27.02.2014 Views

INAUGURAL–DISSERTATION zur Erlangung der Doktorwürde der ...

INAUGURAL–DISSERTATION zur Erlangung der Doktorwürde der ...

INAUGURAL–DISSERTATION zur Erlangung der Doktorwürde der ...

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.

<strong>INAUGURAL–DISSERTATION</strong><br />

<strong>zur</strong><br />

<strong>Erlangung</strong> <strong>der</strong> <strong>Doktorwürde</strong><br />

<strong>der</strong><br />

Naturwissenschaftlich–Mathematischen<br />

Gesamtfakultät<br />

<strong>der</strong><br />

Ruprecht–Karls–Universität<br />

Heidelberg<br />

vorgelegt von<br />

M. Tech. Srikanth Reddy Gopireddy<br />

aus Nizambad, India<br />

Tag <strong>der</strong> mündlichen Prüfung: 06.12.2013


Numerical Simulation of Bi-component Droplet<br />

Evaporation and Dispersion in Spray and<br />

Spray Drying<br />

Gutachter: Prof. Dr. Eva Gutheil<br />

PD Dr. N. Dahmen


Dreams can never become a reality without hard work<br />

– Swamy Vivekananda


Abstract<br />

Spray drying is one of the most widely used drying techniques to convert liquid feed<br />

into a dry pow<strong>der</strong>. The modeling of spray flows and spray drying has been studied<br />

for many years now, to determine the characteristics of the end products, e.g. particle<br />

size, shape, density or porosity. So far, the simulation of polymer or sugar solution<br />

spray drying has not been studied because drying behavior as well as properties are<br />

unknown. Previous studies concentrated on the systems of milk, salt solution, colloids<br />

or other materials for which the thermal and physical properties are well tabulated.<br />

The present study deals with the modeling and simulation of polyvinylpyrrolidone<br />

(PVP)/water and mannitol/water spray flows. PVP is a polymer, widely used as<br />

a pharmaceutical excipient, and mainly manufactured by BASF un<strong>der</strong> several patented<br />

names, whereas mannitol is a sugar, which is used in dry pow<strong>der</strong> inhalers and tablets.<br />

Experimental studies have shown that the pow<strong>der</strong> properties of PVP and mannitol are<br />

significantly influenced by the drying conditions. The growing importance of PVP or<br />

mannitol pow<strong>der</strong>s and the inability of existing studies to predict the effect of drying<br />

conditions on the properties of the end product have prompted the development of a<br />

new reliable model and numerical techniques.<br />

Evaporating sprays have a continuous phase (gas) and a dispersed phase, which<br />

consists of droplets of various sizes that may evaporate, coalesce, or breakup, as well<br />

as have their own inertia and size-conditioned dynamics. A modeling approach which<br />

is more commonly used is the Lagrangian description of the dispersed liquid phase.<br />

This approach gives detailed information on the micro-level, but inclusion of droplet<br />

coalescence and breakup increase computational complexity. Moreover, the Lagrangian<br />

description coupled with the Eulerian equations for the gas phase, assuming a pointsource<br />

approximation of the spray, is computationally expensive. As an alternative to<br />

Lagrangian simulations, several Eulerian methods have been developed based on the<br />

Williams’ spray equation. The Euler – Euler methods are computationally efficient<br />

and independent of liquid mass loading in describing dense turbulent spray flows.<br />

The objective of this thesis is the modeling and simulation of spray flows and<br />

spray drying up to the onset of solid layer formation in an Euler – Euler framework.<br />

The behavior of droplet distribution un<strong>der</strong> various drying conditions in bi-component<br />

evaporating spray flows is examined using, for the first time, direct quadrature method<br />

of moments (DQMOM) in two dimensions. In DQMOM, the droplet size and velocity<br />

distribution of the spray is modeled by approximating the number density function in<br />

terms of joint radius and velocity. Transport equations of DQMOM account for droplet<br />

evaporation, heating, drag, and droplet–droplet interactions.<br />

At first, an evaporating water spray in nitrogen is modeled in one dimension (ax-<br />

I


II<br />

ial direction). Earlier studies in spray flows neglected evaporation or consi<strong>der</strong>ed it<br />

through a simplified model, which is addressed by implementing an advanced droplet<br />

evaporation model of Abramzon and Sirignano, whereas droplet motion and droplet<br />

coalescence are estimated through appropriate sub-models. The assumption of evaporative<br />

flux to be zero or computing it with weight ratio constraints was found to be<br />

unphysical, which is improved by estimating it using the maximum entropy formulation.<br />

The gas phase is not yet fully coupled to the DQMOM but its inlet properties are<br />

taken to compute forces acting on droplets and evaporation. The simulation results<br />

are compared with quadrature method of moments (QMOM) and with experiment at<br />

various cross sections. DQMOM shows better results than QMOM, and remarkable<br />

agreement with experiment.<br />

Next, water spray in air in two-dimensional, axisymmetric configuration is modeled<br />

by extending the one-dimensional DQMOM. The DQMOM results are compared with<br />

those of the discrete droplet model (DDM), which is an Euler – Lagrangian approach.<br />

Droplet coalescence is consi<strong>der</strong>ed in DQMOM but neglected in DDM. The simulation<br />

results are validated with new experimental data. Overall, DQMOM shows a much<br />

better performance with respect to computational effort, even with the inclusion of<br />

droplet coalescence.<br />

Before extending DQMOM to model PVP/water spray flows, a single droplet evaporation<br />

and drying model is developed, because most of the evaporation models available<br />

in the literature are valid for salts, colloids or milk pow<strong>der</strong>. The negligence of solid layer<br />

formation effects on the droplet heating and evaporation is addressed, and treatment<br />

of the liquid mixture as the ideal solution is improved by including the non-ideality<br />

effect. The PVP or mannitol in water droplet evaporation and solid layer formation<br />

are simulated, and the results are compared with new experimental data, which shows<br />

that the present model effectively captures the first three stages of evaporation and<br />

drying of a bi-component droplet.<br />

Finally, PVP/water spray flows in air are simulated using DQMOM including the<br />

developed bi-component evaporation model. Simulation results are compared with new<br />

experimental data at various cross sections and very good agreement is observed.<br />

In conclusion, water and PVP/water evaporating spray flows, and preliminary<br />

stages of PVP/water and mannitol/water spray drying, i.e., until solid layer formation,<br />

are successfully modeled and simulated, and show good agreement with experiment.<br />

Keywords: Sprays, PVP, Mannitol, DQMOM, Bi-component droplet


III<br />

Zusammenfassung<br />

Sprühtrocknung ist eines <strong>der</strong> am häufigsten eingesetzten Verfahren, um eine zugeführte<br />

Flüssigkeit in ein trockenes Pulver umzuwandeln. Die Modellierung von Sprühtrocknungsprozessen<br />

und des Sprays selbst wird seit vielen Jahren betrieben, um die Eigenschaften<br />

<strong>der</strong> Endprodukte, wie z.B. Partikelgröße, Form, Dichte o<strong>der</strong> Porosität, bestimmen<br />

zu können. Die Sprühtrocknung von Polymer- o<strong>der</strong> Zuckerlösungen wurde bisher<br />

noch nicht numerisch untersucht, da <strong>der</strong>en Trocknungsverhalten und Eigenschaften<br />

unbekannt sind. Bislang wurden nur Systeme mit Milch, Salzlösungen o<strong>der</strong> Kolloiden<br />

untersucht, <strong>der</strong>en thermische und physikalische Eigenschaften gut belegt sind.<br />

Die vorliegende Arbeit widmet sich <strong>der</strong> Modellierung und Simulation von Polyvinylpyrrolidon<br />

(PVP)/Wasser und Mannitol/Wasser-Sprays. PVP ist ein Polymer, weit<br />

verbreitet als pharmazeutisches Bindemittel und von <strong>der</strong> BASF unter verschiedenen<br />

patentierten Namen hergestellt, während Mannitol, ein Zucker, hauptsächlich in Trockenpulverinhalatoren<br />

und Tabletten verwendet wird. Experimentelle Studien haben gezeigt,<br />

dass die Eigenschaften von PVP- und Mannitol-Pulvern von den Trocknungsbedingungen<br />

signifikant beeinflusst werden. Die zunehmende Bedeutung von PVP- und<br />

Mannitol-Pulvern und das Fehlen geeigneter Methoden <strong>zur</strong> Bestimmung des Einflusses<br />

<strong>der</strong> Trocknungsbedingungen auf die Eigenschaften <strong>der</strong> Endprodukte haben die Entwicklung<br />

eines neuen zuverlässigen Modells sowie numerischer Methoden angeregt.<br />

Verdampfende Sprays bestehen aus einer kontinuierlichen Phase (Gas), und einer<br />

zerstäubten Phase, die aus Tropfen unterschiedlicher Größe besteht, die verdampfen,<br />

koaleszieren o<strong>der</strong> auch aufbrechen können, die aber auch ihre eigene Trägheit und<br />

größenabhängige Dynamik besitzen. Ein häufig verwendeter Modellierungsansatz ist<br />

die Beschreibung <strong>der</strong> zerstäubten, flüssigen Phase im Lagrangeschen Bezugssystem.<br />

Dieser Ansatz liefert detaillierte Informationen auf Mikroebene, aber Tropfen-Interaktionen<br />

wie Koaleszenz und Aufbrechen sind schwierig zu implementieren. Zudem<br />

ist <strong>der</strong> Lagrange-Ansatz, gekoppelt mit den Gleichungen <strong>der</strong> Gasphase im Eulerschen<br />

Bezugssystem unter Annahme <strong>der</strong> Punktquellen-Annäherung, zeitintensiv. Die Alternative<br />

zu Lagrange-Simulationen sind verschiedene Eulersche Methoden, die auf <strong>der</strong><br />

Basis <strong>der</strong> Williams-Spraygleichung entwickelt wurden. Die Beschreibung von dichten<br />

turbulenten Sprayströmungen ist bei Verwendung dieser Euler – Euler Methoden zeiteffizient<br />

und unabhängig von <strong>der</strong> Massenladung <strong>der</strong> flüssigen Phase.<br />

Die Zielsetzung <strong>der</strong> vorliegenden Arbeit ist die Modellierung und Simulation <strong>der</strong><br />

Sprühtrocknung bis zum Beginn <strong>der</strong> Partikelbildung im Eulerschen Bezugssystem. Zur<br />

Untersuchung des Verhaltens <strong>der</strong> Tropfenverteilung unter verschiedenen Trocknungsbedingungen<br />

wurde erstmals die Methode direct quadrature method of moments (DQ-<br />

MOM) <strong>zur</strong> Betrachtung <strong>der</strong> verdampfenden Zweikomponentensprays eingesetzt. In <strong>der</strong>


IV<br />

DQMOM wird die Tropfengrößen- und Geschwindigkeitsverteilung des Sprays modelliert,<br />

indem die Zahlendichtefunktion angenähert wird. Die Transportgleichungen <strong>der</strong><br />

DQMOM berücksichtigen Tropfenverdampfung, Aufheizung, Wi<strong>der</strong>stand und Tropfen-<br />

Tropfen-Interaktionen.<br />

Zuerst wird ein verdampfendes Wasserspray in Stickstoff in eindimensionaler Konfiguration,<br />

d.h. in axialer Richtung des Sprays, modelliert. Frühere Spraystudien vernachlässigten<br />

Verdampfungseffekte o<strong>der</strong> berücksichtigten diese durch ein vereinfachtes<br />

Modell. In dieser Arbeit wird die Tropfenverdampfung jedoch durch das Modell von<br />

Abramzon und Sirignano beschrieben, während Tropfenbewegung und -koaleszenz mit<br />

geeigneten Modellen abgeschätzt werden. Da die Vernachlässigung des Verdampfungsflusses<br />

o<strong>der</strong> seine Berechnung durch Einschränkungen des Gewichtsverhältnisses sich<br />

als unphysikalisch herausstellte, wurde <strong>der</strong> Fluss hier durch die Maximum-Entropie-<br />

Methode berechnet. Die Gasphase ist noch nicht vollständig an die DQMOM gekoppelt,<br />

stattdessen dienen die Gas-Einlaufbedingungen als Grundlage <strong>zur</strong> Berechnung<br />

<strong>der</strong> Kräfte, die auf Tropfen und Verdampfung wirken. Die Resultate <strong>der</strong> Simulationen<br />

werden mit <strong>der</strong> Quadratur-Momentenmethode (QMOM) und Experimenten an verschiedenen<br />

Querschnitten verglichen. Die DQMOM zeigt bessere Ergebnisse als die<br />

QMOM und auch erstaunliche Übereinstimmung mit dem Experiment.<br />

Als nächstes wird das Wasserspray in umgeben<strong>der</strong> Luft in zweidimensionaler, axialsymmetrischer<br />

Konfiguration durch Erweiterung <strong>der</strong> eindimensionalen DQMOM modelliert.<br />

Die DQMOM-Resultate werden mit denen des diskreten Tropfenmodells (DDM),<br />

ein Euler – Lagrange Ansatz, verglichen. Tropfenkoaleszenz wird in <strong>der</strong> DQMOM<br />

berücksichtigt, in <strong>der</strong> DDM aber vernachlässigt. Die Simulationsergebnisse werden<br />

durch aktuelle experimentelle Daten validiert. Insgesamt zeigt die DQMOM deutlich<br />

bessere Recheneffizienz, sogar unter Einschluss <strong>der</strong> Tropfenkoaleszenz.<br />

Bevor die DQMOM auf PVP/Wasser-Sprays erweitert wird, wird ein Verdampfungsund<br />

Trocknungsmodell für einen Einzeltropfen entwickelt, da die meisten <strong>der</strong> in <strong>der</strong> Literatur<br />

bekannten Verdampfungsmodelle auf Salze, Kolloide o<strong>der</strong> Milchpulver angewendet<br />

werden. Das Modell berücksichtigt die Partikelbildung in Zusammenhang mit <strong>der</strong><br />

Tropfenaufheizung und -verdampfung, und die Behandlung <strong>der</strong> flüssigen Mischung als<br />

ideale Lösung wird durch Einschluss nicht-idealer Effekte verbessert. Die Ergebnisse<br />

<strong>der</strong> Simulation dieses Modells werden mit aktuellen experimentellen Daten verglichen,<br />

und es kann gezeigt werden, dass das entwickelte Modell die ersten drei Phasen <strong>der</strong><br />

Verdampfung und des Trocknens eines Zweikomponententropfen effektiv erfassen kann.<br />

Schließlich wird ein PVP/Wasser-Spray in umgeben<strong>der</strong> Luft mittels DQMOM simuliert<br />

unter Anwendung des entwickelten Zweikomponentenverdampfungsmodells. Die<br />

Ergebnisse werden mit aktuellen experimentellen Daten an mehreren Querschnitten<br />

verglichen, und es konnte eine sehr gute Übereinstimmung festgestellt werden.


V<br />

Letztendlich können verdampfende Wasser- und PVP/Wasser-Sprays und die Frühphasen<br />

<strong>der</strong> Sprühtrocknung von PVP/Wasser- und Mannitol/Wasser-Tropfen, d.h. bis<br />

zum Einsetzen <strong>der</strong> Bildung einer festen Schicht, erfolgreich modelliert und simuliert<br />

werden, unter guter Übereinstimmung mit dem Experiment.<br />

Stichwörter: Sprays, PVP, Mannitol, DQMOM, Zweikomponententropfen


Contents<br />

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

I<br />

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

2. Mathematical Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

2.1 State of the Art . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

2.2 Euler – Lagrangian Approach . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.2.1 Gas Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.2.2 Discrete Droplet Model (DDM) . . . . . . . . . . . . . . . . . . 19<br />

2.3 Euler – Euler Approach . . . . . . . . . . . . . . . . . . . . . . . . . . 20<br />

2.3.1 Treatment of the Spray . . . . . . . . . . . . . . . . . . . . . . . 20<br />

2.3.2 NDF Transport Equation . . . . . . . . . . . . . . . . . . . . . 22<br />

2.3.3 Quadrature Method of Moments (QMOM) . . . . . . . . . . . . 23<br />

2.3.4 Direct Quadrature Method of Moments (DQMOM) . . . . . . . 25<br />

2.4 Single Droplet Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

2.4.1 Droplet Heating and Evaporation . . . . . . . . . . . . . . . . . 29<br />

2.4.1.1 Single Component Droplet . . . . . . . . . . . . . . . . 29<br />

2.4.1.2 Bi-component Droplet . . . . . . . . . . . . . . . . . . 32<br />

2.4.2 Droplet Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

2.4.3 Droplet Breakup . . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

2.4.4 Droplet Coalescence . . . . . . . . . . . . . . . . . . . . . . . . 41<br />

3. Numerical Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

3.1 Finite Difference Method for Bi-component Droplet Evaporation and<br />

Solid Layer Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

3.2 Spray Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47<br />

3.2.1 Finite Volume Method for QMOM . . . . . . . . . . . . . . . . 47<br />

3.2.2 Finite Difference Scheme for DQMOM . . . . . . . . . . . . . . 49<br />

3.2.3 Wheeler Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

3.3 Numerical Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . 53


VIII<br />

Contents<br />

4. Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

4.1 One-dimensional Evaporating Water Spray in Nitrogen . . . . . . . . . 55<br />

4.1.1 Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . . . 56<br />

4.1.2 Initial Data Generation . . . . . . . . . . . . . . . . . . . . . . . 56<br />

4.1.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . 58<br />

4.2 Two-dimensional Evaporating Water Spray in Air . . . . . . . . . . . . 66<br />

4.2.1 Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

4.2.2 Initial Data Generation . . . . . . . . . . . . . . . . . . . . . . . 67<br />

4.2.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . 69<br />

4.3 Single Bi-component Droplet Evaporation and Solid Layer Formation . 76<br />

4.3.1 Vapor-Liquid Equilibrium . . . . . . . . . . . . . . . . . . . . . 77<br />

4.3.2 Non-ideal Liquid Mixture . . . . . . . . . . . . . . . . . . . . . 77<br />

4.3.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . 80<br />

4.4 Two-dimensional Evaporating PVP/Water Spray in Air . . . . . . . . . 92<br />

4.4.1 Experiment and Initial Data Generation . . . . . . . . . . . . . 93<br />

4.4.2 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . 93<br />

5. Conclusions and Future Work . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

Appendix 103<br />

A. Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

B. Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111


List of Tables<br />

4.1 Initial weights and abscissas . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

4.2 Droplet size distribution of water spray . . . . . . . . . . . . . . . . . . 61<br />

4.3 Experimental drying conditions . . . . . . . . . . . . . . . . . . . . . . 81<br />

4.4 Experiment vs simulation . . . . . . . . . . . . . . . . . . . . . . . . . 86


X<br />

List of Tables


List of Figures<br />

1.1 A schematic diagram of the spray drying process [3]. . . . . . . . . . . 2<br />

1.2 Chemical structure of polyvinylpyrrolidone (C 6 H 9 NO) n [8]. . . . . . . . 3<br />

1.3 SEM images of spray dried PVP [11]. . . . . . . . . . . . . . . . . . . . 4<br />

1.4 Chemical structure of mannitol (C 6 H 8 (OH) 6 ) [12]. . . . . . . . . . . . . 4<br />

2.1 Sketch of a pressure-atomized spray formation [69, 70]. . . . . . . . . . 11<br />

2.2 A typical log-normal distribution with different values of µ and σ. . . . 21<br />

2.3 Approximation of NDF in QMOM. . . . . . . . . . . . . . . . . . . . . 24<br />

2.4 NDF approximation in DQMOM. . . . . . . . . . . . . . . . . . . . . . 25<br />

2.5 Schematic diagram of stages in single droplet evaporation and drying. . 34<br />

2.6 Droplet breakup mechanisms based on Weber number [69, 178]. . . . . 41<br />

2.7 Droplet collision regimes: (a) bouncing, (b) coalescence [184]. . . . . . . 42<br />

2.8 Droplet collision regimes: (c) reflexive or crossing separation, (d) stretching<br />

separation [184]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

3.1 Flowchart of the DQMOM computational code. . . . . . . . . . . . . . 52<br />

4.1 Photograph of the water spray formation. . . . . . . . . . . . . . . . . . 56<br />

4.2 Schematic diagram of spray with measurement positions. . . . . . . . . 56<br />

4.3 Experimental surface frequency distribution at cross section 0.14 m (left)<br />

and 0.54 m (right) away from the nozzle exit. . . . . . . . . . . . . . . 57<br />

4.4 Comparison of simulated and measured [199] droplet mass and temperature<br />

profiles for the evaporation of a pure water droplet. . . . . . . . . 58<br />

4.5 Homogeneous and inhomogeneous calculations of DQMOM. . . . . . . 59<br />

4.6 Comparison of QMOM and DQMOM results with experiment. . . . . . 60<br />

4.7 Droplet size distribution of water spray and at t = 0 s, and at t = 1 s<br />

with d 2 law evaporation rate. . . . . . . . . . . . . . . . . . . . . . . . 60<br />

4.8 Velocity profiles of three droplets with different initial radii and velocities<br />

un<strong>der</strong> the influence of drag alone (left) and drag and gravity (right). . . 62<br />

4.9 Effect of liquid inflow rates on Sauter mean diameter computed with<br />

and without coalescence. . . . . . . . . . . . . . . . . . . . . . . . . . . 62


XII<br />

List of Figures<br />

4.10 Profiles of Sauter mean diameter (left) and mean droplet diameter (right)<br />

computed with and without coalescence at surrounding gas temperatures<br />

of 293 K and 313 K. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

4.11 Profiles of specific surface area computed with and without coalescence<br />

at surrounding gas temperatures of 293 K and 313 K. . . . . . . . . . . 64<br />

4.12 Profiles of droplet number density computed with and without coalescence<br />

at surrounding gas temperatures of 293 K and 313 K. . . . . . . 65<br />

4.13 Schematic diagram of the experimental setup. . . . . . . . . . . . . . . 66<br />

4.14 Profile of effective cross-section area of the probe volume for measured<br />

droplet size. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

4.15 Experimental and DQMOM approximation of droplet number density<br />

for a water spray. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68<br />

4.16 Experimental and reconstructed NDF of 80 kg/h water spray at 64.5 mm<br />

from the center of the spray axis. . . . . . . . . . . . . . . . . . . . . . 69<br />

4.17 Experimental and reconstructed NDF of 121 kg/h water spray at 81 mm<br />

from the center of the spray axis. . . . . . . . . . . . . . . . . . . . . . 70<br />

4.18 Experimental and numerical profiles of the Sauter mean diameter of water<br />

spray with 80 kg/h liquid flow rate at the cross section of 0.12 m (left)<br />

and 0.16 m (right) distance from the nozzle exit. . . . . . . . . . . . . . 71<br />

4.19 Experimental and numerical profiles of the Sauter mean diameter at the<br />

cross section of 0.12 m distance from the nozzle exit for 120 kg/h. . . . 71<br />

4.20 Experimental and numerical profiles of the Sauter mean diameter at the<br />

cross section of 0.16 m distance from the nozzle exit for 120 kg/h. . . . 72<br />

4.21 Experimental and numerical profiles of the mean droplet diameter of water<br />

spray with 80 kg/h liquid flow rate at the cross section of 0.12 m (left)<br />

and 0.16 m (right) distance from the nozzle exit. . . . . . . . . . . . . . 73<br />

4.22 Experimental and numerical profiles of the mean droplet diameter of water<br />

spray with 120 kg/h liquid flow rate at the cross section of 0.12 m (left)<br />

and 0.16 m (right) distance from the nozzle exit. . . . . . . . . . . . . . 74<br />

4.23 Experimental and numerical profiles of the mean droplet velocity at the<br />

cross section of 0.12 m distance from the nozzle exit. . . . . . . . . . . 75<br />

4.24 Experimental and numerical profiles of the mean droplet velocity at the<br />

cross section of 0.16 m distance from the nozzle exit. . . . . . . . . . . 75<br />

4.25 Numerical and experimental [214] results of water activity (a w ) in PVP/water<br />

solution at 73.0 ◦ C (left) and 94.5 ◦ C (right). . . . . . . . . . . . . . . . 78<br />

4.26 Numerical results of water activity (a w ) in mannitol/water solution at<br />

94.5 ◦ C and 160 ◦ C. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78


List of Figures<br />

XIII<br />

4.27 Effect of non-ideality on the vapor pressure of water at different temperatures.<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

4.28 Variation of vapor pressure of water with water mass fraction in PVP/water<br />

solution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

4.29 Experimental data of PVP [218] and mannitol [219] saturation solubility<br />

in water. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

4.30 Effect of gas velocity on the evolution of mass and temperature of a<br />

mannitol/water droplet. . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

4.31 Effect of elevated gas temperature on the surface area of a mannitol/water<br />

droplet. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

4.32 Effect of gas temperature on the temporal development of mannitol mass<br />

fraction inside the droplet at 0.5 s (left) and 0.9 s (right). . . . . . . . . 83<br />

4.33 Effect of gas velocity and temperature on porosity and final particle size<br />

of mannitol/water droplet. . . . . . . . . . . . . . . . . . . . . . . . . . 84<br />

4.34 SEM images of mannitol samples spray dried at 70 ◦ C (a), 100 ◦ C (b)<br />

and 90 ◦ C (c). Zoomed images of the surface structures of these particles<br />

at 70 ◦ C (d), 100 ◦ C (e) and 90 ◦ C (f) [225]. . . . . . . . . . . . . . . . . 85<br />

4.35 Effect of gas temperatures of 60 ◦ C and 95 ◦ C and relative humidity of<br />

1% R.H. (left) and 30% R.H. (right) on the droplet surface area. . . . . 86<br />

4.36 Time evolution of mannitol/water droplet surface area computed by<br />

present model and RMM. . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

4.37 Effect of initial droplet temperature on the evaporation rate of mannitol/water<br />

droplet. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

4.38 Effect of gas temperature on solid layer thickness inside the PVP/water<br />

droplet. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

4.39 Effect of gas temperature on the droplet mass and temperature. . . . . 89<br />

4.40 Temporal development of PVP mass fraction profiles inside the droplet<br />

subjected to dry air (left) and hot air with 5% R.H (right). . . . . . . . 90<br />

4.41 Time evolution of PVP/water droplet surface area predicted by present<br />

model and RMM. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90<br />

4.42 Effect of relative humidity on the water evaporation rate from the mannitol/water<br />

droplet. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

4.43 Effect of initial PVP mass fraction on the profiles of droplet radius and<br />

temperature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92<br />

4.44 Photograph of the PVP/water spray formation with 112 kg/h liquid<br />

inflow rate in experiment [206]. . . . . . . . . . . . . . . . . . . . . . . 93<br />

4.45 Experimental and DQMOM approximation of droplet number density<br />

for PVP/water spray. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94


XIV<br />

List of Figures<br />

4.46 Experimental and numerical profiles of the Sauter mean diameter (left)<br />

and mean droplet diameter (right) of PVP/water spray in air at the<br />

cross section of 0.12 m distance from the nozzle exit. . . . . . . . . . . 94<br />

4.47 Experimental and numerical profiles of the Sauter mean diameter (left)<br />

and mean droplet diameter (right) of PVP/water spray in air at the<br />

cross section of 0.16 m distance from the nozzle exit. . . . . . . . . . . 95<br />

4.48 Experimental and numerical profiles of the mean droplet velocity of<br />

PVP/water spray in air at the cross section of 0.12 m distance from<br />

the nozzle exit. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96<br />

4.49 Experimental and numerical profiles of the mean droplet velocity of<br />

PVP/water spray in air at the cross section of 0.16 m distance from<br />

the nozzle exit. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97


1. Introduction<br />

Drying has found applications in many areas of industry such as chemical, food, pharmaceutical,<br />

polymer, ceramics and mineral processing. The basic idea of drying is to<br />

remove liquid by evaporation from the material that has either dissolved or suspended<br />

solids to a dried pow<strong>der</strong>. The objectives can be, to reduce the transportation costs,<br />

to increase the shelf life of a material, which is done for instance in the preservation<br />

of milk, tomato, etc. or because the material has better properties in dry form than<br />

when dissolved.<br />

Spray drying is one of the most widely used drying techniques and it is a process<br />

for converting a liquid feed into a pow<strong>der</strong> by evaporating the solvent. The other drying<br />

techniques to produce pow<strong>der</strong>s mainly include freeze drying, supercritical drying and<br />

vacuum drying. Compared to other evaporation processes, spray drying has the great<br />

advantage that products can be dried without much loss of volatile or thermally unsteady<br />

or degradable compounds. These advantages are especially important in the production<br />

of pharmaceutical bulk materials such as polymers (e.g. polyvinylpyrrolidone),<br />

carbohydrates (e.g. mannitol) and food pow<strong>der</strong>s (e.g. milk and coffee pow<strong>der</strong>s) [1, 2].<br />

Spray dryers are extensively used in industry, usually placed at the end-point of a<br />

process plant, which play an important role in the whole process, not because of their<br />

capital investment, size and operating costs but mainly because of their high energy<br />

efficiency and throughput production.<br />

Spray drying can be structured into several steps primarily consisting of the atomization<br />

of liquid feed into a population of poly-disperse droplets followed by the<br />

convection of droplets and gas, the evaporation of the liquid solvent from the droplets,<br />

droplet-droplet collisions, which might lead to coalescence, aggregation and eventually<br />

breakup. In some cases, chemical reactions are also involved, e.g. in the production<br />

of polymer via monomer polymerization using spray drying. These sub-processes are<br />

inter-dependent. The atomization leads to droplets with specific size distribution and<br />

kinetic energy, thus influencing the droplets convection and probability of collision.<br />

The drying gas in most of the cases is the ambient air heated to a desired temperature<br />

with controlled relative humidity and this hot air is either co-flow or counter-flow<br />

depending on the system requirement and the thermal sensibility of the material. A<br />

schematic diagram of the typical spray drying process in shown in Fig. 1.1 [3]. A review<br />

of available spray drying designs, process types and applications is given by Masters [1].


2 1. Introduction<br />

Fig. 1.1: A schematic diagram of the spray drying process [3].<br />

Spray drying is a complex process and it is very difficult to predict the quality,<br />

properties and characteristics of the product such as particle size, density, porosity,<br />

flowability, shape, compressibility, etc. for the given drying conditions. The industrial<br />

practice to design is always based on the field experience and know-how followed by<br />

experimentation in pilot plant trials, which can be very expensive in case of rare materials<br />

[1, 4]. Problems associated with scale-up and hydrodynamics of the driers have<br />

resulted in limited success. The progress in computational techniques and computing<br />

prowess has given the advantage to develop a robust model of heat and mass transfer<br />

based on the equations of fluid flow within the spray chamber, which can predict the<br />

whole spray drying process i.e., from the liquid feed stock entering into the dryer to the<br />

end product, thereby resulting in cost effective and economic spray drier designs [1, 5].<br />

The modeling of spray flows and the spray drying has been studied for many years<br />

now, which is done to predict the characteristics of the spray drying end products,<br />

e.g. particle size, shape, density or porosity. The previous studies mainly concern the<br />

systems of whole milk, salt solution, colloids or other materials for which the physical<br />

and thermal properties are well tabulated. The simulation of polymer or sugar solution<br />

spray drying is not studied so far because of unknown drying behavior of these materials<br />

and unavailability of properties.<br />

The present study deals with the modeling and simulation of polyvinylpyrrolidone<br />

(PVP)/water and mannitol/water spray flows, with an aim to predict the effect<br />

of drying conditions on the evolution of droplet properties. The synthesis, applications<br />

of PVP and mannitol, and the motivation to choose these systems are elaborated in<br />

the following paragraphs.<br />

PVP is a water-soluble polymer made from the monomer N-vinylpyrrolidone [6, 7].


