ M , since<br />

a surface blowing results in the thickening <strong>of</strong> the laminar boundary. [Schlichting 2000]. Abramzon<br />

<strong>and</strong> Sirignano stated that for an evaporating spherical <strong>droplet</strong>, the radii <strong>of</strong> the thermal <strong>and</strong><br />

diffusional films are the obtained from [Abramzon 1989].<br />

Equation 2.59<br />

∗<br />

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

∗<br />

Nu<br />

⋅<br />

Nu − 2<br />

film,<br />

T = rS<br />

∗<br />

r<br />

∗<br />

Sh<br />

⋅<br />

Sh − 2<br />

film,<br />

M = rS<br />

∗<br />

r<br />

∗<br />

Sh are the modified Nusselt number <strong>and</strong> Sherwood number [Kastner 2001]. To<br />

characterise the change <strong>of</strong> the film thickness relative to the initial film thickness due to the Stefan<br />

flow, the correction factor F is introduced


32 SINGLE DROPLET DRYING<br />

Equation 2.60<br />

FT<br />

FM<br />

δT<br />

=<br />

δ<br />

T , 0<br />

δ<br />

=<br />

δ<br />

M<br />

M , 0<br />

To find T F <strong>and</strong> F M for the film thickness, the model problem <strong>of</strong> the laminar boundary-layer flow<br />

past a vaporizing wedge was considered [Abramzon 1989]. In case <strong>of</strong> isothermal surface <strong>and</strong><br />

constant physical properties <strong>of</strong> the fluid, the correction factors do not depend on local Reynolds<br />

number <strong>and</strong> are practically insensitive to Pr<strong>and</strong>tl number <strong>and</strong> Schmidt numbers <strong>and</strong> can be<br />

approximated as function <strong>of</strong> the Spalding mass transfer number B M <strong>and</strong> the Spalding heat transfer<br />

number B T [Abramzon 1989; Kastner 2001].<br />

Equation 2.61 F ( B ) = ( 1 + B )<br />

T<br />

F M<br />

T<br />

T<br />

0.<br />

7<br />

( B ) = ( 1 + B )<br />

M<br />

M<br />

ln<br />

⋅<br />

0.<br />

7<br />

( 1 + B )<br />

ln<br />

⋅<br />

B<br />

T<br />

T<br />

( 1 + B )<br />

Using these factors the modified Nusselt number <strong>and</strong> Sherwood number are calculated by<br />

Equation 2.62<br />

Nu<br />

Sh<br />

∗<br />

∗<br />

Nu−<br />

2<br />

= 2 +<br />

FT<br />

Sh−<br />

2<br />

= 2 +<br />

F<br />

M<br />

As in other approximate models, the film model assumes that temperature <strong>and</strong> solvent vapour<br />

concentrations along the surface are uniform [Abramzon 1989; Yao et al. 2003]. Finally, the model<br />

yields the following expressions for the <strong>droplet</strong> radius-time course <strong>and</strong> the evaporation coefficient<br />

2 2<br />

Equation 2.63 = r − β ⋅ t<br />

rS S,<br />

0 r , film<br />

2 ⋅ϕ<br />

gas ⋅ DAB<br />

Sh<br />

Equation 2.64 β r,<br />

film = ⋅ ln(<br />

1−<br />

BM<br />

) ⋅<br />

ρ<br />

2<br />

liquid<br />

∗<br />

For a spherical <strong>droplet</strong> in an environment with no natural or forced convection Sh = Sh = 2.<br />

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

Equation 2.64 turns into Equation 2.54. The same considerations are valid for the Nusselt<br />

∗<br />

number Nu = Nu = 2.<br />

0 .<br />

B<br />

M<br />

M<br />


CHAPTER 2 SINGLE DROPLET DRYING 33<br />

2.3.3 Diffusion-controlled evaporation <strong>of</strong> a single <strong>droplet</strong><br />

If it is assumed that the <strong>droplet</strong> is surrounded<br />

by an atmosphere that has approximately the<br />

same temperature as the <strong>droplet</strong> liquid <strong>and</strong> is<br />

low in comparison with the boiling point<br />

temperature <strong>of</strong> the <strong>droplet</strong> liquid, the<br />

evaporation can be described by a simpler<br />

equation than Equation 2.54 [Fuchs 1959;<br />

Niven 1980]. Under these conditions the<br />

evaporation process is dominated by diffusion<br />

in the vapour phase. In this special case υr = 0<br />

<strong>and</strong> the d 2 -law can be derived directly by<br />

integrating Equation 2.47 [Frohn 2000]:<br />

Equation 2.65<br />

dc<br />

2<br />

− m& vap = DAB<br />

⋅ A⋅<br />

= 4π ⋅ r ⋅ DAB<br />

⋅<br />

dr<br />

Equation 2.65 is equal to Fick’s first law for a spherical <strong>droplet</strong> with the surface area<br />

dc<br />

dr<br />

4 ⋅ r<br />

2<br />

π . Upon<br />

integration dc / dr where c ∞ is the vapour concentration for r → ∞ <strong>and</strong> r S the <strong>droplet</strong> radius with<br />

the vapour concentration ( rS<br />

) cS<br />

c = , the rate <strong>of</strong> evaporation is obtained by the expression<br />

Equation 2.66 & = 4π<br />

⋅ r ⋅ D ⋅ ( c − c )<br />

mvap S AB S<br />

Substituting the vapour concentration by the partial vapour pressure <strong>of</strong> the liquid at the surface with<br />

the surface temperature T S <strong>and</strong> in the surrounding air with the temperature T ∞ using the ideal gas<br />

law yields<br />

Equation 2.67 ⎟ M liquid ⎛ pS<br />

p ⎞ ∞<br />

m&<br />

vap = 4π<br />

⋅ rS<br />

⋅ DAB<br />

⋅ ⋅ ⎜ −<br />

R ⎝ TS<br />

T∞<br />

⎠<br />

Figure 2.12: Schematic description <strong>of</strong> the relevant<br />

variables in diffusion-controlled evaporation <strong>of</strong> a single<br />

<strong>droplet</strong>. Droplet shrinkage <strong>and</strong> water vapour diffusion are<br />

in equilibrium at any time.<br />

To show that the d 2 -law follows from this expression, Equation 2.67 is set equal to the mass flux<br />

caused by the shrinkage <strong>of</strong> the evaporating <strong>droplet</strong> following Equation 2.52<br />


34 SINGLE DROPLET DRYING<br />

Equation 2.68 ⎟ 2<br />

dr 2 ⋅ D ⋅ M ⎛ ⎞<br />

S<br />

AB liquid pS<br />

p∞<br />

− =<br />

⋅ ⎜ −<br />

dt ρliquid<br />

⋅ R ⎝ TS<br />

T∞<br />

⎠<br />

Integration leads to decrease <strong>of</strong> the squared diameter with time according to the d 2 -law<br />

2 ⋅ DAB<br />

⋅ M<br />

2 2<br />

liquid ⎛ p p ⎞<br />

S ∞<br />

Equation 2.69 rS<br />

= rS<br />

, 0 −<br />

⋅ ⋅ t<br />

liquid R ⎜ −<br />

TS<br />

T ⎟<br />

ρ ⋅ ⎝<br />

∞ ⎠<br />

Where the evaporation coefficient is given by<br />

Equation 2.70 ⎟ 2 ⋅ DAB<br />

⋅ M liquid ⎛ p ⎞<br />

S p∞<br />

β r =<br />

⋅ ⎜ −<br />

ρliquid<br />

⋅ R ⎝ TS<br />

T∞<br />

⎠<br />

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

If an evaporating <strong>droplet</strong> consists <strong>of</strong> a mixtures <strong>of</strong> different solvents it is important to look at the<br />

vapour concentration <strong>of</strong> the components at the <strong>droplet</strong> surface. If the mass flux <strong>of</strong> liquid<br />

component A is bigger than <strong>of</strong> liquid component B , e.g. ethanol in a water/ethanol mixture, two<br />

different <strong>drying</strong> stages can be seen. In the first stage the evaporation rate is determined by<br />

component A . The <strong>droplet</strong> radius decreases faster with time than in the second stage. When almost<br />

all <strong>of</strong> component A has evaporated its mass flux turns little <strong>and</strong> the evaporation process is<br />

determined by the evaporation <strong>of</strong> component B . In this second stage the radius <strong>of</strong> the <strong>droplet</strong><br />

decreases slower than in the first stage. The surface temperature <strong>of</strong> the mixture <strong>droplet</strong> is always<br />

between the cooling temperatures <strong>of</strong> the liquid component A <strong>and</strong> component B [Rensink 2004].<br />

The differential form describing the decrease <strong>of</strong> the <strong>droplet</strong> radius <strong>of</strong> a solvent mixture with time is<br />

given by<br />

Equation 2.71<br />

2⋅<br />

⎡<br />

⋅ ⎢D<br />

⎣<br />

⎤ 2 r<br />

⎥ − ⋅<br />

⎦ 3 ρ<br />

d<br />

⋅<br />

dT<br />

2<br />

∗ ∗<br />

2<br />

dr ρ<br />

S gas<br />

Sh<br />

Sh<br />

ρ<br />

S liquid dTS<br />

=<br />

AB,<br />

A ⋅ln(<br />

1+<br />

BM,<br />

A ) ⋅ + DAB,<br />

B ⋅ln(<br />

1+<br />

BM,<br />

B ) ⋅<br />

⋅<br />

dt ρ liquid<br />

2<br />

2<br />

liquid S dt<br />

The precise derivation <strong>of</strong> the above equation is shown by Rensink using Raoult’s <strong>and</strong> Dalton’s law<br />

for the partial vapour pressures <strong>and</strong> the activity coefficients by Wilson [1964] for two component<br />

mixtures [Rensink 2004]. In contrast to pure solvent <strong>droplet</strong>s the surface temperature is not constant<br />

during the evaporation time <strong>and</strong> the temperature dependence <strong>of</strong> the liquid density must be<br />

considered within the calculations.


CHAPTER 2 SINGLE DROPLET DRYING 35<br />

2.3.5 Droplet-Gas Interactions<br />

Liquid <strong>droplet</strong>s may experience mechanical interactions with gas flows. Droplet-gas interactions<br />

can produce changes <strong>of</strong> resulting flow velocity <strong>and</strong> surface temperature <strong>of</strong> the <strong>droplet</strong>s that<br />

influence heat <strong>and</strong> mass transfer or lead to distortion <strong>and</strong> break up. For the calculation <strong>of</strong> <strong>droplet</strong><br />

motion it is important to know the forces with which the surrounding fluid acts on the <strong>droplet</strong>s<br />

[Kastner 2001]. These forces are usually expressed in terms <strong>of</strong> the drag coefficient, CD. When the<br />

viscosity <strong>of</strong> a liquid is low the deformation is determined primarily by aerodynamic forces <strong>and</strong><br />

forces resulting from surface tension [Walton 1994]. For the steady motion <strong>of</strong> single rigid spherical<br />

particles CD depends only on the Reynolds number. The results <strong>of</strong> experimental investigation for<br />

this case are known as st<strong>and</strong>ard drag curve or st<strong>and</strong>ard drag coefficient [Frohn 2000]. The more<br />

realistic case <strong>of</strong> non-steady particle motion has been studied by Tempkin [1982].<br />

Considering a one-dimensional gas flow parallel to the <strong>droplet</strong> movement the <strong>droplet</strong> motion<br />

<strong>and</strong> radius reduction are governed by the following equations derived by Abramzon <strong>and</strong> Sirignano<br />

[Abramzon 1989]:<br />

Equation 2.72<br />

du<br />

dt<br />

3 µ<br />

= ⋅<br />

16 ρ<br />

⋅<br />

( u − u )<br />

⋅ Re⋅<br />

D G ∞ <strong>droplet</strong><br />

C<br />

2<br />

D<br />

liquid rS<br />

Using Equation 2.72 the relative velocity between <strong>droplet</strong> u <strong>droplet</strong> <strong>and</strong> gas flow u∞ decreases all time,<br />

according to real situations <strong>of</strong> a non-steady environment. A number <strong>of</strong> experimental investigations<br />

have been made to determine how C D varies with the <strong>droplet</strong> Reynolds number [Schlichting 2000;<br />

Kastner 2001].<br />

( Re)


36 SINGLE DROPLET DRYING<br />

Table 2.2: Correlations to calculate the drag coefficient <strong>of</strong> a rigid sphere atr different Reynolds numbers [Kastner 2001]<br />

Literature Drag coefficient Valid for<br />

Stokes<br />

24<br />

C =<br />

[Schlichting 2000] D<br />

rigid sphere; Re


CHAPTER 2 SINGLE DROPLET DRYING 37<br />

The <strong>drying</strong> first stage – Constant rate period<br />

The <strong>drying</strong> <strong>of</strong> <strong>droplet</strong>s with dissolved or suspended<br />

solids involves a period <strong>of</strong> surface evaporation<br />

comparable to the vaporisation <strong>of</strong> pure liquid <strong>droplet</strong>s<br />

at constant wet bulb temperature. The volume change<br />

with time over the first <strong>drying</strong> stage follows the d 2 -<br />

law, although the <strong>droplet</strong>s contain a solid component<br />

[Kastner et al. 2001]. As long as the surface <strong>of</strong> the<br />

<strong>droplet</strong> is completely wetted the evaporated amount<br />

<strong>of</strong> solvent with time stays constant (Figure 2.13 A-B).<br />

For this reason the first phase in the <strong>drying</strong> process is<br />

also called “constant rate” [Grassmann et al. 1997;<br />

Masters 2002]. Taking the vapour pressure lowering<br />

effect <strong>of</strong> the dissolved solid into account the<br />

evaporation coefficient in the first stage can be<br />

calculated using Equation 2.58. The wet bulb<br />

temperature <strong>of</strong> the solution or suspension presents the<br />

<strong>droplet</strong> temperature within this stage. Much <strong>of</strong> the<br />

available moisture in a <strong>droplet</strong> is removed during the<br />

first period <strong>of</strong> <strong>drying</strong>. Moisture migrates from the<br />

<strong>droplet</strong> interior to the surface at a rate to maintain<br />

surface saturation. Capillary <strong>and</strong> diffusion<br />

mechanisms are involved depending upon the nature<br />

<strong>of</strong> the solids in the <strong>droplet</strong> as a solution or suspension<br />

[Masters 2002].<br />

Figure 2.13: Drying course at constant ambient<br />

conditions. The continuous lines show the <strong>drying</strong><br />

behaviour <strong>of</strong> a hygroscopic substance, the dashed<br />

lines <strong>of</strong> a non-hygroscopic substance. Diagram (a)<br />

to (c) are the most common way to plot <strong>drying</strong><br />

kinetic curves:<br />

(a) Water content with time;<br />

(b) Evaporation rate with time;<br />

(c) Evaporation rate with water content.<br />

[Grassmann et al. 1997]<br />

Due to the loss <strong>of</strong> liquid mass during the evaporation the solid becomes more <strong>and</strong> more<br />

concentrated within the remaining <strong>droplet</strong> <strong>and</strong> eventually a point is reached when the rate <strong>of</strong><br />

migration <strong>of</strong> moisture to the surface becomes the limiting factor in <strong>drying</strong> rate. Surface wetness can<br />

no longer be maintained <strong>and</strong> a falling-<strong>of</strong>f in <strong>drying</strong> rate results [Geankoplis 1993; Masters 2002].<br />

This can lead to a radial distribution <strong>of</strong> the solid <strong>and</strong> furthermore to a crust formation initiated in the<br />

outer region <strong>of</strong> the <strong>droplet</strong> close to the surface [Kastner 2001]. Ranz <strong>and</strong> Marshall explained this<br />

behaviour on the basis, that the rate <strong>of</strong> diffusion <strong>of</strong> the dissolved material back into the <strong>droplet</strong> is<br />

slow compared to the rate <strong>of</strong> solvent evaporation on the surface [Ranz 1952 a; Ranz 1952 b]. The


38 SINGLE DROPLET DRYING<br />

first stage <strong>of</strong> <strong>droplet</strong> <strong>drying</strong> ends when the solid begins to form a crust at the surface. This may<br />

occur before conditions <strong>of</strong> uniform saturation throughout the <strong>droplet</strong> are reached according to the<br />

above considerations by Ranz <strong>and</strong> Marshall [Walton 1994]. To estimate the <strong>drying</strong> time <strong>of</strong> the<br />

constant rate period it is necessary to predict the critical moisture content <strong>of</strong> a formulation, that is<br />

the moisture content at the end <strong>of</strong> the first <strong>drying</strong> stage [Grassmann et al. 1997].<br />

The critical moisture content – Critical point<br />

The change from the first to the second <strong>drying</strong> stage shows a sharp inflexion point within the<br />

evaporation curve Figure 2.13 B. This point is characterized by the crust formation <strong>of</strong> the dissolved<br />

solid at the critical moisture point <strong>and</strong> is therefore termed “critical point” [Elperin 1995; Grassmann<br />

et al. 1997]. At this point the entire surface is no longer wetted <strong>and</strong> the wetted area continually<br />

decreases in this first period <strong>of</strong> the falling rate until the surface is completely dry [Geankoplis<br />

1993]. The crust formation <strong>of</strong> an evaporating <strong>droplet</strong> with dissolved or suspended solids is generally<br />

a function <strong>of</strong> the <strong>drying</strong> conditions (temperature, humidity <strong>and</strong> relative velocity in relation to the<br />

ambient gas), the physical properties <strong>of</strong> the liquid solvent (vapour pressure <strong>and</strong> surface tension), the<br />

chemical <strong>and</strong> physical properties <strong>of</strong> the dissolved or suspended solid (solubility <strong>and</strong> surface<br />

activity) <strong>and</strong> the initial <strong>droplet</strong> size [Kastner et al. 2001; Masters 2002]. Because the evaporation<br />

principles in the second <strong>drying</strong> stage are different to the first stage, it is very important to know the<br />

critical moisture content <strong>of</strong> a solution for the calculation <strong>of</strong> <strong>drying</strong> times<br />

[Walton 1994].<br />

The second <strong>drying</strong> stage – Falling rate<br />

Discussing the second <strong>drying</strong> stage, it has to be differentiated between suspension <strong>droplet</strong>s <strong>and</strong><br />

solutions <strong>droplet</strong>s with dissolved substances. During this stage the volume <strong>of</strong> the formed particle<br />

remains constant, although there is still evaporation <strong>of</strong> liquid solvent from inside the <strong>droplet</strong><br />

[Kastner et al. 2001].<br />

When <strong>drying</strong> suspension <strong>droplet</strong>s, containing an inert solid material, the transition to the next<br />

<strong>drying</strong> stage occurs when capillary mechanisms in the pores <strong>of</strong> the arranging solid structure are<br />

insufficient to maintain completely wet surfaces [Masters 2002]. The falling-rate period begins with<br />

the “funicular state” in Figure 2.14 <strong>and</strong> continues until the liquid fills the pores no longer <strong>and</strong> the<br />

capillary flow ceases. Then vapour diffusion mechanisms dominate in the so-called “pendular state”<br />

conditions until <strong>drying</strong> is completed [Masters 2002].


CHAPTER 2 SINGLE DROPLET DRYING 39<br />

Figure 2.14: Drying phases <strong>of</strong> suspension <strong>droplet</strong>s [Masters 2002]<br />

Drying solution <strong>droplet</strong>s (Figure 2.15), the constant-rate period ceases when the <strong>droplet</strong> moisture<br />

content falls to the critical value, characterized by the initial presence <strong>of</strong> a solid phase formed at the<br />

air-liquid-interphase [Kastner et al. 2001]. Heat for the evaporation is transferred through the solid<br />

zone <strong>of</strong> vaporization, whereas vaporized solvent tends to move through the solid into the air stream.<br />

Movement <strong>of</strong> moisture from the interior to the surface becomes less <strong>and</strong> less, owing to increasing<br />

resistance to the mass transfer caused by the solid phase becoming more extensively [Masters<br />

2002]. This results in a permanently decreasing evaporation rate <strong>and</strong> in the difficulty to remove the<br />

last amounts <strong>of</strong> moisture in order to produce a particle with specified residual moisture content. For<br />

this reason the second <strong>drying</strong> stage is also named “falling rate” Figure 2.13 B-C [Walton 1994].<br />

Within this stage the rate <strong>of</strong> heat transfer exceeds the rate <strong>of</strong> mass transfer. As a result the <strong>droplet</strong><br />

starts to heat up. Sub-surface evaporation occurs if the heat transfer is sufficient high enough to<br />

cause vaporisation within the <strong>droplet</strong> [Masters 2002].<br />

The rate <strong>of</strong> mass transfer during the second <strong>drying</strong> stage is mainly controlled by the following two<br />

mechanisms [Walton 1994]:<br />

1.) The removal <strong>of</strong> solvent vapour from the solid particle surface due to external conditions like<br />

temperature, humidity, transport properties <strong>and</strong> relative velocity <strong>of</strong> the surrounding <strong>drying</strong> gas,<br />

similar to the conditions determining the first <strong>drying</strong> stage.


40 SINGLE DROPLET DRYING<br />

2.) The movement <strong>of</strong> moisture in the solid due to the internal chemical <strong>and</strong> physical nature <strong>of</strong> the<br />

solid <strong>and</strong> the moisture content.<br />

Figure 2.15: Drying stages <strong>of</strong> solution <strong>droplet</strong>s [Farid 2003]<br />

To calculate the <strong>drying</strong> time in the second stage, Ranz <strong>and</strong> Marshall proposed a relationship using<br />

the critical X critical <strong>and</strong> the final moisture content X final [Ranz 1952 b].<br />

Equation 2.73<br />

t<br />

falling rate<br />

hv<br />

⋅ ρ<br />

=<br />

particle<br />

⋅ d S ⋅ critical<br />

12 ⋅ λ ⋅ ΔT<br />

air<br />

( X − X )<br />

They also proposed an equation for the total <strong>drying</strong> time <strong>of</strong> a <strong>droplet</strong> containing dissolved solids<br />

with negligible vapour pressure lowering effect. To calculate the overall <strong>drying</strong> time the knowledge<br />

<strong>of</strong> the relationship between moisture content <strong>and</strong> <strong>droplet</strong> temperature is necessary<br />

Equation 2.74<br />

t<br />

overall<br />

hv<br />

⋅ ρliquid<br />

⋅<br />

=<br />

12 ⋅ λ ⋅<br />

air<br />

2 2 ( d − d )<br />

S 0<br />

S critical<br />

( T − T )<br />

∞<br />

S<br />

h<br />

v<br />

⋅ ρ<br />

particle<br />

⋅ d<br />

critical<br />

12 ⋅ λ<br />

air<br />

⋅<br />

critical<br />

final<br />

( X − X )<br />

The evaporation time <strong>of</strong> <strong>droplet</strong>s where vapour pressure lowering is encountered is usually longer<br />

than for pure liquid <strong>droplet</strong>s <strong>and</strong> <strong>droplet</strong>s with negligible vapour pressure lowering due to a smaller<br />

temperature driving force <strong>and</strong> therefore a decreased mass transfer. [Walton 1994]. Also according<br />

to Equation 2.74 the <strong>droplet</strong> size must be taken into account when evaluating <strong>drying</strong> times.<br />

⋅ ΔT<br />

critical<br />

final


CHAPTER 2 SINGLE DROPLET DRYING 41<br />

As already mentioned the temperature-<strong>drying</strong> time course <strong>of</strong> a <strong>droplet</strong> with solid content is<br />

different to that <strong>of</strong> pure liquid <strong>droplet</strong>s. The surface temperature <strong>of</strong> the evaporating <strong>droplet</strong> remains<br />

constant <strong>and</strong> close to the wet bulb temperature during the constant rate period, followed by a<br />

gradual increase approaching the dry bulb temperature <strong>of</strong> the air during the falling rate [Farid 2003].<br />

Figure 2.16 shows the different temperature stages in relation to Figure 2.15 during <strong>drying</strong> <strong>of</strong> a<br />

<strong>droplet</strong> containing a solid.<br />

Figure 2.16: Surface temperature-time course <strong>of</strong> a solution <strong>droplet</strong> in relation to Figure 2.15 [Farid 2003]<br />

2.4.2 The movement <strong>of</strong> moisture in solids<br />

To evaporate moisture from the exposed surface <strong>of</strong> a formed solid particle in the second <strong>drying</strong><br />

stage, the moisture must move from the depths <strong>of</strong> the particle to the surface. The mechanisms <strong>of</strong> the<br />

movement affect the quality <strong>of</strong> the <strong>drying</strong> process <strong>and</strong> the <strong>drying</strong> time. Some mechanisms advanced<br />

to explain the various types falling rate curves [Geankoplis 1993]<br />

1.) Vapour or liquid diffusion theory<br />

Diffusion <strong>of</strong> liquid moisture or vapour occurs due to a concentration gradient between the<br />

depths <strong>of</strong> the formed particle <strong>and</strong> the surface. This kind <strong>of</strong> transport is <strong>of</strong>ten found in non-<br />

porous solids, where single phase solutions are formed. The moisture diffusivity, D AB , usually<br />

decreases with decreased moisture content, so that the diffusivities are average values over the<br />

range <strong>of</strong> concentrations used [Geankoplis 1993].<br />

2.) Capillary flow<br />

In porous solids the unbound moisture moves through interconnecting pores <strong>and</strong> channels or<br />

voids <strong>of</strong> the solid by capillary action <strong>and</strong> not by diffusion. As the solvent is evaporated <strong>and</strong> the<br />

solid structure is built, a meniscus <strong>of</strong> liquid is formed across each pore in the depths <strong>of</strong> the


42 SINGLE DROPLET DRYING<br />

solid. This sets up capillary forces by interfacial tension between solvent <strong>and</strong> solid, providing<br />

the driving force for the movement <strong>of</strong> the solvent through the pores[Geankoplis 1993]. Small<br />

pores develop greater forces than large pores. The height <strong>of</strong> the meniscus inside a capillary<br />

h can be calculated by = [ ⋅γ<br />

/ ( r ⋅ g ⋅ ρ ) ]<br />

capillary<br />

2 [Frohn 2000].<br />

hcapillary capillary<br />

3.) Repeated evaporation <strong>and</strong> condensation<br />

The mechanism <strong>of</strong> evaporation <strong>and</strong> condensation is similar to the considerations <strong>of</strong> heat<br />

transfer within porous humid substances. Initially the pores are full <strong>of</strong> moisture but then<br />

gradually air pockets appear to replace the moisture lost due to evaporation. The liquid<br />

withdraws to the neck <strong>of</strong> the pores <strong>and</strong> migrates either by creeping along the capillary wall or<br />

by successive evaporation <strong>and</strong> condensation between liquid bridges. This process is also<br />

termed “liquid assisted vapour transfer” [Kneule 1975; Walton 1994].<br />

4.) Effects <strong>of</strong> shrinkage <strong>and</strong> pressure gradient<br />

Moisture flow due to morphological changes is <strong>of</strong>ten affecting the <strong>drying</strong> rate. The way the<br />

morphology is changed depends on the particle forming properties <strong>of</strong> the dissolved substance.<br />

A non-porous skin that is flexible or pliable in nature can lead to particle inflation due to<br />

internal moisture evaporation during the second <strong>drying</strong> stage. If the internal pressure is too<br />

great, even rupture <strong>of</strong> the skin can occur releasing the solvent vapour [Masters 1991]. Another<br />

possible effect is the behaviour <strong>of</strong> a non-porous <strong>and</strong> rigid skin. The particle may crack or even<br />

explode to release the internal pressure caused by the solvent vapour [Walton 1994]. To<br />

reduce these effects <strong>of</strong> particle formation it is <strong>of</strong>ten desirable to dry with moist air to decrease<br />

the <strong>drying</strong> rate resulting in a reduction <strong>of</strong> the shrinkage effects on warping or hardening<br />

[Geankoplis 1993].<br />

2.4.3 Modelling <strong>of</strong> <strong>drying</strong> <strong>of</strong> single <strong>droplet</strong>s containing dissolved or suspended solids<br />

Droplets with dissolved solids<br />

In contrast to the evaporation <strong>of</strong> pure solvent <strong>droplet</strong>s the effects <strong>of</strong> changed heat <strong>and</strong> mass transfer<br />

to <strong>and</strong> from a solution <strong>droplet</strong> in the second <strong>drying</strong> stage have to be taken into account when<br />

modelling the single <strong>droplet</strong> <strong>drying</strong> <strong>of</strong> these <strong>droplet</strong>s. Literature describes different approaches used<br />

to develop an appropriate model. Most <strong>of</strong> them are based on the absence <strong>of</strong> a temperature gradient<br />

within the <strong>droplet</strong> [Farid 2003]. Wijlhuizen et al. [1979] <strong>and</strong> Sano <strong>and</strong> Keey [Sano 1982] studied<br />

the <strong>drying</strong> <strong>of</strong> <strong>droplet</strong>s assuming a uniform temperature <strong>and</strong> diffusion <strong>of</strong> water through the solid with<br />

subsequent evaporation at the surface <strong>of</strong> the <strong>droplet</strong>. They used the diffusion equation with an


CHAPTER 2 SINGLE DROPLET DRYING 43<br />

effective diffusivity in contrast to the molar diffusivity to describe the water diffusion. Also<br />

assuming a uniform temperature the model by Nesic <strong>and</strong> Vodnik [Nesic 1991] was based on the<br />

formation <strong>of</strong> a crust with a receding crust-bulb interface. To describe the diffusion <strong>of</strong> vapour<br />

through the crust the diffusion equation with a concentration dependent diffusivity was used. The<br />

crust’s sensible heat was ignored which will have an effect on the calculated <strong>droplet</strong> temperature at<br />

the later periods <strong>of</strong> <strong>drying</strong> [Farid 2003]. Cheong, Jeffreys <strong>and</strong> Mumford [Cheong et al. 1986]<br />

proposed a receding interface model assuming a linear temperature pr<strong>of</strong>ile between the <strong>droplet</strong><br />

surface <strong>and</strong> centre. The model also ignored the sensible heating <strong>of</strong> the crust but <strong>of</strong>fered the<br />

possibility to solve the diffusion equation in the crust using experimentally evaluated physical<br />

parameters. Based on the calculation <strong>of</strong> the Biot number Bi = h ⋅ rS<br />

/ λ solid , Farid suggested that a<br />

determination <strong>of</strong> the internal temperature pr<strong>of</strong>ile is necessary to model single <strong>droplet</strong> <strong>drying</strong> within<br />

the falling rate due to a decrease <strong>of</strong> the thermal conductivity <strong>of</strong> the <strong>droplet</strong> as <strong>drying</strong> processes<br />

[Farid 2003]. His model considers the <strong>droplet</strong> shrinkage as well as the temperature distribution<br />

within the <strong>droplet</strong> using the wet bulb temperature as an input value.<br />

Droplets with suspended inert solids<br />

Similar to the solution <strong>droplet</strong>s, the <strong>drying</strong> <strong>of</strong> <strong>droplet</strong>s containing inert solids contains at least two<br />

different <strong>drying</strong> stages. Depending on the physical properties <strong>of</strong> the solid the suspended particles<br />

form a massive sphere with homogenous particle distribution or a hollow sphere with<br />

heterogeneous particle distribution. To underst<strong>and</strong> the <strong>drying</strong> process <strong>and</strong> the mechanisms <strong>of</strong><br />

particle formation it is necessary to discuss <strong>and</strong> calculate the different interface <strong>and</strong> volume forces<br />

within the suspension <strong>droplet</strong>. On the basis <strong>of</strong> the continuity equation described by Bird [1960], a<br />

model calculating the change <strong>of</strong> the dimensionless mass concentration c ˆ , = c , / c , 0 <strong>of</strong> the<br />

evaporating suspension <strong>droplet</strong> with time t <strong>and</strong> radial coordinate ξ = r / rS<br />

was set up by Kastner<br />

[2001].<br />

2<br />

dcˆ<br />

M , i dcˆ<br />

M , i ⎡2<br />

⋅ Fo ξ K ⎤ Fo⋅<br />

d cˆ<br />

L<br />

M<br />

Equation 2.75 = ⋅ ⎢ − ⋅<br />

2 2 ⎥ + 2 2<br />

dt dξ<br />

⎣ξ<br />

⋅ rˆ<br />

S 2 ⋅ rˆ<br />

S K L,<br />

0 ⎦ rˆ<br />

S ⋅ dξ<br />

K L contains all factors influencing the evaporation <strong>of</strong> the liquid component except <strong>of</strong> the <strong>droplet</strong><br />

radius r S . It can be derived directly for the model <strong>of</strong> Abramzon <strong>and</strong> Sirignano [Abramzon 1989]<br />

discussed in 2.3.2.<br />

∗<br />

Equation 2.76 K L = ρ<br />

gas ⋅ DAB<br />

⋅ ln( 1 + BM<br />

) ⋅Sh<br />

, i<br />

M i<br />

M i<br />

M i,


44 SINGLE DROPLET DRYING<br />

2.5 Particle formation, size <strong>and</strong> morphology<br />

2.5.1 Spray-dried particles<br />

The size <strong>of</strong> spray-dried particles depends on the size <strong>of</strong> the <strong>droplet</strong>s after atomization <strong>and</strong> the total<br />

solute concentration c in the solution going to be spray-dried. The relationship between these three<br />

parameters is given in Equation 2.77 [Maa et al. 1998 b].<br />

1/<br />

3<br />

⎛ c ⎞<br />

Equation 2.77 d S ( particle)<br />

= ⎜ ⎟ ⋅ d<br />

⎜ ⎟ S ( <strong>droplet</strong>)<br />

⎝ ρ particle ⎠<br />

Elverson et al. [2003] examined the correlation between <strong>droplet</strong> <strong>and</strong> particle size <strong>and</strong> showed an<br />

almost linear relationship. They also observed that the influence <strong>of</strong> the solid content in lactose<br />

solutions <strong>of</strong> 5 to 20% (w/w) resulted in only moderate size increase. The small differences in<br />

particle size implied, that particles obtained from 5% (w/w) solutions had a lower density than<br />

particles from more concentrated solutions (10% <strong>and</strong> 20% w/w). This may be an effect <strong>of</strong> loss <strong>of</strong><br />

<strong>droplet</strong>s/particles in the <strong>drying</strong> tower due to an increased loss as the feed concentration increases<br />

[Elverson et al. 2003]. Masters [1991] reported a deficiency in size difference between particles<br />

with different dry matter contents. He stated that products with film-forming materials show a<br />

significant reduction in particle size when concentrations are low.<br />

Not only the <strong>droplet</strong> size <strong>and</strong> the total solid content <strong>of</strong> the liquid feed but also external<br />

<strong>drying</strong> conditions as well as the material <strong>and</strong> formulation properties have to be considered when<br />

discussing effects on particle formation <strong>and</strong> morphology. Maa et al. [1997] carried out one <strong>of</strong> the<br />

largest studies <strong>of</strong> effects <strong>of</strong> operation <strong>and</strong> formulation variables on the morphology <strong>of</strong> spray dried<br />

<strong>protein</strong> particles. After examining effects <strong>of</strong> the outlet temperature T outlet he stated, that the degree<br />

<strong>of</strong> deformation like holes <strong>and</strong> dimples increased with increasingT outlet . Moreover, it was seen in<br />

SEM pictures that lower outlet temperatures resulted in more spherical particles. The <strong>protein</strong><br />

formulation itself also had a strong influence on particle morphology as well as possible adjuvant<br />

(e.g. sugars) <strong>and</strong> the presence <strong>of</strong> surfactants. Proteins <strong>and</strong> high molecular weight additives change<br />

the balance <strong>of</strong> surface-to-viscous forces <strong>and</strong> result in various particle features [Adler et al. 2000;<br />

Alex<strong>and</strong>er 1978]. Sugar added <strong>protein</strong> <strong>formulations</strong> showed different behaviour. Lactose leads to<br />

spray-dried particles with deep holes or donut-like shapes. With mannitol even a <strong>drying</strong>-timedependent<br />

crystallization occurred [Maa et al. 1997]. Surfactant molecules also may modify the<br />

properties <strong>of</strong> the dry crust by reducing the tendency <strong>of</strong> the spray-dried particles to collapse during


CHAPTER 2 SINGLE DROPLET DRYING 45<br />

the falling-rate period. SEM pictures demonstrated a change in particle shape <strong>and</strong> morphology after<br />

spray-<strong>drying</strong> a <strong>protein</strong>/sugar/surfactant formulation compared to pure <strong>protein</strong> or <strong>protein</strong>/sugar<br />

compositions [Adler et al. 2000; Maa et al. 1997].<br />

If the <strong>drying</strong> air temperature in an evaporation process is above the boiling point <strong>of</strong> the<br />

<strong>droplet</strong> solution <strong>and</strong> the liquid trapped inside the <strong>droplet</strong> crust reaches the boiling point, vapour is<br />

produced causing pressure to begin to build up within the <strong>droplet</strong>. The effect the pressure has on the<br />

subsequent particle formation during the constant rate depends on the physical properties <strong>of</strong> the<br />

solid [Masters 1991]. If the layer represents a porous crust, the vapour inside will be released<br />

through the pores. If the solid forms a non-porous plastic (impervious) film, the <strong>droplet</strong> may<br />

exp<strong>and</strong>, collapse, rupture or even disintegrate. However, <strong>droplet</strong> temperatures do not <strong>of</strong>ten reach<br />

boiling-point levels <strong>of</strong> the solutions even in spray-dryers at highT inlet [Masters 2002]. Most <strong>of</strong> the<br />

products show different <strong>drying</strong> characteristics within the falling-rate. Droplets with an impervious<br />

surface at the critical moisture content result in a sharp fall <strong>of</strong> the evaporation rate in the second<br />

<strong>drying</strong> stage. Whereas solution <strong>droplet</strong>s forming a highly porous surface layer lead to an only<br />

gradually falling <strong>drying</strong> rate because the vapour continues to migrate easily through the <strong>droplet</strong>-air<br />

interface [Masters 2002]. Figure 2.17 <strong>and</strong> Table 2.3 show shapes <strong>and</strong> factors influencing the final<br />

particle shape <strong>of</strong> spray-dried particles<br />

Table 2.3: Factor influencing the final particle shape [Masters 2002]<br />

Surface characteristic<br />

created during initial<br />

<strong>drying</strong> phase (1)<br />

Temperature<br />

environment<br />

Effect on <strong>droplet</strong> during<br />

<strong>drying</strong><br />

Plastic, porous Non plastic, porous Plastic, impervious<br />

Temperature distribution under co-current flow <strong>drying</strong> mode<br />

Shrinkage<br />

Final (2) particle shape Spherical Spherical<br />

Near original size<br />

(Reduction through volatile<br />

loss)<br />

Expansion,<br />

bursting,<br />

collapse on cooling<br />

Misshapen,<br />

fragmented<br />

(1) Dependent upon product formulation, presence <strong>of</strong> additives<br />

(2) Particles subject to size enlargement by agglomeration, self adhesion effects, <strong>and</strong> reduction<br />

through mechanical h<strong>and</strong>ling


46 SINGLE DROPLET DRYING<br />

2.5.2 <strong>Single</strong> <strong>droplet</strong> <strong>drying</strong> experiments<br />

Figure 2.17: Particle<br />

shapes formed during the<br />

spray-<strong>drying</strong> process:<br />

Phase 1: atomisation <strong>and</strong><br />

contact with hot air; phase<br />

2: crust formation <strong>and</strong><br />

beginning <strong>of</strong> the falling<br />

rate period; phase 3:<br />

various shapes <strong>and</strong><br />

structures [Masters 2002].<br />

(1) solid, spherical;<br />

(2) shrivelled, hollow;<br />

(3) hollow, spherical;<br />

(4) cenospherical;<br />

(5) disintegrated.<br />

The photographs show<br />

from top down:<br />

(a) solid, spherical;<br />

(b) misshapen;<br />

(c) cenospherical;<br />

(d) agglomerated;<br />

(e) granulated<br />

<strong>Single</strong> <strong>droplet</strong> <strong>drying</strong> experiments have been used in the study <strong>of</strong> particle morphology in spray<br />

<strong>drying</strong> process since in 1952 Ranz <strong>and</strong> Marshall carried out one <strong>of</strong> the first examinations<br />

[Ranz 1952 a; Ranz 1952 b]. Due to the larger diameters than spray <strong>droplet</strong>s <strong>and</strong> the not allowed<br />

free rotation while <strong>drying</strong>, single <strong>droplet</strong> <strong>drying</strong> methods have been criticised for being unrealistic.<br />

Other reasons are that suspension devices provide as a heat source <strong>and</strong> a site for vapour bubble<br />

nucleation within the drop <strong>and</strong> may act to deform the <strong>droplet</strong>s in later <strong>drying</strong> process. Despite these<br />

disadvantages, the larger <strong>droplet</strong>s <strong>of</strong> single <strong>droplet</strong> <strong>drying</strong> studies are the easy to observe, record<br />

<strong>and</strong> control [Lin 2003]. Table 2.4 gives a brief review on single <strong>droplet</strong> <strong>drying</strong> experiments to study<br />

particle formation <strong>and</strong> morphology.<br />

Charlesworth <strong>and</strong> Marshall were the first who expressed the <strong>drying</strong> behaviour <strong>of</strong> the<br />

<strong>droplet</strong>s in a form <strong>of</strong> a generalized sequence <strong>of</strong> morphological events, leading to classification <strong>of</strong><br />

the final particle morphology. They interpreted the data from weight loss <strong>and</strong> temperature versus<br />

time curves <strong>and</strong> suggested that at the beginning <strong>of</strong> <strong>drying</strong> the temperature was slightly above the<br />

wet bulb temperature <strong>of</strong> the pure solvent due to the vapour pressure lowering effect <strong>of</strong> the dissolved<br />

solid. The temperature rose with the first appearance <strong>of</strong> a solid phase <strong>and</strong> increased rapidly as the<br />

crust formation neared completion [Charlesworth 1960; Walton 1994].


CHAPTER 2 SINGLE DROPLET DRYING 47<br />

Table 2.4: Summary <strong>of</strong> important studies on <strong>drying</strong> kinetics <strong>and</strong> particles morphology <strong>of</strong> single <strong>droplet</strong>s [Lin 2003;<br />

Walton 1994]<br />

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

literature<br />

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

[Ranz 1952 a;<br />

Ranz 1952 b]<br />

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

Marshall<br />

[Charlesworth 1960]<br />

Futura et al.<br />

[Futura et al. 1982]<br />

El-Sayed et al.<br />

[El-Sayed et al. 1990]<br />

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

[Sunkel 1993]<br />

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

[Walton 1999]<br />

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

[Lin 2003]<br />

Suspending method<br />

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

measurement<br />

Glass capillary 80µm;<br />

manganin-constantan<br />

thermocouple<br />

340µm glass filament,<br />

<strong>droplet</strong>s formed from a<br />

microburette;<br />

thermoelement manganinconstantan<br />

wires connected<br />

to a recording<br />

potentiometer<br />

Steady state acoustic<br />

pressure gradient held a<br />

<strong>droplet</strong> in space against<br />

gravitation;<br />

Thermocouple<br />

Annealed glass with a glass<br />

bead d=100-300µm on the<br />

tip;<br />

0.12mm Type-Ethermocouple<br />

Flexible silica capillary<br />

hollow fibre;<br />

type-K thermocouple<br />

Rotating glass filament in a<br />

wind tunnel;<br />

thermocouple<br />

Capillary filament inside a<br />

drop holder;<br />

Thermocouple<br />

Test material Initial size Heating media<br />

(flow rate)<br />

Sodium chloride,<br />

ammonium nitrate<br />

Sodium sulphate,<br />

potassium sulphate,<br />

copper sulphate,<br />

calcium chloride,<br />

sodium acetate,<br />

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

Ammonium chloride,<br />

ammonium nitrate,<br />

ammonium sulphate<br />

Sucrose,<br />

maltodextrine,<br />

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

skim milk<br />

C<strong>of</strong>fee extract,<br />

maltodextrine,<br />

non-fat milk<br />

Different inorganic <strong>and</strong><br />

organic salts,<br />

gelatine,<br />

semi-instant skimmed<br />

milk,<br />

co-dried egg <strong>and</strong><br />

skimmed milk powder,<br />

anionic detergents<br />

Calcium acetate,<br />

sodium acetate,<br />

potassium carbonate,<br />

sodium chloride,<br />

ammonium chloride,<br />

lithium manganous<br />

nitrate,<br />

barium alumino boro<br />

silicate BABS<br />

0.6-1.1 mm<br />

Air at 85-22°C<br />

<strong>and</strong> 300cm/s<br />

1.3-1.8 mm Air at 31-159°C<br />

<strong>and</strong> 39-157cm/s<br />

0.9-1.0 mm Air at 65-95°C at<br />

0.6-1.9m/s<br />

1.4-1.5mm Air at 25-250°C<br />

<strong>and</strong> 10cm/s<br />

0.5-1.8mm Nitrogen at 18°C<br />

<strong>and</strong> 1.0l/min<br />

1.0-2.0mm Air at 70-200°C<br />

<strong>and</strong> 1.0m/s<br />

1.2-1.3mm Compressed air at<br />

60-170°C with<br />

Re=4.9-5.2<br />

Based on the idea <strong>of</strong> a classification system, Walton <strong>and</strong> Mumford carried out one <strong>of</strong> the most<br />

extensive studies on particle morphology [Walton 1999]. They produced particles in a convective


48 SINGLE DROPLET DRYING<br />

<strong>drying</strong> process analogous to spray <strong>drying</strong> using a rotating glass filament in a kind <strong>of</strong> wind tunnel.<br />

Different solid or mixtures <strong>of</strong> solid were dried from solutions, slurries or pastes as single <strong>droplet</strong>s.<br />

Their results were related to the experimental conditions like air temperature, initial solid content<br />

<strong>and</strong> the degree <strong>of</strong> feed aeration <strong>and</strong> to the chemical, physical <strong>and</strong> crust forming properties <strong>of</strong> the<br />

solid. Walton’s <strong>and</strong> Mumford`s studies demonstrated that all <strong>of</strong> their particle morphologies could be<br />

categorized in skin-forming, crystalline or agglomerate materials. The term “skin” must be regarded<br />

as a generalisation used to describe the particle surface structure built <strong>of</strong> polymeric (amorphous)<br />

<strong>and</strong> sub-microcrystalline materials. They showed that both chemical <strong>and</strong> physical nature <strong>of</strong> the<br />

material is important to determine the <strong>drying</strong> behaviour. Regarding external conditions, an increase<br />

in <strong>drying</strong> temperature from 70°C to 200°C resulted in an increased rate <strong>of</strong> heat <strong>and</strong> mass transfer<br />

<strong>and</strong> shorter <strong>drying</strong> times. Due to the more violent <strong>drying</strong> conditions the particles showed a greater<br />

tendency to inflate, shrivel <strong>and</strong> in some cases explode [Walton 1999]. One <strong>of</strong> the latest<br />

investigations <strong>of</strong> <strong>drying</strong> behaviour <strong>and</strong> particles morphology <strong>of</strong> single <strong>droplet</strong>s was done by Lin <strong>and</strong><br />

Gentry [Lin 2003]. Their experiments resulted in the summary that low <strong>drying</strong> temperature <strong>and</strong><br />

material with high latent heat <strong>of</strong> crystallisation, for example endothermic crystallisation, is<br />

favourable for small, dense <strong>and</strong> regularly-shaped particles. Materials forming elastic shell structures<br />

such as ammonium chloride led to hollow particles <strong>and</strong> materials with high solubility to small,<br />

dense <strong>and</strong> irregularly-shaped particles. They also stated that higher initial solute concentration was<br />

favourable for the formation <strong>of</strong> dense particles. Table 2.5 will give an overview on the external <strong>and</strong><br />

internal conditions conducive to form solid or hollow particles [Lin 2003].<br />

Table 2.5: External <strong>and</strong> internal conditions for the formation <strong>of</strong> hollow <strong>and</strong> massive particles [Lin 2003].<br />

Conditions conducive to solid particle formation Conditions conducive to hollow particle formation<br />

- Nuclei-free environment [Leong 1981];<br />

- High solubility solute [Leong 1981];<br />

- Low initial relative saturation <strong>droplet</strong> [Lin 2003]<br />

- low <strong>drying</strong> rates [Lin 2003]<br />

- smaller <strong>droplet</strong> [Leong 1987]<br />

- higher viscosity [Zhang et al. 191]<br />

- lower diffusivity in solvent [El-Sayed et al. 1990]<br />

- Rapid <strong>drying</strong> <strong>of</strong> solution <strong>droplet</strong><br />

[Leong 1981; Leong 1987]<br />

- Decomposition <strong>of</strong> thermolabile particles involving a<br />

gaseous component [Roth 1988]


CHAPTER 3 ACOUSTIC LEVITATION 49<br />

3 Acoustic Levitation<br />

3.1 Basic principles <strong>of</strong> acoustic levitation<br />

3.1.1 Application <strong>of</strong> acoustic levitation<br />

The phenomenon <strong>of</strong> contactless positioning <strong>of</strong> small liquid or solid samples in the pressure nodes <strong>of</strong><br />

acoustic st<strong>and</strong>ing waves is known from the “Kundt’s tube” experiment in acoustics <strong>and</strong> is<br />

commonly termed acoustic or ultrasonic levitation. The gravity acting on the sample is compensated<br />

by the forces <strong>of</strong> the ultrasonic field. The first description <strong>of</strong> acoustic levitation was published in<br />

1934 by King <strong>and</strong> later in 1962 by Gorkh<strong>of</strong>f [Lierke 2002; Yarin et al. 1998] but further technical<br />

development was enforced in the 1970s by the American <strong>and</strong> European space agencies. They were<br />

interested in a reliable space laboratory tool for containerless processing under microgravity<br />

conditions to carry out experiments in material science, physics <strong>and</strong> biology [Lierke 2002; Trinh<br />

1985]. Today acoustic levitation is also available to terrestrial laboratory research under normal<br />

gravity conditions. A summary <strong>of</strong> possible applications <strong>of</strong> acoustic levitation was given by Lierke<br />

[Lierke 2002]. Table 3.1 presents already published applications apart from experiments <strong>of</strong> single<br />

<strong>droplet</strong> <strong>drying</strong> kinetics <strong>and</strong> particle formation. In the 1960s Burdukov <strong>and</strong> Nakoryakov studied the<br />

influence <strong>of</strong> the ultrasonic field on the evaporation <strong>and</strong> mass transfer <strong>of</strong> naphthalene spheres<br />

[Burdukov 1963]. They calculated the acoustic streaming near a small rigid sphere <strong>and</strong> the mass<br />

transfer rate at the surface <strong>of</strong> the sphere positioned in a st<strong>and</strong>ing plane sound wave. Seaver,<br />

Galloway <strong>and</strong> Mannucia [Seaver et al. 1989; Seaver et al. 1990] described the evaporation <strong>of</strong><br />

volatile non-ideal liquid mixtures, the condensation <strong>of</strong> water vapour onto evaporating drops <strong>of</strong> 1butanol<br />

<strong>and</strong> the remote thermometry <strong>of</strong> water drops using laser-induced fluorescence using an<br />

acoustic levitator in a free-jet wind tunnel. They levitated liquid <strong>droplet</strong> in a range from 150µm to<br />

3mm in laminar air streams with velocities from 25 to 350 cm/s. Their evaporation measurements <strong>of</strong><br />

water <strong>droplet</strong>s agreed exactly to the calculated predictions using Equation 2.89 together with a<br />

correlation to take the effects <strong>of</strong> mass transfer by forced convection into account. Gopinath <strong>and</strong><br />

Mills [Gopinath 1993] examined the convective heat transfer from a sphere due to the acoustic<br />

streaming <strong>of</strong> an ultrasonic levitator for large Reynolds numbers. The results obtained were<br />

important for thermal analysis <strong>of</strong> containerless processing in space. They used analytical <strong>and</strong><br />

numerical techniques to obtain Nusselt number correlations for a wide range <strong>of</strong> Pr<strong>and</strong>tl numbers.<br />

Tian <strong>and</strong> Apfel [Tian 1996] introduced a new multiple <strong>droplet</strong> electro-acoustic levitation system to<br />

study the various phenomena associated with <strong>droplet</strong> array <strong>and</strong> the influence <strong>of</strong> the ultrasonic field


50 ACOUSTIC LEVITATION<br />

on the evaporation <strong>of</strong> single <strong>droplet</strong>s. In their experiments they showed that the perturbation <strong>of</strong> the<br />

acoustic field to the <strong>droplet</strong> evaporation process is negligible. Yarin, Brenn, Tropea et al. disagreed<br />

with the statement <strong>of</strong> Tian <strong>and</strong> Apfel. They showed an influence <strong>of</strong> the acoustic streaming on<br />

Sherwood number <strong>and</strong> mass transfer, as well as an enrichment <strong>of</strong> gas at the outer boundary <strong>of</strong> the<br />

acoustic boundary layer by liquid vapour [Yarin et al. 1999]. A more detailed view at these<br />

influencing factors on the evaporation kinetics <strong>of</strong> levitated <strong>droplet</strong>s will be discussed in chapter 3.2.<br />

The opposite effects <strong>of</strong> increased mass transfer due to the acoustic streaming <strong>and</strong> decreased<br />

evaporation rate due to the enrichment <strong>of</strong> liquid vapour around the <strong>droplet</strong> surface compensate each<br />

other. This may lead to a statement <strong>of</strong> Tian <strong>and</strong> Apfel regarding the evaporation process equal to<br />

that <strong>of</strong> undisturbed <strong>droplet</strong> [Kastner 2001]. Kastner <strong>and</strong> Brenn investigated the influence <strong>of</strong><br />

evaporating <strong>droplet</strong>s on the acoustic field <strong>and</strong> vice versa. He examined single <strong>droplet</strong> <strong>drying</strong> <strong>of</strong><br />

solvents <strong>and</strong> suspension <strong>droplet</strong>s under various <strong>drying</strong> conditions using an acoustic tube levitator.<br />

His experimental data was used to verify a model describing the evaporation process <strong>of</strong> binary<br />

mixtures <strong>of</strong> solvents as well as <strong>of</strong> suspension <strong>droplet</strong>s [Kastner 2001; Kastner et al. 2001; Yarin et<br />

al. 2002a]. Further work with same system was done by Yarin, Brenn <strong>and</strong> Rensink to investigate the<br />

effect <strong>of</strong> <strong>droplet</strong> surface oscillations on the evaporation rate <strong>of</strong> pure solvents <strong>and</strong> binary solvent<br />

mixtures [Rensink 2004; Yarin et al 2002 b]. They showed an increased evaporation rate <strong>of</strong> solvent<br />

as well as <strong>of</strong> suspension <strong>droplet</strong> due to oscillations <strong>of</strong> the <strong>droplet</strong> surface. Tuckermann, Bauerecker<br />

<strong>and</strong> Neidhardt used an open levitation system at ambient conditions to look at the evaporation rates<br />

<strong>of</strong> alkanes <strong>and</strong> alkanols. They additionally installed an IR-thermography system to look at the<br />

course <strong>of</strong> the surface temperature <strong>of</strong> the evaporating <strong>droplet</strong>. They used the experimental data to<br />

determine <strong>and</strong> to verify binary diffusion coefficients form literature [Tuckermann et al. 2002 b].<br />

Groenewold et al. also determined the evaporation rates <strong>of</strong> single suspension <strong>droplet</strong>s with γ-Al2O3<br />

within the two <strong>drying</strong> stages. They used an acoustic tube levitator in combination with a dew point<br />

hygrometer to measure the moisture content <strong>of</strong> the outlet air stream [Groenewold et al. 2002; Möser<br />

et al. 2001].


CHAPTER 3 ACOUSTIC LEVITATION 51<br />

Table 3.1: Experimental research in material science, physics <strong>and</strong> biotechnology using an acoustic levitation system<br />

Author <strong>and</strong> literature Research area Details<br />

Trinh <strong>and</strong> Hsu<br />

[Trinh 1986]<br />

Seaver <strong>and</strong> Frost<br />

[Frost 1993]<br />

Tuckermann<br />

[Tuckermann 2002]<br />

Welter <strong>and</strong> Neidhart<br />

[Welter 1997];<br />

Eberhardt <strong>and</strong> Neidhart<br />

[Eberhardt 1999];<br />

Rohling et al.<br />

[Rohling et al. 2000]<br />

Ohsaka et al.<br />

[Ohsaka et al. 2002]<br />

Ohsaka et al.<br />

[Ohsaka et al. 2003]<br />

Kavouras <strong>and</strong> Krammer<br />

[Kavouras 2003]<br />

Sacher <strong>and</strong> Krammer<br />

[Sacher 2005]<br />

Weis <strong>and</strong> Nardozzi<br />

[Weis 2005]<br />

3.1.2 Fundamentals <strong>of</strong> acoustics<br />

Density measurements <strong>of</strong> liquids <strong>and</strong><br />

low density solid metals<br />

Surface tension <strong>of</strong> one-octa-decanol<br />

monolayer<br />

Surface tension <strong>of</strong> surfactantmonolayers<br />

Micro <strong>and</strong> trace analysis in<br />

acoustically levitated <strong>droplet</strong>s<br />

Viscosity measurements <strong>of</strong> highly<br />

viscous, particularly under-cooled<br />

liquids<br />

Thermal diffusion coefficients <strong>of</strong><br />

under-cooled liquids<br />

Gas/solid-reactions <strong>of</strong> Ca(0H)2 <strong>and</strong><br />

humid, HCl-gas<br />

Crystallisation <strong>of</strong> CaCO3<br />

Enzyme kinetics <strong>of</strong> alkaline<br />

phosphatase-catalyzed hydrolysis <strong>of</strong> 4methylum-belliferone<br />

phosphate<br />

Use <strong>of</strong> two different techniques based on the static<br />

equilibrium position <strong>of</strong> levitated samples <strong>and</strong> on<br />

the dynamic interaction <strong>of</strong> a levitated sample <strong>and</strong><br />

the acoustic field<br />

Formation <strong>of</strong> liquid-condensed phase <strong>of</strong> oneoctadecanol<br />

thin films on the surface <strong>of</strong> levitated<br />

drops<br />

Investigation <strong>of</strong> the liquid-condensed state <strong>of</strong><br />

surfactant monolayers <strong>of</strong> levitated <strong>droplet</strong>s using<br />

an infra-red camera<br />

Levitation <strong>of</strong> analytes containing liquid <strong>droplet</strong>s<br />

in combination with absorption <strong>and</strong> fluorescence<br />

measurements for acid-base titrations with<br />

absorption <strong>and</strong> fluorescence indicator.<br />

Levitation <strong>of</strong> liquid samples <strong>and</strong> elongation <strong>of</strong><br />

drops by rotating them beyond the point <strong>of</strong><br />

bifurcation. After the drop was allowed to restore<br />

by surface tension driven relaxation, the viscosity<br />

was determined via a relaxation model.<br />

Levitation <strong>of</strong> laser-heated <strong>droplet</strong> with subsequent<br />

natural cooling by heat loss from the surface.<br />

Infra-red camera measurement <strong>of</strong> the cooling rate<br />

in combination with a radial heat conduction<br />

model enables the calculation <strong>of</strong> the diffusion<br />

coefficient.<br />

Levitation <strong>of</strong> single Ca(OH)2 particles <strong>and</strong><br />

exposition to a defined gas atmosphere.<br />

Measurement <strong>of</strong> the particle weight change to<br />

calculate SO2 <strong>and</strong> HCl absorption.<br />

Investigation <strong>of</strong> the influence <strong>of</strong> temperature,<br />

resident time <strong>and</strong> specific forces on the<br />

crystallisation <strong>of</strong> CaCO3 containing levitated<br />

<strong>droplet</strong>s.<br />

Measurement <strong>of</strong> the rate <strong>of</strong> product formation <strong>of</strong><br />

alkaline phosphatase-catalyzed hydrolysis <strong>of</strong> 4methylum-belliferone<br />

phosphate in super-cooled<br />

levitated <strong>droplet</strong>s.<br />

The velocity <strong>of</strong> sound waves depends on the properties <strong>of</strong> the material or medium through which<br />

the sound waves pass. In a liquid or gas it can be described by the following equation


52 ACOUSTIC LEVITATION<br />

Equation 3.1<br />

ν =<br />

0<br />

EBM<br />

ρ<br />

ρ is the density <strong>of</strong> the liquid <strong>of</strong> gaseous material <strong>of</strong> medium <strong>and</strong> E BM represents the bulk modulus<br />

<strong>of</strong> elasticity. The change <strong>of</strong> sound velocity with the absolute temperature T , the most important<br />

factor, is given by<br />

Equation 3.2<br />

ν 0 ( air)<br />

⋅ R ⋅T<br />

=<br />

M<br />

α<br />

(air)<br />

R is the universal gas constant, M (air) the molecular weight <strong>of</strong> air <strong>and</strong> α the adiabatic index,<br />

which is 1.402 for air [Tipler 2000]. An acoustic field is usually characterized by the gas particle<br />

velocity <strong>and</strong> the sound pressure. The gas particle velocity or short particle velocity B is the real or<br />

imagined velocity <strong>of</strong> a particle in a medium on a longitudinal wave <strong>of</strong> pressure. It can be calculated<br />

as the product <strong>of</strong> particle displacement ξ <strong>and</strong> angular frequency ω .<br />

Equation 3.3 B = ξ ⋅ω<br />

= ξ ⋅ 2 π ⋅ f<br />

The particle velocity level, sometimes called sound velocity level (SVL), gives the ratio <strong>of</strong> a<br />

particle velocity 1 B to the reference particle velocity B ref .<br />

Equation 3.4<br />

= 20 ⋅ log<br />

B<br />

→<br />

1<br />

Lυ 10<br />

ref<br />

Bref<br />

B<br />

= 5⋅10<br />

The unit <strong>of</strong> the SVL is named decibel (dB) <strong>and</strong> is a dimensionless unit expressing the ratio between<br />

two quantities.<br />

The sound pressure, p sound , is the force per area in N/m 2 <strong>of</strong> the root mean square pressure<br />

deviation caused by a sound wave passing through a fixed point. It is calculated as the product <strong>of</strong><br />

medium density ρ , speed <strong>of</strong> sound ν 0 <strong>and</strong> particle velocity B<br />

Equation 3.5 ρ ⋅ν<br />

⋅ B<br />

p sound<br />

= 0<br />

The logarithmic measure <strong>of</strong> the sound pressure related to the pressure <strong>of</strong> a reference noise is called<br />

sound pressure level (SPL) or sound level L P .<br />

−8<br />

m<br />

s


CHAPTER 3 ACOUSTIC LEVITATION 53<br />

1<br />

psound<br />

−5<br />

N<br />

Equation 3.6 Lp = 20 ⋅ log10<br />

→ pref<br />

= 2 ⋅10<br />

2<br />

p<br />

m<br />

ref<br />

Two other st<strong>and</strong>ardized values are important when dealing with acoustics. The sound or acoustic<br />

intensity I <strong>and</strong> the sound or acoustic power W . The sound intensity is calculated as the product <strong>of</strong><br />

sound pressure p sound <strong>and</strong> particle velocity B . The logarithmic measure <strong>of</strong> the sound intensity in<br />

relation to the st<strong>and</strong>ard reference sound intensity is called sound intensity level (SIL) L I <strong>and</strong> given<br />

by<br />

I<br />

−12<br />

W<br />

Equation 3.7 LI = 10 ⋅ log10<br />

→ I ref = 1⋅10<br />

2<br />

I<br />

m<br />

ref<br />

The power is a measure <strong>of</strong> sonic energy per time. It is the product <strong>of</strong> sound intensity I <strong>and</strong> area A .<br />

According to all other values before, the sound power level (SWL) L W is the logarithmic measure<br />

<strong>of</strong> sound power in relation to a st<strong>and</strong>ard reference sound power Wref <strong>and</strong> calculated by<br />

W<br />

−12<br />

Equation 3.8 LW = 10 ⋅ log10<br />

→ Wref<br />

= 1⋅10<br />

W<br />

W<br />

3.1.3 Forces <strong>of</strong> a st<strong>and</strong>ing acoustic wave<br />

ref<br />

The concept <strong>of</strong> st<strong>and</strong>ing waves depends on the reflection <strong>of</strong> sound waves <strong>and</strong> is known from<br />

Kundt’s tube experiment shown in Figure 3.1. The length <strong>of</strong> the tube has to be a integer multiple <strong>of</strong><br />

the half wavelength to create a st<strong>and</strong>ing acoustic wave inside [Tipler 2000].<br />

λn<br />

Equation 3.9 l = n ⋅ n = 1,<br />

2,<br />

3,...<br />

2<br />

According to Equation 3.9 the relation between frequency <strong>of</strong> the ultrasonic transducer <strong>and</strong> the<br />

length <strong>of</strong> the tube can be calculated by [Tipler 2000]<br />

ν<br />

0 ν 0<br />

Equation 3.10 f n = = n ⋅ n = 1,<br />

2,<br />

3,...<br />

λ 2 ⋅ l<br />

n


54 ACOUSTIC LEVITATION<br />

Figure 3.1: Kundt’s tube: apparatus consisting <strong>of</strong> a glass tube supported on a metal base. The clamp on the left side<br />

holds a metal rod with a metal disk attached to one end. Rod <strong>and</strong> disk extend inside the glass tube, whose position can<br />

be adjusted to centre the tube about the disk. The disk must not touch the glass because the vibrations set up in the rod<br />

will break the glass. The tube is closed by a stopper on the other end [Nave 2005]<br />

In the case <strong>of</strong> an acoustic levitator, whose basic<br />

set-up can be seen in Figure 3.2, a st<strong>and</strong>ing wave<br />

is formed between an ultrasonic transducer (e.g. a<br />

piezocrystal) producing a sound wave with the<br />

wavelength λ <strong>and</strong> a transducer at the distance<br />

LR = n ⋅ λ / 2 ( n = 1,<br />

2,<br />

3,...<br />

) . Assuming a one-<br />

dimensional ultrasound wave inside the levitation<br />

system, the one–dimensional wave equation with<br />

appropriate boundary conditions at the sound<br />

source ( x = 0)<br />

<strong>and</strong> the reflector ( x = L ) can be<br />

used for mathematical description according to<br />

Yarin et al. [1998].<br />

2<br />

2<br />

∂ p′ i 2 ∂ p′<br />

i<br />

Equation 3.11 = c 2 0 ⋅ 2<br />

∂t<br />

∂x<br />

x<br />

R<br />

= 0 : pi<br />

= A0<br />

x = L<br />

R<br />

′<br />

⋅ e<br />

∂p′<br />

i : = 0<br />

∂x<br />

−iωt<br />

transducer<br />

reflector<br />

sound particle<br />

velocity<br />

sample<br />

sound pressure<br />

Figure 3.2: Axial sound pressure <strong>and</strong> particle velocity<br />

distribution <strong>of</strong> a st<strong>and</strong>ing acoustic wave leading to<br />

pressure nodes <strong>and</strong> antinodes. Liquid <strong>and</strong> solid samples<br />

can be levitated within the pressure nodes [Tuckermann<br />

et al. 2001; Yarin et al. 1998]


CHAPTER 3 ACOUSTIC LEVITATION 55<br />

p′<br />

i is the pressure perturbation in the incident wave, A 0 its amplitude at the source surface, ω the<br />

angular frequency corresponding to the ultrasonic range, c 0 the speed <strong>of</strong> sound in air, t the time, x<br />

the vertical coordinate <strong>and</strong> i the imaginary unit. The axial position <strong>of</strong> the pressure nodes x within<br />

the st<strong>and</strong>ing acoustic wave can be calculated by<br />

ν 0 ⎛ π ⎞<br />

Equation 3.12 x = LR<br />

− ⋅ ⎜ + π n⎟<br />

n = 0, 1, 2, 3, ..., n<br />

ω ⎝ 2 ⎠<br />

The position <strong>of</strong> the maxima <strong>and</strong> minima can be determined by<br />

ν 0<br />

Equation 3.13 xmax = LR<br />

− ⋅ π n n = 1, 3, 5,...,<br />

n<br />

ω<br />

ν 0<br />

xmin = LR<br />

− ⋅ π n n = 0,<br />

2,<br />

4,...,<br />

n −1<br />

ω<br />

The approximate solution <strong>of</strong> Equation 3.12 for the incident wave in an infinite levitator is given by<br />

Yarin et al. [1998].<br />

−iωt<br />

ω<br />

Equation 3.14 p′<br />

i = A0e<br />

⋅ e ⋅ cos x<br />

ν<br />

A0 e is the effective amplitude <strong>of</strong> the acoustic field <strong>and</strong> is calculated by<br />

Equation 3.15<br />

A0<br />

A0e = −<br />

cos ω R ν 0<br />

0<br />

( L / )<br />

Yarin et. al [1998] as well as by Kastner [2001] state that the approximation (Equation 3.14 <strong>and</strong><br />

Equation 3.15) leads to almost identical results regarding the amplitude <strong>of</strong> the sound wave <strong>and</strong><br />

number <strong>of</strong> pressure nodes as the exact solution <strong>of</strong> Equation 3.11. Only the position <strong>of</strong> the pressure<br />

nodes differs by about 5% from the exact result. Figure 3.3 shows the sketch <strong>of</strong> the incident<br />

acoustic wave, the pressure nodes <strong>and</strong> the acoustic levitation force.


56 ACOUSTIC LEVITATION<br />

Figure 3.3: Incident acoustic wave with levitation force FL on a rigid sphere. Z is the vertical coordinate with z=0 as<br />

position <strong>of</strong> the centre <strong>of</strong> the rigid sphere <strong>and</strong> L the displacement <strong>of</strong> the centre <strong>of</strong> this sphere relative to the pressure<br />

antinode [Yarin et al. 1998]<br />

If a spherical <strong>droplet</strong> or a rigid sphere is levitated within the pressure node <strong>of</strong> a st<strong>and</strong>ing acoustic<br />

wave, a scattered acoustic field appears in addition to the incident one [Yarin et al. 1998]. Due to<br />

interaction the overall pressure perturbation at the <strong>droplet</strong> surface p′ is the amount <strong>of</strong> pressure<br />

perturbation <strong>of</strong> the incident wave p′ i plus the scattered acoustic wave p′ S [Kastner 2001]. The<br />

scattered acoustic pressure is given by<br />

−iωt Equation 3.16 ′ = A ⋅ e ⋅ p ( r)<br />

pS 0 e<br />

S<br />

where the function p S <strong>of</strong> the radius-vector r is found from the Helmholtz equation [Yarin et al.<br />

1998].<br />

⎛ ω ⎞<br />

Equation 3.17 Δ p ⎟ ′<br />

S + ⎜ pS<br />

= 0<br />

⎝ν<br />

0 ⎠<br />

2<br />

The mathematical solution <strong>of</strong> Equation 3.17 can be seen from King [1934] or Yarin et al. [1998].<br />

Using the equation for the acoustic radiation pressure given by L<strong>and</strong>au und Lifshitz [Kastner 2001]


CHAPTER 3 ACOUSTIC LEVITATION 57<br />

Equation 3.18<br />

p<br />

a<br />

2<br />

p′<br />

ρ<br />

= −<br />

2 ⋅ ρ ⋅ν<br />

2<br />

gas<br />

2<br />

0<br />

gas<br />

ν ′<br />

⎟ 2<br />

2 A ⎛ ⎞<br />

0 2 ω<br />

p′<br />

= ⋅ cos ⎜ x<br />

2 ⎝ν<br />

0 ⎠<br />

⎟ 2<br />

2 A0<br />

2⎛<br />

ω ⎞<br />

ν ′ = ⋅ sin ⎜ x<br />

2 2<br />

2 ⋅ ρ gas ⋅ν<br />

0 ⎝ν<br />

0 ⎠<br />

2<br />

ν ′ is the squared acoustic gas velocity time averaged (denoted by ) over a period long<br />

enough compared to the wave cycle [Yarin et al. 1998]. The acoustic radiation pressure produced<br />

by a st<strong>and</strong>ing acoustic wave at the <strong>droplet</strong> or particle surface derived from Equation 3.14, Equation<br />

3.16 <strong>and</strong> Equation 3.18 can be used to calculate the acoustic levitation force according to King<br />

[1934] <strong>and</strong> Yarin et al. [1998].<br />

⎛ A ⎞<br />

2<br />

Equation 3.19 F ρ r ⎜ 0e<br />

L = π gas ⋅ S ⋅ ⎟ ⋅ sin(<br />

2k<br />

Δ xD<br />

) ⋅ f ( Ω )<br />

⎜ ρ gas ν ⎟<br />

⎝ ⋅ 0 ⎠<br />

with = k0 ⋅ rS<br />

= ( ω /ν 0 ) ⋅ rS<br />

Ω where 0 = ω /ν 0<br />

2<br />

2<br />

k is the wave number <strong>and</strong> Δ xD<br />

the distance from the<br />

mass centre <strong>of</strong> the <strong>droplet</strong>s to its next upper pressure node.<br />

The function f ( Ω ) is given by<br />

( ) 1 ( Ω )<br />

⎡ 1 F0F1<br />

+ G0G<br />

Equation 3.20 f ( Ω ) = ⎢ 3 2 2<br />

⎣(<br />

Ω ) H 0 H1<br />

−<br />

+<br />

( Ω )<br />

∞<br />

∑<br />

n=<br />

2<br />

( F F + G G )<br />

⎪⎧<br />

⎨<br />

⎪⎩<br />

2 ! 2 1 2 2 ( Ω )<br />

5 2 2<br />

H1<br />

H 2<br />

2<br />

⎛ ρ<br />

− 3⋅<br />

⎜1<br />

−<br />

⎜<br />

⎝ ρ<br />

gas<br />

liquid<br />

⎞⎪⎫<br />

⎟<br />

⎬<br />

⎠⎪⎭<br />

n ( )<br />

( n + 1)<br />

( Fn+<br />

1Fn<br />

+ Gn+<br />

1Gn<br />

) 2<br />

−1<br />

( Ω ) − n ⋅ ( n + 2)<br />

2n+<br />

3<br />

2 1<br />

( Ω<br />

) H H<br />

n+<br />

1<br />

n<br />

⎤<br />

{ }⎥⎦


58 ACOUSTIC LEVITATION<br />

The values F i , G i <strong>and</strong> H i are functions <strong>of</strong> Ω <strong>and</strong> are defined by<br />

1−i<br />

i<br />

Equation 3.21 F i ( Ω ) = ( Ω ) ni+<br />

1(<br />

Ω ) + n i i<br />

Ω<br />

( )<br />

1−i<br />

i<br />

Gi ( Ω ) ( Ω ) ji<br />

1(<br />

Ω ) − ji<br />

= +<br />

2<br />

2<br />

( ) = F ( Ω ) G ( Ω )<br />

i ( Ω )<br />

[ ] 2 / 1<br />

H i Ω i + i<br />

( )<br />

Ω<br />

( )<br />

Ω<br />

Equation 3.21 is valid for all values <strong>of</strong> Ω . The functions ( Ω )<br />

j <strong>and</strong> n ( Ω ) are the spherical Bessel-<br />

<strong>and</strong> Neumann-Functions [Kastner 2001; King 1934]. In case <strong>of</strong> Ω


CHAPTER 3 ACOUSTIC LEVITATION 59<br />

Equation 3.25<br />

A<br />

0e<br />

=<br />

4 rS<br />

⋅ ρ<br />

⋅<br />

3 f<br />

gas<br />

⋅ ρ<br />

liquid<br />

2<br />

⋅ g ⋅ν<br />

( Ω ) ⋅ sin(<br />

2k<br />

⋅ Δ x )<br />

0<br />

A slow decrease in driving voltage <strong>of</strong> the ultrasonic transducer towards the minimum driving<br />

voltage at unchanged transducer-reflector distance will lead to the positioning <strong>of</strong> the mass balance<br />

point <strong>of</strong> the <strong>droplet</strong> or particle at Δ x = π/ 4Ω<br />

. A further decrease results in a dropout <strong>of</strong> the<br />

D<br />

sample because the acoustic levitation force is now small to compensate the gravitations force <strong>of</strong><br />

the sample. At this so called dropout-point the effective amplitude A0 em can be directly calculated<br />

from the properties <strong>of</strong> the levitated liquid <strong>droplet</strong> or rigid sphere [Yarin et al. 1998].<br />

Equation 3.26<br />

A<br />

0em<br />

=<br />

4 rS<br />

⋅ ρ<br />

⋅<br />

3<br />

liquid<br />

f<br />

⋅ ρ<br />

gas<br />

( Ω )<br />

D<br />

0<br />

2<br />

⋅ g ⋅ν<br />

On the assumption that the oscillation amplitude <strong>of</strong> the transducer varies linearly to the driving<br />

voltage, which was proven by Yarin et al., the relation between both under the conditions <strong>of</strong> <strong>droplet</strong><br />

levitation ( 0 ,U 0<br />

Equation 3.27<br />

A e ) <strong>and</strong> dropout ( A0 em, U 0m<br />

) is<br />

A<br />

0e<br />

= A<br />

0em<br />

U<br />

⋅<br />

U<br />

0<br />

0m<br />

A second possibility to determine the SPL inside the ultrasonic levitation system is using the aspect<br />

ratio <strong>of</strong> a levitated <strong>droplet</strong>. The aspect ratio is the relation between the horizontal <strong>and</strong> the vertical<br />

diameter. The deformation <strong>of</strong> a <strong>droplet</strong> depends on the <strong>droplet</strong> size, the surface tension <strong>of</strong> the liquid<br />

solvent within the surrounding gas <strong>and</strong> the different values <strong>of</strong> the axial <strong>and</strong> radial levitation forces<br />

determined by the SPL . The st<strong>and</strong>ardised radial levitation force F r , also termed Bernoulli force,<br />

results from the particle velocity <strong>and</strong> sound pressure distribution at the levitation axis. It can be<br />

calculated by [Lierke 1996]<br />

Equation 3.28<br />

F<br />

r<br />

kr<br />

= 4 ⋅<br />

k<br />

⋅<br />

J1(<br />

kr<br />

r)<br />

2 ( k ⋅ r)<br />

r<br />

⎡<br />

⋅ ⎢J<br />

⎣<br />

0<br />

( k r)<br />

r<br />

0<br />

2 ⋅ J<br />

−<br />

k<br />

1<br />

r<br />

( k r)<br />

⎤ r<br />

⎥⎦<br />

k r is the radial wave number <strong>and</strong> J i the Bessel function <strong>of</strong> the order i . The ratio <strong>of</strong> the axial <strong>and</strong><br />

radial wave numbers r z k k q = / is a main design parameter <strong>of</strong> an acoustic levitation system. They<br />

are combined in the so called wave number equation [Lierke 2002]<br />

⋅ r


60 ACOUSTIC LEVITATION<br />

Equation 3.29<br />

⎛ ω ⎞<br />

⎜ ⎟<br />

⎝ ⎠<br />

2<br />

2<br />

2 2<br />

k = ⎜ ⎟ = k r + k z<br />

ν 0<br />

According to Lierke, the relation between axial <strong>and</strong> radial levitation<br />

force is 5:1 [Lierke 1996]. This means, that acoustic levitation is<br />

preferably used to compensate gravitational forces. If deformable<br />

samples are investigated, for example liquid solvents, solutions or<br />

suspensions, it leads to a deformation <strong>of</strong> the sample according to<br />

Figure 3.4. The <strong>droplet</strong> will take the shape <strong>of</strong> an oblate ellipsoid.<br />

Deformation increases with an increase <strong>of</strong> SPL until the <strong>droplet</strong><br />

disintegrates.<br />

Theoretical investigations on <strong>droplet</strong> deformation <strong>and</strong> sound<br />

pressure level were done by Marston [1980], Tian <strong>and</strong> Apfel [Tian<br />

1996], Shi <strong>and</strong> Apfel [Shi 1996] <strong>and</strong> Yarin et al. [1998]. Yarin et al.<br />

managed to determine the SPL from the aspect ratio <strong>of</strong> the levitated liquid <strong>droplet</strong> by numerical<br />

solution.<br />

Figure 3.4: Axial <strong>and</strong> radial<br />

levitation forces. The relation <strong>of</strong><br />

5:1 between axial force Fz <strong>and</strong><br />

radial force Fr leads to<br />

deformation <strong>of</strong> the <strong>droplet</strong> to an<br />

oblate ellipsoid [Kastner 2001;<br />

Tuckermann et al. 2001].