3<br />

Fig. 1.2: Chemical structure of polyvinylpyrrolidone (C 6 H 9 NO) n [8].<br />

It is a unique polymer providing a remarkable combination of properties that no other<br />

molecule is yet able to match. PVP offers a variety of properties, such as good initial<br />

tack, transparency, chemical and biological inertness, very low toxicity as well as high<br />

media compatibility and cross linkable flexibility. PVP was first synthesized by Prof.<br />

Walter Reppe and a patent was filed in 1939 for one of the most interesting <strong>der</strong>ivatives<br />

of acetylene chemistry. Soluble PVP is obtained by free-radical polymerization<br />

of vinylpyrrolidone in water or 2-propanol, yielding the chain structure as shown in<br />

Fig. 1.2. There exists several grades of PVP, which are classified based on molecular<br />

weight, and spray drying technology is used in production of all types of PVPs [8].<br />

PVP was initially used as a blood plasma substitute and later in a wide variety<br />

of applications in medicine, pharmacy, cosmetics and industrial production [9]. It is<br />

used as a bin<strong>der</strong> in many pharmaceutical tablets; it simply passes through the body<br />

when taken orally. PVP binds to polar molecules exceptionally well, owing to its<br />

polarity. This has led to its application in coatings for photo-quality ink-jet papers<br />

and transparencies, as well as in inks for ink jet printers. PVP is also used in personal<br />

care products, such as shampoos and toothpastes, in paints, and adhesives that must<br />

be moistened, e.g. old-style postage stamps and envelopes [10]. It has also been used<br />

in contact lens solutions and in steel-quenching solutions. PVP is the basis of the<br />

early formulas for hair sprays and hair gels, and still continues to be a component of<br />

some [10]. As a food additive, PVP is a stabilizer and has E number E1201 [9]. In year<br />

2006, the total world wide production of PVP was 31,000 tonnes, out of which 47% was<br />

used in cosmetics and 27% in pharmaceuticals [9]. Figure 1.3 [11] shows the scanning<br />

electron microscope (SEM) images of PVP pow<strong>der</strong> produced via spray drying. In the<br />

evaporation and drying of PVP dissolved in water droplet, the molecular entanglement<br />

prior to solid layer formation is observed, which is different from the colloids, silica<br />

or salt in water droplet where crust formation is found. It is observed that the spray<br />

drying yields hollow or solid particles with spherical or non-spherical shape but the<br />

initial droplet size, gas temperature and velocity and other drying conditions show<br />

enormous effect on the final pow<strong>der</strong> characteristics such as flowability, particle density,


4 1. Introduction<br />

Fig. 1.3: SEM images of spray dried PVP [11].<br />

porosity and shape [1]. To produce PVP pow<strong>der</strong> with specific required properties<br />

is very important, for example, a uniform particle size distribution of PVP pow<strong>der</strong><br />

in pharmaceuticals not only helps in flowability of the pow<strong>der</strong> but also improves the<br />

appealing of the final product.<br />

Similar to PVP, mannitol has several useful applications. Mannitol is a sugar<br />

alcohol and it is widely used as a carrier particle in tablets. Mannitol is commonly<br />

produced via the hydrogenation of fructose, which is formed from either starch or<br />

sucrose (common table sugar) [13]. Although starch is a cheaper source than sucrose,<br />

the transformation of starch is much more complicated. Hydrogenation of starch yields<br />

a syrup containing about 42% fructose, 52% dextrose, and 6% maltose [13]. Sucrose is<br />

simply hydrolyzed into an invert sugar syrup, which contains about 50% fructose. In<br />

both cases, the syrups are chromatographically purified to contain 90–95% fructose [13].<br />

The fructose is then hydrogenated over a nickel catalyst into mixture of isomers sorbitol<br />

and mannitol with a typical yield of 50% sorbitol and 50% mannitol [13]. The chemical<br />

structure of mannitol is shown in Fig. 1.4 [12].<br />

For many years active pharmaceutical ingredients (APIs) are delivered to the lung<br />

Fig. 1.4: Chemical structure of mannitol (C 6 H 8 (OH) 6 ) [12].


5<br />

via inhalation aerosols. Dry pow<strong>der</strong> inhalers (DPI) are commonly used to achieve<br />

aerosols of a micronized solid API. To guarantee a reliable and constant dosing, the<br />

flow properties of the formulation are of high interest. Since the micronized API with<br />

a particle size of 1 µm - 5 µm [14] exhibits poor flowability, carriers consisting of larger<br />

particles are added to the formulation in or<strong>der</strong> to carry the API particles on their<br />

surface. Due to the sufficiently large size of the carrier particles, the adhesive mixtures<br />

exhibit adequate flowability. In addition to the flowability the surface structure<br />

of the carrier is crucial to the formulation performance [15] and has to be controlled<br />

during development. In the last decade, mannitol was identified as a possible carrier<br />

for DPIs [16] and efforts were made to tailor the surface structure [17–19]. Spray<br />

dried mannitol particles in general have a spherical shape and can consist of two major<br />

polymorphs [17, 20]. Similar to PVP, recent studies [21, 22] of mannitol spray<br />

drying reveal that process parameters like droplet size, gas temperature and relative<br />

humidity exhibit strong correlation with the final pow<strong>der</strong> characteristics. Compared to<br />

PVP/water, evaporation and drying of mannitol/water droplet leads to crust formation<br />

on the droplet surface prior to complete dried particle.<br />

The growing attention and wide applications of PVP and mannitol, as well as the<br />

effects of spray drying process on the final pow<strong>der</strong> characteristics explains the particular<br />

interest towards these systems. The scarcity in thermal and physical properties<br />

of PVP and mannitol, and unknown behavior of evaporation and drying of<br />

PVP/water as well as mannitol/water droplets is a challenge for developing a mathematical<br />

model, numerical technique and validation. This project is part of the German<br />

Science Foundation (DFG) priority program ”SPP1423”, where spray flows and<br />

spray drying of PVP/water and mannitol/water are exclusively studied. This thesis<br />

deals with mathematical modeling and numerical simulation of both mono and bicomponent<br />

(PVP/water and mannitol/water) droplet evaporation and dispersion in<br />

sprays and spray drying with an objective to numerically investigate the effect of drying<br />

conditions such as gas temperature, gas velocity, relative humidity, initial droplet<br />

size and velocity distribution on the evolution of droplet properties, which will enable<br />

in better un<strong>der</strong>standing the spray flows thereby helps in designing the spray dryer.<br />

The computational methods in the area of multiphase flow can primarily be categorized<br />

into two methods, (1) Lagrangian particle tracking method and (2) Euler – Euler<br />

or two continua/fluid methods. In both of these classical approaches, the continuous<br />

gas phase is modeled using the Navier – Stokes equations. Consi<strong>der</strong>ing the presence<br />

of turbulence in the system, the Navier – Stokes equations can be solved on a fine<br />

computational mesh, which allows to capture all the macroscopic structures since all<br />

the consi<strong>der</strong>ed length scales are consi<strong>der</strong>ably larger than the molecular length and time<br />

scales. Such a numerical resolution is defined as direct numerical simulation (DNS).


6 1. Introduction<br />

DNS solves all the characteristic scales of a turbulent flow and it requires no modeling<br />

of scales. DNS is implemented in a numerous flow problems for example particle dispersion<br />

[23, 24], turbulent reacting flows [25], spray flames [26], etc. A complete DNS<br />

of the spray drying is still not imaginary due to its high computational efforts [27, 28].<br />

The alternative to DNS is the large eddy simulation (LES) [29–34] in which the<br />

large eddies are resolved and small eddies are modeled using a subgrid-scale model.<br />

This method requires spatial and temporal resolution of the scales in inertial subrange.<br />

The main disadvantage of LES method is that the accuracy of the flow field depends on<br />

the subgrid-model and filter size. Still, large computational time, and storage analysis<br />

of the huge data sets pose significant problems.<br />

The attractive approach to solve Navier – Stokes equations is the Reynolds-averaged<br />

Navier – Stokes (RANS) numerical simulation [35–37], where the instantaneous Navier –<br />

Stokes equations are averaged with respect to time whereby an instantaneous quantity<br />

is decomposed into its time-averaged and fluctuating quantities. RANS equations together<br />

with the turbulence closure model can be solved to resolve the turbulent flow and<br />

compute the mean flow field quantities [35–37]. There are several RANS turbulence<br />

closure models, which are extensively discussed by Pope [37], and the most notable<br />

turbulence models include k – ɛ and extended k – ɛ model [37].<br />

The other methods like volume of fluid (VOF) [38] and lattice-Boltzmann (LB) [39]<br />

approaches also exist in the multi-phase flows to model the flow around the droplets<br />

or particles, and therefore the fluid flow can be fully resolved. These methods may<br />

be characterized un<strong>der</strong> the DNS method for multi-phase flow problems to define the<br />

interfaces.<br />

In the Lagrangian particle tracking method, the droplets are injected into the gas<br />

and their trajectories are tracked by numerically evaluating the Lagrangian equations<br />

of motion. A typical spray consists of a large number of droplets and with limited<br />

computational resources, numerical parcels are implemented instead of droplets where<br />

each parcel contains of several number of droplets [35].<br />

In Euler – Lagrangian approach, droplet–droplet interactions such as coalescence<br />

and breakup, which occur quite frequently in spray flows are difficult to account for due<br />

to computational complexity. The computational cost can be very expensive due to the<br />

large number of droplets needed to reach the statistical convergence, and computational<br />

cost is also dependent on mass loading of the dispersed phase.<br />

In the two-continua method, also known as Euler – Euler approach, a set of conservation<br />

equations is written for each phase, and the sets are coupled through their<br />

respective source terms. This was first proposed by Elghobashi and Abou-Arab [40]<br />

with the aim to establish a two-phase turbulence model. They <strong>der</strong>ived a two-equation<br />

model based on the principle of k – ɛ model, and the Reynolds-averaged conservation


7<br />

equations are written in terms of volume fraction of each phase such that the sum of<br />

the volume fractions is unity. When the phases are equally distributed in the domain of<br />

interest with only mo<strong>der</strong>ate separation between the phases, the classical Euler – Euler<br />

approach is appropriate.<br />

Euler – Euler methods offer significant advantages over the Euler – Lagrange approach,<br />

e.g. the two-continua method is independent of disperse phase mass loading,<br />

and also the coupling between dispersed and carrier phase does not require averaging<br />

over the parcels unlike in Euler – Lagrangian.<br />

The merits of Euler – Euler methods play an important role when unsteady, turbulent<br />

gas-liquid flows with high dispersed phase mass loading are consi<strong>der</strong>ed. Additionally,<br />

Euler – Euler methods can outperform the Euler – Lagrange in case of unsteady<br />

spray flows and the computational cost do not depend on the droplet mass loading.<br />

Most of the Euler – Euler methods in the field of spray flows are based on the description<br />

of dispersed phase as a number density function (NDF) and the evolution of<br />

this NDF due to physical processes of spray flows are described by the NDF transport<br />

equation, also known as population balance equation (PBE) [41]. This NDF transport<br />

equation is <strong>der</strong>ived based on the kinetic equation [42] similar to the molecular kinetic<br />

theory, and it is known as general particle-dynamic equation in the field of aerosol<br />

science [43, 44].<br />

There exists several Euler – Euler methods based on the kinetic equation such as<br />

Williams’ spray equation and are categorized mainly as multi-fluid methods [45–47] and<br />

moment based methods [48–54]. In the multi-fluid approach, the distribution function<br />

is discretized using a finite volume technique that yields conservation equations for mass<br />

and momentum of droplets in fixed size intervals called sections or fluids [46]. This<br />

approach has recently been extended to higher or<strong>der</strong> of accuracy [55], but discretization<br />

of droplet size space is still a problem that needs to be addressed. On the contrary,<br />

moment based methods such as quadrature method of moments (QMOM) [51, 52, 56–<br />

58] or direct quadrature method of moments (DQMOM) [53, 54, 59] do not pose this<br />

problem and they are found to be efficient and robust in the poly-disperse multiphase<br />

flow problems.<br />

The scope of this work is modeling and simulation of mono and bi-component evaporating<br />

spray flows in an Euler – Euler framework. The focus is on the description of<br />

the characteristics of the spray flows and spray drying process, and the influence of the<br />

droplet size distribution on the droplet properties. In particular, spray inhomogeneity<br />

associated with the atomization process and its transport in the convective medium is<br />

not well un<strong>der</strong>stood. Subtle information is available about the particle formation and<br />

its influence on the properties of the resulting pow<strong>der</strong> in spray drying.<br />

The present study aims to develop a comprehensive spray model, which can be


8 1. Introduction<br />

used to simulate the bi-component droplet evaporation and dispersion in spray flows<br />

and spray drying, and to predict the evolution of the droplet properties. In or<strong>der</strong><br />

to un<strong>der</strong>stand the behavior of droplet distribution un<strong>der</strong> various drying conditions,<br />

the droplet size and velocity distribution of the spray is modeled using DQMOM.<br />

Transport equations of DQMOM should account for droplet evaporation, heating, drag<br />

and droplet–droplet interactions, which are calculated through appropriate sub-models.<br />

The systems of interest for the current study include water spray in nitrogen in one<br />

physical dimension, water and polyvinylpyrrolidone (PVP)/water spray in air in twodimensional,<br />

axisymmetric configuration. These systems were chosen for implementing<br />

and validating the numerical results. The experimental data for water spray in onedimensional<br />

configuration was provided by Dr. R. Wengeler, BASF, Ludwigshafen,<br />

whereas Prof. G. Brenn, TU Graz, Austria, provided the data for the two-dimensional<br />

water and PVP/water spray flows. The system of PVP or mannitol dissolved in water<br />

is also consi<strong>der</strong>ed to verify numerical results of the single bi-component droplet<br />

evaporation and solid layer formation.<br />

Earlier studies based on DQMOM for spray flows either used simplified model for<br />

evaporation [60] or neglected the evaporation itself [61]. In the present study, the evaporation<br />

rate is computed using an advanced droplet evaporation model of Abramzon and<br />

Sirignano [62] for water spray, which accounts for variable liquid and film properties and<br />

includes the convective effects. For bi-component spray, the existing models to compute<br />

the evaporation rate neglect the non-ideality effect induced by non-evaporating<br />

component (e.g. PVP or mannitol) and ignore the solid layer resistance on the evaporation<br />

and droplet heating. In this study, for both PVP/water and mannitol/water<br />

droplet evaporation and solid layer formation, a mathematical model is formulated<br />

consi<strong>der</strong>ing the non-ideality effect and solid layer resistance in the evaporation and<br />

heating. The evaporative flux, which is a point-wise quantity of the number density<br />

function of zero-droplet size, was either assumed to be zero [61, 63] or estimated with<br />

weight ratio constraints [60]. However, the later procedure was found to pose problems<br />

or to behave unphysical in multi-variate distributions [64, 65]. In the present study,<br />

evaporative flux is computed using the maximum entropy formulation [65–67].<br />

The objectives of the current study are: modeling the evaporating water spray<br />

in nitrogen using DQMOM in one physical dimension and comparison of simulation<br />

results with QMOM and experiment, followed by extending the transport equations<br />

of DQMOM to two dimensions to simulate water spray in air, which is done for the<br />

first time, in axisymmetric configuration. The validation of DQMOM results with the<br />

discrete droplet (DDM) model [68], which is a well established Euler – Lagrangian<br />

technique, and with the new experimental data. In these configurations, the various<br />

physical processes due to gas–liquid and droplet–droplet interactions are accounted


9<br />

for through appropriate sub-models. In or<strong>der</strong> to simulate PVP/water spray flows,<br />

existing evaporation model needs modifications to include the effects of non-ideality<br />

and solid layer formation on droplet heating and evaporation, so the present work aims<br />

to develop a mathematical model, which can predict the bi-component single droplet<br />

evaporation and solid layer formation prior to drying with prerequisites to account for<br />

non-ideality of liquid mixture and effect of solid layer resistance on droplet heating and<br />

evaporation. Final objective is to extend DQMOM to simulate PVP/water spray flows<br />

using the developed bi-component evaporation model and subsequent verification of<br />

the simulation results with that of the new experimental data. Complete spray drying<br />

is not yet simulated, which requires coupling of gas phase with the DQMOM and<br />

accounting for droplet temperature in transport equations of DQMOM.<br />

The dissertation is grouped into the following chapters. A review of the numerical<br />

simulation of sprays, governing equations of DDM and QMOM followed by the<br />

development of DQMOM and its transport equations are described in Chapter 2. The<br />

details of individual source terms of the spray flows such as bi-component droplet evaporation,<br />

droplet motion and droplet–droplet interactions and their equations are also<br />

elucidated in Chapter 2. In Chapter 3, numerical schemes to solve the single droplet<br />

evaporation and solid layer formation as well as DQMOM transport equations are explained<br />

along with the solution procedure of QMOM and Euler – Lagrangian approach<br />

DDM. A Wheeler algorithm to compute the initial data for DQMOM simulations and<br />

closure for QMOM unclosed moments is also given in Chapter 3. Chapter 4 presents<br />

the results and discussion starting from the water spray in nitrogen in one-dimension,<br />

followed by water spray in air in two-dimensional, axisymmetric configuration. The single<br />

droplet evaporation and solid layer formation model results are presented for both<br />

mannitol/water and PVP/water droplets. Finally, the results of PVP/water spray are<br />

presented. The conclusions and perspective future work are given in Chapter 5.


10 1. Introduction


2. Mathematical Modeling<br />

Spray constitutes of poly-disperse liquid droplets dispersed in gas medium. A typical<br />

sketch of a pressure-atomized spray breakup and its development is depicted in<br />

Fig. 2.1 [69, 70]. Here poly-disperse means that the properties of the disperse phase<br />

entities can be different for each entity. For example, evaporating sprays have a region<br />

near the nozzle where the liquid jet is not disperse, followed by a region after breakup<br />

of the primary jet that is composed of individual droplets having different properties<br />

such as size, velocity, temperature etc., which are defined as the poly-disperse<br />

droplets [41, 71]. To describe the poly-disperse characteristics of the spray flows, the<br />

mathematical modeling approach of Euler – Lagrangian and Euler – Euler framework<br />

is discussed, and the sub-models for the physical processes of sprays such as evaporation,<br />

forces, coalescence and breakup are elucidated in this chapter. Though the focus<br />

of this work is to model the spray flows using the direct quadrature method of moments<br />

(DQMOM), but the models like quadrature method of moments (QMOM) and<br />

discrete droplet model (DDM), which were used to compare and validate the DQMOM<br />

results, are also presented in this chapter.<br />

2.1 State of the Art<br />

The existing modeling approaches in the area of multiphase flows mainly include, (1)<br />

Lagrangian particle tracking method and (2) Euler – Euler or two continua/fluid meth-<br />

Fig. 2.1: Sketch of a pressure-atomized spray formation [69, 70].


12 2. Mathematical Modeling<br />

ods. In the Lagrangian particle tracking method, the gas phase behavior is typically<br />

predicted by solving the unsteady Reynolds-averaged Navier – Stokes equations with<br />

an appropriate turbulence model and sub-models for various source terms [35, 37, 72].<br />

In this method, droplets are injected into the gas and their trajectories are tracked by<br />

numerically evaluating the Lagrangian equations of motion.<br />

A typical spray consists of a large number of droplets and with limited computational<br />

resources, numerical parcels are implemented instead of droplets where each<br />

parcel contains several number of droplets. The Euler–Lagrangian models are classified<br />

as locally homogeneous flow (LHF) method [73–75] and separated flow (SF)<br />

method [76–78].<br />

The locally homogenous flow approximation of the LHF model for two-phase flow<br />

problems implies that the interphase transport rates are infinitely fast, so that both<br />

phases have same velocity and are in thermal equilibrium at each point of the flow [75].<br />

This model neglects the slip effect between the liquid phase and gas phase. LHF<br />

approximation is the limiting case, which accurately represents spray with very small<br />

droplets [72].<br />

Compared with the LHF model, the SF model has been used more widely in multiphase<br />

flow problems, because it provides the finite rate exchange of mass, momentum<br />

and energy between the phases [72]. The SF model assumes that each phase<br />

displays different properties and flows with different velocities, but the conservation<br />

equations are written only for the combined flow. In addition, the pressure across<br />

any given cross-section of a channel carrying a multiphase flow is assumed to be the<br />

same for both phases [72]. The SF models are further subdivided into discrete droplet<br />

model (DDM) [76–78], continuous droplet model (CDM) and continuous formulation<br />

model (CFM). The differences in these methods are explained by Faeth [72].<br />

In DDM, the spray is divided into representative samples of discrete droplets whose<br />

motion and transport are tracked through the flow field, using a Lagrangian formulation.<br />

This procedure computes the liquid properties based on finite number of particles,<br />

called as parcels which are used to represent the entire spray [72, 76]. The gas phase is<br />

solved using Eulerian formulation, similar to the LHF method. The effect of droplets<br />

on the gas phase is consi<strong>der</strong>ed by introducing appropriate source terms in the gas<br />

phase equations of motion. This type of formulation is is found to be convenient for<br />

consi<strong>der</strong>ing a relatively complete representation of droplet transport processes [72].<br />

The CDM was first introduced by Williams [79]. In this method, droplet properties<br />

are represented by a statistical distribution function defined in terms of droplet<br />

diameter, position, time, velocity, temperature, etc. [80]. Conservation principles yield<br />

a transport equation for the distribution function, which is solved along with the gas<br />

phase equations to deduce the properties of the spray [72, 79, 80]. Similar to DDM,


2.1. State of the Art 13<br />

the governing equations for the gas phase include appropriate source terms to compute<br />

the effects of droplets.<br />

The other important SF method for modeling sprays is the CFM, which employs<br />

a continuum formulation of the conservation equations for both phases [81, 82]. The<br />

motion of both droplets and gas are treated as interpenetrating continua. The work of<br />

Faeth [72] gives an extensive review of all the Euler – Lagrangian models.<br />

The Euler – Lagrangian approach is so far consi<strong>der</strong>ed to be effective in many applications,<br />

which gives detailed information at the micro-level, however it has significant<br />

drawbacks as listed by Archambault [67]. For instance, inclusion of droplet–droplet<br />

interactions such as coalescence and breakup, which occur quite frequently in spray<br />

flows, increases the computational complexity. The computational cost could be very<br />

expensive due to the large number of droplets needed to reach the statistical convergence,<br />

and it may pose difficulties and numerical instabilities in coupling of Lagrangian<br />

description of dispersed phase with the Eulerian equations of the gas phase. The computational<br />

cost is also dependent on mass loading of the dispersed phase. According<br />

to Archambault [67], the vertices of the droplet trajectory and numerical grid of the<br />

gas phase never coincide, hence a sub-grid model is required in or<strong>der</strong> to compute the<br />

exchange rate between the phases [83]. Grid independent solutions are quite difficult to<br />

obtain [84], which could be because of an insufficient number of droplets in a grid cell<br />

leading to a significant error as can be observed in the regions of high droplet number<br />

density.<br />

The study of Garcia et al. [85] and Riber et al. [86] describe and analyze the<br />

comparison of computational time between Euler – Euler and Euler – Lagrangian in<br />

homogeneous and non-homogeneous flows.<br />

There is a tremendous amount of literature available on the Eulerian – Lagrangian<br />

approaches in spray flows and spray drying [76, 87–93], and references therein. As the<br />

focus of the current work is about Euler – Euler approach to spray flows, this section<br />

presents the review of available literature in this area.<br />

A numerous Eulerian models have been recently developed where the disperse phase<br />

described based on a kinetic equation and continuum phase is resolved using Navier–<br />

Stokes equations. The basic idea in kinetic equation based Eulerian methods is that<br />

instead of solving the usual Euler equations for the dispersed phase, the evolution of the<br />

moment transform of the kinetic equation is solved, which resembles Navier – Stokes -<br />

like equation, and this equation is coupled to the continuum phase with the appropriate<br />

source terms. Such a kinetic equation is first <strong>der</strong>ived by Williams [42], known as<br />

Williams’ spray equation which is analogous to Boltzmann’s equation of molecules [94,<br />

95]. The <strong>der</strong>ivation of Williams’ spray equation is given by Archambault [67] and<br />

Ramakrishna [41]. This equation describes the temporal evolution of the probable


14 2. Mathematical Modeling<br />

number of droplets within a range of droplet characteristics such as size, velocity,<br />

temperature and solute mole fraction within the droplet, which constitutes the phasespace<br />

at a spatial location. The solution of this equation coupled with the gas phase<br />

flow field equations provides the average properties of the spray, for example mean<br />

droplet diameter, Sauter mean droplet diameter, mean droplet velocity and many other<br />

statistical properties.<br />

Among the existing Eulerian models, the multi-fluid method allows the detailed<br />

description of poly-disperse droplet size and velocity through correlations. Such an<br />

approach has been shown to be <strong>der</strong>ived from the Williams’ spray equation, Eq. (2.23),<br />

by Laurent and Massot [45] un<strong>der</strong> the mono-kinetic spray assumption, which states<br />

that the velocity dispersion of the spray distribution function at a given time, spatial<br />

location and droplet size has to be zero. This assumption is important since it defines<br />

the validity limit of the multi-fluid model and also results in the ”pressure-less gas<br />

dynamics” structure of the transport equations for conservation of mass and momentum<br />

of droplets [96]. These conservation equations are <strong>der</strong>ived by discretizing the droplet<br />

distribution using a finite volume technique in fixed size intervals called sections or<br />

fluids.<br />

This approach has been extended to higher or<strong>der</strong> of accuracy [46], but discretization<br />

of droplet size phase-space is still a problem that needs to be addressed. The issues<br />

related to the mono-kinetic assumption have to be relaxed if the coalescence of droplets<br />

are to be consi<strong>der</strong>ed, which is addressed by introducing a semi-kinetic equation and<br />

the results are presented for the evaporation and coalescence in spray flows [47, 97].<br />

However, the validation in multi-dimensional configurations and the evaluation of the<br />

level of accuracy of such model versus the reference Lagrangian simulations as well<br />

as the related issue of a detailed study of the effective computational cost of the two<br />

approaches is not yet un<strong>der</strong>stood [98].<br />

Next most notable method is the method of classes (CM) or discretized population<br />

balances (DPB) which is based on the discretization of droplet internal co-ordinates<br />

of the population balance equation [99–102] into a finite series of bins. The CM’s<br />

compute the mean properties of the population such as droplets or particles within<br />

these bins by solving the discretized population balance equation. CM’s are divided<br />

into two categories namely, (1) zero-or<strong>der</strong> methods, and (2) higher or<strong>der</strong> methods. In<br />

zero-or<strong>der</strong> methods, the droplet size distribution (DSD) is consi<strong>der</strong>ed to be constant<br />

in each class, and they are ”extremely stable”. Recently, Vanni [103] reviewed and<br />

compared the wide variety of zero-or<strong>der</strong> CM’s. In higher or<strong>der</strong> methods, the DSD<br />

is defined in a specific functional form for every section of discretization, and these<br />

methods are usually more accurate but less robust [101–103]. The CM’s present the<br />

main disadvantage of requiring a large number of classes to work with good accuracy,


2.1. State of the Art 15<br />

and if the final application of the solution is implementation in a computational fluid<br />

dynamics (CFD) code, then the solution has to be done in every cell of computational<br />

domain, resulting in a very high computational time and memory problems.<br />

On grounds of CM approach to solve the Williams’ spray equation, Tambour [104]<br />

discretized the size axis of the droplet size distribution into sections which are also<br />

known as bins to <strong>der</strong>ive sectional equations. The droplet evaporation, collision and<br />

other physical processes are handled using the source terms among these bins. However,<br />

this method is found to be first or<strong>der</strong> accurate with respect to the droplet size [105],<br />

thus resulting in a strong numerical diffusion when few sections are used. The problem<br />

of finding an appropriate way to improve the accuracy of this method and also<br />

minimizing the number of sections becomes critical, especially when consi<strong>der</strong>ing industrial<br />

codes, which are intended to perform complex three-dimensional simulations.<br />

Later this approach is extended to higher o<strong>der</strong> by Dufour and Villedieu [55], but still<br />

requirement of high number of sections needs to be addressed.<br />

The other option is QMOM [51], which is based on the solution of moment transport<br />

equations of kinetic equation and the evaluation of unknown moments with the<br />

quadrature approximation. In this method, the distribution function is approximated<br />

with n-point Gaussian quadrature [51]. The moments are transported at every time<br />

and space step and the quadrature weights (number density), droplet radius, velocity<br />

and other phase-space variables, termed as abscissas, are computed using the<br />

product-difference (PD) algorithm of Gordon [106]. This method was first introduced<br />

by McGraw [51] in describing the aerosol dynamics to improve the method<br />

of moments (MOM) [107–113], and it is found to be a reliable method than MOM as<br />

the closure problem was observed with MOM [107, 114]. This method is proven to<br />

be promising in the problems of coagulation, aggregation and breakage [48–50], gasparticle<br />

flows [115]. One of the main limitations of QMOM is that since the dispersed<br />

phase is represented through the moments of the size distribution, the phase-average<br />

velocity of different phases must be used to solve the transport equations for the moments.<br />

Thus, in or<strong>der</strong> to use this method in the context of sprays for which the inertia<br />

determines the dynamic behavior of the droplets, it is necessary to extend QMOM<br />

to handle cases where each droplet size is convected by its own velocity [53]. The<br />

efficiency and applicability of such methods [60, 116, 117] for moment inversion in<br />

multi-variate poly-disperse systems have remained a question of interest [115], which<br />

are characteristic in many technical applications. In the work of Marchisio and Fox [71],<br />

a comprehensive review of the existing moment methods is given.<br />

Recently CQMOM was introduced by Yuan and Fox [118], to address the issues<br />

related to moment inversion. CQMOM is a novel moment-inversion algorithm, which<br />

works even for multi-variate moments. One apparent disadvantage of CQMOM (as


16 2. Mathematical Modeling<br />

well as other multi-variate moment-inversion algorithms [115, 119]), according to Yuan<br />

and Fox [118], is the existence of multiple permutations (i.e., the or<strong>der</strong> of conditioning).<br />

In the context of multi-dimensional quadrature, there are a number of interesting open<br />

questions concerning CQMOM. For example, in or<strong>der</strong> to have a realizable quadrature,<br />

the abscissas found from the conditional moments must lie in support of the distribution<br />

function. For one-dimensional distribution functions with compact supports, it is<br />

shown that the abscissas will always lie realizable [118, 120]. However, it still needs<br />

to be found un<strong>der</strong> which conditions boundedness will hold for the abscissas found<br />

from CQMOM for a two-dimensional (or higher) distribution function with compact<br />

support [118]. For certain applications (e.g. turbulent reacting flows), guaranteed<br />

boundedness is critical because the source terms are only defined on the support of the<br />

distribution function. As per revelations made by Yuan and Fox [118], the other open<br />

question is whether further improvements in the moment-inversion algorithm are possible<br />

to increase the number of optimal moments controlled, perhaps up to the maximum<br />

number of degrees of freedom determined by the number of quadrature nodes [118].<br />

In or<strong>der</strong> to overcome these problems, DQMOM has turned out to be an attractive<br />

alternative to QMOM, which was introduced by Fan et al. [121], later extended and<br />

validated by Marchisio and Fox [53]. This approach is found to be a powerful method<br />

in the multiphase flow problems to include all the physical processes of interest. The<br />

principal physical processes that the droplets encounter during the spray flows are (1)<br />

transport in real space or convection, (2) droplet evaporation and drying, (3) acceleration<br />

or deceleration of droplets due to forces induced by the surrounding gas, and<br />

(4) coalescence and collision of droplets leading to poly-dispersity. In DQMOM [53],<br />

the transport equations of weights and abscissas are solved directly rather than the<br />

moment equations, which is done in QMOM, thus avoiding the ”moments” to ”weights<br />

and abscissas” conversion through moment-inversion algorithm, so the word ”direct”<br />

implies. DQMOM also allows each droplet to convect with its own velocity.<br />

The DQMOM is proven to be a robust method in the field of multiphase flows and<br />

it is applied to various research problems other than spray flows after its development<br />

by Marchisio and Fox [53]. DQMOM is adapted and validated for the coagulation and<br />

sintering of particles by extending to bi-variate population balance equations [54]. Recently,<br />

DQMOM has been applied in studying exhaust particle formation and evolution<br />

in the wake of ground vehicle [122], and DQMOM is compared with classes method for<br />

simulations of bubbly flows [123]. In latest studies, DQMOM is employed in modeling<br />

poly-disperse fluidized pow<strong>der</strong>s [124], and modeling of turbulent combustion [125]. It<br />

is also used in combination with micro-mixing model and compared with stochastic<br />

field method for treating turbulent reactions [59].<br />

In the field of spray flows, DQMOM in combination with multi-fluid method is


2.2. Euler – Lagrangian Approach 17<br />

applied to study multi-component (fuel mixture) droplet evaporation [126]. Gumprich<br />

et al. [127, 128] analyzed the dense turbulent sprays using DQMOM, and DQMOM<br />

coupled with Eulerian multi-size moment model [129]. Madsen [61] extended DQMOM<br />

to include droplet coalescence in spray flows by neglecting the effects of evaporation,<br />

whereas Fox et al. [60] further improved DQMOM to model evaporating and coalescing<br />

spray flows but his study assumed simplified models for evaporation and coalescence.<br />

So far DQMOM has not been consi<strong>der</strong>ed to treat the process of spray drying.<br />

In the present study, DQMOM is used to describe the disperse phase consisting of<br />

poly-disperse liquid droplets, whereas the gas phase is not yet resolved but its inlet<br />

flow properties are taken for computing the droplet motion and evaporation. In this<br />

work, the DQMOM is implemented in two dimensions, which is done for the first time,<br />

and applied to study bi-component evaporating spray flows.<br />

2.2 Euler – Lagrangian Approach<br />

In the Euler – Lagrangian approach, the mean field equations are used only for the<br />

continuous gas phase. The droplet properties are defined along the path lines followed<br />

by the droplet. The trajectories of droplets are tracked for each droplet group by using<br />

a set of equations that describe their physical transport in flow field. In the current<br />

study, the discrete droplet model is used to define the droplet phase whereas the gas<br />

phase is modeled using the Navier – Stokes equations.<br />

2.2.1 Gas Flow<br />

The Euler – Lagrangian model DDM includes Euler equations for the gas phase with<br />

source terms for the dilute spray, which is described in Lagrangian coordinates. The<br />

instantaneous Navier – Stokes equations in an axisymmetric, two-dimensional configuration<br />

with no swirl for a dilute spray yield [68, 130]<br />

∂(ρu i )<br />

∂t<br />

+ ∂(ρu iu j )<br />

∂x j<br />

∂ρ<br />

∂t + ∂(ρu j)<br />

= S l,1 , (2.1)<br />

∂x j<br />

= − ∂p<br />

∂x i<br />

+ ∂τ ij<br />

∂x j<br />

+ ρg i + S l,ui , (2.2)<br />

where ρ, u i and p are the density, velocity component and pressure of the gas flow,<br />

respectively. g i is the acceleration due to gravity and the quantities S l,1 and S l,ui are<br />

the source terms due to spray evaporation [130, 131]. τ ij is the viscous stress tensor<br />

given by<br />

( ∂ui<br />

τ ij = µ + ∂u j<br />

− 2 ∂x j ∂x i 3<br />

)<br />

∂u k<br />

δ ij , (2.3)<br />

∂x k


18 2. Mathematical Modeling<br />

where δ is the tensorial Kronecker delta given by<br />

{<br />

1 : i = j<br />

δ ij =<br />

0 : i ≠ j.<br />

(2.4)<br />

Neglecting the processes of radiation, friction heating, Dufour effect, and the viscous<br />

heating, the conservation equation of total stagnant enthalpy can be written as<br />

∂(ρh)<br />

∂t<br />

+ ∂(ρu jh)<br />

∂x j<br />

= ∂p<br />

∂t − ∂J d q,j<br />

∂x j<br />

− ∂J c q,j<br />

∂x j<br />

+ S l,h , (2.5)<br />

where h is the enthalpy of the gas flow and the terms on the right hand side (R.H.S)<br />

are the change rate of the pressure, the heat diffusion term, the heat conduction term<br />

and the source term due to spray evaporation, S l,h , respectively. The heat conduction<br />

term is expressed by the Fourier’s Law<br />

(<br />

)<br />

Jq,j c = −λ ∂T = λ¯Cp ∂h ∑N s<br />

∂Y α<br />

− h α , (2.6)<br />

∂x j ∂x j ∂x j<br />

where λ, T , ¯Cp are thermal conductivity, gas temperature and specific heat capacity,<br />

respectively.<br />

N s refers to the number of chemical species while h α and Y α are the<br />

enthalpy and mass fraction of species α. The heat diffusion term J d q,j is written as<br />

α=1<br />

∑N s<br />

Jq,j d = h α Jα<br />

m<br />

α=1<br />

∑N s<br />

Y α<br />

= − ρh s,α D α,M , (2.7)<br />

∂x j<br />

α=1<br />

where h s,α and D α,M are the specific sensible enthalpy of species α and diffusion coefficient<br />

of species α, respectively. Assuming a unity Lewis number, which is defined as<br />

the ratio of thermal diffusion to mass diffusion, (Le = k/(ρC p D), and equal diffusibility<br />

of all species, the total heat flux is<br />

J q = J c q,j + J d q,j = − λ¯Cp (<br />

∂h<br />

∂x j<br />

−<br />

∂(ρh)<br />

∂t<br />

+ ∂(ρu jh)<br />

∂x j<br />

∑N s<br />

α=1<br />

)<br />

∂Y α<br />

h α −<br />

∂x j<br />

∑N s<br />

α=1<br />

The conservation equation of species mass can be written as<br />

∂(ρY α )<br />

∂t<br />

ρh s,α D α,M<br />

Y α<br />

∂x j<br />

. (2.8)<br />

= ∂p<br />

∂t + ∂ ( )<br />

∂h<br />

Γ h + S l,h . (2.9)<br />

∂x j ∂x j<br />

+ ∂(ρu jY α )<br />

− ∂ ( )<br />

∂Y α<br />

ρD α = S α + δ L,α S l,Yα , (2.10)<br />

∂x j ∂x j ∂x j<br />

where D α is the diffusion coefficient of species α while S α and S l,α are the source terms<br />

due to chemical reactions and spray evaporation, respectively. The mass fraction may<br />

be used to formulate mixture fraction.<br />

The advantage of an appropriately defined


2.2. Euler – Lagrangian Approach 19<br />

mixture fraction is that the source term S α will be zero. In the present work, for the<br />

water and polyvinylpyrrolidone (PVP)/water spray in air, the only possibility is to<br />

define the mixture fraction with reference to hydrogen as oxygen appears in both gas<br />

and liquid. A detailed study of different reference elements are given by Gutheil and<br />

Williams [132]. Thus the mass fraction Z A of element A, where A is either N or H or<br />

O, is defined as<br />

Z A =<br />

n∑<br />

i=1<br />

a IA M A<br />

M I<br />

Y I , (2.11)<br />

where a IA is the mass of element A in molecule I and M A and M I are the molecular<br />

weights of element A and element I, respectively. Using this definition, mixture fraction<br />

can be defined as<br />

ξ =<br />

Z A − Z A,min<br />

Z A,max − Z A,min<br />

. (2.12)<br />

Multiplying Eq. (2.10) by a IAM A<br />

a lA M I<br />

and summing over total number of species un<strong>der</strong> the<br />

assumption of equal diffusivity, the following conservation equation for mixture fraction<br />

is obtained<br />

∂(ρξ)<br />

∂t<br />

+ ∂(ρu iξ)<br />

∂x i<br />

= ∂<br />

∂x i<br />

(<br />

Γ M<br />

∂ξ<br />

∂x i<br />

)<br />

+ S l,ξ , (2.13)<br />

where Γ M = ρD M is the mass diffusion coefficient of the mixture.<br />

Equations (2.1), (2.2), (2.9) and (2.13) are the instantaneous conservation equations<br />

of mass, momentum, energy and mixture fraction. These equations need to be<br />

averaged for application to turbulent flows, and the general averaging types include<br />

time-averaging, ensemble-averaging and Favre- or density-weighted averaging [37]. For<br />

turbulent compressible flows, a density-weighted averaging i.e., Favre-averaging for<br />

Navier – Stokes equations is useful, and for more details about this approach, see [35,<br />

72, 76, 77].<br />

2.2.2 Discrete Droplet Model (DDM)<br />

The discrete droplet model (DDM) is a well established Euler – Lagrange approach<br />

for dilute sprays [72, 133, 134]. The droplet positions and velocities are captured<br />

using Lagrangian particle tracking method, thereby the source terms for the Eulerian<br />

equations of the gas phase are computed. The model captures the trajectories and<br />

dynamics of individual droplets, which are assumed to be parcels [35, 72, 76, 77]. A<br />

parcel refers to a collection of droplets, which are described by a set of properties, i.e.,<br />

(x p,k , r p,k , v p,k , m p,k , T p,k , ∆V ij ), where x p,k is the position, r p,k is the radius, v p,k is the<br />

velocity, m p,k is the liquid mass and T p,k is the temperature of k th parcel in control<br />

volume ∆V ij . By tracking the trajectories of a system of parcels, the model captures


20 2. Mathematical Modeling<br />

the flow properties, i.e., droplet dynamics, evaporation and heating, and these are<br />

described through sub-models. These sub-models for the different physical processes<br />

of interest are given in Subsection 2.4.<br />

The Lagrangian droplet equations are coupled to the gas phase through Eqs. (2.1) –<br />

(2.13), and the spray source terms are formulated using the particle-source-in-cell (PSIC)<br />

model [76, 130]. The system of gas and liquid phase equations is solved through a hybrid<br />

finite volume technique [130] with appropriate initial and boundary conditions. The<br />

DDM simulations were performed by Humza [68], so the details of initial and boundary<br />

conditions and numerical solution procedure of DDM are given in his work [68].<br />

2.3 Euler – Euler Approach<br />

In the Euler – Euler approach, the disperse phase is generally described by a number<br />

density function (NDF). In or<strong>der</strong> to un<strong>der</strong>stand the proposed modeling approach of<br />

DQMOM one needs to be familiar with the definition of the NDF, the NDF transport<br />

equation, and its un<strong>der</strong>lying terms, which are explained in the next subsections.<br />

2.3.1 Treatment of the Spray<br />

The disperse phase constitutes of discrete droplets and each of these droplets can be<br />

identified by a number of properties known as coordinates. In general, the coordinates<br />

are categorized as internal and external. The external coordinates are spatial coordinates.<br />

Internal coordinates refer to the properties of the droplets such as droplet<br />

velocity, mass, volume (or surface area, size), and enthalpy (or temperature). The<br />

NDF contains the information about the population of droplets inside a control volume<br />

[28, 41].<br />

Let us consi<strong>der</strong> a population of droplets in a spray dispersed in a control volume<br />

located at the physical point x = (x 1 , x 2 , x 3 ) where the size of the control volume is<br />

dx = dx 1 dx 2 dx 3 . (2.14)<br />

Let ξ = (ξ 1 , ξ 2 , ..., ξ N ) be the internal-coordinate vector. The NDF n ξ (ξ; x, t) is defined<br />

as the probable number of droplets in the physical volume dx and in phase-space<br />

volume dξ, given by n ξ dxdξ.<br />

The NDF is an average quantity of the dispersed phase, and it has mathematical<br />

characteristics of an averaged function, i.e., it is smooth and differentiable with respect<br />

to time t, physical space dx and phase-space dξ. The number density of droplets<br />

contained in the phase-space volume dξ per unit volume of physical space is n ξ dξ.<br />

By integrating the NDF over all the possible internal-coordinates, different average


2.3. Euler – Euler Approach 21<br />

quantities of interest can be obtained. For instance, the total number density per unit<br />

physical volume N(x, t), which is defined as the total probable number of droplets in<br />

a unit physical volume, and it is also known as zeroth moment, M ξ (0), obtained as<br />

∫<br />

N(x, t) = M ξ (0) = n ξ (ξ; x, t)dξ. (2.15)<br />

Similarly, the first moment defines the mean of the distribution, which is generally<br />

denoted by µ. The other moments which are common in probability and statistics are<br />

central moments [135], which are defined based on the distribution of droplets about<br />

the mean of the NDF. Thus the n th moment is given as<br />

∫<br />

M(n) = (ξ − ¯ξ) n n ξ (ξ; x, t)dξ, (2.16)<br />

where the second moment (n = 2) is known as variance, generally denoted by σ 2 , and<br />

the third and fourth moments define the skewness and kurtosis of the distribution [135],<br />

respectively. Typical log-normal distribution with different values of mean, µ, and<br />

variance, σ 2 , are shown in Fig. 2.2.<br />

Likewise, when the NDF is defined with respect to more than one internal coordinate,<br />

the moments of NDF are known as multi-variate moments, which can be<br />

computed as<br />

∫<br />

k<br />

M ξ (k) = ξ 1 k<br />

1 ξ 2 k<br />

2 ....ξ N N n ξ (ξ)dξ, (2.17)<br />

where k = (k 1 , k 2 , ...k N ) is a vector containing the or<strong>der</strong> of the moments with respect to<br />

each component of ξ. In the description of spray flows, the NDF is in general described<br />

based on the droplet radius and velocities as the internal coordinates, ξ.<br />

1<br />

0.8<br />

µ = 0, σ = 0.5<br />

µ = 0, σ = 1.0<br />

µ = 1, σ = 0.5<br />

0.6<br />

n(ξ)<br />

0.4<br />

0.2<br />

0<br />

0 1 2 3 4 5 6<br />

ξ<br />

Fig. 2.2: A typical log-normal distribution with different values of µ and σ.


22 2. Mathematical Modeling<br />

For the purpose of un<strong>der</strong>standing the average properties of the spray dynamics<br />

from the simulations, the calculation procedure of mean droplet diameter and Sauter<br />

mean diameter is defined below. The consi<strong>der</strong>ed NDF is defined based on the droplet<br />

diameter d. The mean droplet diameter, denoted as d 1,0 or d 10 , can be computed from<br />

the droplet diameter based NDF n d (d)as<br />

d 1,0 = 1 N<br />

∫ ∞<br />

0<br />

dn d (d)dd, (2.18)<br />

where N is the total number density given by<br />

N =<br />

∫ ∞<br />

0<br />

n d (d)dd. (2.19)<br />

Similarly, the Sauter mean diameter, d 3,2 , or simply d 32 , which is a very frequently<br />

used characteristic parameter especially in spray flows and spray drying, is given by<br />

∫ ∞<br />

d 3 n<br />

0 d (d)dd<br />

d 3,2 = ∫ ∞<br />

d<br />

0 2 n d (d)dd . (2.20)<br />

Any other average droplet diameter can be extracted by simply dividing the k + 1 th<br />

moment with k th moment, i.e.,<br />

d k+1,k =<br />

∫ ∞<br />

d k+1 n<br />

0 d (d)dd<br />

∫ ∞<br />

d<br />

0 k n d (d)dd = M k+1<br />

. (2.21)<br />

M k<br />

Here, M k+1 and M k are the moments of droplet diameter based NDF. The equation of<br />

averaged droplet diameter changes with the definition of NDF, i.e., averaged droplet<br />

diameter is different in diameter based NDF from that of volume based NDF, and the<br />

relation between volume based NDF and diameter based NDF is given by<br />

n V (V ; x, t) = k V d 3 n d (d; x, t), (2.22)<br />

where k V<br />

is the volume shape factor.<br />

2.3.2 NDF Transport Equation<br />

The evolution of the NDF due to physical processes is, in general, written in terms of<br />

a transport equation known as population balance equation. This transport equation<br />

is a simple continuity equation written in terms of the NDF, and it can be <strong>der</strong>ived<br />

based on the balance for droplets/dispersed entities in a fixed subregion of internal<br />

coordinates and physical space [41]. This type of equation is known by different names<br />

in different fields. In aerosol dynamics, it is known as particle-dynamics equation,<br />

and in evaporating spray flows it is known as Williams-Boltzmann equation or simply,


2.3. Euler – Euler Approach 23<br />

Williams’ spray equation, which was first introduced by Williams [42]. The <strong>der</strong>ivation<br />

of such an equation is given by Archambault [67] and Ramakrishna [41].<br />

In modeling the spray flows, the principal physical processes that must be accounted<br />

for are transport or convection, droplet evaporation, forces experienced by the droplets,<br />

and droplet–droplet interactions leading to poly-dispersity. Williams’ spray equation<br />

either accounts for these physical processes or it can be easily extended to include,<br />

and it has proven to be a useful starting point for testing novel methods for describing<br />

poly-disperse dense liquid sprays. The Williams’ spray equation [42], is given by<br />

∂f<br />

∂t + ∂(vf)<br />

∂x<br />

(Rf)<br />

= −∂ − ∂(Ff)<br />

∂r ∂v + Q f + Γ f . (2.23)<br />

Equation (2.23) describes the transport of the number density function f(r, v; x, t)<br />

in terms of time, t, and Euclidean space, x. In Eq. (2.23), first term at the left hand<br />

side (L.H.S) accounts for changes in the NDF with time and the second term includes<br />

the convective changes, v and F denote droplet velocity and total forces acting on the<br />

droplet per unit mass, respectively. The first term in the R.H.S includes the effect of<br />

evaporation, where R is the change in the droplet radius with time, i.e., R = dr/dt, and<br />

r is the droplet radius. The last two terms in Eq. (2.23) refer to the droplet–droplet<br />

interactions; Q f represents the increase in f with time due to droplet formation or<br />

destruction by processes such as nucleation or breakup, and Γ f denotes the rate of<br />

change in f due to droplet collisions.<br />

The next subsections present the governing equations of the liquid phase defined<br />

through QMOM and DQMOM, which are used to solve Williams’ spray equation.<br />

2.3.3 Quadrature Method of Moments (QMOM)<br />

Moment methods are an important class of approximate methods <strong>der</strong>ived to solve<br />

kinetic equations, but require closure to truncate the moment set. In QMOM, closure<br />

is achieved by inverting a finite set of moments to reconstruct a point distribution<br />

from which all unclosed moments (e.g.<br />

spatial fluxes) can be related to the finite<br />

moment set [28, 51]. Figure 2.3 shows the typical quadrature approximation of the<br />

NDF in QMOM. The <strong>der</strong>ivation of QMOM starts with the moment transformation<br />

of Williams’ spray equation, which is done in the current study with multi-variate<br />

moments of droplet radius and velocity denoted by M(k 1 , k 2 ; x, t) where k 1 and k 2 are<br />

the moment or<strong>der</strong>s with respect to droplet radius and droplet velocity, respectively .<br />

The moment transformed Williams spray equation is given as [28],<br />

∂M(k 1 , k 2 )<br />

∂t<br />

+ ∂M(k 1, k 2 + 1)<br />

∂x<br />

=<br />

∫ ∞ ∫ ∞<br />

−∞<br />

0<br />

r k 1<br />

v k 2<br />

[<br />

− ∂ (Rf) − ∂(Ff) ]<br />

∂r ∂v + Q f + Γ f drdv.<br />

(2.24)


24 2. Mathematical Modeling<br />

0.4<br />

Number density [(µm) -3 ]<br />

0.2<br />

-5 0 5<br />

Droplet diameter [µm]<br />

Fig. 2.3: Approximation of NDF in QMOM.<br />

In the evolution of every moment M(k 1 , k 2 ) one higher or<strong>der</strong> moment, i.e., M(k 1 , k 2 +1)<br />

appears, see Eq. (2.24), which can be closed by using the product-difference algorithm<br />

(PD) [106]. The PD algorithm is quite efficient in a number of practical cases;<br />

however, it generally becomes less stable as the number of nodes, N, increases. It is<br />

difficult to predict a priori when this will occur, since it depends on the absolute values<br />

of the moments, but typically problems can be expected when N > 10 [71]. Another<br />

issue with PD algorithm is if the distributions with zero mean present, which can occur<br />

when the internal coordinate droplet velocity is ranging between positive and negative<br />

values, in this case the algorithm blows up due to the division by zero in the calculation<br />

of the coefficient matrix of the PD algorithm [71]. The alternative approach to the PD<br />

algorithm, which can handle cases with zero mean and remains stable even in the cases<br />

of N > 10, is the Wheeler algorithm proposed by Sack and Donovan [136]. The step<br />

by step procedure of implementing PD and Wheeler algorithms with corrections for<br />

moment realizability are given by Marchisio and Fox [71] with example calculations.<br />

The substitution of different sub-models for the droplet evaporation, total force<br />

acting on droplets, droplet breakup and collision in the R.H.S of the Eq. (2.24) and with<br />

the help of PD or Wheeler algorithm for the unknown moment terms, yields a closed<br />

transported moment equation, which can then be solved to find the change in moments<br />

with time and spatial location. As discussed in the literature review, see Section 2.1,<br />

this method lacks the ability to handle multi-variate moments as the moment-inversion<br />

algorithm gets cumbersome. The different sub-models to describe the various physical<br />

processes of interest are described in Subsection 2.4. The numerical solution for the<br />

QMOM moment transport equations with initial and boundary conditions, and closure


2.3. Euler – Euler Approach 25<br />

of the unknown moments applying the Wheeler algorithm are explained in Chapter 3.<br />

2.3.4 Direct Quadrature Method of Moments (DQMOM)<br />

In DQMOM, the NDF is approximated as sum of the Dirac-delta functions. Substitution<br />

of this assumed NDF in Williams’ spray equation yields transport equations in<br />

terms of the phase-space [60]. For the present study, a joint droplet radius-velocity<br />

number density function is consi<strong>der</strong>ed, which is approximated in DQMOM as a sum<br />

of the product of weighted Dirac-delta functions [53] of radii and velocities [60],<br />

f(r, v) =<br />

N∑<br />

w n δ(r − r n )δ(v − v n ), (2.25)<br />

n=1<br />

where w n and r n are chosen as N representative quantities of weights and radii, and v n<br />

are the corresponding velocities. Such an approximation with a three-node (N = 3) closure<br />

can be depicted as shown in Fig. 2.4. Application of DQMOM to Williams’ spray<br />

equation results in closed transport equations in terms of droplet weights or number<br />

densities, radii and velocities, which are written as<br />

∂w n<br />

∂t<br />

∂(w n ρ l r n )<br />

∂t<br />

+ ∂(w nv n )<br />

∂x<br />

+ ∂(w nρ l r n v n )<br />

∂x<br />

= a n , (2.26)<br />

= ρ l b n , (2.27)<br />

and<br />

∂(w n ρ l r n v n )<br />

+ ∂(w nρ l r n v n v n )<br />

= ρ l c n , (2.28)<br />

∂t<br />

∂x<br />

where a n , b n and c n are the source terms that account for droplet evaporation, forces<br />

on droplet (drag, buoyancy, lift, basset and virtual mass effect and gravity etc.), coalescence<br />

and breakup. These Eqs. (2.26) – (2.28) form a set of coupled hyperbolic<br />

partial differential equations, which can be solved simultaneously by using appropriate<br />

initial and boundary conditions to find w n (x, t), r n (x, t) and v n (x, t), and thereby the<br />

evolution of droplet distribution function f can be computed.<br />

Fig. 2.4: NDF approximation in DQMOM.


26 2. Mathematical Modeling<br />

The Eqs. (2.26) – (2.28) are closed by modeling the source terms, i.e., a n , b n and c n ,<br />

using the physical models to account for effects of droplet evaporation, forces on droplet,<br />

coalescence and breakup.<br />

These source terms are calculated through the moment<br />

transformation of phase-space terms, which yields the following linear system<br />

∫ [<br />

P k,l = r k v l − ∂(Rf) − ∂(Ff) ]<br />

∂r ∂v + Γ f + Q f drdv. (2.29)<br />

The exact form of the DQMOM linear system relies on the choice of moments, and it<br />

can be generated from<br />

∫<br />

[ ∂f<br />

r k v l ∂t + ∂(vf)<br />

∂x<br />

]<br />

drdv =<br />

+<br />

+<br />

N∑<br />

n=1<br />

N∑<br />

n=1<br />

N∑<br />

n=1<br />

(1 − k)r k nv l 1<br />

1,n v l 2<br />

2,n v l 3<br />

3,n a n<br />

(k − l 1 − l 2 − l 3 )r k−1<br />

n v l 1<br />

1,n v l 2<br />

2,n v l 3<br />

3,n b n<br />

r k nv l 1<br />

1,n v l 2<br />

2,n v l 3<br />

3,n (l 1 v −1<br />

1,nc 1,n + l 2 v −1<br />

2,nc 2,n + l 3 v −1<br />

3,nc 3,n )<br />

+ δ k0 u l 1<br />

1 u l 3<br />

2 u l 3<br />

3 ψ, (2.30)<br />

where ψ is the evaporative flux, and u 1 , u 2 , and u 3 are three components of the gas<br />

velocity. The complete linear system is formed by combining Eqs. (2.29) and (2.30),<br />

which consists of 5N +1 unknowns, a n , b n , c n and ψ. To obtain a solution for this linear<br />

system, the moments are chosen in a way that the resulting coefficient matrix is nonsingular.<br />

Previous validation studies of DQMOM, and comparison of its performance<br />

with QMOM have demonstrated that by using two-node closure (N = 2) approximation<br />

for f is sufficient to track the lower or<strong>der</strong> moments with small errors [48, 49, 54].<br />

Increasing the number of nodes, N, to three (N = 3) have improved the results, and<br />

in general the evaporation and coalescence terms can be accurately approximated with<br />

N = 2–4 [48, 49, 54, 60]. In the present work, a three-node closure is used, i.e., N is<br />

set to be 3, and the corresponding moment set is chosen as [60, 137] k ∈ {1, ..., 2N};<br />

l ∈ {0, 1}, where l is composed of three components l 1 , l 2 , and l 3 . The chosen set of<br />

k and l values conserves the mass and momentum of droplets, and these values are<br />

found to give non-singular source terms matrix [60]. Along with these moments set,<br />

the calculation of the source terms from the linear system requires the mathematical<br />

formulation for the evaporation, forces on droplet and droplet–droplet interactions,<br />

which enter as sub-models and these sub-models.<br />

discussed and mathematical formulation is given in Subsection 2.4.<br />

The sub-models are individually<br />

The evaporation term in Eq. (2.29) can be simplified by evaluating the integral on<br />

the R.H.S, which is given as<br />

∫<br />

− r k v l ∂(Rf)<br />

∂r<br />

= −(r k v l Rf)| r=∞<br />

r=0 + k<br />

∫ ∞<br />

0<br />

r k−1 v l ∂ (Rf) dr. (2.31)<br />

∂r


2.3. Euler – Euler Approach 27<br />

With the assumption that the maximum droplet size is a finite value, the above equation<br />

can be further simplified as<br />

∫<br />

− r k v l ∂(Rf) = δ k0 ψv l + k<br />

∂r<br />

∫ ∞<br />

0<br />

r k−1 v l ∂ (Rf) dr, (2.32)<br />

∂r<br />

where δ k0 is the Kronecker delta, which is defined as δ k0 = 1 if k = 0 and δ k0 = 0 for<br />

any other k value. The quantity ψ = Rf(0) is the evaporative flux, which is a point<br />

wise quantity of the NDF representing the number of droplets having zero size. This<br />

quantity in DQMOM is computed by weight ratio constraints, which are introduced by<br />

Fox et al. [60] where ψ is treated as an additional variable along with a n , b n and c n ’s.<br />

These ratio constraints of weights, radii and velocities [60] are given by<br />

( ) ( )<br />

D wn D rn<br />

= 0;<br />

= 0; (2.33)<br />

Dt w n+1 Dt r n+1<br />

where j is the index for three velocity components.<br />

( )<br />

D vj,n<br />

= 0, (2.34)<br />

Dt v j,n+1<br />

Fox et al. [60] show that the<br />

estimation of evaporative flux via weight ratio constraints is found to give acceptable<br />

results in a stationary one-dimensional configuration. However, Fox et al. [60] suggests<br />

that this calculation procedure is found to pose problems in the case of complicated<br />

distribution functions [64].<br />

In the current study, this is addressed by implementing the maximum entropy (ME)<br />

principle proposed by Mead and Papanicolaou [66] for water and PVP/water spray<br />

flow in air, which estimates the evaporative flux through reconstruction of the droplet<br />

distribution using its moments.<br />

The principle of maximum entropy in the problem of moments is that the distributions<br />

that satisfy the given moment set (also called as constraints), the most likely<br />

or least biased probability density function is the the one whose statistical entropy is<br />

a maximum. This formulation allows the determination of a number density function<br />

from the limited amount of information such as few known moments of a distribution<br />

[66]. The implementation of this method to compute ψ is explained by Massot et<br />

al. [98].<br />

The ME method is first introduced by Mead and Papanicolaou [66] to compute<br />

a distribution for the given moment set based on the maximization of the following<br />

Shannon entropy from the information theory [66],<br />

H[f] ≡ −<br />

∫ rmax<br />

r min<br />

f(x) ln f(x)dx. (2.35)<br />

Mead and Papanicolaou have proven that there exists ME distribution satisfying the<br />

above entropy principle [66] for the case when the vector of moments M belongs to


28 2. Mathematical Modeling<br />

the interior of the moment space M = {M(0), M(1), ..M(N)}. This is a standard constrained<br />

optimization problem where the constraints are to satisfy the given moments.<br />

In the ME method, following the moments satisfaction condition, below equation is<br />

the explicit representation of the ME approximation,<br />

f ME<br />

M (x) ≡ exp ( −Σ N j=0 ξ j x j) , (2.36)<br />

where the coefficients ξ 0 , ξ 1 ...ξ N are the Lagrange multipliers, and N is the number of<br />

moments. These coefficients are computed based on the condition of minimizing the<br />

following convex potential:<br />

∆ ≡<br />

∫ rmax<br />

r min<br />

[<br />

exp<br />

(<br />

−Σ<br />

N<br />

j=0 ξ j x j) − 1 ] dx + Σ N j=0 ξ j M(j). (2.37)<br />

The stationary points of Eq. (2.37) are given by ∂∆<br />

∂ξ i<br />

equation<br />

∫ rmax<br />

≡ 0, which yields the following<br />

r min<br />

x i exp ( −Σ N j=0 ξ j x j) dx ≡ M(i). (2.38)<br />

The solution of the above equation gives ξ i , and substitution of these ξ i in Eq. (2.36)<br />

yields the required NDF. The above equation can be solved numerically using a Newton<br />

method, with the initial guess as ξ ≡ (− ln M(0)/(r min − r max )), 0, ...0), and updated<br />