CHAPTER 3 ACOUSTIC LEVITATION 61<br />

3.2 Influences on <strong>droplet</strong>s inside an acoustic levitator<br />

3.2.1 Interactions with the acoustic field<br />

If a liquid <strong>droplet</strong> containing a pure solvent, a solution or suspension is levitated inside an acoustic<br />

levitator, it is influenced by the st<strong>and</strong>ing acoustic wave in multiple ways. One is the deformation <strong>of</strong><br />

the <strong>droplet</strong>s due to the different strength <strong>of</strong> the axial <strong>and</strong> radial levitation forces, already discussed<br />

in chapter 3.1.3. More important for determining the <strong>drying</strong> kinetics <strong>of</strong> single <strong>droplet</strong>s is the<br />

influence <strong>of</strong> the acoustic streaming field near the levitated <strong>droplet</strong> leading to an acoustic convection<br />

<strong>and</strong> <strong>of</strong> the system <strong>of</strong> the outer toroidal vortices resulting in an accumulation <strong>of</strong> solvent vapour<br />

[Rensink 2004]. Additionally, there is a heating <strong>of</strong> the sample area due to oscillations <strong>of</strong> the<br />

ultrasonic transducer changing the temperature <strong>of</strong> the ambient <strong>drying</strong> air [Kastner 2001].<br />

Furthermore, acoustic radiation pressure increases during evaporation experiments due to a volume<br />

decrease <strong>of</strong> the levitated <strong>droplet</strong> [Kastner 2001]. It is essential to look at all these interactions <strong>of</strong><br />

liquid <strong>droplet</strong> <strong>and</strong> acoustic field precisely to characterise the <strong>drying</strong> kinetics <strong>of</strong> single <strong>droplet</strong>s inside<br />

an ultrasonic field.<br />

3.2.2 Acoustic streaming<br />

The acoustic streaming around a spherical levitated liquid <strong>droplet</strong> can be divided into an inner <strong>and</strong><br />

an outer system <strong>of</strong> vortices also termed the inner <strong>and</strong> outer acoustic streaming, according to Yarin et<br />

al. [1999] <strong>and</strong> Rensink [2004]. This is shown in Figure 3.5. Kastner [2001] termed both system<br />

outer acoustic streaming <strong>and</strong> the acoustic boundary layer inner acoustic streaming. In this work the<br />

terminology <strong>of</strong> Yarin et al. <strong>and</strong> Rensink will be used. Directly at the surface <strong>of</strong> the levitated <strong>droplet</strong><br />

an acoustic boundary layer is formed due to the inner acoustic streaming shown in Figure 3.6. The<br />

radial dimension <strong>of</strong> this acoustic boundary layer was compared to the diffusion boundary layer by<br />

Lee <strong>and</strong> Wang [Lee 1990], Yarin et al [1999] <strong>and</strong> Kastner [2001] using Equation 3.30 for the inner<br />

acoustic boundary <strong>and</strong> Equation 3.31 for the diffusion layer <strong>of</strong> an evaporation <strong>droplet</strong> within an<br />

ultrasonic levitator.


62 ACOUSTIC LEVITATION<br />

Figure 3.5: Acoustic streaming field near a levitated <strong>droplet</strong> with the system <strong>of</strong> outer toroidal vortices. The inner<br />

acoustic streaming is positioned directly at the acoustic boundary layer, whereas the outer acoustic streaming (outer<br />

toroidal vortices) are emerging in space <strong>of</strong> the levitator [Yarin et al. 1999].<br />

Equation 3.30<br />

Equation 3.31<br />

δ<br />

δ<br />

IAS<br />

diff<br />

⎛ 2μ<br />

gas ⎞<br />

= ⎜ ⎟<br />

⎝ ω ⎠<br />

⎛ d S = ⎜<br />

⎝ B / D<br />

AB<br />

1/<br />

2<br />

⎞<br />

⎟ ⎟<br />

⎠<br />

⎛ μgas<br />

⎞<br />

= ⎜ ⎟<br />

⎝ π⋅<br />

f ⎠<br />

1/<br />

2<br />

Yarin et al. calculated the thickness <strong>of</strong> both boundary layers for a water <strong>droplet</strong> with a diameter<br />

−5<br />

2<br />

d = 1.<br />

0 mm in air with the kinematic viscosity μ = 1.<br />

5⋅<br />

10 m /s using an 56 kHz acoustic<br />

S<br />

levitator. Their results showed that the acoustic boundary layer δ = 9.<br />

23µm<br />

is small compared<br />

with the diffusion boundary layer δ = 92μm<br />

[Kastner 2001; Yarin et al. 1999].<br />

diff<br />

1/<br />

2<br />

gas<br />

IAS<br />

Inner acoustic streaming at<br />

the boundary <strong>of</strong> the acoustic<br />

boundary layer<br />

Outer acoustic streaming


CHAPTER 3 ACOUSTIC LEVITATION 63<br />

Figure 3.6: Acoustic streaming as well as acoustic <strong>and</strong> diffusion boundary layer over a small sphere positioned at the<br />

pressure node <strong>of</strong> a st<strong>and</strong>ing acoustic wave [Yarin et al. 1999]<br />

Influence <strong>of</strong> the inner acoustic streaming on mass transfer<br />

The inner acoustic streaming induces a flow field at the surface <strong>of</strong> the <strong>droplet</strong> which is important for<br />

all mass transfer considerations inside an acoustic levitation system. The continuity <strong>of</strong> the shear<br />

stress at the surface leads to an evacuation <strong>of</strong> liquid from the <strong>droplet</strong> also resulting in internal<br />

circulation movement. Yarin et al. calculated the velocity <strong>of</strong> the inner acoustic streaming in case <strong>of</strong><br />

a small spherical <strong>droplet</strong> with a radius r ≤ 0.<br />

137 mm much smaller than the wavelength <strong>of</strong> the<br />

st<strong>and</strong>ing acoustic wave λ 6.<br />

1mm<br />

[Yarin et al. 1999] using the approximation<br />

0 =<br />

δ diff<br />

2<br />

45 B<br />

Equation 3.32 u acoustic = ⋅ sin 2θ<br />

32 ω ⋅ r<br />

SW<br />

S<br />

δ IAS<br />

S<br />

where u acoustic is the velocity <strong>of</strong> the inner acoustic streaming averaged <strong>of</strong> multiple cycles <strong>of</strong> the<br />

st<strong>and</strong>ing acoustic wave <strong>and</strong> θ is the perimeter angle measured from the bottom point <strong>of</strong> the <strong>droplet</strong>.<br />

According to their calculations u acoustic for a water <strong>droplet</strong> at a SPL <strong>of</strong> 165 . 7 dB is<br />

0 . 93m/s<br />

[Rensink 2004; Yarin et al. 1999]. Therefore, the st<strong>and</strong>ing acoustic wave induces an<br />

additional convective blowing during <strong>droplet</strong> evaporation, resulting in an increase in Sherwood<br />

x<br />

x1


64 ACOUSTIC LEVITATION<br />

Number according to mass transfer considerations. The basic formula for the average Sherwood<br />

Number <strong>of</strong> an acoustically levitated <strong>droplet</strong> is derived by Yarin et al. [1999].<br />

Equation 3.33<br />

Sh<br />

= K<br />

acoustic<br />

⋅<br />

( ) 2 / 1<br />

ω ⋅ D<br />

SW<br />

B<br />

AB<br />

where the factor K is given by the acoustic boundary layer considerations <strong>of</strong> the inner acoustic<br />

streaming near the <strong>droplet</strong> surface. The exact derivation can be seen in Yarin et al. [1999] <strong>and</strong><br />

Rensink [2004]. If an approximation is used, K is be calculated by<br />

Equation 3.34<br />

K<br />

acoustic<br />

=<br />

2<br />

π<br />

⎜<br />

⎛<br />

⎝<br />

∫<br />

x<br />

x2<br />

u<br />

u<br />

acoustic<br />

acoustic<br />

r<br />

2<br />

r dx<br />

⎟<br />

⎞<br />

⎠<br />

1/<br />

2<br />

⎛<br />

⎜<br />

ωSW<br />

⋅ r<br />

⋅<br />

⎜<br />

⎝ A0e<br />

/<br />

/ ν<br />

S 0<br />

2 ( ρ ⋅ν<br />

)<br />

r is the distance <strong>of</strong> a point on the <strong>droplet</strong> surface to the semi-minor axis set dimensionless with the<br />

initial <strong>droplet</strong> radius 0<br />

r S . The term u acoustic is set dimensionless with the particle velocity B <strong>and</strong><br />

uacoustic r is averaged over the whole <strong>droplet</strong> surface. Due to a permanently changing <strong>droplet</strong><br />

shape, the factor K acoustic has to be re-calculated continuously [Rensink 2004]. For a small spherical<br />

<strong>droplet</strong> levitated directly within the pressure node <strong>of</strong> a st<strong>and</strong>ing acoustic wave the Sherwood<br />

number is given by<br />

Equation 3.35<br />

Sh<br />

⎛ 45 ⎞<br />

= ⎜ ⎟<br />

⎝ 4 π ⎠<br />

1/<br />

2<br />

⋅<br />

B<br />

( ) 2 / 1<br />

ω ⋅<br />

SW DAB<br />

Influences caused by flows inside the levitated <strong>droplet</strong> are not considered in Equation 3.33 <strong>and</strong><br />

Equation 3.35, though this effect could increase the velocity <strong>of</strong> the inner acoustic streaming up to<br />

10%. However, the Sherwood number would only increase by the factor ( 1.<br />

1)<br />

1.<br />

049<br />

gas<br />

0<br />

⎞<br />

⎟<br />

⎠<br />

1/<br />

2<br />

1 / 2<br />

= [Yarin et<br />

al. 1999]. Therefore, the circulation inside the <strong>droplet</strong>s can be neglected for mass transfer<br />

considerations [Rensink 2004].


CHAPTER 3 ACOUSTIC LEVITATION 65<br />

Influence <strong>of</strong> the outer acoustic streaming on mass transfer<br />

In contrast to the increased mass transfer due to an additional convective influence <strong>of</strong> the inner<br />

acoustic streaming, the outer acoustic streaming leads to a decrease <strong>of</strong> evaporation rate. According<br />

to Figure 3.5, the outer acoustic streaming forms a system <strong>of</strong> toroidal vortices, in which solvent<br />

vapour <strong>of</strong> the evaporated <strong>droplet</strong> is accumulated. If this volume is considered a closed system, heat<br />

<strong>and</strong> mass transfer from the vortices to the outside is repressed. Kastner determined the volume <strong>of</strong><br />

the vortices experimentally to be approximately V = 87µl<br />

[Kastner 2001] <strong>and</strong> showed that the<br />

accumulation <strong>of</strong> solvent vapour proceeded to saturation p / p = 1 within 0.6 seconds for a 2 µl<br />

ethanol <strong>droplet</strong>.<br />

V <strong>droplet</strong> [µl]<br />

2.00<br />

1.98<br />

1.96<br />

1.94<br />

Figure 3.7: Evaporation <strong>of</strong> a 2µl ethanol <strong>droplet</strong> neglecting any convective mass transfer considerations (Sh=2.0).<br />

Shown is the linear decrease <strong>of</strong> the <strong>droplet</strong> volume with time <strong>and</strong> the pressure ratio <strong>of</strong> the solvent vapour pressure in<br />

relation to its saturation vapour pressure at 25°C [Kastner 2001].<br />

To overcome the effect <strong>of</strong> the outer acoustic streaming on reduction <strong>of</strong> the evaporation rate, the<br />

system <strong>of</strong> toroidal vortices needs to be embedded into an axial ventilation air stream. The<br />

accumulated solvent vapour is now blown out <strong>of</strong> the vortices. If the flow <strong>of</strong> ventilation air stream is<br />

large enough, the amount <strong>of</strong> solvent vapour blown out <strong>of</strong> the vortices equals the amount solvent<br />

vapour transported into the vortices from the evaporation <strong>droplet</strong> [Rensink 2004]. To determine the<br />

necessary magnitude <strong>of</strong> the air stream it is also important to consider influences <strong>of</strong> the st<strong>and</strong>ing<br />

acoustic wave. Yarin et al. visualize the air stream <strong>and</strong> the system <strong>of</strong> vortices around a levitated 5µl<br />

n-hexadecane <strong>droplet</strong> inside an acoustic levitator at a SPL <strong>of</strong> approximately 156 dB. The series <strong>of</strong><br />

images (Figure 3.8) indicate that the air stream flow only succeeds in passing around the <strong>droplet</strong> for<br />

vortices<br />

1.92<br />

Droplet volume<br />

Pressure ratio<br />

1.90<br />

0.1<br />

0.0 0.5 1.0 1.5 2.0<br />

t [s]<br />

s<br />

10<br />

1<br />

p / p sat [-]


66 ACOUSTIC LEVITATION<br />

Reynolds numbers exceeding about 100 to 150 calculated at the orifice <strong>of</strong> the air stream [Yarin et<br />

al. 1997].<br />

ReO=70<br />

Figure 3.8: Flowfield around a levitated 5µl n-hexadecane <strong>droplet</strong> at a SPL <strong>of</strong> 156.29dB at different Reynolds numbers<br />

<strong>of</strong> the ventilation air stream [Yarin et al. 1997].<br />

According to Rensink [2004], the minimal flow velocity <strong>of</strong> a ventilation air stream u ventilation<br />

dependent on the SPL <strong>of</strong> the st<strong>and</strong>ing acoustic wave is given by<br />

Equation 3.36<br />

u<br />

ventilation<br />

A0<br />

≥<br />

ρ ⋅ν<br />

gas<br />

0<br />

For a levitated water <strong>droplet</strong> at a SPL <strong>of</strong> 165.7dB (above example) the minimal flow velocity is<br />

u = 4.<br />

3m/s<br />

. Rensink demonstrated that lower air velocities can also be used for ventilation<br />

ventilation<br />

ReO=190<br />

ReO=290<br />

as long as they overcome the effect <strong>of</strong> the st<strong>and</strong>ing acoustic wave shown in Figure 3.8. An air<br />

velocity lower than u ventilation<br />

with a distance <strong>of</strong> 20mm between orifice <strong>and</strong> <strong>droplet</strong> does not reach the<br />

<strong>droplet</strong> surface but is large enough to blow the solvent vapour out <strong>of</strong> the toroidal vortices<br />

completely. This has been shown experimentally by Rensink in Figure 3.9 for different solvent<br />

<strong>droplet</strong>s [Rensink 2004]. At a velocity <strong>of</strong> 1.0m/s the evaporation rate increased by the factor 2.6 for<br />

all tested solvents. At this velocity the solvent vapour is blown out <strong>of</strong> the vortices completely. A<br />

further increase leads to an additional convective influence on the <strong>droplet</strong> surface resulting in a<br />

further increase <strong>of</strong> evaporation rate [Rensink 2004].


CHAPTER 3 ACOUSTIC LEVITATION 67<br />

Figure 3.9: Diagram showing the influence <strong>of</strong> the ventilation air stream on the evaporation rate <strong>of</strong> levitated <strong>droplet</strong>s <strong>of</strong><br />

pure solvents. The y-axis plots the ratio between the evaporation rate with <strong>and</strong> the without ventilation. Therefore all<br />

curves start at y=1.0 for 0.0m/s ventilation velocity [Rensink 2004]<br />

3.2.3 Influence <strong>of</strong> the <strong>droplet</strong> volume<br />

A levitated liquid <strong>droplet</strong> is deformed to an oblate ellipsoid caused by the different values <strong>of</strong> axial<br />

<strong>and</strong> radial levitation forces. During the evaporation <strong>of</strong> this <strong>droplet</strong> the aspect ratio <strong>of</strong> the semi-axis<br />

r hor / rvert<br />

changes due to an increase in liquid pressure in the interior. This pressure increase results<br />

from a decrease in <strong>droplet</strong> radius according to Laplace’s Law<br />

Equation 3.37<br />

2 2 2<br />

) / (d/dt (dS / dS )0<br />

0<br />

2 / dS 0<br />

d/dt (d S<br />

4.0<br />

3.5<br />

3.0<br />

2.5<br />

⋅γ<br />

p =<br />

r<br />

2<br />

Δ<br />

S<br />

Furthermore, a decrease in <strong>droplet</strong> volume leads to an increase in the effective SPL <strong>of</strong> the st<strong>and</strong>ing<br />

acoustic wave. This increase is the effect <strong>of</strong> a resonance shift caused by reflections <strong>of</strong> the acoustic<br />

wave at the <strong>droplet</strong> surface that change with difference <strong>droplet</strong> size [Kastner 2001]. On basis <strong>of</strong><br />

these assumptions, Yarin et al. developed a method numerically calculating the effective SPL from<br />

a <strong>droplet</strong> volume <strong>and</strong> its aspect ratio in an open levitator [Yarin et al. 1998]. Using this<br />

mathematical procedure, Rensink determined the SPL change dependent on the <strong>droplet</strong> volume for<br />

different pure solvent <strong>droplet</strong>s, verifying an increase in SPL with decreasing <strong>droplet</strong> volume<br />

[Rensink 2004]. During the whole <strong>droplet</strong> evaporation the values <strong>of</strong> the effective SPL for water<br />

were much larger than for the other solvents. The reason is the difference in liquid density <strong>of</strong> water<br />

3<br />

ρ water = 1000 kg/m <strong>and</strong> <strong>of</strong> hydrocarbon<br />

2.0<br />

n-decane<br />

1.5<br />

water<br />

n-octane<br />

1.0<br />

2-propanol<br />

methanol<br />

0.5<br />

0.0<br />

ethanol<br />

n-heptane<br />

0.0 0.4 0.8 1.2 1.6<br />

u ventilation [m/s]<br />

3<br />

ρ hydrocarbon<br />

≅ 800 kg/m . The levitation force needed to


68 ACOUSTIC LEVITATION<br />

levitate a 3µl water <strong>droplet</strong> has therefore to be much larger to compensate the gravitational force<br />

than for a hydrocarbon <strong>droplet</strong> with the same volume [Rensink 2004].<br />

SPL [dB]<br />

171<br />

170<br />

169<br />

168<br />

167<br />

166<br />

165<br />

164<br />

163<br />

162<br />

161<br />

0.0 0.5 1.0 1.5 2.0 2.5 3.0<br />

Figure 3.10: Computed evolution <strong>of</strong> the effective SPL during the evaporation process <strong>of</strong> drops <strong>of</strong> water, alcohols <strong>and</strong><br />

alkanes in the acoustic field [Rensink 2004]<br />

3.2.4 Vertical position <strong>of</strong> the levitated <strong>droplet</strong><br />

V <strong>droplet</strong> [µl]<br />

Under the gravitational influence <strong>of</strong> the earth, the position <strong>of</strong> a <strong>droplet</strong>’s centre <strong>of</strong> mass is not<br />

exactly at the pressure node <strong>of</strong> the acoustic field. Due to the mass <strong>of</strong> the <strong>droplet</strong> <strong>and</strong> mass<br />

distributions inside it, the centre <strong>of</strong> mass is displaced to a stable position under the pressure node.<br />

Additionally, the strength <strong>of</strong> the ultrasonic field leads also to a larger or smaller displacement <strong>of</strong> the<br />

<strong>droplet</strong>. A decrease in <strong>droplet</strong> volume during <strong>drying</strong> increases the SPL, resulting in a vertical rise <strong>of</strong><br />

the levitated <strong>droplet</strong> to the adjacent upper pressure node. According to Equation 3.19, the same<br />

effect should be observed if the SPL increases keeping the <strong>droplet</strong> volume <strong>and</strong> <strong>droplet</strong> properties<br />

constant. If the mass <strong>of</strong> a levitated <strong>droplet</strong> is changed, the vertical position <strong>of</strong> the <strong>droplet</strong> has to<br />

change also, according the balance <strong>of</strong> forces <strong>of</strong> the gravitational <strong>and</strong> the levitation force<br />

[Kastner 2001].<br />

4<br />

⎛ A ⎞<br />

3<br />

2<br />

Equation 3.38 r ρ g ρ r ⎜ 0e<br />

π⋅<br />

S ⋅ <strong>droplet</strong> ⋅ = π gas ⋅ S ⋅ ⎟ ⋅ sin(<br />

2 ⋅ k ⋅ ΔxD<br />

) ⋅ f ( Ω )<br />

⎜ ρ gas ν ⎟<br />

0<br />

3<br />

⎝ ⋅ 0 ⎠<br />

2<br />

water<br />

ethanol<br />

methanol<br />

n-heptane


CHAPTER 3 ACOUSTIC LEVITATION 69<br />

3<br />

If the mass <strong>of</strong> the <strong>droplet</strong> ( / 3⋅<br />

π⋅<br />

⋅ ρ ⋅ g)<br />

4 decreases due to evaporation <strong>of</strong> solvent, the<br />

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

position between <strong>droplet</strong> centre <strong>of</strong> mass <strong>and</strong> pressure node ( x )<br />

Δ gets smaller <strong>and</strong> the <strong>droplet</strong> rises<br />

within the stationary ultrasonic field. Equation 3.38 illustrates the influence <strong>of</strong> mass loss on the<br />

vertical position <strong>of</strong> the <strong>droplet</strong> in the case <strong>of</strong> mass loss with a decrease in volume at constant<br />

density, or in the case <strong>of</strong> constant volume with a decreasing density. It can be seen, that changes in<br />

displacement are larger for a decrease in density than in volume [Rensink 2004].<br />

Δ x D [mm]<br />

0.23<br />

0.22<br />

0.21<br />

0.20<br />

0.19<br />

0.18<br />

0.17<br />

0.16<br />

Figure 3.11: Diagram illustrating the change <strong>of</strong> the distance between geometric mass balance point <strong>of</strong> the <strong>droplet</strong> <strong>and</strong><br />

pressure node <strong>of</strong> the st<strong>and</strong>ing acoustic wave with <strong>droplet</strong> mass. The change in density at constant <strong>droplet</strong> volume have a<br />

stronger influence on the vertical position than the decrease in volume at constant <strong>droplet</strong> density [Kastner et al. 2001]<br />

3.2.5 Influences <strong>of</strong> the ultrasonic transducer<br />

1500 1600 1700 1800 1900 2000<br />

m D [µg]<br />

Oscillation <strong>of</strong> the ultrasonic transducer causes heating <strong>of</strong> the interior <strong>of</strong> the levitation chamber. The<br />

increase in temperature leads to an increase in evaporation rate <strong>of</strong> liquid <strong>droplet</strong>s <strong>and</strong> a decrease in<br />

evaporation time. Kastner measured a temperature increase <strong>of</strong> 5°C inside the levitator at an ambient<br />

temperature <strong>of</strong> 20°C without any additional ventilation [Kastner 2001]. This thermal effect has to be<br />

kept in mind for any experiments at ambient temperature without a ventilation air stream. In the<br />

case <strong>of</strong> experiments at elevated temperature the thermal effect can be neglected, because the heat<br />

quantity supplied by external heaters achieves the target value. The total heat quantity needed for<br />

the target temperature is the sum <strong>of</strong> the heat quantity supplied by the external heater <strong>and</strong> the heat<br />

quantity supplied by the ultrasonic transducer itself. Temperature measurements are carried out<br />

inside the levitations system to have precise temperature information.<br />

D<br />

Volume decrease<br />

at constant density<br />

Density decrease<br />

at constant volume


70 ACOUSTIC LEVITATION<br />

3.3 <strong>Single</strong> <strong>droplet</strong> <strong>drying</strong> in an acoustic levitator<br />

3.3.1 Pure solvent <strong>droplet</strong>s<br />

A model to calculate the kinetics for diffusion controlled evaporation <strong>of</strong> a single <strong>droplet</strong> <strong>of</strong> pure<br />

solvents was set up in chapter 2.4.3. The decrease in the squared radius with time is described in<br />

Equation 2.89, where Equation 2.90 gives the evaporation coefficient β r . To account for increased<br />

mass transfer due to the influence <strong>of</strong> the inner acoustic streaming <strong>of</strong> the st<strong>and</strong>ing wave, Equation<br />

2.90 is changed to introduce the Sherwood number<br />

Equation 3.39<br />

2 ⋅ D<br />

β r =<br />

ρ<br />

AB<br />

liquid<br />

⋅ M<br />

⋅ R<br />

liquid<br />

⎛ p<br />

⋅ ⎜<br />

⎝ T<br />

S<br />

S<br />

p<br />

−<br />

T<br />

Together with Equation 3.33 the evaporation coefficient β r is now given by<br />

Equation 3.40 ⎟ B ⋅ M liquid ⎛ D ⎞ ⎛ ⎞<br />

AB pS<br />

p∞<br />

β r = K acoustic⋅<br />

⋅ ⎜<br />

⎟ ⋅ ⎜ −<br />

ρliquid<br />

⋅ R ⎝ ωSW<br />

⎠ ⎝ TS<br />

T∞<br />

⎠<br />

The <strong>droplet</strong> surface temperature T S results form heat transfer from the ambient air to the <strong>droplet</strong><br />

<strong>and</strong> <strong>of</strong> heat loss due to evaporation <strong>of</strong> liquid. Additionally, condensation <strong>of</strong> water vapour from the<br />

ambient air on the <strong>droplet</strong> surface must be taken into account when calculating the surface<br />

temperature. Even though there is little condensation in experiments with low relative humidity <strong>of</strong><br />

the ambient air, it still supplies heat to the <strong>droplet</strong>. According to Yarin et al. [1999] <strong>and</strong> Rensink<br />

[2004], the surface temperature <strong>of</strong> a levitated <strong>droplet</strong> can be calculated by<br />

Equation 3.41<br />

T<br />

S<br />

∞<br />

∞<br />

⎞<br />

⎟ ⋅<br />

⎠<br />

1/<br />

2<br />

Sh<br />

2<br />

( T )<br />

1/<br />

2<br />

⎡ D ⎛ κ<br />

⎤<br />

gas ⎞ psat<br />

S ⋅ M<br />

2 AB<br />

liquid<br />

= Y + ⎢Y<br />

− ⋅ ⎜<br />

⎟ ⋅<br />

⋅ hv<br />

⎥<br />

⎢ λ<br />

⋅<br />

⎣ gas ⎝ DAB<br />

⎠ R TS<br />

⎥⎦<br />

Y<br />

=<br />

1 ⎡<br />

⋅ ⎢T<br />

2 ⎢⎣<br />

1/<br />

2<br />

( T )<br />

D ⎛ κ ⎞<br />

air/water gas ϕ ⋅ psat<br />

∞ ⋅ M water<br />

+ ⋅ ⎜<br />

⎟ ⋅<br />

⋅ h<br />

λgas<br />

⎝ Dair/water<br />

⎠ R ⋅T∞<br />

∞ v water<br />

κ gas <strong>and</strong> λgas are the diffusivity <strong>and</strong> the thermal conductivity <strong>of</strong> the ambient gas, D AB the binary<br />

diffusion coefficient <strong>of</strong> the solvent in the ambience <strong>and</strong> h v its latent heat <strong>of</strong> evaporation. The term<br />

in the second row <strong>of</strong> Equation 3.41 represents the condensation <strong>of</strong> water vapour on the <strong>droplet</strong><br />

surface. Because the substance property values depend on the surface temperature, Equation 3.41<br />

⎤<br />

⎥<br />

⎥⎦


CHAPTER 3 ACOUSTIC LEVITATION 71<br />

has to be solved iteratively to find T S . The calculation <strong>of</strong> the temperature-dependent values <strong>of</strong> T S<br />

can be found in Annex A. The evaporation coefficient for pure solvent <strong>droplet</strong>s under Stefan flow is<br />

derived from Equation 2.64 using Equation 3.33 for the Sherwood number, to yield<br />

B ⋅ ρ gas ⎛ D ⎞ AB<br />

Equation 3.42 β r,<br />

Stefan = K acoustic⋅<br />

⋅ ⎜<br />

⎟ ⋅ ln(<br />

1 − BM<br />

)<br />

ρliquid<br />

⎝ ωSW<br />

⎠<br />

3.3.2 Droplets <strong>of</strong> binary liquid mixtures<br />

The evaporation <strong>of</strong> <strong>droplet</strong>s <strong>of</strong> binary liquid mixtures was described in detail by Yarin et al.<br />

[2002 b] <strong>and</strong> summarized by Rensink [2004]. According to Rensink, the radius-time course is given<br />

by<br />

Equation 3.43<br />

h c1<br />

<strong>and</strong> c2<br />

ρ<br />

liquid<br />

( T )<br />

1/<br />

2<br />

d r ⎡ ⎛ p1,<br />

sat S ⋅ M<br />

S<br />

liquid 1<br />

⎞<br />

= −⎢<br />

hc1<br />

⋅ ⎜<br />

⋅ψ<br />

1 ⋅ Z1(<br />

rS<br />

) − cS∞1<br />

⎟<br />

d t ⎣ ⎝ R ⋅TS<br />

⎠<br />

( T )<br />

⎛ p2,<br />

sat S ⋅ M liquid 2<br />

hc ⋅ ⎜<br />

⋅ψ<br />

2 ⋅ Z<br />

⎝ R ⋅TS<br />

⎞⎤<br />

( rS<br />

) − c ⎟<br />

⎟⎥<br />

⎠⎦<br />

+ 2 2<br />

S∞2<br />

h are the time averaged heat transfer coefficients <strong>of</strong> the whole <strong>droplet</strong> surface <strong>of</strong> the<br />

liquid components 1 <strong>and</strong> 2. They can be calculated using the following equation together with<br />

Equation 3.33 <strong>and</strong> Equation 3.34<br />

Equation 3.44<br />

Sh<br />

i<br />

=<br />

hc1<br />

⋅ 2r<br />

D<br />

AB,<br />

i<br />

S<br />

The activity coefficients ψ 1 <strong>and</strong> ψ 2 are given in Hirata et al. [Yarin et al. 2002 b].<br />

Equation 3.45<br />

1<br />

⎡ ⎛ Ldw<br />

L<br />

ψ i =<br />

⋅ exp⎢−<br />

Z ⋅ ⎜<br />

−<br />

Zi + Lwd<br />

⋅ i ⎣ ⎝ i dw i wd 1<br />

i<br />

( 1 − Z ) ( 1 − Z ) + L ⋅ Z L ⋅ ( − Z )<br />

wd<br />

i<br />

⎞⎤<br />

⎟<br />

⎟⎥<br />

+ Zi<br />

⎠⎦<br />

The coefficients L dw <strong>and</strong> L wd are characteristics <strong>of</strong> a specific liquid pair <strong>and</strong> are given for<br />

water/ethanol mixtures by L dw = 0.<br />

1108 <strong>and</strong> L wd = 0.<br />

956 [Yarin et al. 2002 b]. Z i is the mole<br />

fraction <strong>of</strong> the component i in the binary liquid mixture.


72 ACOUSTIC LEVITATION<br />

The mixture density <strong>of</strong> the <strong>droplet</strong> ρ liquid can be calculated using the density <strong>of</strong> the pure solvents ρ i<br />

<strong>and</strong> their mass fraction Y i [Rensink 2004]<br />

Equation 3.46<br />

ρ<br />

liquid<br />

1<br />

=<br />

Y / ρ +<br />

1<br />

1<br />

( 1 − Y1<br />

) / ρ 2<br />

where Y1 = ( Z1<br />

⋅ M 1)<br />

/ ( Z1<br />

⋅ M 1 + Z 2 ⋅ M 2 ) . To calculate Z 1 ( r = rS<br />

) <strong>and</strong> Z 2 ( r = rS<br />

)<br />

, the diffusion<br />

equation inside the <strong>droplet</strong> must be solved to find the mole (or mass) fraction distributions along the<br />

<strong>droplet</strong> radius at each instant <strong>of</strong> time. The solution <strong>of</strong> this diffusion problem can be seen at Yarin et<br />

al. [2002 b]. The surface temperature <strong>of</strong> a <strong>droplet</strong> <strong>of</strong> binary liquid mixtures is given by Rensink<br />

[2004].<br />

3.3.3 Solution <strong>and</strong> suspension <strong>droplet</strong>s<br />

The <strong>drying</strong> <strong>of</strong> <strong>droplet</strong>s containing dissolved or suspended solids can be divided into two different<br />

stages <strong>and</strong> for hygroscopic substances even in three different stages. The first <strong>drying</strong> stage<br />

resembles the evaporation <strong>of</strong> pure solvent <strong>droplet</strong>s (Figure 3.12 a) <strong>and</strong> can therefore by described by<br />

the same equations but also taking vapour pressure lowering effects into account due to dissolved<br />

substances (see chapter 3.3.1). The decrease <strong>of</strong> <strong>droplet</strong> volume within the first <strong>drying</strong> stage can be<br />

used directly to determine the evaporation rate [Kastner et al. 2001].<br />

Equation 3.47<br />

m&<br />

liquid<br />

= −ρ<br />

liquid<br />

ΔV<br />

⋅<br />

Δ t<br />

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

In this phase the squared diameter or radius decreases almost linearly with time according to the<br />

d 2 -law (Figure 3.12 a). After the critical point, marking the change from the first to the second<br />

<strong>drying</strong> stage, the volume <strong>of</strong> the levitated <strong>droplet</strong> remains constant with time even though there is<br />

still evaporation <strong>of</strong> solvent from inside the <strong>droplet</strong>. This evaporation <strong>of</strong> solvent leads to a decrease<br />

in mass <strong>and</strong> density <strong>of</strong> the particle. This causes a rise <strong>of</strong> the particle within the stationary ultrasonic<br />

field (Figure 3.12 b) that can be used to determine the quantity <strong>of</strong> solvent evaporated in the second<br />

<strong>drying</strong> stage. It is vital that the SPL does not change within the second <strong>drying</strong> stage due to the<br />

change in density <strong>of</strong> the particle. Otherwise determination <strong>of</strong> <strong>drying</strong> kinetics would not be possible.<br />

In chapter 3.2.3 the influence <strong>of</strong> the <strong>droplet</strong> on the st<strong>and</strong>ing acoustic wave <strong>and</strong> SPL was discussed.<br />

Experiments by Kastner [2001] showed that the working voltage <strong>of</strong> the ultrasonic transducer <strong>and</strong>


CHAPTER 3 ACOUSTIC LEVITATION 73<br />

therefore the SPL according to Equation 3.27 is indeed constant during the second <strong>drying</strong> stage<br />

(Figure 3.12 c).<br />

[mm 2<br />

]<br />

2<br />

Δ x D,0 - Δ x D<br />

d S<br />

U [mV]<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

0.05<br />

0.00<br />

-0.05<br />

-0.10<br />

-0.15<br />

-0.20<br />

36<br />

34<br />

32<br />

(a)<br />

(b)<br />

First <strong>drying</strong> stage<br />

(constant rate)<br />

(c)<br />

30<br />

0 50 100 150 200 250<br />

Figure 3.12: Drying characteristics <strong>of</strong> a 1.2 µl suspension <strong>droplet</strong> containing ethanol <strong>and</strong> small glass spheres<br />

dglass=30 µm at 25°C. The solid content was Y0=0.05. (a) decrease <strong>of</strong> the squared diameter with time; (b) vertical<br />

position <strong>of</strong> the <strong>droplet</strong>; (c) voltage at the ultrasonic transducer proportional to the SPL <strong>of</strong> the st<strong>and</strong>ing acoustic wave<br />

[Kastner 2001].<br />

t [s]<br />

Second <strong>drying</strong> stage<br />

(falling rate)


74 ACOUSTIC LEVITATION<br />

The rise <strong>of</strong> the particle within the ultrasonic field is therefore only caused by change <strong>of</strong> the mean<br />

particle density. The <strong>droplet</strong> diameter, the aspect ratio <strong>and</strong> the characteristics <strong>of</strong> the ultrasonic field<br />

at the beginning <strong>of</strong> the <strong>drying</strong> experiment must now be known. The strength <strong>of</strong> the ultrasonic field<br />

<strong>and</strong> the displacement <strong>of</strong> the <strong>droplet</strong> centre <strong>of</strong> mass from the adjacent pressure node are measured.<br />

With this information <strong>and</strong> that from the first <strong>drying</strong> stage, the ΔxD at the critical point is determined.<br />

Assuming a constant SPL within the second <strong>drying</strong> stage, a new constant describing the<br />

strength <strong>and</strong> properties <strong>of</strong> the acoustic field is introduced using Equation 3.19 [Kastner et al. 2001]<br />

2<br />

ρ gas ⋅ rS<br />

A0e<br />

Equation 3.48 K SPL = π<br />

f ( Ω )<br />

g ρ ⋅ c<br />

gas<br />

0<br />

The evaporation rate is then determined via the changing position <strong>of</strong> the centre <strong>of</strong> mass <strong>of</strong> the<br />

particle using [Kastner et al. 2001]<br />

K SPL<br />

Equation 3.49 m& L = ⋅ sin(<br />

2 ⋅ k0<br />

⋅ Δ xD<br />

1)<br />

− sin(<br />

2 ⋅ k0<br />

⋅ Δ xD<br />

2 )<br />

Δ t<br />

2<br />

[ ]<br />

If the evaporation rate is known, then calculation <strong>of</strong> the mean particle density ρ particle , the mean<br />

moisture content <strong>of</strong> the <strong>droplet</strong> X particle <strong>and</strong> the mean particle porosity ε mean is possible [Kastner et<br />

al. 2001]<br />

m<br />

Equation 3.50 ρ () t =<br />

particle<br />

Equation 3.51 X particle () t =<br />

m<br />

Equation 3.52<br />

liquid<br />

solid<br />

V<br />

m<br />

liquid<br />

( t)<br />

msolid<br />

ε mean = 1 −<br />

ρ ⋅V<br />

+ m<br />

particle<br />

solid<br />

liquid ( t)<br />

() t + msolid<br />

paricle<br />

Ultrasonic levitation is therefore one <strong>of</strong> the few techniques to investigate the <strong>drying</strong> process <strong>of</strong><br />

solution or suspension <strong>droplet</strong>s within both <strong>drying</strong> stages. It is suitable to determine the evaporation<br />

rate <strong>and</strong> solvent content <strong>of</strong> a formulation at any desired point <strong>of</strong> time during evaporation. The set-up<br />

also enables measurements at various temperatures <strong>and</strong> humidity to check influences <strong>of</strong> ambient<br />

conditions on the <strong>drying</strong> characteristics. Furthermore, <strong>droplet</strong>s / particles can be removed from the<br />

levitation chamber for further analysis at any time.


CHAPTER 4 MATERIALS AND METHODS 75<br />

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

4.1 Materials<br />

4.1.1 Proteins<br />

Bovine Serum Albumin (bSA)<br />

Bovine serum albumin is a single polypeptide chain consisting <strong>of</strong> about 583 amino acids residues<br />

<strong>and</strong> no carbohydrates [Sigma 2005]. The molecular weight <strong>of</strong> 66.430 kDa is cited by<br />

Hirayama et al. [1990]. At pH 5-7 it contains 17 intrachain disulfide bridges <strong>and</strong> one sulfhydryl<br />

group. None <strong>of</strong> the disulfide bonds was accessible to reducing agents in this pH range [Katchalski et<br />

al. 1957]. Based on hydrodynamic experiments <strong>and</strong> low angle X-ray scattering, serum albumin was<br />

postulated to be an oblate ellipsoid with dimensions <strong>of</strong> 140 x 40 Å, with a secondary structure <strong>of</strong><br />

50-68% alpha-helix <strong>and</strong> 16-18% beta-sheet (Figure 4.1).<br />

Figure 4.1: Classical preception <strong>of</strong> the structure <strong>of</strong> BSA [Friedli 1996]<br />

Studies using a 1H NMR <strong>and</strong> X-ray crystallographic data indicate a heart shaped structure.<br />

According to X-ray crystallography there is no beta-sheet in the structure <strong>of</strong> native serum albumin<br />

but 67% alpha-helix. The remaining polypeptide exists in turns <strong>and</strong> extended flexible regions<br />

between subdomains with no beta-sheet [Friedli 1996]. Albumins in general are characterized by a<br />

low content <strong>of</strong> tryptophan <strong>and</strong> methionine <strong>and</strong> a high content <strong>of</strong> cysteine <strong>and</strong> the charged amino<br />

acids aspartic <strong>and</strong> glutamic acids, lysine <strong>and</strong> arginine. The glycine <strong>and</strong> isoleucine content <strong>of</strong> bSA is<br />

lower than in the average <strong>protein</strong> [Friedli 1996; Peters 1985]. Figure 4.2 shows that the bSA<br />

molecule is made up <strong>of</strong> three homologous domains (I, II, III), divided into nine loops (L1-L9) by<br />

the 17 disulfide bonds. Each domain can be classified into 10 helical segments, 1-6 for subdomain<br />

A <strong>and</strong> 7-10 for sub-domain B.


76 MATERIALS AND METHODS<br />

Figure 4.2: Secondary <strong>and</strong> tertiary structure <strong>of</strong> BSA. (a) Heart shaped structure <strong>of</strong> BSA, according to 1H NMR <strong>and</strong><br />

X-ray crystallography; (b) motif for sub-domain A; (c) motif for sub-domain B [Friedli 1996]<br />

A lyophilized powder <strong>of</strong> the pure <strong>protein</strong>, prepared by a modification <strong>of</strong> Cohn et al. [Cohn et al.<br />

1946], using cold ethanol, pH <strong>and</strong> low temperature precipitation followed by an additional pH<br />

adjustment step prior to final <strong>drying</strong>, was supplied by Sigma (Sigma-Aldrich; 82024 Taufkirchen,<br />

Germany) <strong>and</strong> used without further purification. The product was specified as follows: pH 7.0 in<br />

1% (w/v) aqueous solution, min. 96% (electrophoresis) lyophilised powder [Sigma 2004].<br />

Catalase from bovine liver (CAT)<br />

(a) (b) (c)<br />

Catalase from bovine liver is a tetramer <strong>of</strong> four identical sub-units, each consisting <strong>of</strong> 506 amino<br />

acids. It has an overall molecular weight <strong>of</strong> 250 kDa [Sigma 2005]. Each monomer contains a heme<br />

prosthetic group at the catalytic centre. The enzyme also binds NADP strongly, <strong>of</strong> which the NADP<br />

<strong>and</strong> heme group are within 13.7 Å <strong>of</strong> each other [Sigma 2005]. Only about 60% <strong>of</strong> the catalase<br />

structure is composed <strong>of</strong> regular secondary structure motifs. Alpha-helices account for 26% <strong>of</strong> its<br />

structure <strong>and</strong> beta-sheet for 12%. Irregular structure includes a predominance <strong>of</strong> extended single<br />

st<strong>and</strong>s <strong>and</strong> loops that play a major roll in the assembly <strong>of</strong> the tetramer [Boon et al. 2001].<br />

Considering the tertiary structure, each sub-unit has four domains (Figure 4.3). The first is made up<br />

<strong>of</strong> the amino-terminal 75 residues <strong>and</strong> form an arm with two alpha-helices <strong>and</strong> a large loop<br />

extending from the globular sub-unit. The second <strong>and</strong> largest domain is composed <strong>of</strong> residues 76 to<br />

320 <strong>and</strong> may be classified as a α plus β type domain including a beta-barrel, several helical<br />

segments <strong>of</strong> 3 to 4 turns each <strong>and</strong> various loops. This domain contains the heme moiety. The third<br />

domain is made up by residues 321 to 436 <strong>and</strong> is referred to as wrapping domain. It lacks<br />

discernable secondary structure except for two helices. The carboxy-terminal part <strong>of</strong> catalase


CHAPTER 4 MATERIALS AND METHODS 77<br />

contains residues 437 to 506 <strong>and</strong> forms the surface <strong>of</strong> the enzyme along with 3 alpha-helices <strong>of</strong> the<br />

heme-containing second domain [Boon et al. 2001].<br />

(a) (b)<br />

Figure 4.3: Crystal structure <strong>of</strong> bovine liver catalase by X-ray diffraction. (a) Entire catalase molecule; (b) NADPH<br />

binding site <strong>of</strong> catalase molecule [PDB 2005].<br />

Catalase is present in the peroxisomes <strong>of</strong> nearly all aerobic cells. It protects the cell from the toxic<br />

effects <strong>of</strong> hydrogen peroxide by catalysing its decomposition into molecular oxygen <strong>and</strong> water<br />

without the production <strong>of</strong> free radicals. The mechanism <strong>of</strong> the catalysis can be seen in Figure 4.4.<br />

( III)<br />

− E → H O + O = Fe(<br />

IV)<br />

− E<br />

H 2O<br />

2 + Fe<br />

2<br />

( IV)<br />

− E → H O + O + Fe(<br />

III)<br />

− E<br />

H 2O<br />

2 + O = Fe<br />

2<br />

2<br />

Figure 4.4: Chemistry <strong>of</strong> catalase catalysis. The catalytic process is thought to occur in two stages, where Fe-E<br />

represents the iron centre <strong>of</strong> the heme attached to the rest <strong>of</strong> the enzyme E [Boon et al. 2001]<br />

Catalase does not require any activators, but it can be inhibited by cyanide, azide, hydroxylamine,<br />

cyanogens bromide, 2-mercaptoethanol, dithiothreitol dianisidine <strong>and</strong> nitrate. Catalase activity is<br />

constant over a pH range <strong>of</strong> 4.0-8.5 [Sigma 2005].<br />

The pure <strong>protein</strong> was supplied as powder by Sigma (Sigma-Aldrich; 82024 Taufkirchen,<br />

Germany) <strong>and</strong> used without further purification. The product was specified: 2000-5000 units /mg<br />

<strong>protein</strong>, cell culture tested. One unit will decompose 1.0 µmole <strong>of</strong> H2O2 per minute at pH 7.0 at<br />

25°C while the H2O2 concentration falls from 10.3 to 9.2 mM, measured by the decrease <strong>of</strong><br />

absorption at 240nm using UV-spectroscopy [Sigma 2005].


78 MATERIALS AND METHODS<br />

4.1.2 Excipients <strong>and</strong> Reagents<br />

All excipients <strong>and</strong> reagents used for preparation <strong>of</strong> levitator-dried <strong>formulations</strong>, spray-dried<br />

<strong>formulations</strong> <strong>and</strong> bubble pressure tensiometry <strong>formulations</strong>, as well as all other substances used<br />

during this work are summarized in Table 4.1. Aqueous solutions were prepared in water for<br />

analysis <strong>and</strong> filtered prior to use via a cellulose nitrate membrane filter with a pore diameter <strong>of</strong><br />

0.22 µm.<br />

Table 4.1: Excipients used to formulate the solutions <strong>and</strong> suspensions for the use in an acoustic levitator, spray-dryer<br />

<strong>and</strong> bubble pressure tensiometer.<br />

Excipients Lot-No. Supplied by<br />

Proteins<br />

Albumin, bovine serum, Fraction V 025K1031<br />

103K1373<br />

Sigma, Germany (A-2153)<br />

Catalase from bovine liver<br />

Sugars<br />

013K7055<br />

024K7034<br />

Sigma, Germany (C-1345)<br />

D-Mannitol 100K0129 Sigma, Germany (M-9546)<br />

D (+) Trehalose dehydrate 133K 3775<br />

122K3801<br />

Sigma, Germany (T-5251)<br />

Maltodextrin 15 from maize starch 404054/1 Fluka, Switzerl<strong>and</strong> (31412)<br />

Surfactants<br />

1-Octadecanol, approx. 99% 023K1276 Sigma, Germany (S-5751)<br />

Pluronic F-127 99H1194 Sigma, Germany (P-2443)<br />

Stearic acid S3818073 317 Merck, Germany (8.00673.1000)<br />

Polysorbate 80 440327/1 Fluka, Switzerl<strong>and</strong> (93781)<br />

Other substances<br />

Sodium chloride 10670 Riedel-de Haën (31434)<br />

Trizma pre-set crystals pH 8.0 122K5436 Sigma, Germany (T-8443)<br />

Solvents<br />

Hexane<br />

Ethanol 96% p. a. 10570238 Carl Roth GmbH, Germany<br />

(P-0751)<br />

Water p. a. OC328373<br />

OC522462<br />

Merck, Germany (1.16754.5000)


CHAPTER 4 MATERIALS AND METHODS 79<br />

Table 4.2: Reagents used for catalase activity assay.<br />

Reagents Lot-No. Supplied by<br />

Catalase activity assay<br />

Potassium hydroxide ≥85% 113K0019 Sigma, Germany (P-5958)<br />

Potassium phosphate monobasic 99.0% 024K0050 Sigma, Germany (P-5379)<br />

Hydrogen peroxide 30 wt.% 044K13332 Sigma, Germany (H-1009)<br />

Table 4.3: Reagents used for determination <strong>of</strong> residual water content via Karl-Fischer Titration.<br />

Reagents Lot-No. Supplied by<br />

Karl-Fischer Titration<br />

Nitrogen, gaseous Linde, Germany<br />

Hydranal © -Coulomat AG Oven 4056B Riedel-de Haën (34739)<br />

Hydranal © -Coulomat CG 4268A Riedel-de Haën (34840)<br />

Hydranal © -Humidity Absorber 03390 Riedel-de Haën (34788)<br />

4.1.3 Acoustic levitation system<br />

Acoustic levitator<br />

The acoustic levitator used in this work was a 58 kHz levitator supplied by tec5 AG (61440<br />

Oberursel, Germany). The acrylic process chamber (inner acrylic chamber) was built in the Institute<br />

<strong>of</strong> Fluid Mechanics <strong>of</strong> the Friedrich-Alex<strong>and</strong>er-University in Erlangen, Germany. The original setup<br />

is shown in Figure 4.5. Technical specifications according to tec5 AG are listed in Table 4.4.<br />

The acoustic wave inside the acoustic levitator is created by an oscillating transducer <strong>and</strong> reflected<br />

at a reflector positioned opposite to the transducer. The end plate <strong>of</strong> the transducer has a diameter <strong>of</strong><br />

12mm <strong>and</strong> the transfixion has a diameter <strong>of</strong> 2mm. The horn is attached to a piezo-crystal connected<br />

to a power supply unit. The power input in the transducer depends not only on the working voltage<br />

<strong>of</strong> the power supply unit but also on the distance L R between transducer <strong>and</strong> reflector. If this<br />

distance is set to a multiple <strong>of</strong> half wavelengths, n ⋅ λ / 2,<br />

a st<strong>and</strong>ing wave with equally space<br />

pressure nodes <strong>and</strong> antinodes is created. In every pressure node the levitation <strong>of</strong> a liquid or solid<br />

sample is possible.<br />

0


80 MATERIALS AND METHODS<br />

Table 4.4: Technical specifications <strong>of</strong> the acoustic levitator according to tec5 AG<br />

Technical specifications<br />

Power supply unit<br />

Operation frequency 58 kHz<br />

Particle diameter range ≈15µm to ≈2.5mm<br />

Wavelength <strong>of</strong> the st<strong>and</strong>ing acoustic wave 5.71mm<br />

HF-Power (continuously variable) 0.65 to 5 Watt<br />

Modulation frequency 10Hz – 2kHz<br />

Modulation amplitude 0 – 2 VPP<br />

Modulation input impedance 20kΩ<br />

Operation temperature range 0-70°C<br />

Relative humidity (no condensation) 10-90%<br />

The reflector is positioned opposite the transducer at the<br />

variable distance L R . The distance can be set precisely<br />

with a micrometer screw. Additionally, the horizontal<br />

position can be adjusted with two other micrometer screws.<br />

The reflector consists <strong>of</strong> a round flat metal plate with a<br />

diameter <strong>of</strong> 25.0mm <strong>and</strong> a height <strong>of</strong> 11.0mm. In the centre<br />

<strong>of</strong> the plate is a hole with a diameter <strong>of</strong> 3.6mm <strong>and</strong> a thread<br />

(Figure 4.6). It is used for conditioning <strong>of</strong> the levitation<br />

chamber <strong>and</strong> the ventilation air stream. The levitation<br />

chamber built around the transducer <strong>and</strong> reflector was<br />

made <strong>of</strong> acrylic glass <strong>and</strong> had the size given in Figure<br />

4.7 a. It is the inner acrylic glass box or process chamber.<br />

The base had a small hole for the tube supplying the<br />

ventilation air stream. A second hole in the back with a<br />

diameter <strong>of</strong> 17.3mm enabled the measurement <strong>of</strong><br />

temperature <strong>and</strong> humidity inside the process chamber by<br />

installing a dew point hygrometer. A rectangular recess in<br />

the front was made for the injection <strong>of</strong> the samples by a<br />

gas-chromatography syringe in case <strong>of</strong> liquid samples, or<br />

micro tweezers in case <strong>of</strong> solid samples.<br />

Figure 4.5: Original set-up <strong>of</strong> the acoustic<br />

levitator.


CHAPTER 4 MATERIALS AND METHODS 81<br />

10.7 mm<br />

25 mm<br />

3.75 mm<br />

25 mm<br />

3.75 mm<br />

Figure 4.6: Reflector dimensions <strong>of</strong> the acoustic levitator with borehole for the ventilations air stream.<br />

The acoustic levitator was assembled into an outer acrylic glass box similar to a glove box for<br />

microbiological research to isolate it from environmental conditions. An additional heater (Leister<br />

Heater 8D1-3000W; Leister, Switzerl<strong>and</strong>) was installed to heat the whole unit up to 60°C. The<br />

dimensions <strong>of</strong> the acrylic glass box are shown in Figure 4.7.<br />

124.9 mm<br />

49.7 mm<br />

15.3 mm<br />

62.2 mm<br />

7.1 mm<br />

58 mm 30.6 mm<br />

17.3 mm<br />

95.7 mm<br />

88.6 mm<br />

29.2 mm<br />

(a)<br />

11 mm


82 MATERIALS AND METHODS<br />

55 mm<br />

130 mm<br />

55 mm<br />

145 mm<br />

185 mm 130 mm<br />

380 mm 190 mm 400 mm<br />

580 mm<br />

40 mm<br />

60 mm<br />

Figure 4.7: Dimensions <strong>of</strong> the acrylic glass containers built around the levitation system. Two gloves are fixed at the<br />

two holes in the front side to reach inside the container <strong>and</strong> to inject the sample into the levitator. Through the larger<br />

hole at the back side an additional heated air stream is blown into the container to heat the entire system. The small hole<br />

is used for electric cables <strong>and</strong> the tube for the ventilation air stream. The part on the right side <strong>of</strong> the hatched area<br />

contains a door to pass samples <strong>and</strong> material into the container. The hatched area itself is movable <strong>and</strong> can be drawn out<br />

<strong>of</strong> the container a sluice.<br />

Controlled evaporation mixer (CEM)<br />

The CEM-system is a liquid delivery system that can be applied for atmospheric or vacuum<br />

processes. It consists <strong>of</strong> a liquid flow controller, a mass flow controller for the carrier gas <strong>and</strong> a<br />

temperature controlled mixing <strong>and</strong> evaporation device. All three were made by Bronkhorst (NL)<br />

<strong>and</strong> supplied by Wagner (“Mess- und Regeltechnik”, 63067 Offenbach, Germany). The technical<br />

specifications <strong>and</strong> calibration details <strong>of</strong> the CEM-unit can be seen in Table 4.5 <strong>and</strong> Figure 4.8.<br />

According to the technical specifications it was possible to condition the levitation chamber from<br />

room temperature up to 70°C with a relative humidity from a few ppm to almost 100% depending<br />

on the air flow rate. The relation between evaporated liquid, air flow velocity <strong>and</strong> temperature was<br />

calculated using Fluidat © on the net V1.12/5.59 [Bronkhorst 2005]. Knowing the velocity <strong>and</strong> the<br />

relative humidity <strong>of</strong> the air stream the orifice Reynolds number, Re orifice , at the reflector was<br />

calculated.<br />

50 mm<br />

150 mm<br />

340 mm<br />

380 mm<br />

(b)


CHAPTER 4 MATERIALS AND METHODS 83<br />

Table 4.5: Technical specifications <strong>and</strong> calibration details <strong>of</strong> the controlled evaporation mixer<br />

Technical specifications<br />

Liquid flow controller type L1-FAC-33-0<br />

Serial number M2207627B<br />

Medium water<br />

Feeding pressure 1 bar<br />

Flow range 0.2 to 10.0 g/h<br />

Mass flow controller type F-201C-FAC<br />

Serial number M2207627A<br />

Medium dry air / nitrogen<br />

Feeding pressure 1 bar<br />

Flow range 0.04 to 2.0 ln/min<br />

Temperature controlled mixing unit type W-202-330-P<br />

Serial number M2207627C<br />

Range up to 200°C<br />

compressed<br />

air<br />

cleaning unit<br />

water tank<br />

Mass<br />

flow<br />

Liquid<br />

flow<br />

mixing unit<br />

with heater<br />

Figure 4.8: Components <strong>and</strong> adjustment <strong>of</strong> the controlled evaporation mixer.<br />

to acoustic<br />

levitator


84 MATERIALS AND METHODS<br />

CCD-camera <strong>and</strong> imaging s<strong>of</strong>tware<br />

To record <strong>and</strong> measure the decrease <strong>of</strong> the <strong>droplet</strong> radius with time <strong>and</strong> the change in<br />

position <strong>of</strong> the centre <strong>of</strong> mass <strong>of</strong> the levitated <strong>droplet</strong> during the <strong>drying</strong> experiments, a JAI CV-M4<br />

2/3” monochrome CCD-camera for progressive scan with a bellow <strong>and</strong> a Nikon 60mm objective 2.8<br />

diaphragm was used. For back light illumination either a VLP CIS-50/50-R-24 red LED or a Schott<br />

KL 1500 electronic microscope illumination was used. The CCD camera was connected to a PC via<br />

a PcDIG LVDS frame grabber (32-bit). The pictures were recorded <strong>and</strong> analysed using Image Pro<br />

Plus S<strong>of</strong>tware Version 4.51 (mediaCybernetics © ). All parts <strong>of</strong> the imaging system were supplied by<br />

Weiss Imaging <strong>and</strong> Solutions GmbH, 85232 Günding , Germany. Figure 4.9 gives a summary <strong>of</strong> all<br />

parts <strong>of</strong> the levitation system except for the outer acrylic glass container <strong>and</strong> the additional heater.<br />

In Figure 4.10 the whole set-up can be seen.<br />

CCD-camera<br />

bellows<br />

to PC with<br />

imaging s<strong>of</strong>tware<br />

levitation chamber<br />

macrolens<br />

sample<br />

reflector<br />

micrometer<br />

screw<br />

transducer<br />

piezo-crystal<br />

back light<br />

illumination<br />

micrometer<br />

screw<br />

Controlled<br />

evaporation<br />

mixer<br />

CEM<br />

air flow<br />

liquid flow<br />

Figure 4.9: Sketch <strong>of</strong> the composition <strong>of</strong> all parts <strong>of</strong> the acoustic levitation system used in this work, except <strong>of</strong> the outer<br />

acrylic glass container <strong>and</strong> the additional heater.


CHAPTER 4 MATERIALS AND METHODS 85<br />

(6)<br />

(4)<br />

(5)<br />

(2)<br />

(1)<br />

Figure 4.10: Complete set-up <strong>of</strong> the acoustic levitation system. (1) outer acrylic glass chamber;<br />

(2) acoustic levitator with inner acrylic glass chamber <strong>and</strong> LED; (3) external heating system; (4) CCD-camera system<br />

with bellow <strong>and</strong> macro lens; (5) controlled evaporation mixer with liquid flow <strong>and</strong> mass flow controller; (6) pressure<br />

reducer; (7) PC with imaging s<strong>of</strong>tware; (8) power supply <strong>and</strong> control unit for the acoustic levitator (left) <strong>and</strong> the CEM<br />

(right).<br />

(3)<br />

(7)<br />

(8)


86 MATERIALS AND METHODS<br />

4.2 Methods<br />

4.1.4 Acoustic Levitation<br />

Camera system <strong>and</strong> imaging s<strong>of</strong>tware<br />

To determine the evaporation rate <strong>of</strong> a single <strong>droplet</strong> dried in an acoustic levitator by measuring the<br />

decrease in diameter or by change in position <strong>of</strong> the centre <strong>of</strong> mass, the camera system <strong>and</strong> the<br />

imaging s<strong>of</strong>tware had to be calibrated. For every adjustment <strong>of</strong> the macro lens <strong>and</strong> the bellow the<br />

system was calibrated using a microscope slide with a micrometer. The calibration was verified<br />

measuring levitated spheres <strong>of</strong> polypropylene with a diameter <strong>of</strong> 1500µm, 1850µm <strong>and</strong> 2000µm ±<br />

30µm (Kugel Pompel, 1160 Vienna, Austria) (Spherotech GmbH; 36041 Fulda; Germany). The<br />

experiments were performed with the largest magnification still giving a focused image. The shutter<br />

<strong>of</strong> the camera was set to 1/3000 seconds. Pictures were taken every 0.25, 0.5, 0.75, 1.0, 2.0, 3.0, 4.0<br />

<strong>and</strong> 5.0 s depending on the rate <strong>of</strong> evaporation. The horizontal, vertical <strong>and</strong> Ferret diameter as well<br />

as the aspect ratio <strong>and</strong> the position <strong>of</strong> the centre <strong>of</strong> mass was measured for every picture. The data<br />

were collected from the imaging s<strong>of</strong>tware <strong>and</strong> transferred to Micros<strong>of</strong>t Excel © for further analysis.<br />

Acoustic levitator<br />

The SPL <strong>of</strong> the acoustic levitator was calibrated using a method reported by Rensink [Rensink<br />

2004]. The distance between transducer <strong>and</strong> reflector was initially set to 2.8 mm at ambient<br />

conditions. Then an n-decane <strong>droplet</strong> with an initial volume >2.0 µl was levitated <strong>and</strong> recorded by<br />

the imaging s<strong>of</strong>tware until the volume decreased to exactly 2.0 µl. At this point the magnitude <strong>of</strong><br />

the SPL, adjustable at the power supply, was set to a value such that the aspect ratio <strong>of</strong> the n-decane<br />

<strong>droplet</strong> was 1.5. For a more stable levitation the distance between transducer <strong>and</strong> reflector was<br />

increased by 0.7 mm. This adjustment corresponds to an effective SPL <strong>of</strong> exactly 162.47 dB at<br />

ambient conditions at a temperature <strong>of</strong> about 20°C.<br />

Controlled evaporation mixer (CEM)<br />

The settings <strong>of</strong> the CEM for the desired air flow velocity, the temperature <strong>and</strong> relative humidity<br />

inside the process chamber were determined using Fluidat © on the net V1.12/5.59 [Bronkhorst<br />

2005]. Additionally, the external heater was set to the same temperature to heat the outer acrylic<br />

glass container. Before starting any levitation experiments the system was equilibrated for 1 h. The<br />

temperature <strong>and</strong> the relative humidity inside the process chamber were verified using a Testo 605-


CHAPTER 4 MATERIALS AND METHODS 87<br />

H1 Dew Point Hygrometer. The temperature inside the outer acrylic glass container was determined<br />

using a Testo Mini-Thermometer 150°C (Testo AG; 79853 Lenzkirch, Germany).<br />

Levitation procedure<br />

The sample solvents or solutions used were first filtered through a 0.2 µm filter (Schleicher &<br />

Schull FP 30/0.2 CA-S, sterile, non pyrogenic) <strong>and</strong> then filled into 2µl Clear Crimp Seal Vials<br />

(Supelco; Bellefonte PA, USA) sealed with 11 mm PTFE Liners (Supelco; Bellefonte PA, USA).<br />

Vials containing <strong><strong>protein</strong>s</strong> were put on ice <strong>and</strong> stored in a fridge at 4°C. Immediately before the<br />

experiment the imaging s<strong>of</strong>tware was started. A <strong>droplet</strong> <strong>of</strong> the sample <strong>of</strong> between 1.0 <strong>and</strong> 2.5µl was<br />

positioned into the middle pressure node using a Hamilton 75 ASN 23s 5 µl or a 710 ASN 23s 10 µl<br />

gas chromatography syringe (Hamilton Bonaduz AG; 7402 Bonaduz, Switzerl<strong>and</strong>). During the<br />

<strong>drying</strong> experiments, pictures were taken after a predefined time interval until the <strong>droplet</strong> had either<br />

evaporated completely, dropped out <strong>of</strong> the st<strong>and</strong>ing acoustic wave, or, in the case <strong>of</strong> solutions, the<br />

particle formed did not change shape <strong>and</strong> position for more than 5 minutes. To determine the<br />

evaporation rate <strong>of</strong> the liquid sample the data was analysed according to Chapter 3.3.3.<br />

4.1.5 Maximum bubble pressure tensiometry<br />

Measurement <strong>of</strong> the dynamic interfacial tension was performed using a Lauda MPT2 maximum<br />

bubble pressure tensiometer (Lauda Dr. R. Wobser GmbH, 97955 Lauda-Konigsh<strong>of</strong>en, Germany).<br />

It had a dynamic time range <strong>of</strong> 0.0001 to 100 seconds. The set-up <strong>of</strong> the tensiometer is shown in<br />

Figure 4.11.<br />

Figure 4.11: Lauda MPT 2 maximum bubble<br />

pressure tensiometer:<br />

(1) pump;<br />

(2) filter unit;<br />

(3) measurement volume;<br />

(4) bubble detector;<br />

(5) test vessel<br />

(6) temperature cleading;<br />

(7) mount with capillary;<br />

(8) control unit;<br />

(9) personal computer.<br />

[Lauda 2005]


88 MATERIALS AND METHODS<br />

To determine the dynamic surface tension <strong>of</strong> a liquid, an air<br />

stream is blown through a thin capillary into the liquid<br />

forming bubbles. The pressure at the entrance <strong>of</strong> the<br />

capillary, the volume <strong>of</strong> air flowing through the capillary <strong>and</strong><br />

the time necessary for the formation <strong>of</strong> the bubble are<br />

measured. Depending on the flow volume per time either a jet<br />

or individual bubbles are produced. If the pressure is plotted<br />

against the flow volume per time, curves with different slopes<br />

are obtained. The linear area with the largest slope represents Figure 4.12: Pressure <strong>of</strong> air emerging<br />

the jet area where the air flows laminar through the capillary into water plotted against the air flow rate<br />

to determine the transition between<br />

according to the Hagen-Poiseuille law. If the flow is reduced bubble <strong>and</strong> jet regime [Lauda 2005]<br />

sufficiently, the jet reduces to individual bubbles. The pressure inflating the bubbles remains<br />

independent <strong>of</strong> the flow volume, but depends on their interfacial tension <strong>and</strong> their radius. This<br />

relationship is described by the Laplace equation. The transition between jet range <strong>and</strong> bubble range<br />

leads to a change in the slope <strong>of</strong> the pressure / flow rate curve, which is used to calculate the critical<br />

flow, L C , <strong>and</strong> the corresponding critical pressure, p C (Figure 4.12).<br />

Within the bubble range, the pressure at the outlet <strong>of</strong> the capillary where the bubble is<br />

formed increases from an initial pressure, p initial , to the maximum pressure, p max , (Figure 4.13;<br />

phase 1-3). This maximum pressure is the same pressure as the constant pressure at the inlet <strong>of</strong> the<br />

capillary, p MPT 2 , <strong>and</strong> it occurs when the radius, r , <strong>of</strong> the bubble at the outlet <strong>of</strong> the capillary is<br />

equal to the radius, r cap , <strong>of</strong> the capillary itself (Figure 4.13; phase 3). Shortly after the maximum,<br />

the pressure decreases to a minimum (Figure 4.13; phase 4-5). At this point, the bubble detaches<br />

from the capillary <strong>and</strong> a new bubble is formed (Figure 4.13; phase 5). Using the maximum bubble<br />

pressure, p max , the surface tension <strong>of</strong> the liquid corresponding to the surface age t <strong>of</strong> the bubble is<br />

calculated. The surface age is the time from the beginning <strong>of</strong> the bubble formation to its radius, r cap .<br />

At this point the bubble has the shape <strong>of</strong> a hemisphere. The time necessary to inflate the bubble<br />

from the hemispherical shape to detachment is called dead time, t d . According to Faineman [Lauda<br />

2000] it is calculated by<br />

Equation 4.1 t = t p − td<br />

Equation 4.2<br />

t<br />

d<br />

t<br />

=<br />

p<br />

⋅<br />

( p ⋅ L)<br />

L<br />

C<br />

C<br />

⋅ p


CHAPTER 4 MATERIALS AND METHODS 89<br />

Capillary<br />

1<br />

2<br />

3<br />

4 5<br />

Rejection surface<br />

Figure 4.13: Phases <strong>of</strong> bubble formation <strong>and</strong> internal bubble pressure [Lauda 2000]<br />

10.0 ml aqueous solution <strong>of</strong> the desired concentration was investigated using capillary 41 with a<br />

radius r cap <strong>of</strong> 0.075 mm <strong>and</strong> an immersion depth <strong>of</strong> 2.5 mm. The flow rate was decreased stepwise<br />

from 100 mm 3 /s to 2.5 mm 3 /s. All experiments were performed at 25.0 ± 0.5°C using a Julabo UC-<br />

5B water bath (Julabo GmbH; 77960 Seelbach, Germany). The surface excess <strong>of</strong> the dissolved<br />

sample substance was calculated using the Gibbs Equation [Fuhrling 2004; Hiemenz 1986]<br />

Equation 4.3<br />

4.1.6 Ring tensiometer<br />

pressure<br />

1 ⎛ dγ<br />

⎞<br />

Γ = − ⋅ ⎜ ⎟<br />

R ⋅ T ⎝ d ln c ⎠<br />

Measurement <strong>of</strong> the equilibrium interfacial tension was<br />

performed using a Kruss Digital Tensiometer K10ST (Kruss<br />

GmbH; 22452 Hamburg, Germany) with platinum-iridium<br />

ring. It had a range from 5 to 90mN/m <strong>and</strong> an accuracy <strong>of</strong><br />

±0.1mN/m. The ring is submerged below the interface <strong>and</strong><br />

subsequently raised upwards until contact with the surface is<br />

registered. Then the sample is lowered so that the liquid film<br />

produced beneath the ring is stretched (Figure 4.14). As the<br />

film is stretched, a maximum force is experienced which is<br />

T<br />

1<br />

2<br />

3<br />

4<br />

t<br />

5<br />

time<br />

t p<br />

t d<br />

Figure 4.14: Sketch <strong>of</strong> the ring method<br />

[Kruss 2005]


90 MATERIALS AND METHODS<br />

recorded in the measurement. At the maximum the force vector Fmax r<br />

is exactly parallel to the<br />

direction <strong>of</strong> motion. At this moment the contact angel θ is zero. Figure 4.15 illustrates the change<br />

in force as the distance <strong>of</strong> the ring from the surface increases [Kruss 2005].<br />

Figure 4.15: Change <strong>of</strong> force with ring distance above the interface <strong>of</strong> the liquid sample [Kruss 2005]<br />

The calculation <strong>of</strong> the resulting interfacial tension is made according to the following equation<br />

Equation 4.4<br />

F [mN]<br />

r r<br />

Fmax − FG<br />

γ =<br />

l ⋅ cosθ<br />

with the wetted length l <strong>of</strong> the platinum-iridium ring <strong>and</strong> the weight FG r <strong>of</strong> the volume <strong>of</strong> liquid<br />

lifted [Kruss 2005]. Measurements <strong>of</strong> the equilibrium interfacial tension were performed with<br />

20.0ml <strong>of</strong> aqueous solution at a temperature <strong>of</strong> 25.0 ± 0.5°C using a Julabo UC-5B water bath<br />

(Julabo GmbH; 77960 Seelbach, Germany). Each solution was equilibrated for 180 minutes before<br />

measurement.<br />

4.1.7 Spray-Drying<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

F max<br />

F 1<br />

Lamella<br />

breaks<br />

0.0<br />

0.0 1.0 2.0 3.0 4.0 5.0 6.0<br />

Distance above surface [mm]<br />

Spray-<strong>drying</strong> experiments were performed using a Buchi Mini Spray Dryer B-191 (Buchi AG; 9230<br />

Flawil, Switzerl<strong>and</strong>) <strong>and</strong> a Niro Mobile Minor (Niro A/S; 2860 Soeborg, Denmark). The basic setup<br />

<strong>of</strong> the Buchi B-191 is shown in Figure 4.16. Aqueous solutions in the desired concentration were<br />

prepared in water p.a. <strong>and</strong> spray-dried under st<strong>and</strong>ard conditions in a temperature range comparable<br />

F 3


CHAPTER 4 MATERIALS AND METHODS 91<br />

to the levitator experiments. These conditions can be seen in Table 4.6. For the experiments carried<br />

out with the Buchi B-191 an improved cyclone was used for a better powder recovery [Maury<br />

2005].<br />

Table 4.6: St<strong>and</strong>ard conditions <strong>of</strong> the spray-<strong>drying</strong> experiments.<br />

Figure 4.16: Sketch <strong>of</strong> the Buchi B-191<br />

(1) air inlet;<br />

(2) inlet air heating element;<br />

(3) two-component nozzle;<br />

(4) cyclone separator;<br />

(5) aspirator;<br />

(6) thermocouple for Tinlet;<br />

(7) thermocouple for Toutlet;<br />

(8) product collector;<br />

(9) peristaltic pump;<br />

[Büchi 2005]<br />

Buchi Mini Spray-Dryer B-191 Niro Mobile Minor<br />

Inlet temperature 100°C 115°C<br />

Outlet temperature 60°C 60°C<br />

Drying air flow rate 54 m 3 /h -<br />

Atomization two-fluid nozzle two-fluid nozzle<br />

Atomization flow rate 700 liters/h 2 bar<br />

Nozzle cooling yes no<br />

Nozzle purification yes (4.5 bar) no<br />

Liquid feed 3.0 ml/min 15.0 ml/min<br />

Volume <strong>of</strong> solution 15.0 ml 500 ml<br />

Volume <strong>of</strong> collect vessel 7.1 ml 1000 ml<br />

Flush time after spray-<strong>drying</strong> 5 min 5 min


92 MATERIALS AND METHODS<br />

4.1.8 Karl-Fischer Titration<br />

The residual moisture content <strong>of</strong> the spray-dried <strong>formulations</strong> was determined using a Mitsubishi<br />

Moisture Meter CA-06 Coulometric <strong>and</strong> a Mitsubishi Water Vaporizer VA-06. First the weight <strong>of</strong><br />

an empty glassy sample holder <strong>and</strong> its glassy closure was determined by a Mettler Toledo AT261<br />

Delta Range © analytical balance. The sample holders were transferred into a dry air-purged glove<br />

box. The humidity inside was ≤3.0% all time at room temperature. Samples <strong>of</strong> 80 to 120mg <strong>of</strong> the<br />

spray-dried powder were filled into the sample holders with subsequent sealing still inside the glove<br />

box. Then the samples were transferred into the glass boat pick-up <strong>of</strong> the moisture meter by<br />

connecting the sample holder to the water vaporising unit. Afterwards the glass boat pick-up was<br />

pushed automatically into the heating unit. All measurements were taken at a temperature <strong>of</strong> 130°C,<br />

an initial rate ≤10µg H2O per minute <strong>and</strong> a N2 gas stream <strong>of</strong> 200ml/min. Formulations with a glass<br />

transition temperature lower than 130°C (data obtained from literature [Labuza et al. 1992]), were<br />

measured at a temperature 10 to 20°C lower than their published glass transition temperature. The<br />

percentage <strong>of</strong> moisture content was calculated by using the reweighed sample holders. All<br />

measurements were performed in triplicate.<br />

4.1.9 Enzyme Activity Assay <strong>of</strong> Catalase<br />

Catalase catalyses the decomposition <strong>of</strong> hydrogen peroxide H2O2 to water <strong>and</strong> oxygen according to<br />

Figure 4.4. One unit enzyme will decompose 1.0 µmole <strong>of</strong> H2O2 per minute at pH 7.0 at 25°C,<br />

while the concentration <strong>of</strong> H2O2 decreases from 10.3 to 9.2 mM. The rate <strong>of</strong> disappearance <strong>of</strong> H2O2<br />

is followed by observing the decrease in the absorbance at 240 nm, suggested by Beers <strong>and</strong> Sizer<br />

[Beers 1952]. A PerkinElmer Lambda 25 UV/VIS spectrometer connected to a PC with<br />

PerkinElmer UV WinLab V 5.0 s<strong>of</strong>tware (PerkinElmer LAS GmbH; 63100 Rodgau, Germany) was<br />

used. At first, a 50mM potassium phosphate buffer (reagent A) was prepared using potassium<br />

phosphate monobasic <strong>and</strong> deionised water <strong>and</strong> adjusted to pH 7.0 using 1 M KOH solution. Reagent<br />

A acted as blank solution for calibration <strong>of</strong> the spectrometer at 240 nm. Then a 0.0036% (w/w)<br />

hydrogen peroxide solution (reagent B) was prepared in reagent A as substrate solution <strong>and</strong><br />

equilibrated at 25°C. Reagent B was used when the A240nm was between 0.550 <strong>and</strong> 0.520. Hydrogen<br />

peroxide was added to increase, or reagent A was added to decrease the absorption. Immediately<br />

before use, the catalase solution (reagent C) containing 50 to 100 units enzyme per ml was prepared<br />

in cold reagent A. Catalase was completely dissolved before assaying. In a quartz cuvette 2.9 ml <strong>of</strong><br />

the substrate solution reagent B was mixed with 0.1 ml <strong>of</strong> enzyme solution reagent C by inversion


CHAPTER 4 MATERIALS AND METHODS 93<br />

<strong>and</strong> the time t decrease required for the A240nm to decrease from 0.45 to 0.40 absorbance units was<br />

recorded. The activity <strong>of</strong> catalase was calculated by<br />

Equation 4.5<br />

units / ml enzyme =<br />

t<br />

3.<br />

45⋅<br />

df<br />

⋅ 0.<br />

1<br />

decrease<br />

The factor 3.45 corresponds to the decomposition <strong>of</strong> 3.45 micromoles <strong>of</strong> hydrogen peroxide in a<br />

3.0 ml reaction mixture producing a decrease in the A240nm from 0.45 to 0.40 absorbance units, df<br />

is the dilution factor, <strong>and</strong> the factor 0.1 holds for the volume in millilitre <strong>of</strong> the enzyme used. The<br />

enzyme activity refers to the mass <strong>of</strong> solid or <strong>protein</strong>, <strong>and</strong> is given by<br />

Equation 4.6<br />

Equation 4.7<br />

units / ml enzyme<br />

units / mg solid =<br />

mg solid / ml enzyme<br />

units /ml enzyme<br />

units /mg <strong>protein</strong><br />

=<br />

mg <strong>protein</strong> /ml enzyme<br />

The final assay concentrations in a 3.0 ml reaction mix were 50 mM potassium phosphate, 0.035%<br />

(w/w) hydrogen peroxide <strong>and</strong> 5-10units catalase. Untreated catalase was used as reference with an<br />

activity <strong>of</strong> 100%. All measurements were performed in triplicate.<br />

4.1.10 Kinematic viscosity<br />

The kinematic viscosity was measured using a Schott Ubbelohde Viscometer Type 501 13/Ic<br />

(Schott GmbH; 55014 Mainz, Germany) with a capillary diameter <strong>of</strong> 0.84 mm. Before use it was<br />

cleaned with 15% H2O2 <strong>and</strong> 15% HCl <strong>and</strong> rinsed with deionised water. When the viscometer was<br />

completely dry <strong>and</strong> dust free, 15 ml <strong>of</strong> the liquid sample were transferred through the filling tube<br />

into the reservoir. A vacuum was applied to the capillary tube with simultaneous closing <strong>of</strong> the<br />

venting tube. This caused the subsequent filling <strong>of</strong> the reference level vessel, the capillary tube, the<br />

measuring sphere <strong>and</strong> the pre-run sphere. Then the suction was discontinued <strong>and</strong> the venting tube<br />

was opened again. This caused the liquid column to separate at the lower end <strong>of</strong> the capillary ant to<br />

form the suspended level at the dome-shaped top part. The time it took the leading edge <strong>of</strong> the<br />

meniscus <strong>of</strong> the sample to descend from the upper edge <strong>of</strong> the upper timing mark to the upper edge<br />

<strong>of</strong> the lower timing mark was measured [Schott 2005]. The kinematic viscosity is calculated by<br />

Equation 4.8 ν<br />

= K ⋅ ( t − HC)<br />

viscometer


94 MATERIALS AND METHODS<br />

K viscometer is the viscometer constant <strong>of</strong> approximately 0.03 <strong>and</strong> HC is the kinetic energy correction<br />

dependent on the viscometer <strong>and</strong> flow time. The values for HC can be found in the instruction<br />

manual <strong>of</strong> the viscometer [Schott 2005]. All measurements were performed in triplicate.<br />

4.1.11 Liquid density<br />

Density measurement was performed with a Blaur<strong>and</strong> © Glass Pyknometer (VWR International<br />

GmbH; 64295 Darmstadt, Germany). The pyknometer had a volume <strong>of</strong> 5.370 ± 0.001 cm 3 .The<br />

empty weight <strong>of</strong> the pyknometer <strong>and</strong> its glassy closure with a small capillary inside was determined<br />

using a Mettler Toledo AT261 DeltaRange © analytical balance. The liquid sample, equilibrated at<br />

25.0 ± 0.5°C using a Julabo UC-5B water bath (Julabo GmbH; 77960 Seelbach, Germany) was then<br />

filled into the pyknometer up to a height, that liquid effused out <strong>of</strong> the capillary <strong>of</strong> the glassy<br />

closure when closed. The outside <strong>of</strong> the pyknometer was cleaned before reweighing <strong>of</strong> the<br />

analytical balance. The density was calculated by<br />

Equation 4.9<br />

ρ<br />

liquid<br />

w filled − w<br />

=<br />

V<br />

pyknometer<br />

empty<br />

All measurements were performed in triplicate.<br />

4.1.12 Scanning electron microscopy (SEM)<br />

Particle size <strong>and</strong> morphology <strong>of</strong> all dried particles were examined using an Amray 1810T scanning<br />

electron microscope (Amray; Bedford, USA) at 20 kV. Samples were fixed on aluminium stubs<br />

(G301, Plano) with self-adhesive films <strong>and</strong> spluttered with gold at 20 mA / 5 kV (Hummer JR<br />

Technics; Munich, Germany) for 1.5 min.