ξ’s are estimated by<br />

Here H is the Hessian matrix defined by H i,j ≡<br />

ξ + ≡ ξ − H −1 (M − 〈X〉 ξ<br />

). (2.39)<br />

∂∆<br />

∂ξ i ∂ξ j<br />

≡ 〈x i+j 〉 for i, j ≡ 0, 1, ...N, and<br />

〈X〉 ξ<br />

≡ (〈x 0 〉 ξ<br />

, ... 〈 x N〉 ) is the vector of approximated moments, which are expressed<br />

ξ<br />

as<br />

〈 〉 ∫ rmax<br />

x<br />

k ≡ x i exp ( −Σ N ξ j=0 ξ j x j) dx. (2.40)<br />

r min<br />

The numerical procedure to implement this approach is same as done by Mead and<br />

Papanicolaou [66] and Massot et al. [98], where a double-precision 24-point Gaussian<br />

quadrature method very efficiently produces the required accuracy for 〈 x k〉 ξ . More<br />

details about the <strong>der</strong>ivation of this method and numerical solution procedure are given<br />

by Mead and Papanicolaou [66].<br />

As the systems of interest in the present study are water spray in quiescent air or<br />

nitrogen as well as PVP/water spray in quiescent air, currently the gas phase is not<br />

fully coupled with DQMOM transport equations but its inlet properties taken from<br />

the experiment are used to compute the droplet motion and evaporation.<br />

2.4 Single Droplet Modeling<br />

This section presents the physical processes due to gas–liquid and droplet–droplet interactions,<br />

namely, droplet heating and evaporation, forces acting on the droplet, droplet


2.4. Single Droplet Modeling 29<br />

collisions and breakup, which enter as source terms in QMOM, DQMOM and DDM.<br />

2.4.1 Droplet Heating and Evaporation<br />

In spray flows, particularly in spray drying processes, the droplet evaporation can be<br />

critical because (1) it has direct effect on drying rate of droplets yielding pow<strong>der</strong>, and<br />

(2) it influences the final pow<strong>der</strong> characteristics. The evaporation process can be very<br />

complex un<strong>der</strong> realistic spray drying conditions. Factors that increase the complexity<br />

of the evaporation models are (1) the multi-component character of the liquid solution,<br />

(2) the interaction between droplets in the turbulent gas environment, and (3) large<br />

differences in volatility of solutes. The study of single droplet heating and evaporation<br />

forms a basis for simulating complex spray flows. As stated before, few studies have<br />

been carried out for application of DQMOM on evaporating sprays [60, 126, 138]. However,<br />

these studies consi<strong>der</strong> a simplified evaporation model to calculate the change in<br />

droplet size with time, i.e., either as a linear function of droplet volume or non-linear<br />

function of droplet volume, which is similar to the well established d 2 law. In the<br />

present study, an advanced droplet evaporation model of Abramzon and Sirignano [62]<br />

is used for the single component droplet evaporation, whereas, for the bi-component<br />

PVP/water droplet, the focus is to develop a mathematical model, which can predict<br />

the evaporation and drying of a single bi-component droplet, thereby include the developed<br />

model to study the PVP/water and mannitol/water droplet evaporation and<br />

solid layer formation. The following subsections present the mathematical models for<br />

mono- and bi-component droplet evaporation.<br />

2.4.1.1 Single Component Droplet<br />

The droplet evaporation is a complex process where simultaneous heat and mass transfer<br />

occurs leading to regressing droplet size. Fluid dynamics plays a major role when<br />

there is a relative motion between the droplets and the surrounding gas. The flow properties<br />

have a critical impact on the mass, momentum and energy exchanges between the<br />

gas and the droplets. Droplet evaporation was first studied by Langmuir [139] in 1918.<br />

Earlier studies reported that the droplet surface decreases constantly with time [140],<br />

famously known as d 2 law. After Langmuir [139], several studies were carried out in<br />

this area. Most notable works in droplet evaporation descriptions include the studies<br />

of Chigier [141], Clift et al. [142], Glassman [143], Lefebvre [144], and Williams [145].<br />

A review of existing droplet evaporation models is given by Faeth [72], Law [146] and<br />

Sirignano [140].<br />

The study of Abramzon and Sirignano [62] introduced a model for single component<br />

droplet evaporation, which includes the convection effects, droplet heating, and variable


30 2. Mathematical Modeling<br />

liquid and film properties.<br />

The work of Sirignano [140] classified the single component droplet evaporation<br />

models into six types, and they are given in the or<strong>der</strong> of complexity as, (1) constant<br />

droplet temperature model (also known as the d 2 law), (2) infinite liquid conductivity<br />

model (uniform but time dependent droplet temperature), (3) conduction limit (spherically<br />

symmetric transient droplet heating) model, (4) effective conductivity model, (5)<br />

vortex model of droplet heating, and (6) Navier – Stokes solution. There are various<br />

differences among these models, and some of these models are shown to be limits of<br />

another model [140].<br />

Recent study of Sazhin [147] gives an overview of all the existing droplet evaporation<br />

models, particularly in the field of combustion studies.<br />

For evaporating water spray flow in air, the spatial gradients of the temperature<br />

within the droplet will not be significant when the evaporation conditions are room<br />

temperature and atmospheric pressure. Thus, in the current study, uniform but time<br />

dependent droplet temperature with convective effects can be used to predict the evaporation<br />

rate, and the droplet size regression. Therefore, for evaporating water sprays<br />

un<strong>der</strong> room temperature and pressure conditions, the model of Abramzon and Sirignano<br />

[62] is implemented, which is a uniform temperature model that includes the convective<br />

effects, and consi<strong>der</strong>s the variable liquid and film properties. Here film means<br />

a thin layer across the droplet surface where the saturation of liquid vapor exists, and<br />

this vapor mass fraction is computed based on the vapor-liquid equilibrium.<br />

The rate of change of droplet mass with time due to convective evaporation and<br />

droplet heating in water spray is computed as [62]<br />

ṁ = 2πRρ f D f ˜Sh ln(1 + BM ), (2.41)<br />

where R is the droplet radius, ρ f is the density in the film, D f is the water diffusivity in<br />

the film, ˜Sh is the modified Sherwood number that accounts for the convective effects<br />

of droplet evaporation [62], given as<br />

˜Sh = 2 + Sh − 2<br />

B M<br />

(1 + B M ) ln(1 + B M ). (2.42)<br />

Here, Sh is the Sherwood number, which is defined as the ratio of convective mass<br />

transfer to the diffusion mass transport, and it is generally written in terms of the<br />

droplet Reynolds number, Re d , and Schmidt number, Sc, given by [62]<br />

Sh = 1 + (1 + Re d Sc) 1/3 f(Re d ). (2.43)<br />

The droplet Reynolds number is defined as the ratio of inertial forces to viscous forces,<br />

which is written as Re d = 2rρ g |u − v|/µ f . The Schmidt number is used to characterize


2.4. Single Droplet Modeling 31<br />

the fluid flows in which there is simultaneous momentum and mass diffusion, and<br />

it is defined as the ratio between momentum diffusion and mass diffusion, written<br />

as Sc = µ f /(ρ g D f ).<br />

In Eq. (2.43), the function f(Re d ) depends upon the droplet<br />

Reynolds number, and in case of low Reynolds number, it may be calculated as defined<br />

by Abramzon and Sirignano [62], with f(Re d ) = 1 for Re d ≤ 1 and f(Re d ) = Re d<br />

0.077<br />

for Re d ≤ 400.<br />

In Eq. (2.41), B M is the Spalding mass transfer number, expressed in terms of the<br />

mass fraction of vaporized liquid as,<br />

B M = Y s − Y ∞<br />

1 − Y s<br />

. (2.44)<br />

Here Y s and Y ∞ are mass fractions of the water at the droplet surface and in the bulk<br />

of surrounding gas, respectively.<br />

Y s is computed from the vapor-liquid equilibrium<br />

through the vapor pressure of water, which is written as [148]<br />

Y s =<br />

M w<br />

M w + ¯M(¯p/p w − 1) . (2.45)<br />

The quantities M w and p w denote molar mass and vapor pressure of water while<br />

and ¯p represent molar mass and mean pressure of the surrounding gas, respectively.<br />

Although the initial temperatures of gas and the droplet are equal and are at<br />

room temperature, the droplet temperature is subject to change due to evaporation.<br />

Time evolution of droplet temperature for water spray is computed using the uniform<br />

temperature model [62],<br />

[ ]<br />

dT s<br />

mC pL<br />

dt = Q CpLf (T ∞ − T s )<br />

L = ṁ<br />

− L V (T s ) , (2.46)<br />

B T<br />

where m is the droplet mass, Q L is the net heat transferred to the droplet per unit<br />

time, C pL and C pLf are the specific heat capacity of the liquid and in film, respectively,<br />

T s is the temperature at droplet surface, T ∞ is the temperature of the surrounding<br />

gas, and L V (T s ) is the temperature dependent latent heat of vaporization at T s . B T is<br />

the Spalding heat transfer number, which is calculated in terms of the mass transfer<br />

number using the relation [62]<br />

where the exponent φ is given by [62]<br />

B T = (1 + B M ) φ − 1, (2.47)<br />

φ = C pL<br />

˜Sh 1<br />

C pg Ñu Le . (2.48)<br />

Here C pg is the specific heat capacity of the gas, Le is the Lewis number, and<br />

¯M<br />

Ñu is<br />

the modified Nusselt number, which accounts for convective droplet heating, and it is


32 2. Mathematical Modeling<br />

given by [62]<br />

Ñu = 2 +<br />

Nu<br />

. (2.49)<br />

(1 + B T )<br />

−0.7<br />

The Nusselt number, Nu, defined as the ratio between convective heat transfer to<br />

conductive heat transfer and it is usually expressed in terms of the droplet Reynolds<br />

number, Re d , and Prandtl number, Pr, as<br />

Nu = 1 + (1 + Re d Pr) 1/3 f(Re d ). (2.50)<br />

The Prandtl number, Pr, is defined as the ratio of momentum diffusion rate to the<br />

thermal diffusion rate, written as, Pr = C pLf µ f /k f .<br />

2.4.1.2 Bi-component Droplet<br />

Many studies present the evaporation phenomena associated with pure and multicomponent<br />

droplet, but there is a lack of a mathematical model, which can predict the<br />

evaporation and drying behavior of a droplet containing a polymer or sugar dissolved<br />

in water because of the unknown physical behavior, unavailability of experimental<br />

results and complexity of the problem. The available literature in the area of single<br />

bi-component droplet evaporation and drying is reviewed in the following paragraphs<br />

followed by the development of new mathematical model to compute the evaporation<br />

and solid layer formation of a bi-component droplet.<br />

Charlesworth and Marshall [149] first investigated the process of single droplet<br />

evaporation and drying by measuring the change in droplet mass using the deflection<br />

of a thin, long glass filament. This study [149] also classifies different stages of droplet<br />

evaporation. Later, this experiment with some modifications is consi<strong>der</strong>ed in many<br />

studies. The work of Sano and Keey [150] includes the drying behavior of colloidal<br />

material into a hollow sphere by consi<strong>der</strong>ing the migration of solid matter towards the<br />

center of the droplet through the convection measurement inside the droplet, which is<br />

a challenge to experiment [151].<br />

Most of the experiments concerning the droplet evaporation and drying available<br />

in literature are either related to salts [149, 152–154], milk pow<strong>der</strong>s [155, 156] or some<br />

other colloidal matter [150, 151, 157–159], but none deals with droplets of polymer<br />

or mannitol as a constituent. Previously developed models assume a uniform temperature<br />

gradient within the droplet [151, 152, 157], and neglect the effect of solid<br />

formation [152, 156, 157]. The study of Nesic and Vodnik [151] presents the kinetics<br />

of droplet evaporation to predict the drying characteristics of a colloidal silica droplet,<br />

where the crust formation on the surface occurring in this configuration is consi<strong>der</strong>ed.<br />

The surface vapor concentration of the evaporating solvent is calculated using experimental<br />

material dependent factors, which are not available for every solution including


2.4. Single Droplet Modeling 33<br />

polymer and mannitol solutions in water. Moreover, in case of a polymer, molecular<br />

entanglement leads to solid layer formation. Nesic and Vodnik [151] use a more detailed<br />

description of various stages of droplet evaporation and drying. These stages, as<br />

described by Nesic and Vodnik [151], are that the droplet temperature initially rises<br />

to an equilibrium value and solvent evaporates continuously, which in turn increases<br />

the solute mass fraction within the droplet. When the solute mass fraction at the<br />

droplet surface rises to a critical value, then there starts a thin solid layer formation,<br />

and further drying leads to a dried particle.<br />

Farid [156] shows that the droplet evaporation and drying are controlled by thermaldiffusion<br />

rather than mass-diffusion as assumed by most of the earlier studies [149, 150,<br />

152]. In Farid’s model [156], the time taken for the formation of crust on a colloidal<br />

silica droplet is calculated using the energy balance, which does not account for solvent<br />

and solute composition changes, and the evaporation rate is computed using a<br />

simple relation without accounting for the variation in film and liquid properties. For<br />

droplets with suspended solids inside, the population balance approach is recently developed<br />

[158] to model the nucleation and growth of suspended solids inside an ideal<br />

binary liquid droplet with an assumption that there exist some nuclei of suspended<br />

solids initially. But this method cannot be applied in the present case of droplet<br />

with polymer or sugar, as solute is completely dissolved in water. Golman et al. [160]<br />

presents a model for the evaporation and drying of slurry droplets, which is an improvement<br />

over the receding interface model of Cheong et al [161] for slurry droplets, and<br />

the bi-component liquid mixture is treated as ideal. A detailed review of all existing<br />

theoretical models of evaporation and drying of single droplet containing dissolved and<br />

insoluble solids is given by Mezhericher et al. [162], and a review of evaporation models<br />

in the area of combustion is given by Sazhin [147].<br />

Adhikari et al. [163] and Vehring et al. [164] give a review of the experimental<br />

studies in the area of single droplet evaporation and drying. Tsapis et al. [159] and<br />

Sugiyama et al. [165] have levitated droplets using Leidenfrost phenomenon on a concave<br />

hot plate whereas Yarin et al. [166] levitated droplets using an acoustic levitator.<br />

This technique was successfully used to study shell buckling during particle formation<br />

[165]. A drawback of this approach is that the flow field and temperature field<br />

in the vicinity of the droplet are different from those of a free flowing droplet in a<br />

spray dryer. A chain of mono-disperse free falling droplets has been used by several<br />

experimental groups to study heat and mass transfer, drying, and particle formation<br />

processes. El Golli et al. [154] measured salt droplet evaporation and compared their<br />

results with a theoretical model. A similar technique to study the effect of drying<br />

rates on particle formation was used by Alexan<strong>der</strong> and King [167] and El-Sayed et<br />

al. [168]. Wallack et al. [169] compared measured evaporation rates with a numerical


34 2. Mathematical Modeling<br />

Fig. 2.5: Schematic diagram of stages in single droplet evaporation and drying.<br />

model, achieving fairly good agreement. The droplet generators used in these studies<br />

produce a chain of closely spaced droplets, which leads to droplet–droplet interactions<br />

in processes that are limited by gas phase transport processes.<br />

The aim of the present work is to develop a mathematical model, which can be<br />

applied to predict the evaporation and drying characteristics of droplets of the polymer<br />

PVP dissolved in water and mannitol dissolved in water solution. Prerequisites<br />

of the method are, (1) accounting for the solid layer resistance in mass evaporation<br />

rate and energy calculation, and (2) treatment of the liquid mixture as non-ideal by<br />

computing the activity coefficient of the evaporating component.<br />

The problem un<strong>der</strong> consi<strong>der</strong>ation is the evaporation and drying of an isolated single<br />

spherical droplet consisting of a binary mixture of a liquid and a dissolved solid material<br />

with low or zero vapor pressure.<br />

During the evaporation and drying of the bi-component droplet, the droplet un<strong>der</strong>goes<br />

four stages as explained by Nesic and Vodnik [151], which are depicted in Fig. 2.5.<br />

In the initial stage, the droplet temperature quickly rises to an equilibrium temperature,<br />

which is most often near to the wet bulb temperature for surrounding gas and<br />

humidity, with some solvent evaporation.<br />

In the second stage, the droplet starts to shrink as solvent evaporates causing the<br />

solute mass fraction to increase at the droplet surface; this leads to a slight raise in<br />

the droplet temperature (see Fig. 2.5). The increase in solute mass fraction at the<br />

droplet surface hin<strong>der</strong>s further evaporation as the vapor pressure of the solvent at<br />

the surface drops. The third stage of drying starts when the solute mass fraction at<br />

the surface raises to a threshold value, which most often is equal to the saturation<br />

solubility of the solute in the solvent, whereupon the crust formation starts for salts,<br />

sugar and colloidal material. In the case of polymers, molecular entanglement and


2.4. Single Droplet Modeling 35<br />

gradual increase in concentration lead to solid layer formation at the droplet surface.<br />

In the latter case, the solid layer thickens and develops into the droplet interior as<br />

shown in Fig. 2.5, and a rapid fall in evaporation rate is observed.<br />

In this period,<br />

the heat penetrated into the liquid is used for heating the droplet, which causes the<br />

droplet temperature to rise rapidly. Further drying behavior of droplet depends on the<br />

vapor diffusivity through the solid layer. In the final stage of drying, boiling followed<br />

by particle drying, eventually leading to dried product formation, takes place.<br />

The different assumptions in developing this mathematical model include the following:<br />

1. The droplet remains spherical in shape throughout the evaporation with spherical<br />

symmetry.<br />

2. Solubility of gas in liquid is negligible.<br />

3. Gas phase is in a quasi-steady state.<br />

4. No influence of chemical reactions occurs within and outside the droplet.<br />

5. No heat transfer due to radiation.<br />

6. No mass diffusion by temperature and pressure gradients.<br />

7. No change in droplet radius once the solid layer formation starts.<br />

8. Internal circulation of water and capillary effects are negligible.<br />

The problem of evaporation and drying of a single droplet can be well defined using<br />

the species mass diffusion and heat conduction equations in spherical coordinates. The<br />

diffusion equation for the substance i in the droplet, formulated in terms of mass<br />

fraction Y i , reads<br />

∂Y i<br />

∂t = D [ (<br />

12 ∂<br />

r 2 ∂Y )]<br />

i<br />

, (2.51)<br />

r 2 ∂r ∂r<br />

where D 12 is the binary diffusion coefficient in the liquid, r is the radial coordinate<br />

within the droplet radius, and t stands for time. In this equation i = 1 denotes the<br />

solvent (water) and i = 2 denotes solute (PVP or mannitol). Initially, the droplet is a<br />

homogenous mixture, Y i = Y i0 at t = 0 s. At the droplet center, r = 0 m, the regularity<br />

condition must be satisfied at any time, ∂Y i /∂r = 0. The boundary condition at the<br />

droplet surface must account for the change in droplet size,<br />

∂Y i<br />

−D 12<br />

∂r − Y ∂R<br />

i<br />

∂t =<br />

ṁi<br />

Aρ l<br />

(2.52)<br />

at r = R(t). Here ṁ i is the mass evaporation rate of substance i across the droplet<br />

surface, R(t) and A(t) are time dependent droplet radius and surface area, respectively,


36 2. Mathematical Modeling<br />

and ρ l is the liquid density.<br />

ṁ i is zero for non-evaporating solute (PVP or mannitol),<br />

i = 2. The diffusion process described through Eq. (2.51) provides the mass fraction<br />

profiles inside the droplet. In or<strong>der</strong> to close this equation, the evaporation rate from the<br />

droplet surface, ṁ i is needed, which appears in Eq. (2.52). This rate of evaporation is<br />

determined based on Sherwood analogy of Abramzon and Sirignano’s model [62], and<br />

in the present study, it is used in the extended form for a bi-component liquid mixture<br />

as modified by Brenn et al. [170],<br />

ṁ i = 2πR i ρ f D f ˜Sh ln(1 + BM,i ), (2.53)<br />

where R i is volume equivalent partial radius of component i, based on its corresponding<br />

volume fraction, computed as R i = R(V i /V ) 1/3 , ˜Sh is the modified Sherwood number<br />

defined by Eq. (2.42), which accounts for the effect of convective droplet evaporation<br />

[62], D f is water vapor diffusivity in film, and ρ f is the density in the film. B M,i is<br />

the Spalding mass transfer number for component i, and it is calculated as [62, 171],<br />

B M,i = Y i,s − Y i,∞<br />

1 − Y i,s<br />

, (2.54)<br />

where Y i,s and Y i,∞ are the mass fractions of evaporating component i at the droplet<br />

surface and in the bulk of the gas, respectively. Nesic and Vodnik [151] implemented<br />

a similar approach, but they do not account for the volume fraction based radius in<br />

the calculation of the evaporation rate, i.e., droplet radius R is used instead of R i<br />

in computing ṁ i . The evaporation rate retardation due to solid layer resistance may<br />

be consi<strong>der</strong>ed through modification of Eq. (2.53) by extending the work of Nesic and<br />

Vodnik [151] to yield<br />

ṁ =<br />

∑ N<br />

i=1 2πR iρ f D f ˜Sh ln(1 + BM,i )<br />

1 + ˜ShD<br />

, (2.55)<br />

f δ/[2D s (R − δ)]<br />

where ṁ is the total evaporation rate, δ is the solid layer thickness at the droplet surface<br />

and D s is the diffusivity of vapor in the solid layer. Since the solute vapor pressure is<br />

low or zero and the droplet’s solute evaporation rate is zero or very small, negligence<br />

of the volume correction (using R in the place of R i ) may lead to an artificial increase<br />

in evaporation rate. In the present situation, the summation in Eq. (2.55) is only over<br />

component 1, because the solute (PVP or mannitol) does not evaporate, but for the<br />

sake of generality, the summation is kept.<br />

During the initial and second stage, δ equals zero. But once the solute mass fraction<br />

at the droplet surface reaches a threshold value, which is most often near saturation<br />

solubility level, there is initiation of solid layer. This solid layer on the droplet surface<br />

offers significant resistance to evaporation and is evident from the second term in the<br />

denominator of Eq. (2.55) [172]. The effect of capillary force on water vapor diffusion


2.4. Single Droplet Modeling 37<br />

through solid layer due to pressure difference in pores is not consi<strong>der</strong>ed, and it is the<br />

scope of the future study. Moreover, the influence of internal circulation within the<br />

droplet is neglected, which can be modeled by a correction for the diffusion coefficient<br />

rather than adding a convection term [151].<br />

The heat conduction equation, describing the conductive heat transfer within the<br />

droplet, is written as<br />

∂T<br />

∂t = α [ ( ∂<br />

r 2 ∂T )]<br />

, (2.56)<br />

r 2 ∂r ∂r<br />

where T is the liquid temperature and α denotes the thermal diffusivity. The above<br />

equation is solved with the following initial and boundary conditions: At t = 0 s, the<br />

droplet is at uniform temperature, T = T 0 .<br />

At the droplet center, r = 0 m, zero<br />

gradient condition prevails at any time, ∂T /∂r = 0. The energy balance at the droplet<br />

surface is given through the boundary condition,<br />

∂T<br />

k l<br />

∂r = h(T ∂R<br />

g − T s ) + L V (T s )ρ l<br />

∂t<br />

(2.57)<br />

at r = R, where R is the droplet radius. In Eq. (2.57), T s denotes droplet surface<br />

temperature, T g stands for gas temperature in the bulk, k l is the liquid thermal conductivity,<br />

h is the convective heat transfer coefficient, and L V (T s ) is the latent heat of<br />

vaporization at the surface temperature, T s .<br />

In this work, first Eq. (2.56) is solved numerically with initial and boundary condition<br />

as defined above using a finite difference method. It is observed that the gradient<br />

in droplet temperature from the center to the droplet surface is very small as the computed<br />

Biot number, which is a measure of heat transfer resistances within and outside<br />

the droplet, (Bi = h/k s R = k g /(2k s )Nu), always remains below 0.5. Therefore, in the<br />

remaining simulations, uniform temperature within the droplet is assumed, which is a<br />

valid assumption as per the revelations made by Mezhericher et al. [173]. The droplet<br />

temperature continuously changes due to heat transfer from ambient gas to the binary<br />

liquid droplet, and it is computed using the energy balance across the droplet, which<br />

gives the net heat transferred into the droplet [62], as<br />

[ ]<br />

dT s<br />

mC pL<br />

dt = Q CpLf (T g − T s )<br />

L = ṁ<br />

− L V (T s ) , (2.58)<br />

B T<br />

where m is the total droplet mass, m = Σ N i=1m i , C pL , C pLf are the specific heat capacity<br />

of liquid and in the film, respectively and B T is the Spalding heat transfer number.<br />

This equation can be used to calculate the time evolution of droplet temperature. Here,<br />

the heat transfer number, B T , is calculated in terms of mass transfer number defined<br />

by Eq. (2.47).<br />

Equation (2.58) needs modification in or<strong>der</strong> to account for the solid layer formation<br />

at droplet surface, and this is achieved through the equation written in terms of the


38 2. Mathematical Modeling<br />

solid layer thickness, δ, as<br />

where<br />

mC pL<br />

dT s<br />

dt = Q L + ṁL V (T s )<br />

1 + Ñu k gfδ/[2k s (R − δ)] − ṁL V (T s ), (2.59)<br />

Ñu is the modified Nusselt number defined by Eq. (2.49), which accounts for<br />

the effect of convective droplet heating [62], k s and k gf are the thermal conductivity of<br />

the solid layer and in the film, respectively, and Q L is the net heat transferred to the<br />

droplet [62], given by Eq. (2.58). Similar to Eq. (2.55), the second term in denominator<br />

inside the bracket of Eq. (2.59) denotes the resistance due to solid formation at the<br />

droplet surface, and its effect becomes significant only when the solid layer thickness,<br />

δ, is positive. The difference between heat transfer and mass transfer resistance is<br />

that the ratio of diffusion coefficients D f /D s is larger than the ratio k g /k s [151], which<br />

implies that resistance to mass transfer due to solid layer formation is higher than the<br />

heat transfer.<br />

In the present study, simulations are also carried out with rapid mixing model (RMM)<br />

which is a simple model based on the assumption that the liquid mixture inside the<br />

droplet is always homogeneous (no spatial gradients of mass fraction within the droplet)<br />

and infinity conductivity within the droplet, thus the droplet is at uniform temperature<br />

at every time. In this work, the RMM is extended to account for solid layer resistance<br />

on the droplet evaporation rate and heating, thus the governing equations in the RMM<br />

are Eq. (2.55) and Eq. (2.59), and the time evolution of solute (i = 2) mass fraction is<br />

calculated based on the simple mass balance of solute and solvent mass fraction within<br />

droplet, given as<br />

Y 2,RMM = Y 02 − m 02<br />

m − ṁ , (2.60)<br />

where Y 02 and m 02 are the initial solute mass fraction and solute mass within the<br />

droplet, respectively.<br />

2.4.2 Droplet Motion<br />

The dynamics of liquid droplets in sprays is the basic physical process that needs to<br />

be computed for the coupling of gas–liquid phases due to its strong dependance on<br />

the flow of surrounding gas. The droplet velocity v at position x can be computed as<br />

following<br />

v = dx<br />

dt . (2.61)<br />

The acceleration of droplets due to different forces acting on droplets can be written<br />

as [174]<br />

dv<br />

dt = ΣF + ρ (<br />

g Du<br />

2ρ l Dt − dv )<br />

, (2.62)<br />

dt


2.4. Single Droplet Modeling 39<br />

where the first term in R.H.S includes all the forces such as aerodynamic drag, gravity,<br />

Basset, lift, and buoyancy etc.<br />

and the second term in R.H.S is the added mass<br />

D<br />

force [174]. In Eq. (2.62), is the substantial or material <strong>der</strong>ivative.<br />

Dt<br />

The force experienced by the droplets due to difference in velocities of droplets<br />

and surrounding gas is known as drag force, F d . The droplet velocity evolution by<br />

interactive drag induced by the surrounding gas, and gravity per unit droplet mass is<br />

commuted using the following relation, which describes droplet motion [175]<br />

F d = 3 8<br />

1 ρ g<br />

(u − v)|u − v|C D + g, (2.63)<br />

r ρ l<br />

where ρ g and u are the density and velocity of the surrounding gas, respectively, while<br />

ρ l , C D and g are liquid density, drag coefficient, and gravitational acceleration, respectively.<br />

The dependencies of the drag force are confined to the droplet radius, droplet<br />

shape, droplet density, ρ l , relative velocity between gas and droplet, u − v, gas density,<br />

ρ g , kinematic viscosity of the gas, η g , and surface tension, σ d .<br />

The drag coefficient, C D , is calculated as a function of the droplet Reynolds number,<br />

Re d = 2rρ g |u − v|/µ f , where µ f is the mean dynamic viscosity in the film, as [176]<br />

{<br />

24<br />

(1+ Re<br />

C D =<br />

1 d 6 Re0.687 d ) if Re d


40 2. Mathematical Modeling<br />

surrounding gas on the droplets. Buoyancy force is equal to the weight of the displaced<br />

gas due to droplet motion.<br />

The added mass force, defined by second term in the right hand side of Eq. (2.62),<br />

accounts for the acceleration of the gas due to the droplet motion. When a droplet<br />

accelerates in gas, it implies an acceleration of the surrounding gas at the expense of<br />

the force exerted by the droplet. Since the added mass force depends on the fluid<br />

density, it is often neglected for droplets much denser than the gas [177]. In this work,<br />

the ratio between liquid density and gas density is about 10 3 , so the effect of added<br />

mass can be neglected.<br />

The unsteady behavior of the droplet, buoyancy effects, compressibility of the gas,<br />

rotation effects, the fluid motion within the droplet or other subtle forces are not<br />

consi<strong>der</strong>ed. It can be shown [174] that terms originating from these phenomena are<br />

negligible for large ratios of droplet to gas densities, and for low droplet Mach numbers,<br />

Ma = |u − v|/c < 0.03, where c is the speed of sound in the gas.<br />

2.4.3 Droplet Breakup<br />

Liquid drops generated from the primary breakup of the liquid sheet, moving in the<br />

surrounding gas may un<strong>der</strong>go further breakup or disintegration un<strong>der</strong> certain conditions,<br />

leading to formation of smaller droplets. This phenomenon is called as droplet<br />

breakup or secondary atomization. The exact mechanisms of the droplet breakup is<br />

not yet completely un<strong>der</strong>stood as there are many uncertainties in the quantitative description<br />

of the process. The relative motion between a droplet and the surrounding<br />

gas causes a non-uniform distribution of pressure and shear stress on the droplet surface,<br />

which results in deformation of the droplet and cause it to disintegrate when<br />

they overcome the opposing force of surface tension. The newly formed droplets may<br />

still un<strong>der</strong>go further breakup until surface tension force of the newly formed droplet<br />

is higher than the external forces. The work of Pilch and Erdmann [178] explained<br />

the various regimes of breakup, which are depicted in Fig. 2.6. Faeth et al. [70] and<br />

Faeth [72] give an overview of existing mechanisms of droplet breakup.<br />

According to Faeth et al. [70], the breakup regime transitions are mainly functions<br />

of the gas Weber number, We g , and the Ohnesorge number, Oh. The Weber number<br />

is defined as the ratio between the drag force to surface tension force, written as<br />

We g = 2rρ g|u − v| 2<br />

, (2.67)<br />

σ<br />

where σ is the surface tension and r is the droplet radius. The Ohnesorge number,<br />

represents the ratio of viscous forces to inertial and surface tension forces, given as<br />

Oh =<br />

µ l<br />

√ 2ρl rσ , (2.68)


2.4. Single Droplet Modeling 41<br />

(a) Vibrational breakup, We g ≈ 12 (b) Bag breakup, We g < 20<br />

(c) Bag / streamer breakup, We g < 50 (d) Stripping breakup, We g < 100<br />

(e) Catastrophic breakup, We g > 100<br />

Fig. 2.6: Droplet breakup mechanisms based on Weber number [69, 178].<br />

where µ l is the liquid viscosity. The existing breakup models developed based on<br />

the various mechanisms include, wave breakup (WB) model [179], Taylor analogy<br />

breakup (TAB) model [180], enhanced Taylor analogy breakup (ETAB) model [181],<br />

Rayleigh-Taylor instability (RTI) model [182], and droplet deformation and breakup (DDB)<br />

model [183]. Madsen [61] extended DQMOM to include droplet coalescence and<br />

breakup in spray flows by neglecting the effects of evaporation. In the present study,<br />

the focus is on the influence of droplet coalescence, evaporation and drag on droplet<br />

characteristics, and the study concerns the spray at a distance after the atomization,<br />

which may not breakup further, the droplet breakup is currently neglected.<br />

2.4.4 Droplet Coalescence<br />

The droplets in spray flows when come close enough, they interact with each other<br />

leading to collision of droplets. The collision dynamics of liquid droplets is important<br />

in the evolution of spray flows as they can significantly effect the spray characteristics<br />

such as droplet size and velocity distribution, and in turn influence the final pow<strong>der</strong><br />

characteristics in spray drying process.


42 2. Mathematical Modeling<br />

Fig. 2.7: Droplet collision regimes: (a) bouncing, (b) coalescence [184].<br />

The outcome of the droplet collision is mostly dependent on the size, mass, surface<br />

tension, and velocity of colliding droplets. The collision outcome is classified into four<br />

different types: bounce, coalescence, reflexive separation and stretching separation.<br />

This classification is depicted in Figs. 2.7 and 2.8 [184]. Reflexive separation which<br />

occurs in head on and near-head on collision of droplets from miscible liquids does not<br />

exist for the immiscible liquids [185], and the collision mechanism in immiscible liquids<br />

is identified by Planchette and Brenn, and termed it as crossing separation [185].<br />

The spray models developed to account for the droplet–droplet interactions mostly<br />

assume that there are only two possibilities of collision outcome: the droplets rebound<br />

without any change in droplet size or they coalesce to give a single droplet [186]. These<br />

models are only applicable to the study of two droplets colliding with each other but<br />

Fig. 2.8: Droplet collision regimes: (c) reflexive or crossing separation, (d) stretching<br />

separation [184].


2.4. Single Droplet Modeling 43<br />

not to the spray itself as the extension of these models to dense spray flows is much<br />

more complex [184], because of individual droplet tracking requirement. Implementation<br />

of droplet coalescence models needs tracking of individual droplets as done in<br />

Euler – Lagrangian simulations. In case of Euler – Euler models, droplet distribution<br />

is computed but the individual droplets are not tracked. Hylkema and Villedieu [187]<br />

developed a droplet collision model based on the droplet distribution, which can be<br />

implemented in Euler–Euler methods. In the current study, as the spray flow is modeled<br />

using Eulerian approach where the global droplet distribution is computed, so the<br />

droplet collisions are taken into account as described by Hylkema and Villedieu [187]<br />

and Laurent [46]. To emphasize upon coalescence only, standard assumptions [46] for<br />

droplet coalescence have been employed. These assumptions imply that each binary<br />

collision either leads to coalescence (E c = 1) or rebound (E c = 0), and conservation<br />

of mass and momentum before and after the collision [46] is assured. In addition, the<br />

mean collision time is assumed to be smaller than the inter-collision time. Thus, the<br />

coalescence function can be written in terms of the flux of newly formed droplet, Q + c<br />

and flux of the vanishing droplets, Q − c , given by [187]<br />

Γ f = Q + c + Q − c , (2.69)<br />

where Q + c and Q − c are calculated as<br />

Q − c<br />

∫ ∞<br />

= −<br />

−∞<br />

∫ ∞<br />

0<br />

f(t, x; r, v)f(t, x; r 1 , v 1 ) × B(|v − v 1 |)dr 1 dv 1 , (2.70)<br />

Q + c =<br />

∫ ∞ ∫ ∞<br />

−∞<br />

where B(|v 1 − v 2 |) is given by,<br />

0<br />

1<br />

2 f(t, x; r 1, v 1 )f(t, x; r 2 , v 2 ) × B(|v 1 − v 2 |)dr 1 dv 1 , (2.71)<br />

B(|v 1 − v 2 |) = π(r 1 + r 2 ) 2 |v 2 − v 1 |E c , (2.72)<br />

and B(|v − v 1 |) is defined accordingly. In the above equations, (r, v) refer to postcollision<br />

properties, which are related to pre-collision properties (r 1 , v 1 ) and (r 2 , v 2 )<br />

through the relations [46, 187]<br />

v = r3 1v 1 + r2v 3 2<br />

, (2.73)<br />

r1 3 + r2<br />

3<br />

r 3 = r1 3 + r2. 3 (2.74)<br />

The collision efficiency is computed following the work of O’Rourke [186], which is<br />

written as<br />

E c =<br />

K 2<br />

(K + 1/2) 2 , (2.75)


44 2. Mathematical Modeling<br />

where K is given as [186]<br />

K = 2 ρ l |v 1 − v 2 |r2<br />

2 . (2.76)<br />

9 µ g r 1<br />

For DQMOM, substitution of all these sub-models defined in Subsection 2.4 into the<br />

Eq. (2.29) results in a linear system, which in turn is substituted into Eq. (2.30) to<br />

yield in a linear system of equations.<br />

The solution of this linear system gives the<br />

various source terms, i.e., a n , b n and c n that appear in DQMOM transport Eqs. (2.26),<br />

(2.27) and (2.28), which then constitutes a closed system. The numerical solution<br />

procedure to solve these equations with initial and boundary conditions is explained<br />

in Chapter 3. Droplet–droplet interactions are currently neglected in DDM because of<br />

the computational complexity involved if Lagrangian models are used.<br />

In summary, the new implementations of the present study include implementation<br />

of an advanced droplet evaporation model for water sprays in DQMOM, a new mathematical<br />

model development for the polymer or sugar dissolved in water droplets evaporation<br />

and solid layer formation at the droplet surface, and improvement of the evaporative<br />

flux calculation with maximum entropy formulation. Extension of DQMOM<br />

to simulate two-dimensional system, and implementation of developed bi-component<br />

evaporation model in DQMOM.


3. Numerical Methods<br />

In the numerical simulation, the governing equations are discretized and solved by<br />

computer programs where appropriate numerical algorithms are required. An ideal<br />

numerical algorithm should<br />

• be linearly stable for all cases of interest;<br />

• ensure the positivity property when appropriate;<br />

• be reasonably accurate;<br />

• be computationally efficient.<br />

There are several numerical methods available for the fluid mechanics. The methods<br />

ranging from the most discrete (or particulate) in nature to the most continuous (or<br />

global) include:<br />

• particle methods<br />

• characteristic methods<br />

• Lagrangian finite difference/finite volume method<br />

• Eulerian finite difference/finite volume method<br />

• finite element methods<br />

• spectral methods<br />

Each method has advantages and disadvantages, consequently has the preferable applications.<br />

Usually, it is difficult or inefficient for a stand-alone method to simulate<br />

a complex system. Hybrid method, which is like a bootstrapping process, combines<br />

the advantages of the multiple methods and minimizes their disadvantages. The disadvantage<br />

of hybrid method is that the consistency problem is more serious. Special<br />

strategies are needed to keep consistency between the multiple methods.<br />

In this chapter the numerical methods employed to solve the single droplet evaporation<br />

and drying equations, QMOM and DQMOM transport equations are explained.<br />

In this work, the DDM computations performed by Humza [68] are used to validate the<br />

DQMOM results for water spray in air in two-dimensional, axisymmetric configuration.<br />

Hence, the numerical details of the DDM simulations can be referred to Humza [68].


46 3. Numerical Methods<br />

3.1 Finite Difference Method for Bi-component<br />

Droplet Evaporation and Solid Layer<br />

Formation<br />

The partial differential equation, Eq. (2.51), with initial and boundary conditions is<br />

solved numerically at every time and spatial location within the droplet using second<br />

or<strong>der</strong> explicit finite difference method, given as<br />

Y j+1<br />

i<br />

∆t<br />

− Y j<br />

i<br />

[<br />

r<br />

2<br />

= D i+1 (Y j<br />

i+1 − Y j<br />

i ) − r2 i−1(Y j<br />

i − Y j<br />

i−1 )]<br />

12 , (3.1)<br />

ri 2∆r2 where r i+1 = r i + ∆r, r i−1 = r i − ∆r, and i and j are the spatial location within the<br />

droplet and time step indices, respectively. Equation (3.1) can be simplified to yield<br />

the following equation<br />

Y j+1<br />

i<br />

= Y j<br />

i<br />

+ D 12∆t [<br />

r<br />

2<br />

ri 2 i+1 (Y j<br />

∆r2 i+1 − Y j<br />

i ) − r2 i−1(Y j<br />

i − Y j<br />

i−1 )] . (3.2)<br />

The initial condition to compute Eq. (3.2) is provided as a Dirichlet condition, i.e.,<br />

Y = Y i0 at every location inside the droplet at t = 0. A Neumann boundary condition<br />

is applied at the center of the droplet, i.e., ∂Y i /∂r = 0 at r = 0, which implies the<br />

radial symmetry within the droplet. A Robin boundary condition is employed at the<br />

droplet surface, and it is given by Eq. (2.52).<br />

The energy Eq. (2.59) is an ordinary differential equation, solved using Runge-Kutta<br />

4 th or<strong>der</strong> method. The droplet is discretized into equal distant grid points at any given<br />

time. As the droplet size decreases with time thereby the grid size changes because<br />

grid points are fixed, thus a moving grid problem is solved, and grid independency of<br />

the numerical method is tested using different grid sizes with the number of grid points<br />

varying from 10 to 100. The value of 50 grid nodes is found to perform well.<br />

The numerical stability of the method is tested using various time steps following<br />

the Courant-Friedrichs-Lewy (CFL) condition [188]. The CFL condition defines the<br />

limiting criteria for the numerical grid size when the time step and fluid velocity are<br />

known, and it is defined as<br />

C = u∆t<br />

∆x ≤ C max, (3.3)<br />

where C is the dimensionless number known as Courant number, u is the velocity, ∆t<br />

and ∆x are the time step and grid size, respectively. C max is the maximum possible<br />

Courant number to get a stable numerical solution, and it is generally taken as any<br />

positive value lower than or equal to 0.5 [188]. The step-by-step procedure of Abramzon<br />

and Sirignano [62] is applied to calculate the mass evaporation rate given by Eq. (2.55).<br />

Numerical simulations of pure water, mannitol dissolved in water droplet evaporation


3.2. Spray Modeling 47<br />

and solid layer formation is done to test the implementation of this algorithm and the<br />

numerical results are compared with experimental data. The results are presented in<br />

Chapter 4.<br />

3.2 Spray Modeling<br />

3.2.1 Finite Volume Method for QMOM<br />

In the present study, QMOM is implemented with a three-node (three weights or number<br />

densities, three droplet radii, and three droplet velocities) closure approximation<br />

of the NDF, which requires a total of nine moments of the NDF to compute the initial<br />

data of droplet radii and velocities and corresponding weights. The transport equations<br />

are generated by selecting k 1 ∈ {0, 1, 2, 3} and k 2 ∈ {0, 1} in Eq. (2.24), which<br />

is equivalent to three-node closure. The choice of three-node closure with the mentioned<br />

values of k 1 and k 2 is proven to be accurate in previous studies [48, 49, 51]. The<br />

substitution of k 1 and k 2 values results in the following equations:<br />

∂M(0, 0)<br />

∂t<br />

+<br />

∂M(0, 1)<br />

∂x<br />

=<br />

∫ ∞ ∫ ∞<br />

−∞<br />

0<br />

[<br />

− ∂ (Rf) − ∂(Ff) ]<br />

∂r ∂v + Q f + Γ f drdv, (3.4)<br />

∂M(1, 0)<br />

∂t<br />

∂M(0, 1)<br />

∂t<br />

∂M(1, 1)<br />

∂t<br />

∂M(2, 1)<br />

∂t<br />

+<br />

+<br />

+<br />

+<br />

∂M(1, 1)<br />

∂x<br />

∂M(0, 2)<br />

∂x<br />

∂M(1, 2)<br />

∂x<br />

∂M(2, 2)<br />

∂x<br />

=<br />

=<br />

=<br />

=<br />

∫ ∞ ∫ ∞<br />

−∞<br />

0<br />

∫ ∞ ∫ ∞<br />

−∞<br />

0<br />

∫ ∞ ∫ ∞<br />

−∞<br />

0<br />

∫ ∞ ∫ ∞<br />

−∞<br />

0<br />

[<br />

r − ∂ (Rf) − ∂(Ff) ]<br />

∂r ∂v + Q f + Γ f drdv, (3.5)<br />

[<br />

v − ∂ (Rf) − ∂(Ff) ]<br />

∂r ∂v + Q f + Γ f drdv, (3.6)<br />

[<br />

rv − ∂ (Rf) − ∂(Ff) ]<br />

∂r ∂v + Q f + Γ f drdv, (3.7)<br />

[<br />

r 2 v − ∂ (Rf) − ∂(Ff) ]<br />

∂r ∂v + Q f + Γ f drdv (3.8)<br />

∂M(3, 1)<br />

∂t<br />

+<br />

∂M(3, 2)<br />

∂x<br />

=<br />

∫ ∞ ∫ ∞<br />

−∞<br />

0<br />

[<br />

r 3 v − ∂ (Rf) − ∂(Ff) ]<br />

∂r ∂v + Q f + Γ f drdv. (3.9)<br />

The M(0, 2), M(1, 2), M(2, 2) and M(3, 2) fall away from the selected moment set<br />

defined by k 1 and k 2 values, and these four unclosed moments are computed in terms<br />

of the weights and abscissas:


48 3. Numerical Methods<br />

M(0, 2) = w 1 v 2 1 + w 2 v 2 2 + w 3 v 2 3, (3.10)<br />

M(1, 2) = w 1 r 1 v 2 1 + w 2 r 2 v 2 2 + w 3 r 3 v 2 3, (3.11)<br />

M(2, 2) = w 1 r 2 1v 2 1 + w 2 r 2 2v 2 2 + w 3 r 2 3v 2 3, (3.12)<br />

M(3, 2) = w 1 r 3 1v 2 1 + w 2 r 3 2v 2 2 + w 3 r 3 3v 2 3. (3.13)<br />

These weights and abscissas are computed using the Wheeler algorithm (see Subsection<br />

3.2.3). Similarly, if any of the terms on the right hand side of these Eqs. (3.4) – (3.9)<br />

contain unknown moments, they will be closed in the analogous manner.<br />

To solve Eqs. (3.4) – (3.9) a numerical scheme based on a kinetic transport scheme<br />

to evaluate the spatial fluxes [96, 189] can be employed. A first-or<strong>der</strong>, explicit, finite<br />

volume scheme for these equations can be written for the set of moments<br />

as<br />

M =<br />

M n+1<br />

[<br />

M(0, 0), M(1, 0), M(0, 1), M(1, 1), M(2, 1), M(3, 1)] T<br />

(3.14)<br />

i = Mi n − ∆t [<br />

]<br />

G(Mi n , M n<br />

∆x<br />

i+1) − G(Mi−1, n Mi n ) + ∆tSi n (3.15)<br />

where n is the time step, i is the grid node, S is the right hand side estimate of<br />

Eqs. (3.4)– (3.9), and G is the flux function. Using the velocity abscissas, the movement<br />

of the quadrature node from left to right or right to left is determined. The flux function<br />

at any time step is expressed as [28]<br />

G(M i , M i+1 ) = H + (M i ) + H − (M i+1 ) (3.16)<br />

where<br />

⎛ ⎞<br />

⎛ ⎞<br />

⎛ ⎞<br />

1<br />

1<br />

1<br />

r 1<br />

r 2<br />

r 3<br />

H − (M) = w 1 min(v 1 , 0)<br />

v 1<br />

r 1 v + w 2 min(v 2 , 0)<br />

v 2<br />

1<br />

r 2 v + w 3 min(v 3 , 0)<br />

v 3<br />

2<br />

r 3 v ,<br />

3<br />

⎜<br />

⎝r1v 2 ⎟<br />

⎜<br />

1 ⎠<br />

⎝r2v 2 ⎟<br />

⎜<br />

2 ⎠<br />

⎝r3v 2 ⎟<br />

3 ⎠<br />

r1v 3 1 r2v 3 2 r3v 3 3<br />

(3.17)<br />

and


3.2. Spray Modeling 49<br />

⎛ ⎞<br />

⎛ ⎞<br />

⎛ ⎞<br />

1<br />

1<br />

1<br />

r 1<br />

r 2<br />

r 3<br />

H + (M) = w 1 max(v 1 , 0)<br />

v 1<br />

r 1 v + w 2 max(v 2 , 0)<br />

v 2<br />

1<br />

r 2 v + w 3 max(v 3 , 0)<br />

v 3<br />

2<br />

r 3 v .<br />

3<br />

⎜<br />

⎝r1v 2 ⎟<br />

⎜<br />

1 ⎠<br />

⎝r2v 2 ⎟<br />

⎜<br />

2 ⎠<br />

⎝r3v 2 ⎟<br />

3 ⎠<br />

r1v 3 1 r2v 3 2 r3v 3 3<br />

(3.18)<br />

Higher-or<strong>der</strong> flux schemes can also be developed to control numerical diffusion [190].<br />

However, the key characteristics of the flux function is that the quadrature method<br />

provides a realizable set of weights and abscissas at every grid node that can be used<br />

to determine the node velocities. In the present study, only the steady state solution<br />

of Eqs. (3.4) – (3.9) is needed due to the fact that experimental data provides only the<br />

time averaged droplet properties. The steady form of Eqs. (3.4) – (3.9) is solved using<br />

Runge-Kutta 4 th or<strong>der</strong> method. The QMOM simulations are carried out only for onedimensional<br />

water spray in nitrogen in or<strong>der</strong> to compare and validate DQMOM results,<br />

and the initial data for the QMOM simulations are generated from the experimental<br />

data by calculating the above moment set, and the initial data generation procedure<br />

is outlined in Chapter 4.<br />

3.2.2 Finite Difference Scheme for DQMOM<br />

A generalized model for three-dimensional physical space has been discussed for application<br />

to evaporating sprays [191]. At first, DQMOM is applied to study the steady<br />

spray flows in one physical dimension, i.e., in the axial direction x. Thus, inhomogeneous<br />

formulation also known as steady state form of DQMOM transport Eqs. (2.26) –<br />

(2.28) can be rewritten as below by neglecting the terms containing time, t,<br />

where<br />

∂U n<br />

∂x = S n, (3.19)<br />

U n ∈ {w n v n , w n ρ l r n v n , w n ρ l r n v n v n },<br />

S n ∈ {a n , ρ l b n , ρ l c n }.<br />

Similarly, the homogeneous formulations of DQMOM transport Eqs. (2.26) – (2.28)<br />

can be rewritten as following by ignoring the spatial terms,<br />

∂U n<br />

∂t<br />

= S n , (3.20)


50 3. Numerical Methods<br />

where<br />

U n ∈ {w n , w n ρ l r n , w n ρ l r n v n },<br />

S n ∈ {a n , ρ l b n , ρ l c n }.<br />

To solve these equations, the choice of numerical scheme is important because this<br />

array of equations are strongly coupled. Different finite difference schemes with varying<br />

or<strong>der</strong> of accuracy are tested.<br />

It has been shown [191] that Runge-Kutta 4 th or<strong>der</strong><br />

method can accurately solve the system of inhomogeneous equations represented by<br />

Eq. (3.19) and proven to be computationally efficient for DQMOM in one-dimensional<br />

physical space [191]. In the current study, the NDF is approximated by a three-node<br />

closure in DQMOM, which is proven to be accurate in previous studies [48, 49]. The<br />

three-node approximation of NDF implies that a total of nine coupled equations, which<br />

are generated by substituting n = {1, 2, 3}, in Eq. (3.19). These equations are solved<br />

to find the evolution of NDF, which is achieved by discretizing and estimating these<br />

equations with Runge-Kutta 4 th or<strong>der</strong> method. At every spatial location within the<br />

geometry, the source terms are computed through the models proposed in Chapter 2.<br />

The flowchart shown in Fig. 3.1 outlines the step by step procedure of the computational<br />

code.<br />

The previous studies concerning DQMOM in spray flows, transport equations are<br />

never solved in two-dimensional configuration but only in one dimension. In this study,<br />

the DQMOM transport equations are solved in two-dimensional (axial and radial direction)<br />

geometrical configuration for water and PVP/water sprays in air, by implementing<br />

a finite difference numerical scheme. At each axial and radial location, the<br />

coupled steady state transport equations of DQMOM are solved and the source terms<br />

such as droplet heating, evaporation rate, total forces acting on droplet and droplet coalescence<br />

are computed from the weights and abscissas available from the initial values<br />

at first iteration and from the last computed value in the next iterations. The steady<br />

form of the DQMOM transport Eqs. (2.26) – (2.28) in two dimensions can be written<br />

as<br />

where<br />

∂U n<br />

∂x + ∂E n<br />

∂z = S n, (3.21)<br />

U n ∈ {w n v n , w n ρ l r n v n , w n ρ l r n v n u n , w n ρ l r n v n v n },<br />

E n ∈ {w n u n , w n ρ l r n u n , w n ρ l r n u n u n , w n ρ l r n u n v n }.<br />

(3.22)<br />

In Eq. (3.21), x is the axial direction, z is the radial direction, and the corresponding<br />

velocities are v and u, respectively.<br />

To keep the computational efficiency, ease of


3.2. Spray Modeling 51<br />

application and numerical accuracy, a second or<strong>der</strong> explicit finite difference scheme is<br />

applied to solve steady state form of Eqs. (2.26) – (2.28) [192], which are represented<br />

by Eq. (3.21). Thus the solution formula may be written as [193]<br />

U j+1<br />

n,i<br />

= U j n,i − ∆x [<br />

1.5E<br />

j<br />

i<br />

∆z<br />

− 2Ej i−1 + ] 0.5Ej i−2 + ∆xS<br />

j<br />

i , (3.23)<br />

where i and j are grid nodes in radial and axial directions, respectively.<br />

The above formulation is applied to an equidistant rectangular grid, where the size<br />

of each grid cell is 1.5 × 10 −3 m in radial direction and 1.0 × 10 −4 m in axial<br />

direction, resulting in a maximum of 80 × 1000 grid nodes. The initial data to start<br />

simulations in both the configurations, i.e, one and two-dimensional cases is generated<br />

from the experimental data provided by Dr. R. Wengeler, BASF Ludwigshafen (onedimensional<br />

water spray in nitrogen) and Prof. G. Brenn, TU Graz (two-dimensional<br />

water and PVP/water spray in air) using Wheeler algorithm (see Subsection 3.2.3). The<br />

experimental data closest to the nozzle exit is taken for generating the initial data and<br />

the procedure for calculating this initial data from experiment is explained in Chapter<br />

4 along with brief description about the experimental setup. The boundary conditions<br />

in solving DQMOM include (1) if droplets hit the axis of symmetry, they are reflected,<br />

and (2) Neumann boundary is applied for the lateral sides of the computational domain<br />

and exit plane. The experimental data available at other cross sections away from the<br />

nozzle exit is used to validate the simulation results. The flowchart of the computational<br />

code is illustrated in Fig. 3.1.<br />

3.2.3 Wheeler Algorithm<br />

The Wheeler algorithm developed by Sack and Donovan [136], requires 2N +1 moments<br />

to compute N weights (number density) and N abscissas (droplet radii or velocities).<br />

The moment set is represented as M = [M(0), M(1), ...M(2N + 1)] T . This algorithm<br />

is used to generate the initial data in DQMOM whereas in QMOM it is used to compute<br />

the unknown moments. The first step in Wheeler algorithm is to compute the<br />

coefficients π α based on these 2N + 1 moments of the distribution function n(ξ), given<br />

as<br />

π α+1 (ξ) = ξπ α (ξ). (3.24)<br />

The above recursive relation has the properties of π −1 (ξ) = 0 and π 0 (ξ) = 1. Here, α<br />

is a subset of number of moments 2N + 1, i.e., α ∈ 0, 1, 2..N − 1. From these coefficients<br />

π α (ξ), a symmetric tridiagonal matrix is computed through some intermediate<br />

quantities:<br />

∫<br />

σ α,β = n(ξ)π α (ξ)π β (ξ)dξ, (3.25)


52 3. Numerical Methods<br />

Initialization<br />

(Generate weights and abscissas from the<br />

moments through Wheeler algorithm)<br />

Grid generation<br />

Compute droplet evaporation, Eq. (2.55)<br />

Compute droplet heating, Eq. (2.59)<br />

Compute drag force, Eq. (2.63)<br />

Compute coalescence, Eq. (2.69)<br />

Solve source terms of spray, Eq. (2.30)<br />

Update weights, abscissas,<br />

Eqs. (2.26) – (2.28)<br />

No<br />

Is the<br />

final measurement<br />

position<br />

reached?<br />

Stop<br />

Yes<br />

Fig. 3.1: Flowchart of the DQMOM computational code.


3.3. Numerical Performance 53<br />

where β ∈ α, α + 1, ...2N − α − 1. These quantities, σ α,β , are calculated by initializing<br />

and a 0 = M(1)/M(0), b 0 = 0. The recurrence relation is<br />

σ −1,α = 0, (3.26)<br />

σ 0,α = M(α), (3.27)<br />

σ α,β = σ α−1,β+1 − a α−2 σ α−1,β − b β−1 σ α−2,β , (3.28)<br />

where the tridiagonal matrix components are given as<br />

a α = σ α,α+1<br />

− σ α−1,α<br />

, (3.29)<br />

σ α,α σ α−1,α−1<br />

σ α,α<br />

b α = . (3.30)<br />

σ α−1,α−1<br />

Here, the values of a α are the diagonal elements and b α are the upper and lower diagonal<br />

elements of the symmetric tridiagonal matrix. The eigenvalues of this matrix are the<br />

abscissas (droplet radii, velocities) where as the corresponding eigenvectors are the<br />

weights (number densities). More details about <strong>der</strong>ivation of this algorithm is given<br />

by Gautschi [194], and example calculations are given by Marchisio and Fox [71].<br />

3.3 Numerical Performance<br />

For the DDM computations, which are carried out Humza [68], a hybrid finite volume<br />

method based on the SIMPLER (Semi-Implicit Method for Pressure-Linked Equations<br />

- Revised) algorithm [68, 195] is used to solve the mean conservation equation of the<br />

gas flow, and a Lagrangian stochastic droplet parcel method is used for the spray flow.<br />

The initial and boundary conditions are generated from the experimental data. A<br />

non-equidistant rectangular numerical grid is used, which is finer in the region near<br />

the nozzle exit with a total of 78 × 101 grid nodes. The numerical time step for<br />

the governing gas phase equations is controlled by applying the CFL condition [188].<br />

The solution algorithm and numerical details of the DDM calculation are given by<br />

Humza [68].<br />

The DQMOM simulations are carried out on a PC with two Intel dual core 2.2 GHz<br />

processors having 8 GB RAM. The DDM is simulated on a PC having an AMD quad<br />

Opteron 1.8 GHz processor with 64 GB RAM [68]. The latter PC had several jobs<br />

running simultaneously, so that the available RAM on both the PCs is about identical.<br />

All simulations are run on a single processor. The computations for DQMOM and DDM<br />

take about one hour and three days, respectively. Thus, the DQMOM computations<br />

show a much better performance with respect to the computational cost.


54 3. Numerical Methods


4. Results and Discussion<br />

Results presented in this chapter are categorized into four sections based on the model<br />

development and implementation, which are discussed successively: At first, onedimensional<br />

water spray in nitrogen results are given, followed by the results of twodimensional<br />

evaporating water spray flows in air in axisymmetric configuration. Later,<br />

single bi-component droplet evaporation and solid layer formation results are discussed,<br />

and finally, results of PVP/water spray in air in an axisymmetric configuration are presented.<br />

4.1 One-dimensional Evaporating Water Spray in<br />

Nitrogen<br />

A spray can be generated by pumping the liquid through a nozzle that facilitates<br />

dispersion of liquid into a spray. Nozzles are mainly used to distribute a liquid over an<br />

area thereby liquid surface area is increased. There are three types of nozzles normally<br />

used, which include spinning disk nozzle, single-fluid or centrifugal pressure nozzle,<br />

twin-fluid nozzle [196]. The spinning disk nozzles are also known as rotary atomizers.<br />

The single-fluid nozzles include pressure-swirl nozzle, plain-orifice nozzle, hollow cone<br />

nozzle, etc., whereas the twin-fluid nozzles can be internal-mix or external-mix two-fluid<br />

atomizers [196]. The present study concerns the simulation of water spray generated<br />

using a hollow cone nozzle, which is single fluid nozzle. However, the model presented<br />

in this work is equally applicable to other type of nozzles as the current work focuses<br />

on the simulation of spray after the primary breakup.<br />

Evaporating sprays are of special interest as those occur not only in many industrial<br />

applications but also constitute the defining physical phenomena in spray drying<br />

process. Therefore, having models validated for evaporating sprays motivate their application<br />

in simulations of spray drying. A water spray injected through a hollow cone<br />

Delavan SDX-SE-90 nozzle in a vertical spray chamber and carried by nitrogen is simulated<br />

by DQMOM and the results are compared with the QMOM, and validated with<br />

the experiment.