CHAPTER 5 RESULTS AND DISCUSSION 95<br />

5 Results <strong>and</strong> discussion<br />

5.1 Pre-liminary levitation tests<br />

5.1.1 Size measurement<br />

The calibration <strong>of</strong> the imaging s<strong>of</strong>tware is performed with a microscope slide with micrometer <strong>and</strong><br />

verified using polypropylene spheres with a diameter <strong>of</strong> 1500 µm, 1850 µm <strong>and</strong> 2000 µm ± 30µm.<br />

The precision <strong>of</strong> size measurements <strong>of</strong> a sample within an acoustic levitation system depends not<br />

only on the reproduction scale, the resolution <strong>of</strong> the CCD-chip, the intensity <strong>of</strong> the back<br />

illumination <strong>and</strong> the calibration <strong>of</strong> the system, but also on the dynamics <strong>of</strong> the levitated <strong>droplet</strong> or<br />

particle. Vibrations, air flow or a not ideal adjusted st<strong>and</strong>ing acoustic sound wave can lead to a<br />

deflection <strong>of</strong> the sample out <strong>of</strong> its rest position resulting in oscillations [Schnitzler 1998;<br />

Tuckermann 2002]. To eliminate out-<strong>of</strong>-focus pictures the levitation system is adjusted very<br />

precisely <strong>and</strong> the shutter time <strong>of</strong> the camera is 1/5000 s. The intensity <strong>of</strong> illumination is chosen not<br />

to obscure the outline <strong>of</strong> the levitated sample. Levitated liquid samples are also influenced by<br />

interior oscillation resulting in a continuous deformation <strong>of</strong> the <strong>droplet</strong> shape with change <strong>of</strong> the<br />

aspect ratio. This kind <strong>of</strong> <strong>droplet</strong> oscillation leads to an additional factor complicating the precise<br />

measurement <strong>of</strong> the size <strong>of</strong> liquid <strong>droplet</strong>s. Figure 5.1 shows the larger inaccuracy <strong>of</strong> measurement<br />

<strong>of</strong> a levitated Migliol N812 <strong>droplet</strong> due to the internal oscillation compared to the smaller<br />

inaccuracy <strong>of</strong> a rigid polypropylene sphere. The difference between horizontal <strong>and</strong> vertical diameter<br />

<strong>of</strong> the Migliol N812 <strong>droplet</strong> is caused by the different values <strong>of</strong> the axial to radial levitation force<br />

[Lierke 1996]. The larger st<strong>and</strong>ard deviation <strong>of</strong> the vertical diameters <strong>of</strong> the Migliol <strong>droplet</strong>s shows<br />

that the internal oscillations occur in the axial direction. Another important factor was the influence<br />

<strong>of</strong> the intensity <strong>of</strong> back illumination on size measurement <strong>of</strong> the levitated <strong>droplet</strong> or particle. An<br />

illumination too intense caused an obscuration <strong>of</strong> the outline <strong>of</strong> the sample as shown in Figure 5.1.<br />

The stronger the illumination, the larger is the decrease in diameter determined by the imaging<br />

s<strong>of</strong>tware. This effect is tested with a levitated polypropylene sphere at different intensity <strong>of</strong> back<br />

light illumination. The first three levels do not cause a substantial change in the measured horizontal<br />

diameter, but level 4 to 6 obscures the outline increasingly (Figure 5.3 a). This leads to a smaller<br />

determined diameter as well as to a larger st<strong>and</strong>ard deviation <strong>of</strong> the diameter means (Figure 5.3 b).<br />

Therefore level 2 or level 3 are chosen for the back illumination <strong>of</strong> the levitation experiments.


96 RESULTS AND DISCUSSION<br />

Counts [-]<br />

Counts [-]<br />

200<br />

150<br />

100<br />

50<br />

0<br />

1830 1835 1840 1845 1850 1855 1860 1865<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

(a)<br />

(b)<br />

Diameter [µm]<br />

1500 1550 2050 2100 2150<br />

Diameter [µm]<br />

Horizontal diameter<br />

Vertical diameter<br />

Gauss-Fit horizontal diameter<br />

X mean = 1848.3 µm<br />

SD = 2.8 µm<br />

Gauss-Fit vertical diameter<br />

X mean = 1849,1 µm<br />

SD = 3.2 µm<br />

Horizontal diameter<br />

Vertical diameter<br />

Gauss-Fit vertical diameter<br />

X mean = 2118.7 µm<br />

SD = 12.5 µm<br />

Gauss-Fit horizontal diameter<br />

X = 1530.8 µm<br />

mean<br />

SD = 17.8 µm<br />

Figure 5.1: Statistics <strong>of</strong> the recorded horizontal <strong>and</strong> vertical diameter at 25.3°C <strong>and</strong> 20.4% humidity. (a) Levitated<br />

polypropylene sphere. (b) Levitated Migliol N812 <strong>droplet</strong>.<br />

Level 1 Level 2 Level 3<br />

Level 4 Level 5 Level 6<br />

Figure 5.2: Pictures <strong>of</strong> a levitated poly- propylene sphere at 25°C with different intensity <strong>of</strong> the illumination.


CHAPTER 5 RESULTS AND DISCUSSION 97<br />

Horizontal diameter [µm]<br />

Counts [-]<br />

Figure 5.3: (a) Diagram <strong>of</strong> the diameter determined by the calibrated imaging s<strong>of</strong>tware in relation to the intensity <strong>of</strong> the<br />

back light illumination; (b) Direct statistical correlation between the back illumination at level 2 <strong>and</strong> at level 6.<br />

5.1.2 Injectable volume<br />

To levitate a liquid sample at the pressure node <strong>of</strong> the ultrasonic levitation system, gas<br />

chromatography syringes with different volume are used depending on the viscosity <strong>of</strong> the sample<br />

solution. Due to deformation <strong>of</strong> the levitated <strong>droplet</strong>, volumes <strong>and</strong> surfaces from the recorded<br />

pictures are calculated assuming an oblate ellipsoid using the horizontal radius rhor <strong>and</strong> vertical<br />

radius r vert .<br />

1860<br />

1840<br />

1820<br />

1800<br />

1780<br />

1760<br />

1740<br />

1720<br />

1700<br />

1680<br />

(a)<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

(b)<br />

1 2 3 4 5 6<br />

Back illumination level [-]<br />

1670 1680 1690 1700 1840 1850<br />

Horizontal diameter [µm]<br />

4 2<br />

Equation 5.1 V<strong>droplet</strong> = ⋅ π⋅<br />

rhor<br />

⋅ rvert<br />

3<br />

Back illumination level 2<br />

Back illumination level 6<br />

Gaussian Fit (level 2)<br />

X mean = 1848.2 µm<br />

SD = 2.5 µm<br />

Gaussian Fit (level 6)<br />

X mean = 1685.6 µm<br />

SD = 8.7 µm


98 RESULTS AND DISCUSSION<br />

2<br />

2<br />

2 rvert<br />

⎛1<br />

+ E ⎞<br />

r<br />

Equation 5.2 S <strong>droplet</strong> = 2 π⋅<br />

rhor<br />

+ π⋅<br />

⋅ ln⎜<br />

⎟ with E = 1-<br />

2<br />

E ⎝1<br />

− E ⎠<br />

r<br />

First it is necessary to determine if the injected volume corresponds to the volume calculated from<br />

the pictures recorded by the imaging s<strong>of</strong>tware. Different volumes <strong>of</strong> water, ethanol <strong>and</strong> solutions<br />

with low viscosity using the 5µl Hamilton 75 ASN syringe <strong>and</strong> maltodextrin solution > 200 mg/ml<br />

using the 10µl Hamilton 710 ASN syringe, are levitated at ambient conditions. The injected volume<br />

is determined from the recorded pictures using Equation 5.1. Figure 5.4 shows the relationship<br />

between the injected <strong>and</strong> calculated volume.<br />

Calculated volume [µl]<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

(a)<br />

Water<br />

Kinematic viscosity = 1.00mm 2 /s<br />

1.5 2.0 2.5 3.0<br />

Injected volume [µl]<br />

Figure 5.4: Diagram <strong>of</strong> the volume determined by the imaging s<strong>of</strong>tware versus the injected volume for (a) water <strong>and</strong><br />

(b) maltodextrin solution 20% (w/V) at 25°C.<br />

The smallest injectable <strong>droplet</strong> that separates from the syringe without any problems has a volume<br />

<strong>of</strong> 1.0 µl for both substances. It can be seen that the determined volume matches almost exactly the<br />

injected one, independent <strong>of</strong> the viscosity <strong>of</strong> the liquid sample <strong>and</strong> the syringe.<br />

Secondly, the influence <strong>of</strong> the effective sound pressure level on the precision <strong>of</strong> the size<br />

measurement is tested. 2 µl <strong>of</strong> water are levitated at different effective SPL, determined via the<br />

method <strong>of</strong> aspect ratio explained in Chapter 3 [Yarin et al. 1998]. An increase in the sound pressure<br />

level leads to a strong deformation <strong>of</strong> the <strong>droplet</strong>. This results in an increase <strong>of</strong> the aspect ratio <strong>and</strong><br />

the ellipticity, E , in Equation 5.2. Up to an effective SPL <strong>of</strong> 164.9 dB, corresponding to an aspect<br />

ratio <strong>of</strong> 1.26 for water at 25.1°C, the injected <strong>and</strong> calculated volume match very well (Figure 5.5).<br />

At larger SPLs <strong>and</strong> aspect ratios the determined volume continuously becomes smaller than the<br />

injected one. To overcome this effect <strong>of</strong> the SPL for later examinations, the volume is calculated<br />

Calulated volume [µl]<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

(b)<br />

vert<br />

hor<br />

Maltodextrin solution 20% (w/v)<br />

Kinematic viscosity = 3.75mm 2 /s<br />

1.5 2.0 2.5 3.0<br />

Injected volume [µl]


CHAPTER 5 RESULTS AND DISCUSSION 99<br />

using Equation 5.2 as the volume <strong>of</strong> a surface equivalent sphere. Figure 5.5 shows, that the SPL<br />

does not affect this kind <strong>of</strong> volume determination <strong>of</strong> the liquid sample.<br />

Determined volume [µl]<br />

2.2 Volume via Equation 5.1<br />

Volume <strong>of</strong> a surface<br />

equivalent sphere<br />

2.1 Aspect ratio<br />

2.0<br />

1.9<br />

1.8<br />

Figure 5.5: Diagram <strong>of</strong> the influence <strong>of</strong> the effective SPL on the volume determined by the imaging s<strong>of</strong>tware <strong>and</strong> the<br />

aspect ratio <strong>of</strong> the levitated <strong>droplet</strong> at 25.1°C <strong>and</strong> 26.8% relative humidity. Additionally, the volume <strong>of</strong> a surface<br />

equivalent sphere calculated using Equation 5.2 is plotted into the diagram.<br />

5.1.3 Heating <strong>of</strong> the process chamber<br />

163 164 165 166 167 168<br />

1.0<br />

169<br />

Effective SPL [dB]<br />

Oscillation <strong>of</strong> the ultrasonic transducer causes a heating <strong>of</strong> the process chamber <strong>of</strong> the levitation<br />

system. The increase in temperature inside can lead to an increase in evaporation rate <strong>of</strong> liquid<br />

<strong>droplet</strong>s <strong>and</strong> a decrease in evaporation time [Kastner 2001]. To investigate the influence <strong>of</strong> the<br />

effective SPL, the temperature inside the process chamber in the area <strong>of</strong> the middle pressure node<br />

was measured within the first 15 min after the ultrasonic levitator has been switched on at ambient<br />

conditions. Three different effective SPLs are investigated (Figure 5.6 a). The results show a<br />

dependence <strong>of</strong> the temperature development inside the levitator on the effective SPL. After<br />

10 to 15 min an equilibrium temperature is reached for all given SPLs. Measurements <strong>of</strong> the<br />

evaporation rate <strong>of</strong> solvents or solutions at ambient conditions need therefore at least an<br />

equilibration time <strong>of</strong> 15 min before starting the first experiment. The influence <strong>of</strong> the SPL on the<br />

temperature inside the process chamber with additional heating by the external heater <strong>and</strong><br />

ventilation using the CEM is also tested. The temperature <strong>of</strong> the external heater is set to either 40°C<br />

or 60°C. The conditions at the CEM are set to produce an air flow <strong>of</strong> 40 or 60°C with an orifice<br />

Reynolds number <strong>of</strong> 200 at a relative humidity <strong>of</strong> 0.1%. The results are shown in Figure 5.6 b. As<br />

stated by Kastner [2001] an additional heating with an external heater suppresses the influence <strong>of</strong><br />

the heat input due to the oscillation <strong>of</strong> the ultrasonic transducer. The presence <strong>of</strong> air flow is <strong>of</strong> no<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

Aspect ration [-]


100 RESULTS AND DISCUSSION<br />

account. The experiments also show a temperature fluctuation <strong>of</strong> only ± 0.1°C <strong>of</strong> the preset<br />

temperature value.<br />

Temperature in process chamber [°C]<br />

Temperature in process chamber [°C]<br />

26<br />

25<br />

24<br />

23<br />

22<br />

21<br />

20<br />

(a)<br />

62.0<br />

61.0<br />

60.0<br />

59.0<br />

40.5<br />

40.0<br />

39.5<br />

39.0<br />

(b)<br />

0 2 4 6 8 10 12 14 16<br />

Time [min]<br />

Effective SPL 163.7 dB<br />

Effective SPL 166.4 dB<br />

Effective SPL 167.9 dB<br />

Heater 40°C - SPL 166.3 - Ventilation 0.000 m/s<br />

Heater 40°C - SPL 166.3 - Ventilation 0.885 m/s<br />

Heater 60°C - SPL 166.9 - Ventilation 0.000 m/s<br />

Heater 60°C - SPL 166.9 - Ventilation 0.992 m/s<br />

0 2 4 6 8 10 12 14<br />

Time [min]<br />

Figure 5.6: (a) Diagram <strong>of</strong> the temperatures increase with time at different SPLs inside the process chamber <strong>of</strong> the<br />

acoustic levitation system at ambient conditions <strong>of</strong> approximately 20.4°C <strong>and</strong> 27.8% relative humidity. (b) Diagram <strong>of</strong><br />

the influence <strong>of</strong> the SPL on the temperature inside the process chamber <strong>of</strong> the levitation system in case <strong>of</strong> additional<br />

heating with an external heater <strong>and</strong> ventilation at 40°C <strong>and</strong> 60°C <strong>and</strong> a relative humidity <strong>of</strong> 0.1%.


CHAPTER 5 RESULTS AND DISCUSSION 101<br />

5.1.4 Parametric studies<br />

Levitation capabilities in a single-axis levitator are improved by using a curved reflector <strong>and</strong><br />

enlarging its section so as to levitate dense materials or liquids under large ventilation air streams<br />

[Xie 2001]. To describe the st<strong>and</strong>ing acoustic wave the amplitude distribution <strong>of</strong> an axially<br />

emerging Gaussian stream is used (Figure 5.7) [Schnitzler 1998].<br />

Figure 5.7: Axial dispersion <strong>of</strong> a st<strong>and</strong>ing<br />

acoustic wave in Gaussian approximation<br />

according to Schnitzler [Bauer et al. 1999;<br />

Schnitzler 1998]<br />

Prediction <strong>of</strong> the stability <strong>of</strong> the st<strong>and</strong>ing acoustic wave can be made, based on the geometry <strong>of</strong> the<br />

ultrasonic transducer <strong>and</strong> reflector. The stability depends on their distance, L R , <strong>and</strong> their radius <strong>of</strong><br />

curvature, R i [Tuckermann 2002]. The relation between both is given by<br />

Equation 5.3<br />

L<br />

gi = 1 −<br />

R<br />

R<br />

i<br />

Stable levitation conditions are given for < g g < 1.<br />

In the stability diagram shown in Figure 5.8,<br />

0 1 2<br />

all values <strong>of</strong> g i with stable levitation are marked by the hatched area.<br />

not stable<br />

2<br />

not stable<br />

semi-concentric<br />

flat-flat<br />

1<br />

0<br />

semi-confocal<br />

stable<br />

-2 -1 0<br />

stable<br />

-1<br />

1<br />

symmetricalconfocal<br />

2<br />

symmetricalconcentric<br />

-2<br />

not stable<br />

not stable<br />

Figure 5.8: Stability diagram <strong>of</strong> different transducer-reflector set-ups <strong>of</strong> an acoustic levitation system calculated using<br />

Equation 5.3 [Tuckermann 2002].


102 RESULTS AND DISCUSSION<br />

According to Schnitzler [1998] <strong>and</strong> Tuckermann [2002] the semi-confocal set-up is the least<br />

problematic, because its values are inside the stable levitation area <strong>and</strong> not at any <strong>of</strong> the boundaries.<br />

On these assumptions a semi-confocal set-up is built by manufacturing a reflector with a radius <strong>of</strong><br />

curvature <strong>of</strong> twice the distance between transducer <strong>and</strong> reflector L R (Figure 5.9 [Schnitzler 1998]).<br />

The advantage <strong>of</strong> the semi-confocal set-up compared to the flat-flat one is tested by levitating a<br />

water <strong>droplet</strong> at high ventilation air stream at 60°C <strong>and</strong> monitoring the vertical <strong>and</strong> horizontal<br />

position <strong>of</strong> the centre <strong>of</strong> mass <strong>of</strong> the <strong>droplet</strong>. Figure 5.10 shows the more stable levitation conditions<br />

using the semi-confocal set-up <strong>of</strong> the acoustic levitator at an air velocity <strong>of</strong> 2.0m/s than the flat-flat<br />

one. The semi-confocal set-up the <strong>droplet</strong> varies within an amplitude <strong>of</strong> only ± 100µm, whereas the<br />

amplitude <strong>of</strong> the flat-flat levitator was ± 500µm. Considering the vertical position, the difference<br />

between both set-ups is however not that large. The reason is the internal oscillation <strong>of</strong> the levitated<br />

<strong>droplet</strong> causing large changes <strong>of</strong> its vertical diameter <strong>of</strong> the <strong>droplet</strong>. This is also seen in Figure 5.3<br />

As a result, the imaging s<strong>of</strong>tware determined an “apparent” movement <strong>of</strong> the <strong>droplet</strong> centre <strong>of</strong> mass,<br />

even though the levitation is more stable than that using the flat-flat set-up <strong>of</strong> the levitation system.<br />

Figure 5.9: Sketch <strong>of</strong> the semi-confocal set-up <strong>of</strong> the ultrasonic levitator on the left [Schnitzler 1998] <strong>and</strong> picture <strong>of</strong> the<br />

ultrasonic reflector with a radius <strong>of</strong> curvature <strong>of</strong> twice the transducer-reflector distance one the right.


CHAPTER 5 RESULTS AND DISCUSSION 103<br />

Horizontal distance to initial position [µm]<br />

600<br />

400<br />

200<br />

0<br />

-200<br />

-400<br />

-600<br />

(a)<br />

Figure 5.10: Diagram comparing the semi-confocal with the flat-flat set-up <strong>of</strong> an acoustic levitator at 60.0°C <strong>and</strong> a<br />

relative humidity <strong>of</strong> 20.0%. (a) Levitation experiment in still air without ventilation. (b) To stress the levitated water<br />

<strong>droplet</strong> the ventilation air stream was set to 2.012m/s, equal to an orifice Reynolds number <strong>of</strong> 400.<br />

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

5.2.1 Data analysis<br />

Droplets <strong>of</strong> different initial size are levitated with the acoustic levitation system. It is to be expected<br />

that the evaporation process can be described by the d 2- law [Frohn 2000].<br />

Different initial <strong>droplet</strong> size<br />

Droplets <strong>of</strong> the same liquid but <strong>of</strong> different initial <strong>droplet</strong> size or volume have different evaporation<br />

lifetimes but the same evaporation rate. There are various possibilities to plot the data obtained from<br />

single <strong>droplet</strong> <strong>drying</strong> experiments. Figure 5.11 shows plots <strong>of</strong> the decrease in radius, mean ferret<br />

diameter, squared radius, squared mean ferret diameter, volume <strong>and</strong> surface <strong>of</strong> an evaporating water<br />

<strong>droplet</strong> with time. The plots <strong>of</strong> the squared radius, squared diameter or surface with time each yields<br />

an approximate straight line. On differentiating Equation 2.53, the change <strong>of</strong> <strong>droplet</strong> radius with<br />

time dt<br />

drS / is obtained as dr / dt −(<br />

β / 2r<br />

)<br />

S<br />

= . For constant evaporation coefficient β r the<br />

r<br />

S<br />

change <strong>of</strong> radius per time drS / dt depends only on the momentary radius r S . This explains the<br />

influence <strong>of</strong> the initial <strong>droplet</strong> size on the plots shown in Figure 5.11. If the d 2- law describes the<br />

evaporation process with sufficient accuracy, the influence <strong>of</strong> different initial sizes can be<br />

eliminated by dividing Equation 2.53 by the initial radius r S 0 . The evaporation process is then<br />

described by<br />

semi-confocal<br />

flat-flat<br />

0 20 40 60 80 100 120 140<br />

Time [sec]<br />

Vertical distance form initial point [µm]<br />

600<br />

400<br />

200<br />

0<br />

-200<br />

-400<br />

-600<br />

(b)<br />

semi-confocal<br />

flat-flat<br />

0 20 40 60 80 100 120 140<br />

Time [sec]


104 RESULTS AND DISCUSSION<br />

2<br />

rS<br />

t<br />

Equation 5.4 = 1 − β<br />

2<br />

r ⋅ 2<br />

r<br />

r<br />

S 0<br />

S 0<br />

The evaporation coefficient β r depends on the thermodynamic properties <strong>of</strong> the <strong>droplet</strong> liquid <strong>and</strong><br />

on <strong>drying</strong> air temperature <strong>and</strong> relative humidity [Frohn 2000]. Figure 5.12 shows two plots <strong>of</strong> the<br />

data from Figure 5.11 according to Equation 5.4.<br />

Droplet radius [µm]<br />

Mean Ferret diameter [µm]<br />

Droplet volume [µl]<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

0 200 400 600 800 1000 1200 1400<br />

Evaporation time [s]<br />

V 0 = 2.014 µl<br />

V 0 = 1.618 µl<br />

V 0 = 1.153 µl<br />

0<br />

0 200 400 600 800 1000 1200 1400<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

Evaporation time [s]<br />

V 0 = 2.014 µl<br />

V 0 = 1.618 µl<br />

V 0 = 1.153 µl<br />

0.0<br />

0 200 400 600 800 1000 1200 1400<br />

Evaporation time [s]<br />

V 0 = 2.014 µl<br />

V 0 = 1.618 µl<br />

V 0 = 1.153 µl<br />

0.0<br />

0 200 400 600 800 1000 1200 1400<br />

Figure 5.11: Different diagrams <strong>of</strong> <strong>droplet</strong> size as a function <strong>of</strong> evaporation time. The <strong>droplet</strong>s had different initial<br />

<strong>droplet</strong> radii <strong>and</strong> were levitated at a temperature <strong>of</strong> 25.0°C <strong>and</strong> relative humidity <strong>of</strong> 3.0%. The initial effective SPL was<br />

163.5 dB for the <strong>droplet</strong> with V0= 2.014 µl.<br />

Squared radius [mm 2 ]<br />

Squared mean Ferret diameter [mm 2 ]<br />

Droplet surface [mm 2 ]<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

Evaporation time [s]<br />

V 0 = 2.014 µl<br />

V 0 = 1.618 µl<br />

V 0 = 1.153 µl<br />

0.0<br />

0 200 400 600 800 1000 1200 1400<br />

8.0<br />

7.0<br />

6.0<br />

5.0<br />

4.0<br />

3.0<br />

2.0<br />

1.0<br />

Evaporation time [s]<br />

V 0 = 2.014 µl<br />

V 0 = 1.618 µl<br />

V 0 = 1.153 µl<br />

0.0<br />

0 200 400 600 800 1000 1200 1400<br />

Evaporation time [s]<br />

V 0 = 2.104 µl<br />

V 0 = 1.618 µl<br />

V 0 = 1.153 µl


CHAPTER 5 RESULTS AND DISCUSSION 105<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 200 400 600 800 1000 1200 1400<br />

(a)<br />

Figure 5.12: (a) (rS 2 / rS 0 2 ) as a fuction <strong>of</strong> evaporation time. (b) (rS 2 / rS 0 2 ) as a function <strong>of</strong> (t / rS 0 2 ). The water<br />

<strong>droplet</strong>s with different initial volume were levitated in an acoustic levitator at 25.0°C <strong>and</strong> a relative humidity <strong>of</strong> 3.0%.<br />

The initial effective SPL was 163.5 dB for the <strong>droplet</strong> with V0= 2.014 µl.<br />

The plots <strong>of</strong><br />

2<br />

rS rS<br />

0<br />

2 / as a function <strong>of</strong> time, t , (Figure 5.12 a) result in approximate straight lines<br />

from unity independent <strong>of</strong> the initial <strong>droplet</strong> size. From a linear regression for each measurement<br />

the slope<br />

β / <strong>of</strong> the curves was obtained. If<br />

2<br />

r rS 0<br />

Evaporation time [s]<br />

V 0 = 2.014 µl<br />

V 0 = 1.618 µl<br />

V 0 = 1.153 µl<br />

2<br />

rS rS<br />

0<br />

0.0<br />

0 500 1000 1500 2000 2500<br />

2 / is now plotted against the<br />

2<br />

t rS<br />

0<br />

/ (Figure<br />

5.12 b) the negative slope directly represents the evaporation coefficient β r [Frohn 2000]. Small<br />

differences between the measurements may be caused by slightly differing air conditions or<br />

differing internal oscillation <strong>of</strong> <strong>droplet</strong>s [Rensink 2004]. The evaporation coefficient β r<br />

characterizes the evaporation process for a given pure solvent <strong>droplet</strong> under conditions valid for the<br />

d 2 -law [Frohn 2000].<br />

The acoustically levitated <strong>droplet</strong>s <strong>of</strong> pure liquids can be described as oblate ellipsoid. Thus,<br />

a direct determination <strong>of</strong> the <strong>droplet</strong> radius was not possible. Three possible techniques to obtain the<br />

radius from the measured data are available. First, by calculating the radius <strong>of</strong> a volume equivalent<br />

sphere using Equation 5.1. Secondly, by calculating the radius <strong>of</strong> a surface equivalent sphere<br />

according to Equation 5.2. Thirdly, by direct determination <strong>of</strong> the mean ferret radius <strong>of</strong> the levitated<br />

<strong>droplet</strong> using the imaging s<strong>of</strong>tware. A comparison <strong>of</strong> the plots for all three radii is shown in Figure<br />

5.13.<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

(b)<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

V 0 = 2.014 µl<br />

V 0 = 1.618 µl<br />

V 0 = 1.153 µl


106 RESULTS AND DISCUSSION<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

Figure 5.13: Plot <strong>of</strong> the evaporation experiment <strong>and</strong> the evaporation coefficient <strong>of</strong> a levitated water <strong>droplet</strong> at a<br />

temperature <strong>of</strong> 25.0°C <strong>and</strong> a relative humidity <strong>of</strong> 3.0%. The radius was calculated <strong>of</strong> a surface equivalent sphere,<br />

volume equivalent sphere or the mean ferret diameter. The initial <strong>droplet</strong> volume was 2.014µl at an initial effective SPL<br />

<strong>of</strong> 163.5 dB.<br />

There is a close match <strong>of</strong> the evaporation coefficients calculated from the radius <strong>of</strong> a volume<br />

2<br />

equivalent sphere ( β = 0.<br />

000501mm<br />

/s ) <strong>and</strong> the radius <strong>of</strong> a surface equivalent sphere<br />

r<br />

2<br />

( β = 0.<br />

000505mm<br />

/s).<br />

The evaporation coefficient calculated from the directly determined mean<br />

r<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 500 1000 1500 2000 2500<br />

(a)<br />

2<br />

ferret diameter differs with β = 0.<br />

000470 mm /s . This is caused by fluctuations in the horizontal<br />

r<br />

<strong>and</strong> vertical diameters due to internal oscillation <strong>of</strong> the levitated <strong>droplet</strong>. The determination <strong>of</strong> the<br />

mean ferret diameter always leads to a larger radius than the other two methods, except for the case<br />

<strong>of</strong> a purely spherical <strong>droplet</strong>. A comparison between radii calculated by volume or surface always<br />

leads to almost identical results. Since large SPL leads to inaccuracy determining the actual volume<br />

<strong>of</strong> the <strong>droplet</strong> using Equation 5.1 (Figure 5.5) the data <strong>of</strong> all the following experiments were<br />

calculated using the radius <strong>of</strong> a surface equivalent sphere.<br />

Different <strong>droplet</strong> liquids<br />

Surface equivalent sphere<br />

Volume equivalent sphere<br />

Radius <strong>of</strong> the mean ferret diameter<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Different liquids have different vapour pressures, resulting in different evaporation coefficients β r<br />

under the same <strong>drying</strong> conditions [Frohn 2000]. Figure 5.14 shows the results for the evaporation <strong>of</strong><br />

levitated water, hexane <strong>and</strong> ethanol <strong>droplet</strong>s. The different behaviour <strong>of</strong> the three solvents is clearly<br />

shown by a different evaporation rate under the same <strong>drying</strong> conditions. The shortest <strong>drying</strong> time<br />

2<br />

<strong>and</strong> largest evaporation coefficient β = 0.<br />

008329 mm /s at 25.0°C <strong>and</strong> 3.0% relative humidity is<br />

r<br />

Evaporation coefficient β r [mm 2 /s]<br />

0.00051<br />

0.00050<br />

0.00049<br />

0.00048<br />

0.00047<br />

for hexane, whose vapour pressure at this temperature is p = 18891.<br />

8 Pa [EngineeringToolBox<br />

2005]. The evaporation coefficient <strong>of</strong> the ethanol <strong>droplet</strong> with a vapour pressure <strong>of</strong><br />

(b)<br />

vap<br />

surface volume ferret<br />

Type <strong>of</strong> calculated <strong>droplet</strong> radius


CHAPTER 5 RESULTS AND DISCUSSION 107<br />

2<br />

p = 7208.<br />

7 Pa under the same ambient conditions is β = 0.<br />

002136mm<br />

/s . The flattened course<br />

vap<br />

<strong>of</strong> the ethanol plot near the end <strong>of</strong> the experiment results from condensation <strong>of</strong> water vapour from<br />

the <strong>drying</strong> air onto the <strong>droplet</strong> surface. This effect <strong>and</strong> its influence on the evaporation <strong>of</strong> ethanol<br />

<strong>droplet</strong>s will be discussed in detail later. The longest evaporation time <strong>and</strong> also the smallest<br />

2<br />

evaporation coefficient is found for the <strong>drying</strong> <strong>of</strong> the water <strong>droplet</strong>, with β = 0.<br />

000505mm<br />

/s . The<br />

water vapour pressure at 25.0°C temperature is p = 3174.<br />

5 Pa [EngineeringToolBox 2005].<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 500 1000 1500 2000 2500<br />

vap<br />

Figure 5.14: Variation <strong>of</strong> (rs 2 /rS 0 2 ) with (t/rS 0 2 ) for three <strong>droplet</strong>s, consisting <strong>of</strong> water, ethanol <strong>and</strong> hexane levitated at a<br />

temperature <strong>of</strong> 25.0°C <strong>and</strong> a relative humidity <strong>of</strong> 3.0%. The initial effective SPL was 163.5dB for the water <strong>droplet</strong>,<br />

162.3dB for ethanol <strong>and</strong> 160.5dB for hexane.<br />

Multicomponent <strong>droplet</strong>s<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Water<br />

Ethanol<br />

Hexane<br />

Evaporation coefficient β r<br />

Droplets consisting <strong>of</strong> a mixture <strong>of</strong> two or more pure solvents are used in many technical<br />

applications, especially in combustion processes. Experimental investigations <strong>of</strong> the evaporation<br />

<strong>and</strong> combustion <strong>of</strong> <strong>droplet</strong>s <strong>of</strong> binary mixtures have been performed by Law [1982], Yang et al.<br />

[1989] <strong>and</strong> using the method <strong>of</strong> acoustic levitation by Yarin et al. [2002 b]. In all studies it was<br />

found that the evaporation coefficient β r changed with time under constant <strong>drying</strong> conditions.<br />

Figure 5.15 shows the result <strong>of</strong> evaporation <strong>of</strong> a levitated <strong>droplet</strong> consisting <strong>of</strong> a binary mixture <strong>of</strong><br />

water <strong>and</strong> ethanol at 30°C <strong>and</strong> 0.1% relative humidity.<br />

0.009<br />

0.008<br />

0.007<br />

0.006<br />

0.005<br />

0.004<br />

0.003<br />

0.002<br />

0.001<br />

0.000<br />

r<br />

r<br />

water ethanol hexane<br />

Solvent


108 RESULTS AND DISCUSSION<br />

Figure 5.15: Variation <strong>of</strong> (rS 2 /rS 0 2 ) with (t/rS 0 2 ) for three different <strong>droplet</strong>s, consisting <strong>of</strong> pure water, pure ethanol <strong>and</strong> a<br />

mixture <strong>of</strong> both. The experiments were carried out at 30.0°C <strong>and</strong> a relative humidity <strong>of</strong> 0.1% to prevent water vapour<br />

condensation at the <strong>droplet</strong> surface. The initial effective SPL was 163.5 dB for water, 162.3 dB for ethanol <strong>and</strong><br />

161.8 dB for the mixture.<br />

2 2<br />

2<br />

The plot shows the evolution <strong>of</strong> ( r / r ) as a function <strong>of</strong> ( t / r )<br />

S<br />

S 0<br />

S 0<br />

during the evaporation process.<br />

For pure water <strong>and</strong> ethanol <strong>droplet</strong>s, a linear behaviour is obtained according to the d 2 -law. For the<br />

mixture <strong>of</strong> these solvents a non-linear behaviour is observed. The slope <strong>of</strong> the curve changes with<br />

time. Initially, the momentary β r lies in between the values found for the pure liquids, but more on<br />

the side <strong>of</strong> the more volatile solvent. Towards the end <strong>of</strong> the evaporation process the slope has<br />

become approximately the same as that for the less volatile liquid. The same observations were<br />

found for mixtures <strong>of</strong> pentadecane <strong>and</strong> hexadecane in optically levitated <strong>droplet</strong> [Frohn 2000], <strong>and</strong><br />

for mixtures <strong>of</strong> methanol <strong>and</strong> water in acoustically levitated <strong>droplet</strong>s [Kastner 2001].<br />

The differences between pure solvent <strong>and</strong> mixtures <strong>of</strong> binary liquids can be well illustrated<br />

from a plot <strong>of</strong> the rate <strong>of</strong> size change as a function <strong>of</strong> time according to<br />

Equation 5.5<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

drS<br />

β r = −<br />

dt 2r<br />

S<br />

Figure 5.16 presents the data for water, ethanol <strong>and</strong> the binary mixture plotted using Equation 5.5.<br />

Initially, the evaporation rate <strong>of</strong> the mixture <strong>droplet</strong> is closer to that <strong>of</strong> the more volatile pure<br />

solvent. Towards the end <strong>of</strong> the evaporation process drS /dt converged towards, but does not reach,<br />

the values <strong>of</strong> the less volatile liquid.<br />

2 ) [-]<br />

2 / rS 0<br />

mainly ethanol<br />

evaporation<br />

0.2<br />

0.0<br />

mainly water<br />

evaporation<br />

0 500 1000 1500 2000 2500<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Pure water<br />

Ethanol 60% (V/V)<br />

Pure ethanol


CHAPTER 5 RESULTS AND DISCUSSION 109<br />

dr S / dt [mm/s]<br />

0.000<br />

-0.001<br />

-0.002<br />

-0.003<br />

Figure 5.16: Size change rate as function <strong>of</strong> time for pure water <strong>and</strong> ethanol <strong>droplet</strong>s <strong>and</strong> <strong>droplet</strong>s consisting <strong>of</strong> binary<br />

mixture <strong>of</strong> both substances. The <strong>droplet</strong>s were levitated at a temperature <strong>of</strong> 30.0°C <strong>and</strong> 0.1// relative humidity. The<br />

initial effective SPL was 163.5dB for the water<strong>droplet</strong>, 162.3dB for the ethanol <strong>droplet</strong> <strong>and</strong> 161.8dB for the mixture<br />

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

5.2.2 Evaporation <strong>of</strong> pure water <strong>droplet</strong>s<br />

-0.004<br />

Pure water<br />

Ethanol 60% (V/V)<br />

-0.005<br />

Pure ethanol<br />

0 200 400 600 800 1000<br />

Evaporation time [s]<br />

Water in its pure form is the most universal solvent in spray-<strong>drying</strong> [Masters 2002]. It has a surface<br />

−6<br />

2<br />

tension <strong>of</strong> 72 . 75mN/m<br />

<strong>and</strong> a kinematic viscosity <strong>of</strong> 1.<br />

004 ⋅ 10 m /s at 20°C<br />

[EngineeringToolBox 2005]. Its single molecules are linked by hydrogen bonds to form a kind <strong>of</strong><br />

dynamic cluster structure. The size <strong>of</strong> these clusters depends on the temperature <strong>of</strong> the water itself.<br />

At 0°C a cluster consists <strong>of</strong> about 300 molecules, at 100°C <strong>of</strong> about 20 molecules <strong>and</strong> at the critical<br />

temperature <strong>of</strong> 374.1°C <strong>of</strong> one to two molecules [Bauer et al. 1999; Burger 1998]. The structural<br />

composition can be seen in Figure 5.17.<br />

Figure 5.17: Structural composition <strong>of</strong> water [Burger 1998].


110 RESULTS AND DISCUSSION<br />

The average lifetime <strong>of</strong> the hydrogen bonds that determine the structural composition decreases<br />

with increasing temperature. At room temperature the lifetime is about 10 -4 s. The short lifetime <strong>of</strong><br />

the hydrogen bonds combined with the thermal mobility <strong>of</strong> the molecules lead to a high velocity <strong>of</strong><br />

structural changes without loosing the basic composition principle [Bauer et al. 1999]. These<br />

characteristics determine the solvent properties <strong>of</strong> water. The larger the concentration <strong>of</strong><br />

imperfections in the structural composition, the better is the potential to dissolve other substances.<br />

This explains the high temperature dependence <strong>of</strong> the solvent properties <strong>of</strong> water.<br />

Calculation <strong>of</strong> limits <strong>of</strong> the levitation range for water <strong>droplet</strong>s<br />

Samples with a diameter d ≥ 2 / 3⋅<br />

λ0<br />

cannot be levitated within the st<strong>and</strong>ing acoustic wave <strong>of</strong> an<br />

S<br />

ultrasonic levitator, because the SPL decreases with increasing <strong>droplet</strong> size [Lierke 1996]. The<br />

maximum levitation size <strong>of</strong> liquid samples is also dependent on the surface tension <strong>and</strong> the density<br />

<strong>of</strong> the liquid. Experiments performed by Lierke showed that <strong>droplet</strong>s with an acoustic Bond<br />

Number ≥ 1.5 cannot be levitated. This Bond number is independent <strong>of</strong> the ultrasonic frequency,<br />

which means that these figures are absolute values [Tec5 2002]. It is calculated by Tuckermann<br />

[Tuckermann 2002] that<br />

Equation 5.6 ( Bo)<br />

acoustic<br />

ρ<br />

=<br />

liquid<br />

2<br />

⋅ vˆ<br />

⋅ d<br />

8⋅<br />

γ<br />

max S<br />

The minimum diameter <strong>of</strong> a levitated sample depends on the frequency <strong>of</strong> the ultrasonic levitator<br />

<strong>and</strong> on the viscosity <strong>of</strong> the surrounding gas [Tuckermann 2002]. For water <strong>droplet</strong>s the minimum<br />

<strong>and</strong> maximum <strong>droplet</strong> sizes at 20°C are calculated according to Equation 5.6 <strong>and</strong> are shown in<br />

Table 5.1.<br />

Table 5.1: Overview <strong>of</strong> the minimum <strong>and</strong> maximum diameter <strong>and</strong> volume <strong>of</strong> levitated water <strong>droplet</strong>s for<br />

(Bo)acoustic = 1.5 in an 58 kHz acoustic levitator.<br />

Liquid<br />

Surface<br />

tension γ<br />

[ ]<br />

Density<br />

ρ liquid<br />

3<br />

mN/m [ ]<br />

Maximum<br />

diameter<br />

Maximum<br />

volume<br />

Minimum<br />

diameter<br />

Minimum<br />

volume<br />

g/cm [ mm ] [ µl ] [ µm ] [ µl ]<br />

Water 72.75 0.9982 6.67 155.37 15µm 1.77⋅10 -6


CHAPTER 5 RESULTS AND DISCUSSION 111<br />

The optimal diameter for a stable water <strong>droplet</strong> in an acoustic levitator is determined according to<br />

Lierke [1996], using:<br />

d S opt<br />

Equation 5.7 ( )<br />

c0<br />

=<br />

3⋅<br />

f<br />

At a temperature <strong>of</strong> 20°C with a sound velocity <strong>of</strong> 343 m/s, the optimal stable <strong>droplet</strong> diameter in a<br />

58 kHz levitator is ( ) = 1.<br />

97 mm<br />

d , corresponding to a volume <strong>of</strong> 4.01 µl.<br />

S<br />

opt<br />

Levitation <strong>of</strong> water <strong>droplet</strong>s<br />

The evaporation <strong>of</strong> a pure water <strong>droplet</strong> is performed in the acoustic levitator at 25°C <strong>and</strong> a relative<br />

humidity <strong>of</strong> 3.0% without forced air. The data is plotted in Figure 5.18 a. Figure 5.18 b shows the<br />

course <strong>of</strong> the effective sound pressure level calculated using the method <strong>of</strong> Yarin et al. [1998],<br />

together with <strong>droplet</strong> volume.<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

2 2 1.0 ( r / rS )<br />

S 0<br />

volume<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 250 500 750 1000 1250<br />

(a)<br />

Evaporation time [s]<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

Droplet volume [µl]<br />

0.0<br />

0 250 500 750 1000 1250<br />

Figure 5.18: Evaporation <strong>of</strong> a pure water <strong>droplet</strong> at 25°C <strong>and</strong> 3.0% relative humidity at an initial effective SPL <strong>of</strong><br />

163.25 dB.(a) Size <strong>and</strong> volume change <strong>of</strong> the <strong>droplet</strong> versus time. (b) Influence <strong>of</strong> the volume change on the course <strong>of</strong><br />

the sound pressure level.<br />

The decrease in <strong>droplet</strong> volume has a substantial influence on the development <strong>of</strong> the effective SPL<br />

values. A decreasing <strong>droplet</strong> volume leads to increasing effective SPL due to a resonance shift<br />

caused by changing reflections <strong>of</strong> the acoustic wave at the <strong>droplet</strong> surface [Kastner 2001]. The<br />

theoretical evaporation curve for the diffusion-controlled evaporation <strong>of</strong> a single <strong>droplet</strong> with the<br />

same initial size in still air, <strong>and</strong> wet bulb temperature calculated using Equation 2.69 shown in<br />

Figure 5.19.<br />

Droplet volume [µl]<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

(b)<br />

Droplet volume<br />

Effective SPL<br />

Evaporation time [s]<br />

166<br />

165<br />

164<br />

163<br />

Effective SPL [dB]


112 RESULTS AND DISCUSSION<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 500 1000 1500<br />

(a)<br />

Evaporation time [s]<br />

Diffusion model<br />

Experimental data<br />

1.75<br />

0 500 1000 1500<br />

Figure 5.19: Comparison <strong>of</strong> the evaporation <strong>of</strong> a levitated <strong>droplet</strong> with the diffusion-controlled evaporation model<br />

without convective influence at 25°C <strong>and</strong> 3.0%relative humidity. The <strong>droplet</strong> had an initial volume <strong>of</strong> 2.0 µl <strong>and</strong> the<br />

initial effective SPL was 163.25 dB. (a) (rS 2 /rS 0 2 ) with evaporation time. (b) Comparison <strong>of</strong> the Shwerwood numbers.<br />

The diffusion-controlled evaporation <strong>of</strong> a single <strong>droplet</strong> according to Equation 2.69 has a longer<br />

(4.3 min) evaporation time than the experimental evaporation curve. The Sherwood number<br />

calculated from the experimental data is 2.5, 0.5 larger than expected for diffusion-controlled<br />

evaporation in still air [Masters 1991]. The reason for the increased Sherwood number can be<br />

explained by influences <strong>of</strong> inner acoustic streaming. The continuity <strong>of</strong> shear stress at the surface<br />

leads to extra loss <strong>of</strong> liquid from the <strong>droplet</strong> resulting in the observed increased mass transfer<br />

[Rensink 2004; Yarin et al. 1999]. To account for inner acoustic streaming the data were calculated<br />

again using Equation 3.40 <strong>and</strong> Equation 3.34. Figure 5.20 shows that the calculated evaporation<br />

curve now yields a larger evaporation rate with shorter evaporation time than the experimental<br />

evaporation curve. The Sherwood number is averaged to 4.9, which is about twice as large as the<br />

Sherwood number <strong>of</strong> 2.5 for the experimental data determined only from the d 2 -law. The reason is<br />

to be found in the outer acoustic streaming, in which solvent vapour <strong>of</strong> the evaporated <strong>droplet</strong> is<br />

accumulated. This accumulation decreases heat <strong>and</strong> mass transfer from the outer streaming vortices<br />

to the <strong>drying</strong> air <strong>and</strong> causes a decrease in evaporation rate <strong>and</strong> hence an increase in evaporation time<br />

[Kastner 2001; Yarin et al. 1999]. This hindering effect can be overcome by using a forced <strong>drying</strong><br />

air stream [Rensink 2004]. The influence <strong>of</strong> different <strong>drying</strong> air flow velocities on the <strong>droplet</strong><br />

evaporation is therefore tested <strong>and</strong> compared to the calculated data.<br />

Sherwood number<br />

3.00<br />

2.75<br />

2.50<br />

2.25<br />

2.00<br />

(b)<br />

Evaporation time [s]<br />

Diffusion model<br />

Experimental data


CHAPTER 5 RESULTS AND DISCUSSION 113<br />

S ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 250 500 750 1000 1250<br />

(a)<br />

Evaporation time [s]<br />

Levitation model<br />

Experimental data<br />

Figure 5.20: Comparison between the evaporation model for levitated <strong>droplet</strong>s with increased Sherwood numbers due<br />

to the inner acoustic streaming <strong>and</strong> experimental data <strong>of</strong> a 2.0µl <strong>droplet</strong> in still air at 25°C <strong>and</strong> 3.0% relative humidity.<br />

(a) (rS 2 /rS 0 2 )-course with evaporation time. (b) Comparison <strong>of</strong> the Sherwood numbers.<br />

2.0 µl <strong>droplet</strong>s <strong>of</strong> water are levitated at 25.0°C <strong>and</strong> 3.0% relative humidity <strong>and</strong> evaporated at<br />

different velocities <strong>of</strong> the <strong>drying</strong> air stream. A comparison <strong>of</strong> the experimental results to the model<br />

data calculated using Equation 3.40 <strong>and</strong> Equation 3.34 can be seen in Figure 5.21. A velocity <strong>of</strong><br />

0.41 m/s corresponding to an orifice Reynolds number <strong>of</strong> 100 (Figure 5.21 a) leads to an<br />

evaporation time still larger than the calculated one. The water vapour is evidently not removed <strong>of</strong><br />

the system <strong>of</strong> vortices completely <strong>and</strong> accumulation still occurs. At an air stream <strong>of</strong> 0.82 m/s<br />

(Figure 5.21 b) the experimental data matches almost exactly with the calculated curve. The water<br />

vapour is removed completely from the outer streaming vortices without causing any additional<br />

convective <strong>droplet</strong> <strong>drying</strong>. This result was also found by Rensink, who found a suitable <strong>drying</strong> air<br />

stream <strong>of</strong> 1.0 m/s [Rensink 2004]. Higher <strong>drying</strong> air velocities cause a further decrease in the<br />

evaporation time (Figure 5.21 c), resulting now from additional convective <strong>droplet</strong> <strong>drying</strong> due to the<br />

forced air stream. The Sherwood number for experiments with <strong>drying</strong> air velocities lower than<br />

1.0 m/s are well described by Equations 3.33 <strong>and</strong> Equation 3.34 (Figure 5.21 d). The Ranz-<br />

Marshall-Equation 2.57 provides a better solution for calculation <strong>of</strong> the radius-time course <strong>of</strong><br />

levitated <strong>droplet</strong> at air velocities ≥ 1.0 m/s [Ranz 1952 a; Ranz 1952 b].<br />

Sherwood number<br />

5.5<br />

5.0<br />

4.5<br />

4.0<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

(b)<br />

0 250 500 750 1000 1250<br />

Evaporation time [s]<br />

Levitation model<br />

Experimental data


114 RESULTS AND DISCUSSION<br />

S ) [-]<br />

2 / rS 0<br />

( r S<br />

S ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

(a)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Levitation model<br />

ν air = 0.41 m/s (Re orifice = 100)<br />

0 250 500 750 1000<br />

Evaporation time [s]<br />

0.0<br />

0 250 500 750 1000<br />

(c)<br />

Evaporation model<br />

ν air = 1.23 m/s (Re orifice = 300)<br />

Evaporation time [s]<br />

0.0<br />

0 250 500 750 1000<br />

Figure 5.21: Comparison <strong>of</strong> the model data with increased Sherwood numbers with the evaporation results <strong>of</strong> levitated<br />

water <strong>droplet</strong>s at 25°C, 3.0% relative humidity <strong>and</strong> different ventilation air stream velocities. (a) (rS 2 /rS 0 2 )-evaporation<br />

time course for a levitated <strong>droplet</strong> with a ventilation <strong>of</strong> νair = 0.41 m/s, corresponding to an orifice Reynolds number <strong>of</strong><br />

100. (b) (rS 2 /rS 0 2 )-evaporation time course for a levitated 2.0 µl <strong>droplet</strong> with a ventilation <strong>of</strong> νair = 0.82 m/s,<br />

corresponding to an orifice Reynolds number <strong>of</strong> 200. (c) (rS 2 /rS 0 2 )-evaporation time course for a levitated 2.0 µl <strong>droplet</strong><br />

with a ventilation <strong>of</strong> νair = 1.23 m/s, corresponding to an orifice Reynolds number <strong>of</strong> 300. (d) Comparison <strong>of</strong> the<br />

Sherwood numbers with time between model <strong>and</strong> experiment at Reorifice = 200.<br />

Figure 5.22 compares Equation 2.57 with the experimental data at 25.0°C, 3.0% relative humidity<br />

<strong>and</strong> different air velocities. The air velocity <strong>of</strong> 1.23 m/s corresponds to an orifice Reynolds number<br />

<strong>of</strong> 300, <strong>and</strong> the velocity <strong>of</strong> 1.64 m/s to an orifice Reynolds number <strong>of</strong> 400. It can be seen that the<br />

model <strong>of</strong> Equation 2.57 predicts the radius-time course <strong>and</strong> the Sherwood number almost exactly<br />

for both data sets.<br />

S ) [-]<br />

2 / rS 0<br />

( r S<br />

Sherwood number<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

(b)<br />

6.0<br />

5.0<br />

4.0<br />

3.0<br />

2.0<br />

(d)<br />

Levitation model<br />

ν air = 0.82 m/s (Re orifice = 200)<br />

Evaporation time [s]<br />

Levitation model<br />

ν air = 0.82 m/s (Re orifice = 200)<br />

0 250 500 750 1000<br />

Evaporation time [s]


CHAPTER 5 RESULTS AND DISCUSSION 115<br />

S ) [-]<br />

2 / rS 0<br />

( r S<br />

S ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 100 200 300 400 500 600 700<br />

(a)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Evaporation model Ranz-Marshall<br />

Experiment ν air = 1.23 m/s<br />

Evaporation time [s]<br />

0.0<br />

0 100 200 300 400 500 600 700<br />

(a)<br />

Evaporation model Ranz-Marshall<br />

Experiment ν air = 1.639 m/s<br />

Evaporation time [s]<br />

0 100 200 300 400 500 600 700<br />

Figure 5.22: Comparison <strong>of</strong> the Ranz-Marshall correlation with experimental data from levitation experiments at 25°C,<br />

3.0% relative humidity <strong>and</strong> different ventilation air stream velocities. The initial effective SPL for both data sets was<br />

163.5 dB. (a) (rS 2 /rS 0 2 ) with evaporation time for a 1.51 µl <strong>droplet</strong> at νair = 1.23 m/s (Reorifice = 300). (b) Comparison <strong>of</strong><br />

the Sherwood numbers <strong>of</strong> the experimental data in (a). (c) (rS 2 /rS 0 2 ) with evaporation time for a 1.64 µl <strong>droplet</strong> at<br />

νair = 1.64 m/s (Reorifice = 400). (d) Comparison <strong>of</strong> the Sherwood numbers <strong>of</strong> the experimental data in (c).<br />

The following experimental results with water <strong>droplet</strong>s at various temperatures were compared to<br />

the three different evaporation models outlined above:<br />

- evaporation model for diffusion-controlled evaporation without any convective influence<br />

(Equation 2.69);<br />

- evaporation model (levitation model) including influence <strong>of</strong> the inner acoustic streaming<br />

(Equation 3.40 <strong>and</strong> Equation 3.34);<br />

- evaporation model at high <strong>drying</strong> air velocities using the Ranz-Marshall correlation<br />

(Equation 2.57).<br />

The results at 40°C are given in Figure 5.23, <strong>and</strong> the results at 60°C in Figure 5.24.<br />

Sherwood number<br />

Sherwood number<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

(b)<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

(d)<br />

Evaporation model Ranz-Marshall<br />

Experiment ν air = 1.23 m/s<br />

Evaporation time [s]<br />

Evaporation model Ranz-Marshall<br />

Experiment ν air =1.639 m/s<br />

0 100 200 300 400 500 600 700<br />

Evaporation time [s]


116 RESULTS AND DISCUSSION<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 200 400 600 800 1000<br />

(a)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 200 400 600 800<br />

(c)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Levitation model at 40°C<br />

Experimental data ν air = 0.00 m/s<br />

Evaporation time [s]<br />

0.0<br />

0 200 400 600 800<br />

(a)<br />

Diffusion-controlled evaporation<br />

Experiment in still air at 40.0°C<br />

Evaporation time [s]<br />

Levitation model at 40°C<br />

Experimental data ν air = 0.88 m/s<br />

(Re orifice = 200)<br />

Evaporation time [s]<br />

0 200 400 600 800<br />

Figure 5.23: Comparison <strong>of</strong> the evaporation models with experimental data from levitated water <strong>droplet</strong>s at 40°C <strong>and</strong><br />

0.1% relative humidity. The initial <strong>droplet</strong> volume <strong>of</strong> all experiments was 2.0 µl at an initial effective SPL <strong>of</strong> 166.0 dB.<br />

(a) Comparison <strong>of</strong> the diffusion-controlled evaporation model with experimental data in still air. (b) Comparison <strong>of</strong> the<br />

Sherwood numbers <strong>of</strong> the diffusion-controlled evaporation model <strong>and</strong> experiments in air. (c) Comparison <strong>of</strong> the<br />

levitation model with experimental data in still air. (d) Comparison <strong>of</strong> the Sherwood numbers <strong>of</strong> the levitation model<br />

with experiments in still air. (e) Comparison <strong>of</strong> the levitation model with experimental data at a ventilation velocity <strong>of</strong><br />

0.88 m/s. (f) Comparison <strong>of</strong> the Sherwood number <strong>of</strong> the levitation model with experimental data at a ventilation<br />

velocity <strong>of</strong> 0.88 m/s.<br />

Sherwood number<br />

Sherwood number<br />

Sherwood number<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0 200 400 600 800<br />

(b)<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

(d)<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

Leviation model at 40°C<br />

Experimental data ν air = 0.00 m/s<br />

Evaporation time [s]<br />

2.0<br />

(Re = 200)<br />

orifice<br />

0 200 400 600 800<br />

(f)<br />

Diffusion-controlled evaporation<br />

Experimental data ν air = 0.00 m/s<br />

Evaporation time [s]<br />

Levitation model at 40°C<br />

Experimental data ν air = 0.88 m/s<br />

Evaporation time [s]


CHAPTER 5 RESULTS AND DISCUSSION 117<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 100 200 300<br />

(a)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Diffusion-controlled evaporation<br />

Experimental data ν air = 0.00 m/s<br />

Evaporation time [s]<br />

0.0<br />

0 100 200 300<br />

(c)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Levitation model at 60°C<br />

Experimental data ν air = 0.00 m/s<br />

Evaporation time [s]<br />

0.0<br />

0 100 200 300<br />

(e)<br />

Levitation model at 60°C<br />

Experimental data ν air = 1.0m/s<br />

(Re orifice = 200)<br />

Evaporation time [s]<br />

0 100 200 300<br />

Figure 5.24: Comparison <strong>of</strong> the evaporation models with experimental data from levitated water <strong>droplet</strong>s at 60°C <strong>and</strong><br />

0.1% relative humidity. The initial <strong>droplet</strong> volume <strong>of</strong> all experiments was 1.0 µl at an initial effective SPL <strong>of</strong> 167.0 dB.<br />

(a) Comparison <strong>of</strong> the diffusion-controlled evaporation model with experimental data in still air. (b) Comparison <strong>of</strong> the<br />

Sherwood numbers <strong>of</strong> the diffusion-controlled evaporation model with experimental data in air. (c) Comparison <strong>of</strong> the<br />

levitation model with experimental data in still air. (d) Comparison <strong>of</strong> the Sherwood numbers <strong>of</strong> the levitation model<br />

with experimental data in still air. (e) Comparison <strong>of</strong> the levitation model with experimental data at a ventilation<br />

velocity <strong>of</strong> 1.0 m/s. (f) Comparison <strong>of</strong> the Sherwood number <strong>of</strong> the levitation model with experimental data at a<br />

ventilation velocity <strong>of</strong> 1.0 m/s.<br />

Sherwood number<br />

Sherwood number<br />

Sherwood number<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

(b)<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

(d)<br />

Diffusion-controlled evaporation<br />

Experimental data ν air = 0.00 m/s<br />

Evaporation time [s]<br />

Levitation model at 60°C<br />

Experimental data ν air = 0.00 m/s<br />

0 100 200 300<br />

Evaporation time [s]<br />

2.0<br />

(Re = 200)<br />

orifice<br />

0 100 200 300<br />

(f)<br />

Levitation model at 60°C<br />

Experimental data ν air = 1.0 m/s<br />

Evaporation time [s]


118 RESULTS AND DISCUSSION<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 200 400 600 800<br />

(a)<br />

Ranz-Marshall correlation<br />

Experiment ν air = 1.77 m/s<br />

(Re orifice = 400)<br />

Evaporation time [s]<br />

2.0<br />

(Re = 400)<br />

orifice<br />

0 200 400 600 800<br />

Figure 5.25: Comparison <strong>of</strong> the Ranz-Marshall correlation with experimental data from levitation experiments at 40°C,<br />

0.1% relative humidity <strong>and</strong> a ventilation velocity <strong>of</strong> 1.77 m/s (Reorifice = 400). The initial effective SPL was 166.0 dB<br />

<strong>and</strong> the initial <strong>droplet</strong> volume was 1.7 µl. (a) (rS 2 /rS 0 2 ) with evaporation time. (b) Comparison <strong>of</strong> the Sherwood numbers<br />

<strong>of</strong> the Ranz-Marshall correlation with experimental data.<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 50 100 150 200 250 300<br />

(a)<br />

Evaporation Ranz-Marshall<br />

Experiment ν air = 1.98 m/s<br />

(Re orifice = 400)<br />

Evaporation time [s]<br />

Figure 5.26: Comparison <strong>of</strong> the Ranz-Marshall correlation with experimental data from levitation experiments at 60°C,<br />

0.1% relative humidity <strong>and</strong> a ventilation velocities <strong>of</strong> 1.98 m/s (Reorifice = 400). The initial effective SPL was 166.8 dB<br />

<strong>and</strong> the initial <strong>droplet</strong> volume was 1.7 µl. (a) (rS 2 /rS 0 2 ) with evaporation time. (b) Comparison <strong>of</strong> the Sherwood numbers<br />

<strong>of</strong> the Ranz-Marshall correlation with experimental data.<br />

Figure 5.23 presents that the levitation experiments for water <strong>droplet</strong>s at 40°C <strong>and</strong> 0.1% relative<br />

humidity show agreement with the levitation model, as also observed in Figure 5.21 for the<br />

experiments at 25°C <strong>and</strong> 3.0% relative humidity. The comparison between the Ranz-Marshall<br />

correlation <strong>and</strong> the experimental data at a <strong>drying</strong> velocity <strong>of</strong> 1.77 m/s in Figure 5.25 shows,<br />

however, a larger difference in the later times <strong>of</strong> <strong>droplet</strong> evaporation. This is caused by differing<br />

values <strong>of</strong> the Sherwood number within the later part. According to the Ranz-Marshall correlation<br />

Sherwood number<br />

Sherwood number<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

18.0<br />

16.0<br />

14.0<br />

12.0<br />

10.0<br />

(b)<br />

8.0<br />

6.0<br />

4.0<br />

Ranz-Marshall correlation<br />

Experiment ν air = 1.77 m/s<br />

Evaporation time [s]<br />

2.0<br />

0 50 100 150 200 250 300<br />

(b)<br />

Evaporation Ranz-Marshall<br />

Experiment ν air = 1.98 m/s<br />

(Re orifice = 400)<br />

Evaporation time [s]


CHAPTER 5 RESULTS AND DISCUSSION 119<br />

the Sherwood number decreases to values under 6.0 with decreasing <strong>droplet</strong> diameter Figure 5.25 b.<br />

The Sherwood number does not, however, decrease under 6.0 during the whole experiment, due to<br />

the influence <strong>of</strong> the inner acoustic streaming. Figure 5.23f shows that the Sherwood number will be<br />

> 6.0 at all times for this combination <strong>of</strong> initial <strong>droplet</strong> diameter <strong>and</strong> effective SPL.<br />

The experiments performed at 60°C with a <strong>drying</strong> air velocity up to 0.99 m/s could be<br />

described almost exactly with the levitation model (Figure 5.24 e). At velocities larger than 1.0 m/s<br />

the experimental results do not match with the Ranz-Marshall correlation (Figure 5.26 a). The<br />

experimental data at 1.98 m/s show a higher evaporation rate <strong>and</strong> smaller evaporation time than<br />

predicted by the model. A comparison <strong>of</strong> the Sherwood numbers <strong>of</strong> the levitation model <strong>and</strong> the<br />

Ranz-Marshall correlation shows that these are similar (Figure 5.24 f <strong>and</strong> Figure 5.26 b). The<br />

Sherwood number according to the levitation model is on average about 7.5. The increase in <strong>drying</strong><br />

air velocity from 1.0 m/s to 1.98 m/s does not correlate with an adequate increase in Sherwood<br />

number calculated via the Ranz-Marshall correlation. Consequently the experiments at 60°C, 0.1%<br />

relative humidity <strong>and</strong> a <strong>drying</strong> air velocity > 1.0 m/s cannot be described using the Ranz-Marshall<br />

correlation.<br />

An explanation for the discrepancy between the Ranz-Marshall model <strong>and</strong> experimental data<br />

at 60°C <strong>and</strong> 0.1% relative humidity may be the unstable <strong>droplet</strong> position within the st<strong>and</strong>ing<br />

acoustic wave. In almost dry air <strong>and</strong> at velocities <strong>of</strong> about 2.0 m/s very strong oscillation <strong>of</strong> the<br />

<strong>droplet</strong> occurred. Rensink [2004] showed that oscillations <strong>of</strong> the <strong>droplet</strong> in relation to its<br />

environment as well as internal oscillations <strong>of</strong> the <strong>droplet</strong> surface itself can lead to an increase in the<br />

Reynolds number <strong>and</strong> therefore also to an increase in the evaporation rate. At higher relative<br />

humidity <strong>of</strong> the <strong>drying</strong> air at 60°C both kinds <strong>of</strong> <strong>droplet</strong> oscillations decreased visibly, which led to<br />

a better correlation between the experimental data <strong>and</strong> the predictions <strong>of</strong> the radius-time course by<br />

the Ranz-Marshall correlation (Figure 5.27).


120 RESULTS AND DISCUSSION<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 100 200 300 400<br />

(a)<br />

Ranz-Marshall correlation<br />

at 60°C <strong>and</strong> 40% rel. humidity<br />

Experimental data ν air = 2.04 m/s<br />

Evaporation time [s]<br />

2.0<br />

0 100 200 300 400<br />

Figure 5.27: Comparison <strong>of</strong> the Ranz-Marshall correlation with experimental data from levitation experiments at 60°C,<br />

40% relative humidity <strong>and</strong> a ventilation velocity <strong>of</strong> 2.04 m/s (Reorifice = 400). The initial effective SPL was 166.8 dB <strong>and</strong><br />

the initial <strong>droplet</strong> volume was 1.7 µl. (a) (rS 2 /rS 0 2 ) with evaporation time. (b) Comparison <strong>of</strong> the Sherwood numbers <strong>of</strong><br />

the Ranz-Marshall correlation with experimental data.<br />

The calculation <strong>of</strong> a single evaporation coefficient describing the whole radius-time course <strong>of</strong><br />

<strong>droplet</strong> evaporation is thus not possible because <strong>of</strong> the influence <strong>of</strong> both inner acoustic streaming<br />

<strong>and</strong> <strong>drying</strong> air velocity on the Sherwood number. As already discussed, the relation between the<br />

volume <strong>of</strong> a levitated <strong>droplet</strong> <strong>and</strong> the effective SPL determines indirectly the Sherwood number via<br />

Equation 3.33 <strong>and</strong> Equation 3.34. If each experiment starts with the same <strong>droplet</strong> volume at the<br />

same effective SPL, a range can be chosen in which the evaporation coefficient can be calculated<br />

for differing <strong>drying</strong> conditions. Droplets with an initial volume <strong>of</strong> 1.5 µl were levitated at the<br />

pressure node <strong>of</strong> the st<strong>and</strong>ing acoustic wave at an initial effective SPL <strong>of</strong> 165.5 dB. The influences<br />

<strong>of</strong> temperature, humidity <strong>and</strong> <strong>drying</strong> air velocity on the evaporation coefficient, as well as a<br />

correlation to the different models were investigated. Typical raw data <strong>of</strong> these <strong>droplet</strong> <strong>drying</strong><br />

experiments at different temperature <strong>and</strong> humidity without any <strong>drying</strong> air stream are shown in<br />

Figure 5.28. The evaporation coefficients were calculated in the range between 1.0 µl <strong>and</strong> 0.5 µl <strong>of</strong><br />

<strong>droplet</strong> volume from the d 2 -law using Equation 5.4. The influence <strong>of</strong> temperature <strong>and</strong> humidity can<br />

clearly be seen.<br />

Sherwood number<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

(b)<br />

Ranz-Marshall correlation<br />

at 60°C <strong>and</strong> 40% rel. humidity<br />

Experimental data ν air = 2.04 m/s<br />

Evaporation time [s]


CHAPTER 5 RESULTS AND DISCUSSION 121<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 500 1000 1500 2000 2500<br />

Figure 5.28: Raw data <strong>of</strong> evaporation experiments at different temperatures <strong>and</strong> relative humidity <strong>of</strong> the <strong>drying</strong> air. The<br />

initial effective SPL was 165.5 dB <strong>and</strong> the initial <strong>droplet</strong> volume was 1.5 µl. All experiments were performed without<br />

ventilation air stream. (a) (rS 2 / rS 0 2 ) with (t / rS 0 2 ) at different temperatures <strong>and</strong> 0.1% relative humidity <strong>of</strong> the <strong>drying</strong> air.<br />

(b) (rS 2 / rS 0 2 ) with (t / rS 0 2 ) at 60°C <strong>and</strong> different humidity <strong>of</strong> the <strong>drying</strong> air.<br />

Evaporation coefficient β r [mm 2 /s]<br />

(a)<br />

0.0030<br />

0.0025<br />

0.0020<br />

0.0015<br />

0.0010<br />

0.0005<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

25°C<br />

40°C<br />

50°C<br />

60°C<br />

70°C<br />

0.0000<br />

20 30 40 50 60 70<br />

(a)<br />

Rel. humidity 0.1%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

Rel. humidity 60%<br />

Rel. humidity 80%<br />

Drying temperature [°C]<br />

0 1000 2000 3000 4000 5000 6000<br />

Figure 5.29: Influence <strong>of</strong> temperature <strong>and</strong> relative humidity <strong>of</strong> the <strong>drying</strong> air on the evaporation coefficient <strong>of</strong> levitated<br />

water <strong>droplet</strong>s inside an acoustic levitator. The evaporation coefficients were determined between 1.0 µl <strong>and</strong> 0.5 µl <strong>of</strong><br />

the <strong>droplet</strong> volume. The initial effective SPL for all experiments used in this comparison was 165.5 dB <strong>and</strong> the initial<br />

<strong>droplet</strong> volume was 1.5 µl. (a) evaporation coefficients with <strong>drying</strong> temperature at different humidity. (b) Evaporation<br />

coefficients with the relative humidity <strong>of</strong> the <strong>drying</strong> air at different temperature.<br />

To compare the various experiments performed at differing <strong>drying</strong> velocity, the influence <strong>of</strong><br />

changing diameter on the Reynolds <strong>and</strong> Sherwood numbers has to be taken into account. In these<br />

experiments the evaporation coefficient was determined at <strong>droplet</strong> diameter between 1200 µm <strong>and</strong><br />

800 µm. The influences <strong>of</strong> temperature, humidity <strong>and</strong> <strong>drying</strong> air velocity on the evaporation rate <strong>of</strong><br />

levitated water <strong>droplet</strong>s are determined starting with an initial volume <strong>of</strong> 2.0 µl <strong>and</strong> at an initial<br />

effective SPL <strong>of</strong> 165.5 dB. Representative raw data as well as evaporation coefficient are plotted in<br />

Figure 5.30 <strong>and</strong> Figure 5.31.<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

Evaporation coefficient β r [mm 2 /s]<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

(b)<br />

0.0030<br />

0.0025<br />

0.0020<br />

0.0015<br />

0.0010<br />

0.0005<br />

0.0000<br />

(b)<br />

Drying temperature 60°C<br />

Rel. humidity 0.1%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

Rel. humidity 60%<br />

Rel. humidity 80%<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

0 20 40 60 80<br />

Relative humidity <strong>of</strong> <strong>drying</strong> air [%]<br />

25°C<br />

40°C<br />

50°C<br />

60°C<br />

70°C


122 RESULTS AND DISCUSSION<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

Figure 5.30: Typical raw data <strong>of</strong> evaporation experiments at different temperatures <strong>and</strong> <strong>drying</strong> air velocity <strong>of</strong> the <strong>drying</strong><br />

air. The initial effective SPL was 165.5 dB <strong>and</strong> the initial <strong>droplet</strong> volume was 2.0 µl. All experiments were performed at<br />

a relative humidity <strong>of</strong> 0.1%. (a) (rS 2 / rS 0 2 ) with (t / rS 0 2 ) at different ventilation velocities at a temperature <strong>of</strong> 40° <strong>of</strong> the<br />

<strong>drying</strong> air. (b) (rS 2 / rS 0 2 ) with (t / rS 0 2 ) at different temperatures with a ventilation velocity corresponding to an orifice<br />

Reynolds number <strong>of</strong> 400.<br />

Evaporation coefficient β r [mm 2 /s]<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 200 400 600 800 1000 1200 1400<br />

(a)<br />

0.006<br />

0.005<br />

0.004<br />

0.003<br />

0.002<br />

0.001<br />

(a)<br />

25°C<br />

40°C<br />

60°C<br />

T∞ = 40°C; rel. humidity 0.1%<br />

Ventilation ν air = 0.00 m/s<br />

Ventilation ν air = 0.88 m/s<br />

Ventilation ν air = 1.77 m/s<br />

Ventilation ν air = 2.65 m/s<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

0.0 0.5 1.0 1.5 2.0 2.5 3.0<br />

Ventilation velocity ν air [m/s]<br />

0.0 0.5 1.0 1.5 2.0 2.5 3.0<br />

Figure 5.31: Influence <strong>of</strong> temperature <strong>and</strong> relative humidity <strong>of</strong> the <strong>drying</strong> air on the evaporation coefficient <strong>of</strong> levitated<br />

water <strong>droplet</strong>s inside an acoustic levitator at different ventilation velocities. The evaporation coefficients were<br />

determined between 1200 µm <strong>and</strong> 800 µm <strong>of</strong> the <strong>droplet</strong> diameter. The initial effective SPL for all experiments used in<br />

this comparison was 165.5 dB <strong>and</strong> the initial <strong>droplet</strong> volume was 2.0 µl. (a) evaporation coefficients with ventilation<br />

velocity at different <strong>drying</strong> temperature <strong>and</strong> 0.1% relative humidity. (b) Evaporation coefficients with ventilation<br />

velocity at 60.0°C <strong>and</strong> different relative humidity.<br />

An increase in <strong>drying</strong> air velocity or an increase in temperature or a decrease in relative humidity <strong>of</strong><br />

the <strong>drying</strong> air lead to an increase in the evaporation coefficient β r . The largest jump <strong>of</strong> the values<br />

<strong>of</strong> the evaporation coefficient is seen between the experiments at 0.0 m/s <strong>and</strong> 0.88 m/s ventilation<br />

velocity. This shows clearly the effect <strong>of</strong> the accumulation <strong>of</strong> water vapour in the toroidal vortices<br />

<strong>of</strong> the outer acoustic streaming <strong>and</strong> its influence on the evaporation time <strong>of</strong> a single levitated<br />

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

2 ) [-]<br />

2 / rS 0<br />

Evaporation coefficient β r [mm 2 /s]<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 200 400 600 800<br />

(b)<br />

0.006<br />

0.005<br />

0.004<br />

0.003<br />

0.002<br />

0.001<br />

(b)<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Drying temperature 60°C<br />

Rel. humidity 0.1%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

Re orifice = 400; rel. humidity 0.1%<br />

25°C; ν air = 1.67 m/s<br />

40°C; ν air = 1.77 m/s<br />

60°C; ν air = 1.98 m/s<br />

Vemtilation velocity ν air [m/s]


CHAPTER 5 RESULTS AND DISCUSSION 123<br />

A comparison between the data from <strong>droplet</strong> experiments with <strong>drying</strong> air velocity larger than about<br />

0.8 m/s <strong>and</strong> the Ranz-Marshall correlation for the Sherwood number is given in Figure 5.33. The<br />

results at 25°C with a relative humidity <strong>of</strong> 0.1% (Figure 5.33 a) show a close match to the predicted<br />

values calculated from the Ranz-Marshall equation for all tested ventilation velocities. The results<br />

<strong>of</strong> the experiments at 60°C <strong>and</strong> 0.1% relative humidity show, however, poor match to the Ranz-<br />

Marshall equation, as already discussed above (result not shown). At a higher relative humidity <strong>of</strong><br />

the 60°C hot <strong>drying</strong> air (Figure 5.33 b), the experimental evaporation coefficients match quite<br />

closely to the calculated ones using the Ranz-Marshall correlation. The <strong>droplet</strong>s at 60°C <strong>and</strong> 40%<br />

humidity are visibly more stable <strong>and</strong> do not oscillate in the ultrasonic field, accounting for the better<br />

fit than at 0.1% relative humidity.<br />

Evaporation coefficient β r [mm 2 /s]<br />

0.0018<br />

0.0016<br />

0.0014<br />

0.0012<br />

Drying temperature 25°C<br />

Rel. humidity 0.1%<br />

0.0010<br />

0.0008<br />

Model Ranz-Marshall<br />

Model acoustic levitation<br />

Experimental data<br />

0.8 1.2 1.6 2.0 2.4 2.8<br />

(a)<br />

Ventilation velocity ν air [m/s]<br />

0.0020<br />

0.0018<br />

0.0016<br />

0.0014<br />

0.0012<br />

1.0 1.2 1.4 1.6 1.8 2.0 2.2<br />

Figure 5.32: Comparison <strong>of</strong> the evaporation coefficients <strong>of</strong> experimental data with the coefficients predicted using the<br />

Ranz-Marshall correlation. The evaporation coefficients were determined between 1200 µm <strong>and</strong> 800 µm <strong>of</strong> the <strong>droplet</strong><br />

diameter. The initial effective SPL for all experiments used in this comparison was 165.5 dB <strong>and</strong> the initial <strong>droplet</strong><br />

volume was 2.0 µl. (a) evaporation coefficients with ventilation velocity at 25°C <strong>and</strong> 0.1% relative humidity.<br />

(b) Evaporation coefficients with ventilation velocity at 60°C <strong>and</strong> 40% relative humidity.<br />

Evaporation coefficient β r [mm 2 /s]<br />

(b)<br />

Drying temperature 60°C<br />

Relative humidity 40%<br />

Model Ranz-Marshall<br />

Experimental data<br />

Ventilation velocity ν air [m/s]


124 RESULTS AND DISCUSSION<br />

5.2.3 Evaporation <strong>of</strong> pure ethanol <strong>droplet</strong>s<br />

Calculation <strong>of</strong> limits <strong>of</strong> the levitation range for ethanol <strong>droplet</strong>s<br />

The maximum <strong>droplet</strong> diameter for the levitation <strong>of</strong> ethanol <strong>droplet</strong>s depends on the surface tension<br />

<strong>and</strong> density <strong>of</strong> ethanol as well as on the effective SPL <strong>of</strong> the st<strong>and</strong>ing acoustic wave [Lierke 1996;<br />

Tuckermann 2002]. The minimum diameter depends on the frequency <strong>of</strong> the levitator <strong>and</strong> the<br />

viscosity <strong>of</strong> the surrounding gas [Tuckermann 2002]. Table 5.2 gives an overview <strong>of</strong> the size range<br />

for ethanol <strong>droplet</strong>s in a 58 kHz levitation system.<br />

Table 5.2: Overview <strong>of</strong> the minimum <strong>and</strong> maximum diameters <strong>and</strong> volume <strong>of</strong> levitated ethanol <strong>droplet</strong>s for<br />

(Bo)acoustic = 1.5 in an 58 kHz acoustic levitator [EngineeringToolBox 2005; Tuckermann 2002].<br />

Liquid<br />

Surface<br />

tension γ<br />

[ mN/m ]<br />

Density<br />

ρ liquid<br />

3<br />

[ g/cm ]<br />

Maximum<br />

diameter<br />

[ mm ]<br />

Maximum<br />

volume<br />

[ µl ]<br />

Minimum<br />

diameter<br />

[ µm ]<br />

Minimum<br />

volume<br />

[ µl ]<br />

Ethanol 22.3 0.785 4.21 39.03 15µm 1.77⋅10 -6<br />

At a temperature <strong>of</strong> 20°C with a sound velocity <strong>of</strong> 343 m/s the optimal <strong>droplet</strong> diameter in a 58 kHz<br />

levitator is ( ) = 1.<br />

97 mm<br />

d corresponding to a volume <strong>of</strong> 4.01µl, according to Equation 5.7. The<br />

S<br />

opt<br />

−6<br />

2<br />

kinematic viscosity <strong>of</strong> ethanol at 20°C is 1.<br />

52 ⋅ 10 m /s [EngineeringToolBox 2005].<br />

Levitation <strong>of</strong> pure ethanol <strong>droplet</strong>s<br />

The evaporation <strong>of</strong> a pure ethanol <strong>droplet</strong> at 25.0°C <strong>and</strong> a relative humidity <strong>of</strong> 4.1% without<br />

ventilation air stream is analysed here according to chapter 5.2.1 <strong>and</strong> is plotted in Figure 5.33 a. The<br />

influence <strong>of</strong> the decreasing <strong>droplet</strong> volume on the effective sound pressure level <strong>of</strong> the st<strong>and</strong>ing<br />

acoustic wave is shown in Figure 5.33 b calculated using the method <strong>of</strong> Yarin et al. [1998]. As with<br />

water, a decrease in <strong>droplet</strong> volume leads to an increase <strong>of</strong> the effective SPL. The values <strong>of</strong> the SPL<br />

for ethanol <strong>droplet</strong> volume are smaller than those for water <strong>droplet</strong>s with the same volume. This<br />

effect was also shown by Rensink [2004] <strong>and</strong> is caused by the different densities <strong>of</strong> the two liquid<br />

solvents. The experimental data is first compared to the model for diffusion-controlled evaporation<br />

<strong>of</strong> single ethanol <strong>droplet</strong>. The surface temperature <strong>of</strong> the model <strong>droplet</strong> with the same initial <strong>droplet</strong><br />

size is then calculated using Equation 3.41. The comparison between model <strong>and</strong> experiment can be<br />

seen in Figure 5.34.