56 4. Results and Discussion<br />

Fig. 4.1: Photograph of the water<br />

spray formation.<br />

Fig. 4.2: Schematic diagram of spray<br />

with measurement positions.<br />

4.1.1 Experimental Setup<br />

Experiments have been carried out by Dr. R. Wengeler at BASF, Ludwigshafen, where<br />

a water spray is injected into a cylindrical spray chamber. The carrier gas is nitrogen<br />

at room temperature. Three different experiments are conducted by keeping the<br />

spray inflow rate at 80, 150 and 200 kg/h while the gas volumetric flow rate is fixed<br />

at 200 Nm 3 /h. The droplet size distribution is recorded at sections of 0.14, 0.54,<br />

and 0.84 m distance from the nozzle exit using laser Doppler anemometry (LDA).<br />

Measurements at 0.14 m are taken as a starting point for initial data generation for<br />

computations. Figure 4.1 shows the photograph of water spray formation in experiment,<br />

and the schematic representation of spray with dimensions and measurement<br />

positions in experiment is shown in Fig. 4.2. The spray column has a diameter of 1 m.<br />

The present simulations concern the experimental data generated using the Delavan<br />

nozzle SDX-SE-90 with an internal diameter of 2 mm and an outer diameter of 12 mm<br />

at the nozzle throat and 16 mm at the top.<br />

4.1.2 Initial Data Generation<br />

The experimental data provide the cumulative volume frequency of different droplet<br />

sizes. These volume frequencies are converted into surface frequencies by dividing the<br />

individual volume frequency with the corresponding diameter. Figure 4.3 shows the<br />

surface frequencies at the distance of 0.14 m (left) and 0.54 m (right) from the nozzle<br />

exit, respectively, as obtained from the experimental data. At 0.14 m distance away<br />

from the nozzle, there is a higher number of small-sized droplets shown in the left side<br />

of Fig. 4.3, whereas at 0.54 m distance, an increased number of larger size droplets<br />

is found, see right part of Fig. 4.3. The droplet velocities are not measured in the


4.1. One-dimensional Evaporating Water Spray in Nitrogen 57<br />

Surface frequency [µm ­1 ]<br />

0.12<br />

1<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

0<br />

10 1 10 2 10 3<br />

Droplet diameter [µm]<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Cumulative surface frequency [µm ­1 ]<br />

Surface frequency [µm ­1 ]<br />

0.12<br />

1<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

0<br />

10 1 10 2 10 3<br />

Droplet diameter [µm]<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Cumulative surface frequency [µm ­1 ]<br />

Fig. 4.3: Experimental surface frequency distribution at cross section 0.14 m (left) and<br />

0.54 m (right) away from the nozzle exit.<br />

experiments, and they are calculated using the relation<br />

√<br />

ρ l − ρ g<br />

v = 1.74 gd (4.1)<br />

ρ g<br />

given by Stieß [197], where d is droplet diameter, ρ l and ρ g are the liquid and gas<br />

densities, respectively, and g is the acceleration due to gravity. This relation is the<br />

estimate of the terminal velocity of the droplets [197], and it is proven to give accurate<br />

value of the droplet velocity [191, 198].<br />

The moment sets are calculated by means<br />

of these droplet radius, velocity and surface frequency, which are used as initial data<br />

for QMOM whereas for DQMOM these moments are in turn used to calculate the<br />

weights (representing surface frequencies), radii and velocities through the Wheeler<br />

algorithm [136] as explained in Chapter 3. These data (weights, radii and velocities)<br />

are then used as initial data to start the computations.<br />

Tab. 4.1 lists these initial<br />

values with three-node approximation for 80 kg/h and 150 kg/h water inflow rate.<br />

Tab. 4.1: Initial weights and abscissas<br />

Liquid flow rate [kg/h] Weights [(µm) −1 ] Radii [µm] Velocities [m/s]<br />

80 0.638 24.424 1.09<br />

0.276 86.432 1.94<br />

0.086 143.29 2.76<br />

150 0.733 21.739 1.03<br />

0.223 79.706 1.84<br />

0.044 128.350 2.72


58 4. Results and Discussion<br />

4.1.3 Results and Discussion<br />

In this subsection, the simulation results of DQMOM in one dimension are presented,<br />

and compared with experimental data and with the simpler model QMOM. Computations<br />

are carried out consi<strong>der</strong>ing different ambient gas temperatures, i.e. 293 K,<br />

as in experiments, and 313 K, and different inflow rates of liquid as in experiments<br />

in or<strong>der</strong> to investigate the effect of evaporation and drag force along with gravity on<br />

droplet characteristics and spray dynamics. The gas inflow rate remains fixed in accordance<br />

with experiments. Although the surrounding gas velocity is fixed as 0.078 m/s<br />

in experiments, the simulations are performed using different velocities of surrounding<br />

gas in or<strong>der</strong> to analyze the effect of drag force on the droplet dynamics. Also, the<br />

cases of spray with and without coalescence are compared to analyze the influence of<br />

coalescence on droplet distribution.<br />

First, the implementation of the droplet evaporation model is tested for a single pure<br />

water droplet. The numerical predictions and experimental results of water droplet<br />

mass and temperature are shown in Fig. 4.4. The experimental data are taken from<br />

Werner [199], and refer to a 6 µl water droplet evaporation in air at 40 ◦ C, 3.75%<br />

relative humidity (R.H.) and flow velocity of 0.3 m/s. The droplet mass continuously<br />

decreases due to water evaporation, and initially there is no significant increase in<br />

droplet temperature. When the droplet mass reduces to a negligible value (less than<br />

5% of its initial mass), temperature raises quickly to the gas temperature, and there is<br />

good agreement between the simulation and the experiment.<br />

6<br />

60<br />

Droplet mass [mg]<br />

5<br />

4<br />

3<br />

2<br />

Mass [Simulation]<br />

Mass [Experiment]<br />

T [Simulation]<br />

T [Experiment]<br />

50<br />

40<br />

30<br />

20<br />

Droplet temperature [°C]<br />

1<br />

10<br />

0<br />

0 100 200 300 400 500 600 700<br />

0<br />

800<br />

Time [s]<br />

Fig. 4.4: Comparison of simulated and measured [199] droplet mass and temperature<br />

profiles for the evaporation of a pure water droplet.


4.1. One-dimensional Evaporating Water Spray in Nitrogen 59<br />

3E+08<br />

Number density [m ­3 ]<br />

2.8E+08<br />

2.6E+08<br />

2.4E+08<br />

2.2E+08<br />

2E+08<br />

1.8E+08<br />

1.6E+08<br />

1.4E+08<br />

1.2E+08<br />

Inhomogeneous<br />

Homogeneous<br />

Experiment<br />

1E+08<br />

8E+07<br />

6E+07<br />

0.14 0.28 0.42 0.56 0.7 0.84<br />

Position [m]<br />

0 1 2 3 4 5<br />

Time [s]<br />

Fig. 4.5: Homogeneous and inhomogeneous calculations of DQMOM.<br />

Before carrying out simulations with an inhomogeneous system of DQMOM transport<br />

Eq. (3.19), homogenous formulations of these equations, given by Eq. (3.20), are<br />

simulated and the results are compared with inhomogeneous computations. Fig. 4.5<br />

shows the computed and experimental profiles of number density at different cross sections<br />

of the spray chamber with homogeneous and inhomogeneous system of DQMOM<br />

equations (see Eq. (3.19)) for 80 kg/h water spray in nitrogen flowing with 0.078 m/s<br />

velocity at 293 K. In the homogeneous results, the time axis of the model is matched<br />

to experimental position through the droplet velocity. The number density decreases<br />

along the spray axis due to droplet evaporation, and the predictions with inhomogeneous<br />

formulation captures the physics of the spray more accurately [198, 200]. Thus,<br />

the present work includes the numerical solutions of the inhomogeneous linear system,<br />

which are formed through application of DQMOM in one physical dimension (axial<br />

direction).<br />

Figure 4.6 displays the results of Sauter mean diameter showing the comparison of<br />

QMOM and DQMOM, where DQMOM results are shown for both lower and higher<br />

Reynolds number. Here, the liquid flow rate is 80 kg/h. It can be seen that the<br />

QMOM results strongly deviate from the experiment whereas DQMOM improves the<br />

results of QMOM significantly even for lower droplet Reynolds numbers, and with<br />

higher Reynolds number, which is the case in this simulations, the agreement between<br />

DQMOM and experiment is very good [191]. A general intuitive question could be<br />

”why the Sauter mean diameter increases in spite of evaporation in simulations as well<br />

as the experiment?”. The answer is elaborated with an example by consi<strong>der</strong>ing the<br />

droplet size distribution of a water spray shown in Fig. 4.7, which displays the droplet


60 4. Results and Discussion<br />

Sauter mean diameter [µm]<br />

140<br />

135<br />

130<br />

125<br />

120<br />

115<br />

110<br />

DQMOM (Re>1)<br />

DQMOM (Re


4.1. One-dimensional Evaporating Water Spray in Nitrogen 61<br />

Tab. 4.2: Droplet size distribution of water spray<br />

d(t = 0 s) [µm] f [(µm) −1 ] d(t = 1 s) [µm] d(t = 50 s) [µm] ]<br />

10.643 0.409 0.000 0.000<br />

12.913 4.091 0.000 0.000<br />

15.588 11.434 0.000 0.000<br />

18.745 11.762 0.000 0.000<br />

22.468 8.472 0.000 0.000<br />

26.858 7.112 0.000 0.000<br />

32.037 7.093 5.138 0.000<br />

38.145 7.481 21.332 0.000<br />

45.350 7.304 32.506 0.000<br />

53.848 6.609 43.584 0.000<br />

63.870 5.439 55.492 0.000<br />

75.692 4.491 68.770 0.000<br />

89.635 4.073 83.872 0.000<br />

106.080 4.319 101.257 0.000<br />

125.478 4.574 121.427 0.000<br />

148.355 3.579 144.946 0.000<br />

175.337 1.490 172.462 0.000<br />

207.163 0.255 204.735 0.000<br />

244.697 0.015 242.646 99.381<br />

The k value is usually estimated from the material properties such as density, diffusivity,<br />

etc., and in general, it has a value in the range of 10 −7 to 10 −11 . Just for the sake of<br />

explanation, k is assumed to be 1.0E-09 m 2 /s. The change in the droplet diameter is<br />

computed at t = 1 s using d 2 law is given in Tab. 4.2, see the third column. Comparing<br />

the values of droplet diameter at t = 0 s and t = 1 s, it shows that the droplet diameter<br />

decreases and lower size droplets vanish. Using these data, the computed d 32 at t =<br />

1 s is 117.126 µm, which shows an increase from initial value. This increase continues<br />

till certain evaporation time (see last column in Tab. 4.2), whereupon the Sauter mean<br />

diameter starts to decrease because most of the smaller size droplets vanish and only<br />

few droplets have finite size. The Sauter mean diameter of this distribution decreases<br />

to 99.381 µm after 50 s of d 2 law evaporation rate (see the last column in Tab. 4.2).<br />

Figure 4.8 shows the plots of droplet velocities subjected to only drag force (left),<br />

and drag force with gravity (right) for three different sized droplets, respectively. In<br />

case of only drag caused by the surrounding gas with initial velocity of 0.078 m/s (experimental<br />

value), the velocity decreases at first due to drag force and later the droplets


62 4. Results and Discussion<br />

3<br />

3<br />

Veloctiy [m/s]<br />

2.5<br />

2<br />

1.5<br />

1<br />

u[1] = 1.09 m/s, r[1] = 24 µm<br />

u[2] = 1.94 m/s, r[2] = 86 µm<br />

u[3] = 2.76 m/s, r[3] = 143 µm<br />

Velocity [m/s]<br />

2.5<br />

2<br />

1.5<br />

1<br />

u[1] = 1.09 m/s, r[1] = 24 µm<br />

u[2] = 1.94 m/s, r[2] = 86 µm<br />

u[3] = 2.76 m/s, r[3] = 143 µm<br />

0.5<br />

0.5<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Time [s]<br />

0<br />

0 1 2 3 4 5<br />

Time [s]<br />

Fig. 4.8: Velocity profiles of three droplets with different initial radii and velocities<br />

un<strong>der</strong> the influence of drag alone (left) and drag and gravity (right).<br />

follow the streamlines of the gas after reaching a steady value, cf. left side of Fig. 4.8.<br />

On the other hand, when the droplets encounter gravity in addition to drag force applied<br />

by the surrounding gas, the droplet velocity initially decreases due to drag and<br />

then increases linearly due to gravity as seen in right part of Fig. 4.8. In a previous<br />

study [198, 200], it has been shown that the mo<strong>der</strong>ate droplet evaporation un<strong>der</strong> the<br />

present conditions does not significantly influence droplet velocity.<br />

140<br />

130<br />

Experiment - 80 kg/h<br />

Evaporation - 80 kg/h<br />

Evaporation and Coalescence - 80 kg/h<br />

Experiment - 150 kg/h<br />

Evaporation - 150 kg/h<br />

Evaporation and Coalescence - 150 kg/h<br />

Sauter mean diameter [µm]<br />

120<br />

110<br />

100<br />

90<br />

80<br />

70<br />

0.14 0.28 0.42 0.56 0.7 0.84<br />

Position [m]<br />

Fig. 4.9: Effect of liquid inflow rates on Sauter mean diameter computed with and<br />

without coalescence.


4.1. One-dimensional Evaporating Water Spray in Nitrogen 63<br />

The variations in Sauter mean diameter with axial position of the spray for two different<br />

liquid inflow rates of 80 kg/h and 150 kg/h are shown in Fig. 4.9. The results for<br />

80 kg/h show an increasing Sauter mean diameter with evaporation. Inclusion of coalescence<br />

in addition to evaporation leads to excellent agreement between computational<br />

and experimental results. On the contrary, the computational results for 150 kg/h at<br />

x = 0.54 m seem to be deviating far away from the experimental data. The observed<br />

deviation is due to inconsistency in experimental data, which is evident from the fact<br />

that the experimental flow rate does not match the prescribed value of 150 kg/h at<br />

0.54 m. Therefore, the results from 80 kg/h will be discussed for the remaining part of<br />

this section.<br />

Figure 4.10 displays the profiles of the Sauter mean diameter (left) and mean droplet<br />

diameter (right) of water spray subjected to evaporation at 293 K and 313 K temperatures<br />

of surrounding gas as well as with and without coalescence. As expected, Sauter<br />

mean diameter increases substantially with evaporation that causes the decrease and<br />

eventual loss of small size droplets. Higher temperature imposes a rise in evaporation,<br />

which consi<strong>der</strong>ably accelerates the rate of increase of Sauter mean diameter. A comparison<br />

with experimental data reveals the importance of modeling the droplet coalescence,<br />

which not only improves the simulation results but also has excellent agreement with<br />

experiment (see left side of Fig. 4.10).<br />

Similar to Sauter mean diameter, the mean droplet diameter is an important physical<br />

quantity for several applications such as particle size analysis of pow<strong>der</strong> sampling<br />

in food and pharmaceutical industries [201]. Mean droplet diameter of a number density<br />

based distribution can be computed using the Eq. (2.18). Since very small size<br />

160<br />

150<br />

Experiment - 293 K<br />

Evaporation - 293 K<br />

Evaporation and Coalescence - 293 K<br />

Evaporation - 313 K<br />

Evaporation and Coalescence - 313 K<br />

90<br />

85<br />

Experiment - 293 K<br />

Evaporation - 293 K<br />

Evaporation and Coalescence - 293 K<br />

Evaporation - 313 K<br />

Evaporation and Coalescence - 313 K<br />

Sauter mean diameter [µm]<br />

140<br />

130<br />

120<br />

110<br />

D 10<br />

[µm]<br />

80<br />

75<br />

70<br />

65<br />

60<br />

55<br />

100<br />

0.14 0.28 0.42 0.56 0.7 0.84<br />

Position [m]<br />

50<br />

0.14 0.28 0.42 0.56 0.7 0.84<br />

Position [m]<br />

Fig. 4.10: Profiles of Sauter mean diameter (left) and mean droplet diameter (right)<br />

computed with and without coalescence at surrounding gas temperatures of<br />

293 K and 313 K.


64 4. Results and Discussion<br />

droplets may completely evaporate, leading to decreased total droplet number at any<br />

cross section, cf. Fig. 4.12, the mean value of droplet diameters increases. Therefore,<br />

as the droplets move and start to evaporate and vanish completely, the mean droplet<br />

diameter of the spray starts increasing at cross sections away from the nozzle although<br />

individual droplet diameters decrease (see right part of Fig. 4.10). This observation<br />

is in agreement with the behavior of the Sauter mean diameter shown in left side of<br />

Fig. 4.10. Coalescence causes an increase of droplet diameter as anticipated.<br />

Figure 4.11 shows the results of the droplet specific surface area. The specific<br />

surface area is an important parameter, which is used particularly to characterize<br />

pow<strong>der</strong> materials, and it is defined as the ratio of total surface area of the individual<br />

droplets/particles to the total volume [201]. It can be seen that the specific surface<br />

area decreases as a result of evaporation, which leads to decrease in number density<br />

of droplets. This is evident from the behavior of specific surface area at a higher<br />

temperature. A comparison with experimental data confirms the role of coalescence in<br />

improving the results as observed in case of the Sauter mean diameter and the mean<br />

droplet diameter displayed in Fig. 4.10.<br />

Figure 4.12 shows the plots of total droplet number density in axial direction. Since<br />

the geometric configuration consi<strong>der</strong>ed for the numerical solution is one-dimensional,<br />

the integral value of droplet number density over the corresponding cross sections<br />

is displayed. It can be seen that evaporation causes the droplet number density to<br />

decrease as the spray develops. This decrease is much pronounced at the higher temperature<br />

(313 K) due to enhanced evaporation. It is worthwhile to note that inclusion<br />

6.0E+04<br />

Experiment - 293 K<br />

Evaporation - 293 K<br />

Evaporation and Coalescence - 293 K<br />

Evaporation - 313 K<br />

Evaporation and Coalescence - 313 K<br />

Specific surface area [m 2 /m 3 ]<br />

5.5E+04<br />

5.0E+04<br />

4.5E+04<br />

4.0E+04<br />

3.5E+04<br />

0.14 0.28 0.42 0.56 0.7 0.84<br />

Position [m]<br />

Fig. 4.11: Profiles of specific surface area computed with and without coalescence at<br />

surrounding gas temperatures of 293 K and 313 K.


4.1. One-dimensional Evaporating Water Spray in Nitrogen 65<br />

3E+08<br />

2.5E+08<br />

Experiment - 293 K<br />

Evaporation - 293 K<br />

Evaporation and Coalescence - 293 K<br />

Evaporation - 313 K<br />

Evaporation and Coalescence - 313 K<br />

Number density [m -3 ]<br />

2E+08<br />

1.5E+08<br />

1E+08<br />

5E+07<br />

0.14 0.28 0.42 0.56 0.7 0.84<br />

Position [m]<br />

Fig. 4.12: Profiles of droplet number density computed with and without coalescence<br />

at surrounding gas temperatures of 293 K and 313 K.<br />

of coalescence affects the calculation of droplet number density significantly as it can<br />

be inferred through comparison of the numerical results with experimental data. This<br />

may be un<strong>der</strong>stood by the fact that only coalescence is consi<strong>der</strong>ed in the present work<br />

and processes of breakup, reflexive and stretching separation along with formation of<br />

satellite droplets is neglected in the present simulations, which leads to a lower droplet<br />

number density at any given position. This may be improved by including a more<br />

advanced droplet–droplet interaction model [184]. Moreover, in these computations<br />

the evaporating flux at zero droplet size is computed through the ratio constraints of<br />

weights, radii and velocities given by Eqs. (2.33)– (2.34), which are <strong>der</strong>ived based on<br />

a smooth and continuous density function [60]. This approach is prone to errors and<br />

may be rectified by implementing an maximum entropy model [66] explained in Section<br />

2.3.4, which is done in the case of two-dimensional water and PVP/water in air<br />

spray flows.<br />

The successful implementation of DQMOM in studying the one-dimensional water<br />

spray flow in nitrogen and the good agreement with experimental data has led to the<br />

extension of DQMOM to two dimensions in or<strong>der</strong> to model the evaporating water<br />

spray in air in two-dimensional configuration. The DQMOM extension is outlined in<br />

Section 2.3.4. The next section presents the results of two-dimensional water spray in<br />

air in axisymmetric configuration.


66 4. Results and Discussion<br />

4.2 Two-dimensional Evaporating Water Spray in<br />

Air<br />

A water spray injected into air through a hollow cone Delavan SDX-SE-90 nozzle in<br />

a vertical spray chamber, is modeled by DQMOM and DDM. The one-dimensional<br />

transport equations of DQMOM [191] are extended to two-dimensional to model the<br />

spray flow in axisymmetric configuration [202, 203]. The starting data for the simulations<br />

are taken from experimental data, where the experiments are conducted by the<br />

group of Prof. G. Brenn at TU Graz, Austria. The experimental setup is explained in<br />

the next section. The generation of initial data is discussed in the following section.<br />

The simulation results of DQMOM are compared with the results of DDM and both<br />

these model results are validated with the experiment [202, 203].<br />

4.2.1 Experimental Setup<br />

A series of experiments is carried out at TU Graz by the group of Prof. G. Brenn where<br />

a water spray in air is studied for different liquid mass inflow rates. The droplet sizes<br />

and velocities are recorded at various cross sections for different liquid inflow rates<br />

using phase Doppler anemometry (PDA) [204]. The present simulations concern the<br />

experimental data generated using a Delavan nozzle SDX-SE-90 having an internal<br />

diameter of 0.002 m, an outer diameter of 0.012 m at the nozzle throat and 0.016 m<br />

at the top, for liquid inflow rates of 80 kg/h and 120 kg/h. A water spray is injected<br />

into a cylindrical spray chamber of diameter 1 m. The carrier gas is air at room<br />

temperature and atmospheric pressure. Measurements are recorded at cross sections of<br />

0.08 m, 0.12 m and 0.16 m. Figure 4.13 illustrates the schematic of the experimental<br />

Fig. 4.13: Schematic diagram of the experimental setup.


4.2. Two-dimensional Evaporating Water Spray in Air 67<br />

setup. The data at 0.08 m are taken as starting point for initial data generation for<br />

computations, and results are compared at later cross sections [205].<br />

4.2.2 Initial Data Generation<br />

The experimental data at the closest position to the nozzle is used to generate initial<br />

data for the numerical computations of DQMOM. The nearest experimental position<br />

is 0.08 m from the nozzle, where the measurements are available at radial positions<br />

separated by 1.5 × 10 −3 m distance. The PDA data at every radial position consists<br />

of droplet radius, velocities in axial and radial directions, and the time elapsed for<br />

each measurement, which gives the total time carried out over a period. These data<br />

are grouped into 100 droplet size classes [206]. The effective cross sectional area of<br />

the probe volume is computed, which is done to eliminate errors in measuring volume<br />

due to nonlinearity in phase/diameter relationship in large size droplets because of the<br />

nonuniform beam intensity [207]. The result of the calculation for a water flow rate of<br />

80 kg/h, at a position of 0.066 m from the center is shown in Fig. 4.14. The trajectory<br />

length exhibits strong fluctuations, and fluctuations increase with the droplet size.<br />

Furthermore, the number of droplets in the size classes for the larger diameters is<br />

typically much lower than in the smaller size classes. Therefore, the properties such<br />

as droplet trajectory lengths through the probe volume are statistically unreliable for<br />

drops with sizes greater than a certain threshold value [206, 207]. In particular, the<br />

decrease of the effective probe volume size with increasing droplet size such as from<br />

0.45<br />

Effective cross­section area [mm 2 ]<br />

0.4<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

x<br />

x<br />

x x x xx x x<br />

xxx<br />

x x x<br />

x x x x<br />

Experiment<br />

Corrected<br />

Trend line<br />

x<br />

x<br />

x x<br />

x<br />

x x xxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxxx<br />

x x x<br />

x<br />

x<br />

x x<br />

x<br />

x x x x x<br />

x x<br />

0<br />

0 100 200 300 400<br />

Drop size class [µm]<br />

Fig. 4.14: Profile of effective cross-section area of the probe volume for measured droplet<br />

size.


68 4. Results and Discussion<br />

200 µm as shown in Fig. 4.14 is invalid as the effective cross-sectional area should<br />

increase with droplet size [207]. The effective cross-section area is therefore calculated<br />

using a linear trend line from a threshold diameter. In the first step, the linear trend<br />

line is calculated using a linear regression scheme based on the data in the droplet size<br />

classes up to 60% of the maximum droplet size.<br />

In the second step, for all droplet size classes larger than 40% of the maximum<br />

droplet size class for this experimental position, the values of the effective cross sectional<br />

area are obtained as values of the linear trend line. Therefore, there is an<br />

overlap of the size class ranges used for computing the trend line and those whose<br />

probe volume cross-section areas are calculated using the trend line. Once the effective<br />

cross-sectional area probe volume is corrected, the number density is corrected<br />

correspondingly. Then, the moment sets of droplet size and velocities are computed,<br />

which in turn are used to calculate the initial weights (number densities), radii and<br />

velocities using the Wheeler algorithm [136]. In the present study, the spray distribution<br />

is approximated by a three-node closure, which is proven to be accurate in<br />

previous studies [48, 49, 191, 202]. The three-node approximation of NDF implies that<br />

the required number of moments is 12 (3 each: weights, droplet radii, axial velocities,<br />

radial velocities). The same procedure is followed at every radial position for the crosssection<br />

of 0.08 m. Figure 4.15 shows the experimental distribution of droplets and<br />

DQMOM approximation at 0.066 m from the center of the spray for 80 kg/h water<br />

flow rate. The problem of negative moments is handled by employing the adaptive<br />

Wheeler algorithm [208].<br />

Number density [ (µm) ­1 ]<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

Experiment<br />

DQMOM<br />

0<br />

0 50 100 150<br />

Droplet radius [ µm ]<br />

Fig. 4.15: Experimental and DQMOM approximation of droplet number density for a<br />

water spray.


4.2. Two-dimensional Evaporating Water Spray in Air 69<br />

The DDM simulation are carried out only for water spray in two-dimensional configuration<br />

by Humza [68]. In his work, the same experimental data is used to generate<br />

a system of parcels for DDM, where the properties of the k th parcel are denoted by<br />

(x k , r k , u k , v k , m k ) for the present two-dimensional configuration. The liquid mass of<br />

k th parcel is computed assuming the spherical symmetry of the droplets, i.e.,<br />

m k =<br />

N∑<br />

i=1<br />

4<br />

3 πρ lr 3 i , (4.3)<br />

where N refers to the number of droplets in the parcel. The number of parcels for the<br />

inflow rate of 80 kg/h is 3,704 and for 120 kg/h, it is 3,464. A non-equidistant rectangular<br />

grid with 7,878 grid points (78 in radial and 101 in axial direction, respectively)<br />

is used [68].<br />

4.2.3 Results and Discussion<br />

At first the implementation of maximum entropy (ME) method for the calculation of<br />

evaporative flux is studied. Fig. 4.16 shows the computed NDF at radial distance of<br />

64.5 mm from the center and 0.08 m axial distance from nozzle for 80 kg/h liquid flow<br />

rate using ME method and its comparison with experiment, where a good agreement<br />

between the ME approximated NDF and experiment can be found. The evaporative<br />

flux computed using the weight ratio constraints, which are defined by Eqs. (2.33)–<br />

(2.34), for this position is found to be 0.39, where as with ME method it is 0.022 and<br />

0.05<br />

0.04<br />

Flow rate: 80 kg/h<br />

Axial position: 80 mm<br />

Radial position: 64.5 mm<br />

ME reconstruction<br />

Experiment<br />

NDF [(µm) ­1 ]<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

0 25 50 75 100 125 150 175<br />

Droplet radius [µm]<br />

Fig. 4.16: Experimental and reconstructed NDF of 80 kg/h water spray at 64.5 mm<br />

from the center of the spray axis.


70 4. Results and Discussion<br />

0.03<br />

Flow rate: 121 kg/h<br />

Axial position: 80 mm<br />

Radial position: 81 mm<br />

NDF [(µm) -1 ]<br />

0.02<br />

0.01<br />

ME reconstruction<br />

Experiment<br />

0<br />

0 25 50 75 100 125 150 175<br />

Droplet radius [µm]<br />

Fig. 4.17: Experimental and reconstructed NDF of 121 kg/h water spray at 81 mm from<br />

the center of the spray axis.<br />

in experiment the same is about 0.03 (see Fig. 4.16). Similar observation is made with<br />

higher liquid flow rate as well, Fig. 4.17 shows the experimental and reconstructed<br />

NDF for water spray of 121 kg/h at 81 mm radial position from the center and 0.08 m<br />

away from the nozzle. Thus, the ME approach improves the DQMOM evaporative flux<br />

calculation procedure and it has excellent agreement with the experiment. Therefore,<br />

in the current study, the ME method is used for the ψ calculation in two-dimensional<br />

evaporating water and PVP/water spray flows.<br />

In numerical simulation of water spray in two-dimensional configuration, average<br />

droplet properties such as mean droplet diameter, Sauter mean diameter and mean<br />

droplet velocity are computed using both the methods, i.e., DQMOM and DDM, and<br />

the simulation results are compared with the experiment at the cross sections of 0.12 m<br />

and 0.16 m away from the nozzle exit. Figure 4.18 shows the computed and experimental<br />

profiles of the Sauter mean diameter at cross sections of 0.12 m (left) and<br />

0.16 m (right) downstream to the nozzle orifice for 80 kg/h. The DDM simulation<br />

results match quite well with the experiment at the center of the spray at 0.12 m<br />

away from the nozzle exit, but slightly un<strong>der</strong>predicts towards the periphery of the<br />

spray whereas good agreement is observed at 0.16 m cross section between DDM and<br />

experiment.<br />

The DQMOM simulation results are in good agreement with experiment at 0.12 m<br />

downstream the nozzle exit, and it is closer to the experimental data at higher radial<br />

distance as well. Further downstream, at 0.16 m from the nozzle orifice (see right part<br />

of the Fig. 4.18), the DQMOM simulations reveal some scattering near the centerline,


4.2. Two-dimensional Evaporating Water Spray in Air 71<br />

180<br />

180<br />

Sauter mean diameter [µm]<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

DQMOM<br />

DDM<br />

Experiment<br />

Sauter mean diameter [µm]<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

DQMOM<br />

DDM<br />

Experiment<br />

20<br />

20<br />

0<br />

-100 -50 0 50 100<br />

Radial Position [mm]<br />

0<br />

-100 -50 0 50 100<br />

Radial position [mm]<br />

Fig. 4.18: Experimental and numerical profiles of the Sauter mean diameter of water<br />

spray with 80 kg/h liquid flow rate at the cross section of 0.12 m (left) and<br />

0.16 m (right) distance from the nozzle exit.<br />

while at higher radial distances, they un<strong>der</strong>predict the experimental results. This<br />

discrepancy may be the result of numerical scheme, which employs an explicit finite<br />

difference method to solve the transport equations of DQMOM; the results can be<br />

improved by implementing an implicit method. The post-processing of experimental<br />

data, which is explained in Subsection 4.2.2, may be the reason of the deviation, too.<br />

For an elevated liquid inflow rate of 120 kg/h, the computed and experimental<br />

180<br />

160<br />

DQMOM<br />

DDM<br />

Experiment<br />

Sauter mean diameter [µm]<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

-120 -80 -40 0 40 80 120<br />

Radial position [mm]<br />

Fig. 4.19: Experimental and numerical profiles of the Sauter mean diameter at the cross<br />

section of 0.12 m distance from the nozzle exit for 120 kg/h.


72 4. Results and Discussion<br />

160<br />

140<br />

DQMOM<br />

DDM<br />

Experiment<br />

Sauter mean diameter [µm]<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

-120 -80 -40 0 40 80 120<br />

Radial position [mm]<br />

Fig. 4.20: Experimental and numerical profiles of the Sauter mean diameter at the cross<br />

section of 0.16 m distance from the nozzle exit for 120 kg/h.<br />

profiles of Sauter mean diameter at cross sections of 0.12 m and 0.16 m away from<br />

the nozzle exit are shown in Figs. 4.19 and 4.20. An increased liquid flow rate causes<br />

a somewhat decreased droplet size: For a given liquid, increased mass flow rate leads<br />

to higher pressure drop in the atomizer, which decreases liquid sheet size and breakup<br />

length to yield smaller particles as can be seen when compared with Fig. 4.18.<br />

At the cross section of 0.12 m, it can be seen that DQMOM performs better than the<br />

DDM results as DDM overpredicts the experimental values. The scattering behavior<br />

of DQMOM simulation results near the centerline is observed in this case, too. As the<br />

droplets move to the next cross section, a decrease in large size droplets is evident,<br />

which is predicted by both DQMOM and DDM. The results show that the DQMOM<br />

shows better agreement with experiment, while DDM predicts somewhat higher values<br />

than the experiment at corresponding radial positions [205].<br />

The overall shape of a hollow cone spray is captured quite nicely by both methods,<br />

although some deviations are observed, particularly in DQMOM as compared to<br />

experimental profile. This is possibly due to the post-processing of the experimental<br />

data as explained in Subsection 4.2.2, which is done to correct the number frequency<br />

at every measuring position to rule out the fluctuations in the effective cross sectional<br />

area of the measuring volume for the larger droplet sizes [207]. This correction of experimental<br />

data is position dependent, whereas DQMOM and DDM results account<br />

for these corrections for the initial condition but not at positions further downstream.<br />

Another reason for the discrepancies in the DQMOM results may be due to the fact<br />

that the spray equations are not yet fully coupled to the gas phase.