CHAPTER 5 RESULTS AND DISCUSSION 125<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

2 2<br />

1.0 ( r / rS )<br />

S 0<br />

Droplet volume<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0.0<br />

0 100 200 300 400<br />

(a)<br />

Evaporation time [s]<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

Droplet volume [µl]<br />

161<br />

0.0<br />

0 100 200 300 400<br />

Figure 5.33: Evaporation <strong>of</strong> a pure ethanol <strong>droplet</strong> at 25°C <strong>and</strong> 4.1% relative humidity in still air. The initial volume <strong>of</strong><br />

the <strong>droplet</strong> was 2.2 µl at an initial effective SPL <strong>of</strong> 161.9 dB. (a) Change <strong>of</strong> size <strong>and</strong> volume with evaporation time. (b)<br />

Influence <strong>of</strong> the volume on the effective SPL.<br />

The diffusion-controlled evaporation <strong>of</strong> a single ethanol <strong>droplet</strong> calculated according to<br />

Equation 2.89 has an approx. 2.2 minutes longer evaporation time than the experimental<br />

determination <strong>of</strong> a levitated ethanol <strong>droplet</strong> with the same initial <strong>droplet</strong> size. The Sherwood<br />

number calculated from the experimental data starts at 4.0 <strong>and</strong> decreases to 2.2 at the end <strong>of</strong> the<br />

evaporation process. As seen with water, the Sherwood number increases with shear stress at the<br />

<strong>droplet</strong> surface caused by the inner acoustic streaming [Rensink 2004; Yarin et al. 1999]. To<br />

account for increased mass transfer via inner acoustic streaming, the evaporation rate <strong>and</strong> the<br />

Sherwood number are calculated again using the levitation model <strong>of</strong> Equation 3.40 <strong>and</strong><br />

Equation 3.34. The results <strong>and</strong> the comparison with experimental data at 25°C <strong>and</strong> 4.7% relative<br />

humidity are shown in Figure 5.35.<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 100 200 300 400<br />

(a)<br />

Diffusion-controlled model<br />

Experimental data ν air = 0.0 m/s<br />

Evaporation time [s]<br />

Figure 5.34: Comparison <strong>of</strong> experimental ethanol data with the model for diffusion controlled evaporation <strong>of</strong> a single<br />

<strong>droplet</strong> at 25°C <strong>and</strong> 4.7% relative humidity. The initial <strong>droplet</strong> volume was 1.3 µl at an initial effective SPL <strong>of</strong><br />

162.0 dB. (a) (rS 2 / rS 0 2 ) vs. evaporation time. (b) Comparison <strong>of</strong> the Sherwood numbers.<br />

Effective SPL [dB]<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

165<br />

164<br />

163<br />

162<br />

(b)<br />

5.0<br />

4.0<br />

3.0<br />

2.0<br />

(b)<br />

Evaporation time [s]<br />

Effective SPL<br />

Droplet volume<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

Diffusion-controlled model<br />

Experimental data ν air = 0.0 m/s<br />

Droplet volume [µl]<br />

0 100 200 300 400<br />

Evaporation time [s]


126 RESULTS AND DISCUSSION<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 100 200 300 400<br />

(a)<br />

Levitation model at 25.0°C<br />

Experimental data ν air = 0.0 m/s<br />

Evaporation time [s]<br />

0 100 200 300 400<br />

Figure 5.35: Comparison <strong>of</strong> experimental data <strong>of</strong> a pure ethanol <strong>droplet</strong> with the levitation model for evaporation <strong>of</strong> a<br />

single <strong>droplet</strong> at 25°C <strong>and</strong> 4.7% relative humidity with increased Sherwood number due to he inner acoustic streaming.<br />

The initial <strong>droplet</strong> volume was 1.3 µl at an initial effective SPL <strong>of</strong> 162.0 dB. (a) (rS 2 / rS 0 2 ) with evaporation time. (b)<br />

Comparison <strong>of</strong> the Sherwood numbers.<br />

The levitation model predicts a faster evaporation than the experimentally measured one. The<br />

accumulation <strong>of</strong> solvent vapour in the toroidal vortices caused by the outer acoustic streaming now<br />

decreases the evaporation rate <strong>of</strong> the pure ethanol <strong>droplet</strong>. As seen with water this effect can be<br />

overcome by introducing an axial ventilation air stream with a velocity <strong>of</strong> about 1.0 m/s. Figure<br />

5.36 shows the comparison between the levitation model <strong>and</strong> experimental data at a ventilation<br />

velocity <strong>of</strong> 0.83 m/s. At larger ventilation velocity the radius-time course <strong>of</strong> the evaporating ethanol<br />

<strong>droplet</strong> can be modelled using the Ranz-Marshall correlation for the Sherwood number. A<br />

comparison with experimental data at a velocity <strong>of</strong> 1.67 m/s corresponding to an orifice Reynolds<br />

number <strong>of</strong> 400 can be found in Figure 5.37.<br />

Sherwood number<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

(b)<br />

Levitation model at 25.0°C<br />

Experimental data ν air = 0.0 m/s<br />

Evaporation time [s]


CHAPTER 5 RESULTS AND DISCUSSION 127<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

Figure 5.36: Comparison <strong>of</strong> the levitation model with experimental data <strong>of</strong> a pure ethanol <strong>droplet</strong> at a ventilation<br />

velocity <strong>of</strong> 0.83 m/s, 25°C <strong>and</strong> 4.7% relative humidity. The initial <strong>droplet</strong> volume was 1.3 µl at an initial effective SPL<br />

<strong>of</strong> 162.0 dB. (a) (rS 2 / rS 0 2 ) with evaporation time. (b) Comparison <strong>of</strong> the Sherwood numbers.<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 50 100 150<br />

(a)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Ranz-Marshall correlation<br />

Experimental data ν air = 1.67 m/s<br />

(Re orifice = 400)<br />

0.0<br />

0 50 100 150 200<br />

(a)<br />

Levitation model at 25.0°C<br />

Experimental data ν air = 0.83 m/s<br />

Evaporation time [s]<br />

Evaporation time [s]<br />

2.0<br />

(Re = 400)<br />

orifice<br />

0 50 100 150 200<br />

Figure 5.37: Comparison <strong>of</strong> the Ranz-Marshall correlation with experimental data from a levitated ethanol <strong>droplet</strong> at a<br />

ventilation velocity <strong>of</strong> 1.67 m/s, corresponding to an orifice Reynolds number <strong>of</strong> 400. The temperature <strong>of</strong> the <strong>drying</strong> air<br />

was 25°C with a relative humidity <strong>of</strong> 0.1%. The initial <strong>droplet</strong> volume was 1.3 µl at an initial effective SPL <strong>of</strong><br />

162.0 dB. (a) (rS 2 / rS 0 2 ) with evaporation time. (b) Comparison <strong>of</strong> the Sherwood numbers.<br />

The influence <strong>of</strong> elevated temperatures was now investigated using either the levitation model at<br />

ventilation velocities <strong>of</strong> about 1.0 m/s or the Ranz-Marshall correlation at larger velocities. All<br />

experimental comparisons were performed at 40°C (Figure 5.38) <strong>and</strong> 60°C (Figure 5.39) with a<br />

relative humidity <strong>of</strong> 0.1%.<br />

Sherwood number<br />

Sherwood number<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

(b)<br />

12.0<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

(b)<br />

Levitation model at 25.0°C<br />

Experimental data ν air = 0.83 m/s<br />

0 50 100 150<br />

Evaporation time [s]<br />

Ranz-Marshall correlation<br />

Experimental data ν air = 1.67 m/s<br />

Evaporation time [s]


128 RESULTS AND DISCUSSION<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 50 100 150 200<br />

(a)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Diffusion-controlled model<br />

Experimental data ν air = 0.0 m/s<br />

Evaporation time [s]<br />

0.0<br />

0 50 100 150 200<br />

(c)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Levitation model at 40°C<br />

Experimental data ν air = 0.0 m/s<br />

Evaporation time [s]<br />

0.0<br />

0 50 100 150 200<br />

(e)<br />

Levitation model at 40°C<br />

Experimental data ν air =0.88 m/s<br />

(Re orifice = 200)<br />

Evaporation time [s]<br />

0 50 100 150 200<br />

Figure 5.38: Comparison <strong>of</strong> diffusion model (a-b) <strong>and</strong> levitation model (c-f) with experimental data from levitated pure<br />

ethanol <strong>droplet</strong>s at 40°C <strong>and</strong> 0.1% relative humidity. The initial <strong>droplet</strong> volume <strong>of</strong> all experiments was 1.6 µl at an<br />

initial effective SPL <strong>of</strong> 163.0 dB. (a) Comparison <strong>of</strong> the radius-time course <strong>of</strong> the diffusion-controlled evaporation<br />

model with experimental data in still air. (b) Comparison <strong>of</strong> the Sherwood numbers <strong>of</strong> the diffusion-controlled model<br />

with experimental data in still air. (c) Comparison <strong>of</strong> the radius-time course <strong>of</strong> the levitation model with experimental<br />

data in still air. (d) Comparison <strong>of</strong> the Sherwood numbers <strong>of</strong> the levitation model with experimental data in still air. (e)<br />

Comparison <strong>of</strong> the radius-time course <strong>of</strong> the levitation model with experimental data at a ventilation velocity <strong>of</strong><br />

0.88 m/s. (f) Comparison <strong>of</strong> the Sherwood numbers <strong>of</strong> the levitation model with experimental data at a ventilation<br />

velocity <strong>of</strong> 0.88 m/s.<br />

Sherwood number<br />

Sherwood number<br />

Sherwood number<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

(b)<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

(d)<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

Diffusion-controlled model<br />

Experimental data ν air = 0.0 m/s<br />

Evaporation time [s]<br />

Levitation model at 40°C<br />

Experimental data ν air = 0.0 m/s<br />

0 50 100 150 200<br />

Evaporation time [s]<br />

2.0<br />

(Re = 200)<br />

orifice<br />

0 50 100 150 200<br />

(f)<br />

Levitation model at 40°C<br />

Experimental data ν air =0.88 m/s<br />

Evporation time [s]


CHAPTER 5 RESULTS AND DISCUSSION 129<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Diffusion-controlled model at 60°C<br />

Experimental data ν air = 0.00 m/s<br />

0.0<br />

0 20 40 60 80 100 120<br />

(a)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Evaporation time [s]<br />

0.0<br />

0 20 40 60 80 100 120<br />

(c)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Levitation model at 60°C<br />

Experimental data ν air = 0.00 m/s<br />

Evaporation time [s]<br />

0.0<br />

0 20 40 60 80 100 120<br />

(e)<br />

Levitation model at 60°C<br />

Experimental data ν air = 0.99 m/s<br />

Re orifice = 200<br />

Evaporation time [s]<br />

0 20 40 60 80 100 120<br />

Figure 5.39: Comparison <strong>of</strong> diffusion model (a-b) <strong>and</strong> levitation model (c-f) with experimental data from levitated pure<br />

ethanol <strong>droplet</strong>s at 60°C <strong>and</strong> 0.1% relative humidity. The initial <strong>droplet</strong> volume <strong>of</strong> all experiments was 1.5 µl at an<br />

initial effective SPL <strong>of</strong> 163.25 dB. (a) Comparison <strong>of</strong> the radius-time course <strong>of</strong> the diffusion-controlled evaporation<br />

model with experimental data in still air. (b) Comparison <strong>of</strong> the Sherwood numbers <strong>of</strong> the diffusion-controlled model<br />

with experimental data in still air. (c) Comparison <strong>of</strong> the radius-time course <strong>of</strong> the levitation model with experimental<br />

data in still air. (d) Comparison <strong>of</strong> the Sherwood numbers <strong>of</strong> the levitation model with experimental data in still air. (e)<br />

Comparison <strong>of</strong> the radius-time course <strong>of</strong> the levitation model with experimental data at a ventilation velocity <strong>of</strong><br />

0.99 m/s. (f) Comparison <strong>of</strong> the Sherwood numbers <strong>of</strong> the levitation model with experimental data at a ventilation<br />

velocity <strong>of</strong> 0.99 m/s.<br />

Sherwood number<br />

Sherwood number<br />

Sherwood number [-]<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

(b)<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

(d)<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

Diffusion-controlled model at 60°C<br />

Experimental data ν air = 0.00 m/s<br />

Evaporation time [s]<br />

Levitation model at 60°C<br />

Experimental data ν air = 0.00 m/s<br />

0 20 40 60 80 100 120<br />

Evaporation time [s]<br />

2.0<br />

Re = 200<br />

orifice<br />

0 20 40 60 80 100 120<br />

(f)<br />

Levitation model at 60°C<br />

Experimental data ν air = 0.99 m/s<br />

Evaporation time [s]


130 RESULTS AND DISCUSSION<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 50 100 150<br />

(a)<br />

Ranz-Marshall correlation<br />

Experimental data ν air = 1.77 m/s<br />

(Re orifice = 400)<br />

Evaporation time [s]<br />

2.0<br />

0 50 100 150<br />

Figure 5.40: Comparison <strong>of</strong> the Ranz-Marshall correlation with experimental data from a levitated pure ethanol <strong>droplet</strong><br />

at 40°C <strong>and</strong> 0.1% humidity. The ventilation velocity υair was 1.77 m/s (Reorifice = 400). The initial <strong>droplet</strong> volume was<br />

1.6 µl at an initial effective SPL <strong>of</strong> 163.0 dB. (a) (rS 2 / rS 0 2 ) vs. evaporation time <strong>of</strong> the Ranz-Marshall correlation <strong>and</strong><br />

experimental data at υair = 1.77 m/s (b) Comparison <strong>of</strong> the Sherwood numbers between the Ranz-Marshall correlation<br />

<strong>and</strong> experimental data at υair = 1.77 m/s.<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 20 40 60 80 100<br />

(a)<br />

Ranz-Marshall model at 60°C<br />

Experimental data ν air = 1.98 m/s<br />

(Re orifice = 200)<br />

Evaporation time [s]<br />

Figure 5.41: Comparison <strong>of</strong> the Ranz-Marshall correlation with experimental data from a levitated pure ethanol <strong>droplet</strong><br />

at 60°C <strong>and</strong> 0.1% humidity. The ventilation velocity νair was 1.98 m/s (Reorifice = 400). The initial <strong>droplet</strong> volume was<br />

1.3 µl at an initial effective SPL <strong>of</strong> 163.25 dB. (a) (rS 2 / rS 0 2 ) vs. evaporation time <strong>of</strong> the Ranz-Marshall correlation <strong>and</strong><br />

experimental data at νair = 1.98 m/s (b) Comparison <strong>of</strong> the Sherwood numbers between the Ranz-Marshall correlation<br />

<strong>and</strong> experimental data at νair = 1.98 m/s.<br />

By considering the effects <strong>of</strong> inner <strong>and</strong> outer acoustic streaming, the radius-time course <strong>of</strong> an<br />

evaporating pure ethanol <strong>droplet</strong> at temperatures <strong>of</strong> 40°C <strong>and</strong> 60°C with a ventilation velocity <strong>of</strong><br />

approximately 1.0 m/s can be accurately described using Equation 3.40 <strong>and</strong> Equation 3.34 (Figure<br />

5.38, Figure 5.39). At larger ventilation velocities the Ranz-Marshall correlation according to<br />

Equation 2.57 provides a suitable model to predict the radius-time course in the first part <strong>of</strong> <strong>droplet</strong><br />

evaporation (Figure 5.40; Figure 5.41). With decreasing <strong>droplet</strong> size a difference between predicted<br />

Sherwood number<br />

Sherwood number<br />

14.0<br />

12.0<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

(b)<br />

12.0<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

Ranz-Marshall correlation<br />

Experimental data ν air = 1.77 m/s<br />

(Re orifice = 400)<br />

Evaporation time [s]<br />

Ranz-Marshall model at 60°C<br />

Experimental data ν air = 1.98 m/s<br />

2.0<br />

(Re = 200)<br />

orifice<br />

0 20 40 60 80 100<br />

(b)<br />

Evaporation time [s]


CHAPTER 5 RESULTS AND DISCUSSION 131<br />

radius-time course <strong>and</strong> experimental data occurs. This can be explained by looking at the Sherwood<br />

numbers predicted by the levitation model <strong>and</strong> the Ranz-Marshall correlation. In the first part <strong>of</strong><br />

evaporation the Sherwood number calculated by the Ranz-Marshall equation exceeds the Sherwood<br />

number caused by the inner acoustic streaming. With decreasing <strong>droplet</strong> size the Sherwood number<br />

by Ranz-Marshall decreases also due to a decrease in the <strong>droplet</strong> Reynolds number. At 40°C, for<br />

example, the Sherwood number decreases to values lower than 6.0. Figure 5.39 f shows that the<br />

Sherwood number caused by the inner acoustic streaming is always larger than 6.0 during the whole<br />

evaporation experiment. In the final evaporation phase the acoustic Sherwood number exceeds the<br />

ventilation Sherwood number, leading to faster evaporation than predicted by the Ranz-Marshall<br />

correlation.<br />

An altered <strong>drying</strong> behaviour <strong>of</strong> ethanol <strong>droplet</strong>s occurs with a relative humidity <strong>of</strong> the <strong>drying</strong><br />

air ≥ 5.0%. The radius-time course differs from the nearly linear style in dry air.<br />

Figure 5.42 shows a comparison <strong>of</strong> the raw data at 50°C <strong>and</strong> relative humidity form 0.1 to 80%.<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

Drying temperature 50.0°C<br />

0 50 100 150 200 250<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Rel. humidity 0.1%<br />

Rel. humidity 20.0%<br />

Rel. humidity 40.0%<br />

Rel. humidity 60.0%<br />

Rel. humidity 80.0%<br />

Figure 5.42: Raw data from the evaporation <strong>of</strong> ethanol <strong>droplet</strong>s in still air at 50°C <strong>and</strong> different relative humidity. The<br />

initial <strong>droplet</strong> size <strong>of</strong> all experiments was 1.5 µl at an initial effective SPL <strong>of</strong> 162.25 dB.<br />

Two different evaporation stages occur for the pure ethanol <strong>droplet</strong>s at higher relative humidity <strong>of</strong><br />

the <strong>drying</strong> air. An initial phase with a large evaporation rate is followed by a transition phase with a<br />

constantly decreasing rate. Afterwards, a second phase with a lower evaporation rate compared to<br />

the first is detected. The vaporisation <strong>of</strong> ethanol leads to a cooling <strong>of</strong> the <strong>droplet</strong> surface resulting in<br />

attainment <strong>of</strong> the wet bulb temperature. If there is sufficient humidity in the air <strong>and</strong> the wet bulb<br />

temperature falls below the dew point. There condensation <strong>of</strong> water vapour at the <strong>droplet</strong> surface


132 RESULTS AND DISCUSSION<br />

occurs. Law et al. [1987] showed that the vaporization <strong>of</strong> pure volatile alcohols in humid air is<br />

accompanied by the simultaneous condensation <strong>of</strong> water vapour on the <strong>droplet</strong> surface <strong>and</strong> its<br />

subsequent diffusion into the <strong>droplet</strong> interior. In the first phase, the evaporation <strong>of</strong> pure ethanol<br />

dominates. An increasing humidity results, however, in an increasing evaporation rate <strong>of</strong> the ethanol<br />

<strong>droplet</strong>s in this first part <strong>of</strong> evaporation. The associated heat release at the <strong>droplet</strong> surface due to the<br />

condensation increases the initial evaporation rate <strong>of</strong> the ethanol. The mixture <strong>of</strong> ethanol <strong>and</strong> water<br />

formed has a higher wet bulb temperature than pure ethanol, resulting in a larger evaporation rate.<br />

Figure 5.43 a shows the values <strong>of</strong> calculated dew point temperature at the <strong>droplet</strong> surface with<br />

increasing humidity <strong>of</strong> the <strong>drying</strong> air at 40°C, 50°C <strong>and</strong> 60°C. Figure 5.43 b shows the initial<br />

evaporation rate <strong>of</strong> the ethanol <strong>droplet</strong>s vaporizing in still air. Due to the effect <strong>of</strong> the condensed<br />

water vapour on the surface temperature, the evaporation rate in the first phase increases not only<br />

with increasing temperature but also with increasing humidity.<br />

Calculated surface temperature [°C]<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

40°C <strong>drying</strong> air<br />

50°C <strong>drying</strong> air<br />

60°C <strong>drying</strong> air<br />

0 10 20 30 40<br />

Relative humidity <strong>of</strong> <strong>drying</strong> air [%]<br />

0 20 40 60 80<br />

Figure 5.43: Surface temperature <strong>and</strong> initial evaporation coefficient <strong>of</strong> ethanol <strong>droplet</strong>s in still air without ventilation at<br />

different temperature <strong>and</strong> humidity. The initial <strong>droplet</strong> size was 1.5 µl at an initial effective SPL <strong>of</strong> 162.25 dB. (a) Plot<br />

<strong>of</strong> the calculated surface temperature <strong>of</strong> the ethanol <strong>droplet</strong>s. (b) Initial evaporation coefficient with relative humidity<br />

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

After all <strong>of</strong> the ethanol has evaporated, the subsequent radius-time-course is dominated by the<br />

evaporation <strong>of</strong> the condensed water vapour. The evaporation rate in this stage is equal to that <strong>of</strong><br />

pure water <strong>droplet</strong>s under the same ambient conditions. Figure 5.44 shows the terminal evaporation<br />

coefficient <strong>of</strong> the ethanol experiments at elevated humidity at the end <strong>of</strong> the second evaporation<br />

phase. The evaporation coefficient was determined starting at 0.05 µl residual <strong>droplet</strong> volume to the<br />

end <strong>of</strong> the evaporation process.<br />

Initial evaporation rate β r [mm 2 /s]<br />

0.016<br />

0.012<br />

0.008<br />

0.004<br />

0.000<br />

40°C <strong>drying</strong> air<br />

50°C <strong>drying</strong> air<br />

60°C <strong>drying</strong> air<br />

70°C <strong>drying</strong> air<br />

Relative humidity <strong>of</strong> <strong>drying</strong> air [%]


CHAPTER 5 RESULTS AND DISCUSSION 133<br />

Terminal evaporation rate β r [mm 2 /s]<br />

0.0016<br />

0.0012<br />

0.0008<br />

0.0004<br />

0.0000<br />

0 20 40 60 80<br />

Relative humidity <strong>of</strong> <strong>drying</strong> air [%]<br />

40°C<br />

50°C<br />

60°C<br />

70°C<br />

Figure 5.44: Terminal evaporation coefficient <strong>of</strong> the ethanol experiments at different humidity <strong>and</strong> temperature <strong>of</strong> the<br />

<strong>drying</strong> air. The coefficients were determined in a range <strong>of</strong> 50 to 0 nl <strong>droplet</strong> volume at the end <strong>of</strong> each experiment. The<br />

initial <strong>droplet</strong> volume was 1.5 µl at an initial effective SPL <strong>of</strong> 162.25 dB.<br />

A comparison with the values from the pure water <strong>droplet</strong> experiments in the same size range <strong>and</strong><br />

still air under the same ambient conditions shows almost perfect match <strong>of</strong> the evaporation<br />

coefficients at all tested temperatures. The data for 40°C <strong>and</strong> 60°C at different humidity is shown in<br />

Figure 5.45. The experiments at 60°C were repeated with forced air flow. The evaporation<br />

coefficient <strong>of</strong> each experiment was determined initially at a volume <strong>of</strong> 1.5 to 1.3 µl <strong>and</strong> in the end at<br />

a volume <strong>of</strong> 50 nl to 0 nl. The radius-time course <strong>of</strong> the raw data can be seen in Figure 5.46.<br />

Terminal evaporation constant β r [mm 2 /s]<br />

0.0014<br />

0.0012<br />

0.0010<br />

0.0008<br />

0.0006<br />

0.0004<br />

0.0002<br />

Condensed water vapour at 40°C<br />

Water <strong>droplet</strong> at 40°C<br />

Condensed water vapour at 60°C<br />

Water <strong>droplet</strong> at 60°C<br />

0.0000<br />

0 20 40 60 80<br />

Relative humidity <strong>of</strong> <strong>drying</strong> air [%]<br />

Figure 5.45: Comparison between the terminal evaporation coefficient <strong>of</strong> ethanol <strong>droplet</strong>s <strong>and</strong> the terminal evaporation<br />

coefficient <strong>of</strong> pure water <strong>droplet</strong>s with different humidity at 40°C <strong>and</strong> 60°C. The coefficients were determined in a<br />

range <strong>of</strong> 50 to 0 nl <strong>of</strong> the <strong>droplet</strong> volume at the end <strong>of</strong> each experiment. The initial <strong>droplet</strong> volume was 1.5 µl at an<br />

initial effective SPL <strong>of</strong> 162.25 dB


134 RESULTS AND DISCUSSION<br />

0.0<br />

0 50 100 150 200 250 300 350<br />

Figure 5.46: Raw data <strong>of</strong> evaporation experiments <strong>of</strong> ethanol <strong>droplet</strong>s at 60°C <strong>and</strong> different humidity <strong>of</strong> the <strong>drying</strong> air.<br />

The ventilation velocity was 1.0 m/s <strong>and</strong> corresponded to an orifice Reynolds number <strong>of</strong> 200. The initial <strong>droplet</strong> volume<br />

<strong>of</strong> each experiment was 1.5 µl at an initial effective SPL <strong>of</strong> 163.25 dB.<br />

As before with elevated humidity in still air, the radius-time course with <strong>drying</strong> air velocity shows<br />

two different evaporation stages. The fast size-change at the beginning is caused by the evaporation<br />

<strong>of</strong> ethanol. Subsequently, the evaporation rate decreases to almost constant values in the second<br />

evaporation stage. This stage is dominated by evaporation <strong>of</strong> the condensed water vapour at the<br />

<strong>droplet</strong> surface. Figure 5.47 shows the initial evaporation coefficient as function <strong>of</strong> <strong>drying</strong> air<br />

velocity at different temperature <strong>and</strong> humidity. As seen before in still air, the initial evaporation<br />

coefficient increases with increasing humidity <strong>and</strong> temperature <strong>of</strong> the <strong>drying</strong> air.<br />

Evaporation coefficient β r [mm 2 /s]<br />

0.014<br />

0.012<br />

0.010<br />

0.008<br />

0.006<br />

0.004<br />

0.002<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

25°C; 0.1%<br />

40°C; 0.1%<br />

60°C; 0.1%<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0 0.5 1.0 1.5 2.0<br />

Drying air velocity ν air [m/s]<br />

60.0°C; 0.1% rel. humidity<br />

60.0°C; 5.0% rel. humidity<br />

60.0°C; 20.0% rel. humidity<br />

60.0°C; 40.0% rel. humidity<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

0.0 0.5 1.0 1.5 2.0<br />

Figure 5.47: Initial evaporation coefficient <strong>of</strong> ethanol <strong>droplet</strong>s as a function <strong>of</strong> <strong>drying</strong> air velocity at different<br />

temperature <strong>and</strong> humidity <strong>of</strong> the <strong>drying</strong> air. The initial <strong>droplet</strong> volume <strong>of</strong> each experiment was 1.5 µl at an initial<br />

effective SPL <strong>of</strong> 162.75 dB to 164.15. (a) Initial evaporation coefficient versus <strong>drying</strong> air velocity for different<br />

temperatures. (b) Initial evaporation coefficient versus <strong>drying</strong> air velocity for different humidity.<br />

Evaporation coefficient β r [mm 2 /s]<br />

0.006<br />

0.005<br />

0.004<br />

0.003<br />

0.002<br />

0.001<br />

Drying temperature 60°C<br />

Rel. humidity 0.1%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

Drying air velocity ν air [m/s]


CHAPTER 5 RESULTS AND DISCUSSION 135<br />

Figure 5.48 presents the initial <strong>and</strong> terminal evaporation coefficients at 60°C <strong>and</strong> 1.0 m/s ventilation<br />

velocity, as well as a comparison <strong>of</strong> the terminal evaporation coefficient with values <strong>of</strong> pure water<br />

under the same conditions. The terminal evaporation coefficient correlates to the values for pure<br />

water under the same ambient conditions, <strong>and</strong> decreases with increasing humidity. A simple<br />

analytical prediction <strong>of</strong> the evaporation process using Equation 3.45 is not possible, because a<br />

rapid-mixing is used to describe the time-distribution <strong>of</strong> the two components. This assumes that all<br />

transport processes <strong>of</strong> the components inside the <strong>droplet</strong> occur much faster than the change in<br />

concentration at the <strong>droplet</strong> surface. This results in a homogeneous distribution <strong>of</strong> the components<br />

over the radial axis <strong>of</strong> the <strong>droplet</strong>. In case <strong>of</strong> condensation <strong>of</strong> water vapour on the <strong>droplet</strong> surface,<br />

however, rapid-mixing cannot describe this behaviour, because the condensed water vapour is<br />

initially located only at the <strong>droplet</strong> surface [Kastner 2001].<br />

Initial evaporation coefficient β r [mm 2 /s]<br />

Terminal evaporation coefficient β r [mm 2 /s]<br />

0.0260<br />

0.0240<br />

0.0220<br />

0.0200<br />

0.0180<br />

0.0160<br />

0.0140<br />

0.0120<br />

0.0100<br />

(a)<br />

0.0022<br />

0.0020<br />

0.0018<br />

0.0016<br />

0.0014<br />

0.0012<br />

0.0010<br />

0.0008<br />

0.0006<br />

(b)<br />

Initial β r <strong>of</strong> ethanol<br />

Surface temperature<br />

0 10 20 30 40<br />

Relative humidity <strong>of</strong> <strong>drying</strong> air [%]<br />

Terminal β r <strong>of</strong> water <strong>droplet</strong>s<br />

β r <strong>of</strong> condensed water vapour<br />

0 10 20 30 40<br />

Relative humidity <strong>of</strong> <strong>drying</strong> air [%]<br />

Figure 5.48: Initial <strong>and</strong> terminal evaporation coefficient <strong>of</strong> ethanol <strong>droplet</strong>s at 60°C <strong>and</strong> different humidity <strong>of</strong> the<br />

<strong>drying</strong> air. The ventilation velocity was 1.0 m/s corresponding to an orifice Reynolds number <strong>of</strong> 200. The initial <strong>droplet</strong><br />

volume <strong>of</strong> each experiment was 1.5 µl at an initial effective SPL <strong>of</strong> 163.25 dB. (a) Initial evaporation coefficient <strong>and</strong><br />

<strong>droplet</strong> surface temperature with humidity. (b) Comparison <strong>of</strong> the terminal evaporation coefficient <strong>of</strong> ethanol <strong>droplet</strong> at<br />

elevated humidity <strong>of</strong> the <strong>drying</strong> air <strong>and</strong> water <strong>droplet</strong>s.<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Calculated initial surface temperature [°C]


136 RESULTS AND DISCUSSION<br />

5.2.4 Evaporation <strong>of</strong> ethanol-water mixtures<br />

To investigate the evaporation behaviour <strong>of</strong> multi-component solvent <strong>droplet</strong>s inside the acoustic<br />

levitation system, different mixtures <strong>of</strong> ethanol <strong>and</strong> water were levitated <strong>and</strong> compared to the results<br />

obtained from pure ethanol <strong>and</strong> pure water. Figure 5.49 shows the evaporation behaviour <strong>of</strong> the<br />

mixture <strong>droplet</strong>s in still air, neglecting the accumulation <strong>of</strong> solvent vapour in the toroidal vortices.<br />

The decrease in the squared radius or diameter cannot be described by the d 2 -law. As with the<br />

ethanol <strong>droplet</strong>s at high humidity, the ethanol-water mixture shows a radius-time course that is split<br />

into two parts. At the beginning both components evaporate simultaneously, with domination <strong>of</strong> the<br />

more volatile one (EtOH). Then a transition period controlled by the evaporation <strong>of</strong> both<br />

components in equal measure follows. Afterwards, the more volatile solvent has almost completely<br />

evaporated, <strong>and</strong> the subsequent evaporation is dominated by the less volatile solvent (water)<br />

[Kastner 2001].<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

40°C<br />

0.0<br />

0 200 400 600 800 1000<br />

(a)<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

water:ethanol = 10:0<br />

water:ethanol = 8:2<br />

water:ethanol = 4:6<br />

water:ethanol = 2:8<br />

water:ethanol = 0:10<br />

60°C<br />

0.0<br />

0 200 400 600<br />

Figure 5.49: Evaporation <strong>of</strong> multi-component solvent <strong>droplet</strong>s at different temperature <strong>and</strong> a humidity <strong>of</strong> 0.1%. The<br />

initial <strong>droplet</strong> volume <strong>of</strong> all experiments was 1.5 µl at an initial effective SPL <strong>of</strong> 164.0 dB. (a) Radius-time course <strong>of</strong><br />

<strong>droplet</strong>s at different water-ethanol ratio at 40°C. (b) Radius-time course <strong>of</strong> <strong>droplet</strong>s with different water-ethanol ratio at<br />

60°C.<br />

The evaporation coefficients in the initial evaporation phase <strong>and</strong> at the end <strong>of</strong> each experiment were<br />

calculated by using the first 5 data points in a size range between 1.5 µl <strong>and</strong> 1.3 µl <strong>and</strong> the last data<br />

points starting from 0.05 µl. The results are compared to the values obtained form pure water <strong>and</strong><br />

ethanol <strong>droplet</strong>s in the same size range <strong>and</strong> under the same ambient conditions. Figure 5.50 shows<br />

increasing evaporation coefficients in the initial phase with increasing amount <strong>of</strong> ethanol in the<br />

solvent mixture. The values for the mixtures are between the coefficients <strong>of</strong> pure water <strong>and</strong> pure<br />

ethanol. In the terminal part, the evaporation coefficients show almost constant values independent<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

(b)<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

water:ethanol = 10:0<br />

water:ethanol = 8:2<br />

water:ethanol = 4:6<br />

water:ethanol = 2:8<br />

water:ethanol = 0:10


CHAPTER 5 RESULTS AND DISCUSSION 137<br />

<strong>of</strong> the initial composition <strong>of</strong> the <strong>droplet</strong>. The comparison with pure water <strong>droplet</strong>s shows good<br />

agreement. This demonstrates again that the size change <strong>of</strong> the <strong>droplet</strong>s in the second evaporation<br />

phase is dominated by the evaporation <strong>of</strong> water. The experimental values in Figure 5.50 <strong>and</strong> also the<br />

raw data at different temperature show that an increase in <strong>drying</strong> temperature leads to an increase in<br />

the evaporation rate in every phase <strong>of</strong> evaporation. The different evaporation coefficients at the<br />

beginning <strong>and</strong> end <strong>of</strong> pure water <strong>droplet</strong>s are caused by the effects <strong>of</strong> inner acoustic streaming on<br />

the shrinking <strong>droplet</strong> surface, as reported in detail in chapter 5.2.2.<br />

Evaporation coefficient β r [mm 2 /s]<br />

0.005<br />

0.004<br />

0.003<br />

0.002<br />

0.001<br />

0.000<br />

(a)<br />

Initial phase at 40°C<br />

Terminal phase at 40°C<br />

10:0 8:2 6:4 4:6 2:8 0:10<br />

Water : Ethanol [-]<br />

10:0 8:2 6:4 4:6 2:8 0:10<br />

Figure 5.50: Initial <strong>and</strong> terminal evaporation coefficients <strong>of</strong> mixture <strong>droplet</strong>s <strong>of</strong> ethanol <strong>and</strong> water at 40°C <strong>and</strong> 60°C<br />

<strong>drying</strong> temperature <strong>and</strong> a humidity <strong>of</strong> 0.1%. The initial <strong>droplet</strong> diameter <strong>of</strong> each experiment was 1.5 µl at an initial<br />

effective SPL <strong>of</strong> 164.0 dB. (a) Evaporation coefficients at 40°C. (b) Evaporation coefficients at 60°C.<br />

2 ) [-]<br />

2 / rS 0<br />

( r S<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Drying temperature 25°C<br />

Drying temperature 40°C<br />

Drying temperature 60°C<br />

0.0<br />

Water : EtOH = 40 : 60<br />

0 200 400 600 800 1000<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Figure 5.51: Raw data, initial <strong>and</strong> terminal evaporation coefficients <strong>of</strong> ethanol-water mixture <strong>droplet</strong>s at different<br />

temperatures <strong>and</strong> 0.1% relative humidity. The initial <strong>droplet</strong> volume <strong>of</strong> each experiment was 1.5 µl at an effective initial<br />

SPL <strong>of</strong> 164.0 dB. (a) Radius-time course <strong>of</strong> ethanol-water (6:4) <strong>droplet</strong>s at different temperature. (b) Comparison <strong>of</strong> the<br />

initial <strong>and</strong> terminal evaporation coefficients at different temperatures<br />

Evaporation coefficient β r [mm 2 /s]<br />

Evaporation coefficient β r [mm 2 /s]<br />

0.010<br />

0.008<br />

0.006<br />

0.004<br />

0.002<br />

0.000<br />

0.010<br />

0.008<br />

0.006<br />

0.004<br />

0.002<br />

0.000<br />

(b)<br />

Initial phase at 60°C<br />

Terminal phase at 60°C<br />

Initial β r at 40°C<br />

Initial β r at 60°C<br />

Water : Ethanol [-]<br />

Terminal β r at 40°C<br />

Terminal β r at 60°C<br />

10:0 8:2 6:4 4:6 2:8 0:10<br />

Water : Ethanol [-]


138 RESULTS AND DISCUSSION<br />

5.3 Evaporation <strong>of</strong> solution <strong>droplet</strong>s<br />

5.3.1 Evaporation <strong>of</strong> maltodextrin solution <strong>droplet</strong>s<br />

Properties <strong>of</strong> maltodextrin<br />

Maltodextrin is produced by controlled hydrolysis <strong>of</strong> starch using acids or enzymes with subsequent<br />

purification <strong>and</strong> spray-<strong>drying</strong>. Different grades with a variable ratio <strong>of</strong> glucose, maltose <strong>and</strong><br />

polysaccharides are available [Burger 1998]. The use <strong>of</strong> maltodextrines as <strong>drying</strong> aids in spray<strong>drying</strong><br />

has been practiced since the 1970s, mostly in the food industry [Adhikari et al. 2004;<br />

Brennan et al. 1971]. In combination with gum Arabic <strong>and</strong> modified starch it serves a basis as a<br />

wall material in micro-encapsulation processes <strong>of</strong> oils [Bh<strong>and</strong>ari et al. 1992; Krishnan et al. 2005;<br />

Soottitantawat et al. 2003], fatty acids [Minemoto et al. 2002], flavours <strong>and</strong> other volatile<br />

substances [Bh<strong>and</strong>ari et al. 1992; Buffo 2000]. In pharmaceutical research maltodextrin has been<br />

used to produce model <strong>protein</strong> particles with BSA for the preparation <strong>of</strong> aerosol <strong>formulations</strong> by<br />

spray-<strong>drying</strong> [Lucas et al. 1997]. As an additive to penicillium occitanis cellulases <strong>formulations</strong>, it<br />

had a positive effect on enzyme stabilisation during spray-<strong>drying</strong> [Belghith et al. 2001]. For the<br />

levitation experiments the liquid properties <strong>of</strong> four maltodextrin solutions with different solid<br />

content were determined <strong>and</strong> shown in Table 5.3.<br />

Table 5.3: Properties <strong>of</strong> maltodextrin solutions with different solid content.<br />

Solid<br />

content<br />

Surface<br />

tension γ<br />

[ mg/ ml]<br />

[ ]<br />

Density<br />

ρ liquid<br />

3<br />

mN/m [ ]<br />

Maximum<br />

diameter<br />

Maximum<br />

volume<br />

Minimum<br />

diameter<br />

Minimum<br />

volume<br />

g/cm [ mm ] [ µl ] [ µm ] [ µl ]<br />

50 72.8 ± 0.12 1.015 6.57 148.49 15 1.77⋅10 -6<br />

100 73.2 ± 0.06 1.036 6.46 141.15 15 1.77⋅10 -6<br />

200 73.6 ± 0.06 1.071 6.29 130.30 15 1.77⋅10 -6<br />

400 74.5 ± 0.12 1.142 5.97 111.41 15 1.77⋅10 -6<br />

Data analysis <strong>of</strong> maltodextrin experiments<br />

Analysis <strong>of</strong> the <strong>drying</strong> data <strong>of</strong> solution <strong>droplet</strong>s is divided into at least two different stages <strong>and</strong><br />

differs therefore from the analysis <strong>of</strong> pure solvent <strong>droplet</strong>s. In the first stage, corresponding to the<br />

constant rate period [Kastner et al. 2001; Masters 1991], the squared radius <strong>of</strong> the <strong>droplet</strong> decreases


CHAPTER 5 RESULTS AND DISCUSSION 139<br />

linearly with evaporation time, similar to pure solvent <strong>droplet</strong>s. The evaporation rate is determined<br />

directly from the horizontal <strong>and</strong> vertical diameters <strong>of</strong> the levitated <strong>droplet</strong>. The data is analysed<br />

according to Chapter 5.2.1 <strong>and</strong> the evaporation rate calculated using Equation 3.47. After the<br />

critical point, which marks the transition from the first to the second <strong>drying</strong> stage, the volume <strong>of</strong> the<br />

levitated <strong>droplet</strong> remains constant even though there is still evaporation <strong>of</strong> solvent from inside the<br />

<strong>droplet</strong> [Kastner et al. 2001; Walton 1999]. Determination <strong>of</strong> the evaporation rate in this falling rate<br />

period is performed from the rise <strong>of</strong> the <strong>droplet</strong> within the stationary ultrasonic field towards its<br />

adjacent upper pressure node, <strong>and</strong> the evaporation rate calculated with Equation 3.49. To determine<br />

the critical point the plot <strong>of</strong> the squared radius with time as well as the aspect ratio is used. The<br />

effective SPL at the critical point necessary for calculation <strong>of</strong> the evaporation rate in the second<br />

<strong>drying</strong> stage is determined levitating a water <strong>droplet</strong> with the same size <strong>and</strong> under the same<br />

conditions as the solid particle at the critical point. This eliminates the effects <strong>of</strong> surface tension <strong>and</strong><br />

kinematic viscosity on the shape <strong>of</strong> the <strong>droplet</strong>. Figure 5.52 shows the raw data <strong>of</strong> a 200 mg/ml<br />

solution <strong>droplet</strong> obtained by the imaging s<strong>of</strong>tware versus evaporation time. Both, the rise in position<br />

as well as the change in aspect ratio provide a useful tool to determine the critical point between the<br />

first <strong>and</strong> second <strong>drying</strong> stages. The change in aspect ratio <strong>of</strong> the <strong>droplet</strong> / particle at the critical point<br />

is caused by influences <strong>of</strong> the st<strong>and</strong>ing acoustic wave. While the horizontal diameter does not<br />

change further due to the formation <strong>of</strong> a solid crust at the surface, the vertical diameter continues to<br />

decrease after the critical point. The section is marked by the ellipse in Figure 5.52. This is caused<br />

by the existing 5:1 ratio between axial <strong>and</strong> radial forces acting in the ultrasonic field [Lierke 1996].<br />

The particle is thus flattened during this phase <strong>of</strong> <strong>drying</strong>, resulting in an increase in aspect ratio. The<br />

deformation holds until the end <strong>of</strong> the <strong>drying</strong> process <strong>and</strong> can be seen in the front illumination<br />

picture gallery taken during the <strong>drying</strong> process (Figure 5.53) as well as in SEM pictures <strong>of</strong> the final<br />

particle. The flattening <strong>of</strong> the top <strong>of</strong> the <strong>droplet</strong> / particle during formation <strong>of</strong> a solid crust can be<br />

seen clearly. The dry particle has a translucent glassy appearance.


140 RESULTS AND DISCUSSION<br />

2 ) [-] Aspect ratio [-]<br />

2 / rS 0<br />

( r S<br />

Diameter [µm]<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

Constant rate<br />

Falling rate<br />

0.4<br />

0.2<br />

Droplet / particle size<br />

Aspect ratio<br />

Vertical position<br />

-300<br />

0.0<br />

-400<br />

0 200 400 600 800 1000<br />

1600<br />

1400<br />

1200<br />

1000<br />

Evaporation time [s]<br />

100<br />

0<br />

-100<br />

-200<br />

800<br />

600<br />

Horizontal diameter<br />

Vertical diameter<br />

Aspect ratio<br />

1.1<br />

400<br />

1.0<br />

0 200 400 600 800 1000<br />

Evaporation time [s]<br />

Figure 5.52: Raw data form the levitation <strong>of</strong> a maltodextrin solution <strong>droplet</strong> with 200 mg/ml solid content without<br />

ventilation. The <strong>drying</strong> air had a temperature <strong>of</strong> 40°C with 0.1% relative humidity. The initial <strong>droplet</strong> size was 1.6 µl at<br />

an initial effective SPL <strong>of</strong> 164.5 dB.<br />

The influences <strong>of</strong> SPL on the shape <strong>and</strong> the aspect ratio <strong>of</strong> particles are examined in <strong>drying</strong><br />

experiments under the same ambient conditions with various strengths <strong>of</strong> the ultrasonic field. The<br />

aspect ratio at the beginning <strong>of</strong> the levitation experiment, at the critical point, <strong>and</strong> at the end <strong>of</strong> the<br />

<strong>drying</strong> process is shown in Figure 5.54 for a maltodextrin solution with 200 mg/ml solid content.<br />

With an increase in the effective SPL the aspect ratio increases at all determined points examined<br />

during the <strong>drying</strong> process. This results in a greater flattening <strong>of</strong> the dry particle with increasing<br />

effective SPL. Figure 5.54 a shows the different values <strong>of</strong> the aspect ratio with initial effective SPL.<br />

The aspect ratio <strong>of</strong> the <strong>droplet</strong> decreases directly before the critical point. The formation <strong>of</strong> a solid<br />

crust is influenced by the axial levitation force, resulting in the flattening <strong>of</strong> the particle already<br />

1.4<br />

1.3<br />

1.2<br />

Distance to adjacent pressure node [µm]<br />

Aspect ratio [-]


CHAPTER 5 RESULTS AND DISCUSSION 141<br />

mentioned. The increase in the aspect ratio starts at the critical point <strong>and</strong> is completed by the end <strong>of</strong><br />

the <strong>drying</strong> process. The initial effective SPL that determines the starting aspect ratio does therefore<br />

influence the shape <strong>of</strong> the final particle (Figure 5.54 b).<br />

Figure 5.53: Evaporation <strong>of</strong> a maltodextrin solution with 100 mg/ml solid content at 27°C <strong>and</strong> 23% relative humidity<br />

without ventilation. Pictures were taken with the CCD camera every 20 seconds. The gallery starts 10 minutes after<br />

injection <strong>of</strong> the <strong>droplet</strong>. The initial <strong>droplet</strong> volume was 2.0 µl at an effective SPL <strong>of</strong> 164.75 dB.<br />

Aspect ratio [-]<br />

2.2<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

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

Critical point<br />

Final praticle<br />

1.0<br />

163.0 164.0 165.0 166.0 167.0 168.0 169.0<br />

(a)<br />

Initial effective SPL [dB]<br />

1.4<br />

1.2<br />

Experimental data<br />

Linear fit ( R<br />

1.0 1.2 1.4 1.6 1.8 2.0<br />

2 = 0.98 )<br />

Figure 5.54: Aspect ratio <strong>of</strong> levitated maltodextrin solution <strong>droplet</strong>s with 200 mg/ml solid content <strong>and</strong> dried<br />

maltodextrin particles at 60°C <strong>and</strong> 0.1% relative humidity. The initial <strong>droplet</strong> size was 1.8 µl at different initial effective<br />

SPL. (a) Aspect ratio with initial effective SPL. (b) Correlation between initial aspect ratio <strong>of</strong> the <strong>droplet</strong> <strong>and</strong> the final<br />

aspect ratio <strong>of</strong> the formed particle.<br />

Aspect ratio <strong>of</strong> the final particle [-]<br />

2.2<br />

2.0<br />

1.8<br />

1.6<br />

(b)<br />

Aspect ratio <strong>of</strong> the initial <strong>droplet</strong> [-]


142 RESULTS AND DISCUSSION<br />

There are a number <strong>of</strong> possibilities <strong>of</strong> plotting the evaporation rate <strong>of</strong> the maltodextrin solution<br />

<strong>droplet</strong>s, as already seen with the pure solvent <strong>droplet</strong>s in Chapter 5.2.1. The first is to plot the<br />

evaporation rate in mass per time unit on the y-axis versus the time on the x-axis (Figure 5.55 a). A<br />

decreasing evaporation rate with time is seen due to both the decrease in surface area in the first<br />

<strong>drying</strong> stage <strong>and</strong> the impeded solvent evaporation in the second stage. This plot shows different<br />

values <strong>of</strong> evaporation rate with different initial volume <strong>of</strong> the <strong>droplet</strong>. This is normalised in a plot <strong>of</strong><br />

the evaporation rate in mass per time <strong>and</strong> surface area versus time divided by the squared initial<br />

radius (Figure 5.55 b). The calculation <strong>of</strong> an average curve is now possible, if the influence <strong>of</strong> the<br />

inner acoustic streaming on the <strong>droplet</strong> surface is neglected. Figure 5.55b shows an almost constant<br />

evaporation rate in the first <strong>drying</strong> stage, a clear break at the critical point, <strong>and</strong> falling values<br />

subsequently in the second <strong>drying</strong> stage.<br />

Evaporation rate [mg/s]<br />

6.0<br />

5.0<br />

4.0<br />

3.0<br />

2.0<br />

1.0<br />

0.0<br />

0 200 400 600 800 1000<br />

(a)<br />

Evaporation time [s]<br />

Initial volume 2.249 µl<br />

Initial volume 1.815 µl<br />

Initial volume 1.657 µl<br />

Figure 5.55: Evaporation rate <strong>of</strong> maltodextrin <strong>droplet</strong>s with 200 mg/ml solid content dried at 40°C <strong>and</strong> a relative<br />

humidity <strong>of</strong> 0.1% without ventilation air stream. Initial <strong>droplet</strong> volume <strong>and</strong> initial effective SPL differed at the three<br />

experiments. (a) Evaporation rate in evaporated mass per time unit with the evaporation time. (b) Evaporation rate in<br />

evaporated mass per time unit <strong>and</strong> surface area with evaporation time divided by the squared initial radius.<br />

The influence <strong>of</strong> the solid content on the radius-time course <strong>and</strong> the evaporation rate is shown in<br />

Figure 5.56. With increasing solid concentration the evaporation time <strong>of</strong> the first <strong>drying</strong> stage<br />

decreases leading to a shift in the critical point to earlier times. The 40% solution does not show a<br />

clear transition between the two <strong>drying</strong> stages. The evaporation rate decreases also due to water<br />

vapour pressure depression by the dissolved molecules. The decline in the evaporation rate in the<br />

second <strong>drying</strong> stage becomes shallower with higher solid content. The duration <strong>of</strong> the second <strong>drying</strong><br />

stage increases therefore with higher solid content (Table 5.4).<br />

Evaporation rate [mg/s/mm 2 ]<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1 Constant Falling<br />

0.0<br />

rate rate<br />

0 500 1000 1500 2000<br />

(b)<br />

Critical point<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Initial volume 2.249 µl<br />

Initial volume 1.815 µl<br />

Initial volume 1.657 µl


CHAPTER 5 RESULTS AND DISCUSSION 143<br />

Table 5.4: Constant <strong>and</strong> falling rate stages for 1.8 µl maltodextrin <strong>droplet</strong>s with different solid content dried at a<br />

temperature <strong>of</strong> 40°C <strong>and</strong> 0.1% relative humidity.<br />

Solid content<br />

[ ]<br />

Evaporation rate in<br />

constant rate period<br />

−1 −2<br />

mg/ml [ µg s ⋅ mm ]<br />

⋅ [ ]<br />

Critical point tcrit Duration <strong>of</strong> falling<br />

rate period tcrit<br />

s [] s<br />

50 0.75-0.63 445.3 162.4<br />

100 0.71-0.63 353.9 278.3<br />

200 0.62-0.58 283.6 364.6<br />

400 0.58-0.50 91.5 708.6<br />

Evaporation rate [µg/s/mm 2 ]<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 500 1000 1500 2000<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

50 mg/ml<br />

100 mg/ml<br />

200 mg/ml<br />

400 mg/ml<br />

Figure 5.56: Evaporation rate <strong>of</strong> maltodextrin solutions with different solid content at 40°C <strong>and</strong> a relative humidity <strong>of</strong><br />

0.1% without ventilation air stream. The initial <strong>droplet</strong> volume <strong>of</strong> all experiments was 1.8 µl at an initial effective SPL<br />

<strong>of</strong> 164.75 dB.<br />

The temperature <strong>of</strong> the <strong>drying</strong> air has also an influence on the evaporation time <strong>and</strong> the evaporation<br />

rate <strong>of</strong> solution <strong>droplet</strong>s. Higher temperature leads to an increase in evaporation rate <strong>and</strong> a decrease<br />

in evaporation time in the first <strong>drying</strong> stage Figure 5.57. The evaporation rate is almost constant in<br />

this phase. The higher the temperature is, the earlier the critical point is reached <strong>and</strong> therefore the<br />

formation <strong>of</strong> a solid crust leading to a decrease in the duration <strong>of</strong> the first <strong>and</strong> an increase in the<br />

second <strong>drying</strong> stage Table 5.5. At 25°C the transition from constant to falling rate stage does not<br />

result in a sharp point compared to 40°C <strong>and</strong> 60°C.


144 RESULTS AND DISCUSSION<br />

Table 5.5: Duration <strong>of</strong> constant <strong>and</strong> falling rate for 1.8 µl maltodextrin <strong>droplet</strong>s with 200 mg/ml solid content dried at a<br />

different temperature <strong>and</strong> 0.1% relative humidity.<br />

Drying temperature<br />

[ ]<br />

Constant rate period<br />

⎡ µg ⎤<br />

mg/ml ⎢ 2<br />

⎣s⋅<br />

mm ⎥<br />

⎦<br />

Critical point<br />

tcrit<br />

Duration <strong>of</strong> falling<br />

rate period tcrit<br />

[ s ]<br />

[] s<br />

25°C 0.39-0.36 502.5 628.0<br />

40°C 0.62-0.58 283.6 364.6<br />

60°C 1.22-1.19 111.7 333.9<br />

Evaporation rate [mg/s/mm 2 ]<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 500 1000 1500 2000<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Drying temp. 25°C<br />

Drying temp. 40°C<br />

Drying temp. 60°C<br />

Figure 5.57: Evaporation rate <strong>of</strong> maltodextrin solutions with 200 mg/ml solid content at different temperature <strong>and</strong> a<br />

relative humidity <strong>of</strong> 0.1% without ventilation air stream. The initial <strong>droplet</strong> volume <strong>of</strong> all experiments was 1.8 µl at an<br />

initial effective SPL <strong>of</strong> 164.75 dB.<br />

The formed particles differ in size. Figure 5.58 shows that the particle size not only increases with<br />

solid content <strong>of</strong> the maltodextrin solution (Figure 5.58 a), but also with the temperature <strong>of</strong> the<br />

<strong>drying</strong> air (Figure 5.58 b). For a 1.8 µl <strong>droplet</strong> with 50 mg/ml solid content the volume <strong>of</strong> the<br />

formed particle increases from 0.107 µl at 25°C to 0.129 µl at 40°C. A comparison <strong>of</strong> the final<br />

particle densities shows decreasing density with higher <strong>drying</strong> temperature (Figure 5.58 c). The<br />

density also slightly increases with greater solid content. The difference between the final particle<br />

size <strong>and</strong> density values at 40°C <strong>and</strong> 60°C are, however, small compared to the values at 25°C. SEM<br />

pictures <strong>of</strong> maltodextrin particles produced in the levitation system <strong>and</strong> in the Buchi spray dryer<br />

with different solid content are shown in Figure 5.59. The maltodextrin solution was sprayed at


CHAPTER 5 RESULTS AND DISCUSSION 145<br />

100°C inlet <strong>and</strong> 60°C outlet temperature. The levitated particles show a similar morphology<br />

compared to the larger particles <strong>of</strong> the spray dried powder. The interior <strong>of</strong> the particles has a<br />

massive solid structure without any holes. The surface is indented by small flat craters which are<br />

not combined with capillaries in the interior.<br />

2 ) [-]<br />

( r S / r S 0<br />

Particle density [g/cm 3 ]<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

50 mg/ml<br />

100 mg/ml<br />

200 mg/ml<br />

0.0<br />

400 mg/ml<br />

0 500 1000 1500<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

[ Temp. 40°C - rel. humidity 0.1%]<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Drying temp. 25°C<br />

Drying temp. 40°C<br />

Drying temp. 60°C<br />

0.6<br />

0 10 20 30 40<br />

Maltodextrin solid content (m/V) [%]<br />

0 10 20 30 40<br />

Figure 5.58: Comparison <strong>of</strong> the resulting particle size <strong>and</strong> density <strong>of</strong> maltodextrin solutions with different solid content<br />

dried at different temperature <strong>and</strong> 0.1% relative humidity. The initial <strong>droplet</strong> volume <strong>of</strong> all experiments was 1.8 µl at an<br />

initial effective SPL between 164.0 <strong>and</strong> 164.75 dB. (a) Radius-time course <strong>of</strong> Maltodextrin solutions with different solid<br />

content at 40°C. (b) Particle size compared to the initial <strong>droplet</strong> volume wit different solid content <strong>and</strong> temperature.<br />

(c) Resulting particle density with different solid content <strong>and</strong> temperature.<br />

Particle radius relativ to initial <strong>droplet</strong> radius [%]<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

Drying temp. 25°C<br />

Drying temp. 40°C<br />

Drying temp. 60°C<br />

Maltodextrin solid content (m/V) [%]


146 RESULTS AND DISCUSSION<br />

(a)<br />

(b)<br />

(c)<br />

Figure 5.59: SEM pictures <strong>of</strong> maltodextrin particles from the levitation system <strong>and</strong> the Buchi B-191 Mini Spray Dryer.<br />

(a-c) Maltodextrin particles from a solution with 100 mg/ml solid content dried at 60°C <strong>and</strong> 0.1% relative humidity.<br />

Shown are pictures with a magnification <strong>of</strong> 75.5-times (a), 180-time (b) <strong>and</strong> 1000-times (c). (d-f) Spray-dried<br />

maltodextrin particles from solutions with 100 mg/ml (d), 200 mg/ml (e) <strong>and</strong> 400 mg/ml (f) solid content (m/V) at an<br />

inlet temperature <strong>of</strong> 100°C <strong>and</strong> an outlet temperature <strong>of</strong> 60°C. The pictures were taken with 2000-times magnification.<br />

(d)<br />

(e)<br />

(f)


CHAPTER 5 RESULTS AND DISCUSSION 147<br />

5.3.2 Evaporation <strong>of</strong> trehalose solution <strong>droplet</strong>s<br />

Properties <strong>of</strong> trehalose<br />

Trehalose (α,α-trehalose) is a disaccharide formed by the 1,1-linkage <strong>of</strong> two D-glucose molecules<br />

[Richards et al. 2002]. It has the molecular formula C12H22O11 <strong>and</strong> a molecular weight <strong>of</strong><br />

342.31 g/mol. Commercially <strong>and</strong> purified, it is usually found as dihydrate [Sigma 2005]. Trehalose<br />

is a non-reducing sugar, that first melts at 97°C. Additional heat drives <strong>of</strong>f the water <strong>of</strong><br />

crystallisation until the material resolidifies at 130°C. The anhydrous Trehalose melts at 203°C<br />

[Richards et al. 2002]. The natural occurrence <strong>of</strong> α,α-trehalose is listed in over 80 different species<br />

<strong>of</strong> plants, algae, fungi, yeasts, bacteria insects <strong>and</strong> other vertebrates. The biosynthesis <strong>of</strong> trehalose in<br />

brewer’s yeast is catalysed by enzymes facilitating the reaction <strong>of</strong> UDP-D-glucose with D-<br />

glucose 6-phosphate. The phosphate <strong>of</strong> the resulting α,α-trehalose 6-phosphate is enzymatically<br />

removed leaving a trehalose molecule [Richards et al. 2002]. The role <strong>of</strong> trehalose in nature varies<br />

from energy source during certain stages <strong>of</strong> development or the flight <strong>of</strong> insects to structural<br />

component [Elbein 1974]. One very important observation is the participation <strong>of</strong> trehalose in<br />

stabilisation <strong>of</strong> life processes in organisms that can survive either freezing or dehydration [Newman<br />

et al. 1993]. It can stabilize the dehydrated form <strong>of</strong> biological structures <strong>and</strong> is able to restore them<br />

functional as soon as the hydration <strong>and</strong> temperature conditions return to normal [Lins et al. 2004].<br />

Two different stabilization levels are discussed [Lins et al. 2004]: one is the stabilization <strong>of</strong><br />

membranes <strong>and</strong> lipid assemblies at very low hydration [Hoekstra et al. 1997] <strong>and</strong> the other is the<br />

stabilisation <strong>of</strong> biological macromolecules in the folded state under denaturation conditions [Xie<br />

1997]. Due to the wide appearance in different biostructural classes, the protective mechanisms <strong>of</strong><br />

trehalose against stress by dehydration, heating or freezing is said to be a non-specific process<br />

[Crowe et al. 1990]. Three main hypothesis are discussed describing the stabilizing mechanisms <strong>of</strong><br />

trehalose [Lins et al. 2004]. The first one is called the water-replacement hypothesis. It states<br />

direct interactions between the disaccharide molecules <strong>and</strong> the protected biostructure through<br />

hydrogen bonds [Arakawa et al. 2001]. The second mechanism is the water-layer hypothesis,<br />

discussing a trapping <strong>of</strong> the water molecules close to the biomolecular surface layer [Gil et al.<br />

1996]. The third <strong>and</strong> last is called mechanical entrapment hypothesis. It favours the entrapment <strong>of</strong><br />

a particular biomolecular confirmation in a high viscosity trehalose glass [Gil et al. 1996; Lins et al.<br />

2004]. Among the various osmolytes in nature used to overcome deleterious environmental effects,<br />

trehalose seems to be exceptional compared to other sugars <strong>and</strong> polyols [Kaushik 2003]. It is more<br />

stable to wide ranges <strong>of</strong> pH <strong>and</strong> heat than most other sugars <strong>and</strong> its viscosity is relatively low with a


148 RESULTS AND DISCUSSION<br />

40% solution being approximately 5.7 centipoises [Richards et al. 2002]. Recent studies by Kaushik<br />

<strong>and</strong> Bhat [Kaushik 2003] stated that trehalose increases the transition temperature ( T )<br />

Δ <strong>of</strong> <strong><strong>protein</strong>s</strong><br />

to a maximal extent among the compatible osmolytes <strong>of</strong> the sugar <strong>and</strong> polyol series. They also<br />

observed a decrease in the heat capacity <strong>of</strong> denaturation for all there tested <strong><strong>protein</strong>s</strong>. This parameter<br />

is said to reflect upon the subtle changes in <strong>protein</strong>-solvent interactions [Liu 1996]. Trehalose also<br />

increases the surface tension <strong>of</strong> water to a larger extent than the other sugars <strong>and</strong> polyols. A linear<br />

correlation between the increased surface tension <strong>and</strong> the increase in the transition temperature<br />

( T )<br />

Δ <strong>of</strong> <strong>protein</strong> was found by Kaushik <strong>and</strong> Bhat [Kaushik 2003]. It was also reported, that the<br />

m<br />

presence <strong>of</strong> trehalose leads to larger changes <strong>of</strong> the thermodynamic properties <strong>of</strong> water than for<br />

other sugars, like the increase in partial molal heat capacity <strong>and</strong> volume. Trehalose is said to have a<br />

larger hydrated volume compared to other disaccharides [Sola-Penna 1998]. These increased<br />

parameters have been said to lead to stronger hydrogen bonding between hydroxyl groups <strong>of</strong> water<br />

molecules <strong>and</strong> trehalose. In such a solution additional energy would be needed for <strong>protein</strong><br />

denaturation to accommodate its increased surface area [Kaushik 2003]. Additionally, there is also a<br />

depletion <strong>of</strong> the surface tension increasing cosolvents at the <strong>protein</strong>-solvent interface leading to the<br />

preferential hydration <strong>of</strong> the <strong>protein</strong> [Timasheff 2002].<br />

For the levitation experiments four trehalose solutions with different solid content were<br />

prepared <strong>and</strong> their liquid properties were determined before starting the experiments. The data is<br />

shown in Table 5.6<br />

Table 5.6: Liquid properties <strong>and</strong> levitation size range for trehalose solutions with different solid content.<br />

Solid<br />

content<br />

Surface<br />

tension γ<br />

[ mg/ml ] [ ]<br />

Density<br />

ρ liquid<br />

3<br />

mN/m [ ]<br />

Maximum<br />

diameter<br />

Maximum<br />

volume<br />

Minimum<br />

diameter<br />

m<br />

Minimum<br />

volume<br />

g/cm [ mm ] [ µl ] [ µm ] [ µl ]<br />

50.0 72.80 ± 0.14 1.015 6.56 148.1 15µm 1.77⋅10 -6<br />

100.0 73.13 ± 0.08 1.032 6.49 142.8 15µm 1.77⋅10 -6<br />

200.0 74.02 ± 0.20 1.068 6.33 133.4 15µm 1.77⋅10 -6<br />

400.0 75.09 ± 0.20 1.140 6.02 114.5 15µm 1.77⋅10 -6


CHAPTER 5 RESULTS AND DISCUSSION 149<br />

Data analysis <strong>of</strong> trehalose experiments<br />

The analysis <strong>of</strong> the experimental data <strong>of</strong> the trehalose solution <strong>droplet</strong>s has to be divided at least<br />

into two different <strong>drying</strong> stages, similar to the maltodextrin analysis. The critical point between<br />

constant <strong>and</strong> falling rate can be determined using the course <strong>of</strong> the aspect ratio with evaporation<br />

time. The trehalose particles flattens during the formation <strong>of</strong> a solid crust due to the influence <strong>of</strong> the<br />

st<strong>and</strong>ing acoustic wave <strong>and</strong> the 5:1 relation between axial <strong>and</strong> radial levitation force. While the<br />

vertical diameter decreases under the influence <strong>of</strong> the axial levitation force after the critical point,<br />

the horizontal diameter stays nearly constant. This leads to an increase in aspect ratio <strong>and</strong> therefore<br />

provides a possibility to determine the duration <strong>of</strong> the first <strong>drying</strong> stage <strong>and</strong> the point <strong>of</strong> time <strong>of</strong> the<br />

transition to the second <strong>drying</strong> stage. In all experiments the minimal effective SPL for a stable<br />

levitation <strong>of</strong> the <strong>droplet</strong> / particle was chosen to keep the influence <strong>of</strong> the axial levitation force to a<br />

minimum. Figure 5.60 shows the raw data <strong>of</strong> the <strong>drying</strong> <strong>of</strong> a trehalose solution with 100 mg/ml<br />

solid content at 60°C <strong>and</strong> 5.0% relative humidity.<br />

Figure 5.60 shows that there is a change in aspect ratio <strong>and</strong> a rise <strong>of</strong> the formed particle after<br />

the critical point in the second <strong>drying</strong> stage. In contrast to maltodextrin, both diameters change<br />

substantially after the critical point, whereas the influence <strong>of</strong> the axial levitation force on the<br />

particle is much larger than the influence <strong>of</strong> the radial one resulting in an increasing aspect ratio <strong>and</strong><br />

a flattening <strong>of</strong> the particle during the falling rate. There is no sharp break in the curve <strong>of</strong> the squared<br />

radius with time at the critical point according to the considerations <strong>of</strong> chapter 2.4.2 <strong>and</strong> literature<br />

[Masters 1991]. Small size changes still occur in the second <strong>drying</strong> stage rounding the plotted<br />

curve. The deformation <strong>of</strong> the particle holds up till the end <strong>of</strong> the <strong>drying</strong> process is achieved. The<br />

gallery <strong>of</strong> photos taken from a levitated trehalose <strong>droplet</strong> with 200 mg/ml solid content with front<br />

illumination shows the deformation <strong>of</strong> the <strong>droplet</strong> / particle during the <strong>drying</strong> process very clearly.<br />

The impression <strong>of</strong> the top is not as for the maltodextrin <strong>droplet</strong>s.<br />

As with the maltodextrin solution, there are different ways to plot the evaporation rate<br />

calculated from the experimental data. To compensate for different initial volumes <strong>of</strong> the <strong>droplet</strong>s<br />

the data is best plot as evaporated liquid mass per time <strong>and</strong> surface area with evaporation time<br />

divided by the square <strong>of</strong> the initial radius. For trehalose solution <strong>droplet</strong>s this is show in Figure<br />

5.62. All trehalose experiments were performed at 60°C. The humidity <strong>of</strong> the <strong>drying</strong> air was varied<br />

from 5.0% to a maximum <strong>of</strong> 40%. The influence <strong>of</strong> the <strong>drying</strong> air humidity for the trehalose<br />

solutions with different solid content can be seen in Figure 5.63. An increasing humidity leads to an<br />

increase in evaporation time <strong>and</strong> a shift <strong>of</strong> the critical point to later points <strong>of</strong> time in the <strong>drying</strong><br />

process.


150 RESULTS AND DISCUSSION<br />

2 ) [-] Aspect ratio [-]<br />

2 / rS 0<br />

Diameter [µm]<br />

( r S<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Constant rate<br />

Falling rate<br />

Droplet / particle size<br />

Aspect ratio<br />

Vertical position<br />

0.0<br />

-200<br />

0 200 400 600 800 1000<br />

(a)<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

Evaporation time [s]<br />

Figure 5.60: Raw data <strong>of</strong> the levitation <strong>of</strong> a trehalose solution <strong>droplet</strong> with 200 mg/ml solid content without ventilation.<br />

The <strong>drying</strong> air had a temperature <strong>of</strong> 60°C <strong>and</strong> a relative humidity <strong>of</strong> 5.0%. The initial <strong>droplet</strong> size was 1.6 µl at an initial<br />

effective SPL <strong>of</strong> 165.1 dB. The dashed red line marks the position <strong>of</strong> the critical point.<br />

This is caused by a decrease <strong>of</strong> the water vapour pressure gradient between <strong>droplet</strong> surface <strong>and</strong><br />

environment. The evaporation rate in the first <strong>drying</strong> stage is almost constant for all humidities <strong>of</strong><br />

the <strong>drying</strong> air <strong>and</strong> all different solid concentrations, except for the trehalose solution with<br />

400 mg/ml solid content. The decline in the evaporation rate after the critical point is steeper for<br />

experiments at 5% <strong>and</strong> 10% relative humidity than for the experiments at 20% <strong>and</strong> 40%. The<br />

trehalose solutions with 400 mg/ml solid content do not show a clear first <strong>drying</strong> stage any more.<br />

The decrease <strong>of</strong> the evaporation rate occurred very early in the <strong>drying</strong> process due to high solid<br />

concentration <strong>and</strong> the almost immediate crust formation after the injection <strong>of</strong> the <strong>droplet</strong> into the<br />

acoustic levitation system.<br />

50<br />

0<br />

-50<br />

-100<br />

-150<br />

600<br />

1.2<br />

400<br />

Horizontal diameter<br />

200<br />

Vertical diameter<br />

Aspect ratio<br />

1.1<br />

0<br />

1.0<br />

0 200 400 600 800 1000<br />

Evaporation time [s]<br />

1.6<br />

1.5<br />

1.4<br />

1.3<br />

Distance to adjacent pressure node [µm]<br />

Aspect ratio [-]


CHAPTER 5 RESULTS AND DISCUSSION 151<br />

Figure 5.61: Evaporation <strong>of</strong> a trehalose solution <strong>droplet</strong> with 200 mg/ml solid content at 25°C <strong>and</strong> 26% relative<br />

humidity in still air. Pictures were taken with the CCD-camera every 30 seconds. The gallery starts with the 5 minutes<br />

after the injection <strong>of</strong> the <strong>droplet</strong>. The initial <strong>droplet</strong> volume was 2.1µl at an initial effective SPL <strong>of</strong> 164.28 dB.<br />

Evaporation rate [µg/s]<br />

10.0<br />

9.0<br />

Initial volume 1,663 µl<br />

8.0<br />

7.0<br />

6.0<br />

5.0<br />

4.0<br />

3.0<br />

2.0<br />

1.0<br />

0.0<br />

Initial volume 2,123 µl<br />

Initial volume 2,437 µl<br />

0 200 400 600 800 1000<br />

(a)<br />

Evaporation time [s]<br />

Figure 5.62: Different possibilities to plot the evaporation rate <strong>of</strong> trehalose solution <strong>droplet</strong>s with 100 mg/ml solid<br />

content at a <strong>drying</strong> temperature <strong>of</strong> 60°C <strong>and</strong> a relative humidity <strong>of</strong> 5%. Initial <strong>droplet</strong> volume <strong>and</strong> initial SPL differed<br />

between the three experiments. (a) Evaporation rate as evaporated mass per time unit with evaporation time. (b)<br />

Evaporation rate as evaporated mass per time unit <strong>and</strong> surface area with time divided by the squared initial <strong>droplet</strong><br />

radius.<br />

Evaporation rate [µg/s/mm 2 ]<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

Critical point<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Initial volume 1,667 µl<br />

Initial volume 2,123 µl<br />

Initial volume 2,437 µl<br />

0.2<br />

Constant Falling<br />

0.0<br />

rate rate<br />

0 200 400 600 800 1000 1200 1400<br />

(b)


152 RESULTS AND DISCUSSION<br />

The influence <strong>of</strong> the solid content on the course <strong>of</strong> the evaporation rate with time at different<br />

humidity <strong>of</strong> the <strong>drying</strong> air is shown in Figure 5.64. With increasing trehalose concentration the<br />

evaporation rate in the first <strong>drying</strong> stage is decreasing due to the amount <strong>of</strong> dissolved molecules <strong>and</strong><br />

the related water vapour pressure depression. The evaporation rate shows almost constant values in<br />

this phase, except for the solution with 400 mg/ml trehalose. The duration <strong>of</strong> the constant rate was<br />

decreasing with increasing solid content, leading to a shift <strong>of</strong> the critical point to earlier point <strong>of</strong><br />

time in the <strong>drying</strong> curve. The decline <strong>of</strong> the evaporation rate after the crust formation is steeper for<br />

lower concentrated trehalose solutions than for higher ones. Due to this effect, the duration <strong>of</strong> the<br />

second <strong>drying</strong> stage is slightly increasing for increasing solid content <strong>of</strong> the trehalose solutions.<br />

Table 5.7 gives an example comparison <strong>of</strong> the durations <strong>of</strong> the several <strong>drying</strong> stages <strong>and</strong> the<br />

position <strong>of</strong> the critical point for trehalose solutions with different solid content dried at 60°C <strong>and</strong><br />

5.0% relative humidity <strong>of</strong> the <strong>drying</strong> air. The experiments with 200 mg/ml trehalose content had an<br />

unexplainable course compared to all other experimental data.<br />

Evaporation rate [µg/s/mm 2 ]<br />

Evaporation rate [µg/s/mm 2 ]<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 500 1000 1500 2000<br />

(a)<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Rel. humidity 5%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

[Solid 50 mg/ml - 60°C]<br />

2 2<br />

( t / r ) [s/mm ]<br />

0<br />

0.0<br />

0 500 1000 1500 2000<br />

(c)<br />

Rel. humidity 5%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

[ Solid 200 mg/ml - 60°C<br />

2 2<br />

( t / r ) [s/mm ]<br />

0<br />

Evaporation rate [µg/s/mm 2 ]<br />

Evaporation rate [µg/s/mm 2 ]<br />

0.0<br />

0 500 1000 1500 2000<br />

Figure 5.63: Evaporation rate <strong>of</strong> trehalose solutions with different solid content at a <strong>drying</strong> temperature <strong>of</strong> 60°C <strong>and</strong><br />

different relative humidity. The initial <strong>droplet</strong> volume <strong>of</strong> the experiments was between 1.7 <strong>and</strong> 2.0 µl at an initial<br />

effective SPL between 163.75 <strong>and</strong> 165.1 dB. (a) 50 mg/ml trehalose solid content. (b) 100 mg/ml trehalose solid<br />

content. (c) 200 mg/ml trehalose solid content. (d) 400 mg/ml trehalose solid content.<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

(b)<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Rel. humidity 5%<br />

Rel. humidity 10%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

[ Solid 100 mg/ml - 60°C]<br />

2 2<br />

( t / r ) [s/mm ]<br />

0<br />

Rel. humidity 5%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

[ Solid 400 mg/ml - 60°C]<br />

0.0<br />

0 500 1000 1500 2000<br />

(d)<br />

2 2<br />

( t / r ) [s/mm ]<br />

0


CHAPTER 5 RESULTS AND DISCUSSION 153<br />

Table 5.7: Duration <strong>of</strong> constant <strong>and</strong> falling rate for 1.8 µl trehalose <strong>droplet</strong>s with different solid content dried at a<br />

temperature <strong>of</strong> 60°C <strong>and</strong> a relative humidity <strong>of</strong> 5%. The initial effective SPL was between 163.8 <strong>and</strong> 165.1 dB.<br />

Evaporation rate [µg/s/mm 2 ]<br />

Evaporation rate [µg/s/mm 2 ]<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Solid content<br />

50 mg/ml<br />

100 mg/ml<br />

200 mg/ml<br />

400 mg/ml<br />

[60°C - rel. humidity 40%]<br />

0.0<br />

0 500 1000 1500 2000<br />

(c)<br />

[ ]<br />

Constant rate period<br />

⎡ µg ⎤<br />

mg/ml ⎢ 2<br />

⎣s⋅<br />

mm ⎥<br />

⎦<br />

2 2<br />

( t / r ) [s/mm ]<br />

0<br />

Critical point tcrit<br />

Duration <strong>of</strong> detectable<br />

falling rate period<br />

[ s ]<br />

[] s<br />

50.0 1.20 – 1.10 274.5 234.0<br />

100.0 1.06 – 1.03 261.1 344.2<br />

200.0 1.06 – 1.03 - -<br />

400.0 0.99 – 0.98 138.9 668.6<br />

0.0<br />

0 500 1000 1500 2000<br />

(a)<br />

2 2<br />

( t / r ) [s/mm ]<br />

0<br />

50 mg/ml<br />

100 mg/ml<br />

200 mg/ml<br />

400 mg/ml<br />

[60°C - rel. humdity 5%]<br />

Evaporation rate [µg/s/mm 2 ]<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 500 1000 1500 2000<br />

(b)<br />

50 mg/ml<br />

100 mg/ml<br />

200 mg/ml<br />

400 mg/ml<br />

[60°C - rel. humidity 20%]<br />

2 2<br />

( t / r ) [s/mm ]<br />

0<br />

Figure 5.64: Evaporation rate <strong>of</strong> trehalose<br />

solutions with different solid content at a <strong>drying</strong><br />

temperature <strong>of</strong> 60°C <strong>and</strong> different relative<br />

humidity. The initial <strong>droplet</strong> volume <strong>of</strong> the<br />

experiments was between 1.7 <strong>and</strong> 2.0 µl at an<br />

initial effective SPL between 163.75 <strong>and</strong><br />

165.1 dB.<br />

(a) Evaporation rate <strong>of</strong> trehalose solutions at 60°C<br />

<strong>and</strong> 5% relative humidity.<br />

(b) Evaporation rate <strong>of</strong> trehalose solutions at 60°C<br />

<strong>and</strong> 20% relative humidity.<br />

(c) Evaporation rate <strong>of</strong> trehalose solutions at 60°C<br />

<strong>and</strong> 40% relative humidity.


154 RESULTS AND DISCUSSION<br />

A comparison <strong>of</strong> the evaporation coefficient β r confirms the statement concluded from the<br />

diagrams <strong>of</strong> the evaporation rate with time <strong>of</strong> Figure 5.63 <strong>and</strong> Figure 5.64. The evaporation<br />

coefficient is determined using the <strong>droplet</strong> size in a range between 1.8 µl <strong>and</strong> 1.5 µl in the first<br />

<strong>drying</strong> stage. The results are plotted in Figure 5.65 With increasing trehalose content <strong>and</strong> increasing<br />

humidity a decrease in the evaporation coefficient can be determined. This shows again very clearly<br />

the influence <strong>of</strong> the amount <strong>of</strong> dissolved molecules as well as the influence <strong>of</strong> the water vapour<br />

pressure gradient between surface <strong>and</strong> ambience.<br />

Evaporation coefficient β r [mm 2 /s]<br />

0.0022<br />

0.0020<br />

0.0018<br />

0.0016<br />

0.0014<br />

0.0012<br />

0.0010<br />

0.0008<br />

0.0006<br />

0.0004<br />

Rel. humidity 5%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

[Drying temp. 60°C]<br />

0 100 200 300 400<br />

Trehalose solid content [mg/ml]<br />

Figure 5.65: Evaporation coefficient βr in the first <strong>drying</strong> stage <strong>of</strong> trehalose solutions as a function <strong>of</strong> trehalose solid<br />

content at 60°C <strong>and</strong> different relative humidity in still air. The value are determined in a size range between 1.8 µl <strong>and</strong><br />

1.5 µl. The initial effective SPL was between 163.9 dB <strong>and</strong> 162.9 dB.<br />

Investigations <strong>of</strong> the resulting particle size <strong>and</strong> particle density are carried out similar to the<br />

maltodextrin solutions. The data is shown in Figure 5.66. An increasing solid content leads to an<br />

increase in the particle size due to the higher solid concentration in the initial <strong>droplet</strong>. It can also be<br />

seen that the particle radius in relation to the initial <strong>droplet</strong> radius shows the larges values for the<br />

faster evaporation process at 60°C <strong>and</strong> 5% relative humidity <strong>and</strong> the smallest values at 60°C <strong>and</strong><br />

40% humidity. The density confirms this conclusion. Trehalose particles formed <strong>of</strong> solutions with<br />

the same solid content but at different humidity <strong>of</strong> the <strong>drying</strong> air have an increasing resulting<br />

particle density with increase in <strong>drying</strong> time, which is equal to an increasing humidity <strong>of</strong> the <strong>drying</strong><br />

air. The data shows that the density <strong>of</strong> the 400 mg/ml trehalose solution particles is lower compared<br />

to the less concentrated samples under the same <strong>drying</strong> conditions. Figure 5.67 shows SEM pictures<br />

<strong>of</strong> trehalose particles produced in the levitation system <strong>and</strong> in the Buchi Mini Spray-Dryer.<br />

Trehalose particles from the levitation system show a folded structure on their top (Figure 5.67 b).


CHAPTER 5 RESULTS AND DISCUSSION 155<br />

This effect can also be seen during the <strong>drying</strong> experiment itself using front light illumination. The<br />

particles stay more spherical during the second <strong>drying</strong> stage than the maltodextrin ones. Structural<br />

similarities to the spray-dried particles cannot be found at this magnification.<br />

Particle radius relativ to initial <strong>droplet</strong> radius [%]<br />

80<br />

70<br />

60<br />

50<br />

40<br />

Rel. humidity 5%<br />

Rel. humidity 20%<br />

30<br />

Rel. humidity 40%<br />

0 100 200 300 400<br />

(a)<br />

Trehalose solution solid content [mg/ml]<br />

0.9<br />

0 100 200 300 400<br />

Figure 5.66: Comparison <strong>of</strong> the resulting particle size <strong>and</strong> density <strong>of</strong> trehalose solutions with different solid content,<br />

dried 60°C <strong>and</strong> different humidity without ventilation. The initial <strong>droplet</strong> volume <strong>of</strong> all experiments was between 1.7 µl<br />

<strong>and</strong> 2.0 µl at an initial effective SPL <strong>of</strong> 163.8 dB to 165.1 dB. (a) Particle radius compared to the initial <strong>droplet</strong> radius<br />

with different trehalose solution solid content at different humidity. (b) Resulting particle density with trehalose<br />

solution solid content at different humidity.<br />

To overcome the influence <strong>of</strong> the accumulation <strong>of</strong> solvent vapour in the toroidal vortices <strong>of</strong><br />

the outer acoustic streaming, a trehalose solution with 100 mg/ml solid content is tested under the<br />

influence <strong>of</strong> different ventilation velocities. With increasing flow rate the <strong>droplet</strong> becomes more <strong>and</strong><br />

more unstable. This is most notable in the second <strong>drying</strong> stage after the formation <strong>of</strong> the solid crust,<br />

when the mass <strong>of</strong> the particle is decreasing due to the proceeding evaporation <strong>of</strong> solvent. The<br />

determination <strong>of</strong> the evaporation rate in this stage is very difficult <strong>and</strong> sometimes, if the oscillations<br />

<strong>and</strong> the movement <strong>of</strong> the particle is too large, impossible. In contrast to the experiments in still air<br />

neglecting the effects <strong>of</strong> the inner acoustic streaming on the Sherwood number, the experiments in<br />

moving air cannot be plotted to compensate for the different initial <strong>droplet</strong> size. The <strong>droplet</strong> <strong>and</strong><br />

resulting particle size has a direct influence on the Reynolds number at the <strong>droplet</strong> / particle <strong>and</strong><br />

therefore a direct influence on the Sherwood number calculated by the Ranz-Marshall equation.<br />

Because different initial <strong>droplet</strong> sizes result in different particle diameters under the same ambient<br />

conditions, the Sherwood number <strong>and</strong> the mass transfer would be different in the second <strong>drying</strong><br />

stage. To compare experiments, it is necessary to inject solution <strong>droplet</strong>s with the same initial<br />

volume into a st<strong>and</strong>ing acoustic wave with a constant effective SPL. Figure 5.68 a shows the course<br />

Particle density [g/cm 3 ]<br />

1.3<br />

1.2<br />

1.1<br />

1.0<br />

(b)<br />

Rel. humidity 5%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

Trehalose solution solid content [mg/ml]


156 RESULTS AND DISCUSSION<br />

<strong>of</strong> the Sherwood number at a ventilation velocity <strong>of</strong> 1.5 m/s ( Re = 300)<br />

orifice calculated with the<br />

Ranz-Marshall equation <strong>and</strong> the Sherwood number at a ventilation velocity <strong>of</strong> 1.0 m/s<br />

( Re = 200)<br />

orifice calculated with the levitation model. The influence <strong>of</strong> the ventilation on the<br />

evaporation rate <strong>of</strong> a 1.7 µl <strong>droplet</strong> with 100 mg/ml trehalose content compared to experiments in<br />

still air at 60°C <strong>and</strong> a relative humidity <strong>of</strong> 5% is plotted in Figure 5.68 b. With increasing flow rate<br />

the evaporation rate in the first <strong>drying</strong> stage also increases. The formation <strong>of</strong> a solid crust appears<br />

earlier at higher ventilation velocity than in still air, resulting in a decrease in evaporation time <strong>of</strong><br />

the constant rate. The decline <strong>of</strong> the evaporation rate after the critical point is steeper with<br />

increasing ventilation. Increasing the humidity <strong>of</strong> the ventilation air stream leads to effects on the<br />

evaporation rate equal to Figure 5.64 for the experiments in still air. At a velocity <strong>of</strong> 1.0 m/s <strong>and</strong><br />

increasing humidity the evaporation rate decreases leading to a increase in the evaporation time <strong>and</strong><br />

a shift <strong>of</strong> the critical point to later points <strong>of</strong> time in the evaporation process (Figure 5.68 c).<br />

(a)<br />

(b)<br />

Figure 5.67: SEM pictures <strong>of</strong> trehalose particles from the levitation system <strong>and</strong> the Buchi B-191 Mini Spray Dryer.<br />

(a-b) Trehalose particles from a solution with 200 mg/ml solid content dried at 60°C <strong>and</strong> 5% relative humidity. Shown<br />

are pictures with a magnification <strong>of</strong> 75.5-times (a), 1000-times (b). (c-d) Spray-dried trehalose particles from solutions<br />

with 100 mg/ml solid content at an inlet temperature <strong>of</strong> 100°C <strong>and</strong> an outlet temperature <strong>of</strong> 60°C. The pictures were<br />

taken with 1000-times (c) <strong>and</strong> 2000-times (d) magnification.<br />

(c)<br />

(d)


CHAPTER 5 RESULTS AND DISCUSSION 157<br />

Sherwood number [-]<br />

Evaporation rate [µg/s/mm 2 ]<br />

10.0<br />

9.0<br />

8.0<br />

7.0<br />

6.0<br />

5.0<br />

4.0<br />

3.0<br />

Levitation model at 166 dB<br />

2.0<br />

Ranz-Marshall correlation at 1.5 m/s<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0<br />

(a)<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

0 200 400 600 800 1000<br />

(b)<br />

60°C<br />

Droplet volume [µl]<br />

Ventilation 0.0 m/s<br />

Ventilation 1.0 m/s<br />

Ventilation 1.5 m/s<br />

[60°C - rel. humidity 5%]<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Figure 5.68: (a) Sherwood number as a function <strong>of</strong><br />

<strong>droplet</strong> volume at 60°C <strong>and</strong> 0.1% relative humidity.<br />

(b) Evaporation rate <strong>of</strong> trehalose solution <strong>droplet</strong>s<br />

with 100 mg/ml solid content at 60°C <strong>and</strong> 5%<br />

relative humidity. The initial <strong>droplet</strong> volume <strong>of</strong> the<br />

experiments was 1.7 µl at an initial effective SPL <strong>of</strong><br />

166.0 dB. The <strong>drying</strong> air velocity was varied from<br />

still air to 1.5 m/s. (c) Evaporation rate <strong>of</strong> trehalose<br />

solution <strong>droplet</strong>s with 100 mg/ml solid content at<br />

60°C <strong>and</strong> different relative humidity. All<br />

experiments were carried out with a ventilation <strong>of</strong><br />

1.0 m/s. The initial <strong>droplet</strong> volume <strong>of</strong> the<br />

experiments was 1.7 µl at an initial effective SPL<br />

between 164.8 dB <strong>and</strong> 165.1 dB.<br />

0.0<br />

0 250 500 750 1000 1250 1500<br />

The effects <strong>of</strong> the air flow rate on the particle formation <strong>and</strong> the particle morphology was examined.<br />

With increasing ventilation the particle radius relative to the initial <strong>droplet</strong> radius increases, whereas<br />

the density decreases. A <strong>drying</strong> <strong>of</strong> the particles at a relative humidity <strong>of</strong> 20% <strong>and</strong> 40% <strong>of</strong> the <strong>drying</strong><br />

air leads to a decrease <strong>of</strong> the particle size <strong>and</strong> an increase in density compared to the results from<br />

the experimental data at 5% relative humidity. It seems that a decreasing evaporation time,<br />

regardless <strong>of</strong> which is the causing factor, <strong>drying</strong> temperature, humidity <strong>of</strong> the <strong>drying</strong> air or<br />

magnitude <strong>of</strong> the ventilation velocity, is leading to larger <strong>and</strong> less dense particles. The results can be<br />

seen in Figure 5.69. A composition <strong>of</strong> SEM pictures <strong>of</strong> the resulting particles formed under a<br />

ventilation velocity <strong>of</strong> 1.0 m/s <strong>and</strong> a humidity <strong>of</strong> 5% <strong>and</strong> 20% can be seen in Figure 5.70. It seems<br />

that the fast <strong>drying</strong> process at 60°C, 1.0 m/s <strong>and</strong> 5% humidity causes a porous structure visible at<br />

the particle surface by the small pores Figure 5.70 b. In contrast, the experiments with longer<br />

evaporation time at 60°C, 1.0 m/s <strong>and</strong> 20% humidity show a smooth surface <strong>and</strong> the folded<br />

structure (Figure 5.70 c / d) equal to the particles produces in still air (Figure 5.67 b).<br />

Evaporation rate [µg/s/mm 2 ]<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

(c)<br />

Rel. humidity 5%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

[60°C - ventilation 1.0 m/s]<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0


158 RESULTS AND DISCUSSION<br />

Particle radius relative to initial <strong>droplet</strong> radius [%]<br />

54.0<br />

52.0<br />

50.0<br />

48.0<br />

46.0<br />

44.0<br />

42.0<br />

(a)<br />

[Drying temp. 60°C]<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6<br />

Ventilation velocity [m/s]<br />

Rel. humidity 5%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

Figure 5.69: Resulting particle size <strong>and</strong> density <strong>of</strong> trehalose solutions with 100 mg/ml solid content, dried 60°C <strong>and</strong><br />

different humidity at a ventilation velocity between 0.0 <strong>and</strong> 1.5 m/s. The initial <strong>droplet</strong> volume <strong>of</strong> all experiments was<br />

between 1.7 µl <strong>and</strong> 2.0 µl at an initial effective SPL <strong>of</strong> 164.8 dB to 165.1 dB. (a) Particle size versus ventilation velocity<br />

at different humidity. (b) Final particle density with ventilation velocity at different humidity.<br />

(a) (b)<br />

(c) (d)<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6<br />

Figure 5.70: SEM pictures <strong>of</strong> trehalose particles with 100 mg/ml solid content from the levitation system. (a-b)<br />

Trehalose particles formed at 60°C <strong>and</strong> 5.0% relative humidity with a ventilation <strong>of</strong> 1.0 m/s. Shown are pictures with a<br />

magnification <strong>of</strong> 100-times (a) <strong>and</strong> 1000-times (b). (c-d) Trehalose particles formed at 60°C <strong>and</strong> 20.0% relative<br />

humidity with a ventilation <strong>of</strong> 1.0 m/s. Shown are pictures with a magnification <strong>of</strong> 100-times (c) <strong>and</strong> 1000-times (d)<br />

Particle density [g/cm 3 ]<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

(b)<br />

Rel. humidity 5%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

[Drying temp. 60°C]<br />

Ventilation velocity [m/s]


CHAPTER 5 RESULTS AND DISCUSSION 159<br />

5.3.3 Evaporation <strong>of</strong> mannitol solution <strong>droplet</strong>s<br />

Properties <strong>of</strong> mannitol<br />

Mannitol is a hexavalent sugar alcohol with the molecular formula C6H14O6 <strong>and</strong> a molecular weight<br />

<strong>of</strong> 182.17 g/mol [Sigma 2005]. It melts at a temperature <strong>of</strong> 166-168°C [Windholz et al. 1983] <strong>and</strong><br />

has a maximum solubility <strong>of</strong> 182 mg/ml in water at 25°C [McEvoy 1992]. D-mannitol is a common<br />

excipients in spray- <strong>and</strong> freeze-<strong>drying</strong> with a good chemical stability [Yu et al. 1998]. It does not<br />

undergo hydrolysis at high or low pH. Even mannitol has the strong tendency to crystallize [Yu et<br />

al. 1998] it also exists in fully <strong>and</strong> partially amorphous state in certain <strong>formulations</strong> [Bosquillon et<br />

al. 2004; Costantino et al. 1998]. A crystalline state <strong>of</strong> excipients is detrimental for <strong>protein</strong> stability<br />

because <strong>of</strong> phase separation <strong>and</strong> loss <strong>of</strong> excipients-<strong>protein</strong> interaction [Carpenter et al. 1997].<br />

Constantino et al. [Costantino et al. 1998] stabilized recombinant humanized Anti-IgE using<br />

mannitol <strong>and</strong> sodium phosphate. The presence <strong>of</strong> sodium phosphate was successful in inhibiting the<br />

mannitol crystallization upon spray <strong>drying</strong> resulting in a significant lowering <strong>of</strong> solid state<br />

aggregation. During long-term storage some crystallization in the spray-dried formulation was<br />

observed [Costantino et al. 1998]. Spray-dried <strong>formulations</strong> <strong>of</strong> albumin together with mannitol <strong>and</strong><br />

dipalmitoylphosphatidylcholin also showed a fully amorphous structure <strong>of</strong> the produced particles<br />

[Bosquillon et al. 2004].<br />

A 100 mg/ml was prepared for single <strong>droplet</strong> <strong>drying</strong> experiments in the levitations system.<br />

Experiments were carried out at 60°C <strong>and</strong> different humidity as well as at different ventilation<br />

velocity. The liquid properties were determined before starting the levitation procedure. The data is<br />

shown in Table 5.8.<br />

Table 5.8: Liquid properties <strong>and</strong> levitation size range for mannitol solutions with different solid content.<br />

Solid<br />

content<br />

Surface<br />

tension γ<br />

[ mg/ml ] [ ]<br />

Density<br />

ρ liquid<br />

3<br />

mN/m [ ]<br />

Maximum<br />

diameter<br />

Maximum<br />

volume<br />

Minimum<br />

diameter<br />

Minimum<br />

volume<br />

g/cm [ mm ] [ µl ] [ µm ] [ µl ]<br />

100.0 73.49 ± 0.08 1.030 6.53 145.8 15 1.77⋅10 -6


160 RESULTS AND DISCUSSION<br />

Data analysis <strong>of</strong> mannitol solution experiments<br />

The analysis <strong>of</strong> the experimental data <strong>of</strong> the mannitol solution <strong>droplet</strong>s has to be divided into two<br />

different <strong>drying</strong> stages. In contrast to maltodextrin <strong>and</strong> trehalose, the mannitol <strong>droplet</strong>s behave<br />

different at the transition between constant <strong>and</strong> falling rate. An immediate particle formation occurs<br />

at the critical point resulting in a sharp break <strong>of</strong> the curve plotting the size versus evaporation time.<br />

The horizontal <strong>and</strong> vertical diameters do not change any more in the second <strong>drying</strong> stage leading to<br />

a constant aspect ratio <strong>and</strong> a constant size <strong>of</strong> the particle. The rise <strong>of</strong> the particle in the falling rate<br />

can be determined without taking account <strong>of</strong> any impressions caused by the st<strong>and</strong>ing acoustic wave<br />

after the initial particle formation. Even though an influence <strong>of</strong> the SPL on changes <strong>of</strong> the particle<br />

shape within the second <strong>drying</strong> stage cannot be seen, the minimal value for a stable levitation was<br />

chosen. In some experiments the immediate particle formation at the critical point leads to strong<br />

oscillations <strong>of</strong> the new formed particle resulting in the impossibility to calculate the evaporation rate<br />

in the second <strong>drying</strong> stage or even a drop-out <strong>of</strong> the particle <strong>of</strong> its pressure node. In contrast,<br />

levitation experiments with stable <strong>droplet</strong> / particle position at the same <strong>drying</strong> conditions can also<br />

be found. It seems that the shape <strong>of</strong> the resulting particle has an influence on the stability <strong>of</strong> the<br />

formed particle at the critical point <strong>and</strong> within the second <strong>drying</strong> stage. Figure 5.71 shows the raw<br />

data <strong>of</strong> the <strong>drying</strong> <strong>of</strong> a mannitol solution with 100 mg/ml solid content at 60°C <strong>and</strong> 10% relative<br />

humidity with the stable behaviour <strong>of</strong> the particle at the transition from first to second <strong>drying</strong> stage.<br />

The evaporation rate in the falling rate is again calculated using the rise <strong>of</strong> the particle in the<br />

constant stationary ultrasonic field. The end <strong>of</strong> solvent evaporation is characterized by a constant<br />

particle size <strong>and</strong> a constant particle position. A gallery <strong>of</strong> front illumination pictures <strong>of</strong> the <strong>drying</strong><br />

process for mannitol at room temperature can be seen in Figure 5.72. In contrast to maltodextrin <strong>and</strong><br />

trehalose that form glassy, almost transparent particles, the particles <strong>of</strong> mannitol are white <strong>and</strong> nontransparent.<br />

A size reduction or a flattening <strong>of</strong> the particle due to influences <strong>of</strong> the ultrasonic field<br />

cannot be seen in Figure 5.72. The final particles show a fragile structure. Low mechanic pressure<br />

causes the break <strong>of</strong> the particle <strong>and</strong> the formation <strong>of</strong> a small amount <strong>of</strong> white powder.<br />

A ventilation <strong>of</strong> 1.0 m/s at 60°C <strong>and</strong> 5% humidity <strong>of</strong> the <strong>drying</strong> air leads to an increase in<br />

the evaporation rate from 1.4 µg/s/mm 2 to almost 2.4 µg/s/mm 2 . Therefore, a decrease in the<br />

duration <strong>of</strong> the first <strong>drying</strong> stage is observed. As with still air, the evaporation rate in the first <strong>drying</strong><br />

stage gives higher values for mannitol than for trehalose. Experiments at a humidity <strong>of</strong> 20% <strong>and</strong><br />

40% <strong>and</strong> a ventilation velocity <strong>of</strong> 1.0 m/s do not show a second <strong>drying</strong> stage at all. The position <strong>of</strong><br />

the particle stays constant after the critical point <strong>and</strong> no tendency <strong>of</strong> a rise <strong>of</strong> the centre <strong>of</strong> mass


CHAPTER 5 RESULTS AND DISCUSSION 161<br />

towards its adjacent upper pressure node can be observed. The raw data <strong>of</strong> mannitol experiments at<br />

20% <strong>and</strong> 40% humidity are presented in Figure 5.71.<br />

2 ) [-] Aspect ratio [-]<br />

2 / rS 0<br />

Droplet / particle diameter [µm]<br />

( r S<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Droplet / particle size<br />

Aspect ratio<br />

Vertical position<br />

0.0<br />

-100<br />

0 200 400 600 800 1000<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

Constant rate<br />

Falling rate<br />

Evaporation time [s]<br />

Horizontal diameter<br />

Vertical diameter<br />

Aspect ratio<br />

Figure 5.71: Raw data <strong>of</strong> the levitation <strong>of</strong> a mannitol solution <strong>droplet</strong> with 100 mg/ml solid content without ventilation.<br />

The <strong>drying</strong> air had a temperature <strong>of</strong> 60°C <strong>and</strong> a relative humidity <strong>of</strong> 10%. The initial <strong>droplet</strong> size was 1.7 µl at an initial<br />

effective SPL <strong>of</strong> 164.8 dB. The dashed red line marks the position <strong>of</strong> the critical point.<br />

20<br />

0<br />

-20<br />

-40<br />

-60<br />

-80<br />

400<br />

1.0<br />

0 200 400 600 800 1000<br />

Evaporation time [s]<br />

1.5<br />

1.4<br />

1.3<br />

1.2<br />

1.1<br />

Particle distance to adjacent node [µm]<br />

Aspect ratio [-]


162 RESULTS AND DISCUSSION<br />

Figure 5.72: Evaporation <strong>of</strong> a mannitol solution <strong>droplet</strong> with 100 mg/ml solid content at 25°C <strong>and</strong> 28% relative<br />

humidity without ventilation. Pictures were taken with the CCD-camera every 20 seconds. The gallery starts with the<br />

3.5 minutes after the injection <strong>of</strong> the <strong>droplet</strong>. The initial <strong>droplet</strong> volume was 1.7 µl at an initial effective SPL <strong>of</strong><br />

164.8 dB.<br />

The evaporation rate <strong>of</strong> mannitol <strong>droplet</strong>s with a solid content <strong>of</strong> 100 mg/ml at a temperature <strong>of</strong><br />

60°C <strong>and</strong> different relative humidity is determined using the method compensating the different<br />

initial <strong>droplet</strong> volumes. The plot is presented in Figure 5.73. The evaporation rate in the first <strong>drying</strong><br />

stage decreased with increasing humidity <strong>of</strong> the <strong>drying</strong> air. This leads to an increase in duration <strong>of</strong><br />

the first <strong>drying</strong> stage equal to maltodextrin <strong>and</strong> trehalose. In contrast to the other two sugars,<br />

mannitol shows a sharp break in the curve for the evaporation rate <strong>of</strong> all tested ambient conditions,<br />

even at high humidity. Also interesting are the larger values <strong>of</strong> the evaporation rate for mannitol in<br />

the first <strong>drying</strong> stage than for maltodextrin <strong>and</strong> trehalose. Even though a 100 mg/ml solution<br />

contains more dissolved mannitol molecules than a trehalose solution with the same solid<br />

concentration, the evaporation rate for mannitol is higher than for trehalose. This is observed at all<br />

tested humidity conditions <strong>of</strong> the <strong>drying</strong> air.


CHAPTER 5 RESULTS AND DISCUSSION 163<br />

Evaporation rate [µg/s/mm 2 ]<br />

Figure 5.73: Evaporation rate <strong>of</strong> mannitol solutions with 100 mg/ml solid content at a <strong>drying</strong> temperature <strong>of</strong> 60°C <strong>and</strong><br />

different humidity. The initial <strong>droplet</strong> volume <strong>of</strong> the experiments was 1.7 µl at an initial effective SPL 164.8 dB.<br />

Evaporation rate [µg/s/mm 2 ]<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0 500 1000 1500<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

Critical point<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Rel. humidity 5%<br />

Rel. humidity 10%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

Ventilation 0.0 m/s<br />

Ventilation 1.0 m/s<br />

[60°C - rel. humidity 5%]<br />

0.0<br />

0 200 400 600 800<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Figure 5.74: Comparison <strong>of</strong> the evaporation rate <strong>of</strong> mannitol solutions with 100 mg/ml solid content at a <strong>drying</strong><br />

temperature <strong>of</strong> 60°C <strong>and</strong> 5% humidity in still air <strong>and</strong> with ventilation <strong>of</strong> 1.0 m/s. The initial <strong>droplet</strong> volume <strong>of</strong> the<br />

experiments was 1.7 µl at an initial effective SPL 164.8 dB.<br />

In contrast to maltodextrin <strong>and</strong> trehalose, the ventilation <strong>of</strong> the mannitol particle with humid air<br />

seems to prevent further <strong>drying</strong>. A comparison <strong>of</strong> the amount <strong>of</strong> solvent evaporated in the first<br />

<strong>drying</strong> stage shows an increase from 63.4 ± 2.7% (n=3) at 5% humidity to 81.9 ± 1.3% (n=3) at<br />

20% humidity to 92.9 ± 0.2% (n=3) at 40% humidity. Calculations <strong>of</strong> evaporated solvent for pure<br />

water <strong>droplet</strong>s at the same conditions do not show an additional evaporation <strong>of</strong> condensed water<br />

vapour. Furthermore, the final particles also show a moist behaviour when trying to take SEM<br />

pictures. According to the evaporated solvent in the first <strong>drying</strong> stage <strong>and</strong> the non-existence <strong>of</strong> a<br />

second <strong>drying</strong> stage in the experimental data, the particles produced at 40% humidity have


164 RESULTS AND DISCUSSION<br />

theoretically a lower residual humidity than the particles dried at 20% humidity. More probable is<br />

that the rise <strong>of</strong> the particle is so minimal <strong>and</strong> at the same time the oscillations <strong>of</strong> the particle due to<br />

the ventilation are that large, that the second <strong>drying</strong> stage cannot be detected. The particles are<br />

affected very easily by the air stream in the second <strong>drying</strong> stage due to an unbalanced particle shape<br />

causing turbulences in the area <strong>of</strong> the pressure node. The strong oscillations starting at the critical<br />

point can be seen in Figure 5.75 marked by the ellipse. The higher the <strong>drying</strong> air velocity, the more<br />

abstract the shape becomes. Additionally, the densities <strong>of</strong> the particles show the lowest values for<br />

all examined sugars at <strong>drying</strong> air conditions 1.0 m/s <strong>and</strong> 60°C. Equal to maltodextrin <strong>and</strong> trehalose,<br />

an increase in particle size <strong>and</strong> a decrease in particle density with decreasing humidity <strong>and</strong><br />

increasing ventilation can be observed.<br />

2 ) [-] Aspect ratio [-]<br />

2 / rS 0<br />

( r S<br />

Instable particle position after the critical point<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

Droplet / particle size<br />

Aspect ratio<br />

Vertical position<br />

0.2<br />

[60°C - humidity 20% - 1.0 m/s]<br />

-200<br />

0 150 300 450 600 750<br />

(a)<br />

Evaporation time [s]<br />

Figure 5.75: Raw data mannitol solution <strong>droplet</strong>s with 100 mg/ml solid content. The <strong>drying</strong> air had a temperature <strong>of</strong><br />

60°C <strong>and</strong> a ventilation velocity <strong>of</strong> 1.0 m/s. The dashed red line marks the position <strong>of</strong> the critical point.<br />

(a) Relative humidity 20%. The initial <strong>droplet</strong> size was 1.4 µl at an initial effective SPL <strong>of</strong> 162.5 dB. (b) Relative<br />

humidity 40%. The initial <strong>droplet</strong> size was 1.7 µl at an initial effective SPL <strong>of</strong> 166.6 dB.<br />

Elversson <strong>and</strong> Millqvist-Fureby [Elversson 2005] determined the effective particle density <strong>of</strong> spray-<br />

dried mannitol ( = 200 ° C; T = 90°<br />

C; Liquid feed = 5ml/min)<br />

T outlet<br />

50<br />

0<br />

-50<br />

-100<br />

-150<br />

Particle distance to adjacent node [µm]<br />

2 ) [-] Aspect ratio [-]<br />

2 / rS 0<br />

( r S<br />

Particle distance to adjacent node [µm]<br />

Instable particle position after the critical point<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0 150<br />

50<br />

0<br />

-50<br />

Droplet / particle size<br />

Aspect ratio<br />

-100<br />

Vertical position<br />

-150<br />

[60°C - humidity 40% - 1.0 m/s]<br />

-200<br />

300 450 600 750<br />

Evaporation time [s]<br />

inlet calculated from <strong>droplet</strong>-to-<br />

particle measurements as 0.24 g/cm 3 for a 5% (w/w) solution <strong>and</strong> 0.30 g/cm 3 for a 15% (w/w)<br />

solution. The calculated particle density <strong>of</strong> the dried mannitol solution <strong>droplet</strong>s with 100 mg/ml at<br />

60°C, 5% humidity <strong>and</strong> <strong>drying</strong> air velocity <strong>of</strong> 1.0 m/s in this work is 0.32 ± 0.01 g/cm 3 (n = 3).<br />

They determined the apparent particle density using a gas pycnometer to 1.50 g/cm 3 for the<br />

5% (w/w) solution <strong>and</strong> 1.49 g/cm 3 for the 15% (w/w). Due to blow holes in many <strong>of</strong> their mannitol<br />

particles the measured density corresponded well to the true density <strong>of</strong> mannitol (1.51 g/cm 3 ). The<br />

BET surface area <strong>of</strong> the spray-dried mannitol powder was determined as 4.63 m 2 /g at a particle size<br />

<strong>of</strong> 4.00 µm (5% (w/w)) solution to 4.04 m 2 /g at a particles size <strong>of</strong> 5.01 µm (15% (w/w))<br />

[Elversson 2005].<br />

(b)


CHAPTER 5 RESULTS AND DISCUSSION 165<br />

Particle radius relative to initial <strong>droplet</strong> radius [%]<br />

70<br />

65<br />

60<br />

55<br />

50<br />

Ventilation 0.0 m/s<br />

Ventilation 1.0 m/s<br />

0 5 10 15 20 25 30 35 40 45<br />

Relative humidity <strong>of</strong> <strong>drying</strong> air at 60°C [%]<br />

Figure 5.76: Comparison <strong>of</strong> the resulting particle size <strong>and</strong> density <strong>of</strong> mannitol solutions with 100 mg/ml solid content,<br />

dried 60°C with different humidity <strong>and</strong> <strong>drying</strong> air velocity. The initial <strong>droplet</strong> volume <strong>of</strong> all experiments was between<br />

1.5 µl <strong>and</strong> 2.0 µl at an initial effective SPL <strong>of</strong> 162.5 dB to 166.6 dB. (a) Particle radius compared to the initial <strong>droplet</strong><br />

radius with ventilation velocity at different humidity. (b) Final particle density with ventilation velocity at different<br />

humidity.<br />

The SEM pictures <strong>of</strong> the final particles show a mostly spicular shape <strong>of</strong> the particle surface<br />

depending on the temperature, humidity <strong>and</strong> velocity <strong>of</strong> the <strong>drying</strong> air. The strongest distinctive<br />

spicular morphology is observed for particles dried in still air at room temperature (Figure 5.77 a/b).<br />

The particles are only partially spherical with protuberances at their surface.<br />

(a) (b)<br />

0.2<br />

0 5 10 15 20 25 30 35 40 45<br />

Figure 5.77: SEM pictures <strong>of</strong> mannitol particles with 100 mg/ml solid content from the levitation system dried at 25°C<br />

<strong>and</strong> 20% relative humidity in still air. Shown are pictures with a magnification <strong>of</strong> 100-times (a) <strong>and</strong> 350-times (b).<br />

Particle density [g/cm 3 ]<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

Ventilation 0.0 m/s<br />

Ventilation 1.0 m/s<br />

Relative humidity <strong>of</strong> <strong>drying</strong> air at 60°C [%]


166 RESULTS AND DISCUSSION<br />

In literature [Elversson 2005] solid state characterization <strong>of</strong> spray-dried mannitol solutions<br />

( = 200 ° C; T = 90°<br />

C; Liquid feed = 5ml/min)<br />

T outlet<br />

inlet showed an 89 to 100% crystalline mannitol<br />

fraction. This leads to the conclusion that sharp critical point in the evaporation curve <strong>of</strong> the<br />

levitated <strong>droplet</strong>s <strong>and</strong> the spicular surface morphology is caused by the crystallization <strong>of</strong> mannitol<br />

during the <strong>drying</strong> process.<br />

The levitation experiments with faster solvent evaporation at 60°C <strong>and</strong> 5% relative humidity<br />

in still air (Figure 5.78) lead to particles with a smoother surface compared to the particles produced<br />

at room temperature (Figure 5.77). Particles from different experiments all under the same <strong>drying</strong><br />

air conditions show different particle morphology (Figure 5.78 a <strong>and</strong> c). No two particles with the<br />

same shape could be found. The investigated particles all differ in their protuberances but all show a<br />

hollow interior with blow holes in their surface (Figure 5.78 c, Figure 5.79). A substantial influence<br />

<strong>of</strong> the <strong>drying</strong> air humidity on the structure cannot be seen. It seems that the prolongation <strong>of</strong> the<br />

<strong>drying</strong> time due in higher humidity environment increases the spicular particle morphology (Figure<br />

5.80 <strong>and</strong> Figure 5.81). A constant particle shape at higher humidity cannot be found, either. A<br />

comparison <strong>of</strong> the surface morphology with spray-dried mannitol is difficult due to the insufficient<br />

magnification <strong>of</strong> the small particles <strong>of</strong> the spray-dried product (Figure 5.81). The SEM pictures <strong>of</strong><br />

spray-dried mannitol in largest magnification (Figure 5.81 b) show particles with a smooth surface.<br />

A spicular surface structure cannot be determined. Equal to the levitated particles, hollow particle<br />

interior <strong>and</strong> blow holes in the surface can be seen. Many fragments <strong>and</strong> broken particles can be<br />

seen. Exceptional particles can be found when <strong>drying</strong> under <strong>drying</strong> air velocity conditions (Figure<br />

5.82) at 60°C <strong>and</strong> 0.1% <strong>drying</strong> air humidity. The final particles have the structure <strong>of</strong> a bowl (Figure<br />

5.82 a). The bottom side <strong>of</strong> the particle facing the air stream has an almost smooth whereas<br />

protuberances are formed on the top side (Figure 5.82 d). The particles are flatter than under still air<br />

conditions. An increase in <strong>drying</strong> air humidity <strong>and</strong>, therefore, an increase in <strong>drying</strong> time also seems<br />

to lead to the development <strong>of</strong> more spicular surface structure than at 0.1% humidity (Figure 5.83).<br />

Still the morphology <strong>of</strong> particles dried under the same <strong>drying</strong> air conditions varies in a way that a<br />

clear influence <strong>of</strong> the ambient conditions cannot be seen.


CHAPTER 5 RESULTS AND DISCUSSION 167<br />

(a) (b)<br />

blow holes<br />

(c) (d)<br />

Figure 5.78: SEM pictures <strong>of</strong> mannitol particles with 100 mg/ml solid content from the levitation system dried at 60°C<br />

<strong>and</strong> 5% humidity in still air. Shown are pictures <strong>of</strong> two different particles with a magnification <strong>of</strong> 100-times (a), 300times<br />

(b), 100-times (c) <strong>and</strong> 350-times (d).<br />

(a) (b)<br />

hollow<br />

interior<br />

Figure 5.79: SEM pictures <strong>of</strong> mannitol particles with 100 mg/ml solid content from the levitation system dried at 60°C<br />

<strong>and</strong> 10% relative humidity in still air. Shown are pictures with a magnification <strong>of</strong> 100-times (a) <strong>and</strong> 300-times (b).


168 RESULTS AND DISCUSSION<br />

(a) (b)<br />

(c) (d)<br />

Figure 5.80: SEM pictures <strong>of</strong> mannitol particles with 100 mg/ml solid content from the levitation system dried at 60°C<br />

<strong>and</strong> 20% relative humidity in still air. Shown are pictures with a magnification <strong>of</strong> 100-times (a), 600-times (b), 500times<br />

(c) <strong>and</strong> 1000-times (d).<br />

(a) (b)<br />

Figure 5.81: SEM pictures <strong>of</strong> spray-dried mannitol particles with 100 mg/ml solid content in a Buchi-Mini-Spray-<br />

Dryer B 190 with Tinlet = 100°C, Toutlet = 60°C <strong>and</strong> a liquid feed <strong>of</strong> 3 ml/min. Shown are pictures with a magnification <strong>of</strong><br />

1000-times (a) <strong>and</strong> 2500-times (b).


CHAPTER 5 RESULTS AND DISCUSSION 169<br />

(a) (b)<br />

(c) (d)<br />

Figure 5.82: SEM pictures <strong>of</strong> mannitol particles with 100 mg/ml solid content from the levitation system dried at 60°C<br />

<strong>and</strong> 0.% relative humidity with a <strong>drying</strong> air velocity <strong>of</strong> 1.0 m/s. Shown are pictures with a magnification <strong>of</strong> 80-times (a),<br />

600-times (b), 150-times (c) <strong>and</strong> 500-times (d).<br />

(a) (b)<br />

Figure 5.83: SEM pictures <strong>of</strong> mannitol particles with 100 mg/ml solid content from the levitation system dried at 60°C<br />

<strong>and</strong> 20% relative humidity in still air. Shown are pictures with a magnification <strong>of</strong> 60-times (a) <strong>and</strong> 1000-times (b).


170 RESULTS AND DISCUSSION<br />

5.3.4 Evaporation <strong>of</strong> bSA solution <strong>droplet</strong>s<br />

Properties <strong>of</strong> bSA<br />

The single-chain polypeptide bSA is <strong>of</strong>ten used as a model <strong>protein</strong> in, for example, processing<br />

examinations <strong>and</strong> physical-state investigations in spray-<strong>drying</strong> [Adler et al. 2000; Bosquillon et al.<br />

2004; Lucas et al. 1998; Witschi 1999], microencapsulation [Carrasquillo et al. 2001; Elversson<br />

2005 b] <strong>and</strong> freeze-<strong>drying</strong> [Izutsu et al. 2004; Tattini et al. 2005]. Its structure <strong>and</strong> characteristics<br />

are described in detail in Chapter 4.1.1. Lucas et al. [1998] co-processed bSA with maltodextrin to<br />

produce model <strong>protein</strong> particles by spray-<strong>drying</strong> to prepare aerosol <strong>formulations</strong> <strong>and</strong> examined the<br />

influence <strong>of</strong> fine particle multiplets as performance modifiers for <strong>protein</strong> deposition from dry<br />

powder inhalers. The surface composition <strong>of</strong> spray-dried particles <strong>of</strong> bSA containing trehalose <strong>and</strong><br />

different surfactants was examined using electron spectroscopy for chemical analysis by Adler et al.<br />

[2000]. They found that an increasing amount <strong>of</strong> polysorbate 80 in the spray-dried solution reduces<br />

<strong>protein</strong> adsorption at the water/air-interface. Competitive <strong>protein</strong> absorption between bSA <strong>and</strong> β-<br />

lactoglobulin during spray-<strong>drying</strong> was studied by L<strong>and</strong>strom et al. [2000]. The powder surface was<br />

examined using fluorescence quenching <strong>of</strong> pyrene labelled <strong><strong>protein</strong>s</strong>. The surface analysis after<br />

spray-<strong>drying</strong> <strong>of</strong> solutions with lower <strong>protein</strong> concentrations seemed independent <strong>of</strong> the other <strong>protein</strong><br />

present. The aerodynamic behaviour, surface composition <strong>and</strong> physical state <strong>of</strong> powder aerosols<br />

containing albumin, dipalmitoylphosphatidylcholine <strong>and</strong> a <strong>protein</strong> stabilizer like lactose, trehalose<br />

or mannitol <strong>and</strong> prepared by spray-<strong>drying</strong> was investigated by Bosquillon et. al [2004]. Witschi<br />

[1999] prepared starch, chitosan or Carbopol © microspheres containing bSA by spray-<strong>drying</strong>. Their<br />

particles in a size range between 2-4 µm were characterized by scanning electron mircroscopy <strong>and</strong><br />

laser diffraction. The <strong>protein</strong> release pr<strong>of</strong>iles were determined using an open-membrane system <strong>and</strong><br />

polarized Calu-3 cell sheets were taken to evaluate relative bioadhesion. Recent studies in the area<br />

<strong>of</strong> <strong>protein</strong> encapsulation by spray-<strong>drying</strong> was done by Elversson <strong>and</strong> Millqvist-Fureby [Elversson<br />

2005 b]. The aim was to examine to what extent an aqueous two-phase system could improve<br />

<strong>protein</strong> stability during spray-<strong>drying</strong> by preventing interactions with the air-liquid interface. The<br />

spray-dried solution contained bSA, polyvinylalcohol (PVA), dextran <strong>and</strong> in some experiments<br />

trehalose. Their results showed that although bSA was well encapsulated by PVA, a partial loss <strong>of</strong><br />

native structure was observed. Other research in <strong>protein</strong> microencapsulation performed by<br />

Carrasquillo et al. [2001] used excipients-stabilized spray-freeze dried BSA in a non-aqueous oil-inoil<br />

encapsulation methodology for the development <strong>of</strong> sustained release microspheres. As wall<br />

material poly(D,L-lactide-co-glycolide) (PLG) was used. FTIR analysis showed that bSA released


CHAPTER 5 RESULTS AND DISCUSSION 171<br />

from the microspheres had similar monomer content <strong>and</strong> the same secondary structure as<br />

unencapsuled BSA. In freeze <strong>drying</strong> bSA was used as model <strong>protein</strong> to examine the protective<br />

effects <strong>of</strong> saccharides with different molecular weight against lyophilisation induced structural<br />

perturbation [Izutsu et al. 2004]. A decreasing structure stabilizing effect <strong>of</strong> the saccharides with<br />

increasing number <strong>of</strong> the ssaccharide units was found. Recently, PEGylated bSA was chosen by<br />

Tattini et al. [2005] to investigate the influence <strong>of</strong> lyophilisation on structure <strong>and</strong> phase changes.<br />

A 100 mg/ml pure bSA solution <strong>and</strong> a 100 mg/ml bSA-Trehalose (1:1) mixture were<br />

prepared for single <strong>droplet</strong> <strong>drying</strong> experiments in the levitations system. Experiments were carried<br />

out at 60°C <strong>and</strong> different humidity as well as at different <strong>drying</strong> air velocity. The liquid properties<br />

were determined before starting the levitation procedure. The data are shown in Table 5.9.<br />

Table 5.9: Liquid properties <strong>and</strong> levitation size range for bSA <strong>and</strong> bSA-Trehalose solutions with 100 mg/ml solid<br />

content.<br />

Solid<br />

content<br />

100 mg/ml<br />

Surface<br />

tension<br />

[ ]<br />

Density<br />

ρ liquid<br />

3<br />

mN/m [ ]<br />

Maximum<br />

diameter<br />

Maximum<br />

volume<br />

Minimum<br />

diameter<br />

Minimum<br />

volume<br />

g/cm [ mm ] [ µl ] [ µm ] [ µl ]<br />

bSA 58.1 1.023 5.20 73.5 15 1.77⋅10 -6<br />

bSA-<br />

Trehalose<br />

58.5 1.027 5.21 74.2 15 1.77⋅10 -6<br />

Data analysis <strong>of</strong> bSA solution experiments<br />

The data analysis <strong>of</strong> pure bSA <strong>droplet</strong>s / particles is more difficult than the analysis <strong>of</strong> the sugar<br />

data. As seen with maltodextrin, trehalose <strong>and</strong> mannitol, the <strong>drying</strong> process <strong>of</strong> a pure bSA solution<br />

can be divided into constant <strong>and</strong> falling rates. The evaporation rate in the first <strong>drying</strong> stage is<br />

calculated using the decrease in <strong>droplet</strong> volume [Frohn 2000; Kastner et al. 2001], whereas the<br />

evaporation in the second <strong>drying</strong> stage after the crust formation is calculated using the rise <strong>of</strong> the<br />

particle within the constant stationary ultrasonic field. In contrast to the saccharides, the<br />

<strong>droplet</strong> / particle behaviour shortly after the critical point differs from the experiments reported so<br />

far. In almost all experiments, the new formed particle starts to oscillate after the formation <strong>of</strong> a<br />

solid crust at the <strong>droplet</strong> surface. The amplitude <strong>of</strong> the oscillations increases such that the st<strong>and</strong>ing<br />

acoustic wave cannot stabilize the particle any more <strong>and</strong> it drops out <strong>of</strong> its levitation position.<br />

Additionally, breaking <strong>of</strong> the particle within the falling rate can sometimes be seen in the photo<br />

stream taken by the imaging s<strong>of</strong>tware. The results <strong>of</strong> pure bSA presented are therefore calculated


172 RESULTS AND DISCUSSION<br />

from the few experiments with a stable <strong>droplet</strong> / particle levitation. Figure 5.84 presents the raw<br />

data <strong>of</strong> a bSA solution with 100 mg/ml solid content dried at 60°C <strong>and</strong> 5% relative humidity in still<br />

air. It can be seen that change in aspect ratio after the formation <strong>of</strong> the solid crust can be used to<br />

determine the critical point <strong>and</strong> the transition between first <strong>and</strong> second <strong>drying</strong> stages.<br />

2 ) Aspect ratio<br />

2 / rS 0<br />

( r S<br />

Diameter [µm]<br />

Constant rate Falling rate<br />

1.6<br />

100<br />

1.4<br />

0<br />

Distance<br />

1.2<br />

to<br />

1.0<br />

-100<br />

Particle size<br />

0.8<br />

Aspect ratio<br />

adjacent<br />

Vertical position -200<br />

0.6<br />

pressure<br />

-300<br />

0.4<br />

node<br />

0.2<br />

-400<br />

0 100 200 300 400 500 600<br />

[µm]<br />

1200<br />

1100<br />

1000<br />

900<br />

800<br />

700<br />

600<br />

500<br />

Evaporation time [s]<br />

1500 Horizontal diameter<br />

1400<br />

Vertical diameter<br />

1300<br />

Aspect ratio<br />

Change in both diameters<br />

after critical point<br />

400<br />

1.0<br />

0 100 200 300 400 500 600<br />

Evaporation time [s]<br />

Figure 5.84: Raw data <strong>of</strong> the levitation <strong>of</strong> a pure bSA solution <strong>droplet</strong> with 100 mg/ml solid content in still air. The<br />

<strong>drying</strong> air had a temperature <strong>of</strong> 60°C <strong>and</strong> a relative humidity <strong>of</strong> 5%. The initial <strong>droplet</strong> size was 1.1 µl at an initial<br />

effective SPL <strong>of</strong> 166.25 dB. The dashed red line marks the position <strong>of</strong> the critical point.<br />

1.6<br />

1.5<br />

1.4<br />

1.3<br />

1.2<br />

1.1<br />

Aspect ratio


CHAPTER 5 RESULTS AND DISCUSSION 173<br />

There is a substantial change in vertical as well as in horizontal diameter after the critical point.<br />

Since the effective SPL depends on the particle size, <strong>and</strong> calculation <strong>of</strong> the evaporation rate in the<br />

second <strong>drying</strong> stage requires a constant SPL, the determination <strong>of</strong> the evaporation coefficient is not<br />

possible. Calculation <strong>of</strong> the SPL at the critical point <strong>and</strong> at the end <strong>of</strong> the size change <strong>of</strong> the particle<br />

in the falling rate using the method <strong>of</strong> Yarin et al. [1998] shows a difference <strong>of</strong> 0.025 from the<br />

maximum <strong>of</strong> 0.10 dB. This small difference has no substantial influence on the calculated<br />

evaporation rate data <strong>and</strong> can therefore be neglected. Figure 5.85 shows front illumination pictures<br />

<strong>of</strong> a levitation experiment <strong>of</strong> pure bSA solution at 60°C <strong>and</strong> 5% humidity in still air. As with<br />

maltodextrin, the particle flattens within the second <strong>drying</strong> stage. The line <strong>of</strong> breakage <strong>of</strong> the<br />

particle after the critical point can also be seen <strong>and</strong> is marked by the white arrows in Figure 5.86.<br />

Figure 5.85: Evaporation <strong>of</strong> a pure bSA solution <strong>droplet</strong> with 100 mg/ml solid content at 60°C <strong>and</strong> 10% relative<br />

humidity without <strong>drying</strong> air stream. Pictures were taken with the CCD-camera every 10 seconds. The initial <strong>droplet</strong><br />

volume was 1.9 µl at an initial effective SPL <strong>of</strong> 164.2 dB.


174 RESULTS AND DISCUSSION<br />

before breakage after breakage<br />

(a) (b) line <strong>of</strong> breakage<br />

Figure 5.86: Front illumination pictures <strong>of</strong> pure bSA particles with 100 mg/ml solid content dried at 60°C <strong>and</strong> 10%<br />

relative humidity in still air. Shown are pictures directly before (a) <strong>and</strong> after (b) breakage <strong>of</strong> the particle.<br />

hole in the<br />

particle top<br />

(a) (b)<br />

Figure 5.87: SEM pictures <strong>of</strong> pure bSA particles with 100 mg/ml solid content dried at 60°C <strong>and</strong> 10% relative<br />

humidity in still air. Shown are pictures with a magnification <strong>of</strong> 100-times (a) <strong>and</strong> 250-times (b).<br />

Figure 5.87 shows SEM pictures <strong>of</strong> the broken particle from the front illumination experiment <strong>of</strong><br />

Figure 5.85. It can be seen that the <strong>drying</strong> <strong>of</strong> the pure bSA solutions forms a hollow particle with a<br />

massive solid crust. The crust does not show any capillary structure. The half <strong>of</strong> a large hole (“blow<br />

hole”) can be seen in the top side <strong>of</strong> the bSA particle, whereas the bottom side is intact. Two<br />

different ways <strong>of</strong> particle formation <strong>of</strong> the pure bSA solution can be observed. The first way is the<br />

(Figure 5.87) already presented formation <strong>of</strong> holes in the particle surface during the second <strong>drying</strong><br />

stage with a occasional breakage <strong>of</strong> the particle into two parts. The second way is the formation <strong>of</strong><br />

an elliptical to almost spherical particle with intact surface <strong>and</strong> also a massive crust, that always<br />

breaks in the second <strong>drying</strong> stage. Particles obtained from the second way <strong>of</strong> particle formation are<br />

presented in Figure 5.88.


CHAPTER 5 RESULTS AND DISCUSSION 175<br />

(a) (b)<br />

Figure 5.88: SEM pictures <strong>of</strong> pure BSA particles with 100 mg/ml solid content dried at 60°C <strong>and</strong> 10% relative<br />

humidity in still air. Shown are pictures with a magnification <strong>of</strong> 100-times (a) <strong>and</strong> 500-times (b).<br />

Both the breakage <strong>of</strong> the particle as well as the formation <strong>of</strong> a large hole in the particle crust leading<br />

to the formation <strong>of</strong> an “interior surface” have an influence on the results <strong>and</strong> the calculation <strong>of</strong> the<br />

evaporation rate. In the first case, breakage <strong>of</strong> the particle can result in complete disintegration <strong>and</strong><br />

drop-out <strong>of</strong> the remaining halves out <strong>of</strong> the st<strong>and</strong>ing acoustic wave. A determination <strong>of</strong> the<br />

evaporation rate within the falling rate is then impossible. If the particle breaks, but the two parts<br />

stick together, water vapour from the inside can easily reach the exterior without having to penetrate<br />

the crust <strong>of</strong> solid bSA. In the second case, the formation <strong>of</strong> a hole in the crust <strong>and</strong> the development<br />

<strong>of</strong> an interior surface enlarges the surface area <strong>of</strong> the particle, which results in higher evaporation<br />

rates [Frohn 2000]. Neither <strong>of</strong> the two cases can be unequivocally recorded by the imaging-s<strong>of</strong>tware<br />

during the progression <strong>of</strong> the levitation experiment. The back light illumination produces only a<br />

shadow image <strong>of</strong> the <strong>droplet</strong> / particle. Use <strong>of</strong> front light illumination is difficult, due to the poor<br />

contrast between <strong>droplet</strong> <strong>and</strong> background. Therefore, the pictures must be taken with back light<br />

illumination <strong>and</strong> the evaporation rate calculated as with maltodextrin, trehalose or mannitol. The<br />

evaporation rate <strong>of</strong> pure bSA solutions with 100 mg/ml solid content at 60°C <strong>and</strong> different relative<br />

humidity in still air can be seen in Figure 5.89. Increasing <strong>drying</strong> air humidity results in an increase<br />

in the duration <strong>of</strong> the first <strong>drying</strong> stage <strong>and</strong> a shift <strong>of</strong> the critical point (Figure 5.89). The<br />

evaporation rate in the second <strong>drying</strong> stage does not decrease with evaporation time continuously.<br />

An increase in the evaporation rate within the falling rate period occurs. The starting evaporation<br />

rate increase is concurrent to the rupture or break <strong>of</strong> the particle. The easier solvent evaporation<br />

through holes <strong>and</strong> the enlarged surface area leads to the increase in the curve in Figure 5.89. The<br />

maximum in the evaporation rate in the falling rate period occurs later with increasing <strong>drying</strong> air<br />

humidity.


176 RESULTS AND DISCUSSION<br />

Evaporation rate [µg/s/mm 2 ]<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

Critical point<br />

0.4<br />

Rel. humidity 5%<br />

0.2<br />

Rel. humidity 10%<br />

Rel. humidity 20%<br />

0.0<br />

[Drying temperature 60°C]<br />

0 200 400 600 800 1000<br />

2 2<br />

t / r [s/mm ]<br />

S 0<br />

Evaporation rate<br />

increase in<br />

falling rate period<br />

Figure 5.89: Evaporation rate <strong>of</strong> pure BSA solutions with 100 mg/ml solid content at a <strong>drying</strong> temperature <strong>of</strong> 60°C <strong>and</strong><br />

different relative humidity. The initial <strong>droplet</strong> volume <strong>of</strong> the experiments was between 1.5 <strong>and</strong> 2.0 µl at an initial<br />

effective SPL between 163.75 <strong>and</strong> 165.1 dB.<br />

Analysing the SEM pictures <strong>of</strong> the broken pure bSA particles, the thickness <strong>of</strong> the solid crust is<br />

determined to 31.75% <strong>of</strong> the particle radius. Therefore, a spherical particle with a radius <strong>of</strong> 500 µm<br />

has an apparent volume <strong>of</strong> 0.524 µl <strong>and</strong> a void volume <strong>of</strong> 0.166 µl. The outer surface area <strong>of</strong> the<br />

particle <strong>of</strong> 3.14 mm 2 is enlarged by an interior surface <strong>of</strong> 1.46 mm 2 to an overall surface area <strong>of</strong><br />

4.60 mm 2 during the second <strong>drying</strong> stage. Blow holes in the surface are neglected in these<br />

considerations. No statements about the temporal development <strong>of</strong> the interior surface can be made<br />

analysing the back light illuminated photo stream <strong>of</strong> the bSA <strong>droplet</strong> <strong>drying</strong> experiment. A<br />

correction <strong>of</strong> the surface in the falling rate period is not possible. Pictures <strong>of</strong> the front light<br />

illuminated single <strong>droplet</strong> <strong>drying</strong> experiments <strong>of</strong> the mixture <strong>of</strong> 50 mg/ml bSA <strong>and</strong> 50 mg/ml<br />

trehalose can be seen in Figure 5.90. The molar ratio between bSA <strong>and</strong> trehalose is approximately<br />

1:194. The <strong>drying</strong> behaviour <strong>of</strong> the mixture is different to the pure bSA <strong>droplet</strong>s. There is no change<br />

in the horizontal particle diameter after the critical point but small changes in the vertical diameter<br />

(Figure 5.91). Therefore, a small flattening <strong>of</strong> the particle in the second <strong>drying</strong> stage can be seen<br />

from the aspect ratio <strong>and</strong> the live pictures. In contrast to the pure bSA solution, an increase in the<br />

evaporation rate within the second <strong>drying</strong> stage is not determined for the mixture with trehalose.<br />

With increasing evaporation time in the falling rate, the values <strong>of</strong> the evaporated mass per time <strong>and</strong><br />

surface area get continuously smaller until the end <strong>of</strong> the <strong>drying</strong> process. The plot <strong>of</strong> the<br />

evaporation rate versus time at 60°C <strong>and</strong> different relative humidity in still air is shown in Figure


CHAPTER 5 RESULTS AND DISCUSSION 177<br />

5.92. As with all other tested substances, an increase in <strong>drying</strong> air humidity leads to an increase in<br />

duration <strong>of</strong> the first <strong>drying</strong> stage.<br />

Figure 5.90: Evaporation <strong>of</strong> a mixture bSA-trehalose (1:1) solution <strong>droplet</strong> with 100 mg/ml solid content at Room<br />

temperature <strong>and</strong> 27% relative humidity without <strong>drying</strong> air stream. Pictures were taken with the CCD-camera every 30<br />

seconds. The initial <strong>droplet</strong> volume was 2.0 µl at an initial effective SPL <strong>of</strong> 164.4 dB. The gallery starts 15 minutes<br />

after <strong>droplet</strong> injection.<br />

The <strong>droplet</strong>s / particles <strong>of</strong> the bSA-trehalose (1:1) mixture show more stable levitation behaviour<br />

during the evaporation experiments than pure bSA. Oscillations associated with the particle<br />

formation at the critical point cannot be determined. Considering the course <strong>of</strong> the evaporation rate<br />

versus time, Figure 5.91 emphasizes this statement. The SEM pictures <strong>of</strong> the final bSA-trehalose<br />

particles show slightly flattened particle shape with a folded top side (Figure 5.93 a, b). A breakage<br />

<strong>of</strong> the particles into two or more parts cannot be seen. Instead <strong>of</strong> that, many small cracks appeared<br />

on the bottom side <strong>of</strong> the surface (Figure 5.93 c, e <strong>and</strong> f). It is obvious that those cracks do not have<br />

the same influence on the evaporation rate in the second <strong>drying</strong> stage than the breakage <strong>of</strong> the pure<br />

bSA particles has (Figure 5.92). The particle surface itself is rougher for the mixture particles than<br />

for any other <strong>of</strong> the tested substances (Figure 5.93 d). Destruction <strong>of</strong> the particles shows a massive<br />

particle interior. Blow holes cannot be found. Therefore, no additional interior surface has to be<br />

taken into account when analysing the evaporation rate versus time. An influence <strong>of</strong> the <strong>drying</strong> air<br />

humidity on the particle morphology cannot be seen.


178 RESULTS AND DISCUSSION<br />

(r / r 0 ) 2 Aspect ratio<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

Droplet / particle size<br />

Aspect ratio<br />

Vertical position<br />

0.2<br />

-250<br />

0 500 1000 1500<br />

Evaporation time [s]<br />

Figure 5.91: Raw data <strong>of</strong> the levitation <strong>of</strong> bSA-Trehalose (1:1) mixture solution <strong>droplet</strong> with 100 mg/ml solid content<br />

without ventilation. The <strong>drying</strong> air had a temperature <strong>of</strong> 60°C <strong>and</strong> a relative humidity <strong>of</strong> 5%. The initial <strong>droplet</strong> size was<br />

2.8 µl at an initial effective SPL <strong>of</strong> 166.4 dB. The dashed red line marks the position <strong>of</strong> the critical point.<br />

Evaporation rate [µg/s/mm 2 ]<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Diameter [µm]<br />

2200<br />

2000<br />

1800<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

Constant rate Falling rate<br />

Horizontal diameter<br />

Vertical diameter<br />

Aspect ratio<br />

constant horizontal diameter<br />

50<br />

0<br />

-50<br />

-100<br />

-150<br />

-200<br />

600<br />

0 500 1000<br />

1.0<br />

1500<br />

Evaporation time [s]<br />

Rel. humidity 5%<br />

Rel. humidity 10%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

[Drying air temp. 60°C]<br />

0.0<br />

0 500 1000 1500<br />

2 2<br />

t / r [s/mm ]<br />

0<br />

1.6<br />

1.5<br />

1.4<br />

1.3<br />

1.2<br />

1.1<br />

Distance to adjecent pressure node [mm]<br />

Aspect ratio<br />

Figure 5.92: Evaporation rate <strong>of</strong> bSA-<br />

Trehalose (1:1) mixture solutions with<br />

100 mg/ml solid content at a <strong>drying</strong><br />

temperature <strong>of</strong> 60°C <strong>and</strong> different relative<br />

humidity. The initial <strong>droplet</strong> volume <strong>of</strong><br />

the experiments was between 1.5 <strong>and</strong><br />

2.0 µl at an initial effective SPL between<br />

163.8 <strong>and</strong> 165.9 dB.


CHAPTER 5 RESULTS AND DISCUSSION 179<br />

(a) (b)<br />

(c) (d)<br />

(e) (f)<br />

Figure 5.93: SEM pictures <strong>of</strong> BSA-Trehalose (1:1) particles with 100 mg/ml solid content dried at 60°C <strong>and</strong> 5%<br />

relative humidity (a-b) or 20% humidity (c-f) in still air. Shown are pictures with a magnification <strong>of</strong> 100-times (a,c),<br />

250-times (e), 1000-times (b, d) <strong>and</strong> 2000-times (f).


180 RESULTS AND DISCUSSION<br />

The comparison <strong>of</strong> the evaporation rate <strong>of</strong> single <strong>droplet</strong>s <strong>of</strong> pure bSA, the mixture <strong>of</strong> bSAtrehalose<br />

(1:1) <strong>and</strong> pure trehalose at 60°C <strong>and</strong> 10% relative humidity in still air is shown in Figure<br />

5.94. The evaporation rate <strong>of</strong> bSA has larger values in the first <strong>drying</strong> stage than the mixture <strong>and</strong> the<br />

pure trehalose due to the lowest amount <strong>of</strong> dissolved molecules. A 100 mg/ml bSA solution has a<br />

molar concentration <strong>of</strong> 1.51 x 10 -3 mol/l, whereas the pure trehalose has a molar concentration <strong>of</strong><br />

0.29 mol/l. The mixture <strong>of</strong> bSA <strong>and</strong> trehalose has a molar concentration <strong>of</strong> approximately<br />

0.15 mol/l. The first <strong>drying</strong> stage increases with addition <strong>of</strong> trehalose to the solution under the same<br />

<strong>drying</strong> conditions <strong>and</strong> with the same initial <strong>droplet</strong> size. The critical point shifts to later points <strong>of</strong><br />

time in the <strong>drying</strong> process. The decline in the evaporation rate in the falling rate period is shallower<br />

for the pure trehalose than for the mixture. In contrary, the values <strong>of</strong> the evaporation rate after the<br />

decline are larger in the pure trehalose experiments than in the pure bSA or mixture runs. The<br />

percentage <strong>of</strong> solvent evaporated in the first <strong>drying</strong> stage at 60°C in still air is listed in Table 5.10.<br />

65.1% <strong>of</strong> solvent are evaporated in the constant rate <strong>of</strong> pure bSA solutions, whereas 87.8% <strong>of</strong><br />

solvent are evaporated form the bSA-trehalose (1:1) mixture at 5% relative humidity <strong>of</strong> the <strong>drying</strong><br />

air. The values for the bSA-trehalose (1:1) mixture <strong>and</strong> the pure trehalose solution are almost<br />

identical. No influence <strong>of</strong> the <strong>drying</strong> air humidity in the range between 5% <strong>and</strong> 20% on the amount<br />

<strong>of</strong> evaporated solvent can be seen either for these to system. The calculated values at 40% are larger<br />

than at lower humidity caused by the difficulty to determine the exact critical point due to the slow<br />

<strong>drying</strong> process. The st<strong>and</strong>ard deviation at 40% humidity is larger than at all other <strong>drying</strong> air<br />

humidities. The pure bSA solution shows a decrease in the amount <strong>of</strong> evaporated solvent in the<br />

constant rate with increasing <strong>drying</strong> air humidity.<br />

Table 5.10: Percentage <strong>of</strong> evaporated solvent in the first <strong>drying</strong> stage at 60°C <strong>and</strong> different relative humidity in still air.<br />

Relative humidity Percentage <strong>of</strong> evaporated solvent in the first <strong>drying</strong> stage (n=3)<br />

<strong>of</strong> <strong>drying</strong> air bSA bSA-Trehalose (1:1) Trehalose<br />

mg/ml<br />

mg/ml<br />

100 mg/ml<br />

[ % ]<br />

[ 100 ]<br />

[ 100 ]<br />

[ ]<br />

5 65.1 ± 1.9 87.8 ± 0.7 85.6 ± 1.8<br />

10 64.3 ± 1.1 87.6 ± 0.7 87.8 ± 1.7<br />

20 61.3 ± 1.8 87.9 ± 0.5 86.5 ± 1.9<br />

40 - 92.2 ± 2.5 91.6 ± 3.1


CHAPTER 5 RESULTS AND DISCUSSION 181<br />

Evaporation rate [mg/s/mm 2 ]<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Figure 5.94: Comparison <strong>of</strong> the evaporation rate <strong>of</strong> pure bSA, bSA-trehalose (1:1) mixture <strong>and</strong> pure trehalose solutions<br />

with 100 mg/ml solid content at a <strong>drying</strong> temperature <strong>of</strong> 60°C <strong>and</strong> 10% relative humidity in still air. The initial <strong>droplet</strong><br />

volume <strong>of</strong> the experiments was 1.5 µl an initial effective SPL between 164.1 <strong>and</strong> 164.5 dB. Due to the different density<br />

<strong>of</strong> the solutions, the initial effective SPL differed at constant initial <strong>droplet</strong> volume. This effect is neglected in the<br />

considerations <strong>of</strong> the experiment.<br />

For comparison with the spray-dried product, bSA <strong>and</strong> bSA-trehalose (1:1) solutions with a solid<br />

content <strong>of</strong> 100 mg/ml are spray-dried using the Buchi Mini Spray Dryer B-191. The SEM pictures<br />

<strong>of</strong> the final powders are presented in Figure 5.95 <strong>and</strong> compared to the spray-dried pure trehalose<br />

particles <strong>of</strong> Figure 5.67. bSA shows particles with a smooth surface <strong>and</strong> a donut-like shape with<br />

dimples <strong>and</strong> holes. In contrast to this result, Maa et al. [1997] found a wrinkled or raisin-like surface<br />

morphology for spray-dried pure bSA, produced under similar process conditions<br />

[Tinlet = 90°<br />

C, Toutlet<br />

= 53°<br />

C, Liquid feed =<br />

Critical point <strong>and</strong> end <strong>of</strong> first <strong>drying</strong> stage<br />

0.0<br />

0 250 500 750 1000 1250 1500 1750 2000<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

5ml/min]<br />

bSA pure<br />

bSA-Trehalose (1:1)<br />

Trehalose pure<br />

larger evaporation rate<br />

with increasing<br />

trehalose content<br />

. The addition <strong>of</strong> trehalose to the bSA solution<br />

changed the appearance <strong>of</strong> the spray-dried particles. As found for rhGH [Maa et al. 1997] <strong>and</strong><br />

granulocyte stimulation factor [Niven et al. 1993], the particles <strong>of</strong> the formulated <strong>protein</strong> show a<br />

wrinkled surface shape. The spray-dried pure trehalose particles are mostly spherical with a smooth<br />

surface (Figure 5.67). The change in surface morphology from the smooth <strong>and</strong> spherical trehalose to<br />

the raisin-like structure <strong>of</strong> the bSA-trehalose mixture is an effect <strong>of</strong> the addition on an high<br />

molecular weight substance, which alters the balance <strong>of</strong> surface-to-viscous forces controlling the<br />

<strong>droplet</strong> shape during the <strong>drying</strong> process [Alex<strong>and</strong>er 1978]. ESCA (Electron Spectroscopy for<br />

Chemical Analysis) <strong>of</strong> spray-dried pure bSA, pure trehalose <strong>and</strong> bSA-trehalose (5:95) mixture<br />

particles performed by Adler et al. [2000] gave an approximately 9 times higher N-level for the<br />

mixture than expected from a homogeneous distribution <strong>of</strong> bSA within the trehalose particles. They


182 RESULTS AND DISCUSSION<br />

attributed this increase to the adsorption <strong>of</strong> bSA to the liquid-air interface <strong>of</strong> the atomized <strong>droplet</strong>s<br />

before they dry.<br />

(a) (b)<br />

Figure 5.95: (a) SEM picture <strong>of</strong> spray-dried bSA particles with 100 mg/ml solid content in a Buchi-Mini-Spray-Dryer<br />

B 190 with Tinlet = 100°C, Toutlet = 60°C <strong>and</strong> a liquid feed <strong>of</strong> 3 ml/min. Shown is a picture with a magnification <strong>of</strong><br />

3000-times (b) SEM picture <strong>of</strong> spray-dried bSA-trehalose (1:1) mixture particles with 100 mg/ml solid content in a<br />

Buchi-Mini-Spray-Dryer B 190 with Tinlet = 100°C, Toutlet = 60°C <strong>and</strong> a liquid feed <strong>of</strong> 3 ml/min. Shown is a pictures<br />

with a magnification <strong>of</strong> 2000-times.<br />

The influence <strong>of</strong> trehalose on the size <strong>and</strong> density <strong>of</strong> the final particles resulting from the levitated<br />

solutions at 60°C <strong>and</strong> different relative humidity in still air can be seen in Figure 5.96. As with all<br />

other tested substances before, the particle size <strong>of</strong> bSA <strong>and</strong> the bSA-trehalose (1:1) mixture<br />

decreases with increasing <strong>drying</strong> air humidity <strong>and</strong> increasing <strong>drying</strong> time. The final particle density,<br />

however, increases with increasing <strong>drying</strong> air humidity. The final particle density <strong>of</strong> the bSA<br />

particles is smaller at the same ambient conditions than the density <strong>of</strong> the bSA-trehalose (1:1)<br />

mixture. At 60°C <strong>and</strong> 5% relative humidity <strong>drying</strong> conditions the bSA particles have a density <strong>of</strong><br />

averaged 0.62 g/cm 3 , whereas the mixture particles with trehalose have a density <strong>of</strong> 0.95 g/cm 3 .<br />

Figure 5.96 presents that the <strong>drying</strong> <strong>of</strong> the bSA-trehalose (1:1) mixture <strong>and</strong> the pure trehalose yields<br />

almost the same particle size <strong>and</strong> density. The pure trehalose particles are only marginally smaller<br />

in size <strong>and</strong> only marginally larger in density than the bSA-trehalose (1:1) mixture. The course <strong>of</strong> the<br />

density as a function <strong>of</strong> evaporation time for all three <strong>formulations</strong> at 60°C <strong>and</strong> 5% relative<br />

humidity in still air can be seen in Figure 5.97. All curves start at approximately the same point<br />

because <strong>of</strong> their similar liquid density. After an initial increase in density due to an increasing solid<br />

content <strong>and</strong> a decrease in <strong>droplet</strong> volume, the density decreases as soon as a constant <strong>droplet</strong><br />

volume is reached due to evaporation <strong>of</strong> liquid from inside the particle. Using this plot it can be<br />

seen that there are still changes in volume after the critical point, because <strong>of</strong> a further increase in<br />

density within the second <strong>drying</strong> stage. The maximum values <strong>of</strong> the density during the <strong>drying</strong>


CHAPTER 5 RESULTS AND DISCUSSION 183<br />

experiments are determined to 1.064 g/cm 3 for pure bSA solution, 1.244 g/cm 3 for the bSAtrehalose<br />

(1:1) mixture <strong>and</strong> 1.451 g/cm 3 for the pure trehalose solution. The final particle densities<br />

are equal to the data <strong>of</strong> Figure 5.96.<br />

Particle radius relative to initial <strong>droplet</strong> radius [%]<br />

56.0<br />

54.0<br />

52.0<br />

50.0<br />

48.0<br />

46.0<br />

44.0<br />

42.0<br />

0 10 20 30 40<br />

(a)<br />

Pure bSA (100 mg/ml)<br />

bSA-Trehalose (100 mg/ml)<br />

Pure trehalose (100 mg/ml)<br />

Relative humidity <strong>of</strong> <strong>drying</strong> air [%]<br />

0 10 20 30 40<br />

Figure 5.96: Comparison <strong>of</strong> the resulting particle size <strong>and</strong> density <strong>of</strong> pure bSA, bSA-trehalose (1:1) <strong>and</strong> pure trehalose<br />

solutions with 100 mg/ml solid content, dried 60°C with different humidity in still air. The initial <strong>droplet</strong> volume <strong>of</strong> all<br />

experiments was between 1.5 µl <strong>and</strong> 2.0 µl at an initial effective SPL <strong>of</strong> 163.1 dB to 165.9 dB. (a) Particle radius<br />

compared to the initial <strong>droplet</strong> radius versus <strong>drying</strong> air humidity. (b) Final particle density versus <strong>drying</strong> air humidity.<br />

Droplet / particle density [g/cm 3 ]<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

[60°C - 5% rel.humidity]<br />

Figure 5.97: Droplet / particle density as a function <strong>of</strong> (t / rS 0 2 ) for pure bSA, bSA-trehalose (1:1) mixture <strong>and</strong> pure<br />

trehalose solutions with a solid content <strong>of</strong> 100 mg/ml at 60°C <strong>and</strong> 5% relative humidity in still air. The initial <strong>droplet</strong><br />

volume <strong>of</strong> the experiments was 1.5 µl an initial effective SPL between 164.1 <strong>and</strong> 164.5 dB.<br />

Final particle density [g/cm 3 ]<br />

1.2<br />

1.1<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

(b)<br />

0 500 1000 1500 2000<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Pure bSA (100 mg/ml)<br />

bSA-Trehalose (100 mg/ml)<br />

Pure trehalose (100 mg/ml)<br />

Relative humidity <strong>of</strong> <strong>drying</strong> air [%]<br />

pure trehalose<br />

bSA-trehalose<br />

pure bSA


184 RESULTS AND DISCUSSION<br />

Levitation experiments with <strong>drying</strong> air velocity cannot be performed for the pure bSA solution. As<br />

with the experiments in still air, the <strong>droplet</strong> / particle shows an instable behaviour after the<br />

formation <strong>of</strong> a solid crust. In combination with the stress related to the direct air streaming the<br />

<strong>droplet</strong> is blown out <strong>of</strong> the st<strong>and</strong>ing acoustic wave into the environment in almost every experiment.<br />

In case that the particle stays levitated the <strong>droplet</strong> oscillations are so strong that a determination <strong>of</strong><br />

the evaporation rate in the second <strong>drying</strong> stage using the rise <strong>of</strong> the centre <strong>of</strong> mass <strong>of</strong> the particle to<br />

its next adjacent pressure node is not possible. The particles <strong>of</strong> these experiments are only used for<br />

taking SEM pictures (Figure 5.98). As with the particles obtained under still air conditions, the pure<br />

bSA particles show a hollow interior, whereas the mixture <strong>of</strong> bSA <strong>and</strong> trehalose has a massive<br />

particle structure. Analysing the SEM picture <strong>of</strong> a pure bSA particle dried at 60°C <strong>and</strong> 5% relative<br />

humidity in a <strong>drying</strong> air velocity <strong>of</strong> 1.0 m/s, the solid crust has a thickness <strong>of</strong> 22.34% <strong>of</strong> the particle<br />

radius. This is much smaller than the 31.75% obtained in still air. Assuming a spherical particle<br />

with a radius <strong>of</strong> 500 µm, the void volume inside the particle would be 0.245 µl at a total volume <strong>of</strong><br />

0.524 µl. With increasing <strong>drying</strong> air humidity <strong>and</strong> increasing <strong>drying</strong> time, the solid crust <strong>of</strong> the<br />

hollow bSA particle seems to get larger in relation to the particle radius. A large magnification <strong>of</strong><br />

the solid particle crust enables to see two to three different shell areas. The top level shell looks like<br />

a thin particle coating skin with small circular slots on the surface. Persisting capillaries cannot be<br />

seen. The next level is a porous area with lots <strong>of</strong> very small cavities. This thickness <strong>of</strong> this area is<br />

increasing with increasing <strong>drying</strong> air humidity. It is not found at 5% <strong>drying</strong> air humidity.<br />

(a)<br />

Figure 5.98: SEM pictures <strong>of</strong> pure bSA particles with 100 mg/ml solid content dried at 60°C <strong>and</strong> 5% relative humidity<br />

at a <strong>drying</strong> air velocity <strong>of</strong> 1.0 m/s. Shown are pictures with a magnification <strong>of</strong> 70-times (a) <strong>and</strong> 4000-times.<br />

(b)


CHAPTER 5 RESULTS AND DISCUSSION 185<br />

The innermost <strong>and</strong> also largest shell area appears massive <strong>and</strong> uniform on the SEM pictures. It does<br />

not show any holes, capillaries or inhomogeneity in its structure. The appearance <strong>of</strong> this area in<br />

Figure 5.99 d is associated to a bad waste edge. The discussed overall structure <strong>of</strong> the particle shell<br />

area is only seen at the outer surface <strong>of</strong> the particle <strong>and</strong> not at the inner surface developing during<br />

the <strong>drying</strong> process. The inner surface is formed by the third massive shell area.<br />

(a) (b)<br />

(c) (d)<br />

Figure 5.99: SEM pictures <strong>of</strong> pure bSA particles with 100 mg/ml solid content dried at 60°C <strong>and</strong> 1.0 m/s <strong>drying</strong> air<br />

velocity in 20% relative humidity (a-b) or 40% humidity (c-d). Shown are pictures with a magnification <strong>of</strong> 125-times<br />

(a), 1000-times (b), 100-times (c) <strong>and</strong> 500-times (d).