4.2. Two-dimensional Evaporating Water Spray in Air 73<br />

Comparing the maximum values of the Sauter mean diameter at the two cross<br />

sections displayed in Figs. 4.18, 4.19 and 4.20, a decrease in large size droplets is<br />

observed as the droplets move away from the nozzle. Even though the process of<br />

evaporation is consi<strong>der</strong>ed in the present models, the major reason for the decrease in<br />

droplet size may be attributed to the influence of drag force applied by the surrounding<br />

gas, because significant evaporation may not occur at the present room temperature<br />

condition. This decrease is more evident in the large droplet size region, where the<br />

dynamic interaction of droplet with surrounding gas dominates, as observed in profiles<br />

of mean droplet velocity (see Figs. 4.23 and 4.24).<br />

Besides the Sauter mean diameter, in many technical applications such as particle<br />

size analysis of pow<strong>der</strong> sampling in food and pharmaceutical industries, the mean<br />

droplet diameter is an important physical quantity [201]. Radial profiles of the mean<br />

droplet diameter compared with experiment are shown in Fig. 4.21 for 80 kg/h at<br />

0.12 m (left) and 0.16 m (right) distance away from nozzle. DDM results are in very<br />

good agreement with the experiment. A slight decrease in the mean droplet diameter<br />

is observed as the droplets move away from nozzle indicating some mass transfer from<br />

liquid to gas, which is attributable to gas–liquid interactions. The DQMOM results<br />

are in excellent agreement with experiment at the cross section of 0.12 m near the<br />

centerline, and DQMOM results show better agreement than DDM results (see left part<br />

of Fig. 4.21). At 75 mm radial position, the DQMOM results are below experimental<br />

values, which may stem from the explicit finite difference technique. At the cross<br />

section of 0.16 m, a good agreement is observed between DQMOM and experiment<br />

near the axis of symmetry, even though some scattering behavior is found (see right<br />

100<br />

100<br />

Mean droplet diameter [µm]<br />

80<br />

60<br />

40<br />

20<br />

DQMOM<br />

DDM<br />

Experiment<br />

Mean droplet diameter [µm]<br />

80<br />

60<br />

40<br />

20<br />

DQMOM<br />

DDM<br />

Experiment<br />

0<br />

-100 -50 0 50 100<br />

Radial position [mm]<br />

0<br />

-100 -50 0 50 100<br />

Radial Position [mm]<br />

Fig. 4.21: Experimental and numerical profiles of the mean droplet diameter of water<br />

spray with 80 kg/h liquid flow rate at the cross section of 0.12 m (left) and<br />

0.16 m (right) distance from the nozzle exit.


74 4. Results and Discussion<br />

120<br />

DQMOM<br />

DDM<br />

Experiment<br />

80<br />

DQMOM<br />

DDM<br />

Experiment<br />

100<br />

Mean droplet diameter [µm]<br />

80<br />

60<br />

40<br />

Mean droplet diameter [µm]<br />

60<br />

40<br />

20<br />

20<br />

0<br />

-120 -80 -40 0 40 80 120<br />

Radial position [mm]<br />

0<br />

-120 -80 -40 0 40 80 120<br />

Radial position [mm]<br />

Fig. 4.22: Experimental and numerical profiles of the mean droplet diameter of water<br />

spray with 120 kg/h liquid flow rate at the cross section of 0.12 m (left) and<br />

0.16 m (right) distance from the nozzle exit.<br />

side of Fig. 4.21).<br />

In Fig. 4.21, deviations from the experiment occur in the large droplet size region,<br />

which is due to the fact that the numerical technique captures the distribution function<br />

globally, and there could be some local discrepancies as well. This may be improved<br />

by solving the gas phase equations for DQMOM, which is not yet done in the present<br />

study, where the inlet gas flow properties are used to calculate the source terms for<br />

transport equations for DQMOM [209].<br />

Figure 4.22 shows the computed and experimental profiles of the mean droplet<br />

diameter at cross sections of 0.12 m (left) and 0.16 m (right) away from the nozzle exit<br />

for liquid inflow rate of 120 kg/h. Similar to Sauter mean diameter, elevated liquid<br />

flow rate leads to somewhat decreased droplet size (compare Fig. 4.22 with 4.21) . At<br />

0.12 m away from the nozzle exit, both DDM and DQMOM agree well with each other<br />

near the centerline, where they show relatively higher values than the experiment. At<br />

the radial positions away from the centerline, DQMOM is in good agreement with the<br />

experiment, and it is better than the DDM results. As the droplets move away from<br />

the nozzle exit, a decrease in size can be observed at the cross section of 0.16 m away<br />

from the the nozzle exit (see right part of the Fig. 4.22), which is similar to the case<br />

of liquid flow rate of 80 kg/h. Near the centerline at 0.16 m away from the nozzle<br />

exit, both DQMOM and DDM show the same behavior and predict slightly higher<br />

values than experiment. At higher radial positions, DDM values are higher compared<br />

to DQMOM and experiment, whereas DQMOM coincides with the experimental data.<br />

In Figs. 4.23 and 4.24, the radial profiles of mean droplet velocity are displayed<br />

at different cross sections. It can be seen that the droplet velocity is higher for larger


4.2. Two-dimensional Evaporating Water Spray in Air 75<br />

Mean droplet velocity [m/s]<br />

6<br />

5<br />

4<br />

3<br />

DQMOM<br />

DDM<br />

Experiment<br />

2<br />

-100 -50 0 50 100<br />

Radial position [mm]<br />

Fig. 4.23: Experimental and numerical profiles of the mean droplet velocity at the cross<br />

section of 0.12 m distance from the nozzle exit.<br />

droplets as anticipated. Interestingly, the small size droplets near the axis of symmetry<br />

also move at a higher velocity as observed in the experiment and thus causing the<br />

velocity profile bimodal, which is predicted quite nicely by both models.<br />

A closer look reveals that the width of the jet is captured by the DQMOM, whereas<br />

the DDM predicts somewhat broa<strong>der</strong> profiles with a lower maximum value at the centerline.<br />

At the spray edge, a judgement of the numerical methods is difficult, since the<br />

Mean droplet velocity [m/s]<br />

6<br />

5<br />

4<br />

3<br />

DQMOM<br />

DDM<br />

Experiment<br />

2<br />

-100 -50 0 50 100<br />

Radial position [mm]<br />

Fig. 4.24: Experimental and numerical profiles of the mean droplet velocity at the cross<br />

section of 0.16 m distance from the nozzle exit.


76 4. Results and Discussion<br />

experimental data are somewhat spread at 0.12 m from the nozzle exit. At 0.16 m, the<br />

slopes of the numerical results deviate from the experimental data, particularly in large<br />

size droplets region where the effective cross sectional area shows strong fluctuations<br />

in experiment as shown in Fig. 4.14. This implies that the post-processing of experimental<br />

data plays an important role in the corrections of number density and thereby<br />

the droplet properties [207]. Comparing the velocity profiles at the two different cross<br />

sections, it is seen that the velocity decreases as droplets move away from the nozzle.<br />

This is because the droplets are strongly decelerated by the dynamic interaction with<br />

the surrounding gas. The gas around the spray stagnates and is driven into motion<br />

only due to the spray entrainment. The gas motion driven by the spray arises at the<br />

expense that the droplet loses momentum.<br />

The droplet properties are predicted quite well by the present simulations, which<br />

confirms their applicability for spray flows. There are some deviations between simulation<br />

and experimental results, which are attributable to the post-processing of the<br />

experimental data as mentioned before. In case of DDM, neglecting droplet–droplet<br />

interactions may need reconsi<strong>der</strong>ation. For DQMOM, the improved numerical scheme<br />

and the simultaneous solution of the gas phase equations may improve the simulation<br />

results.<br />

Based on these simulation results and comparison with the experiment, it can be<br />

concluded that the DQMOM is a robust method, which can predict the spray flows<br />

accurately. This led to the implementation of DQMOM to study bi-component evaporating<br />

spray, i.e., PVP/water spray flow in two dimensions. In or<strong>der</strong> to perform<br />

simulations of PVP/water spray flows, the predictability and efficiency of developed bicomponent<br />

droplet evaporation and solid layer formation model (see Subsection 2.4.1.2)<br />

needs to be verified un<strong>der</strong> different drying conditions. The next section presents the<br />

numerical simulation of single bi-component droplet evaporation and solid layer development,<br />

and comparison of simulation results with experiment.<br />

4.3 Single Bi-component Droplet Evaporation and<br />

Solid Layer Formation<br />

The model presented in Subsection 2.4.1.2 to predict the evaporation and solid layer<br />

formation for PVP/water droplet and mannitol/water droplet is simulated with different<br />

conditions such as initial solute mass fraction, gas temperature and velocity,<br />

relative humidity, initial droplet size etc. In the next subsections, the vapor-liquid<br />

equilibrium calculation followed by non-ideality effect caused by the solute (PVP or<br />

mannitol) presence on the droplet heating and evaporation rate is explained. Finally,<br />

the single droplet evaporation and solid layer development results are presented.


4.3. Single Bi-component Droplet Evaporation and Solid Layer Formation 77<br />

4.3.1 Vapor-Liquid Equilibrium<br />

The vapor-liquid equilibrium for the evaporating component i is needed for the calculation<br />

of Spalding’s mass transfer number, B M,1 , cf. Eq. (2.54), for both PVP/water<br />

and mannitol/water droplet evaporation and solid layer development cases, in which<br />

the mass fraction, Y 1,s , of the evaporating component appears; this mass fraction is<br />

calculated through the mole fraction, X i , i = 1, of the evaporating component water,<br />

in terms of the activity coefficient<br />

X i = p vap,i<br />

p m<br />

γ i X L,i , (4.4)<br />

where p vap,1 is the vapor pressure of pure water and p m is the total mixture pressure,<br />

which is equal to the ambient gas pressure, and in the present study it equals the<br />

atmospheric pressure. Here, X L,i is the mole fraction of evaporating component i in<br />

liquid phase, and γ i is the activity coefficient of evaporating component i, which is<br />

calculated through equation given as,<br />

γ i = a wY L,i<br />

X L,i<br />

. (4.5)<br />

Here a w is the water activity, Y L,i with i = 1, is the water mass fraction within the<br />

droplet. The calculation of water activity coefficient is described in next subsection.<br />

4.3.2 Non-ideal Liquid Mixture<br />

The presence of polymer or mannitol with water leads to non-ideal liquid behavior,<br />

which must be accounted for in calculating the mole fraction of water vapor at the<br />

droplet surface. In this work, the liquid mixture is treated as non-ideal by determining<br />

the influence of individual components on each other through their activity coefficients.<br />

The universal functional activity coefficient (UNIFAC) method is the accurate and<br />

most extensively used procedure [210], which estimates the activity coefficient as a<br />

sum of combinatorial and residual terms. This method, however, cannot be applied for<br />

polymer solutions as they have significant difference in accessible volume for a molecule<br />

in the solution [211].<br />

The work of Oishi and Prausnitz [211] extended UNIFAC method to account for<br />

such differences in accessible volume by introducing a free-volume term, which enabled<br />

the UNIFAC approach to be applied to polymer solution systems. However, it is<br />

proven that their model fails for aqueous polymer systems because of the inadequacy<br />

of its free-volume term [212]. In the current study, the activity coefficient of water<br />

in PVP/water solution is computed using the UNIFAC-van <strong>der</strong> Waals-Free Volume<br />

method known as UNIFAC-vdW-FV method [212], which accounts for the free-volume


78 4. Results and Discussion<br />

1<br />

1<br />

0.8<br />

0.8<br />

Water activity [­]<br />

0.6<br />

0.4<br />

Temperature = 73.0°C<br />

Water activity [-]<br />

0.6<br />

0.4<br />

Temperature = 94.5°C<br />

0.2<br />

UNIFAC<br />

UNIFAC­vdW­FV<br />

Experiment<br />

0.2<br />

UNIFAC<br />

UNIFAC-vdW-FV<br />

Experiment<br />

0<br />

0 0.05 0.1 0.15 0.2 0.25<br />

Water mass fraction [­]<br />

0<br />

0 0.05 0.1 0.15 0.2 0.25<br />

Water mass fraction [-]<br />

Fig. 4.25: Numerical and experimental [214] results of water activity (a w ) in PVP/water<br />

solution at 73.0 ◦ C (left) and 94.5 ◦ C (right).<br />

effect in aqueous polymer solutions. In case of mannitol/water droplet evaporation<br />

study, the activity coefficient of water is calculated using the analytical solution of<br />

groups (ASOG) contribution method [213], as it is proven to perform better than the<br />

UNIFAC method [213].<br />

Before implementation of the UNIFAC-vdW-FV method into the current PVP/water<br />

droplet code, it has been verified by comparing the water activity (a w ) computed using<br />

the UNIFAC method [210]. Results from these two methods are compared with<br />

1<br />

0.96<br />

T = 94.5°C<br />

T = 160°C<br />

Water activity [­]<br />

0.92<br />

0.88<br />

0.84<br />

0 0.1 0.2 0.3 0.4 0.5 0.6<br />

Mannitol mass fraction [­]<br />

Fig. 4.26: Numerical results of water activity (a w ) in mannitol/water solution at 94.5 ◦ C<br />

and 160 ◦ C.


4.3. Single Bi-component Droplet Evaporation and Solid Layer Formation 79<br />

experimental results [214]. Molecular properties data such as van <strong>der</strong> Waals volume<br />

and radii for PVP are taken from Bondi [215], Danner and High [216], and the interaction<br />

parameters of individual molecules required in the UNIFAC-vdW-FV method<br />

are taken from Daubert and Danner [217]. In Fig. 4.25, variation of the weight based<br />

water activity with water mass fraction in PVP/water solution is exemplarily shown<br />

at a temperature of 73.0 ◦ C (left) and 94.5 ◦ C (right), respectively.<br />

The results reveal that the UNIFAC-vdW-FV method improves the UNIFAC method<br />

results, and the UNIFAC-vdW-FV predictions are in excellent agreement with the experimental<br />

data. Therefore, in the current study, the UNIFAC-vdW-FV method is<br />

implemented to compute water activity in PVP/water solution.<br />

The change in water activity with mannitol mass fraction in mannitol/water solutions<br />

at a temperature of 94.5 ◦ C and at 160 ◦ C computed using ASOG method is<br />

displayed in Fig. 4.26. The results show that the water activity in mannitol/water solution<br />

decreases not only with increased mannitol mass fraction but also with increased<br />

liquid temperature.<br />

The effect of non-ideality through activity coefficient on the reduction of vapor<br />

pressure of water in PVP/water solution for different mass fractions of PVP dissolved in<br />

water at different temperatures is shown in Fig. 4.27. For the sake of comparison, ideal<br />

condition is also shown where the activity coefficient always remains at unity so that the<br />

vapor pressure is independent of solute mass fraction. It can be clearly observed that<br />

the liquid mixture strongly deviates from ideal behavior and the deviation increases<br />

with the increasing PVP mass fraction in water.<br />

Figure 4.28 shows the effect of non-ideality through PVP presence in PVP/water<br />

1<br />

0.8<br />

Ideal<br />

Y PVP<br />

= 0.2 (non­ideal)<br />

Y PVP<br />

= 0.3 (non­ideal)<br />

1<br />

0.8<br />

T = 50°C [P s<br />

= 0.12 bar]<br />

T = 100°C [P s<br />

= 1 bar]<br />

Vapor pressure [bar]<br />

0.6<br />

0.4<br />

Vapor pressure [bar]<br />

0.6<br />

0.4<br />

0.2<br />

0.2<br />

0<br />

0 20 40 60 80 100<br />

Temperature [K]<br />

0<br />

0 0.1 0.2 0.3 0.4<br />

Water mass fraction [­]<br />

Fig. 4.27: Effect of non-ideality on the vapor<br />

pressure of water at different<br />

temperatures.<br />

Fig. 4.28: Variation of vapor pressure of<br />

water with water mass fraction<br />

in PVP/water solution.


80 4. Results and Discussion<br />

solution on the water vapor pressure at 50 ◦ C and 100 ◦ C of liquid temperature. The<br />

vapor pressure of pure water at 50 ◦ C is about 0.12 bar whereas at 100 ◦ C it is 1 bar.<br />

An increase in water mass fraction increases the vapor pressure and it equals the pure<br />

water pressure when the water mass fraction is above 0.4. Thus, it infers that the<br />

role of water activity coefficient is important when the water mass fraction within the<br />

droplet falls below 0.4 at 50 ◦ C and 100 ◦ C, which occurs in the present simulations.<br />

4.3.3 Results and Discussion<br />

The simulation of evaporation and solid layer development of single droplet containing<br />

PVP or mannitol in water is carried out un<strong>der</strong> various drying conditions such as<br />

surrounding gas temperature, gas velocity, and relative humidity to investigate their<br />

effect on drying characteristics. The effect of the initial solute (PVP or mannitol) mass<br />

fraction on the final particle characteristics is also studied. The droplet is assumed to<br />

be spherical during the entire evaporation and drying process. The simulations are<br />

also carried out with rapid mixing model (RMM), which is a simple model based on<br />

the assumptions that the liquid mixture inside the droplet is always homogeneous and<br />

infinity conductivity within the droplet thus the droplet is at uniform temperature at<br />

every time. The governing equations of RMM are presented in Subsection 2.4.1.2.<br />

The thermal properties of PVP and mannitol are taken from Dakroury et al. [220],<br />

and mass diffusivity of PVP in water is obtained from Metaxiotou and Nychas [221],<br />

whereas the mass diffusivity of mannitol in water is taken from Grigoriev and Mey-<br />

0.7<br />

0.6<br />

Solute mass fraction [­]<br />

0.5<br />

0.4<br />

0.3<br />

PVP<br />

Mannitol<br />

0.2<br />

0.1<br />

0 20 40 60 80 100<br />

Temperature [°C]<br />

Fig. 4.29: Experimental data of PVP [218] and mannitol [219] saturation solubility in<br />

water.


4.3. Single Bi-component Droplet Evaporation and Solid Layer Formation 81<br />

Tab. 4.3: Experimental drying conditions<br />

Drying condition<br />

Values<br />

Initial solute mass fraction 0.075, 0.05 and 0.15<br />

Initial droplet radius 70 µm<br />

Initial droplet temperature<br />

Gas temperature<br />

Gas velocity<br />

20 and 70 ◦ C<br />

60, 67, 95, 100, 160 and 210 ◦ C<br />

0.05, 0.65 and 10 m/s<br />

Relative humidity (R.H.) 0.5 1.0, 2.0 and 30%<br />

likhov [222]. The critical temperature and pressure of PVP and mannitol are taken<br />

from Daubert and Danner [217]. The vapor diffusion coefficient through the solid layer<br />

of PVP or mannitol, D s , and solid thermal conductivity, k s are not available in literature,<br />

therefore, they are computed similar to the work of Nesic and Vodnik [151]. The<br />

physical and thermal properties in the film are estimated at the reference composition<br />

using the 1/3 rule [223]. The PVP/water and mannitol/water solution physical and<br />

thermal properties are computed with the standard rules of mixing. The variation<br />

of saturation solubility of PVP in water and mannitol in water with temperature is<br />

taken from measurements [218, 219], and it is shown in Fig. 4.29. The solid layer at<br />

the droplet surface is presumed to develop when the PVP mass fraction at the droplet<br />

surface reaches 20% above its saturation solubility limit, and in the case of mannitol,<br />

it is assumed that the crust and solid layer formation begins when the mannitol mass<br />

fraction reaches 0.9, which is much higher than the saturation solubility, in or<strong>der</strong> to<br />

avoid re-dissolution of solid layer with increased temperature as it shows large variation<br />

of solubility with temperature, see Fig. 4.29.<br />

The numerical results presented refer to a droplet of initial radius 70 µm at 20 ◦ C<br />

containing 0.15 PVP or mannitol initial mass fraction subjected to air with 0.5% relative<br />

humidity (R.H.) flowing at 0.65 m/s with 100 ◦ C initial gas temperature [172].<br />

The various drying conditions for numerical simulations taken from the experimental<br />

study of Littringer et al. [21] and Sedelmayer et al. [224], are listed in Tab. 4.3 and<br />

numerical results are compared with available experimental data [21, 224].<br />

Figure 4.30 shows the change in mannitol/water droplet mass and temperature<br />

with time for the above conditions and for increased initial gas velocity (U g = 10 m/s).<br />

Initially, there is no significant increase in droplet temperature, and droplet mass reduces<br />

due to continuous water evaporation. After an initial heating period, the droplet<br />

temperature rises very quickly indicating the formation of solid layer whereupon the<br />

rate of evaporation is reduced due to added resistance coming from solid layer, which is<br />

reflected in the droplet mass profile. The higher gas flow rate increases convection and


82 4. Results and Discussion<br />

Droplet mass [µg]<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

Mass [ U g<br />

= 0.65 m/s ]<br />

Mass [ U g<br />

= 10 m/s ]<br />

T [ U g<br />

= 0.65 m/s ]<br />

T [ U g<br />

= 10 m/s ]<br />

120<br />

100<br />

80<br />

60<br />

40<br />

Droplet temperature [°C]<br />

0.2<br />

0 0.5 1 1.5 2<br />

20<br />

2.5<br />

Time [s]<br />

Fig. 4.30: Effect of gas velocity on the evolution of mass and temperature of a mannitol/water<br />

droplet.<br />

thereby the water evaporation, hence there is quicker development of the solid layer.<br />

The solid layer forms in about 1.7 s with U g = 0.65 m/s, whereas with U g = 10 m/s, the<br />

solid layer forms in about 0.75 s. A closer look reveals that there is higher droplet mass<br />

at any given time after solid layer formation when compared with lower gas velocity<br />

situation, which means that increased gas velocity would lead to larger particle and<br />

the porosity, defined as the ratio of the volume occupied by water at the instance of<br />

1<br />

0.9<br />

0.8<br />

U g<br />

= 0.65 m/s<br />

T g<br />

= 67°C<br />

T g<br />

= 100°C<br />

T g<br />

= 160°C<br />

(d/d 0<br />

) 2 [-]<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0 0.5 1 1.5 2 2.5 3 3.5<br />

Time [s]<br />

Fig. 4.31: Effect of elevated gas temperature on the surface area of a mannitol/water<br />

droplet.


4.3. Single Bi-component Droplet Evaporation and Solid Layer Formation 83<br />

solid layer formation over that of the whole particle volume, would be higher in case<br />

of increased gas velocity (see Fig. 4.33).<br />

Figure 4.31 shows the effect of initial gas temperature on the temporal change of<br />

the dimensionless surface area of a mannitol/water droplet. Elevated gas temperature<br />

leads to higher energy transfer from the gas to the droplet, and thereby, an increase in<br />

the rate of droplet evaporation and drying. The surface area continuously decreases due<br />

to water evaporation until the beginning of solid layer formation whereupon particle<br />

size remains constant, which is reflected in Fig. 4.31.<br />

The higher the gas temperature the quicker the time taken for the solid layer formation:<br />

In case of 67 ◦ C the solid layer develops in about 2.9 s and with 100 ◦ C the solid<br />

layer forms in 1.7 s, whereas with 160 ◦ C, the same is observed in about 0.9 s. There<br />

is larger surface area at the time of solid layer formation with higher gas temperature,<br />

which means that elevated gas temperature would give larger particles towards the end<br />

of the drying process (see Fig. 4.33).<br />

The effect of gas temperature on the development of mannitol mass fraction profiles<br />

inside the droplet of initial radius 70 µm subjected to dry air with 0.5% R.H., flowing at<br />

0.65 m/s with temperatures of 67, 100 and 160 ◦ C is shown in Fig. 4.32 at 0.5 s (left)<br />

and at 0.9 s (right), respectively. Initially, the droplet interior has a homogenous<br />

mannitol mass fraction distribution of 0.15 (not shown here) and with time, there is<br />

development of mannitol mass fraction gradients inside the droplet, and the droplet<br />

size reduces due to continuous water evaporation. For 100 ◦ C initial gas temperature,<br />

the droplet radius is 62 µm at 0.5 s whereas at 0.9 s it reduces to 56 µm, which is<br />

seen in Fig. 4.32, respectively. The increased initial gas temperature yields higher mass<br />

fraction gradients inside the droplet mainly due to the decreased activity coefficient of<br />

0.35<br />

Time = 0.5 s<br />

0.8<br />

Mannitol mass fraction [­]<br />

0.3<br />

0.25<br />

0.2<br />

T g<br />

= 67°C<br />

T g<br />

= 100°C<br />

T g<br />

= 160°C<br />

Mannitol mass fraction [­]<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

Time = 0.9 s<br />

T g<br />

= 67°C<br />

T g<br />

= 100°C<br />

T g<br />

= 160°C<br />

0.15<br />

0 10 20 30 40 50 60 70<br />

Radial position inside the droplet [µm]<br />

0.2<br />

0 10 20 30 40 50 60 70<br />

Radial position inside the droplet [µm]<br />

Fig. 4.32: Effect of gas temperature on the temporal development of mannitol mass<br />

fraction inside the droplet at 0.5 s (left) and 0.9 s (right).


84 4. Results and Discussion<br />

Porosity [-]<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

Porosity [ U g<br />

= 0.65 m/s ]<br />

Porosity [ U g<br />

= 10 m/s ]<br />

Radius [ U g<br />

= 0.65 m/s ]<br />

Radius [ U g<br />

= 10 m/s ]<br />

50<br />

49<br />

48<br />

47<br />

46<br />

45<br />

44<br />

Particle radius [µm]<br />

0.2<br />

43<br />

0.1<br />

42<br />

60 90 120 150 180 210 240<br />

Gas temperature [°C]<br />

Fig. 4.33: Effect of gas velocity and temperature on porosity and final particle size of<br />

mannitol/water droplet.<br />

water in mannitol (see Fig. 4.26) and enhanced heat transfer. The activity coefficient<br />

decreases not only with increase in temperature but also with mannitol mass fraction,<br />

which is quite clearly seen at later times, i.e., 0.9 s shown in right part of Fig. 4.32.<br />

The effect of elevated gas velocity and temperature on the final particle porosity<br />

and particle radius is shown in Fig. 4.33. The porosity increases with higher gas temperature<br />

and velocity because of quicker solid layer formation, thereby yielding larger<br />

particles. The computed porosity of mannitol particle with 160 ◦ C gas temperature<br />

and 0.65 m/s gas velocity is 0.39 and the corresponding value in experiment is found<br />

to be 0.41 [21]. The final particle radius is reported as 42 µm in experiments [21],<br />

which can be compared to the corresponding computed value of 44.8 µm, showing a<br />

very good agreement.<br />

In experiments [21], it is reported that increased gas temperature leads to less<br />

porous particle with shriveled or non-spherical shape as seen in Fig. 4.34, which shows<br />

the the scanning electron microscope (SEM) images of the mannitol samples spray dried<br />

un<strong>der</strong> different drying temperatures. For the outlet temperature of 70 ◦ C, a spherical<br />

mannitol particle is obtained, see Fig. 4.34(a). For the higher outlet temperature<br />

of 100 ◦ C the particle shape changes from spherical to a ’raisin like’ structure, cf.<br />

Fig. 4.34(b), that may occur due to inflation of a drying shell and this transition is<br />

observed at 90 ◦ C, see Fig. 4.34(c). Higher temperatures lead to faster evaporation<br />

of the water, leading to less time to form a stable structure on the droplets surface.<br />

From the high porosity in combination with cuts of spray dried mannitol particles in<br />

previous studies [21], the formation of a particle with an outer shell is evident [225].


4.3. Single Bi-component Droplet Evaporation and Solid Layer Formation 85<br />

Fig. 4.34: SEM images of mannitol samples spray dried at 70 ◦ C (a), 100 ◦ C (b) and<br />

90 ◦ C (c). Zoomed images of the surface structures of these particles at<br />

70 ◦ C (d), 100 ◦ C (e) and 90 ◦ C (f) [225].<br />

This shell formation is in good accordance with the simulations performed so far.<br />

The particle surface consists of small, needle shaped structures in case of low drying<br />

temperatures as shown in Fig. 4.34(d), and smaller, non-needle shaped structures for<br />

higher drying temperatures, cf. Fig. 4.34(e), and the shift from needle shape to nonneedle<br />

structures is seen in Fig. 4.34(f). The increased gas temperature not only effects<br />

the final particle shape but also internal structure [225]. In computations, the change<br />

in particle shape is not accounted for, and it is assumed to be spherical throughout the<br />

evaporation and drying period, therefore, the present numerical results show increase<br />

in porosity with temperature as anticipated, see Fig. 4.33. This behavior will change<br />

when the final drying step is added to the present model, and if non-spherical particle<br />

formation will be consi<strong>der</strong>ed.<br />

Figure 4.35 shows the effect of gas temperatures of 60 ◦ C and 95 ◦ C and relative<br />

humidity of 1% R.H. (left) and 30% R.H. (right), respectively, on the droplet surface<br />

area and comparison with experimental data. The experiments are carried out by<br />

Sedelmayer et al. [224] at the University of Hamburg in an acoustic levitator. The<br />

simulation results show excellent agreement with the experiment. The droplet surface<br />

area continuously decreases due to water evaporation until a critical value where the<br />

solid layer formation starts, which is quite nicely predicted by the simulation. Increased<br />

temperature increases the evaporation rate and thereby quicker solid layer formation<br />

as seen in left part of Fig. 4.35, whereas increased humidity increases the solid layer<br />

formation time, i.e., at 60 ◦ C at 1% R.H. the solid layer forms in about 65 s and with


86 4. Results and Discussion<br />

1<br />

0.8<br />

R.H. = 1%<br />

Simulation [ T g<br />

= 60°C ]<br />

Experiment [ T g<br />

= 60°C ]<br />

Simulation [ T g<br />

= 95°C ]<br />

Experiment [ T g<br />

= 95°C ]<br />

1<br />

0.8<br />

R.H. = 30%<br />

Simulation [ T g<br />

= 60°C ]<br />

Experiment [ T g<br />

= 60°C ]<br />

Simulation [ T g<br />

= 95°C ]<br />

Experiment [ T g<br />

= 95°C ]<br />

(d/330 µm) 2 [-]<br />

0.6<br />

0.4<br />

(d/360 µm) 2 [-]<br />

0.6<br />

0.4<br />

0.2<br />

0.2<br />

0<br />

0 10 20 30 40 50 60 70 80<br />

Time [s]<br />

0<br />

0 50 100 150 200 250<br />

Time [s]<br />

Fig. 4.35: Effect of gas temperatures of 60 ◦ C and 95 ◦ C and relative humidity of<br />

1% R.H. (left) and 30% R.H. (right) on the droplet surface area.<br />

30% R.H. the same observed in about 205 s, see right part of Fig. 4.35. The profiles of<br />

the normalized droplet surface, (d/d 0 ) 2 , shown in the Fig. 4.35, reveal that the droplet<br />

evaporation rate prior to solid layer formation in the present case deviates from the<br />

linear decrease with time as would be expected from the classical d 2 law, where a<br />

constant evaporation constant is assumed.<br />

These experiments are carried out with different initial droplet radius for every<br />

experiment, and Tab. 4.4 gives the initial droplet radii (R 0 ) and particle size at the<br />

time of solid layer formation (t s ) in every experiment and its corresponding computed<br />

value from simulation.<br />

The comparison between rapid mixing model (RMM) and the present model is given<br />

in Fig. 4.36, which shows the time evolution of mannitol/water droplet surface area for<br />

initial droplet radius of 70 µm at 20 ◦ C temperature and subjected to hot air of 160 ◦ C<br />

with 0.5% R.H. and flowing at 0.65 m/s. Even though there is little difference between<br />

RMM and the present approach during the initial time period, however, in the later<br />

Tab. 4.4: Experiment vs simulation<br />

T g R.H. R 0 Particle radius at t s [µm]<br />

[ ◦ C] [%] [µm] Simulation Experiment<br />

60 1.0 330 145.1 147.9<br />

30.0 360 155.2 155.1<br />

95 1.0 215 95.0 109.4<br />

30.0 280 122.5 121.4


4.3. Single Bi-component Droplet Evaporation and Solid Layer Formation 87<br />

1<br />

0.9<br />

0.8<br />

T g<br />

= 160°C, U g<br />

= 0.65 m/s<br />

RMM<br />

Present model<br />

(d/d 0<br />

) 2 [­]<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

Time [s]<br />

Fig. 4.36: Time evolution of mannitol/water droplet surface area computed by present<br />

model and RMM.<br />

time period RMM overpredicts the decrease in droplet surface and thereby the time of<br />

the solid layer formation caused by the fact that the assumption of homogeneous liquid<br />

mixture within the droplet. This assumption leads to more water to be evaporated,<br />

which increases the solute mass fraction to the critical value so that the formation of<br />

solid layer begins.<br />

The effect of initial droplet temperatures of 20 ◦ C and 70 ◦ C on the evaporation<br />

1.6<br />

T g<br />

= 160°C<br />

120<br />

Droplet mass [µg]<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

Mass [ T 0<br />

= 20°C ]<br />

T [ T 0<br />

= 20°C ]<br />

Mass [ T 0<br />

= 70°C ]<br />

T [ T 0<br />

= 70°C ]<br />

100<br />

80<br />

60<br />

40<br />

Droplet temperature [°C]<br />

0.2<br />

0 0.2 0.4 0.6 0.8 1<br />

20<br />

1.2<br />

Time [s]<br />

Fig. 4.37: Effect of initial droplet temperature on the evaporation rate of mannitol/water<br />

droplet.