186 RESULTS AND DISCUSSION<br />

Interfacial behaviour <strong>and</strong> surface excess <strong>of</strong> bSA<br />

To investigate the earlier critical point <strong>of</strong> the pure bSA solutions <strong>and</strong> the smaller amount <strong>of</strong> solvent<br />

evaporated in the first stage <strong>of</strong> the <strong>drying</strong> process in comparison to the pure trehalose <strong>and</strong> the bSAtrehalose<br />

solution, the surface activities <strong>of</strong> the different substances were analysed using a bubble<br />

pressure tensiometer. Due to their dual hydrophobic / hydrophilic nature, dissolved <strong><strong>protein</strong>s</strong> tend to<br />

orient themselves so that the exposure <strong>of</strong> the hydrophobic parts <strong>of</strong> the <strong>protein</strong> to the aqueous<br />

solution is minimized [R<strong>and</strong>olph 2002]. In systems with air / water interfaces, <strong><strong>protein</strong>s</strong> tend to<br />

accumulate at these interfaces exposing their hydrophilic portion to water. This kind <strong>of</strong> orientation<br />

<strong>and</strong> surface adsorption can also occur at solid / water interfaces, such as those found during the<br />

<strong>drying</strong> process between the solid crust at the <strong>droplet</strong> /particle surface <strong>and</strong> the liquid interior<br />

[Fuhrling 2004; R<strong>and</strong>olph 2002]. If according to classical thermodynamics, the surface tension <strong>of</strong><br />

the interface exceeds the internal tension <strong>of</strong> the <strong>protein</strong>, the surface area <strong>of</strong> the <strong>protein</strong> must increase<br />

for example by unfolding, until the two values are equal [Verger et al. 1973]. If only an equilibrium<br />

monolayer would be formed by surface adsorption <strong>and</strong> unfolding <strong>of</strong> the <strong>protein</strong>, the amount <strong>of</strong><br />

adsorbed <strong>protein</strong> would be small [R<strong>and</strong>olph 2002; Verger et al. 1973]. However, additional<br />

processes like gas-to-liquid surface transitions, surface precipitation <strong>and</strong> the formation <strong>of</strong> surface<br />

sublayers can occur, depending on the degree <strong>of</strong> hydrophobicity <strong>and</strong> the characteristics <strong>of</strong> the<br />

<strong>protein</strong>. According to de Jongh et al. [2004], the events describing the adsorption process <strong>of</strong><br />

<strong><strong>protein</strong>s</strong> at air-liquid interfaces are firstly bulk diffusion that is determined by the molecular size.<br />

Secondly, net adsorption, which is related to the kinetic barrier determining the ratio <strong>of</strong> sticking to<br />

or bouncing from the interface. Thirdly, conformational changes that are determined by the<br />

activation energy required to disrupt part <strong>of</strong> the <strong>protein</strong>’s globular stability. Fourthly, <strong>protein</strong><br />

networking, determined by the chemical reactivity or physiochemical activity <strong>of</strong> side chains that are<br />

related to local system conditions like the dielectric constant. Fifthly, response to externally applied<br />

shear conditions. Additionally, the orientation <strong>of</strong> <strong>protein</strong> molecules in the adsorbed phase must be<br />

considered because <strong>of</strong> their non-uniformity in properties or structure across their exterior surface.<br />

Adsorbed <strong><strong>protein</strong>s</strong> are not free to rotate, because <strong>of</strong> multiple bonding between <strong>protein</strong> <strong>and</strong> adsorbed<br />

surface <strong>and</strong> therefore a fixed array <strong>of</strong> exterior amino acids residues <strong>of</strong> the <strong>protein</strong> is exposed to the<br />

bulk phase [Horbett 1992]. The driving force <strong>of</strong> <strong>protein</strong> adsorption at interfaces is a decrease in<br />

entropy <strong>of</strong> the water molecules that are ordered around the <strong>protein</strong>’s hydrophobic parts when the<br />

<strong>protein</strong> is in bulk solution. It is determined not only by the hydrophobicity but also by the flexibility<br />

<strong>of</strong> the <strong>protein</strong>. A flexible <strong>protein</strong> can expose addition non-polar residues, leading to an increase in<br />

strength <strong>of</strong> the binding to the surface [Tripp 1995]. Adsorption <strong>and</strong> unfolding <strong>of</strong> the <strong>protein</strong> can lead<br />

to nearly a complete loss <strong>of</strong> native activity [R<strong>and</strong>olph 2002; Verger et al. 1973]. In contrast, a gain


CHAPTER 5 RESULTS AND DISCUSSION 187<br />

<strong>of</strong> structure, for example by an increase in α-helical content, is reported in literature for some<br />

<strong><strong>protein</strong>s</strong> [Caessens et al. 1999]. When <strong><strong>protein</strong>s</strong> are adsorbed at interfaces, the interfacial tension will<br />

decrease dynamically from the value <strong>of</strong> the pure solvent to the lower equilibrium value. Tripp <strong>and</strong><br />

Magda [Tripp 1995] divide this process into three different dynamic interfacial kinetic regimes.<br />

First, the induction time with little or no apparent decrease in the interfacial tension. Secondly, a<br />

rapid interfacial tension decrease when approximately 50% <strong>of</strong> the monolayer surface coverage is<br />

reached. Thirdly, the meso-equilibrium interfacial tension that results in the steady-state interfacial<br />

tension when the <strong>protein</strong> molecules have achieved their equilibrium confirmation <strong>and</strong> surface<br />

concentration at the interface.<br />

Figure 5.100 shows the air / water interfacial tension versus log time isotherms obtained<br />

with bSA solutions with concentrations varied from 0.1 to 4.0 mmol for a solution volume <strong>of</strong> 10 ml<br />

in a bubble pressure tensiometer. Even after an interfacial lifetime <strong>of</strong> only 1 ms (log t = -3.0) the<br />

values for the dynamic interfacial tension are lower than that <strong>of</strong> 72.8 mN/m measured for pure<br />

water. For low bSA concentration (0.1 mmol) the values after this time lies only marginally below<br />

pure water. With increasing solid concentration <strong>and</strong> increasing interfacial lifetime the interfacial<br />

tension decreases. Because <strong>protein</strong> adsorption <strong>and</strong> rearrangement extends over long time <strong>of</strong> a few<br />

hours to some days before equilibrium is reached <strong>and</strong> long-time approximation is regarded more<br />

accurate than short-time, recent studies on <strong><strong>protein</strong>s</strong> dynamic interfacial tension have measured at<br />

interfacial lifetimes ≥ 1 s <strong>and</strong> applied long-time approximation to describe the interfacial tension<br />

versus time [Chen et al. 1998; Tripp 1995]. In spray-<strong>drying</strong> however, the relevant time scale for<br />

atomisation <strong>and</strong> <strong>droplet</strong> <strong>drying</strong> is < 1 s [Nurnberg 1980], in which it is expected that short time<br />

approximation is valid. The time range <strong>of</strong> a single <strong>droplet</strong> <strong>drying</strong> experiment in the acoustic<br />

levitation system at 60°C is from seconds to a few minutes depending on the <strong>drying</strong> air humidity<br />

<strong>and</strong> velocity.<br />

The Ward <strong>and</strong> Tordai Equation considers diffusion-controlled adsorption to the surface <strong>of</strong><br />

a <strong>protein</strong> or surfactant solution in equilibrium with its bulk which is suddenly exp<strong>and</strong>ed or<br />

compressed. They integrated the diffusion equation using Green’s function with the boundary<br />

condition for conservation <strong>of</strong> mass at the surface <strong>and</strong> the boundary conditions far away from the<br />

surface [Joos 1999].


188 RESULTS AND DISCUSSION<br />

Dynamic interfacial tension [mN/m]<br />

74.0<br />

72.0<br />

70.0<br />

68.0<br />

66.0<br />

64.0<br />

bSA solution<br />

0.1 mmol<br />

62.0<br />

0.5 mmol<br />

1.0 mmol<br />

60.0<br />

2.0 mmol<br />

58.0<br />

4.0 mmol<br />

0.0 1.0 2.0 3.0 4.0 5.0<br />

Effective time [s]<br />

Figure 5.100: Dynamic air / water interfacial tension as a function <strong>of</strong> time <strong>of</strong> bSA solutions with concentration <strong>of</strong><br />

minimum 0.1 mmol to 4.0 mmol measured at 25°C.<br />

1/<br />

2<br />

t<br />

⎛ D ⎞ ⎡<br />

⎤<br />

Equation 5.8 Γ () t = Γ0<br />

+ 2⎜ ⎟ ⋅ ⎢c0<br />

t − ∫ cS<br />

⋅ ( t −υ<br />

) ⋅ d υ ⎥<br />

⎝ π ⎠ ⎢⎣<br />

0<br />

⎥⎦<br />

Γ is the surface excess concentration, D the diffusion coefficient, c the concentration <strong>and</strong> υ the<br />

relaxation. In case that Γ 0 , then at early times the diffusion-controlled process 0 ≈ c <strong>and</strong> the<br />

0 =<br />

convolution integral <strong>of</strong> Ward <strong>and</strong> Tordai can be neglected, resulting in a “short-time”<br />

approximation [Joos 1999]<br />

⎛ D ⋅ t ⎞<br />

Equation 5.9 Γ () t = 2 ⋅ c0<br />

⋅ ⎜ ⎟<br />

⎝ π ⎠<br />

1/<br />

2<br />

In terms <strong>of</strong> surface pressure, Π () t γ − γ ( t)<br />

= Γ ( t)RT<br />

⎛ D ⋅ t ⎞<br />

Equation 5.10 Π () t = γ 0 − γ () t = 2 RTc0⎜<br />

⎟<br />

⎝ π ⎠<br />

= 0 , this can be rearranged to<br />

The measured dynamic surface pressure, Π ( t)<br />

, should therefore increase linearly with t in the<br />

early times <strong>of</strong> the adsorption process [Joos 1999].<br />

For a “long-time” approximation the Ward <strong>and</strong> Tordai equation is reformed as<br />

1/<br />

2 t<br />

⎛ D ⎞<br />

= ⎜ ⎟<br />

S<br />

⎝ π ∫ with Δ cS = cS<br />

− c0<br />

⎠ 0<br />

Equation 5.11 Γ () t −2<br />

⋅ Δ c ( t −υ<br />

) d t<br />

Dynamic interfacial tenion [mN/m]<br />

74.0<br />

72.0<br />

70.0<br />

68.0<br />

66.0<br />

64.0<br />

62.0<br />

60.0<br />

58.0<br />

bSA solution<br />

0.1 mmol<br />

0.5 mmol<br />

1.0 mmol<br />

2.0 mmol<br />

4.0 mmol<br />

-3.0 -2.0 -1.0 0.0 1.0<br />

Log <strong>of</strong> time in seconds [-]<br />

S


CHAPTER 5 RESULTS AND DISCUSSION 189<br />

Over the interval at large times, Δ cS<br />

changes little <strong>and</strong> Equation 5.11 can be given by<br />

1/<br />

2<br />

⎛ 4Dt<br />

⎞<br />

Equation 5.12 Γ () t = ⎜ ⎟ ( c0<br />

− cS<br />

)<br />

⎝ π ⎠<br />

If the change in concentration is linearized with respect to the interfacial tension, γ () t , <strong>and</strong> if the<br />

Gibbs equation dγ = −RT<br />

⋅ Γ ⋅ dc / c0<br />

is used, the “long-time” approximation to Ward <strong>and</strong> Tordai<br />

becomes<br />

Equation 5.13 γ () t<br />

γ<br />

− ∞<br />

0<br />

1/<br />

2<br />

2<br />

RT ⋅ Γ ⎛ π ⎞<br />

= ⋅ ⎜ ⎟<br />

c ⎝ 4Dt<br />

⎠<br />

Figure 5.101 shows the relationship between surface pressure <strong>and</strong> interfacial tension, to convert the<br />

above equations to representations <strong>of</strong> surface pressure. At low concentrations, when<br />

Π () t / Π ≤ 0.<br />

1 the short-time approximation is valid [Joos 1999]. The long-time approximation<br />

∞<br />

describes bulk-to-surface diffusion as single adsorption mechanism [Fuhrling 2004]. For short-time<br />

approximation the plot <strong>of</strong> Π () t / Π ∞ versus t should be linear, whereas for long-time<br />

approximation the plot <strong>of</strong> Π ∞ () t / Π ∞ versus 1 / t should be linear.<br />

Π (t) Π∞ (t)<br />

γ 0 γ (t) γ∞<br />

Π∞<br />

Figure 5.101: Relationship between surface pressure <strong>and</strong> interfacial tension [Fuhrling 2004].<br />

In Figure 5.102, Π () t / Π ∞ is plotted versus t for the bSA solutions for diffusion –controlled<br />

adsorption. The values <strong>of</strong> Π ∞ were determined by equilibrium ring tensiometry after an<br />

equilibration time <strong>of</strong> 3 hours <strong>and</strong> corrected using the method by Harkins <strong>and</strong> Jordan [Harkins 1930].<br />

The results are shown in Table 5.11. In comparison to data published by Tripp [1995] who found an<br />

meso-equilibrium interfacial tension <strong>of</strong> 52.7 mN/m for an aqueous 0.015 mmol bSA solution at


190 RESULTS AND DISCUSSION<br />

25°C <strong>and</strong> recently by Wen [2001], who found an equilibrium interfacial tension <strong>of</strong> 50 mN/m for an<br />

aqueous bSA solution <strong>of</strong> the same concentration, the experimental values in this work are higher.<br />

Table 5.11: Values for equilibrium interfacial tension after 3 hours for aqueous bSA solutions at 25°C ambient<br />

temperature using ring tensiometry.<br />

Concentration <strong>of</strong><br />

bSA solution<br />

Equilibrium<br />

interfacial tension<br />

[ mmol ] 0.1 0.5 1.0 2.0 4.0<br />

[ mN/m ] 57.4 56.5 56.1 54.9 54.4<br />

Π (t) / Π ∞<br />

Figure 5.102: Π (t) / Π ∞ as a function <strong>of</strong> time ½ <strong>of</strong> bSA solutions with different molar concentration at 25°C measured<br />

with bubble pressure tensiometry.<br />

In Figure 5.102 the plot <strong>of</strong> the short-time approximation shows a biphasic behaviour with two<br />

almost linear ranges separated by a clear break point at approximately 500 ms ( t = 0.<br />

7 ). The<br />

slopes <strong>of</strong> the initial section at t ≤ 500 ms increase with higher <strong>protein</strong> concentration, but are almost<br />

equal for the three most concentrated bSA solutions. The linearity occurs at longer adsorption times<br />

than that <strong>of</strong> Π () t / Π ≤ 0.<br />

1 stated by Joos [1999]. The slopes in the second linear section for<br />

∞<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

4.0 mmol<br />

2.0 mmol<br />

1.0 mmol<br />

0.5 mmol<br />

0.1 mmol<br />

0.0<br />

[bSA solution at 25°C]<br />

0.0 0.5 1.0 1.5 2.0 2.5 3.0<br />

Time 1/2 [s 1/2 ]<br />

t ≥ 500 ms are equal for all bSA concentrations. Π ( t ) / Π ∞ increases much slower than in the slope<br />

in the initial section. Thus, the diffusional adsorption is evidently reduced in rate at the latter points<br />

in the curves <strong>of</strong> Figure 5.102. According to Fuhrling [2004], it seems that a barrier to adsorption is<br />

formed.


CHAPTER 5 RESULTS AND DISCUSSION 191<br />

Figure 5.103 presents a plot <strong>of</strong> the dynamic interfacial tension versus the total molarity <strong>of</strong><br />

the bSA solution at increasing time. The values decrease with increasing concentration <strong>of</strong> bSA <strong>and</strong><br />

increasing time. Only the 1 ms curve shows tendency <strong>of</strong> a levelling-out at approximately 2 mmol/l.<br />

The curves at 10 ms, 100 ms <strong>and</strong> 1000 ms show a further decrease in interfacial tension with<br />

increasing solution molarity <strong>of</strong> bSA.<br />

Figure 5.103: Dynamic air / water interfacial tension as a function <strong>of</strong> solution molarity <strong>of</strong> bSA solutions at different<br />

times.<br />

The total interfacial excess concentration <strong>of</strong> bSA is calculated using the Gibbs equation.<br />

Transformation <strong>of</strong> the original equation leads to a form that shows that the slope <strong>of</strong> the plot <strong>of</strong> the<br />

dynamic interfacial tension versus ln c defines the interfacial excess <strong>of</strong> a solute [Hiemenz 1986].<br />

Equation 5.14<br />

Dynamic interfacial tension [mN/m]<br />

74.0<br />

72.0<br />

70.0<br />

68.0<br />

66.0<br />

64.0<br />

62.0<br />

60.0<br />

[bSA solutions at 25°C]<br />

0.0 1.0 2.0 3.0 4.0 5.0<br />

1 ⎛ dγ<br />

⎞<br />

Γ = − ⋅ ⎜ ⎟<br />

RT ⎝ d ln c ⎠<br />

Solution molarity [mmol/l]<br />

T<br />

10 ms<br />

100 ms<br />

1000 ms<br />

Figure 5.104 shows the calculated surface excess concentration plotted versus solution molarity <strong>of</strong><br />

the bSA solutions for different interfacial lifetime. As expected, an increase in Γ (t)<br />

can be seen<br />

with increasing <strong>protein</strong> concentration. The formation <strong>of</strong> a plateau with increasing bSA<br />

concentration cannot be seen from Figure 5.104. Additionally to the plots <strong>of</strong> the surface tension<br />

with time or concentration, the curve for the extrapolated values at 10 s lifetime is also shown. After<br />

1 s interfacial lifetime the surface excess value increases up to 0.00132 mmol/m 2 for a 4.0 mmol/l<br />

bSA solution. After 10 s the surface excess is increased to 0.00146 mmol/m 2 . These values are<br />

equivalent to 88 mg/m 2 – 97 mg/m 2 . Chen et al. [1998] determined a long time equilibrium surface<br />

excess value for bSA <strong>of</strong> approximately 12 mg/m 2 using the Gibbs equation. They noted that this<br />

1 ms


192 RESULTS AND DISCUSSION<br />

value corresponds to a close-packed monolayer <strong>of</strong> the <strong>protein</strong>. The values in this work are seven<br />

times larger than that <strong>of</strong> Chen et al.. Repeated measurement <strong>of</strong> the bSA solutions using a different<br />

batch <strong>of</strong> lyophilized bSA produced the same results. Figure 5.105 presents the surface excess<br />

concentration versus interfacial lifetime with different solid concentration. In the time range<br />

between 1 s <strong>and</strong> 10 s a further increase in surface excess concentration can be seen for all bSA<br />

concentrations.<br />

Surface excess [mmol/m 2 ]<br />

0.0016<br />

0.0014<br />

0.0012<br />

0.0010<br />

0.0008<br />

0.0006<br />

0.0004<br />

0.0002<br />

0.0000<br />

10000 ms<br />

0.000 0.001 0.002 0.003 0.004 0.005<br />

Solution molarity [mmol/m 3 ]<br />

1000 ms<br />

100 ms<br />

10 ms<br />

Figure 5.104: Surface excess concentration Γ (t) as a function <strong>of</strong> bSA solution molarity at 25°C for different lifetimes.<br />

Surface excess [mmol/m 2 ]<br />

0.0016<br />

0.0014<br />

0.0012<br />

0.0010<br />

0.0008<br />

0.0006<br />

0.0004<br />

0.0002<br />

0.0000<br />

Figure 5.105: Surface excess concentration as a function <strong>of</strong> time <strong>of</strong> bSA solutions at various concentration at 25°C.<br />

1 ms<br />

[bSA solution at 25°C]<br />

4.0 mmol<br />

2.0 mmol<br />

1.0 mmol<br />

0.5 mmol<br />

0.0 1.0 2.0 9.0 10.0 11.0 12.0<br />

Time [s]<br />

0.1 mmol<br />

[bSA solution at 25°C]


CHAPTER 5 RESULTS AND DISCUSSION 193<br />

A comparison <strong>of</strong> the surface excess concentration <strong>of</strong> bSA solutions with other <strong>protein</strong> solutions is<br />

presented in Figure 5.106. Trypsinogen from bovine pancreas with molecular weight <strong>of</strong> 23.7 kDa<br />

[Sigma 2005] <strong>and</strong> salmon calcitonin with a molecular weight <strong>of</strong> 3.43 kDa [Sigma 2005] were<br />

examined by Fuhrling [2004]. The plots show the surface excess concentration <strong>of</strong> a 4.0 mmol/l<br />

solution versus time for all three <strong><strong>protein</strong>s</strong> <strong>and</strong> the surface excess concentration with solution<br />

molarity for a lifetime <strong>of</strong> 1 s.<br />

Surface excess [mmol/m 2 ]<br />

0.0014<br />

0.0012<br />

0.0010<br />

0.0008<br />

0.0006<br />

0.0004<br />

0.0002<br />

bSA (this work)<br />

0.0000<br />

Trypsinogen [Fuhrling 2004]<br />

Calcitonin [Fuhrling 2004]<br />

0.000 0.001 0.002 0.003 0.004<br />

(a)<br />

[Lifetime 1s at 25°C]<br />

Solution molarity [mmol/m 3 ]<br />

0.0014<br />

0.0012<br />

0.0010<br />

0.0008<br />

0.0006<br />

0.0004<br />

0.0002<br />

0.0000<br />

Figure 5.106: Surface excess concentration <strong>of</strong> bSA, trypsinogen, <strong>and</strong> salmon calcitonin. (a) Surface excess versus<br />

solution molarity at 1 s lifetime. (b) Surface excess verusus time for a 4.0 mmol/l solution.<br />

At an interfacial lifetime <strong>of</strong> 1 s the surface excess values <strong>of</strong> trypsinogen increase to<br />

0.0006 mmol/m 2 for 1 mmol/kg, <strong>and</strong> 0.0008 mmol/m 2 for 8 mmol/kg. These values are equivalent<br />

to 14 mg/m 2 – 19mg/m 2 [Fuhrling 2004]. According to Figure 5.106 <strong>and</strong> influence <strong>of</strong> the molecular<br />

weight <strong>of</strong> the <strong>protein</strong> on the surface excess cannot be seen, because bSA with a higher molecular<br />

weight than trypsinogen as well as calcitonin with a lower molecular weight than trypsinogen show<br />

a larger surface excess concentration.<br />

Interfacial behaviour <strong>and</strong> surface excess <strong>of</strong> Pluconic F 127<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

To assess the potential <strong>of</strong> interfacial activity <strong>of</strong> the bSA solutions, Pluronic F 127 solutions with the<br />

same solution molarity are prepared <strong>and</strong> tested in a bubble pressure tensiometer the same way as the<br />

bSA solutions before. Pluronic F 127 is chosen due to its polymeric character with a molecular<br />

weight <strong>of</strong> 12.6 kDa [Seth 2005]. Figure 5.107 shows the surface tension versus log time (Figure<br />

5.107 a) <strong>and</strong> solution molarity (Figure 5.107 b). The equilibrium interfacial tension values measured<br />

by ring tensiometry are listed in Table 5.12.<br />

Surface excess [mmol/m 2 ]<br />

(b)<br />

bSA (this work)<br />

Trypsinogen [Fuhrling 2004]<br />

Calcitonin [Fuhrling 2004]<br />

[Solution molarity 4.0mmol at 25°C]<br />

Time [s]


194 RESULTS AND DISCUSSION<br />

Table 5.12: Values for equilibrium interfacial tension after 3 hours for aqueous Pluronic F 127 solutions at 25°C<br />

ambient temperature using ring tensiometry.<br />

Concentration <strong>of</strong><br />

Pluronic solution<br />

Equilibrium<br />

interfacial tension<br />

[ mmol ] 0.1 0.5 1.0 2.0 4.0<br />

[ mN/m ] 50.1 43.9 41.0 38.3 37.0<br />

The Pluronic F 127 solutions show a higher surface activity at equivalent concentrations than does<br />

the bSA. In all plots the interfacial tension decreases continually as seen in Figure 5.107 a. At<br />

lifetimes ≥ 1 s the slope <strong>of</strong> the plot, γ / d logt<br />

, tends to increase with decreasing surfactant<br />

d C<br />

concentration. With increasing solution molarity <strong>and</strong> lifetime the interfacial tension decreases,<br />

indicating a higher amount <strong>of</strong> surface coverage. Due to the higher degree <strong>of</strong> surface coverage at<br />

larger Pluronic concentration, the adsorption rate is reduced.<br />

Dynamic interfacial tension [mN/m]<br />

65.0<br />

60.0<br />

55.0<br />

50.0<br />

45.0<br />

40.0<br />

(a)<br />

0.1 mmol<br />

0.5 mmol<br />

1.0 mmol<br />

2.0 mmol<br />

4.0 mmol<br />

[Pluronic F 127 solutions at 25°C]<br />

-3 -2 -1 0 1<br />

Log time in seconds [-]<br />

35.0<br />

0.0 1.0 2.0 3.0 4.0 5.0<br />

Solution molarity [mmol/l]<br />

Figure 5.107: Dynamic interfacial tension <strong>of</strong> Pluronic F 127 solutions at 25°C. (a) Dynamic interfacial versus log time<br />

in seconds for different Pluronic concentrations. (b) Dynamic interfacial tension versus solution molarity at different<br />

lifetime.<br />

The plots <strong>of</strong> the calculated surface excess concentration versus solution molarity <strong>of</strong> the<br />

Pluronic F 127 solution <strong>and</strong> interfacial lifetime are presented in Figure 5.108. The Pluronic curves<br />

show a greater surface excess concentration with increasing solution molarity <strong>and</strong> increasing<br />

interfacial lifetime. The experimental results are smaller than those published by Fuhrling [2004]<br />

for equivalent concentration <strong>of</strong> the solution. The surface excess values for Pluronic F 127 are larger<br />

than the bSA ones with the same molar concentration.<br />

Dynamic interfacial tension [mN/m]<br />

75.0<br />

70.0<br />

65.0<br />

60.0<br />

55.0<br />

50.0<br />

45.0<br />

40.0<br />

(b)<br />

[Pluronic F 127 solutions at 25°C]<br />

1 ms<br />

10 ms<br />

100 ms<br />

1000 ms


CHAPTER 5 RESULTS AND DISCUSSION 195<br />

Surface excess [mmol/m 2 ]<br />

0.0025<br />

0.0020<br />

0.0015<br />

0.0010<br />

0.0005<br />

100 ms<br />

10 ms<br />

1 ms<br />

0.0000<br />

[Pluronic F 127 solutions at 25°C]<br />

0.000 0.001 0.002 0.003 0.004 0.005<br />

(a)<br />

Solution molarity [mmol/m 3 ]<br />

1000 ms<br />

Figure 5.108: Surface excess concentration <strong>of</strong> Pluronic F 127 solutions at 25°C. (a) Surface excess concentration<br />

versus solution molarity at different interfacial lifetimes. (b) Surface excess concentration versus interfacial lifetime for<br />

different solid concentrations.<br />

Influence <strong>of</strong> trehalose on the interfacial behaviour <strong>of</strong> bSA<br />

Surface excess [mmol/m 2 ]<br />

0.0025 4.0 mmol<br />

2.0 mmol<br />

1.0 mmol<br />

0.0020<br />

0.5 mmol<br />

0.0015<br />

0.0010<br />

0.0005<br />

0.0000<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4<br />

Time [s]<br />

0.1 mmol<br />

[Pluronic F 127 solutions at 25°C]<br />

The effect <strong>of</strong> trehalose on the surface tension <strong>and</strong> other physico-chemical properties <strong>of</strong> water has<br />

been studied <strong>and</strong> discussed in detail by Kaushik <strong>and</strong> Bhat [Kaushik 2003]. It has been observed, that<br />

trehalose increases the surface tension <strong>of</strong> water by a larger amount than other sugars <strong>and</strong> polyols do.<br />

The thermodynamic properties <strong>of</strong> water like partial molal heat capacity <strong>and</strong> volume show a<br />

markable increase in the presence <strong>of</strong> trehalose [Kaushik 2003]. A larger hydrated volume compared<br />

to other sugars [Sola-Penna 1998] <strong>and</strong> a linear correlation <strong>of</strong> the M T Δ ( T M = midpoint <strong>of</strong> thermal<br />

denaturation) with increasing surface tension due to trehalose has also been reported<br />

[Kaushik 2003]. Other effects <strong>of</strong> trehalose on the stabilisation <strong>of</strong> <strong><strong>protein</strong>s</strong> have already been<br />

discussed in Chapter 5.3.2. Fuhrling [2004] showed an increase in the dynamic air / water<br />

interfacial tension with increasing solid concentration. For a 50 mg/ml solution he found an<br />

interfacial tension <strong>of</strong> averaged 73.1 mN/m increasing to approximately 73.8 mN/m for a 400 mg/ml<br />

solution. Due to this increase in surface tension with increasing solid content, the calculated surface<br />

excess was negative (-0.002 mmol/m 2 ) for a trehalose solution with 100 mg/ml solid content at an<br />

interfacial lifetime <strong>of</strong> 1 s. The results <strong>of</strong> pure trehalose solutions with a solid content <strong>of</strong> 50 mg/ml,<br />

100 mg/ml, 200 mg/ml <strong>and</strong> 400 mg/ml are presented in Table 5.13. The values <strong>of</strong> the equivalent<br />

solutions are larger than the above. To investigate the influence <strong>of</strong> trehalose on the interfacial<br />

tension <strong>of</strong> bSA solutions within this work, the same molar concentration <strong>of</strong> bSA with 200 mg/ml<br />

trehalose were prepared <strong>and</strong> tested equal to the pure bSA <strong>and</strong> Pluronic F 127. Therefore, the density<br />

<strong>of</strong> the 1 mmol/l bSA solution increased from 1.01 g/cm 3 without trehalose to 1.09 g/cm 3 with<br />

(b)


196 RESULTS AND DISCUSSION<br />

trehalose. Figure 5.109 shows the influence <strong>of</strong> trehalose on the dynamic interfacial tension <strong>of</strong> bSA<br />

solutions as a function <strong>of</strong> solution molarity for an interfacial lifetime <strong>of</strong> 0.1 s <strong>and</strong> 1.0 s as well as the<br />

dynamic interfacial tension as a function <strong>of</strong> interfacial lifetime for a solution molarity <strong>of</strong> 2.0 mmol/l<br />

<strong>and</strong> 4.0 mmol/l.<br />

Table 5.13: Equilibrium interfacial tension <strong>of</strong> trehalose solutions with different solid content measured with bubble<br />

pressure tensiometry at 25°C.<br />

Concentration <strong>of</strong><br />

trehalose solution<br />

Average interfacial<br />

tension (n = 3)<br />

Dynamic interfacial tension [mN/m]<br />

Dynamic interfacial tension [mN/m]<br />

76.0<br />

74.0<br />

72.0<br />

70.0<br />

68.0<br />

66.0<br />

64.0<br />

62.0<br />

(a)<br />

74.0<br />

72.0<br />

70.0<br />

68.0<br />

66.0<br />

64.0<br />

62.0<br />

(c)<br />

[ mg/ml ]<br />

50 100 200 400<br />

[ mN/m ] 72.80 ± 0.14 73.13 ± 0.08 74.02 ± 0.20 75.09 ± 0.20<br />

Interfacial lifetime 100 ms<br />

bSA solution<br />

bSA solution + 200 mg/ml trehalose<br />

0.0 1.0 2.0 3.0 4.0<br />

bSA solution molarity [mmol/l]<br />

Solution bSA molarity 2.0 mmol/l<br />

bSA solution<br />

bSA solution + 200 mg/ml trehalose<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

Time [s]<br />

0.0 1.0 2.0 3.0 4.0<br />

Figure 5.109: Dynamic interfacial tension <strong>of</strong> bSA solution <strong>and</strong> bSA solution with 200 mg/ml trehalose content a<br />

function <strong>of</strong> solution molarity <strong>and</strong> interfacial lifetime. (a) Dynamic interfacial tension versus solution molarity for an<br />

interfacial lifetime <strong>of</strong> 100 ms. (b) Dynamic interfacial tension versus solution molarity for an interfacial lifetime <strong>of</strong><br />

1000 ms. (c) Dynamic interfacial tension versus interfacial lifetime for a bSA solution molarity <strong>of</strong> 2.0 mmol/l. (d)<br />

Dynamic interfacial tension versus interfacial lifetime for a bSA solution molarity <strong>of</strong> 4.0 mmol/l.<br />

Dynamic interfacial tension [mN/m]<br />

Dynamic interfacial tension [mN/m]<br />

76.0<br />

74.0<br />

72.0<br />

70.0<br />

68.0<br />

66.0<br />

64.0<br />

62.0<br />

(b)<br />

74.0<br />

72.0<br />

70.0<br />

68.0<br />

66.0<br />

64.0<br />

62.0<br />

(d)<br />

bSA solution<br />

bSA solution + 200 mg/ml Trehalose<br />

Interfacial lifetime 1000 ms<br />

bSA solution molarity [mmol/l]<br />

bSA solution<br />

bSA solution + 200 mg/ml trehalose<br />

Solution bSA molarity 4.0 mmol/l<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

Time [s]


CHAPTER 5 RESULTS AND DISCUSSION 197<br />

The presence <strong>of</strong> trehalose reduces the adsorption <strong>of</strong> bSA to the surface. After an interfacial lifetime<br />

<strong>of</strong> 1 s an increase in the dynamic interfacial tension <strong>of</strong> 3.67 mN/m for the 2.0 mmol/l bSA solution<br />

<strong>and</strong> <strong>of</strong> 4.42 mN/m for the 4.0 mmol/l bSA solution can be detected. Similar results are obtained<br />

plotting the dynamic interfacial tension as a function <strong>of</strong> bSA solution molarity. Trehalose increases<br />

the slope d γ () t / dc <strong>of</strong> the plots for interfacial lifetimes <strong>of</strong> 100 ms <strong>and</strong> 1 s. Figure 5.110 presents the<br />

surface excess concentration versus bSA solution molarity <strong>and</strong> interfacial lifetime for bSA solutions<br />

<strong>and</strong> the bSA-trehalose mixture with a trehalose content <strong>of</strong> 200 mg/ml. Trehalose decreases the<br />

surface excess <strong>of</strong> bSA for every solution molarity at interfacial lifetimes <strong>of</strong> 100 ms <strong>and</strong> 1 s. For a<br />

4.0 mmol/l solution the addition <strong>of</strong> trehalose reduces the surface excess <strong>of</strong> bSA to 41% <strong>of</strong> the value<br />

<strong>of</strong> pure bSA. The plot <strong>of</strong> the surface excess versus interfacial lifetime in Figure 5.110 c <strong>and</strong> d shows<br />

a decrease in slope d Γ () t / dt with addition <strong>of</strong> trehalose at equivalent bSA concentration. This leads<br />

to a reduction <strong>of</strong> bSA at the surface from 87 mg/m 2 for the pure 4.0 mmol/l bSA solution at an<br />

interfacial lifetime <strong>of</strong> 1 s to 36 mg/m 2 for the mixture with 200 mg/ml trehalose. The values at 10 s<br />

are extrapolated from the experimental data. The role <strong>of</strong> the surface tension in the stabilization <strong>of</strong><br />

RNase A with trehalose was investigated by Lin <strong>and</strong> Timasheff [Lin 1996] <strong>and</strong> the preferential<br />

interactions <strong>of</strong> trehalose <strong>and</strong> ribonuclease as well as the thermal unfolding <strong>of</strong> the <strong>protein</strong> was<br />

examined by Xie <strong>and</strong> Timasheff [Xie 1997]. Preferential interaction measurements demonstrated<br />

that trehalose unlike other sugars is totally excluded from the first hydration shell <strong>of</strong> the <strong>protein</strong><br />

[Lin 1996]. On a scale comparing the preferential interactions <strong>of</strong> co-solvents published by<br />

Timasheff [2002], trehalose shows the most preferential hydration <strong>of</strong> the <strong>protein</strong> with the highest<br />

amount <strong>of</strong> preferential exclusion. The results were supported by Wyman linkage analysis done by<br />

Kaushik [2003]. The negative values <strong>of</strong> the Wyman slope <strong>of</strong> the tangent from -7 to -4 for various<br />

<strong><strong>protein</strong>s</strong> at 1.5 M trehalose concentration indicated the preferential exclusion <strong>of</strong> trehalose from the<br />

hydration shell <strong>of</strong> the <strong>protein</strong> upon denaturation. Lins et al. [2004] set-up a model for <strong>protein</strong>trehalose<br />

interactions in aqueous solution on the nanosecond timescale. Initially, trehalose is<br />

distributed homogeneously in solution. The presence <strong>of</strong> a <strong>protein</strong> induces clustering <strong>of</strong> trehalose<br />

molecules <strong>and</strong> the trapping <strong>of</strong> a thin water-layer at the <strong><strong>protein</strong>s</strong> surface leading to preferential<br />

hydration. The trehalose molecules in the coating layer then compete with the <strong>protein</strong> for forming<br />

hydrogen bonds with the water molecules in the trapped layer. Recapitulating, preferential<br />

hydration <strong>of</strong> the <strong>protein</strong> in a triphasic system <strong>of</strong> water, <strong>protein</strong> <strong>and</strong> co-solvent postulates, that the cosolvent<br />

(e.g. a sugar) is excluded from vicinal water that composes the solvation layer <strong>of</strong> the<br />

<strong>protein</strong>. The <strong>protein</strong> then becomes preferentially hydrated with a decrease in radius <strong>of</strong> the salvation<br />

layer <strong>and</strong> the apparent volume <strong>of</strong> the <strong>protein</strong> [Sola-Penna 1998]. This leads to more stable <strong>protein</strong><br />

confirmation <strong>and</strong> enhances the maintenance <strong>of</strong> the native state, since the unfolded state becomes


198 RESULTS AND DISCUSSION<br />

thermodynamically even less favourable [Lee 1981]. This mechanism would increase the free<br />

energy <strong>of</strong> adsorption <strong>of</strong> the <strong>protein</strong> to the interface <strong>and</strong> hence reduces its extend <strong>of</strong> adsorption<br />

[Fuhrling 2004]. A reduced extend or slower rate <strong>of</strong> adsorption <strong>of</strong> the <strong>protein</strong> to the air / water<br />

interface could be a possible explanation for the increase in the amount evaporated in the first<br />

<strong>drying</strong> stage <strong>and</strong> the shift <strong>of</strong> the critical point in the single <strong>droplet</strong> <strong>drying</strong> experiment <strong>of</strong> pure bSA,<br />

bSA-trehalose (1:1) <strong>and</strong> pure trehalose, as presented in Figure 5.94.<br />

Surface excess [mmol/m 2 ]<br />

Surface excess [mmol/m 2 ]<br />

0.0016<br />

0.0014<br />

0.0012<br />

0.0010<br />

0.0008<br />

0.0006<br />

0.0004<br />

0.0002<br />

0.0000<br />

(a)<br />

0.0016<br />

0.0014<br />

0.0012<br />

0.0010<br />

0.0008<br />

0.0006<br />

0.0004<br />

0.0002<br />

0.0000<br />

0.000 0.001 0.002 0.003 0.004 0.005<br />

(c)<br />

Interfacial lifetime 100 ms<br />

bSA solution<br />

bSA solution + 200 mg/ml trehalose<br />

bSA solution molarity [mmol/m 3 ]<br />

bSA solution molarity 0.002 mmol/m 3<br />

bSA solution<br />

bSA solution + 200 mg/ml trehalose<br />

0.0 0.5 1.0 9.0 10.0<br />

Time [s]<br />

Surface excess [mmol/m 2 ]<br />

Surface excess [mmol/m 2 ]<br />

0.0016<br />

0.0014<br />

0.0012<br />

0.0010<br />

0.0008<br />

0.0006<br />

0.0004<br />

0.0002<br />

0.0000<br />

Interfacial lifetime 1000 ms<br />

bSA solution<br />

bSA solution + 200 mg/ml trehalose<br />

0.000 0.001 0.002 0.003 0.004 0.005<br />

bSA solution molarity [mmol/m 3 ]<br />

Figure 5.110: Surface excess concentration <strong>of</strong> bSA solution <strong>and</strong> bSA solution with 200 mg/ml trehalose content a<br />

function <strong>of</strong> solution molarity <strong>and</strong> interfacial lifetime. (a) Surface excess concentration versus solution molarity for an<br />

interfacial lifetime <strong>of</strong> 100 ms. (b) Surface excess concentration versus solution molarity for an interfacial lifetime <strong>of</strong><br />

1000 ms. (c) Surface excess concentration versus interfacial lifetime for a bSA solution molarity <strong>of</strong> 0.002 mmol/m 3 .<br />

(d) Surface excess concentration versus interfacial lifetime for a bSA solution molarity <strong>of</strong> 0.004 mmol/m 3 .<br />

(b)<br />

0.0016<br />

0.0014<br />

0.0012<br />

0.0010<br />

0.0008<br />

0.0006<br />

0.0004<br />

0.0002<br />

0.0000<br />

(d)<br />

bSA solution molarity 0.004 mmol/m 3<br />

bSA solution<br />

bSA solution + 200 mg/ml trehalose<br />

0.0 0.5 1.0 9.0 10.0<br />

Time [s]


CHAPTER 5 RESULTS AND DISCUSSION 199<br />

Determination <strong>of</strong> bSA monolayer<br />

A monomolecular layer, monolayer, <strong>of</strong> an air / water interfacial active substance can slow down the<br />

evaporation <strong>of</strong> the <strong>droplet</strong> solvent [Tuckermann 2002]. The formation <strong>and</strong> the characteristics <strong>of</strong><br />

such monolayers are usually investigated using a film balance. The determination via acoustical<br />

levitation <strong>of</strong> <strong>droplet</strong>s was first described by Frost <strong>and</strong> Seaver [Frost 1993] <strong>and</strong> experimentally<br />

studied <strong>and</strong> discussed in detail for 1-octadecanol, stearic acid <strong>and</strong> stearylamin by Tuckermann<br />

[2002]. The addition <strong>of</strong> an interfacial active substance to a levitated pure solvent <strong>droplet</strong> results in a<br />

decrease in surface tension <strong>and</strong> a stronger deformation <strong>of</strong> the <strong>droplet</strong>. The aspect ratio increases due<br />

to the larger vertical than horizontal levitation forces. The process <strong>of</strong> monolayer formation <strong>of</strong><br />

surfactant molecules is described by Hiemenz [1986] <strong>and</strong> shown in Figure 5.111 a to e. In the first<br />

<strong>and</strong> second state, the so called gaseous state <strong>and</strong> the liquid-exp<strong>and</strong>ed state, the surfactant molecules<br />

are too far away from each other <strong>and</strong> consequently the surface tension does not decrease very much<br />

(Figure 5.111, regions G <strong>and</strong> L1-G). With the evaporation <strong>of</strong> solvent <strong>of</strong> a <strong>droplet</strong> containing<br />

interfacial active substances <strong>and</strong>, therefore, a decrease in <strong>droplet</strong> volume <strong>and</strong> surface area, the<br />

surfactant molecules draw nearer causing a strong decrease in surface tension. This state is called<br />

the liquid condensed state (L1) <strong>and</strong> the solid state (L2, S) (Figure 5.111). Between the two states<br />

there is a break, known as the intermediate or transition state, <strong>and</strong> indicated by the capital I. Further<br />

evaporation <strong>of</strong> solvent <strong>and</strong> further compression leads to a collapse <strong>of</strong> the monolayer (Figure<br />

5.111 e). The pressure at which this occurs is somewhere in the vicinity <strong>of</strong> the equilibrium<br />

spreading pressure [Hiemenz 1986].<br />

Figure 5.111: Phases <strong>of</strong> the formation <strong>of</strong> monolayers <strong>of</strong> interfacial active substances on the surface <strong>of</strong> a levitated<br />

<strong>droplet</strong>. (left) Sketch <strong>of</strong> the surfactant orientation on the <strong>droplet</strong> surface. (right) States <strong>of</strong> the surfactant at the <strong>droplet</strong><br />

surface as a function <strong>of</strong> molecule area [Hiemenz 1986].


200 RESULTS AND DISCUSSION<br />

The surface tension <strong>of</strong> the <strong>droplet</strong> cannot be measured directly in the acoustic levitation system. But<br />

the transition from the gaseous to the liquid condensed state <strong>and</strong> the decrease in surface tension can<br />

be determined monitoring the aspect ratio <strong>and</strong> the surface temperature <strong>of</strong> the levitated <strong>droplet</strong><br />

[Tuckermann 2002]. Additionally, the evaporation rate decreases due to an evaporation repressing<br />

effect <strong>of</strong> the monolayer.<br />

Within this work, the formation <strong>of</strong> the liquid-condensed state <strong>of</strong> monolayers is examined<br />

only be monitoring the course <strong>of</strong> the aspect ratio <strong>of</strong> the levitated <strong>droplet</strong> with time. To examine if<br />

the determination <strong>of</strong> the point <strong>of</strong> formation is also possible without IR-thermocamera, 1octadecanol<br />

<strong>and</strong> stearic acid were examined <strong>and</strong> compared to the results published by Tuckermann<br />

[2002]. Therefore, a 1-octadecanol solution in hexane with a concentration <strong>of</strong> 0.143 g/l was<br />

prepared <strong>and</strong> dispersed in water in a relation <strong>of</strong> water to 1-octadecanol as 46.2 to 1. Stearic acid was<br />

also dissolved in hexane in a concentration <strong>of</strong> 0.l g/l <strong>and</strong> dispersed in water in a relation <strong>of</strong> water to<br />

Stearic acid as 22 to 1. The results <strong>and</strong> the comparison to literature are given in Table 5.14. A very<br />

good correlation is found so that a determination <strong>of</strong> the limiting area per molecule in a <strong>droplet</strong><br />

surface monolayer at the transition to the liquid-condensed phase can only be determined from the<br />

course <strong>of</strong> the aspect ratio with time.<br />

Table 5.14: Literature <strong>and</strong> experimental values for the limiting area per molecule in a <strong>droplet</strong> surface monolayer at the<br />

formation <strong>of</strong> the liquid-condensed state L1.<br />

Interfacial active<br />

substance<br />

Langmuir<br />

film balance<br />

[Stosch 2000]<br />

Langmuir<br />

film balance<br />

[Mingotaud et al. 1993]<br />

Acoustic<br />

levitator<br />

[Tuckermann 2002]<br />

Acoustic<br />

levitator<br />

(this work, n = 3)<br />

1-octadecanol - 0.21 nm 2 0.22 ± 0.03 nm 2 0,21 ± 0,02 nm 2<br />

Stearic acid 0.25 nm 2 0.25 nm 2 0.21 ± 0.03 nm 2 0.23 ± 0.04 nm 2<br />

To determine the area per bSA molecule at the formation <strong>of</strong> the liquid-condensed state <strong>of</strong> a bSA<br />

monolayer, a dilution row <strong>of</strong> bSA solutions is prepared <strong>and</strong> tested by evaporating the bSA solution<br />

<strong>droplet</strong> at room temperature <strong>and</strong> 0.3% humidity, <strong>and</strong> monitoring the size <strong>and</strong> aspect ratio <strong>of</strong> the<br />

levitated <strong>droplet</strong>s. When evaporating a <strong>droplet</strong> with an initial bSA concentration <strong>of</strong> 0.02 nmol/l <strong>and</strong><br />

an initial volume <strong>of</strong> 2.8 µl, an aspect ratio pr<strong>of</strong>ile similar to the pr<strong>of</strong>ile shown in Figure 5.111 for<br />

the change in surface tension versus molecule area is obtained. The plot is presented in Figure<br />

5.112. First, with decreasing <strong>droplet</strong> size, the aspect ratio decreases until the bSA molecules on the<br />

<strong>droplet</strong> surface draw that close that the liquid-condensed state is reached <strong>and</strong> the surface tension<br />

decreases. This results in an increase in aspect ratio due to the force distribution in the acoustic<br />

levitator. Equal to Figure 5.111 the plot <strong>of</strong> the aspect ratio versus <strong>drying</strong> time shows an increase in


CHAPTER 5 RESULTS AND DISCUSSION 201<br />

slope until a maximum is reached. After the maximum, the value <strong>of</strong> the aspect ratio decreases<br />

rapidly to almost 1.0. This decrease could possibly be caused by the collapse <strong>of</strong> the monolayer<br />

reported by Hiemenz [1986]. A decrease <strong>of</strong> the evaporation coefficient β r can also be seen in<br />

Figure 5.112 marked by the two red lines. The value determined for β r before the formation <strong>of</strong> the<br />

bSA monolayer in a size range from 1.0 µl to 0.5 µl is 0.000518 mm 2 /s which corresponds to the<br />

values <strong>of</strong> the pure water <strong>droplet</strong> experiments at 25°C <strong>and</strong> 0.1% relative humidity in still air. The<br />

decrease in β r is not associated with the formation <strong>of</strong> the liquid-condensed state, but with the<br />

transition form the liquid-condensed state to the solid state. β r decreases to 0.000273 mm 2 /s. If the<br />

aspect ratio is plotted versus the surface area per molecule, Figure 5.112 b is obtained. If the<br />

transition from gaseous state G to liquid-condensed state L1 is taken for the determination <strong>of</strong> the<br />

area per bSA molecule in the monolayer on the <strong>droplet</strong> surface, a value <strong>of</strong> approximately 100 nm<br />

per bSA molecule is obtained (Figure 5.112 a). Comparison with literature shows a very good<br />

correlation <strong>of</strong> the data [MacRitchie 1986].<br />

( r S / r S 0 ) 2 [-]<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

[bSA solution 0.02 nmol/l]<br />

0.2<br />

Droplet size<br />

Aspect ratio<br />

1.1<br />

0.0<br />

1.0<br />

0 200 400 600 800 1000 1200 1400<br />

(a)<br />

Evaporation time [s]<br />

1.6<br />

1.5<br />

1.4<br />

1.3<br />

1.2<br />

Figure 5.112: (a) Droplet size (rS /rS 0) 2 <strong>and</strong> <strong>droplet</strong> aspect ratio as a function <strong>of</strong> evaporation time for a 0.02 nmol/l<br />

aqueous bSA solution at 25°C <strong>and</strong> 0.3% relative humidity in still air. The initial <strong>droplet</strong> size was 2.8 µl at an initial<br />

effective SPL <strong>of</strong> 165.38 dB. (b) Aspect ratio <strong>of</strong> the levitated 0.02 nmol/l aqueous bSA solution as a function <strong>of</strong> surface<br />

area per bSA molecule. The experiment was performed at 25°C <strong>and</strong> 0.3% relative humidity in still air. The initial<br />

<strong>droplet</strong> volume was 2.8 µl at an initial effective SPL <strong>of</strong> 165.38 dB.<br />

Aspect ratio [-]<br />

Aspect ratio [-]<br />

1.6<br />

1.5<br />

1.4<br />

1.3<br />

1.2<br />

1.1<br />

[bSA solution 0.02 nmol/l]<br />

1.0<br />

300 250 200 150 100 50 0<br />

Surface area per bSA molecule [nm 2 (b)<br />

]<br />

G<br />

L 1<br />

L 2<br />

S


202 RESULTS AND DISCUSSION<br />

5.3.5 Evaporation <strong>of</strong> catalase solution <strong>droplet</strong>s<br />

Properties <strong>of</strong> catalase solutions<br />

The tetramer catalase consist <strong>of</strong> four equal sub-unit, each with 506 amino acids <strong>and</strong> an overall<br />

molecular weight <strong>of</strong> 250 kDa [Sigma 2005]. Catalase is poorly soluble in water, but soluble to<br />

1 mg/ml in 50 mM potassium phosphate buffer at pH 7.0 [Sigma 2005]. In this work a 50 mM tris<br />

buffer at pH 8.0 was used to prepare solutions with a solid content <strong>of</strong> 200 mg/ml. Only few<br />

publications are found on the spray-<strong>drying</strong> <strong>of</strong> catalase. One recent study was done by Liao et al.<br />

[2005] who investigated the physical stability <strong>of</strong> spray-dried <strong><strong>protein</strong>s</strong> (lysozyme <strong>and</strong> catalase)<br />

within surfactant-free hydr<strong>of</strong>luoroalkane (HFA) pressurized metered dose inhalers (pMDIs). Their<br />

model <strong><strong>protein</strong>s</strong> were spray-dried, stabilized in the presence <strong>of</strong> excipients <strong>and</strong> suspended within<br />

HFA. Catalase activity was determined equal to the method described in Chapter 4.2.6. Forbes et al.<br />

[1998] performed water vapour sorption studies on the physical stability <strong>of</strong> spray-dried<br />

<strong>protein</strong> / sugar powders using catalase, insulin <strong>and</strong> ribonuclease A as model <strong><strong>protein</strong>s</strong>. The <strong><strong>protein</strong>s</strong><br />

were spray-dried pure <strong>and</strong> formulated with lactose or mannitol in a Buchi Mini Spray-Dryer at<br />

T = 200° C <strong>and</strong> T = 90° C with an air flow <strong>of</strong> 800 l/h. Powder X-ray diffraction characterized<br />

inlet<br />

outlet<br />

the extent <strong>of</strong> cristallinity in the spray-dried product. The pure spray-dried <strong><strong>protein</strong>s</strong> <strong>and</strong> the<br />

<strong>protein</strong> / lactose-mixtures showed no presence <strong>of</strong> cristallinity, whereas the same <strong><strong>protein</strong>s</strong> co-spraydried<br />

with mannitol showed evidence <strong>of</strong> mannitol component cristallinity except <strong>of</strong> the<br />

RNase/mannitol mixture. When exposed to humidity most <strong>of</strong> their powders showed a transition<br />

from amorphous to crystalline state. However, catalase appeared to inhibit lactose crystallisation<br />

from the amorphous matrix to a greater extend than insulin in short-term humidity exposure test.<br />

Using freeze-<strong>drying</strong>, Liao et al. [2002] investigated the influence <strong>of</strong> type <strong>and</strong> amount <strong>of</strong> small<br />

molecular excipients (glycerol, sorbitol, sucrose <strong>and</strong> trehalose) on the preservation <strong>of</strong> the native<br />

<strong>protein</strong> structure <strong>and</strong> the biological activity <strong>of</strong> catalase <strong>and</strong> lysozyme. Catalase activity was<br />

determined equal to the method <strong>of</strong> Chapter 4.2.6 <strong>and</strong> FTIR-spectroscopy was used to examine<br />

secondary structure <strong>of</strong> the <strong>protein</strong> in the dry state. The native secondary structure <strong>of</strong> catalase<br />

determined in the dry state by FTIR had a weaker correlation with the residual catalase activity than<br />

the degree <strong>of</strong> retained native structure in solution determined by circular dichroism (CD). Jiang <strong>and</strong><br />

Nail [Jiang 1998] used catalase as well as β-galactosidase <strong>and</strong> lactate dehydrogenase (LDH) to<br />

investigate the maximization <strong>of</strong> recovery <strong>of</strong> <strong>protein</strong> activity after freeze-<strong>drying</strong> by manipulating <strong>of</strong><br />

freeze-dry process condition in the absence <strong>of</strong> protecting co-solutes.


CHAPTER 5 RESULTS AND DISCUSSION 203<br />

Pure catalase solutions as well as mixtures <strong>of</strong> catalase <strong>and</strong> trehalose were prepared in<br />

50 mM tris buffer at pH 8.0. All solutions had an overall solid content <strong>of</strong> 100 mg/ml. Experiments<br />

were carried out at 60°C <strong>and</strong> different relative humidity <strong>of</strong> the <strong>drying</strong> air. The liquid properties were<br />

determined before starting the levitation procedure. The data are shown in Table 5.9.<br />

Table 5.15: Liquid properties <strong>and</strong> levitation size range for a pure catalase solution with 100 mg/ml solid content at<br />

25°C.<br />

Solid<br />

content<br />

100 mg/ml<br />

Surface<br />

tension<br />

[ ]<br />

Density<br />

ρ liquid<br />

3<br />

mN/m [ ]<br />

Maximum<br />

diameter<br />

Maximum<br />

volume<br />

Minimum<br />

diameter<br />

Minimum<br />

volume<br />

g/cm [ mm ] [ µl ] [ µm ] [ µl ]<br />

Catalase 48.6 1.028 4.33 42.4 15 1.77⋅10 -6<br />

Figure 5.113 presents the liquid density <strong>of</strong> catalase solutions in a concentration range from<br />

0.05 mmol/l to 1.0 mmol/l with different amount <strong>of</strong> trehalose.<br />

Liquid density [g/cm 3 ]<br />

1.08<br />

1.07<br />

1.06<br />

1.05<br />

1.04<br />

1.03<br />

1.02<br />

1.01<br />

1.00<br />

Pure catalase solutions at 25°C<br />

in 50 mM tris buffer pH 8.0<br />

50 mg/ml<br />

25 mg/ml<br />

12,5 mg/ml<br />

100 mg/ml<br />

250 mg/ml<br />

0.99<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

Catalase concentration [mmol/l]<br />

(a)<br />

0 100 200 300 400<br />

Figure 5.113: Liquid density <strong>of</strong> catalase solution at 25°C. Catalase was dissolved in 50 mM tris buffer at pH 8.0.<br />

(a) Liquid density <strong>of</strong> pure catalase solutions as a function <strong>of</strong> molar concentration. (b) Liquid density <strong>of</strong> catalasetrehalose<br />

mixtures as a function <strong>of</strong> trehalose content for different molar catalase concentrations.<br />

Liquid density [g/cm 3 ]<br />

1.18<br />

1.16<br />

1.14<br />

1.12<br />

1.10<br />

1.08<br />

1.06<br />

1.04<br />

1.02<br />

1.00<br />

(b)<br />

Catalase-trehalose solutions at 25°C<br />

in 50 mM tris buffer pH 8.0<br />

Trehalose content [mg/ml]<br />

Catalase 0.05 mmol/l<br />

Catalase 0.1 mmol/l<br />

Catalase 0.2 mmol/l<br />

Catalase 0.4 mmol/l<br />

Catalase 1.0 mmol/l


204 RESULTS AND DISCUSSION<br />

Data analysis <strong>of</strong> catalase solution experiments<br />

The data analysis <strong>of</strong> the pure catalase <strong>droplet</strong>s / particles is different to the pure sugar solutions. The<br />

evaporation process can be divided into two different stages. The constant rate, where the<br />

evaporation rate is determined using the decrease in <strong>droplet</strong> volume with time [Frohn 2000; Kastner<br />

et al. 2001] <strong>and</strong> the falling rate, where the evaporation rate is calculated using the rise <strong>of</strong> the centre<br />

<strong>of</strong> mass <strong>of</strong> the particle towards its adjacent upper pressure node [Kastner et al. 2001]. The raw data<br />

for a levitation experiment <strong>of</strong> a pure catalase solution with a solid content <strong>of</strong> 100 mg/ml at 60°C <strong>and</strong><br />

5% relative humidity in still air can be seen in Figure 5.114. The behaviour <strong>of</strong> the pure catalase<br />

solutions is as with the pure bSA solutions with the same initial concentration. After the formation<br />

<strong>of</strong> a solid crust at the <strong>droplet</strong> surface, the particles start oscillating what leads to difficulties in<br />

determination <strong>of</strong> the evaporation rate in the second <strong>drying</strong> stage or even to a drop-out <strong>of</strong> the particle<br />

out <strong>of</strong> the st<strong>and</strong>ing acoustic wave. The disturbed particle position can be seen from fluctuation in<br />

the curve <strong>of</strong> the aspect ratio versus <strong>drying</strong> time shortly after the critical point (Figure 5.114). Even a<br />

breakage <strong>of</strong> the levitated particles into two parts can be observed during the experiments. The pure<br />

catalase results in this work are calculated from the few experiments with stable <strong>droplet</strong> / particle<br />

levitation. There are substantial size changes in horizontal <strong>and</strong> vertical diameter after the critical<br />

point. The critical point itself is determined using the increase in aspect ratio <strong>of</strong> the<br />

<strong>droplet</strong> / particle. As discussed in chapter 3.2.3, size changes <strong>of</strong> the levitated <strong>droplet</strong> or particle lead<br />

to changes in the values <strong>of</strong> the effective SPL. Using the method by Kastner et al. [2001] to<br />

determine the evaporation rate, a constant effective SPL in the second <strong>drying</strong> stage is necessary.<br />

The difference <strong>of</strong> the effective SPL at the critical point <strong>and</strong> at the end <strong>of</strong> the size change was<br />

maximum 0.17 dB. The effective SPL is calculated using the method by Yarin et al. [1998]. The<br />

small difference in SPL is neglected when calculating the evaporation rate in the second <strong>drying</strong><br />

stage. Particles resulting form the <strong>drying</strong> <strong>of</strong> pure catalase solution are shown in Figure 5.115. It<br />

seems that the particle formation <strong>of</strong> catalase is more strongly influenced by the magnitude <strong>of</strong> the<br />

effective SPL than the particle formation <strong>of</strong> any other <strong>of</strong> the tested substance. A high effective SPL<br />

<strong>of</strong> 166.5 dB leads to very flat particles with a donut like shape (Figure 5.115 a), whereas an<br />

effective SPL <strong>of</strong> approximately 164.6 dB leads to particles with a more spherical shape but with a<br />

hollow interior (Figure 5.115 b-d). The more spherical particles are almost similar to the pure bSA<br />

particles with a slightly impressed particle top. They show a hole in their top side <strong>and</strong> an<br />

undamaged bottom side. When destroying the final particles a hollow interior was found.


CHAPTER 5 RESULTS AND DISCUSSION 205<br />

2 ) Aspect ratio<br />

2 / rS 0<br />

( r S<br />

Diameter [µm]<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

Droplet / particle size<br />

Aspect ratio<br />

Vertical position<br />

0.2<br />

-400<br />

0 100 200 300 400 500 600 700<br />

(a)<br />

Constant rate Falling rate<br />

Evaporation time [s]<br />

1800 Horizontal diameter<br />

Vertical diameter<br />

1600<br />

Aspect ratio<br />

1400<br />

1200<br />

1000<br />

800<br />

Figure 5.114: Raw data <strong>of</strong> a levitated pure catalase solution <strong>droplet</strong> with 100 mg/ml solid content in 50 mM tris buffer<br />

pH 8.0 dried at 60°C <strong>and</strong> 5% relative humidity in still air. The initial <strong>droplet</strong> size was 1.8 µl at an initial effective SPL<br />

<strong>of</strong> 166.5 dB. (a) Aspect ratio, (rS 2 /rS 0 2 ) <strong>and</strong> position as a function <strong>of</strong> evaporation time. (b) Horizontal diameter, vertical<br />

diameter <strong>and</strong> aspect ratio as a function <strong>of</strong> evaporation time.<br />

The surface holes have a mean diameter <strong>of</strong> 45.1 ± 4.3% (n = 3) <strong>of</strong> the horizontal particle diameter.<br />

The solid crust has a thickness <strong>of</strong> approximately 18.3% <strong>of</strong> the final particle radius. Assuming a<br />

spherical particle with a radius <strong>of</strong> 500 µm the overall volume would be 0.524 µl with a void volume<br />

<strong>of</strong> 0.286 µl. Neglecting any holes in the surface, the surface area <strong>of</strong> the internal void sphere would<br />

be calculated to 2.10 mm 2 , the outer surface area <strong>of</strong> the spherical particle to 3.14 mm 2 . A<br />

comparison to the spray-dried pure catalase product dried in the Buchi Mini Spray-Dryer is<br />

presented in Figure 5.116 <strong>and</strong> shows very good correlation. The spray-dried particles seem to have<br />

100<br />

0<br />

-100<br />

-200<br />

-300<br />

600<br />

1.0<br />

0 100 200 300 400 500 600 700<br />

(b)<br />

Evaporation time [s]<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

Distance to adjacent pressure node [µm]<br />

Aspect ratio [-]


206 RESULTS AND DISCUSSION<br />

impressed surfaces with one ore multiple holes in it, too. Additionally, undamaged particles as well<br />

as particle fragments can be found.<br />

(a) (b)<br />

(c) (d)<br />

Figure 5.115: SEM pictures <strong>of</strong> pure catalase particles with 100 mg/ml solid content in 50 mM tris buffer pH 8.0 dried<br />

at 60°C <strong>and</strong> 5% relative humidity in still air. (a) Pure catalase particle levitated at an initial effective SPL <strong>of</strong> 166.5 dB.<br />

(b-d) Pure catalase particles levitated at an initial effective SPL <strong>of</strong> 164.6 dB. Shown are pictures with a magnification <strong>of</strong><br />

75-times (a), 45-times (b), 80-times (c) <strong>and</strong> 100-times (d).<br />

Figure 5.116: SEM pictures <strong>of</strong> spray-dried pure catalase<br />

particles with 100 mg/ml solid content in 50 mM tris<br />

buffer pH 8.0. Tinlet = 100°C, Toutlet = 60°C, liquid feed<br />

3 ml/min. The picture shown has a magnification <strong>of</strong><br />

3000-times.


CHAPTER 5 RESULTS AND DISCUSSION 207<br />

Due to the similar behaviour <strong>of</strong> the pure catalase solution <strong>droplet</strong> during the <strong>drying</strong> process in the<br />

acoustic levitator compared to the pure bSA solution <strong>and</strong> the similar way <strong>of</strong> particle formation <strong>and</strong><br />

surface morphology, the plot <strong>of</strong> the evaporation rate as a function <strong>of</strong> time is expected to be similar<br />

to that <strong>of</strong> pure bSA. The obtained results proved this assumption <strong>and</strong> are presented in Figure 5.117.<br />

With increasing relative humidity <strong>of</strong> the <strong>drying</strong> air, the values <strong>of</strong> the evaporation rate in the first<br />

<strong>drying</strong> stage decrease from 0.93 µg/s/mm 2 at 5% humidity to 0.41 µg/s/mm 2 at 40% humidity. In<br />

return, the duration <strong>of</strong> the first <strong>drying</strong> stage increases leading to a shift <strong>of</strong> the critical point. After a<br />

first decrease <strong>of</strong> the evaporation rate in the falling rate, an increase in the evaporation rate can be<br />

determined resulting in values almost as large as at the end <strong>of</strong> the constant rate period, equal to<br />

bSA. This increase in evaporation rate is synchronous with the formation <strong>of</strong> the hole in the surface<br />

on the top side <strong>of</strong> the particle. After the maximum in the second <strong>drying</strong> stage is reached, the<br />

evaporation rate decreases to very low values. The decline in evaporation rate is shallower at 40%<br />

relative humidity <strong>of</strong> the <strong>drying</strong> air than at 5%.<br />

Evaporation rate [µg/s/mm 2 ]<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Critical point<br />

0.0<br />

[Drying temp. 60°C]<br />

0 200 400 600 800 1000 1200 1400<br />

2 2<br />

t / r [s/mm ]<br />

S 0<br />

Rel. humidity 5%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

Evaporation rate<br />

increase in falling<br />

rate period<br />

Figure 5.117: Evaporation rate in mass per time <strong>and</strong> surface area <strong>of</strong> pure catalase solutions as a function <strong>of</strong> (t /rS 0 2 ) at<br />

60°C <strong>and</strong> different relative humidity in still air. The initial <strong>droplet</strong> volume <strong>of</strong> the experiments was between 1.5 <strong>and</strong><br />

2.0 ml at an initial effective SPL <strong>of</strong> 164.6 dB to 166.5 dB.<br />

A mixture series <strong>of</strong> catalase <strong>and</strong> trehalose is prepared in 50 mM tris buffer at pH 8.0. The overall<br />

solid content <strong>of</strong> 100 mg/ml is kept constant. The weight ratio <strong>and</strong> the molar ratio <strong>of</strong> the different<br />

catalase-trehalose mixtures are listed in Table 5.16.


208 RESULTS AND DISCUSSION<br />

Table 5.16: Weight ratio <strong>and</strong> molar ratio <strong>of</strong> mixtures with different content <strong>of</strong> catalase <strong>and</strong> trehalose.<br />

Catalase-trehalose [ mg/ml ]<br />

Weight ratio<br />

80:20<br />

4:1<br />

60:40<br />

3:2<br />

40:60<br />

2:3<br />

20:80<br />

Molar ratio 1:183 1:487 1:1096 1:2921<br />

The results <strong>of</strong> the evaporation rate as a function <strong>of</strong> time for the mixture series at a <strong>drying</strong> air <strong>of</strong><br />

60°C <strong>and</strong> 5% relative humidity are presented in Figure 5.118. An analysis <strong>of</strong> a diagram containing<br />

the plots <strong>of</strong> the evaporation rates <strong>of</strong> all tested catalase-trehalose mixtures is difficult. Therefore,<br />

only the pure catalase, the catalase-trehalose (6:4) <strong>and</strong> the catalase-trehalose (2:8) mixture at 5%<br />

<strong>and</strong> 20% humidity are plotted. Due to the longer evaporation time at increased humidity, the events<br />

<strong>of</strong> the <strong>drying</strong> process are protracted <strong>and</strong> better detectable. A tendency <strong>of</strong> decreasing values <strong>of</strong> the<br />

evaporation rate in the first <strong>drying</strong> stage with increasing trehalose content can be seen. An increase<br />

in trehalose leads to an increase in the overall amount <strong>of</strong> dissolved molecules. As with bSA <strong>and</strong><br />

bSA-trehalose (1:1) mixtures, the duration <strong>of</strong> the first <strong>drying</strong> stage increases with increasing<br />

trehalose content. The catalase-trehalose (6:4) mixture at 60°C <strong>and</strong> 5% humidity shows an outlier<br />

behaviour compared to all other results. The shift <strong>of</strong> the critical point due to the increasing trehalose<br />

content becomes more obvious when analysing the results at increased <strong>drying</strong> air humidity. As with<br />

the pure catalase solutions, the increase <strong>of</strong> the evaporation rate in the second <strong>drying</strong> stage due to the<br />

formation <strong>of</strong> holes in the particle surface can still be seen for the catalase-trehalose (8:2) mixture. A<br />

further increase in trehalose content leads to disappearance <strong>of</strong> this behaviour <strong>and</strong> to a continuously<br />

falling evaporation rate in the second <strong>drying</strong> stage without maximum. The SEM pictures <strong>of</strong> the final<br />

particles confirm these results. As the catalase-trehalose (8:2) mixture particles still show holes in<br />

their surface, the final particles <strong>of</strong> the catalase-trehalose (6:4) solution show an undamaged surface<br />

with a massive particle interior. If the trehalose content is increased further, the particle morphology<br />

adapts to that <strong>of</strong> the pure trehalose particles shown in Figure 5.67.<br />

Similar to the effect <strong>of</strong> increasing <strong>drying</strong> air humidity seen in Figure 5.117, increasing<br />

trehalose content leads to a shallower decline <strong>of</strong> the evaporation rate in the falling rate period. Only<br />

marginal differences in the evaporation rate curves <strong>of</strong> the mixtures with a catalase-trehalose ratio <strong>of</strong><br />

4:6 <strong>and</strong> 2:8 <strong>and</strong> the pure trehalose curve can be seen (not shown).<br />

1:4


CHAPTER 5 RESULTS AND DISCUSSION 209<br />

Evaporation rate [µg/s/mm 2 ]<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

Drying temp. 60°C<br />

Rel. humidity 5%<br />

0 250 500 750 1000 1250 1500<br />

Figure 5.118: Evaporation rate <strong>of</strong> pure catalase <strong>and</strong> catalase-trehalose mixtures in 50 mM tris buffer pH 8.0 as a<br />

function <strong>of</strong> (t /rS 0 2 ) at 60°C <strong>and</strong> different relative humidity in still air. The initial <strong>droplet</strong> volume <strong>of</strong> the experiments was<br />

between 1.5 <strong>and</strong> 2.0 µl at an initial effective SPL <strong>of</strong> 164.6 <strong>and</strong> 167.2 dB. (a) Drying conditions 60°C <strong>and</strong> 5% relative<br />

humidity in still air. (b) Drying conditions 60°C <strong>and</strong> 20% relative humidity in still air.<br />

The values <strong>of</strong> the catalase-trehalose (2:8) curve in Figure 5.118 are much lower <strong>and</strong> the evaporation<br />

time is much longer than for pure trehalose shown in Figure 5.63 <strong>and</strong> Figure 5.64. This is caused by<br />

different solvents. In the experiments <strong>of</strong> Figure 5.118 all <strong>formulations</strong> are dissolved in 50 mM tris<br />

buffer at pH 8.0 due to the poor solubility <strong>of</strong> catalase in water, whereas in Figure 5.63 <strong>and</strong> Figure<br />

5.64 trehalose is in aqueous solution without any additives. Due to the increase in dissolved<br />

molecules caused by the buffer, the evaporation rate values <strong>of</strong> Figure 5.118 are lower <strong>and</strong> the<br />

evaporation time is longer. The values <strong>of</strong> the evaporation rate at later points in second <strong>drying</strong> stage<br />

are larger with increasing trehalose content. A comparison <strong>of</strong> the AUCs <strong>of</strong> the different experiments<br />

show the largest values for the pure trehalose <strong>and</strong> the lowest values for the pure catalase, implying a<br />

more efficient <strong>drying</strong> with increasing trehalose content. Because a determination <strong>of</strong> the residual<br />

water content <strong>of</strong> the single particles was not possible, the pure substances <strong>and</strong> the <strong>formulations</strong><br />

presented in Table 5.16 are spray-dried in the Buchi Mini Spray-Dryer at T = 100° C ,<br />

T = 60° C <strong>and</strong> a liquid feed <strong>of</strong> 3 ml/min <strong>and</strong> the residual water content <strong>of</strong> the product was<br />

outlet<br />

Critical point<br />

Rupture <strong>of</strong> the<br />

pure catalase<br />

particle<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

Catalase-trehalose (10:0)<br />

Catalase-trehalose (6:4)<br />

Catalase-trehalose (2:8)<br />

Larger values<br />

with increasing<br />

trehalose content<br />

1.0<br />

Critical point<br />

Catalase-trehalose (10:0)<br />

Catalase-trehalose (6:4)<br />

0.9<br />

Catalase-trehalose (2:8)<br />

0.8<br />

0.7<br />

0.6<br />

Rupture <strong>of</strong> the<br />

pure catalase particle<br />

0.5<br />

0.4<br />

Larger values<br />

with increasing<br />

0.3<br />

0.2<br />

trehalose content<br />

0.1 Drying temp. 60°C<br />

0.0<br />

Rel. humidity 20%<br />

0 250 500 750 1000 1250 1500<br />

analysed via Karl-Fischer titration. The results are shown in Figure 5.119. The relative humidity <strong>of</strong><br />

the spray-dryer environment is additionally plotted. A clear decrease <strong>of</strong> the residual moisture<br />

content with increasing trehalose <strong>and</strong> decreasing catalase content in the formulation cannot be seen.<br />

All values are within a range <strong>of</strong> 0.3%. The lowest residual moisture content is found for the pure<br />

trehalose powder. However, the laboratory humidity which also has an substantial influence on the<br />

residual moisture <strong>of</strong> spray-dried products [Fitzner 2003; Masters 1991] has been lowest in this<br />

Evaporation rate [µg/s/mm 2 ]<br />

2 2<br />

( t / r ) [s/mm ]<br />

S 0<br />

inlet


210 RESULTS AND DISCUSSION<br />

experiment. Further investigations with spray-<strong>drying</strong> in conditioned environment or with<br />

conditioned <strong>drying</strong> air has to be done to examine influences <strong>of</strong> formulation variables on the residual<br />

moisture more detailed. References on the moisture content <strong>of</strong> spray-dried catalase-trehalose<br />

<strong>formulations</strong> could not be found. Tzannis <strong>and</strong> Prestrelski [1999] investigated the moisture effects <strong>of</strong><br />

sucrose when spray-<strong>drying</strong> trypsinogen. They found decreasing residual moisture values with<br />

decreasing sucrose content. Their optimum formulation in terms <strong>of</strong> lowest moisture content had<br />

equal mass <strong>of</strong> sucrose <strong>and</strong> trypsinogen.<br />

Residual moisture content [%]<br />

6.0<br />

5.8<br />

5.6<br />

5.4<br />

5.2<br />

5.0<br />

4.8<br />

(n = 3)<br />

lab humidity<br />

0 20 40 60 80 100<br />

Trehalose content in formulation [% solid]<br />

Figure 5.119: Residual water content <strong>of</strong> spray-dried catalase-trehalose solutions in 50 mM tris buffer at pH 8.0 as a<br />

function <strong>of</strong> trehalose solid content <strong>of</strong> formulation. The residual water content was determined via Karl-Fischer titration.<br />

The final particle size <strong>and</strong> density <strong>of</strong> the catalase-trehalose <strong>formulations</strong> <strong>and</strong> the pure solutions are<br />

presented in Figure 5.120. As with bSA <strong>and</strong> its mixtures with trehalose, the final particle size<br />

relative to the initial <strong>droplet</strong> size decreases with increasing trehalose content <strong>and</strong> increasing<br />

humidity <strong>of</strong> the <strong>drying</strong> air. In return, the particles get denser with increasing trehalose content <strong>and</strong><br />

increasing <strong>drying</strong> air humidity. This behaviour is caused by the different length <strong>of</strong> the first <strong>drying</strong><br />

stage <strong>and</strong> the shift <strong>of</strong> the critical point in Figure 5.118. The longer the first <strong>drying</strong> stage <strong>and</strong> the later<br />

the formation <strong>of</strong> the solid crust at the <strong>droplet</strong> surface, the smaller <strong>and</strong> denser the final particles<br />

become. The effect has been seen for all tested substances <strong>and</strong> <strong>formulations</strong>. With increasing<br />

trehalose content, an increase in duration <strong>of</strong> the first <strong>drying</strong> stage <strong>and</strong> in the amount <strong>of</strong> water<br />

evaporated can be detected. The data are listed in Table 5.17. Only 61.9% <strong>of</strong> solvent is evaporated<br />

in the first <strong>drying</strong> stage from the pure catalase solution <strong>droplet</strong>s. This portion increases to 85.9%<br />

solvent for the pure trehalose solution. The values at 20% relative humidity <strong>of</strong> the <strong>drying</strong> air are in<br />

average 1.7% larger than at 5% humidity. A decrease in the percentage <strong>of</strong> solvent evaporated from<br />

28.0<br />

26.0<br />

24.0<br />

22.0<br />

20.0<br />

18.0<br />

16.0<br />

Relative humidity <strong>of</strong> spray-<strong>drying</strong> environment


CHAPTER 5 RESULTS AND DISCUSSION 211<br />

the pure catalase solution in the constant rate period with increasing humidity, as seen from the bSA<br />

results, cannot be detected. Both <strong>of</strong> the pure <strong>protein</strong> solutions show approximately the same values.<br />

The values <strong>of</strong> the pure trehalose solution in 50 mM tris buffer at pH 8.0 differ only marginally form<br />

the aqueous trehalose solution.<br />

Particle size relative to initial <strong>droplet</strong> size [%]<br />

65<br />

60<br />

55<br />

50<br />

45<br />

40<br />

(a)<br />

Figure 5.120: Particle size <strong>and</strong> final particle density as a function <strong>of</strong> trehalose solid content in the formulation at 60°C<br />

<strong>and</strong> different relative humidity in still air. The ignition <strong>droplet</strong> volume <strong>of</strong> the experiments was between 1.5 µl <strong>and</strong> 2.0 µl<br />

at an initial effective SPL between 164.6 dB to 167.2 dB. (a) Particle size relative to initial <strong>droplet</strong> size versus trehalose<br />

content in formulation. (b) Final particle density versus trehalose content in formulation.<br />

Table 5.17: Percentage <strong>of</strong> evaporated solvent in the first <strong>drying</strong> stage depending on a change <strong>of</strong> trehalose content <strong>and</strong><br />

relative humidity <strong>of</strong> the <strong>drying</strong> air. The experiments were performed at 60°C in still air.<br />

Relative humidity Percentage <strong>of</strong> evaporated solvent in the first <strong>drying</strong> stage (n=3)<br />

<strong>of</strong> <strong>drying</strong> air catalase : trehalose ratio (50 mM tris buffer pH 8.0)<br />

[ % ]<br />

Rel. humidity 5%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

0 20 40 60 80 100<br />

10:0<br />

Drying temp. 60°C<br />

Trehalose content in formulation [% solid]<br />

8:2<br />

6:4<br />

0 20 40 60 80 100<br />

4:6<br />

2:8<br />

0:10<br />

5 61.9 ± 2.8 64.6 ± 3.0 71.2 ± 2.9 76.7 ± 1.8 85.5 ± 2.8 85.9 ± 0.7<br />

20 63.2 ± 4.0 68.4 ± 0.7 72.3 ± 2.1 78.1 ± 1.9 87.1 ± 1.8 87.0 ± 2.8<br />

In Figure 5.121 <strong>and</strong> Figure 5.122 the final particles <strong>of</strong> the catalase-trehalose (6:4) mixture from<br />

spray-<strong>drying</strong> <strong>and</strong> levitation can be seen. With increasing trehalose content the holes <strong>and</strong> impressions<br />

in the particle surface get less <strong>and</strong> the particles become more spherical. Almost ideal spheres are<br />

obtained for pure trehalose, whereas the catalase-trehalose (4:6) <strong>and</strong> (2:8) formulation show the<br />

raisin like morphology <strong>of</strong>ten mentioned in literature [Maa et al. 1997]. The levitated catalasetrehalose<br />

(6:4) particles have a broken surface similar to the bSA-trehalose (1:1) mixture. A folded<br />

surface morphology on the top side <strong>of</strong> the particle cannot be seen in that degree as with bSAtrehalose<br />

solutions. Destruction showed a massive internal particle structure volume.<br />

Particle density [g/cm 3 ]<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

(b)<br />

Rel. humidity 5%<br />

Rel. humidity 20%<br />

Rel. humidity 40%<br />

Drying temp. 60°C<br />

Trehalose content in formulation [% solid]


212 RESULTS AND DISCUSSION<br />

(a) (b) (c)<br />

(d) (e) (f)<br />

Figure 5.121: SEM pictures <strong>of</strong> spray-dried pure catalase-trehalose solutions with 100 mg/ml solid content from a Buchi<br />

Mini Spray-Dryer B-191 with Tinlet=100°C, Toutlet=60°C <strong>and</strong> a liquid feed <strong>of</strong> 3 ml/min. All pictures have a 3500-times<br />

magnification. (a) Pure catalase. (b) Catalase-trehalose (8:2). (c) Catalase-trehalose (6:4). (d) Catalase-trehalose (4:6).<br />

(e) Catalase-trehalose (2:8). (f) Pure trehalose.<br />

(a) (b)<br />

(c) (d)<br />

Figure 5.122: SEM pictures <strong>of</strong> final particles from levitated catalase-trehalose (6:4) solutions with 100 mg/ml solid<br />

content in 50 mM tris buffer pH 8.0 dried at 60°C <strong>and</strong> 5% relative humidity in still air. Shown are pictures with a<br />

magnification <strong>of</strong> 100-times (a), 1000-times (b), 100-times (c) <strong>and</strong> 200-times (d).


CHAPTER 5 RESULTS AND DISCUSSION 213<br />

Interfacial behaviour <strong>and</strong> surface excess <strong>of</strong> catalase<br />

The interfacial behaviour <strong>and</strong> the surface excess <strong>of</strong> catalase are tested to investigate the earlier<br />

critical point in the <strong>drying</strong> process <strong>of</strong> the pure catalase solution compared to the mixtures <strong>of</strong> catalase<br />

<strong>and</strong> trehalose. The pure 50 mM tris buffer at pH 8.0 has no influence on the surface tension <strong>of</strong><br />

water. Due to the much larger molecular weight <strong>of</strong> catalase the solution molarity chosen is in a<br />

range between 0.01 mmol/l <strong>and</strong> 1.0 mmol/l. A solution with a molar catalase concentration <strong>of</strong><br />

0.4 mmol/l is equal to 100 mg/ml. Figure 5.123 a shows the air / water interfacial tension versus log<br />

time in seconds obtained for pure catalase solutions at 25°C using a bubble pressure tensiometer.<br />

Figure 5.123 b presents the plot <strong>of</strong> Π () t / Π ∞ versus t according to the short-time approximation<br />

<strong>of</strong> the Ward <strong>and</strong> Tordai Equation. After an interfacial lifetime <strong>of</strong> only 1 ms ( log = −3.<br />

0)<br />

t the values<br />

for the dynamic interfacial tension are lower than 72.8 mN/m measured for the pure buffer solution.<br />

For low catalase concentration (0.05 mmol) the value after 1 ms lies only marginally below pure<br />

water. For the 0.01 mmol/l solution no difference to water at 1 ms can be detected. With increasing<br />

solid concentration <strong>and</strong> increasing interfacial lifetime the interfacial tension decreases. For the<br />

1.0 mmol/l solution a surface tension <strong>of</strong> 47.5 mN/m is found after an interfacial tension <strong>of</strong> 1 s.<br />

Dynamic interfacial tension [mN/m]<br />

75.0<br />

70.0<br />

65.0<br />

60.0<br />

55.0<br />

50.0<br />

45.0<br />

(a)<br />

Catalase solutions in<br />

50mM tris buffer pH 8.0 at 25°C<br />

0.05 mmol<br />

0.1 mmol<br />

0.2 mmol<br />

0.4 mmol<br />

1.0 mmol<br />

-3 -2 -1 0 1 2<br />

log time in seconds<br />

0.01 mmol<br />

0.2<br />

0.0<br />

Catalase solutions in<br />

50mM tris buffer pH 8.0 at 25°C<br />

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0<br />

Figure 5.123: (a) Dynamic air / water interfacial tension as a function <strong>of</strong> time <strong>of</strong> different catalase solutions in 50 mM<br />

tris buffer at pH 8.0 with concentration <strong>of</strong> minimum 0.01 mmol to a maximum <strong>of</strong> 1.0 mmol/l measured at 25°C. (b)<br />

Π (t) / Π ∞ as a function <strong>of</strong> time ½ <strong>of</strong> catalase solutions in 50mM tris buffer pH 8.0 with different molar concentration at<br />

25°C measured with bubble pressure tensiometry.<br />

In Figure 5.123 b Π () t / Π ∞ is plotted versus t for the catalase solutions for diffusion–controlled<br />

Π t / Π ∞<br />

adsorption. The values <strong>of</strong> Π ∞ were determined by equilibrium ring tensiometry after an<br />

equilibration time <strong>of</strong> 3 hours <strong>and</strong> corrected using the method by Harkins <strong>and</strong> Jordan [1930]. The<br />

results are shown in Table 5.18. Clarkson et al. [1999] found a surface tension <strong>of</strong> 57.0 mN/m for a<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

(b)<br />

Time 1/2 [s 1/2 ]<br />

1.0 mmol<br />

0.4 mmol<br />

0.2 mmol<br />

0.1 mmol<br />

0.05 mmol<br />

0.01 mmol


214 RESULTS AND DISCUSSION<br />

0.01 mmol/l catalase solution in 0.1 mM phosphate buffer at pH 7.5 using ring tensiometry. For a<br />

0.04 mmol/l catalase solution in the same solvent they found a value <strong>of</strong> 54.5mN/m.<br />

Table 5.18: Values for equilibrium interfacial tension after 3 hours for catalase solutions in 50 mM tris buffer pH 8.0 at<br />

25°C ambient temperature using ring tensiometry.<br />

Concentration <strong>of</strong><br />

catalase solution<br />

Equilibrium<br />

interfacial tension<br />

[ mmol ] 0.01 0.05 0.1 0.2 0.4 1.0<br />

[ mN/m ] 55.0 51.1 50.4 48.6 47.0 45.7<br />

The plot <strong>of</strong> the short-time approximation in Figure 5.123 b shows biphasic behaviour. In contrast to<br />

bSA, the catalase solutions do not show a linear shape <strong>of</strong> the plot in the fist part <strong>and</strong> a clear break to<br />

separate the two phases. It seems that the break point is equal to bSA at approximately 500 ms<br />

( t = 0.<br />

7 ). The slopes <strong>of</strong> the initial section for t ≤ 500 ms increase with higher <strong>protein</strong><br />

concentration. The slopes in the second linear section for t ≥ 500 ms are equal for all catalase<br />

concentrations. Π () t / Π ∞ increases much slower in the second phase than in the initial one. Thus,<br />

the diffusional adsorption is evidently reduced in rate at the later points in the curves <strong>of</strong> Figure<br />

5.123 b. According to Fuhrling [2004],it seems that a barrier to adsorption is formed.<br />

Dynamic interfacial tension [mN/m]<br />

75.0<br />

72.5<br />

70.0<br />

67.5<br />

65.0<br />

62.5<br />

60.0<br />

57.5<br />

55.0<br />

52.5<br />

50.0<br />

47.5<br />

Catalase solutions in<br />

50 mM tris buffer at pH 8.0<br />

10 ms<br />

100 ms<br />

1000 ms<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

Catalase concentration [mmol/l]<br />

Figure 5.124: Dynamic air / water interfacial tension as a function <strong>of</strong> solution molarity <strong>of</strong> catalase solutions in 50 mM<br />

tris buffer at pH 8.0 at different times.<br />

1 ms


CHAPTER 5 RESULTS AND DISCUSSION 215<br />

Figure 5.124 presents the plot <strong>of</strong> the dynamic interfacial tension versus the total molarity <strong>of</strong> the<br />

catalase solution at different lifetime. The values decrease with increasing concentration <strong>of</strong> catalase<br />

<strong>and</strong> increasing lifetime. None <strong>of</strong> the curves shows tendency <strong>of</strong> a levelling-out. The interfacial<br />

tension <strong>of</strong> all curves decreases further with increasing solution molarity <strong>of</strong> catalase.<br />

The total interfacial excess concentration <strong>of</strong> catalase is calculated using the Gibbs Equation<br />

5.14 [Hiemenz 1986]. Figure 5.125 a shows the calculated surface excess concentration plotted<br />

versus solution molarity for the catalase solutions at different interfacial lifetime. As expected, an<br />

increase in Γ (t)<br />

is found with increasing <strong>protein</strong> concentration. The formation <strong>of</strong> a plateau with<br />

increasing catalase concentration cannot be seen. After 1 s interfacial lifetime the surface excess<br />

values increase to 0.00226 mmol/m 2 for a 0.04 mmol/l catalase solution <strong>and</strong> to 0.00268 mmol/m 2<br />

for a 0.1 mmol/l catalase solution. These values are equivalent to 565 mg/m 2 – 670 mg/m 2 . In<br />

comparison, a pure bSA solution with a concentration <strong>of</strong> 0.1 mmol/l has a surface excess<br />

concentration <strong>of</strong> 0.00113 mmol/m 2 equivalent to 75.1 mg/m 2 at an interfacial lifetime <strong>of</strong> 1 s. This<br />

means, that catalase has a more than twice as large surface excess concentration based on the<br />

amount <strong>of</strong> molecules than bSA resulting in an almost nine times larger mass per area unit.<br />

Reference values from literature approving the experimental catalase values could not be found.<br />

Figure 5.125 b presents the surface excess concentration versus interfacial lifetime with different<br />

solid concentration. With increasing catalase content the slope <strong>of</strong> the plots in a time range between<br />

100 ms <strong>and</strong> 1 s increases. A levelling-out in this time range can only be seen for the 0.01 mmol/l<br />

catalase solution.<br />

Surface excess [mmol/m 2 ]<br />

0.0030<br />

Catalase solutions in<br />

50 mM tris buffer pH 8.0<br />

0.0025<br />

0.0020<br />

0.0015<br />

0.0010<br />

0.0005<br />

100 ms<br />

10 ms<br />

1 ms<br />

0.0000<br />

0.0000 0.0002 0.0004 0.0006 0.0008 0.0010 0.0012<br />

(a)<br />

Solution molarity [mmo/m 3 ]<br />

1000 ms<br />

0.0030<br />

0.0025<br />

0.0020<br />

0.0015<br />

0.0010<br />

0.0005<br />

0.0000<br />

0.4 mmol<br />

0.2 mmol<br />

0.1 mmol<br />

0.05 mmol<br />

0.01 mmol<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

Figure 5.125: (a) Surface excess concentration Γ (t) as a function <strong>of</strong> catalase solution molarity in 50 mM tris buffer at<br />

pH 8.0 at 25°C for different lifetimes. (b) Surface excess concentration as a function <strong>of</strong> time <strong>of</strong> catalase solutions in<br />

50 mM tris buffer at pH 8.0 at various concentration at 25°C.<br />

Surface excess [mmol/m 2 ]<br />

(b)<br />

Catalase solutions in<br />

50 mM tris buffer pH 8.0<br />

Time [s]<br />

1.0 mmol


216 RESULTS AND DISCUSSION<br />

Influence <strong>of</strong> trehalose on the interfacial behaviour <strong>of</strong> catalase<br />

To investigate the influence <strong>of</strong> trehalose on the interfacial tension <strong>of</strong> catalase solutions, the above<br />

molar concentration <strong>of</strong> catalase were prepared with different amounts <strong>of</strong> trehalose. The density <strong>of</strong> a<br />

0.04 mmol/l catalase solution increases from 1.03 g/cm 3 without trehalose to 1.09 g/cm 3 with<br />

200 mg/ml trehalose. Figure 5.126 shows the influence <strong>of</strong> trehalose on the dynamic interfacial<br />

tension <strong>of</strong> catalase solutions with a catalase concentration <strong>of</strong> 0.04 mmol/l equivalent to 100 mg/ml<br />

as a function <strong>of</strong> log time. The surface tension at an interfacial lifetime increases from 55.9 mN/m<br />

for the pure catalase to 59.8 mN/m for catalase with 200 mg/ml trehalose <strong>and</strong> to 64.4 mN/m with<br />

400 mg/ml trehalose. Figure 5.127 presents the dynamic interfacial tension <strong>of</strong> catalase solutions<br />

with different amount <strong>of</strong> trehalose as a function <strong>of</strong> solution molarity (a, b) <strong>and</strong> interfacial lifetime<br />

(c, d). The addition <strong>of</strong> 400 mg/ml trehalose to a 1.0 mmol/l catalase solution increases the dynamic<br />

interfacial tension by 7.47 mN/m at an interfacial lifetime <strong>of</strong> 100 ms <strong>and</strong> by 8.27 mN/m at an<br />

interfacial lifetime <strong>of</strong> 1000 ms. With increasing solution molarity a decrease in the dynamic<br />

interfacial tension is determined for all tested <strong>formulations</strong>.<br />

Dynamic interfacial tension [mN/m]<br />

75.0<br />

70.0<br />

65.0<br />

60.0<br />

55.0<br />

50.0<br />

at 25°C in<br />

50 mM tris buffer at pH 8.0<br />

-3 -2 -1 0 1<br />

log time in seconds<br />

Catalase 0,40 mmol (pure)<br />

Catalase 0,40 mmol + 100mg/ml Trehalose<br />

Catalase 0,40 mmol + 200mg/ml Trehalose<br />

Catalase 0,40 mmol + 400mg/ml Trehalose<br />

Figure 5.126: Dynamic interfacial tension <strong>of</strong> 0.04 mmol/l catalase solutions in 50 mM tris buffer at pH 8.0 with<br />

different trehalose content as a function <strong>of</strong> log time in seconds at 25°C.<br />

Increasing trehalose content leads to a slight increase in the slope d γ ( t)<br />

/ dc . The plot <strong>of</strong> the<br />

dynamic interfacial tension versus interfacial lifetime does not show a levelling-out at larger<br />

interfacial lifetime for the catalase solutions even at the highest trehalose content <strong>of</strong> 400 mg/ml. It<br />

seems that there is a slight increase in d γ ( t)<br />

/ dt with increasing trehalose concentration in the range


CHAPTER 5 RESULTS AND DISCUSSION 217<br />

between 100 ms <strong>and</strong> 1000 ms. The interactions between <strong><strong>protein</strong>s</strong> <strong>and</strong> trehalose are discussed in<br />

detail in Chapter 5.3.4.<br />

Figure 5.128 presents the surface excess concentration <strong>of</strong> catalase solutions with a<br />

concentration <strong>of</strong> 0.4 mmol/l <strong>and</strong> 1.0 mmol/l <strong>and</strong> different trehalose content versus interfacial<br />

lifetime. Trehalose decreases the surface excess <strong>of</strong> catalase for both solution molarities at interfacial<br />

lifetimes <strong>of</strong> 100 ms <strong>and</strong> 1000 ms. For a 0.4 mmol/l solution the addition <strong>of</strong> 200 mg/ml trehalose<br />

reduces the surface excess <strong>of</strong> pure catalase from 0.0023 mmol/m 2 to 0.0016 mmol/m 2 <strong>and</strong> the<br />

addition <strong>of</strong> 400 mg/ml trehalose results in a surface excess decrease to 0.0015 mmol/m 2 at an<br />

interfacial lifetime <strong>of</strong> 1000 ms. That is equivalent with a decrease in catalase concentration at the<br />

surface from 575 mg/m 2 to 375–400 mg/m 2 .<br />

Dynamic interfacial tension [mN/m]<br />

Dynamic interfacial tension [mN/m]<br />

75.0<br />

70.0<br />

65.0<br />

60.0<br />

55.0<br />

50.0<br />

75.0<br />

70.0<br />

65.0<br />

60.0<br />

55.0<br />

50.0<br />

Catalase + trehalose at 25°C<br />

Interfacial lifetime 100 ms<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4<br />

Catalase solution molarity [mmol/l]<br />

+ 400 mg/ml<br />

+ 200 mg/ml<br />

+ 100 mg/ml<br />

pure<br />

Catalase solution molarity 0.4 mmol/l<br />

Catalase + trehalose at 25°C<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4<br />

Time [s]<br />

+ 400 mg/ml<br />

+ 200 mg/ml<br />

+ 100 mg/ml<br />

pure<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4<br />

Catalase solution molarity [mmol/l]<br />

Figure 5.127: Dynamic interfacial tension <strong>of</strong> catalase solution in 50 mM tris buffer at pH 8.0 with different amount <strong>of</strong><br />

trehalose as a function <strong>of</strong> solution molarity <strong>and</strong> interfacial lifetime. (a) Dynamic interfacial tension versus solution<br />

molarity for an interfacial lifetime <strong>of</strong> 100 ms. (b) Dynamic interfacial tension versus solution molarity for an interfacial<br />

lifetime <strong>of</strong> 1000 ms. (c) Dynamic interfacial tension versus interfacial lifetime for a catalase solution molarity <strong>of</strong><br />

0.4 mmol/l. (d) Dynamic interfacial tension versus interfacial lifetime for a catalase solution molarity <strong>of</strong> 1.0 mmol/l.<br />

Dynamic interfacial tension [mN/m]<br />

Dynamic interfacial tension [mN/m]<br />

75.0<br />

70.0<br />

65.0<br />

60.0<br />

55.0<br />

50.0<br />

75.0<br />

70.0<br />

65.0<br />

60.0<br />

55.0<br />

50.0<br />

Interfacial lifetime 1000 ms<br />

Catalase + trehalose at 25°C<br />

+ 400 mg/ml<br />

+ 200 mg/ml<br />

+ 100 mg/ml<br />

pure<br />

Catalase solution molarity 1.0 mmol/l<br />

Catalase + trehalose at 25°C<br />

+ 400 mg/ml<br />

+ 200 mg/ml<br />

+ 100 mg/ml<br />

pure<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4<br />

Time [s]


218 RESULTS AND DISCUSSION<br />

The plots <strong>of</strong> the surface excess versus interfacial lifetime show almost similar values for the three<br />

different trehalose <strong>formulations</strong>. At a catalase solution molarity <strong>of</strong> 0.4 mmol/l they are within a<br />

range <strong>of</strong> 0.0015 mmol/m 2 to 0.0017 mmol/m 2 . For the 1.0 mmol/l Catalase solution, equivalent to<br />

250 mg/ml, the addition <strong>of</strong> 200 mg/ml or 400 mg/ml trehalose reduces the surface excess <strong>of</strong> catalase<br />

from 0.00268 mmol/m 2 to approximately 0.0017 mmol/m 2 . Therefore the amount <strong>of</strong> catalase at the<br />

surface decreases from 670 mg/m 2 to 425 mg/m 2 . The slope d ( t)<br />

/ dt<br />

Γ <strong>of</strong> the plots between 100 ms<br />

<strong>and</strong> 1000 ms at 0.4 mmol/l catalase solution molarity as well as at 1.0 mmol/l catalase solution<br />

molarity decreases with increasing trehalose content.<br />

Surface excess [mmol/m 2 ]<br />

0.0030<br />

0.0025<br />

0.0020<br />

0.0015<br />

0.0010<br />

0.0005<br />

0.0000<br />

Catalase solution molarity 0.0004 mmol/m 3<br />

Catalase pure<br />

Catalase + 100 mg/ml trehalose<br />

Catalase + 200 mg/ml trehalose<br />

Catalase + 400 mg/ml trehalose<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

Time [s]<br />

Figure 5.128: Surface excess concentration <strong>of</strong> catalase solutions in 50 mM tris buffer at pH 8.0 with different trehalose<br />

content as a function <strong>of</strong> interfacial lifetime at 25°C. (a) Surface excess concentration versus interfacial lifetime for a<br />

0.4mmol/l catalase solution. (b) Surface excess concentration versus interfacial lifetime for a 1.0 mmol/l catalase<br />

solution.<br />

Comparison <strong>of</strong> surface tension <strong>and</strong> surface excess <strong>of</strong> bSA <strong>and</strong> catalase<br />

0.0030 Catalase solution molarity 0.001 mmol/m 3<br />

0.0025<br />

0.0020<br />

0.0015<br />

0.0010<br />

0.0005<br />

0.0000<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

To compare the bSA <strong>and</strong> catalase results only the surface tension <strong>and</strong> the surface excess <strong>of</strong> the<br />

1.0 mmol/l <strong>protein</strong> solutions are presented in Figure 5.130. At an interfacial lifetime <strong>of</strong> 1.0 s the<br />

dynamic interfacial tensions <strong>of</strong> the pure catalase <strong>and</strong> the catalase with 200 mg/ml trehalose have<br />

lower values than the bSA solutions. The slopes d γ / dt <strong>of</strong> both catalase plots between 100 ms <strong>and</strong><br />

1000 ms are smaller than for bSA. The same results are obtained for the surface excess <strong>of</strong> the two<br />

<strong><strong>protein</strong>s</strong>. Catalase <strong>and</strong> the catalase-trehalose formulation show a larger surface excess at any<br />

interfacial lifetime than bSA. The slopes d Γ / dt <strong>of</strong> the plots between 100 ms <strong>and</strong> 1000 ms are<br />

larger for pure catalase than for pure bSA. The same behaviour can be observed for the <strong>protein</strong>-<br />

trehalose <strong>formulations</strong>. The slope d Γ / dt between 100 ms <strong>and</strong> 1000 ms is lager for catalase with<br />

200 mg/ml trehalose than for bSA solution with the same trehalose concentration. The 1.0 mmol/l<br />

Surface excess [mmol/m 2 ]<br />

Catalase pure<br />

Catalase + 100 mg/ml trehalose<br />

Catalase + 200 mg/ml trehalose<br />

Catalase + 400 mg/ml trehalose<br />

Time [s]


CHAPTER 5 RESULTS AND DISCUSSION 219<br />

catalase solution has a surface excess concentration <strong>of</strong> 0.00268 mmol/m 2 at an interfacial lifetime <strong>of</strong><br />

1.0 s, what is equivalent to 670 mg/m 2 . In contrast, bSA has a surface excess concentration <strong>of</strong><br />

0.00113 mmol/m 2 at an interfacial lifetime <strong>of</strong> 1.0 s what is equivalent to 75.1 mg/m 2 . Catalase has a<br />

more than twice as large surface excess concentration based on the amount <strong>of</strong> molecules than bSA<br />

resulting in an almost nine times larger mass per area unit. Although catalase has an approximately<br />

3.8-times larger molecular weight, it shows a more than two times larger surface excess<br />

concentration than bSA at any interfacial lifetime.<br />

Dynamic interfacial tension [mN/m]<br />

75.0<br />

70.0<br />

65.0<br />

60.0<br />

55.0<br />

50.0<br />

bSA pure<br />

bSA + 200 mg/ml trehalose<br />

Catalase pure<br />

Catalase + 200 mg/ml trehalose<br />

Protein solution molarity 1.0 mmol/l<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

Time [s]<br />

Figure 5.129: (a) Dynamic interfacial tension <strong>of</strong> aqueous bSA solutions <strong>and</strong> catalase solutions in 50 mM tris buffer at<br />

pH 8.0 as a function <strong>of</strong> interfacial lifetime for a <strong>protein</strong> concentration <strong>of</strong> 1.0 mmol/l without <strong>and</strong> with 200 mg/ml<br />

trehalose. (b) Surface excess concentration <strong>of</strong> aqueous bSA solutions <strong>and</strong> catalase solutions in 50 mM tris buffer at<br />

pH 8.0 as a function <strong>of</strong> interfacial lifetime for a <strong>protein</strong> concentration <strong>of</strong> 1.0 mmol/l without <strong>and</strong> with 200 mg/ml<br />

trehalose.<br />

Enzyme activity <strong>of</strong> levitated catalase solution <strong>droplet</strong>s<br />

Surface excess [mmol/m 2 ]<br />

0.0030<br />

0.0025<br />

0.0020<br />

0.0015<br />

0.0010<br />

0.0005<br />

0.0000<br />

bSA pure<br />

bSA + 200 mg/ml trehalose<br />

Catalase pure<br />

Catalase + 200 mg/ml trehalose<br />

Protein solution molarity 0.001 mmol/m 3<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

Time [s]<br />

To investigate influences <strong>of</strong> the <strong>drying</strong> conditions <strong>and</strong> the ultrasonic levitation system on the<br />

activity <strong>of</strong> catalase, the solutions <strong>of</strong> the mixture series <strong>of</strong> catalase <strong>and</strong> trehalose with an overall solid<br />

content <strong>of</strong> 100 mg/ml are levitated at 60°C <strong>and</strong> different relative humidity in still air <strong>and</strong> the<br />

residual enzyme activity <strong>of</strong> the final particles is determined. Therefore, the <strong>formulations</strong> are freshly<br />

prepared in ice-cooled 50 mM tris buffer at pH 8.0 <strong>and</strong> divided into two portions. One is frozen at -<br />

80°C <strong>and</strong> used to determine the initial enzyme activity. The other portion is put on ice <strong>and</strong> directly<br />

used for the levitation experiments. The final particles are taken out <strong>of</strong> the st<strong>and</strong>ing acoustic wave<br />

as soon as no more changes in volume <strong>and</strong> vertical position can be determined for 30 seconds. The<br />

particles are also frozen at -80°C for later determination <strong>of</strong> the residual enzyme activity. Due to the<br />

low particle mass <strong>and</strong> the lack <strong>of</strong> accuracy when weighing the particles, the amount <strong>of</strong> catalase is


220 RESULTS AND DISCUSSION<br />

determined using the volume <strong>of</strong> the initial <strong>droplet</strong> calculated from the back illumination pictures<br />

taken by the CCD-camera <strong>and</strong> the density <strong>of</strong> the liquid solution, assuming a particle with no<br />

residual moisture content. To determine the catalase activity, the rate <strong>of</strong> disappearance <strong>of</strong> H2O2 was<br />

followed by monitoring the decrease in absorbance at 240 nm in 50 mM phosphate buffer at pH 7.0<br />

at 25°C. The method is described in detail in Chapter 4.2.6. The experiments are compared to the<br />

results obtained from spray-dried catalase <strong>formulations</strong>. 20 ml solutions are prepared the same way<br />

as for the levitation experiments <strong>and</strong> divided into 5 ml for the determination <strong>of</strong> the initial enzyme<br />

activity <strong>and</strong> to 15 ml for the spray-<strong>drying</strong> experiments in the Buchi Mini Spray-Dryer B-191 at<br />

T = 100° C , T = 60° C <strong>and</strong> a liquid feed <strong>of</strong> 3 ml/min. After the experiment the product is<br />

inlet<br />

outlet<br />

divided into two portions, one to determine the residual moisture content via Karl-Fischer titration<br />

<strong>and</strong> the other to determine the residual enzyme activity. All results are presented in Figure 5.130.<br />

Regarding the spray-<strong>drying</strong> results, the pure trehalose looses 2.9% enzyme activity. Compared to<br />

lactate dehydrogenase (LDH) which looses approximately 11% when spray-dried at T = 90° C<br />

<strong>and</strong> T = 60° C [Adler 1999] the loss in enzyme activity <strong>of</strong> catalase is very low under these<br />

outlet<br />

conditions. Formulating catalase with trehalose increases the activity <strong>of</strong> catalase. The addition <strong>of</strong><br />

20% trehalose based on the solid content is sufficient to achieve a residual activity <strong>of</strong> 99.1%.<br />

Further increase in trehalose content has no significant influence. In contrast, the residual activity <strong>of</strong><br />

the levitated particles at <strong>drying</strong> air conditions <strong>of</strong> 60°C <strong>and</strong> 5% relative humidity is determined to<br />

87.7%. As for spray-<strong>drying</strong>, the residual catalase activity increases with the addition <strong>of</strong> trehalose to<br />

the catalase formulation. 20% trehalose based on the solid content is sufficient to achieve a residual<br />

activity <strong>of</strong> 95.5% under the same <strong>drying</strong> conditions. Further increase in the trehalose content has no<br />

additional influence on the activity. An increase in <strong>drying</strong> air humidity results in a decrease in<br />

catalase activity. The residual enzyme activity determined for the pure catalase particles decreases<br />

to 81.2% at <strong>drying</strong> air humidity <strong>of</strong> 20% <strong>and</strong> to 77.2% at <strong>drying</strong> air humidity <strong>of</strong> 40%. The increased<br />

humidity increases the <strong>droplet</strong> surface temperature from 24.9°C at 60°C <strong>and</strong> 5% humidity to 43.2°C<br />

at 60°C <strong>and</strong> 40% humidity. As for the other results before, the addition <strong>of</strong> 20% trehalose based on<br />

the solid content increases the residual catalase activity to 92.2% at 20% relative humidity <strong>and</strong> to<br />

91.0% at 40% relative humidity. Under these conditions a further increase in trehalose content<br />

seems to result in a slight increase in residual enzyme activity. Although examinations on the<br />

determination <strong>of</strong> enzyme kinetics in levitated water <strong>droplet</strong>s state, that an influence <strong>of</strong> the st<strong>and</strong>ing<br />

acoustic wave <strong>and</strong> the ultrasonic field on the residual activity can be excluded [Weis 2005], the<br />

values for catalase dried in the levitator are lower than for all spray-dried solutions. The larger loss<br />

in activity using the levitation system can be caused by the longer h<strong>and</strong>ling <strong>and</strong> <strong>drying</strong> time<br />

compared to spray-<strong>drying</strong>. A slight decrease in residual catalase activity by averaged 2.4% (n = 3)<br />

inlet


CHAPTER 5 RESULTS AND DISCUSSION 221<br />

can be detected with an increase in effective initial SPL from 164.9 dB to 166.2 dB. Untreated<br />

dissolved catalase shows a 15% loss in activity in 2 hours at 20°C. Further investigations on<br />

enzyme activity <strong>of</strong> more sensitive <strong><strong>protein</strong>s</strong>, for example LDH, have to be done to obtain a better<br />

comparison <strong>of</strong> the two <strong>drying</strong> processes. Altogether a stabilizing effect <strong>of</strong> trehalose on the <strong>protein</strong><br />

can be seen for both <strong>drying</strong> techniques. The <strong>protein</strong>-sugar interactions as well as the effects <strong>and</strong> the<br />

exceptional role <strong>of</strong> trehalose in <strong>protein</strong> stabilization have been discussed in detail in Chapter 5.3.2<br />

<strong>and</strong> Chapter 5.3.4.<br />

Residual catalase activity [%]<br />

100.0<br />

95.0<br />

90.0<br />

85.0<br />

80.0<br />

75.0<br />

Spray-dried T in / T out = 100 / 60°C<br />

Levitated at 60°C <strong>and</strong> 5% humidity<br />

Levitated at 60°C <strong>and</strong> 20% humidity<br />

Levitated at 60°C <strong>and</strong> 40% humidity<br />

0.0 20.0 40.0 60.0 80.0<br />

Trehalose content [% solid]<br />

Figure 5.130: Residual catalase activity as a function <strong>of</strong> trehalose solid content in the formulation with catalase. Shown<br />

are results from spray-<strong>drying</strong> experiments at Tinlet=100°C, Toutlet=60°C <strong>and</strong> a liquid feed <strong>of</strong> 3 ml/min as well as from<br />

levitation experiments at 60°C <strong>and</strong> different relative humidity in still air. The initial <strong>droplet</strong> volume was 1.5 µl at an<br />

initial effective SPL between 164.6 dB <strong>and</strong> 165.1 dB.


CHAPTER 6 CONCLUSIONS 222<br />

6 Conclusions<br />

The present thesis deals with the set-up <strong>of</strong> an acoustic levitation system to examine pharmaceutical<br />

spray-<strong>drying</strong> <strong>formulations</strong> <strong>of</strong> peptides <strong>and</strong> <strong><strong>protein</strong>s</strong>. Spray-<strong>drying</strong> is a well known, rapid <strong>and</strong> easy<br />

method to transform peptide <strong>and</strong> <strong>protein</strong> solutions into dry powders for pulmonary <strong>and</strong> parenteral<br />

application. However, there are effects reported in literature with regard to <strong>protein</strong> stability, which<br />

limit the overall applicability <strong>of</strong> spray-<strong>drying</strong>. To investigate the influence <strong>of</strong> the formulation<br />

excipients <strong>and</strong> the <strong>drying</strong> conditions on the <strong>drying</strong> kinetics <strong>and</strong> the particle formation, single <strong>droplet</strong><br />

<strong>drying</strong> methods have been studied in the last couple <strong>of</strong> years. Acoustic levitation is the containerless<br />

processing <strong>of</strong> small liquid or solid samples in the pressure nodes <strong>of</strong> a st<strong>and</strong>ing acoustic wave<br />

between an ultrasonic transducer <strong>and</strong> a reflector. The characteristics <strong>of</strong> the acoustic levitator enable<br />

the determination <strong>of</strong> the single <strong>droplet</strong> <strong>drying</strong> kinetics within the constant rate <strong>and</strong> the falling rate.<br />

The goal <strong>of</strong> this work was first <strong>of</strong> all to set up an acoustic levitation system for single <strong>droplet</strong> <strong>drying</strong><br />

experiments under varying <strong>drying</strong> conditions. Secondly, the evaporation behaviour <strong>of</strong> single<br />

<strong>droplet</strong>s <strong>of</strong> pure solvents or solvent mixtures should be investigate <strong>and</strong> influences <strong>of</strong> the levitation<br />

system on the <strong>droplet</strong> <strong>drying</strong> kinetics should be taken into account. The experimental results had to<br />

be compared to simple gas-phase models <strong>and</strong> data from the literature. Furthermore, the evaporation<br />

behaviour <strong>of</strong> single solution <strong>droplet</strong>s containing sugars, peptides or <strong><strong>protein</strong>s</strong> <strong>and</strong> complex <strong>protein</strong><br />

<strong>formulations</strong> had to be investigated <strong>and</strong> compared to data form spray-<strong>drying</strong> experiments <strong>and</strong><br />

literature.<br />

In the first part <strong>of</strong> the thesis a single-walled acoustic levitator with a flat-flat transducerreflector<br />

design was set up into a self-constructed acrylic glass container. A mass flow <strong>and</strong> liquid<br />

flow controller in combination with a controlled evaporation mixer <strong>and</strong> an additional heater were<br />

introduced to the system to set predefined <strong>and</strong> accurate <strong>drying</strong> conditions. Thus, experiments in a<br />

temperature range from room temperature up to 70°C with a relative humidity from a few ppm to<br />

almost 100% were possible. Different connection possibilities <strong>of</strong> the <strong>drying</strong> air supply enabled<br />

evaporation in still air as well as under relative velocity conditions up to approximately 3.0 m/s. A<br />

variable positioning <strong>of</strong> the CCD-camera <strong>and</strong> the light source <strong>of</strong>fered further possibility <strong>of</strong> taking<br />

back light <strong>and</strong> front light illumination pictures to get a better underst<strong>and</strong>ing <strong>of</strong> the <strong>drying</strong> process<br />

<strong>and</strong> the particle formation. Comparison <strong>of</strong> levitated Migliol N812 <strong>droplet</strong>s with rigid polypropylene<br />

spheres showed a deformation <strong>of</strong> the Migliol N812 <strong>droplet</strong> to an oblate ellipsoid caused by the<br />

different values <strong>of</strong> axial <strong>and</strong> radial levitation forces. Furthermore, internal <strong>droplet</strong> oscillations<br />

occurring in vertical direction resulted in a larger st<strong>and</strong>ard deviation <strong>of</strong> the mean vertical diameter


CHAPTER 6 CONCLUSIONS 223<br />

<strong>of</strong> the <strong>droplet</strong> than <strong>of</strong> the rigid sphere. Due to the deformation <strong>of</strong> liquid samples the calculation <strong>of</strong><br />

<strong>droplet</strong> volume could not be performed assuming a spherical geometry. Experiments comparing the<br />

injected to the calculated volume showed the best correlation for an oblate rotational ellipsoid. An<br />

increasing effective sound pressure level (SPL) had a strong influence on the deformation <strong>of</strong> the<br />

<strong>droplet</strong> <strong>and</strong> leading to an increasing aspect ratio. Vibrations, air flow or a not ideal adjusted<br />

st<strong>and</strong>ing acoustic sound wave could lead to a deflection <strong>of</strong> the sample out <strong>of</strong> its rest position<br />

resulting in unstable <strong>droplet</strong> behaviour. Levitation capabilities could be improved by using a curved<br />

reflector <strong>and</strong> enlarging its section as to levitate dense materials or liquids under large ventilation air<br />

streams. The semi-confocal set-up with a reflector having a radius <strong>of</strong> curvature <strong>of</strong> twice the distance<br />

between transducer <strong>and</strong> reflector was the least problematic. Experiments at 60°C <strong>and</strong> 20% relative<br />

humidity in still air <strong>and</strong> with a <strong>drying</strong> air velocity <strong>of</strong> 2.0 m/s confirmed the superior <strong>and</strong> more stable<br />

<strong>droplet</strong> levitation <strong>of</strong> the semi-confocal set-up than the flat-flat one.<br />

In the second part <strong>of</strong> this thesis the evaporation <strong>of</strong> pure water <strong>and</strong> ethanol <strong>droplet</strong>s <strong>and</strong><br />

binary mixtures <strong>of</strong> both was analysed in the acoustic levitator <strong>and</strong> the results were compared to<br />

different mathematical single <strong>droplet</strong> <strong>drying</strong> models. The plot <strong>of</strong><br />

2<br />

rS rS<br />

0<br />

2 / versus<br />

2<br />

t rS<br />

0<br />

/ was the most<br />

suitable for data analysis, because it compensated different initial <strong>droplet</strong> volumes. The slope was<br />

defined as evaporation coefficient β r <strong>and</strong> characterized the evaporation process for a given pure<br />

solvent <strong>droplet</strong> under conditions valid for the d 2 -law. Experiments with water <strong>droplet</strong>s at 25°C in<br />

still air showed a faster evaporation than predicted by the diffusion-controlled evaporation model<br />

<strong>and</strong> a slower evaporation than predicted by the levitation model. Influence <strong>of</strong> the st<strong>and</strong>ing acoustic<br />

wave affected the evaporation process. The inner acoustic streaming led to an increased evaporation<br />

due to shear stress at the <strong>droplet</strong> surface, whereas the accumulation <strong>of</strong> solvent vapour in the toroidal<br />

vortices <strong>of</strong> the outer acoustic streaming had a hindering effect on the solvent evaporation. A force<br />

<strong>drying</strong> air stream <strong>of</strong> 1.0 m/s was determined to be suitable to remove the solvent vapour out <strong>of</strong> the<br />

vortices completely. Experimental results obtained at 25°C <strong>and</strong> under a <strong>drying</strong> air velocity <strong>of</strong><br />

1.0 m/s matched perfectly with the levitation model. The Sherwood number was determined to 5.3<br />

for an initial <strong>droplet</strong> volume <strong>of</strong> 2.0 µl. The same correlation with levitation model was obtained at<br />

40°C <strong>and</strong> 60°C with a <strong>drying</strong> air humidity <strong>of</strong> 0.1%. The initial Sherwood number at 40°C <strong>and</strong> 0.1%<br />

humidity was found to be 8.0 for an initial <strong>droplet</strong> volume <strong>of</strong> 2.0 µl. Drying air velocities > 1.0 m/s<br />

resulted in an additional convective influence on the <strong>droplet</strong> surface <strong>and</strong> could not be described by<br />

the levitation model any more. The Ranz-Marshall correlation provided a suitable equation to<br />

predict the existing Sherwood numbers <strong>and</strong> therefore the radius-time course <strong>of</strong> the <strong>drying</strong><br />

experiment. The plot <strong>of</strong> the evaporation coefficient in still air showed increasing values with


CHAPTER 6 CONCLUSIONS 224<br />

increasing temperature <strong>and</strong> decreasing <strong>drying</strong> air humidity. Under relative velocity conditions the<br />

evaporation coefficient determined in a diameter range between 1200 µm <strong>and</strong> 800 µm increased<br />

from 0.00193 mm 2 /s at 60°C <strong>and</strong> 0.1% humidity in still air to 0.00458 mm 2 /s under the same<br />

conditions but with a <strong>drying</strong> air velocity <strong>of</strong> 1.98 m/s. At the same time the evaporation time <strong>of</strong> a<br />

1.0 µl <strong>droplet</strong> was reduced from 225 s to approximately 80 s. Increasing <strong>drying</strong> air velocity <strong>and</strong><br />

decreasing humidity led to increasing values <strong>of</strong> the evaporation coefficient β r under relative<br />

velocity conditions.<br />

The evaporation <strong>of</strong> pure ethanol <strong>droplet</strong>s resulted in a faster evaporation <strong>and</strong> larger β r<br />

values than water due to the higher vapour pressure <strong>of</strong> ethanol at equal temperature. At 60°C <strong>and</strong><br />

0.1% humidity in still air, the evaporation coefficient <strong>of</strong> ethanol was determined to 0.00542 mm 2 /s.<br />

Influences <strong>of</strong> inner <strong>and</strong> outer acoustic streaming were as with pure water <strong>droplet</strong>s. The levitation<br />

model <strong>and</strong> the Ranz-Marshall correlation could also be used to determine the radius-time course at<br />

different temperatures <strong>and</strong> humidities ≤ 3.0%. With an increase in <strong>drying</strong> air humidity biphasic<br />

evaporation behaviour <strong>of</strong> the ethanol <strong>droplet</strong> evaporation could be seen. The vaporization <strong>of</strong> pure<br />

volatile alcohols in humid air was accompanied by the simultaneous condensation <strong>of</strong> water vapour<br />

on the <strong>droplet</strong> surface <strong>and</strong> its subsequent diffusion into the <strong>droplet</strong> interior. In the first phase, the<br />

evaporation <strong>of</strong> pure ethanol dominated. An increasing humidity resulted, however, in an increasing<br />

evaporation rate <strong>of</strong> the ethanol <strong>droplet</strong>s in this first part <strong>of</strong> evaporation. The associated heat release<br />

at the <strong>droplet</strong> surface due to the condensation increased the initial evaporation rate <strong>of</strong> the ethanol.<br />

The mixture <strong>of</strong> ethanol <strong>and</strong> water formed had a higher wet bulb temperature than pure ethanol,<br />

resulting in a larger evaporation rate. The wet bulb temperature <strong>of</strong> ethanol <strong>droplet</strong>s at 60°C <strong>drying</strong><br />

temperature increased from 13.3°C at 0.1% humidity to 32.6°C at 40% humidity. This effect led to<br />

increased evaporation coefficients for ethanol experiments not only with temperature but also with<br />

<strong>drying</strong> air humidity. After all <strong>of</strong> the ethanol was evaporated, the subsequent radius-time-course was<br />

dominated by the evaporation <strong>of</strong> the condensed water vapour. It could be shown, that the<br />

evaporation coefficient in this stage was equal to that <strong>of</strong> pure water <strong>droplet</strong>s under the same ambient<br />

conditions. Under relative velocity conditions an increase in <strong>drying</strong> air velocity resulted in an<br />

increase in evaporation coefficient. At 60°C <strong>and</strong> 0.1% humidity the evaporation coefficient<br />

increased to 0.01308 mm 2 /s resulting in a decrease <strong>of</strong> evaporation time for a 1.0 µl <strong>droplet</strong> from<br />

83.0 s in still air to 27.3 s at a velocity <strong>of</strong> 1.7 m/s.<br />

The binary mixture <strong>of</strong> water-ethanol also showed biphasic behaviour. The results obtained at<br />

60°C <strong>and</strong> 0.1% humidity showed increasing evaporation coefficients in the initial phase with<br />

increasing amount <strong>of</strong> ethanol in the solvent mixture. The values for the mixtures were between the


CHAPTER 6 CONCLUSIONS 225<br />

coefficients <strong>of</strong> pure water <strong>and</strong> pure ethanol under same ambient conditions. In the terminal part, the<br />

evaporation coefficients showed almost constant values independent <strong>of</strong> the initial composition <strong>of</strong><br />

the <strong>droplet</strong>. The comparison with pure water <strong>droplet</strong>s gave a good agreement.<br />

The third part <strong>of</strong> this thesis describes the application <strong>of</strong> the acoustic levitation system to<br />

determine the single <strong>droplet</strong> <strong>drying</strong> kinetics <strong>of</strong> solution <strong>droplet</strong>s. The <strong>drying</strong> behaviour <strong>and</strong> particle<br />

formation <strong>of</strong> sugar solutions relevant in spray-<strong>drying</strong> (maltodextrin, trehalose <strong>and</strong> mannitol), <strong>protein</strong><br />

solution (bSA <strong>and</strong> catalase) <strong>and</strong> <strong>protein</strong>-sugar <strong>formulations</strong> were investigated. Experiments<br />

performed with maltodextrin solutions showed a strong influence <strong>of</strong> the effective SPL on the shape<br />

<strong>of</strong> the resulting particle. With increasing effective SPL the initial aspect ratio <strong>of</strong> the levitated<br />

solution <strong>droplet</strong>s <strong>and</strong> the final particles increased. The initial <strong>droplet</strong> aspect ratio <strong>and</strong> the final<br />

particle aspect ratio showed a linear relationship (R 2 = 0.98). Maltodextrin solutions showed clear<br />

transition from constant to falling rate. Only small size changes <strong>of</strong> the vertical diameter could be<br />

determined in the second <strong>drying</strong> stage, resulting in flatting <strong>of</strong> the particle. Different initial <strong>droplet</strong><br />

sizes in still air experiments could be compensated by plotting the evaporation rate in evaporated<br />

mass per time <strong>and</strong> surface area versus<br />

2<br />

t rS<br />

0<br />

/ . Increased maltodextrin solid content resulted in a<br />

decrease in the duration <strong>of</strong> the first <strong>drying</strong> stage <strong>and</strong> a shallower decline <strong>of</strong> the evaporation rate in<br />

the second <strong>drying</strong> stage. For a 1.8 µl <strong>droplet</strong> dried at 40°C <strong>and</strong> 0.1% relative humidity in still air,<br />

the constant rate decreased from 445.3 s at 50 mg/ml solid content to 91.5 s at 400 mg/ml solid<br />

content whereas the falling rate period increased from 162.4 s to 708.6 s. Due to the different<br />

amount <strong>of</strong> dissolved molecules the evaporation rate in the first <strong>drying</strong> stage decreased from<br />

0.75 µg/s/mm 2 to 0.58 µg/s/mm 2 . Particle size analysis showed an increasing particle size with<br />

increasing solid content. The final particle density <strong>of</strong> the experiments at 25°C gave almost constant<br />

density values <strong>of</strong> approximately 0.85 g/cm 3 with increasing maltodextrin concentration, whereas a<br />

slight increase in particle density from approximately 0.7 g/cm 3 to 0.75 g/cm 3 could be determined<br />

at 40°C <strong>and</strong> 60°C <strong>drying</strong> temperature. SEM pictures <strong>of</strong> the final levitator particles showed a particle<br />

surface indented by small craters which were not combined with capillaries in the interior. The<br />

interior had a massive solid structure without any holes. The levitated particles showed a similar<br />

morphology compared to the larger particles <strong>of</strong> the spray dried powder.<br />

For trehalose solution <strong>droplet</strong>s no sharp transition from constant to falling rate could be seen<br />

form the graph <strong>of</strong> the volume as a function <strong>of</strong> evaporation time. The critical point could only be<br />

detected using the plot <strong>of</strong> the aspect ratio versus time. A flattening as for maltodextrin particles<br />

could not be seen. With increasing solid content an increase in the first <strong>and</strong> a decrease in the second<br />

<strong>drying</strong> stage could be determined. For an initially 1.8 µl <strong>droplet</strong> dried at 60°C <strong>and</strong> 5% relative


CHAPTER 6 CONCLUSIONS 226<br />

humidity in still air, the critical point appeared after 274.4 s for a 50 mg/ml solution <strong>and</strong> after<br />

138.9 s for a 400 mg/ml solution. The detectable falling rate period increased from 234.0 s to<br />

668.6 s. With increasing solid content <strong>and</strong> increasing <strong>drying</strong> air humidity the decline <strong>of</strong> the<br />

evaporation rate after the critical point became shallower. For a 400 mg/ml trehalose solution<br />

<strong>droplet</strong> at 60°C <strong>and</strong> 40% humidity, no transition between first <strong>and</strong> second <strong>drying</strong> stage could be<br />

detected any more. The final particle size in relation to the initial <strong>droplet</strong> size increased from 37.4%<br />

for a 50 mg/ml trehalose solution (60°C; 5% humidity; still air) with increasing solid content to<br />

75.5% for a 400 mg/ml solution. Increasing humidity showed a decreasing particle size. The final<br />

particle density was found to decrease with increasing trehalose content <strong>and</strong> decreasing humidity.<br />

The final particle density values for the above conditions were 1.03 g/cm 3 (50 mg/ml trehalose) <strong>and</strong><br />

0.97 g/cm 3 (400 mg/ml trehalose). Drying under relative velocity conditions showed a decrease in<br />

duration <strong>of</strong> the first <strong>drying</strong> stage <strong>and</strong> a less shallow decline <strong>of</strong> the evaporation rate after the critical<br />

point with increasing <strong>drying</strong> air velocity. The values <strong>of</strong> the evaporation rate in the first <strong>drying</strong> stage<br />

for a 100 mg/ml trehalose solution increased from 1.1 µg/s/mm 2 at 60°C <strong>and</strong> 5% humidity in still<br />

air to almost 3.0 µg/s/mm 2 at the same temperature <strong>and</strong> humidity but with a <strong>drying</strong> air velocity <strong>of</strong><br />

1.5 m/s. The faster evaporation process under relative velocity conditions than in still air resulted in<br />

larger <strong>and</strong> less dense particles.<br />

Mannitol solutions showed a sharp critical point <strong>and</strong> a constant particle size <strong>and</strong> aspect ratio<br />

within the second <strong>drying</strong> stage due to an abrupt particle formation. Influences <strong>of</strong> <strong>drying</strong><br />

temperature, relative humidity <strong>and</strong> velocity on the evaporation rate in the constant <strong>and</strong> falling rate<br />

were as with maltodextrin <strong>and</strong> trehalose. A determination <strong>of</strong> the single <strong>droplet</strong> <strong>drying</strong> kinetics under<br />

air flow conditions could only be performed up to a <strong>drying</strong> air velocity <strong>of</strong> 1.0 m/s. Larger velocities<br />

resulted in strong <strong>droplet</strong> oscillations starting in the second <strong>drying</strong> stage. The abrupt particle<br />

formation <strong>and</strong> unbalanced particle shape in combination with the very low particle density caused<br />

turbulences <strong>of</strong> the <strong>drying</strong> air in the area <strong>of</strong> the pressure node <strong>and</strong> made the particle susceptible for a<br />

blow-out. Mannitol showed the lowest particle density <strong>of</strong> all examined substances <strong>and</strong> <strong>formulations</strong>.<br />

Under the fasted <strong>drying</strong> conditions at 60°C <strong>and</strong> 5% humidity in still air a final particle density <strong>of</strong><br />

0.50 g/cm 3 was obtained at a <strong>drying</strong> air velocity <strong>of</strong> 1.0 m/s even 0.32 g/cm 3 . The SEM pictures <strong>of</strong><br />

particles produced at room temperature showed only partially spherical particles with protuberances<br />

at their surface <strong>and</strong> a spicular morphology. Experiments with faster solvent evaporation led to<br />

particles with a smoother surface compared to the particles produced at room temperature. The<br />

particles had a hollow interior with blow hole in their surface. The SEM pictures <strong>of</strong> spray-dried<br />

mannitol in largest magnification showed particles with a smooth surface. A spicular surface


CHAPTER 6 CONCLUSIONS 227<br />

structure could not be determined. Bowl like particles could be found under air flow conditions with<br />

a smooth bottom side into the direction <strong>of</strong> the air stream <strong>and</strong> protuberances on the top side.<br />

The <strong>drying</strong> <strong>of</strong> bSA solution in the acoustic levitation system showed a behaviour different to<br />

the sugar solutions. Breakage <strong>of</strong> the new formed particle with hollow interior as well as formation<br />

<strong>of</strong> large holes in the surface after the critical point led to instable particle behaviour in the second<br />

<strong>drying</strong> stage with strong oscillations. Analysis <strong>of</strong> the SEM pictures determined a particle crust with<br />

the thickness <strong>of</strong> 31.75% <strong>of</strong> the particle radius for a bSA solution with 100 mg/ml solid content<br />

(60°C; 10% humidity; still air). For a spherical particle with a radius <strong>of</strong> 500 µm a void volume <strong>of</strong><br />

0.166 µl was calculated enlarging the outer surface area <strong>of</strong> the particle by an interior surface <strong>of</strong><br />

1.46 mm 2 to an overall surface area <strong>of</strong> 4.60 mm 2 . This change in surface area within the second<br />

<strong>drying</strong> stage resulted in an increase <strong>of</strong> the evaporation rate to value almost as high as in the constant<br />

rate period. Formulation <strong>of</strong> bSA with trehalose in a bSA-trehalose (1:1) mixture with 100 mg/ml<br />

solid content did not show increasing evaporation rate values in the falling rate period. Blow holes<br />

or hollow particle interiors could not be found on the SEM, either. A comparison <strong>of</strong> the pure<br />

<strong>protein</strong>, the mixture <strong>and</strong> the pure trehalose showed an increase in duration <strong>of</strong> the first <strong>drying</strong> stage<br />

with increasing trehalose content., 65.1% <strong>of</strong> the solvent was evaporated in the constant rate from the<br />

pure bSA solution <strong>droplet</strong>s, whereas the percentage <strong>of</strong> solvent evaporated for the mixture <strong>and</strong> the<br />

pure trehalose was almost equal with 87.8 to 85.6% (60°C; 5% humidity; still air). The evaporation<br />

rate in the first <strong>drying</strong> stage at 60°C <strong>and</strong> 10% humidity in still air decreased from 1.21 µg/s/mm 2 for<br />

the pure bSA to 0.91 µg/s/mm 2 for the pure trehalose. The mixture had a value <strong>of</strong> approximately<br />

1.0 µg/s/mm 2 . With disregard <strong>of</strong> the pure bSA solution, the decline <strong>of</strong> the evaporation rate after the<br />

critical point became shallower with increasing trehalose content. A comparison to the spray-dried<br />

bSA <strong>and</strong> bSA-trehalose product showed very good correlation. Due to the hollow particle interior<br />

the density <strong>of</strong> 0.62 g/cm 3 <strong>of</strong> the pure bSA particles produced in the levitation system was<br />

approximately 0.3 units lower than for the mixture with trehalose. A comparison <strong>of</strong> the bSAtrehalose<br />

(1:1) mixture <strong>and</strong> the pure trehalose particles resulted in almost the same values for size<br />

<strong>and</strong> density with slightly larger mixture particles with a minimal lower density. Investigations on<br />

interfacial behaviour <strong>of</strong> bSA <strong>and</strong> its mixture with 200 mg/ml trehalose using bubble pressure<br />

tensiometry showed that trehalose increased the dynamic interfacial tension <strong>of</strong> a 2.0 mmol/l bSA<br />

solution by 3.67 mN/m <strong>and</strong> <strong>of</strong> a 4.0 mmol/l bSA solution by even 4.42 mN/m (interfacial lifetime<br />

1.0 s). Calculation <strong>of</strong> the surface excess concentration using the Gibbs Equation showed a reduction<br />

<strong>of</strong> bSA at the surface from 87 mg/m 2 to 36 mg/m 2 on the addition <strong>of</strong> 200 mg/ml trehalose to a


CHAPTER 6 CONCLUSIONS 228<br />

4.0 mmol/l bSA solution at an interfacial lifetime <strong>of</strong> 1.0 s. The area per molecule in a bSA<br />

monolayer was determined to be 100 nm 2 at the transition from gaseous to liquid-condensed state.<br />

For pure catalase solutions in 50 mM tris buffer at pH 8.0 levitated at 60°C in still air<br />

particles with surface holes <strong>of</strong> a mean diameter <strong>of</strong> 45.1 ± 4.3% <strong>of</strong> the horizontal particle diameter<br />

were found. The solid crust had a thickness <strong>of</strong> approximately 18.3% <strong>of</strong> the final particle radius. A<br />

spherical particle with a radius <strong>of</strong> 500 µm had an overall volume <strong>of</strong> 0.524 µl with a void volume <strong>of</strong><br />

0.286 µl. The surface area <strong>of</strong> the internal void sphere was calculated to 2.10 mm 2 <strong>and</strong> enlarged the<br />

outer surface to an overall surface area <strong>of</strong> 5.24 mm 2 . The increase in surface area resulted in a<br />

meanwhile increase in evaporation rate within the second <strong>drying</strong> stage. <strong>Single</strong> <strong>droplet</strong> <strong>drying</strong><br />

investigation on a catalase-trehalose mixture series with an overall solid content <strong>of</strong> 100 mg/ml<br />

showed that a catalase-trehalose relation <strong>of</strong> 6:4 was needed to obtain massive particles without<br />

surface holes <strong>and</strong> to overcome the effects <strong>of</strong> an additional surface on the evaporation rate in the<br />

second <strong>drying</strong> stage. With increasing trehalose content, the duration <strong>of</strong> the first <strong>drying</strong> stage <strong>and</strong><br />

therefore the percentage <strong>of</strong> solvent evaporated in this phase increased from 61.9% for the pure<br />

catalase at to 85.5% for a catalase-trehalose (2:8) mixture (60°C; 5% humidity; still air). The initial<br />

evaporation rate in the first <strong>drying</strong> stage decreased from 0.95 µg/s/mm 2 for pure catalase to<br />

0.91 µg/s/mm 2 for the catalase-trehalose (2:8) mixture under the above conditions. The pure<br />

trehalose solution in tris buffer had an initial value <strong>of</strong> 0.89 µg/s/mm 2 . With increasing trehalose<br />

content the decline in evaporation after the critical point became shallower. The morphological<br />

correlation <strong>of</strong> levitated <strong>and</strong> spray-dried product was very good. The density <strong>of</strong> the levitated particles<br />

increased from 0.43 g/cm 3 for the pure catalase with increasing trehalose content to 0.84 g/cm 3 for<br />

the catalase-trehalose (2:8) mixture (60°C; 5% humidity; still air). Bubble pressure tensiometry<br />

showed that 200 mg/ml trehalose increased the dynamic interfacial tension <strong>of</strong> a 0.40 mmol/l<br />

catalase solution, equivalent to 100 mg/ml, by 3.9 mN/m at an interfacial lifetime <strong>of</strong> 1.0 s.<br />

Furthermore, the calculated surface excess concentration was reduced from 575 mg/mm 2 to<br />

400 mg/mm 2 . Investigations on enzyme kinetics in acoustically levitated <strong>droplet</strong>s did not correlate<br />

to the results obtained from the same spray dried catalase-trehalose <strong>formulations</strong>, due to the long<br />

evaporation time at 60°C <strong>and</strong> the large <strong>droplet</strong> size <strong>of</strong> the samples in the levitation system.<br />

However, a stabilizing influence <strong>of</strong> trehalose on the residual catalase activity could be shown.<br />

Addition <strong>of</strong> 20% trehalose in the solid content, equal to a molar ratio catalase-trehalose ratio <strong>of</strong><br />

1:183, increased the residual activity from 87.7% to 95.5% at <strong>drying</strong> conditions <strong>of</strong> 60°C <strong>and</strong> 5%<br />

relative humidity in still air. An increase in <strong>drying</strong> time due to an increase in humidity resulted in a<br />

slightly smaller residual catalase activity.


CHAPTER 6 CONCLUSIONS 229<br />

In summary, the acoustic levitator appears to be a promising tool to investigate the single<br />

<strong>droplet</strong> <strong>drying</strong> behaviour <strong>of</strong> pure solvents, solvent mixtures, suspensions <strong>and</strong> solutions. Even though<br />

influences <strong>of</strong> the st<strong>and</strong>ing acoustic wave on mass transfer <strong>and</strong> particle formation could be<br />

demonstrated, it was possible to predict the radius-time course <strong>of</strong> the evaporation process in the<br />

levitation system with mathematical models <strong>and</strong> to show a very good morphological correlation <strong>of</strong><br />

the levitated particles with the spray-dried powder <strong>of</strong> the same formulation. These results <strong>and</strong> the<br />

possibility to set almost every possible <strong>drying</strong> condition within a given temperature range make the<br />

levitator an interesting tool to perform preliminary test in spray-<strong>drying</strong> research, especially if there<br />

is only little amount <strong>of</strong> substance available. Reconstruction <strong>of</strong> the system with the integration <strong>of</strong> a<br />

cooled ultrasonic transducer <strong>and</strong> an automatic <strong>droplet</strong> injection via a micro-pump system would<br />

<strong>of</strong>fer the possibility to dry at temperature higher than 70°C with a smaller initial <strong>droplet</strong> size <strong>and</strong> to<br />

obtain an improved correlation to the spray-<strong>drying</strong> experiments. The introduction <strong>of</strong> an IRthermocamera<br />

would enable to determine the surface temperature course <strong>of</strong> solvents or solutions<br />

with evaporation time what could provide a tool to verify the countless models in literature to<br />

calculate the wet-bulb temperature. Further application areas arise beside the determination <strong>of</strong><br />

<strong>drying</strong> kinetics, particle formation or <strong>protein</strong> stability. It is quite conceivable to use the ultrasonic<br />

levitation system for basic pharmaceutical investigations <strong>of</strong> distribution processes inside a liquid<br />

<strong>droplet</strong>, for in-situ micro-encapsulation or polymerisation or to examine the quality <strong>of</strong> encapsulation<br />

formulation, for example <strong>of</strong> volatile substances, not only in pharmaceutical but also in food<br />

technology.


CHAPTER 7 ZUSAMMENFASSUNG 230<br />

7 Zusammenfassung<br />

Die vorliegende Arbeit beschäftigt sich mit dem Aufbau eines akustischen Levitationssystems zur<br />

Untersuchung von pharmazeutischen Peptid- und Proteinformulierungen im Bereich der<br />

Sprühtrocknung. Die Sprühtrocknung ist eine bekannte, schnelle und einfache Methode Peptidoder<br />

Proteinlösungen in trockene Pulver für parenterale oder pulmonale Applikation zu überführen.<br />

Jedoch wird in Veröffentlichungen immer wieder auf Einflüsse bezüglich der Proteinstabilität<br />

hingewiesen, die die Anwendbarkeit der Sprühtrocknung in diesem Bereich limitieren. Um<br />

Auswirkungen der Formulierungszusammensetzung und der Trocknungsbedingungen auf die<br />

Trocknungskinetik und die Partikelbildung zu untersuchen, wurde in den letzten Jahren die<br />

Trocknung einzelner Tropfen mit verschiedenen Methoden zunehmend untersucht. Unter<br />

akustischer Levitation versteht man das berührungslose Arbeiten mit kleinen, festen oder flüssigen<br />

Proben, die im Druckknoten einer stehenden akustischen Welle zwischen einem Ultraschallerreger<br />

und einem Reflektor schweben. Die Eigenschaften dieser stehenden Welle ermöglichen die<br />

Bestimmung der Trocknungskinetik eines einzelnen Tropfens in beiden Trocknungsabschnitten.<br />

Das Ziel dieser Arbeit war zuerst der Aufbau eines akustischen Levitationssystems mit dem es<br />

möglich ist, unter variierenden Bedingungen die Trocknungskinetik einzelner Tropfen zu<br />

untersuchen. Anschließend wurde das Trocknungsverhalten reiner Lösungsmittel und derer<br />

Mischungen untersucht. Dabei sollten Einflüsse des Ultraschallfeldes auf die Trocknung genau<br />

betrachtet werden. Die Ergebnisse mussten anschließend mit einfachen Trocknungsmodellen und<br />

bereits publizierten Daten verglichen werden. Als drittes sollte das Trocknungsverhalten von<br />

Zucker- und Proteinlösungen im Ultraschall-Levitator untersucht und mit Daten aus<br />

Sprühtrocknungsexperimenten und Publikationen verglichen werden.<br />

Im ersten praktischen Teil dieser Arbeit wurde um den einw<strong>and</strong>igen, akustischen Levitator<br />

mit flacher Geometrie von Ultraschallw<strong>and</strong>ler und Reflektor ein Acrylglascontainer ähnlich einer<br />

Glove-Box konstruiert, um Einflüsse durch die Umgebungsluft zu minimieren. Durchflussmesser<br />

für Gas und Flüssigkeiten wurden mit einem Konditioniersystem kombiniert und eine zusätzliche<br />

Heizung in das System eingebaut. Es war somit möglich definierte Trocknungsbedingungen von<br />

Raumtemperatur bis 70°C mit Luftfeuchten von annähernd trockener Luft bis fast 100%<br />

einzustellen. Verschiedene Zuluftverbindungen zum Levitator ermöglichten die Trocknung sowohl<br />

in ruhiger Luft als auch mit Luftgeschwindigkeiten bis ungefähr 3,0 m/s. Variable Positionen der<br />

CCD-Kamera und der Lichtquelle boten die Möglichkeit sowohl Gegenlicht als auch<br />

Stirnlichtbilder während des Trocknungsprozesses aufzunehmen. Der Vergleich eines levitierten


CHAPTER 7 ZUSAMMENFASSUNG 231<br />

Migliol N812-Tropfens mit einer festen Polypropylenkugel zeigte deutlich die durch<br />

unterschiedliche Werte für die axiale und radiale Leviationskraft verursachte Verformung des<br />

Tropfens. Des Weiteren konnte an der größeren St<strong>and</strong>ardabweichung des Mittelwertes des<br />

horizontalen Tropfendurchmessers von Migliol N812 im Vergleich zur Polypropylenkugel, das<br />

Auftreten interner Tropfenschwingungen in vertikaler Richtung deutlich gemacht werden. Die<br />

Deformation der Flüssigkeitsproben ergab für den geometrischen Körper eines Rotationsellipsoiden<br />

die beste Übereinstimmung bei der Größen- und Volumenberechnung des Tropfens. Mit steigendem<br />

Schalldruck nahm das Achsenverhältnis aus horizontalem und vertikalem Tropfendurchmesser zu.<br />

Gehäusevibrationen, Trocknungsluft oder eine nicht optimal eingestellte stehende akustische Welle<br />

konnten zu einer Auslenkung des Tropfenschwerpunktes aus seiner Ruheposition und zu<br />

Instabilitäten während des Experimentes führen. Um das Leistungsvermögen des akustischen<br />

Levitators hinsichtlich einer stabilen Probenposition zu erhöhen, wurde die ursprünglich flache<br />

Geometrie von Ultraschallw<strong>and</strong>ler und Reflektor gegen eine semikonfokale ausgetauscht. Bei<br />

einem Kurvenradius des Reflektors mit dem doppelten Abst<strong>and</strong> zwischen W<strong>and</strong>ler und Reflektor<br />

ergab sich die größte Probenstabilität.<br />

Im zweiten Teil der Arbeit wurde die Verdampfung von reinem Wasser und Ethanol und<br />

derer Gemische untersucht und die experimentellen Ergebnisse mit Trocknungsmodellen<br />

verglichen. Dabei zeigte sich die graphische Darstellung des Tropfenradius<br />

2<br />

t rS<br />

0<br />

2<br />

rS rS<br />

0<br />

2 / gegen die Zeit<br />

/ zur Kompensation unterschiedlich großer Anfangsvolumina am besten geeignet. Unter<br />

Bedingungen des d 2 -Gesetzes ergab die Steigung des Graphen den für die Verdampfung des reinen<br />

Lösungsmittels charakteristischen Trockungskoeffizienten β r . Experimente mit Wasser bei 25°C<br />

zeigten eine schnellere Verdampfung des Tropfens als mit dem „Diffusionskontrollierten<br />

Trocknungsmodell“ vorhergesagt wurde und eine langsamere Verdampfung als mit dem<br />

„Levitationsmodell“ berechnet. Hieraus ließen sich die Einflüsse der stehenden akustischen Welle<br />

auf den Trocknungsprozess deutlich erkennen. Während die innere akustische Strömung durch<br />

Scherbeanspruchung der Tropfenoberfläche den Massenübergang und die Sherwood-Zahl erhöhte,<br />

wurde durch Dampfanreicherung im Wirbelsystem der äußeren akustischen Strahlung der<br />

Trocknungsprozess hingegen verlangsamt. Die Einbettung des Tropfens in einen<br />

Trocknungsluftstrahl mit einer Geschwindigkeit von ungefähr 1.0 m/s zeigte sich als ausreichend,<br />

um das den Tropfen umgebende Wirbelsystem frei zu blasen. Experimente bei 25°C mit einer<br />

Luftgeschwindigkeit von 1.0 m/s stimmten mit dem vorhergesagten Trocknungsverlauf des<br />

„Levitationsmodelles“ überein. Für einen anfänglich 2.0 µl großen Tropfen ergab sich unter diesen<br />

Bedingungen eine anfängliche Sherwood-Zahl von 5.3. Dieselbe Übereinstimmung experimenteller


CHAPTER 7 ZUSAMMENFASSUNG 232<br />

Daten mit dem „Levitationsmodell“ wurde auch bei 40°C und 60°C Trocknungstemperatur mit<br />

einer Luftfeuchte von 0.1% gefunden. Für einen 2.0 µl großen Tropfen bei 40°C erhöhte sich dabei<br />

die anfängliche Sherwood-Zahl auf 8.0. Trocknungsexperimente mit einer Luftgeschwindigkeit<br />

> 1.0 m/s konnten aufgrund eines zusätzlichen konvektiven Einflusses nicht mehr durch das<br />

„Levitationsmodell“ beschrieben werden. Mittels der Ranz-Marshall Gleichung war eine<br />

Vorhersage der vorliegenden Sherwood-Zahl und damit des Trocknungsverlaufes möglich. In<br />

unbewegter Trocknungsluft nahm der Trocknungskoeffizient mit steigender Temperatur und<br />

abnehmender Luftfeuchte zu. Mit zusätzlicher Luftgeschwindigkeit stieg β r von 0,00193 mm 2 /s<br />

(unbewegte Luft / 60°C / 0,1% Feuchte) auf 0,00458 mm 2 /s (Luftgeschwindigkeit 1,98 m/s /<br />

60°C / 0,1% Feuchte) an. Die Werte wurden bei einem Tropfendurchmesser zwischen 1200 µm und<br />

800 µm bestimmt. Gleichzeitig nahm die Trocknungszeit eines Tropfens mit einem Volumen von<br />

1,0 µl von 225 s auf 80 s ab. Experimente mit direkter Luftströmung ergaben eine Zunahme von β r<br />

mit zunehmender Luftgeschwindigkeit und abnehmender Feuchte.<br />

Aufgrund des höheren Dampfdrucks bei gleicher Temperatur zeigten Ethanoltropfen eine<br />

schnellere Verdampfung und größere β r -Werte als Wassertropfen. In unbewegter Luft bei 60°C<br />

und 0,1% Feuchte stieg der Trocknungskoeffizient für Ethanol auf 0.00542 mm 2 /s an. Die Einflüsse<br />

der inneren und äußeren akustischen Strahlung hatten die gleiche Auswirkung auf das<br />

Trocknungsverhalten von Ethanoltropfen. Das „Levitationsmodell“ und die Ranz-Marshall<br />

Gleichung konnten zur Vorhersage des Trocknungsverlaufes bei verschiedenen Temperaturen und<br />

Luftfeuchten ≤ 3.0% verwendet werden. Mit Erhöhung der Luftfeuchte wurde ein zweiphasiges<br />

Verdampfungsverhalten bei Ethanol beobachtet, das von einer Kondensation von Feuchte aus der<br />

Trocknungsluft an der Tropfenoberfläche mit nachfolgender Diffusion ins Tropfeninnere<br />

gekennzeichnet war. In der ersten Phase, geprägt durch die Verdampfung von Ethanol, führte eine<br />

Erhöhung der relativen Feuchte der Trocknungsluft zu einer Erhöhung der Verdampfungsrate des<br />

Ethanols. Die durch die Kondensation freiwerdende Energie erhöhte die anfängliche Verdampfung<br />

des Ethanols pro Zeiteinheit. Die Mischung aus Ethanol und kondensierter Feuchte hatte eine<br />

höhere Kühlgrenztemperatur als der reine Ethanol, was wiederum zu größeren Verdampfungsraten<br />

und damit zu einem größeren β r führte. Die Kühlgrenztemperatur am Ethanoltropfen bei 60°C<br />

Trocknungsluft steigt dabei von 13.3°C bei 0,1% Feuchte auf 32,6°C bei 40% Feuchte an. Nach der<br />

Verdampfung des Ethanols wurde der Trocknungsverlauf durch die Verdampfung der<br />

kondensierten Feuchte dominiert. Vergleichende Betrachtungen der zweiten Verdampfungsphase<br />

von Ethanoltropfen in feuchter Luft mit Wassertropfen zeigten sehr gute Übereinstimmung.<br />

Luftgeschwindigkeiten von 1.7 m/s bei 60°C und 0.1% relativer Feuchte erhöhten den


CHAPTER 7 ZUSAMMENFASSUNG 233<br />

Verdampfungskoeffizienten auf 0.01308 mm 2 /s. Dabei sank die Verdampfungszeit eines 1.0 µl<br />

Tropfens von 83.0 s in unbewegter Luft auf 27.3 s ab.<br />

Ein zweiphasiges Verdampfungsverhalten wurde auch für die binäre Mischung aus Wasser<br />

und Ethanol gefunden. In der ersten Phase zeigte sich mit zunehmendem Ethanolanteil eine<br />

Zunahme von β r . Die Werte lagen dabei zwischen denen der reinen Lösungsmittel. Der zweite Teil<br />

war durch ein konstantes β r unabhängig von der ursprünglichen Zusammensatzung des Tropfens<br />

gekennzeichnet. Ein Vergleich des zweiten Abschnitts mit reinen Wassertropfen ergab eine sehr<br />

gute Übereinstimmung.<br />

Im dritten Teil dieser Arbeit wurde das akustische Levitationssystem für die Untersuchung<br />

der Trocknungskinetik von Lösungstropfen verwendet. Dabei lag das Hauptinteresse auf<br />

Substanzen und Formulierungen, die in pharmazeutischen Bereichen der Sprühtrocknung<br />

Anwendung finden, wie Zucker (Maltodextrin, Trehalose und Mannitol), Proteine (bovines Serum<br />

Albumin und Katalase) und deren Mischungen. Untersuchungen zum Einfluss des Ultraschallfeldes<br />

auf Maltodextrinlösungen zeigten einen Zusammenhang zwischen der Stärke des Ultraschallfeldes,<br />

den Achsenverhältnissen der Tropfen und der daraus entstehenden Partikeln. Zwischen beiden<br />

letzteren konnte ein linearer Zusammenhang gefunden werden (R 2 = 0.98). Zur Minimierung der<br />

Einflüsse auf die Partikelbildung wurde die für eine stabile Levitation nötige, niedrigste Stärke des<br />

Ultraschallfeldes gewählt. Bezüglich ihres Trocknungsverhaltens zeigten Maltodextrinlösungen am<br />

Übergang vom ersten zum zweiten Trocknungsabschnitt einen deutlichen kritischen Punkt. Kleine<br />

Veränderungen des vertikalen Durchmessers in der zweiten Trocknungsphase führten zu einer<br />

Abflachung des resultierenden Partikels. Die Darstellung der Trocknungsgeschwindigkeit in<br />

verdampfter Lösungsmittelmenge pro Zeit und Oberfläche gegen<br />

2<br />

t rS<br />

0<br />

/ bot die Möglichkeit<br />

Experimente mit unterschiedlich großen Anfangstropfen zu vergleichen. Mit Erhöhung der<br />

Maltodextrinkonzentration in der Lösung kam es zu einer Verkürzung des ersten und einem<br />

flacheren Abfallen der Trocknungsgeschwindigkeit im zweiten Trocknungsabschnitt. Für einen<br />

1,8 µl Tropfen, getrocknet bei 40°C und 0,1% Feuchte in unbewegter Luft, verkürzte sich die erste<br />

Trocknungsphase von durchschnittlich 445,3 s bei einem Festst<strong>of</strong>fgehalt von 50 mg/ml auf 91,5 s<br />

bei 400 mg/ml. Im Gegenzug nahm die Dauer des zweiten Trocknungsabschnittes von 162,4 s auf<br />

708,6 s zu. Die Trocknungsrate sank von 0,75 µg/s/mm 2 auf 0,58 µg/s/mm 2 ab. Eine Erhöhung der<br />

Trocknungstemperatur von 25°C auf 60°C verkürzte die Trocknungszeit beider Phasen. Mit<br />

zunehmendem Festst<strong>of</strong>fgehalt in der Lösung nahm die Größe der getrockneten Partikel zu, wobei<br />

die Dichte bei 25°C mit 0,85 g/cm 3 in etwa konstant blieb. Bei 40°C und 60°C hingegen wurde mit


CHAPTER 7 ZUSAMMENFASSUNG 234<br />

steigender Maltodextrinkonzentration ein leichter Anstieg der Dichte von 0,7 g/cm 3 auf 0,75 g/cm 3<br />

festgestellt. Die rasterelektronenmikroskopischen Aufnahmen (REM) zeigten eine eingedellte und<br />

mit kleinen Kratern übersähte Oberfläche. Eine Verbindung dieser Krater zu Kapillaren im Inneren<br />

konnte nicht festgestellt werden. Die resultierenden Partikel waren massiv ohne Hohlräume und<br />

zeigten deutliche morphologische Ähnlichkeit zu den größeren Partikeln der sprühgetrockneten<br />

Pulver.<br />

Trehaloselösungen zeigten am Übergang vom ersten zum zweiten Trocknungsabschnitt<br />

keinen deutlichen kritischen Punkt. Lediglich ein beginnender Anstieg des Halbachsenverhältnisses<br />

deutete auf eine Krustenbildung hin. Ein Abflachen der Trehalosepartikel ähnlich wie beim<br />

Maltodextrin konnte nicht beobachtet werden. Ansonsten zeigten Trehaloselösungen ein ähnliches<br />

Trocknungsverhalten. Mit zunehmender Konzentration kam es zu einer Verkürzung des ersten und<br />

einer Verlängerung des zweiten Trocknungsabschnittes. Die Trocknungszeit eines 1,8 µl Tropfen<br />

bei 60°C und 5% relativer Feuchte in unbewegter Luft nahm von durchschnittlich 274,4 s bei<br />

50 mg/ml Trehalose auf 138,9 s bei 400 mg/ml Trehalose ab. Im Gegenzug stieg die Dauer des<br />

zweiten Trocknungsabschnittes von 234,0 s auf 668,6 s an. Mit zunehmender Konzentration zeigte<br />

sich eine langsamere Abnahme der Trocknungsgeschwindigkeit nach dem kritischen Punkt. Bei<br />

400 mg/ml Trehalose getrocknet bei 60°C in unbewegter Luft mit einer Feuchte von 40% konnte<br />

kein kritischer Punkt und damit kein Übergang vom ersten zum zweiten Trocknungsabschnitt mehr<br />

erkannt werden. Bei gleicher Temperatur mit 5% Feuchte in unbewegter Luft nahm die<br />

Partikelgröße von 37,4% der ursprünglichen Tropfengröße bei 50 mg/ml Trehalose auf 75,5% bei<br />

400 mg/ml zu. Zunehmende Luftfeuchte führte zu einer abnehmenden Partikelgröße. Die<br />

Partikeldichte nahm mit zunehmendem Festst<strong>of</strong>fgehalt und abnehmender Feuchte ab. Unter den<br />

obigen Bedingungen bei 5% Feuchte ergaben sich Dichten von 1,03 g/cm 3 (50 mg/ml Trehalose)<br />

und 0,97 g/cm 3 (400 mg/ml Trehalose). Ein Trocknen mit Luftströmung führte mit zunehmender<br />

Luftgeschwindigkeit zu einer Verkürzung des ersten Trocknungsabschnittes und zu einem steileren<br />

Abfall der Trocknungsgeschwindigkeit nach dem kritischen Punkt. Im Falle einer 100 mg/ml<br />

Trehaloselösung nahm die Trocknungsgeschwindigkeit bei 60°C und 5% Feuchte von 1,1 µg/s/mm 2<br />

in unbewegter Luft auf fast 3,0 µg/s/mm 2 bei einer Strömung von 1,5 m/s zu. Die schnellere<br />

Trocknung führte zu größeren und weniger dichten Trehalosepartikeln.<br />

Mannitollösungen zeigten einen sehr deutlichen kritischen Punkt am Übergang vom ersten<br />

zum zweiten Trocknungsabschnitt, sowie eine konstante Partikelgröße und ein konstantes<br />

Halbachsenverhältnis in der zweiten Phase. Die Einflüsse der Trocknungstemperatur, Luftfeuchte<br />

und Luftströmung auf die Trocknungsgeschwindigkeit waren wie bei Maltodextrin und Trehalose.


CHAPTER 7 ZUSAMMENFASSUNG 235<br />

Allerdings konnten Versuche in bewegter Luft nur bis zu einer Luftgeschwindigkeit von 1,0 m/s<br />

durchgeführt werden. Stärkere Luftströmung führte zu starken Tropfenschwingungen im zweiten<br />

Trocknungsabschnitt. Die schlagartige Partikelbildung und die ungleichmäßige Partikelform in<br />

Kombination mit einer sehr niedrigen Dichte führten zur Ausbildung von Turbulenzen im Bereich<br />

des Druckknotens und zu einem Herausblasen des Partikels aus der stehenden akustischen Welle.<br />

Bei 60°C und 5% Feuchte in unbewegter Luft wurde eine Partikeldichte von 0,50 g/cm 3 erhalten, in<br />

Luftgeschwindigkeit von 1,0 m/s sogar 0,32 g/cm 3 . REM-Bilder von Partikeln entst<strong>and</strong>en bei<br />

Raumtemperatur zeigten eine nur teilweise runde Partikelgestalt mit vielen Ausstülpungen und<br />

einer nadelförmigen Oberflächenmorphologie. Eine schnellere Trocknung führte zu runderen<br />

Partikeln mit glätterer Oberfläche als bei Raumtemperatur. Die Mannitolpartikel waren hohl mit<br />

Löchern in ihrer Oberfläche. Die Morphologie konnte im sprühgetrocknete Produkt in höchster<br />

Vergrößerung nicht gefunden werden. In bewegter Luft wurden im Levitationssystem Partikel mit<br />

„Schüsselgestalt“ erhalten. Die Partikel hatten eine glatte, der Luftströmung zugew<strong>and</strong>te und eine<br />

von Ausstülpungen übersähte obere Hälfte.<br />

Die Trocknung von bSA-Lösungen zeigte ein den Zuckerlösungen unterschiedliches<br />

Verhalten. Sowohl der Bruch des neu gebildeten hohlen Partikels als auch das Entstehen großer<br />

Löcher in der Oberfläche nach dem kritischen Punkt führten im zweiten Trocknungsabschnitt zu<br />

instabilem Verhalten mit starken Partikelschwingungen. Die Analyse der REM-Bilder ergab eine<br />

Partikelkruste mit einer Breite von 31,75% des Partikelradius für eine 100 mg/ml konzentrierte bSA<br />

Lösung getrocknet bei 60°C und 10% Feuchte in unbewegter Luft. Unter Annahme einer<br />

sphärischen Geometrie mit einem Radius von 500 µm betrug das Hohlraumvolumen 0,166 µl und<br />

die Oberfläche wird durch Entstehen einer 1,46 mm 2 großen inneren Oberfläche auf insgesamt<br />

4,60 mm 2 vergrößert. Hieraus resultierte eine Zunahme der Trocknungsgeschwindigkeit im zweiten<br />

Trocknungsabschnitt auf Werte annähernd so groß wie am kritischen Punkt. Formulierungen von<br />

bSA mit Trehalose in einer bSA-Trehalose (1:1) Mischung mit einer Konzentration von 100 mg/ml<br />

zeigten diesen Anstieg der Trocknungsgeschwindigkeit in der zweiten Phase nicht. Weder konnten<br />

auf den REM-Bildern Löcher in der Kruste noch eine hohle Partikelstruktur gefunden werden. Im<br />

Vergleich mit den beiden Reinsubstanzen bei gleicher Konzentration kam es mit zunehmendem<br />

Trehalosegehalt zu einer Verlängerung des ersten Trocknungsabschnittes. Bei 60°C und 5%<br />

relativer Feuchte in unbewegter Luft nahm der Anteil an verdampftem Lösungsmittel von 65,1%<br />

bei der reinen bSA-Lösung auf 87,8% für die Mischung und 85,6% für die reine Trehaloselösung<br />

zu. Die Trocknungsgeschwindigkeit in unbewegter Luft sank dabei von 1,21 µg/s/mm 2 für reines<br />

bSA auf 0,91 µg/s/mm 2 für reine Trehalose ab. Die bSA-Trehalose-Mischung hatte unter gleichen


CHAPTER 7 ZUSAMMENFASSUNG 236<br />

Bedingungen eine anfängliche Trocknungsgeschwindigkeit von 1,0 µg/s/mm 2 . Unter<br />

Vernachlässigung der reinen bSA-Lösung sank die Trocknungsgeschwindigkeit nach dem<br />

kritischen Punkt mit zunehmendem Trehalosegehalt langsamer ab. Die morphologische<br />

Übereinstimmung zwischen Levitationspartikeln und sprühgetrocknetem Produkt war sehr gut.<br />

Wegen ihrer hohlen Struktur war die Dichte der levitierten Partikel mit 0,62 g/cm 2 (60°C, 5%<br />

Feuchte, unbewegte Luft) um etwa 0,3 Einheiten kleiner als die der bSA-Trehalose Mischung.<br />

Diese hingegen unterschieden sich nur durch minimal kleinere Werte von der Dichte reiner<br />

Trehalosepartikel. Untersuchungen der Grenzflächenaktivität von bSA-Lösungen und ihrer<br />

Mischungen mit 200 mg/ml Trehalose mittels Blasendrucktensiometrie zeigten, dass der Zusatz von<br />

Trehalose die dynamische Oberflächenspannung einer 2,0 mmol/l bSA Lösung um 3,67 mN/m und<br />

die einer 4,0 mmol/l bSA-Lösung sogar um 4,42 mN/m erhöht (Blasenlebenszeit 1,0 s). Berechnung<br />

der bSA-Konzentration in der Oberfläche unter Verwendung der Gibbs-Gleichung ergab für eine<br />

4,0 mmol/l bSA-Lösung durch Trehalosezusatz (200 mg/ml) eine Reduktion von 87 mg/m 2 auf<br />

36 mg/m 2 . Die Fläche eines bSA-Moleküls in einem Monolayer auf der Tropfenoberfläche wurde<br />

zu 100 nm 2 bestimmt.<br />

Reine Katalaselösungen in 50 mM Tris Puffer bei pH 8,0 getrocknet im Ultraschall-<br />

Levitator führten zu Partikeln mit Löchern von durchschnittlich 45,1% ± 4,3% des horizontalen<br />

Durchmessers (60°C; 5% Feuchte; unbewegte Luft). Die Kruste hatte dabei eine Breite von 18,3%<br />

des Partikelradius. Ausgehend von einer sphärischen Geometrie mit einem Radius von 500 µm<br />

ergab sich ein Hohlraumvolumen von 0,286 µl mit einer Oberflächenvergrößerung von 2,10 mm 2<br />

auf eine Gesamtfläche von 5,24 mm 2 . Die Oberflächenvergrößerung führte, wie auch beim bSA, zu<br />

einem zwischenzeitlichen Anstieg der Trocknungsgeschwindigkeit innerhalb des zweiten<br />

Trocknungsabschnittes. Die systematische Untersuchung einer Mischungsreihe aus Katalase und<br />

Trehalose zeigte, dass ab einem Katalase-Trehalose Verhältnis von 6 zu 4 massive Partikel ohne<br />

Hohlraum und Löcher in der Kruste entstehen und ein Anstieg der Trocknungsgeschwindigkeit im<br />

zweiten Trocknungsabschnitt nicht mehr beobachtet werden kann. Mit zunehmendem<br />

Trehalosegehalt nahm die Dauer des ersten Trocknungsabschnittes zu. Ebenso erhöhte sich auch der<br />

Anteil an verdampftem Lösungsmittel von 61,9% für die reine Katalaselösung auf 85,5% für die<br />

Katalase-Trehalose (2:8) Mischung (60°C; 5% Feuchte; unbewegte Luft). Die<br />

Trocknungsgeschwindigkeit sank unter den obigen Bedingungen von 0,95 µg/s/mm 2 für reine<br />

Katalase-Lösung auf 0,91 µg/s/mm 2 für die Katalase-Trehalose (2:8) Mischung ab. Die reine<br />

Trehaloselösung in Tris Puffer hatte eine Trocknungsgeschwindigkeit von 0,89 µg/s/mm 2 . Mit<br />

zunehmendem Trehalosegehalt nahm bei den Katalase-Lösungen die Trocknungsgeschwindigkeit


CHAPTER 7 ZUSAMMENFASSUNG 237<br />

nach dem kritischen Punkt langsamer ab. Es konnte eine sehr gute morphologische Ähnlichkeit zum<br />

sprühgetrockneten Produkt festgestellt werden. Die Dichte der levitierten Partikel nahm von<br />

0,43 g/cm 3 für reine Katalase mit zunehmendem Trehalosegehalt auf 0,84 g/cm 3 für die Katalase-<br />

Trehalose (2:8) Mischung zu (60°C, 5% Feuchte, unbewegte Luft). Mittels Blasendrucktensiometrie<br />

konnte gezeigt werden, dass es bei einer 0,40 mmol/l Katalaselösung (äquivalent zu 100 mg/ml)<br />

durch Zusatz von 200 mg/ml Trehalose zur Erhöhung der dynamischen Oberflächenspannung um<br />

3,9 mN/m bei einer Blasenlebenszeit von 1,0 s kommt. Die Grenzflächenkonzentration von<br />

Katalase reduzierte sich dabei von 575 mg/mm 2 auf 400 mg/mm 2 . Untersuchungen zur Restaktivität<br />

von Katalase in akustisch levitierten Partikel ergaben wegen der vergleichsweise großen Partikel<br />

und langen Trocknungszeit bei 60°C mit Ergebnissen der Sprühtrocknung keine Übereinstimmung.<br />

Jedoch konnte der stabilisierende Einfluss der Trehalose auf die Aktivität der Katalase während der<br />

Trocknung nachgewiesen werden. Der Zusatz von 20% Trehalose im Festst<strong>of</strong>fgehalt steigerte die<br />

Restaktivität der bei 60°C und 5% Feuchte in unbewegter Luft getrockneten Partikel von 87,7% auf<br />

95,5%. Das molare Verhältnis dieser Lösung aus Katalase zu Trehalose war 1:183. Eine durch<br />

zunehmende Luftfeuchte verursachte Zunahme der Trocknungszeit führte zu leichtem<br />

Aktivitätsverlust bei allen Formulierungen.<br />

Zusammenfassend stellt der akustische Levitator ein viel versprechendes Werkzeug zur<br />

Untersuchung des Trocknungsverhaltens einzelner Lösungsmittel-, Suspensions- und<br />

Lösungstropfen dar. Obwohl Einflüsse der stehenden akustischen Welle auf Massenübergang und<br />

Partikelbildung demonstriert werden konnten, war es möglich den Verlauf der Tropfengröße mit der<br />

Zeit im Levitationssystem mittels mathematischer Modelle vorherzusagen und eine gute<br />

morphologische Übereinstimmung mit den sprühgetrockneten Produkten gleicher<br />

Zusammensetzung zu zeigen. Diese Ergebnisse und die Möglichkeit fast jede mögliche<br />

Trocknungsbedingung innerhalb eines bauartbedingten Temperaturbereiches einzustellen, machen<br />

den Ultraschall-Levitator zum interessanten Werkzeug für Voruntersuchungen in der<br />

Sprühtrocknungsforschung, vor allem dann, wenn nur eine geringe Menge Substanz, zum Beispiel<br />

aus Preisgründen, verfügbar ist. Der Umbau des Systems um die Möglichkeit zu erhalten, den<br />

Ultraschallw<strong>and</strong>ler zu kühlen, sowie die Injektion und Dosierung der Tropfen mittels eines<br />

Mikropumpsystems würde die Möglichkeit eröffnen Trocknungsversuche auch bei Temperaturen<br />

über 70°C und mit kleinerer Anfangstropfengröße durchzuführen. Somit könnte eine noch bessere<br />

Vergleichbarkeit zur Sprühtrocknung geschaffen werden. Die Verwendung einer Infrarot-<br />

Thermokamera würde zusätzlich eine Bestimmung der Oberflächentemperatur der Lösungsmittelund<br />

Lösungstropfen und damit die Überprüfung mathematischer Modelle zur Berechnung


CHAPTER 7 ZUSAMMENFASSUNG 238<br />

zugänglich machen. Weitere Einsatzgebiete zeigen sich auch auf dem pharmazeutischen Sektor.<br />

Eine Verwendung des Levitationssystem zur Untersuchung von Verteilungs-, Mikroverkapselungsoder<br />

Polymerisationsprozessen direkt im levitierten Tropfen, sowie für die Untersuchung der<br />

Qualität von Verkapselungsformulierungen flüchtiger Substanzen ist durchaus denkbar.


Annex<br />

Saturation pressure <strong>of</strong> water vapour<br />

ANNEX 239<br />

For water vapour at atmospheric pressure the saturation pressure, p saturation , in mbar<br />

(1mbar = 10 2 N/m 2 ) for given temperature TS in degrees Celcius is calculated by<br />

[Seaver et al. 1989]<br />

Equation A.1 p a + T ⋅ [ a + T ⋅ ( a + T ⋅{<br />

a + T ⋅ [ a + T ⋅ ( a + a ⋅T<br />

) ] } ) ]<br />

saturation = 0 S 1 S 2 S 3 S 4 S 5 6<br />

Table A.1: Parameters used to calculate the saturation pressure <strong>of</strong> water vapour [Seaver et al. 1989].<br />

a0<br />

a1<br />

a2<br />

a3<br />

a4<br />

a5<br />

a6<br />

Saturation pressure <strong>of</strong> different liquids<br />

6.107799961<br />

4.436518521 x 10 -1<br />

1.428945805 x 10 -2<br />

2.650648731 x 10 -4<br />

3.031240396 x 10 -6<br />

2.034080948 x 10 -8<br />

6.136820929 x 10 -11<br />

The saturation pressure in bar (1 bar = 10 -5 N/m 2 ) for the liquids listed in Table A.2 is given as a<br />

function <strong>of</strong> temperature TS in Kelvin by [Reid et al. 1987]<br />

Equation A.2<br />

p = 0 . 001333224 ⋅ e<br />

saturation<br />

⎛ B ⎞<br />

⎜ A−<br />

⎟<br />

⎝ TS<br />

+ C ⎠<br />

Table A.2: Liquid parameters used to calculate the saturation pressure <strong>of</strong> vapour [Reid et al. 1987].<br />

Liquid A B C<br />

methanol 18.5875 3626.55 -34.29<br />

ethanol 18.9919 3803.98 -41.68<br />

2-Propanol 17.5439 3166.38 -80.15<br />

n-heptane 15.8738 2911.32 -56.51<br />

n-octane 15.9426 3120.29 -63.63<br />

n-decane 16.0114 3456.80 -78.67<br />

S


240 ANNEX A<br />

Latent heat <strong>of</strong> vaporisation <strong>of</strong> water<br />

The latent heat <strong>of</strong> vaporisation <strong>of</strong> water, hv , in cal/g at given Temperature TS in Kelvin is calculated<br />

by [Seaver et al. 1989]<br />

−3<br />

Equation A.3 h 595.<br />

3 + T ⋅ ( 0.<br />

483 + T ⋅1.<br />

2 ⋅10<br />

)<br />

Thermal conductivity <strong>of</strong> air<br />

v = S<br />

S<br />

The thermal conductivity <strong>of</strong> air, ka , in cal/cm/°C/s at given temperature TS in degrees Celcius is<br />

calculated by [Seaver et al. 1989]<br />

−5 −7<br />

Equation A.4 ka = 5 . 8⋅10<br />

+ 1.<br />

6 ⋅10<br />

⋅TS<br />

Binary diffusion coefficient<br />

The binary diffusion coefficient, DAB , in cm 2 /s at given temperature TS in degrees Celcius <strong>and</strong><br />

pressure, p , in atm is calculated by<br />

Equation A.5<br />

D<br />

AB<br />

1.<br />

00 ⋅10<br />

=<br />

p ⋅<br />

−3<br />

⋅T<br />

1.<br />

75<br />

( 1/<br />

M + 1/<br />

M )<br />

[ ( ) ( ) ] 2<br />

S<br />

A<br />

1/<br />

3<br />

1/<br />

3<br />

∑ υ + ∑ υ<br />

A i<br />

MA <strong>and</strong> MB are the molecular weights <strong>of</strong> the substance, whereas ∑ υ<br />

A i <strong>and</strong> ∑ υ<br />

B i are the summed<br />

special diffusion parameters over atoms, groups <strong>and</strong> structural features <strong>of</strong> the diffusion species<br />

[Fuller et al. 1966].<br />

Table A.3: Molecular weight <strong>and</strong> summed diffusion parameters for air / water <strong>and</strong> air / ethanol systems<br />

[Fuller et al. 1966].<br />

System MA [ g/mol ] MB [ g/mol ] ∑ υ<br />

A i ∑ υ<br />

B i<br />

Air / water 28.96 18.03 20.1 12.7<br />

Air / ethanol 28.96 46.07 20.1 50.4<br />

B i<br />

B


REFERENCES 241<br />

References<br />

[Abramzon 1989] Abramzon, B. <strong>and</strong> Sirignano, W. A. (1989). "Droplet Vaporisation<br />

Model for Spray Combustion Calculations."<br />

Int. J. Heat Mass Transfer 32(9): 1605-1618.<br />

[Adjei 1994] Adjei, A. <strong>and</strong> Gupta, P. (1994). "Pulmonary Delivery <strong>of</strong> Therapeutic<br />

Peptides <strong>and</strong> Proteins."<br />

Journal <strong>of</strong> Controlled Release 29(3): 361-373.<br />

[Adler 1999] Adler, M. <strong>and</strong> Lee, G. (1999). "Stability <strong>and</strong> surface activity <strong>of</strong> lactate<br />

dehydrogenase in spray-dried trehalose."<br />

J Pharm Sci 88(2): 199-208.<br />

[Adler et al. 2000] Adler, M., Unger, M. <strong>and</strong> Lee, G. (2000). "Surface composition <strong>of</strong><br />

spray-dried particles <strong>of</strong> bovine serum albumin/trehalose/surfactant."<br />

Pharm Res 17(7): 863-70.<br />

[Alex<strong>and</strong>er 1978] Alex<strong>and</strong>er, K. (1978). Factors governing surface morphology in the<br />

spray <strong>drying</strong> <strong>of</strong> foods.<br />

Berkley, University <strong>of</strong> California, USA. Ph.D. Thesis.<br />

[Almer et al. 2002] Almer, L. O., Wollmer, P., Jonson, B. <strong>and</strong> Troedsson, A. (2002).<br />

"Insulin inhalation with absorption enhancer at meal-times results in<br />

almost normal postpr<strong>and</strong>ial insulin pr<strong>of</strong>iles."<br />

Clin Physiol Funct Imaging 22(3): 218-21.<br />

[Andya et al. 1999] Andya, J. D., Maa, Y. F., Costantino, H. R., Nguyen, P. A.,<br />

Dasovich, N., Sweeney, T. D., Hsu, C. C. <strong>and</strong> Shire, S. J. (1999). "The<br />

effect <strong>of</strong> formulation excipients on <strong>protein</strong> stability <strong>and</strong> aerosol<br />

performance <strong>of</strong> spray-dried powders <strong>of</strong> a recombinant humanized anti-<br />

IgE monoclonal antibody." Pharm Res 16(3): 350-8.<br />

[Arakawa et al. 2001] Arakawa, T., Prestrelski, S. J., Kenney, W. C. <strong>and</strong> Carpenter, J. F.<br />

(2001). "Factors affecting the short-term <strong>and</strong> long-term stabilities <strong>of</strong><br />

<strong><strong>protein</strong>s</strong>." Advances Drug Delivery Reviews 46(1-3): 307-326.<br />

[Baehr 1998] Baehr, H. D. <strong>and</strong> Stephan, K. (1998). Wärme- und St<strong>of</strong>fübertragung.<br />

Berlin, Heidelberg (Germany), Springer Verlag<br />

[Banga 1995] Banga, A. K. (1995). Therapeutic Peptides <strong>and</strong> Proteins. Lancester,<br />

Technomic Publishing Company<br />

[Barmatz 1985] Barmatz, M. <strong>and</strong> Collas, P. (1985). "Acoustic radiation potential on a<br />

sphere in plane, cylindrical <strong>and</strong> spherical st<strong>and</strong>ing wave fields."<br />

Journal <strong>of</strong> Acoustic Society <strong>of</strong> America 77: 926-945.


242 REFERENCES<br />

[Beers 1952] Beers, R. F. J. <strong>and</strong> Sizer, I. W. (1952). "A spectrophotometric method<br />

for measuring the breakdown <strong>of</strong> hydrogen peroxide by catalase."<br />

Journal <strong>of</strong> Biological Chemistry 195: 133-140.<br />

[Bird 1960] Bird, R. B. (1960). Transport Phenomena. New York, John Wiley<br />

[Boon et al. 2001] Boon, E. M., Downs, A. <strong>and</strong> Marcey, D. (2001). Catalase:<br />

Oxidoreductase. Gambier; Ohio (USA), Kenyon College<br />

[Bosnjakovic 1997] Bosnjakovic, F. (1997). Technische Thermodynamik - Teil 2.<br />

Darmstadt, Germany, Dr. Dietrich Steinkopff Verlag GmbH & Co. KG<br />

[Bosquillon et al. 2004] Bosquillon, C., Rouxhet, P. G., Ahimou, F., Simon, D., Culot, C.,<br />

Preat, V. <strong>and</strong> Vanbever, R. (2004). "Aerosolization properties, surface<br />

composition <strong>and</strong> physical state <strong>of</strong> spray-dried <strong>protein</strong> powders."<br />

J Control Release 99(3): 357-67.<br />

[Bronkhorst 2005] Bronkhorst (2005). http://www.fluidat.com.<br />

[Buchi 2005] Buchi (2005). http://www.buchi.com.<br />

[Burdukov 1963] Burdukov, A. P. <strong>and</strong> Nakoryakov, V. E. (1963). "Some mass transfer<br />

problems in an ultrasonic field."<br />

Kinetika Goreniya Iskop. Topliv: 97-103.<br />

[Burdukov 1967] Burdukov, A. P. <strong>and</strong> Nakoryakov, V. E. (1967). "Effect <strong>of</strong> vibrations on<br />

mass transfer from a sphere at large Pr<strong>and</strong>tl numbers."<br />

Prikladnaya Mekhanika Tekhnicheskaya Fizika (3): 158-160.<br />

[Caessens et al. 1999] Caessens, P. W., De Jongh, H. H., Norde, W. <strong>and</strong> Gruppen, H. (1999).<br />

"The adsorption-induced secondary structure <strong>of</strong> beta-casein <strong>and</strong> <strong>of</strong><br />

distinct parts <strong>of</strong> its sequence in relation to foam <strong>and</strong> emulsion<br />

properties." Biochem Biophys Acta 1430(1): 73-83.<br />

[Carpenter et al. 1997] Carpenter, J. F., Pikal, M. J., Chang, B. S. <strong>and</strong> R<strong>and</strong>olph, T. W. (1997).<br />

"Rational design <strong>of</strong> stable lyophilized <strong>protein</strong> <strong>formulations</strong>: some<br />

practical advice." Pharm Res 14(8): 969-75.<br />

[Carrasquillo et al. 2001] Carrasquillo, K. G., Stanley, A. M., Aponte-Carro, J. C., De Jesus, P.,<br />

Costantino, H. R., Bosques, C. J. <strong>and</strong> Griebenow, K. (2001). "Nonaqueous<br />

encapsulation <strong>of</strong> excipient-stabilized spray-freeze dried BSA<br />

into poly(lactide-co-glycolide) microspheres results in release <strong>of</strong> native<br />

<strong>protein</strong>." J Control Release 76(3): 199-208.<br />

[Charlesworth 1960] Charlesworth, D. H. <strong>and</strong> Marshall, W. R. (1960). "Evaporation from<br />

drops containing dissolved solids." AIChE Journal 6(1): 9-19.<br />

[Chen et al. 1998] Chen, P., Prokop, R. M., Susnar, S. S. <strong>and</strong> Neumann, A. W. (1998).<br />

Interfacial tension <strong>of</strong> <strong>protein</strong> solutions using axisymmetric drop shape<br />

analysis. New York, Elsevier


REFERENCES 243<br />

[Cheong et al. 1986] Cheong, H. W., Jeffreys, G. V. <strong>and</strong> Mumford, C. J. (1986). "A receding<br />

interface model for the <strong>drying</strong> <strong>of</strong> slurry <strong>droplet</strong>s."<br />

AIChE Journal 32(8): 1334-1346.<br />

[Clarkson et al. 1999] Clarkson, J. R., Cui, Z. F. <strong>and</strong> Darton, R. C. (1999). "Protein<br />

Denaturation in Foam." J Colloid Interface Sci 215(2): 333-338.<br />

[Cohn et al. 1946] Cohn, E. J., Strong, L. E., Hughes, W. L., Mulford, D. J.,<br />

Ashworth, M. M. <strong>and</strong> Taylor, H. L. (1946). "Preparation <strong>and</strong> properties<br />

<strong>of</strong> serum <strong>and</strong> plasma <strong><strong>protein</strong>s</strong>. IV. A system for the separation into<br />

fractions <strong>of</strong> the <strong>protein</strong> <strong>and</strong> lipo<strong>protein</strong> components <strong>of</strong> biological tissues<br />

<strong>and</strong> fluids."<br />

Journal <strong>of</strong> the American Chemical Society 68: 459-475.<br />

[Constantino et al. 1998] Costantino, H. R., Andya, J. D., Nguyen, P. A., Dasovich, N.,<br />

Sweeney, T. D., Shire, S. J., Hsu, C. C. <strong>and</strong> Maa, Y. F. (1998). "Effect<br />

<strong>of</strong> mannitol crystallization on the stability <strong>and</strong> aerosol performance <strong>of</strong> a<br />

spray-dried pharmaceutical <strong>protein</strong>, recombinant humanized anti-IgE<br />

monoclonal antibody."<br />

J Pharm Sci 87(11): 1406-11.<br />

[Crowe et al. 1990] Crowe, J. H., Carpenter, J. F., Crowe, L. M. <strong>and</strong> Andchordoguy, T.<br />

(1990). "Are freezing <strong>and</strong> dehydration similar stress factors? A<br />

comparison <strong>of</strong> modes <strong>of</strong> interaction <strong>of</strong> stabilizing solutes with<br />

biomolecules." Cryobiology 27: 219-231.<br />

[de Jongh et al. 2004] de Jongh, H. H., Kosters, H. A., Kudryashova, E., Meinders, M. B.,<br />

Tr<strong>of</strong>imova, D. <strong>and</strong> Wierenga, P. A. (2004). "Protein adsorption at airwater<br />

interfaces: a combination <strong>of</strong> details."<br />

Biopolymers 74(1-2): 131-5.<br />

[Earle 2004] Earle, R. L. <strong>and</strong> Earle, M. D. (2004).<br />

Unit operations in food processing. Pergamon Press.<br />

[Eberhardt 1999] Eberhardt, R. <strong>and</strong> Neidhart, B. (1999). "Acoustic levitation device for<br />

sample pretreatment in microanalysis <strong>and</strong> trace analysis."<br />

Fresenius Journal <strong>of</strong> Analytical Chemistry 6(475-479).<br />

[El-Sayed et al. 1990] El-Sayed, T. M., Wallack, D. A. <strong>and</strong> King, C. J. (1990). "Change in<br />

particle morphology during the <strong>drying</strong> <strong>of</strong> drops <strong>of</strong> carbohydrate<br />

solutions <strong>and</strong> food liquids."<br />

Industrial & Engineering Chemistry 29(12): 2346-2354.<br />

[Elbein 1974] Elbein, A. D. (1974). Metabolism <strong>of</strong> alpha, alpha-trehalose.<br />

New York, Academic Press<br />

[Elperin 1995] Elperin, T. <strong>and</strong> Krasovitov, B. (1995). "Evaporation <strong>of</strong> liquids<br />

containing small solid particles."<br />

Int. J. Heat Mass Transfer 38(12): 2259-2267.


244 REFERENCES<br />

[Elversson et al. 2003] Elversson, J., Millqvist-Fureby, A., Alderborn, G. <strong>and</strong> El<strong>of</strong>sson, U.<br />

(2003). "Droplet <strong>and</strong> Particle Size Relationship <strong>and</strong> Shell Thickness <strong>of</strong><br />

Inhalable Lactose Particles during Spray Drying." J Pharm Sci 92(4):<br />

900-909.<br />

[Elversson 2005 a] Elversson, J. <strong>and</strong> Millqvist-Fureby, A. (2005). "Particle size <strong>and</strong> density<br />

in spray <strong>drying</strong>-effects <strong>of</strong> carbohydrate properties."<br />

J Pharm Sci 94(9): 2049-60.<br />

[Elversson 2005 b] Elversson, J. <strong>and</strong> Millqvist-Fureby, A. (2005). "Aqueous two-phase<br />

systems as a formulation concept for spray-dried <strong>protein</strong>."<br />

Int J Pharm 294(1-2): 73-87.<br />

[EngineeringToolbox 2005] EngineeringToolBox (2005). http://www.engineeringtoolbox.com.<br />

[Engineersedge 2005] Engineersedge (2005). Mechanical Engineering, Manufacturing <strong>and</strong><br />

Design Database <strong>and</strong> Tools, http://www.engineersedge.com.<br />

[Faeth 1977] Faeth, G. M. (1977). "Current Status <strong>of</strong> Droplet <strong>and</strong> Liquid<br />

Combustion." Prog. Energy Combustion Sci. 3: 191-224.<br />

[Farid 2003] Farid, M. (2003). "A new approach to modelling <strong>of</strong> single <strong>droplet</strong><br />

<strong>drying</strong>." Chemical Engineering Science 58(13): 2985-2993.<br />

[Fitzner 2003] Fitzner, M. (2003). Chemical <strong>and</strong> physiochemical stability <strong>of</strong> spraydried<br />

rhIL-11 <strong>formulations</strong>. Institute for Pharmaceutical Science.<br />

Erlangen, Germany, Friedrich-Alex<strong>and</strong>er University. Ph.D. thesis.<br />

[Forbes et al. 1998] Forbes, R. T., Davis, K. G., Hindle, M., Clarke, J. G. <strong>and</strong> Maas, J.<br />

(1998). "Water vapour sorption studies on the physical stability <strong>of</strong> a<br />

series <strong>of</strong> spray-dried <strong>protein</strong>/sugar powders for inhalation."<br />

J Pharm Sci 87(11): 1316-21.<br />

[Friedli 1996] Friedli, G. L. (1996). Interaction <strong>of</strong> deamidated soluble wheat <strong>protein</strong><br />

(SWP) with other food <strong><strong>protein</strong>s</strong> <strong>and</strong> metals.<br />

University <strong>of</strong> Surrey; Great Britain. Ph.D. thesis.<br />

[Froessling 1927] Froessling, N. (1927). "The Evaporation <strong>of</strong> Falling Drops."<br />

Gerl<strong>and</strong>s Beitraege zur Geophysik 52: 170-216.<br />

[Frohn 2000] Frohn, A. <strong>and</strong> Roth, N. (2000). Dynamics <strong>of</strong> Droplets.<br />

Berlin, Heidelberg, Springer<br />

[Frost 1993] Frost, A. E. <strong>and</strong> Seaver, M. (1993). "Delayed appearance <strong>of</strong> the liquidcondensed<br />

phase in 1-octadecanol films on levitated waterdrops."<br />

J. Chem. Phys. 100(4): 3268-3275.<br />

[Fuchs 1959] Fuchs, N. A. (1959). Evaporation <strong>and</strong> Droplet Growth in Gaseous<br />

Media. London, Pergamon Press.


REFERENCES 245<br />

[Fuhrling 2004] Fuhrling, C. (2004). Interactions between <strong><strong>protein</strong>s</strong>, sugars <strong>and</strong><br />

surfactants - dynamic studies on absorption at interfaces.<br />

Pharmaceutical Technology. Erlangen, Friedrich-Alex<strong>and</strong>er University.<br />

Ph.D. thesis.<br />

[Fuller et al. 1966] Fuller, E. N., Schettler, P. D. <strong>and</strong> Giddings, J. C. (1966). "A new<br />

method for the prediction <strong>of</strong> binary gas-phase diffusion coefficients."<br />

Industrial & Engineering Chemistry 58(5): 19-27.<br />

[Futura et al. 1982] Futura, T., Okazaki, M., Toei, R. <strong>and</strong> Crosby, E. J. (1982). Formation<br />

<strong>of</strong> Crystals on the Surface <strong>of</strong> Non-Supported Droplet in Drying.<br />

New York, USA, Hemisphere-McGraw-Hill.<br />

[Geankoplis 1993] Geankoplis, C. J. (1993). Transport processes <strong>and</strong> unit operations.<br />

New Jersey, USA, P T R Prentice-Hall Inc.<br />

[Gil et al. 1996] Gil, A. M., Belton, P. S. <strong>and</strong> Felix, V. (1996). "Spectroscopic studies <strong>of</strong><br />

solid alpha-alpha trehalose."<br />

Spectrochimica Acta Part A 52(12): 1649-1659.<br />

[Goolcharran et al. 2000] Goolcharran, C., Khossravi, M. <strong>and</strong> Borchardt, R. T. (2000). Chemical<br />

Pathways <strong>of</strong> Protein <strong>and</strong> Peptide Degradation.<br />

London, Taylor & Francis<br />

[Gopinath 1993] Gopinath, A. <strong>and</strong> Mills, A. F. (1993). "Convective heat transfer from a<br />

sphere due to acoustic streaming."<br />

Journal <strong>of</strong> Heat Transfer 115(2): 332-341.<br />

[Grassmann et al. 1997] Grassmann, P., Widmer, F. <strong>and</strong> Sinn, H. (1997). Einführung in die<br />

thermische Verfahrenstechnik. Berlin (Germany),<br />

Walter de Gruyter & Co.<br />

[Groenewold et al.2002] Groenewold, C., Möser, C., Groenewold, H. <strong>and</strong> Tsotsas, E. (2002).<br />

"Determination <strong>of</strong> <strong>Single</strong>-Particle Drying Kinetics in an Acoustic<br />

Levitator." Chemical Engineering Journal 86(1-2): 217-222.<br />

[Haas 2002] Haas, U. (2002). Physik fuer Pharmazeuten. Stuttgart (Germany),<br />

Wissenschaftliche Verlagsgesellschaft mbH<br />

[Hampe et al. 2001] Hampe, M. J., Buhn, J., Geisbusch, D., Heller, O., Kanzamar, M. <strong>and</strong><br />

Kayembe, J. (2001). Höhere St<strong>of</strong>fübertragung. Darmstadt (Germany),<br />

Technische Universität Darmstadt, Thermische Verfahrenstechnik<br />

[Hartkins 1930] Harkins, W. D. <strong>and</strong> Jordan, H. F. (1930). "A method for the<br />

determination <strong>of</strong> surface tension from the maximum pull on a ring."<br />

Journal <strong>of</strong> the American Chemical Society 52: 1751-1772.<br />

[Hiemenz 1986] Hiemenz, P. C. (1986). Principles <strong>of</strong> colloid <strong>and</strong> surface chemistry.<br />

New York, USA, Marcel Dekker, Inc.


246 REFERENCES<br />

[Hirayama et al. 1990] Hirayama, K., Akashi, S., Furuya, M. <strong>and</strong> Fukuhara, K. (1990). "Rapid<br />

confirmation <strong>and</strong> revision <strong>of</strong> the primary structure <strong>of</strong> bovine serum<br />

albumin by ESIMS <strong>and</strong> Frit-FAB LC/MS." Biochemical <strong>and</strong><br />

Biophysical Research Communications 173(2): 639-646.<br />

[Hoekstra et al. 1997] Hoekstra, F. A., Wolkers, W. F., Buitink, J., Golovina, E. A.,<br />

Crowe, J. H. <strong>and</strong> Crowe, L. (1997). "Membrane stabilization in the dry<br />

state." Comp. Biochem. Physiol. 117(3): 335-341.<br />

[Horbett 1992] Horbett, T. A. (1992). Adsorption <strong>of</strong> <strong><strong>protein</strong>s</strong> <strong>and</strong> peptides at interfaces.<br />

New York, Plenum Press<br />

[ISB 2005] ISB (2005). Information Secretary <strong>of</strong> Biotechnology. www.i-s-b.org.<br />

[Izutsu et al. 2004] Izutsu, K., Aoyagi, N. <strong>and</strong> Kojima, S. (2004). "Protection <strong>of</strong> <strong>protein</strong><br />

secondary structure by saccharides <strong>of</strong> different molecular weights<br />

during freeze-<strong>drying</strong>." Chem Pharm Bull (Tokyo) 52(2): 199-203.<br />

[Jiang 1998] Jiang, S. <strong>and</strong> Nail, S. L. (1998). "Effect <strong>of</strong> process conditions on<br />

recovery <strong>of</strong> <strong>protein</strong> activity after freezing <strong>and</strong> freeze-<strong>drying</strong>."<br />

Eur J Pharm Biopharm 45(3): 249-57.<br />

[Joos 1999] Joos, P. (1999). Dynamic surface phenomena.<br />

Utrecht, Netherl<strong>and</strong>s, VSP<br />

[Kastner 2001] Kastner, O. (2001). Theoretische und experimentelle Untersuchungen<br />

zum St<strong>of</strong>fübergang von Einzeltropfen in einem akustischen<br />

Rohrlevitator. Institute <strong>of</strong> Fluid Mechanics. Friedrich Alex<strong>and</strong>er<br />

University, Erlangen, Germany.<br />

[Kastner et al. 2001] Kastner, O., Brenn, G., Rensink, D. <strong>and</strong> Tropea, C. (2001). "The<br />

Acoustic Levitator - A Novel Device for Determining the Drying<br />

Kinetics <strong>of</strong> <strong>Single</strong> Droplets." Chem. Eng. Technol. 24(4): 335-339.<br />

[Katchalski et al. 1957] Katchalski, E., Benjamin, G. S. <strong>and</strong> Gross, V. (1957). "The availability<br />

<strong>of</strong> the disulfide bonds <strong>of</strong> human <strong>and</strong> bovine serum albumin <strong>and</strong> <strong>of</strong><br />

bovine gamma-globulin to reduction by thioglycolic acid."<br />

Journal <strong>of</strong> the American Chemical Society 79: 4096-4099.<br />

[Kaushik 1998] Kaushik, J. K. <strong>and</strong> Bhat, R. (1998). "Thermal stability <strong>of</strong> <strong><strong>protein</strong>s</strong> in<br />

aqueous polyol solutions: role <strong>of</strong> the surface tension <strong>of</strong> the water in the<br />

stabilizing effect <strong>of</strong> polyols."<br />

Journal <strong>of</strong> Physical Chemistry 102(36): 7058-7066.<br />

[Kaushik 2003] Kaushik, J. K. <strong>and</strong> Bhat, R. (2003). "Why is trehalose an exceptional<br />

<strong>protein</strong> stabilizer? An analysis <strong>of</strong> thermal stability <strong>of</strong> <strong><strong>protein</strong>s</strong> in the<br />

presence <strong>of</strong> the compatible osmolyte trehalose."<br />

Journal <strong>of</strong> Biological Chemistry 29(18): 26458-26465.


REFERENCES 247<br />

[Kavouras 2003] Kavouras, A. <strong>and</strong> Krammer, G. (2003). "Ultrasonic levitation for the<br />

examination <strong>of</strong> gas/solid reactions."<br />

Review <strong>of</strong> Scientific Instruments 74(10): 4468-4473.<br />

[Kawahara et al. 2000] Kawahara, N., Yarin, A. L., Brenn, G., Kastner, O. <strong>and</strong> Durst, F.<br />

(2000). "Effect <strong>of</strong> the acoustic streaming on the mass transfer from a<br />

sublimating sphere." Physic <strong>of</strong> Fluids 12: 912-923.<br />

[King 1934] King, L. V. (1934). "On the acoustic radiation pressure on spheres."<br />

Proceedings <strong>of</strong> the Royal Society (London) A 147: 212-240.<br />

[Kneule 1975] Kneule, F. (1975). Das Trocknen. Aarau (Germany), Sauerländer AG<br />

[Krishna 2000] Krishna, R. <strong>and</strong> Wesselingh, J. A. (2000). Mass transfer in<br />

multicomponent mixtures. Delft, Delft Univ. Press<br />

[Kruss 2005] Kruss (2005). http://www.kruss.info.<br />

[Labuza et al. 1992] Labuza, T., Nelson, K. <strong>and</strong> Coppersmith, C. (1992). Glass transition<br />

temperature <strong>of</strong> food systems. Department <strong>of</strong> Food Science <strong>and</strong><br />

Nutrition. University <strong>of</strong> Minnesota, USA, http://fscn.che.umn.edu.<br />

[L<strong>and</strong>strom et al. 2000] L<strong>and</strong>strom, K., Alsins, J. <strong>and</strong> Bergebstahl, B. (2000). "Competitive<br />

<strong>protein</strong> adsorption between bovine serum albumin <strong>and</strong> betalactoglobulin<br />

during spray-<strong>drying</strong>." Food Hydrocolloids 14: 75-82.<br />

[Lauda 2000] Lauda (2000). Instruction manual Lauda MPT 2.<br />

Lauda-Konigsh<strong>of</strong>en, Germany<br />

[Lauda 2005] Lauda (2005). http://www.lauda.de.<br />

[Law 1982] Law, C. K. (1982). "Recent advances in <strong>droplet</strong> vaporisation <strong>and</strong><br />

combustion." Prog. Energy Combustion Sci. 8(3): 171-201.<br />

[Lee 1990] Lee, C. P. <strong>and</strong> Wang, T. G. (1990). "Outer acoustic streaming."<br />

Journal <strong>of</strong> Acoustical Society <strong>of</strong> America 88: 2367-2375.<br />

[Lee 2002] Lee, G. (2002). "Spray-<strong>drying</strong> <strong>of</strong> <strong><strong>protein</strong>s</strong>."<br />

Pharm Biotechnol 13: 135-58.<br />

[Lee 1981] Lee, J. C. <strong>and</strong> Timasheff, S. N. (1981). "The stabilization <strong>of</strong> <strong><strong>protein</strong>s</strong> by<br />

sucrose." J Biol Chem 256(14): 7193-201.<br />

[Leong 1981] Leong, K. H. (1981). "Morphology <strong>of</strong> aerosol particles generated from<br />

the evaporation <strong>of</strong> solution drops."<br />

Journal <strong>of</strong> Aerosol Science 12(5): 417-435.<br />

[Leong 1987] Leong, K. H. (1987). "Morphological control <strong>of</strong> particles generated<br />

from evaporation <strong>of</strong> solution <strong>droplet</strong>s."<br />

Journal <strong>of</strong> Aerosol Science 18(5): 525-552.


248 REFERENCES<br />

[Liao et al. 2002] Liao, Y. H., Brown, M. B., Quader, A. <strong>and</strong> Martin, G. P. (2002).<br />

"Protective mechanism <strong>of</strong> stabilizing excipients against dehydration in<br />

the freeze-<strong>drying</strong> <strong>of</strong> <strong><strong>protein</strong>s</strong>." Pharm Res 19(12): 1854-61.<br />

[Liao et al. 2005] Liao, Y. H., Brown, M. B., Jones, S. A., Nazir, T. <strong>and</strong> Martin, G. P.<br />

(2005). "The effects <strong>of</strong> polyvinyl alcohol on the in vitro stability <strong>and</strong><br />

delivery <strong>of</strong> spray-dried <strong>protein</strong> particles from surfactant-free HFA<br />

134a-based pressurised metered dose inhalers." Int J Pharm 304(1-2):<br />

29-39.<br />

[Lierke 1995] Lierke, E. G. (1995). "Comparative considerations <strong>of</strong> aerodynamic,<br />

acoustic <strong>and</strong> electrostatic single-drop levitation systems."<br />

Forschung im Ingenieurwesen 61(7/8): 210-216.<br />

[Lierke 1996] Lierke, E. G. (1996). "Akustische Positionierung - Ein umfassender<br />

Überblick über Grundlagen und Anwendung."<br />

Acta Acoustica 82: 220-237.<br />

[Lierke 2002] Lierke, E. G. (2002). "Deformation <strong>and</strong> displacement <strong>of</strong> liquid drops in<br />

an optimized acoustic st<strong>and</strong>ing wave levitator."<br />

Acta Acoustica united with Acoustica 88(2): 206-207.<br />

[Lin 1996] Lin, T. Y. <strong>and</strong> Timasheff, S. N. (1996). "On the role <strong>of</strong> surface tension<br />

in the stabilization <strong>of</strong> globular <strong><strong>protein</strong>s</strong>." Protein Sci 5(2): 372-81.<br />

[Lin 2003] Lin, J. C. <strong>and</strong> Gentry, J. W. (2003). "Spray <strong>drying</strong> drop morphology:<br />

experimental study." Aerosol Science <strong>and</strong> Technology 37(1): 15-32.<br />

[Lins et al. 2004] Lins, R. D., Pereira, C. S. <strong>and</strong> Hunenberger, P. H. (2004). "Trehalose<strong>protein</strong><br />

interactions in aqueous solution."<br />

Proteins: Structure, Function <strong>and</strong> Bioinformatics. 55: 177-186.<br />

[Liu 1996] Liu, Y. <strong>and</strong> Sturtevant, J M. (1996). "The observed change in heat<br />

capacity accompanying the thermal unfolding <strong>of</strong> <strong><strong>protein</strong>s</strong> depends on<br />

the composition <strong>of</strong> the solution <strong>and</strong> on the method employed to change<br />

the temperature <strong>of</strong> unfolding." Biochemistry 35(9): 3059-3062.<br />

[Lucas et al. 1998] Lucas, P., Anderson, K. <strong>and</strong> Staniforth, J. N. (1998). "Protein<br />

deposition from dry powder inhalers: fine particle multiplets as<br />

performance modifiers." Pharm Res 15(4): 562-9.<br />

[Maa et al. 1997] Maa, Y. F., Costantino, H. R., Nguyen, P. A. <strong>and</strong> Hsu, C. C. (1997).<br />

"The effect <strong>of</strong> operating <strong>and</strong> formulation variables on the morphology<br />

<strong>of</strong> spray-dried <strong>protein</strong> particles." Pharm Dev Technol 2(3): 213-23.<br />

[Maa et al 1998 a] Maa, Y. F., Nguyen, P. A., Andya, J. D., Dasovich, N., Sweeney, T. D.,<br />

Shire, S. J. <strong>and</strong> Hsu, C. C. (1998). "Effect <strong>of</strong> spray <strong>drying</strong> <strong>and</strong><br />

subsequent processing conditions on residual moisture content <strong>and</strong><br />

physical/biochemical stability <strong>of</strong> <strong>protein</strong> inhalation powders."<br />

Pharm Res 15(5): 768-75.


REFERENCES 249<br />

[Maa et al. 1998 b] Maa, Y. F., Nguyen, P. A., Sit, K. <strong>and</strong> Hsu, C. C. (1998 b). "Spray<strong>drying</strong><br />

performance <strong>of</strong> a bench-top spray dryer for <strong>protein</strong> aerosol<br />

powder preparation." Biotechnol Bioeng 60(3): 301-9.<br />

[MacRitchie 1986] MacRitchie, F. (1986). "Spread monolayers <strong>of</strong> <strong><strong>protein</strong>s</strong>."<br />

Adv Colloid Interface Sci 25(4): 341-85.<br />

[Marston 1980] Marston, P. L. (1980). "Shape oscillation <strong>and</strong> static deformation <strong>of</strong><br />

drops <strong>and</strong> bubbles driven by modulated radiation stresses - theory."<br />

Journal <strong>of</strong> Acoustic Society <strong>of</strong> America 67(1): 15-26.<br />

[Masters 1991] Masters, K. (1991). Spray Drying H<strong>and</strong>book.<br />

Essex, Great Britain, Longman Scientific & Technical<br />

[Masters 2002] Masters, K. (2002). Spray Drying in Practice.<br />

Denmark, SprayDryConsult International ApS<br />

[Maury 2005] Maury, M. (2005). Aggregation <strong>and</strong> structural changes in spray dried<br />

immunoglobulins. Institute <strong>of</strong> Pharmaceutics. Erlangen, Friedrich-<br />

Alex<strong>and</strong>er University. Ph.D. thesis.<br />

[Maury et al. 2005 a] Maury, M., Murphy, K., Kumar, S., Mauerer, A. <strong>and</strong> Lee, G. (2005).<br />

"Spray-<strong>drying</strong> <strong>of</strong> <strong><strong>protein</strong>s</strong>: effects <strong>of</strong> sorbitol <strong>and</strong> trehalose on<br />

aggregation <strong>and</strong> FT-IR amide I spectrum <strong>of</strong> an immunoglobulin G."<br />

Eur J Pharm Biopharm 59(2): 251-61.<br />

[Maury et al. 2005 b] Maury, M., Murphy, K., Kumar, S., Shi, L. <strong>and</strong> Lee, G. (2005). "Effects<br />

<strong>of</strong> process variables on the powder yield <strong>of</strong> spray-dried trehalose on a<br />

laboratory spray-dryer." Eur J Pharm Biopharm 59(3): 565-73.<br />

[McCabe et al. 2005] McCabe, W. L., Smith, J. C. <strong>and</strong> Harriott, P. (2005). Unit operations <strong>of</strong><br />

chemical engineering. Boston (USA), McGraw-Hill<br />

[Mc Evoy 1992] McEvoy, G. K. (1992). AHFS drug information. USA<br />

[Mingotaud et al. 1993] Mingotaud, A. F., Mingotaud, C. <strong>and</strong> Patterson, L. K. (1993).<br />

H<strong>and</strong>book <strong>of</strong> monolayers. Boca Raton, CRC Press<br />

[Moser et al. 2001] Möser, C., Groenewold, C., Groenewold, H. <strong>and</strong> Tsotsas, E. (2001).<br />

"Study <strong>of</strong> kinetics <strong>of</strong> separation processes in an acoustic levitator:<br />

advantages <strong>and</strong> drawbacks." Chemie Ingenieur Technik 73(8): 1012-<br />

1017.<br />

[Musonge 2005] Musonge, P. <strong>and</strong> Möller, K. (2005).<br />

"Mass transfer across gas-liquid-solid interfaces using film theory."<br />

http://ugs.chemeng.uct.ac.za/file.php/32/FilmCoefficients.pdf.<br />

[Nave 2005] Nave, C. R. (2005). Hyperphysics - Department <strong>of</strong> Physics <strong>and</strong><br />

Astronomy. Georgia State University.<br />

[Nektar 2005] Nektar (2005). http://www.nektar.com.


250 REFERENCES<br />

[Nesic 1991] Nesic, S. <strong>and</strong> Vodnik, J. (1991). "Kinetics <strong>of</strong> <strong>droplet</strong> evaporation."<br />

Chemical Engineering Science 46(3): 527-537.<br />

[Newman et al. 1993] Newman, Y. M., Ring, S. G. <strong>and</strong> Colaco, C. (1993). "The role <strong>of</strong><br />

trehalose <strong>and</strong> other carbohydrates in biopreservation."<br />

Biotechnology <strong>and</strong> Genetic Engineering Reviews 11: 263-294.<br />

[Niro 2005] Niro (2005). "Spray Drying." Company Journal.<br />

[Niven et al. 1993] Niven, R. W., Lott, F. D. <strong>and</strong> Cribbs, J. M. (1993). "Pulmonary<br />

absorption <strong>of</strong> recombinant methionyl human granulocyte colony<br />

stimulating factor (r-hG-CSF) after intratracheal instillation to the<br />

hamster."<br />

Pharm Res 10(11): 1604-10.<br />

[Niven et al. 1994] Niven, R. W., Lott, F. D., Ip, A. Y. <strong>and</strong> Cribbs, J. M. (1994).<br />

"Pulmonary delivery <strong>of</strong> powders <strong>and</strong> solutions containing recombinant<br />

human granulocyte colony-stimulating factor (rhG-CSF) to the rabbit."<br />

Pharm Res 11(8): 1101-9.<br />

[Niven 1980] Niven, W. D. (1980). The Scientific Papers <strong>of</strong> James Clerc Maxwell.<br />

Cambridge, Cambridge University Press<br />

[Novartis 2005] Novartis (2005). http://www.novartis-pharma.com.<br />

[Nurnberg 1980] Nurnberg, E. (1980). "Darstellung und Eigenschaften pharmazeutisch<br />

relevanter Sprühtrocknungsprodukte, eine Übersicht."<br />

Acta Pharmaceutica Technologica 26(1): 39-67.<br />

[Ohsaka et al. 2002] Ohsaka, K., Rednikov, A., Sadhal, S. S. <strong>and</strong> Trinh, E. H. (2002).<br />

"Noncontact technique for determining viscosity from the shape<br />

relaxation <strong>of</strong> ultrasonically levitated <strong>and</strong> initially elongated drops."<br />

Review <strong>of</strong> Scientific Instruments 73(5): 2091-2096.<br />

[Ohsaka et al. 2003] Ohsaka, K., Rednikov, A. <strong>and</strong> Sadhal, S. S. (2003). "Noncontact<br />

technique for determining the thermal diffusivity coefficient on<br />

acoustically levitated liquid drops."<br />

Review <strong>of</strong> Scientific Instruments 74(2): 1107-1112.<br />

[Patton 1997] Patton, J. S. (1997). "Deep-lung delivery <strong>of</strong> therapeutic <strong><strong>protein</strong>s</strong>."<br />

Chemtech 27(12): 34-38.<br />

[PDB 2005] PDB (2005). "Protein Data Bank." http://www.rcsb.org/pdb/.<br />

[Peters 1985] Peters, T., Jr. (1985). "Serum albumin."<br />

Advances in Protein Chemistry 37: 161-245.<br />

[Popiel 1998] Popiel, C. O. <strong>and</strong> Wojtkowiak, J. (1998). "Simple formulas for<br />

thermophysical properties <strong>of</strong> liquid water fro heat transfer calculations<br />

(from 0°C to 150°C)." Heat Transfer Engineering 19(3): 87-101.


REFERENCES 251<br />

[R<strong>and</strong>olph 2002] R<strong>and</strong>olph, T. W. <strong>and</strong> Jones, L. S. (2002). "Surfactant-<strong>protein</strong><br />

interactions." Pharm Biotechnol 13: 159-75.<br />

[Ranz 1952 a] Ranz, W. E. <strong>and</strong> Marshall, W. R. (1952). "Evaporation from Drops -<br />

Part 1." Chemical Engineering Progress 48(3): 141-146.<br />

[Ranz 1952 b] Ranz, W. E. <strong>and</strong> Marshall, W. R. (1952). "Evaporation from Drops -<br />

Part 2." Chemical Engineering Progress 48(4): 173-180.<br />

[Reid 1987] Reid, R. C., Prausnitz, J. M. <strong>and</strong> Poling, B. E. (1987).<br />

The properties <strong>of</strong> gases <strong>and</strong> liquids. McGraw-Hill<br />

[Rensink 2004] Rensink, D. (2004). Verdunstung akustisch levitierter schwingender<br />

Tropfen aus homogenen und heterogenen Medien. Institute <strong>of</strong> Fluid<br />

Mechanics. Erlangen, Germany, University <strong>of</strong> Erlangen. Ph.D. Thesis.<br />

[Richards et al. 2002] Richards, A. B., Krakowka, S., Dexter, L. B., Schmid, H., Wolterbeek,<br />

A. P. M., Waalkens-Berendsen, D. H., Shigoyuki, M. <strong>and</strong> Kurimoto, M.<br />

(2002). "Trehalose: a 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."<br />

Food <strong>and</strong> Chemical Toxicology 40: 871-898.<br />

[Roche 2005] Roche (2005). http://www.roche.de.<br />

[Rohling et al. 2000] Rohling, O., Weitkamp, C. <strong>and</strong> Neidhart, B. (2000). "Experimental<br />

setup for the determination <strong>of</strong> analytes contained in ultrasonic levitated<br />

<strong>droplet</strong>s."<br />

Fresenius Journal <strong>of</strong> Analytical Chemistry 368(2-3): 125-129.<br />

[Roth 1988] Roth, C. <strong>and</strong> Koebrich, R. (1988). "Production <strong>of</strong> hollow spheres."<br />

Journal <strong>of</strong> Aerosol Science 19(7): 939-942.<br />

[Sacher 2005] Sacher, S. <strong>and</strong> Krammer, G. (2005). "Study <strong>of</strong> crystallisation processes<br />

in an ultrasonic levitator." Chemie Ingenieur Technik 77(3): 290-294.<br />

[Sano 1982] Sano, Y. <strong>and</strong> Keey, R. B. (1982). "The <strong>drying</strong> <strong>of</strong> a spherical particle<br />

containing colloidal material into a hollow sphere."<br />

Chemical Engineering Science 37(6): 881-889.<br />

[Schlichting 2000] Schlichting, H. <strong>and</strong> Gersten, K. (2000). Boundary-Layer Theory.<br />

Berlin, Heidelberg, New York, Springer<br />

[Schnitzler 1998] Schnitzler, A. (1998). Aufbau einer akustischen Falle. Department <strong>of</strong><br />

Physics, University <strong>of</strong> Mainz, Germany. State examination work.<br />

[Schott 2005] Schott (2005). "Operating instructions Ubbelohde Viscometer."<br />

http://www.schott.com/labinstruments.<br />

[Seaver et al. 1989] Seaver, M., Galloway, A. <strong>and</strong> Manuccia, T. J. (1989). "Acoustic<br />

levitation in a free-jet wind tunnel."<br />

Review <strong>of</strong> Scientific Instruments 60(11): 3452-3459.


252 REFERENCES<br />

[Seaver et al. 1990] Seaver, M., Galloway, A. <strong>and</strong> Manuccia, T. J. (1990). "Water<br />

condensation onto an evaporating drop <strong>of</strong> 1-butanol."<br />

Aerosol Science <strong>and</strong> Technology 12(3): 741-744.<br />

[Seth 2005] Seth, O. (2005). http://strucbio.biologie.uni-konstanz.de/det/det2.htm.<br />

[Shi 1996] Shi, W. T. <strong>and</strong> Apfel, R. E. (1996). "Deformation <strong>and</strong> position <strong>of</strong><br />

acoustically levitated liquid drops."<br />

Journal <strong>of</strong> Acoustical Society <strong>of</strong> America 99(4): 1977-1984.<br />

[Sigma 2005] Sigma (2005). http://www.sigma-aldrich.com.<br />

[Sirignano 1999] Sirignano, W. A. (1999). Fluid Dynamics <strong>and</strong> Transport <strong>of</strong> Droplets<br />

<strong>and</strong> Sprays. Cambridge, Cambridge University Press<br />

[Sola-Penna 1998] Sola-Penna, M. <strong>and</strong> Meyer-Fernades, J. R. (1998). "Stabilization<br />

against thermal inactivation promoted by sugars on enzyme structure<br />

<strong>and</strong> function. Why is trehalose more effective than other sugars?"<br />

Archives <strong>of</strong> Biochemistry 360(1): 10-14.<br />

[Stahl 1980] Stahl, H. P. (1980). Feuchtigkeit und Trocknen in der pharmazeutischen<br />

Technologie. Darmstadt, Germany, Dr. Dietrich Steinh<strong>of</strong>f Verlag<br />

[Stosch 2000] Stosch, R. <strong>and</strong> Cammenga, H. K. (2000). "Molecular interactions in<br />

mixed monolayers <strong>of</strong> octadecanoic acid <strong>and</strong> three related amphiphiles."<br />

J Colloid Interface Sci 230(2): 291-297.<br />

[Stuart 1966] Stuart, J. T. (1966). "Double boundary layers in oscillatory viscous<br />

flow." Journal <strong>of</strong> Fluid Mechanics 24: 673.<br />

[Sunkel 1993] Sunkel, J. M. <strong>and</strong> King, C. J. (1993). "Influence <strong>of</strong> the development <strong>of</strong><br />

particle morphology upon rates <strong>of</strong> loss <strong>of</strong> volatile solutes during <strong>drying</strong><br />

<strong>of</strong> drops."<br />

Industrial & Engineering Chemistry Research 32(10): 2357-2364.<br />

[Tattini et al. 2005] Tattini, V., Jr., Parra, D. F., Polakiewicz, B. <strong>and</strong> Pitombo, R. N. (2005).<br />

"Effect <strong>of</strong> lyophilisation on the structure <strong>and</strong> phase changes <strong>of</strong><br />

PEGylated-bovine serum albumin." Int J Pharm 304(1-2): 124-34.<br />

[Temkin 1882] Temkin, S. <strong>and</strong> Metha, H. K. (1982). "Droplet drag in an accelerated<br />

flow." Journal <strong>of</strong> Fluid Mechanics 116: 279-313.<br />

[Tian 1996] Tian, Y. <strong>and</strong> Apfel, R. E. (1996). "A novel multiple drop levitator for<br />

the study <strong>of</strong> drop arrays." Journal <strong>of</strong> Aerosol Science 27(5): 721-737.<br />

[Timasheff 2002] Timasheff, S. N. (2002). "Protein hydration, thermodynamic binding,<br />

<strong>and</strong> preferential hydration." Biochemistry 41(46): 13473-13482.<br />

[Tipler 2000] Tipler, P. A. (2000). Physics. Heidelberg, Berling (Germany),<br />

Spektrum Akademischer Verlag


REFERENCES 253<br />

[Trinh 1985] Trinh, E. H. (1985). "Compact acoustic levitation device studies in fluid<br />

dynamics <strong>and</strong> material science in the laboratory <strong>and</strong> microgravity."<br />

Review <strong>of</strong> Scientific Instruments 56(11): 2059-2065.<br />

[Trinh 1986] Trinh, E. H. <strong>and</strong> Hsu, C. J. (1986). "Acoustic levitation methods for<br />

density measurements."<br />

Journal <strong>of</strong> the Acoustic Society <strong>of</strong> America 80(6): 1757-1761.<br />

[Tripp 1995] Tripp, B. C. <strong>and</strong> Magda, J. J. (1995). "Adsorption <strong>of</strong> globular <strong><strong>protein</strong>s</strong><br />

at the air/water interface as measured via dynamic surface tension:<br />

concentration dependence, mass transfer considerations <strong>and</strong> adsorption<br />

kinetics." Journal <strong>of</strong> Colloid <strong>and</strong> Interface Science 173(1): 16-27.<br />

[Tuckermann 2002] Tuckermann, R. (2002). Gase, Aerosole, Tropfen und Partikel in<br />

stehenden Ultraschallfeldern: Eine Untersuchung zur Anreicherung<br />

schwerer Gase, Verdampfung levitierter Tropfen, Kristall- und<br />

Partikelbildung. Faculty <strong>of</strong> Science, Technical University Carolo-<br />

Wilhelmina, Braunschweig (Germany). Ph.D. Thesis.<br />

[Tuckermann et al. 2002] Tuckermann, R., Bauerecker, S. <strong>and</strong> Neidhart, B. (2002). "Evaporation<br />

rates <strong>of</strong> alkanes <strong>and</strong> alkanols from acoustically levitated <strong>droplet</strong>s."<br />

Analytical <strong>and</strong> Bioanalytical Chemistry 372(1): 122-127.<br />

[Tuckermann et al 2001] Tuckermann, R., Bauerecker, S. <strong>and</strong> Neidhart, B. (2001). "Levitation in<br />

ultrasonic fields - suspended <strong>droplet</strong>s."<br />

Physik in unserer Zeit 32(2): 69-75.<br />

[Tzannis 1999 a] Tzannis, S. T. <strong>and</strong> Prestrelski, S. J. (1999). "Activity-stability<br />

considerations <strong>of</strong> trypsinogen during spray <strong>drying</strong>: effects <strong>of</strong> sucrose."<br />

J Pharm Sci 88(3): 351-9.<br />

[Tzannis 1999 b] Tzannis, S. T. <strong>and</strong> Prestrelski, S. J. (1999). "Moisture effects on<br />

<strong>protein</strong>-excipient interactions in spray-dried powders. Nature <strong>of</strong><br />

destabilizing effects <strong>of</strong> sucrose." J Pharm Sci 88(3): 360-70.<br />

[Verger et al. 1973] Verger, R., Mieras, M. C. <strong>and</strong> de Haas, G. H. (1973). "Action <strong>of</strong><br />

phospholipase A at interfaces." J Biol Chem 248(11): 4023-34.<br />

[Walton 1994] Walton, D. E. (1994). The Morphology <strong>of</strong> Spray-Dried Particles.<br />

Birmingham, University <strong>of</strong> Aston. Doctor <strong>of</strong> Philosophy.<br />

[Walton 1999] Walton, D. E. <strong>and</strong> Mumford, C. J. (1999). "The Morphology <strong>of</strong> Spray-<br />

Dried Particles." Trans IChemE 77(A5): 442-460.<br />

[Weis 2005] Weis, D. D. <strong>and</strong> Nardozzi, J. D. (2005). "Enzyme kinetics in<br />

acoustically levitated <strong>droplet</strong>s <strong>of</strong> supercooled water: a novel approach<br />

to cryoenzymology." Analytical Chemistry 77(8): 2558-2563.<br />

[Welter 1997] Welter, E. <strong>and</strong> Neidhart, B. (1997). "Acoustically levitated <strong>droplet</strong>s - a<br />

new tool for micro <strong>and</strong> trace analysis."<br />

Fresenius Journal <strong>of</strong> Analytical Chemistry 357(3): 345-350.


254 REFERENCES<br />

[Wen 2001] Wen, X. <strong>and</strong> Franses, E. I. (2001). "Adsorption <strong>of</strong> bovine serum<br />

albumin at the air/water interface <strong>and</strong> its effects on the formation <strong>of</strong><br />

DPPC surface film." Colloids <strong>and</strong> Surfaces A 190: 319-332.<br />

[Wijlhuizen et al. 1979] Wijlhuizen, A. E., Kerkh<strong>of</strong>, P. J. M. <strong>and</strong> Bruin, S. (1979). "Theoretical<br />

study <strong>of</strong> the inactivation <strong>of</strong> phosphatase during spray <strong>drying</strong> <strong>of</strong> skimmilk."<br />

Chemical Engineering Science 34(5): 651-660.<br />

[Wilson 1964] Wilson, G. M. (1964). "Vapour-Liquid Equilibrium - A new expression<br />

for the excess free energy <strong>of</strong> mixing."<br />

J. Am. Chem. Soc. 86(2): 127-130.<br />

[Windholz et al. 1983] Windholz, M., Budavari, S., Blumetti, R. F. <strong>and</strong> Otterbein, E. S. (1983).<br />

"The Merck Index." www.merck.com.<br />

[Witschi 1999] Witschi, C. <strong>and</strong> Mrsny, R. J. (1999). "In vitro evaluation <strong>of</strong><br />

microparticles <strong>and</strong> polymer gels for use as nasal platforms for <strong>protein</strong><br />

delivery." Pharm Res 16(3): 382-90.<br />

[Xie 1997] Xie, G. <strong>and</strong> Timasheff, S. N. (1997). "The thermodynamic mechanism<br />

<strong>of</strong> <strong>protein</strong> stabilization by trehalose." Biophysical Chemistry 64: 25-43.<br />

[Xie 2001] Xie, W. J. <strong>and</strong> Wei, B. (2001). "Parametric studies <strong>of</strong> single-axis<br />

acoustic levitation." Applied Physics Letters 79(6): 881-883.<br />

[Yang et al. 1989] Yang, W. J., Guo, K. H. <strong>and</strong> Uemura, T. (1989). "Theory <strong>and</strong><br />

experiments on evaporating sessile drops <strong>of</strong> binary liquid mixtures."<br />

Int. J. Heat Mass Transfer 32(7): 1197-1205.<br />

[Yao et al. 2003] Yao, G. F., Abdel-Khalik, S. I. <strong>and</strong> Ghiaasiaan, S. M. (2003). "An<br />

investigation <strong>of</strong> simple evaporation models used in spray simulations."<br />

Journal <strong>of</strong> Heat Transfer 125: 179-182.<br />

[Yarin et al. 1997] Yarin, A. L., Brenn, G., Keller, J., Pfaffenlehner, M., Ryssel, E. <strong>and</strong><br />

Tropea, C. (1997). "Flowfield characteristics <strong>of</strong> an aerodynamic<br />

acoustic levitator." Physic <strong>of</strong> Fluids 9(11): 3300-3314.<br />

[Yarin et al 1998] Yarin, A. L., Pfaffenlehner, M. <strong>and</strong> Tropea, C. (1998). "On the acoustic<br />

levitation <strong>of</strong> <strong>droplet</strong>s." Journal <strong>of</strong> Fluid Mechanics 356: 65-91.<br />

[Yarin et al. 1999] Yarin, A. L., Kastner, O., Brenn, G., Rensink, D. <strong>and</strong> Tropea, C.<br />

(1999). "Evaporation <strong>of</strong> acoustically levitated <strong>droplet</strong>s."<br />

Journal <strong>of</strong> Fluid Mechanics 399: 151-204.<br />

[Yarin et al. 2002 a] Yarin, A. L., Brenn, G., Kastner, O. <strong>and</strong> Tropea, C. (2002). "Drying <strong>of</strong><br />

acoustically levitated <strong>droplet</strong>s <strong>of</strong> liquid-solid suspensions: Evaporation<br />

<strong>and</strong> crust formation." Physic <strong>of</strong> Fluids 14(7): 2289-2298.


REFERENCES 255<br />

[Yarin et al. 2002 b] Yarin, A. L., Brenn, G. <strong>and</strong> Rensink, D. (2002). "Evaporation <strong>of</strong><br />

acoustically levitated <strong>droplet</strong>s <strong>of</strong> binary liquid mixtures."<br />

International Journal <strong>of</strong> Heat <strong>and</strong> Fluid Flow 23(4): 471-486.<br />

[Yarin et al. 2002 c] Yarin, A. L., Weiss, D. A., Brenn, G. <strong>and</strong> Rensink, D. (2002c).<br />

"Acoustically levitated drops: drop oscillation <strong>and</strong> break-up driven<br />

ultrasound modulation."<br />

International Journal <strong>of</strong> Multiphase Flow 28(6): 887-910.<br />

[Yu et al. 1998] Yu, L., Mishra, D. S. <strong>and</strong> Rigsbee, D. R. (1998). "Determination <strong>of</strong><br />

glass properties <strong>of</strong> D-mannitol using sorbitol as an impurity."<br />

Journal <strong>of</strong> Pharmaceutical Science 87(6): 774-777.<br />

[Zhang et al. 1991] Zhang, S. C., Messing, G. L. <strong>and</strong> Huebner, W. (1991). "Yttrium barium<br />

cooper oxide superconductor power synthesis by spray pyrolysis <strong>of</strong><br />

organic acid solutions." Journal <strong>of</strong> Aerosol Science 22(5): 585-599.


256 CURRICULUM VITAE<br />

PERSONAL DATA<br />

Curriculum vitae<br />

Name: Heiko Alex<strong>and</strong>er Schiffter<br />

Date <strong>of</strong> birth: 02. April 1976<br />

Place <strong>of</strong> birth: Heilbronn-Neckargartach, Germany<br />

Nationality: German<br />

Family status: single<br />

Address: Friedrich-Ebert-Street 4, 74177 Bad Friedrichshall, Germany<br />

SCHOOL EDUCATION<br />

1982-1986 Basic primary school, Bad Friedrichshall-Hagenbach, Germany<br />

1986-1995 Grammar school – Moenchsee-Gymnasium, Heilbronn, Germany<br />

06/1995 Diploma from German secondary school qualifying for<br />

university administration <strong>and</strong> matriculation<br />

CIVILIAN SERVICE<br />

1995-1996 Malteser emergency service, Heilbronn, Germany<br />

• Alternative civilian service in care <strong>and</strong> transportation<br />

<strong>of</strong> disabled people<br />

PROFESSIONAL AND ACADEMIC TRAINING<br />

1996-2001 Ruprechts-Karls University, Heidelberg, Germany<br />

• Undergraduate studies in pharmacy<br />

09/1998 1 st part <strong>of</strong> state examination in pharmacy<br />

08-10/1999 German Cancer Research Centre, Heidelberg, Germany<br />

• Research activity in cell biology with interest in<br />

chemoprevention <strong>of</strong> cancer<br />

04/2001 2 nd part <strong>of</strong> state examination in pharmacy<br />

2001-2002 Central-Apotheke, Ravensburg, Germany<br />

• Internship in a community-based pharmacy<br />

05/2002 3 rd part <strong>of</strong> state examination in pharmacy<br />

05/2002 Licensed pharmacist<br />

2002-2005 Department <strong>of</strong> pharmaceutics (Pr<strong>of</strong>. Dr. Ge<strong>of</strong>frey Lee),<br />

Friedrich-Alex<strong>and</strong>er University, Erlangen, Germany<br />

• Ph.D. student in pharmaceutics <strong>and</strong> assistant teacher for<br />

undergraduate students (solid dosage forms <strong>and</strong> biopharmaceutics)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!