88 4. Results and Discussion<br />

Solid layer thickness [µm]<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

U g<br />

= 0.65 m/s<br />

T g<br />

= 100°C<br />

T g<br />

= 160°C<br />

5<br />

0<br />

0.5 0.75 1 1.25 1.5 1.75 2 2.25<br />

Time [s]<br />

Fig. 4.38: Effect of gas temperature on solid layer thickness inside the PVP/water<br />

droplet.<br />

rate is shown in Fig. 4.37. In both the cases, the initial droplet radius is 70 µm, and<br />

it is subjected to hot air flowing at 0.65 m/s with 160 ◦ C. The droplet with 20 ◦ C<br />

initial temperature quickly raises to an equilibrium temperature, which is most often<br />

equal to the wet bulb temperature, whereupon no significant rise in temperature is<br />

found. Whereas with 70 ◦ C, the wet bulb temperature for the gas temperature of<br />

160 ◦ C and 0.5% R.H., is lower than the initial droplet temperature (70 ◦ C), so the<br />

droplet temperature decreases until it equals the wet bulb temperature, and remains<br />

almost constant in further development. Similarly the droplet evaporation rate is<br />

higher in this initial period, and it is reflected in the reduction of droplet mass as seen<br />

in Fig. 4.37. In the later time period, the final particle temperature is same, and it is<br />

equal to 105 ◦ C.<br />

Similar trends are observed for PVP/water evaporation and solid layer formation.<br />

The effect of elevated gas temperature on the temporal development of solid layer<br />

thickness in PVP/water droplet is shown in Fig. 4.38 for the same conditions that are<br />

studied for mannitol/water. Increased gas temperature of T g = 160 ◦ C leads to higher<br />

energy transfer and earlier molecular entanglements of PVP and solid layer formation,<br />

with 100 ◦ C the solid layer forms in about 1.4 s whereas with 160 ◦ C, the same is<br />

observed in 0.7 s, see Fig. 4.38.<br />

Comparison of PVP/water droplet evaporation and solid layer formation with that<br />

of mannitol/water un<strong>der</strong> the same drying conditions reveals that the solid layer forms<br />

quicker in case of PVP/water (in about 1.5 s with 100 ◦ C, see Fig. 4.38) than mannitol/water<br />

(about 1.7 s with 100 ◦ C, see Fig. 4.30). This is due to the fact that the


4.3. Single Bi-component Droplet Evaporation and Solid Layer Formation 89<br />

1.5<br />

Mass [ T g<br />

= 100°C ]<br />

Mass [ T g<br />

= 160°C ]<br />

T [ T g<br />

= 100°C ]<br />

T [ T g<br />

= 160°C ]<br />

100<br />

Droplet mass [µg]<br />

1.25<br />

1<br />

0.75<br />

0.5<br />

80<br />

60<br />

40<br />

Droplet temperature [°C]<br />

0.25<br />

0 0.5 1 1.5 2<br />

20<br />

2.5<br />

Time [s]<br />

Fig. 4.39: Effect of gas temperature on the droplet mass and temperature.<br />

required solute mass fraction for initiation of the solid layer formation is less in case of<br />

PVP (about 0.78 at 100 ◦ C, see Fig. 4.29) compared to mannitol, which is fixed to 0.9.<br />

Figure 4.39 shows the effect of gas temperature on the temporal evolution of<br />

PVP/water droplet mass and temperature when the droplet is subjected to 100 ◦ C<br />

and 160 ◦ C gas temperatures. Elevated temperature leads to higher energy transfer<br />

from the gas to the droplet, and thereby, an increase in the rate of droplet evaporation<br />

and drying, which is reflected in Fig. 4.39. The higher the gas temperature the quicker<br />

the time taken to see molecular entanglement leading to solid layer formation: in case<br />

of 160 ◦ C, the solid layer develops in about 0.7 s whereas with 100 ◦ C , the same is<br />

observed in about 1.5 s, which is in agreement with Fig. 4.38. This means that an<br />

increase in gas temperature would give larger particles towards the end of the drying<br />

process.<br />

Figure 4.40 shows the temporal development of PVP mass fraction profiles inside<br />

the PVP/water droplet of initial radius 70 µm subjected to hot air flowing at 0.65 m/s<br />

with 100 ◦ C temperature and no humidity, i.e., dry air (left) and with 5% R.H. (right),<br />

respectively. Initially, the droplet has a homogenous PVP mass fraction of 0.15 and<br />

with time the droplet size decreases, and there is development of PVP mass fraction<br />

gradients inside the droplet due to continuous water evaporation, which can be seen at<br />

later times in both the figures. The PVP mass fraction at the droplet surface reaches<br />

the value of 0.78 in about 1.4 s with dry air, as seen left side of Fig. 4.40, which is<br />

equivalent to 20% above the saturation solubility whereas the same is achieved after<br />

1.8 s with 5% R.H., see right part of Fig. 4.40. This indicates that the increase in<br />

humidity prolongs the drying period because of the reduced driving force for water


90 4. Results and Discussion<br />

0.8<br />

0.8<br />

Time<br />

PVP mass fraction [­]<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

Time<br />

0 s<br />

0.5 s<br />

1.0 s<br />

1.4 s<br />

PVP mass fraction [­]<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0 s<br />

0.5 s<br />

1.0 s<br />

1.5 s<br />

1.8 s<br />

0.2<br />

0.2<br />

0.1<br />

0 10 20 30 40 50 60 70<br />

Radial position inside the droplet [µm]<br />

0.1<br />

0 10 20 30 40 50 60 70<br />

Radial position inside the droplet [µm]<br />

Fig. 4.40: Temporal development of PVP mass fraction profiles inside the droplet subjected<br />

to dry air (left) and hot air with 5% R.H (right).<br />

evaporation, cf. Eq. (2.54). It is also observed that there is a lower PVP mass fraction<br />

gradient within the droplet for 5% R.H., cf. Fig. 4.40, when compared with dry air<br />

before solid layer develops, implying that humidity leads to smaller size particles with<br />

less porosity. Thus, it appears that the relative humidity plays a major role in the<br />

mass fraction gradients development within the droplet.<br />

Figure 4.41 shows the comparison of present model predictions of PVP/water<br />

droplet surface and that of RMM. The behavior is similar to the revelations made in<br />

1<br />

0.9<br />

T g<br />

= 100°C, U g<br />

= 0.65 m/s<br />

RMM<br />

Present model<br />

0.8<br />

(d/d 0<br />

) 2 [­]<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0 0.3 0.6 0.9 1.2 1.5 1.8<br />

Time [s]<br />

Fig. 4.41: Time evolution of PVP/water droplet surface area predicted by present model<br />

and RMM.


4.3. Single Bi-component Droplet Evaporation and Solid Layer Formation 91<br />

Evaporated water mass [µg]<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

T g<br />

= 100°C, U g<br />

= 0.65 m/s<br />

R.H. = 0.5%<br />

R.H. = 2.0%<br />

0<br />

0 0.25 0.5 0.75 1 1.25 1.5 1.75 2<br />

Time [s]<br />

Fig. 4.42: Effect of relative humidity on the water evaporation rate from the mannitol/water<br />

droplet.<br />

mannitol/water droplet case, i.e., the RMM overpredicts the decrease in droplet surface<br />

area, and thereby the time required for solid layer formation due to the assumption<br />

of homogeneous liquid mixture within the droplet, which delays the formation of the<br />

solid layer.<br />

Figure 4.42 shows the effect of relative humidity on the water evaporation rate from<br />

a mannitol/water droplet subjected to air with 0.5% and 2.0% R.H. For low relative<br />

humidity, the mass fraction of water vapor in the bulk of the air, Y 1,∞ , cf. Eq. (2.54),<br />

is decreased, leading to a higher driving mass transfer rate, and this would eventually<br />

cause faster water evaporation, and thereby somewhat quicker solid layer formation.<br />

With 0.5% R.H. the solid layer develops in about 1.7 s whereas with 2% R.H., the same<br />

is observed in 1.9 s.<br />

Figure 4.43 shows the effect of modification of initial PVP mass fraction on the evolution<br />

of droplet radius and temperature for 0.075 and 0.15 PVP initial mass fractions.<br />

All other conditions remain fixed. Less initial PVP mass fraction implies that there<br />

is more water to evaporate leading to smaller size particle with longer drying time.<br />

With an initial PVP mass fraction of 0.15, the droplet radius reduces to 46.4 µm in<br />

about 1.4 s whereas with 0.075 PVP initial mass fraction, the droplet radius decreases<br />

to 38.5 µm in about 1.7 s before the solid layer formation begins, which is indicated<br />

by the quick rise in droplet temperature reaching the same value in both cases as<br />

seen in Fig. 4.43, showing that initial mass fraction of PVP does not affect the final<br />

temperature of the particle.<br />

Though the final drying step is not yet added to this model, the results presented


92 4. Results and Discussion<br />

70<br />

65<br />

Radius [ Y PVP<br />

= 0.075 ]<br />

Radius [ Y PVP<br />

= 0.15 ]<br />

T [ Y PVP<br />

= 0.075 ]<br />

T [ Y PVP<br />

= 0.15 ]<br />

100<br />

Droplet radius [µm]<br />

60<br />

55<br />

50<br />

45<br />

80<br />

60<br />

40<br />

Droplet temperature [°C]<br />

40<br />

35<br />

0 0.5 1 1.5 2<br />

20<br />

2.5<br />

Time [s]<br />

Fig. 4.43: Effect of initial PVP mass fraction on the profiles of droplet radius and temperature.<br />

here for PVP/water and mannitol/water single droplet evaporation and solid layer formation<br />

is very promising. The first three stages of bi-component droplet evaporation<br />

and drying, i.e., till the solid layer development on the droplet surface is effectively<br />

predicted by present model and the comparison of the numerical results with the experiment<br />

exhibits very good agreement. Thus, this model is included in DQMOM<br />

mathematical formulation in or<strong>der</strong> to simulate bi-component PVP/water spray flows<br />

in an axisymmetric, two-dimensional configuration and the results of this system are<br />

presented in the next section.<br />

4.4 Two-dimensional Evaporating PVP/Water<br />

Spray in Air<br />

This section presents the numerical and experimental results of bi-component evaporating<br />

spray flows. Though the developed model can be applied to simulate PVP/water<br />

and mannitol/water spray flows, but here only the results of PVP/water spray flows<br />

are presented as the initial data with respect to mannitol/water spray is not available.<br />

The experimental setup and initial data generation to start numerical simulations are<br />

presented in the next subsection followed by the results and discussion.


4.4. Two-dimensional Evaporating PVP/Water Spray in Air 93<br />

Fig. 4.44: Photograph of the PVP/water spray formation with 112 kg/h liquid inflow<br />

rate in experiment [206].<br />

4.4.1 Experiment and Initial Data Generation<br />

The PVP/water spray in air experiments have been carried out by the group of<br />

Prof. G. Brenn by spraying a solution of 20% PVP and 80% water (by mass) through<br />

the Delavan nozzle SDX-SE-90 at room temperature. The droplet sizes and velocities<br />

are recorded at the cross sections of 0.08, 0.12 and 0.16 m away from nozzle orifice<br />

using PDA, which provides both droplet size and velocity distributions, similar to the<br />

water spray in air measurements. The liquid mass flow rate of these experiments is<br />

112 kg/h and other conditions of the experiment such as gas velocity, gas temperature<br />

and pressure are same as the water spray in air. Figure 4.44 displays the PVP/water<br />

spray formation in experiment [206]. To generate the initial data for simulating the<br />

PVP/water spray in air, the same procedure as outlined in two-dimensional water spray<br />

in air is followed here. Figure 4.45 shows the experimental droplet size distribution<br />

and the corresponding DQMOM approximation at the radial position 0.036 m from<br />

the spray axis and 0.08 m downstream of the nozzle orifice.<br />

4.4.2 Results and Discussion<br />

PVP/water spray flow in air is modeled using the DQMOM where the bi-component<br />

droplet evaporation of PVP/water droplets [172, 225, 226] are accounted through the<br />

single droplet evaporation and solid formation model presented in Chapter 2 and the<br />

results are discussed in Subsection 4.3.3. For droplet motion, droplet coalescence, the


94 4. Results and Discussion<br />

0.15<br />

0.12<br />

Experiment<br />

DQMOM<br />

NDF [(µm) ­1 ]<br />

0.09<br />

0.06<br />

0.03<br />

0<br />

0 20 40 60 80 100 120<br />

Droplet radius [µm]<br />

Fig. 4.45: Experimental and DQMOM approximation of droplet number density for<br />

PVP/water spray.<br />

same sub-models as employed in water spray are applied here, see Subsection 2.4.<br />

In Fig. 4.46, computed and experimental profiles of Sauter mean diameter (left) and<br />

mean droplet diameter (right) of PVP/water spray for a mass inflow rate of 112 kg/h<br />

at 0.12 m away from the nozzle exit are shown. Similar to the water spray, the spray<br />

distribution assumes a hollow-cone shape, and it is nicely predicted by DQMOM. In<br />

both the figures, a closer look reveals that across all the radial positions, the DQMOM<br />

Sauter mean diameter [µm]<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

DQMOM<br />

Experiment<br />

Mean droplet diameter [µm]<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

DQMOM<br />

Experiment<br />

20<br />

-75 -60 -45 -30 -15 0 15 30 45 60 75<br />

Radial position [mm]<br />

0<br />

-75 -60 -45 -30 -15 0 15 30 45 60 75<br />

Radial position [mm]<br />

Fig. 4.46: Experimental and numerical profiles of the Sauter mean diameter (left) and<br />

mean droplet diameter (right) of PVP/water spray in air at the cross section<br />

of 0.12 m distance from the nozzle exit.


4.4. Two-dimensional Evaporating PVP/Water Spray in Air 95<br />

Sauter mean diameter [µm]<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

DQMOM<br />

Experiment<br />

Mean droplet diameter [µm]<br />

100<br />

80<br />

60<br />

40<br />

20<br />

DQMOM<br />

Experiment<br />

20<br />

-100 -80 -60 -40 -20 0 20 40 60 80 100<br />

Radial position [mm]<br />

0<br />

-100 -80 -60 -40 -20 0 20 40 60 80 100<br />

Radial position [mm]<br />

Fig. 4.47: Experimental and numerical profiles of the Sauter mean diameter (left) and<br />

mean droplet diameter (right) of PVP/water spray in air at the cross section<br />

of 0.16 m distance from the nozzle exit.<br />

un<strong>der</strong>predicts the experimental results and towards the periphery of the spray some<br />

deviation is observed particularly in profiles of Sauter mean diameter as compared to<br />

the experiment. This can possibly be explained by the fact that the DQMOM predicts<br />

the global droplet distribution, but there could be local discrepancies induced by the<br />

gas phase, which is not resolved in the present study. Coupling of DQMOM with the<br />

gas phase would eventually improve the simulation results.<br />

Figure 4.47 displays the Sauter mean diameter (left) and mean droplet diameter<br />

(right) at further downstream the nozzle exit, i.e., at the cross section 0.16 m.<br />

Comparing the maxima in Fig. 4.46 and 4.47 reveals that there is an increase in the<br />

Sauter mean diameter and mean droplet diameter, which is converse to the the water<br />

spray where decrease in droplet size is found. At a given temperature, the evaporation<br />

rate of water from pure water droplets is higher than from the droplets containing PVP<br />

dissolved in water due to the non-ideality effect (see Fig. 4.25). An analysis of droplet<br />

coalescence reveals that it occurs 1.5 times more often in PVP/water spray compared<br />

to water spray, which also contributes to an increased droplet size in the PVP/water<br />

spray. The elevated viscosity of PVP leading to higher viscous PVP/water droplets<br />

compared to pure water droplets influences the droplet coalescence. The present model<br />

is suitable to capture these effects, and a good agreement between the experiment and<br />

simulation is found [209].<br />

The mean droplet velocity of PVP/water spray with 112 kg/h liquid inflow rate at<br />

0.12 m away from the nozzle is shown in Fig. 4.48. Increased liquid flow rate leads to<br />

higher droplet velocity (compare Figs. 4.23 and 4.48), which increases the chances of


96 4. Results and Discussion<br />

23<br />

21<br />

DQMOM<br />

Experiment<br />

Mean droplet velocity [m/s]<br />

19<br />

17<br />

15<br />

13<br />

11<br />

9<br />

7<br />

-75 -60 -45 -30 -15 0 15 30 45 60 75<br />

Radial position [mm]<br />

Fig. 4.48: Experimental and numerical profiles of the mean droplet velocity of<br />

PVP/water spray in air at the cross section of 0.12 m distance from the<br />

nozzle exit.<br />

collision. The smaller size droplets that lie closer to centerline of the spray are moving<br />

at higher velocity than the larger size droplets, which is in quite contrast with that of<br />

water spray (see Fig. 4.23) where both the larger and smaller size droplets move with<br />

higher velocity. This may be because initially the gas around the spray is stagnant<br />

and the droplets decelerate by aerodynamic drag. The surrounding gas acquires the<br />

momentum lost by the droplets, and this creates a flow field in which gas is continually<br />

entrained into the spray. As the entrained gas enters the spray, it drags small liquid<br />

drops at the outer regions of the spray inward, and the momentum lost by the droplets<br />

at the periphery of the spray is larger than the ones that lie closer to the axis of<br />

symmetry, which explains the smaller velocity of larger droplets [227, 228].<br />

Further downstream of the nozzle exit, i.e., at the cross section of 0.16 m away<br />

from the nozzle exit, the retardation of the droplet velocity in large size droplet region<br />

is observed (see Fig. 4.49) similar to water spray as this effect is dependent on initial<br />

liquid flow rate, where low liquid flow rate leads to larger droplets, which take more<br />

time to follow the streamlines of the gas than the smaller size droplets [209, 225].<br />

The simulation results are in good agreement with the experiment, particularly in<br />

smaller size droplets region whereas towards spray edge there is deviation, which can<br />

be attributed to the post-processing of the experimental data and the non-resolved gas<br />

phase.<br />

Concerning the differences in evaporation characteristics for water and PVP/water<br />

droplet evaporation in air, it is found that for a given liquid flow rate and axial po-


4.4. Two-dimensional Evaporating PVP/Water Spray in Air 97<br />

23 DQMOM<br />

Experiment<br />

21<br />

Mean droplet velocity [m/s]<br />

19<br />

17<br />

15<br />

13<br />

11<br />

9<br />

7<br />

-80 -60 -40 -20 0 20 40 60 80<br />

Radial position [mm]<br />

Fig. 4.49: Experimental and numerical profiles of the mean droplet velocity of<br />

PVP/water spray in air at the cross section of 0.16 m distance from the<br />

nozzle exit.<br />

sition from the nozzle exit, droplet size is larger in water spray (120 kg/h) than the<br />

PVP/water spray (112 kg/h), compare Fig. 4.19 with left side of Fig. 4.46, which is<br />

because of the high viscosity of PVP/water solution. Moreover, at a given temperature,<br />

the evaporation rate from pure water droplet is higher than from PVP/water droplet<br />

due to the non-ideality effect caused by polymer presence.


98 4. Results and Discussion


5. Conclusions and Future Work<br />

The objective of the present work is modeling and simulation of polymer or sugar solution<br />

spray drying until the solid layer formation at the droplet surface, and dispersion<br />

in bi-component evaporating spray flows in an Eulerian framework.<br />

In or<strong>der</strong> to un<strong>der</strong>stand the behavior of droplet distribution un<strong>der</strong> various drying<br />

conditions, the direct quadrature method of moments (DQMOM) is implemented, for<br />

the first time, in two dimensions to study the bi-component evaporating spray flows. In<br />

DQMOM, the droplet size and velocity distribution of the spray is modeled by approximating<br />

the number density function in terms of joint radius and velocity. The DQMOM<br />

has been extended to accommodate gas–liquid interactions such as convective droplet<br />

evaporation, drag force and gravity as well as droplet–droplet interactions by including<br />

coalescence. The effect of these physical processes on the evolution of droplet size<br />

distribution and kinetic properties is analyzed and validated with the experiments.<br />

The DQMOM simulation results are also compared with the quadrature method of<br />

moments (QMOM) in one-dimensional configuration whereas in two-dimensional axisymmetric<br />

configuration DQMOM is compared with discrete droplet model (DDM),<br />

which is a well known Euler – Lagrangian approach.<br />

First, evaporating water spray in nitrogen is modeled using DQMOM in one physical<br />

dimension, and the simulation results are compared with QMOM. The water evaporation<br />

is accounted through convective evaporation model of Abramzon and Sirignano,<br />

which accounts for variable liquid and film properties. The drag and droplet coalescence<br />

are included through appropriate sub-models. The gas phase is not yet fully coupled<br />

with DQMOM but its inlet flow properties are used to compute droplet evaporation<br />

and drag. The initial data to start simulations is generated from the experimental data,<br />

which were provided by Dr. R. Wengeler, BASF Ludwigshafen. The simulation results<br />

are validated with experiment at various cross sections. The influence of individual<br />

physical processes is analyzed. It is demonstrated that the model reflects the evaporation<br />

to have a pronounced effect on the parameters pertaining to droplet size. More<br />

importantly, when evaporation is consi<strong>der</strong>ed in combination with droplet coalescence,<br />

the numerical results are improved significantly and show excellent agreement with<br />

experiments. The droplet velocity is largely influenced by the drag force and gravity.<br />

Based on the successful of implementation of DQMOM, it is then extended to model<br />

evaporating water in air in two-dimensional, axisymmetric configuration. The same


100 5. Conclusions and Future Work<br />

system is also modeled using DDM. In DQMOM, the source terms are computed same<br />

as done in the one-dimensional case, and the evolution of droplet size and velocity distributions<br />

are analyzed with both DDM and DQMOM. Droplet collisions are included<br />

in DQMOM by modeling the droplet coalescence. The DDM does not include droplet<br />

collisions due to computational complexity such as redistribution of droplet classes<br />

and increased computational effort. For initialization and validation of the simulation<br />

results, experimental data is used, which was provided by Prof. G. Brenn TU Graz,<br />

measured using PDA. The experimental data contains droplet size and velocity in axial<br />

and radial direction and this data is post-processed in or<strong>der</strong> to eliminate errors in the<br />

large size droplets region. The experimental data at the cross section closest to the<br />

nozzle exit are used for the generation of initial conditions for the simulations, and the<br />

numerical results of DQMOM are compared with experimental data at the later cross<br />

sections, and with DDM.<br />

Overall, both the methods i.e., DQMOM and DDM show good agreement with<br />

the experiment. Some deviations between DQMOM and experiment are observed that<br />

might result from the present DQMOM formulation, which is not yet fully coupled<br />

with the gas phase equations. Concerning the experimental data, a post-processing of<br />

the raw data has been performed in or<strong>der</strong> to correct the number density of large size<br />

droplets with respect to the effective cross section area, leading to different correction<br />

factors for different axial positions in experimental data away from the nozzle exit,<br />

which may also lead to discrepancy between numerical and experimental results. The<br />

DDM performs somewhat better in the periphery of the spray and DQMOM near the<br />

centerline. However, DQMOM shows an excellent numerical performance, and droplet<br />

coalescence is included with relative ease compared to DDM. Therefore, the DQMOM<br />

is further extended to simulate PVP/water sprays in air.<br />

Before simulating the PVP/water spray in air using DQMOM, a model to describe<br />

the bi-component droplet evaporation and solid layer formation is developed. The<br />

system un<strong>der</strong> consi<strong>der</strong>ation is governed by the continuity (diffusion) and energy equations.<br />

Brenn’s model is modified to include the resistance from the solid layer, and<br />

this extended formulation is used to compute the evaporation rate of water from the<br />

bi-component droplet. The temperature inside the droplet appears to be uniform, and<br />

the change in droplet temperature due to heat exchange between the droplet surface<br />

and the surrounding gas is calculated with similar modifications used for mass evaporation<br />

rate to account for the resistance from the solid layer. The variable physical<br />

and thermal properties and the volume fraction based radius are introduced based on<br />

Brenn’s model. The predictability and efficiency of the developed single droplet model<br />

is first verified by simulating PVP/water and mannitol/water droplets. The liquid<br />

mixture is treated as non-ideal with the activity coefficient calculation using the im-


101<br />

proved UNIFAC-vdW-FV method for PVP in water and ASOG contribution method<br />

for mannitol/water solution.<br />

The effect of various drying conditions on the evolution of single droplet characteristics<br />

is analyzed and the results are compared with experimental data. These drying<br />

conditions include the effect of gas temperature, gas velocity, initial solute (PVP or<br />

mannitol) mass fraction and relative humidity, which are found to have significant<br />

effect on the evaporation and drying characteristics of PVP/water as well as mannitol/water<br />

droplet. The study reveals that an increase in gas velocity and temperature<br />

cause earlier formation of the solid layer and faster drying, leading to larger particles<br />

with higher porosity. Humidity in air leads to smaller size particles with less porosity.<br />

The lower initial solute (PVP or mannitol) mass fraction implies that there is more<br />

water to evaporate resulting in smaller size particle with longer drying time. The variation<br />

of the activity coefficient of water with PVP mass fraction is much higher than<br />

that of mannitol, causing a stronger retardation of water evaporation rate from the<br />

PVP/water droplet compared to mannitol/water, which results in a faster solid layer<br />

formation for mannitol/water droplets. The present model successfully predicts the<br />

first three stages of droplet evaporation and drying, i.e., until solid layer formation at<br />

the droplet surface, which can readily be incorporated into an overall model of spray<br />

drying.<br />

The developed bi-component single droplet evaporation and drying model is then<br />

included in the DQMOM to simulate evaporating PVP/water spray flows in air in<br />

two-dimensional, axisymmetric configuration. Even though the developed spray model<br />

is equally applicable to simulate PVP/water and mannitol/water spray flows but only<br />

computations of PVP/water spray flows are performed as mannitol/water spray flow<br />

experimental data is not available. The physical processes of the spray such as drag<br />

and coalescence are included through the appropriate sub-models. Numerical results<br />

are compared with experimental data at different cross sections, and results are found<br />

to be in good agreement with experiment. Some deviations between DQMOM and<br />

experiment in case of mean droplet diameter are observed that might have originated<br />

from the present DQMOM formulation, which is not yet fully coupled with the gas<br />

phase equations. Moreover, the employed numerical technique uses an explicit finite<br />

difference method to solve the DQMOM transport equations – an implicit scheme may<br />

lead to consi<strong>der</strong>able improvement. Additionally, the Schiller–Naumann correlation for<br />

drag coefficient may need revision. In conclusion, mono- and bi-component evaporating<br />

spray flows and preliminary stages of spray drying are successfully modeled and<br />

simulated using DQMOM.<br />

During spray drying, elevated gas temperature enhances droplet heating and evaporation<br />

thus advanced formulation of DQMOM is required in or<strong>der</strong> to model spray


102 5. Conclusions and Future Work<br />

drying. Even though the current DQMOM method includes the droplet radius and<br />

velocity as internal variables, inclusion of droplet temperature and analogous terms<br />

in the Williams’ spray equation will enable modeling of spray drying. The gas phase<br />

equations should be resolved and coupled to the DQMOM transport equations, which<br />

would eventually enable the model to predict the spray drying process, i.e., from the<br />

atomized liquid droplets to the dried end product.<br />

Concerning the bi-component single droplet evaporation and drying, the developed<br />

model needs further extension to account for the capillary force, internal circulation,<br />

shriveling effect or disorientation of particle shape towards the final stages of drying as<br />

observed in experiment. This would further improve the model predictability of solid<br />

layer formation, porosity within the solid layer and final particle size.


Appendix


A. Nomenclature


106 A. Nomenclature<br />

Symbol Unit Description<br />

a w - Water activity<br />

a α , b α - Variables in Wheeler algorithm<br />

a n<br />

Source term in DQMOM<br />

A(t) m 2 Time dependent droplet surface area<br />

B M - Spalding mass transfer number<br />

B M,i - Spalding mass transfer number of coefficient i<br />

B T - Spalding heat transfer number<br />

Bi - Biot number<br />

b n<br />

Source term in DQMOM<br />

c m/s Speed of the sound in gas medium<br />

C D - Drag coefficient in spray model<br />

C pg J/(kg K) Specific heat capacity of gas<br />

C pL J/(kg K) Specific heat capacity of liquid<br />

C pLf J/(kg K) Specific heat capacity in the film<br />

c n<br />

Source term in DQMOM<br />

c 1,n , c 2,n , c 3,n<br />

Source terms in DQMOM<br />

d m Droplet diameter<br />

D 12 m 2 /s Binary diffusion coefficient of liquid mixture<br />

d 1,0 , d 10 m Mean droplet diameter<br />

d 3,2 , d 32 m Sauter mean diameter<br />

d k+1,k m General definition of mean diameter<br />

D f m 2 /s Mass diffusion coefficient in the film<br />

D s m 2 /s Mass diffusion coefficient of vapor in solid layer<br />

E c - Efficiency of the droplet coalescence<br />

F m/s 2 Total force per unit mass on droplets<br />

F h m/s 2 History term or Basset force per unit mass<br />

F L m/s 2 Lift force per unit mass<br />

f<br />

Droplet distribution function<br />

f m −3 Number density function<br />

fM ME(x)<br />

m−3 Maximum entropy method approximation of number density<br />

function<br />

g m/s 2 Acceleration due to gravity<br />

G<br />

Moment flux term defined in finite volume method<br />

h W/(m 2 K) Convective heat transfer coefficient<br />

h J/kg Total specific enthalpy<br />

H - Hessian matrix<br />

H[f]<br />

Shannon entropy


107<br />

J c q J/(m 2 s) Heat flux due to thermal conductivity<br />

J d q J/(m 2 s) Heat flux due to molecular mass diffusion<br />

k m 2 /s 2 Turbulent kinetic energy<br />

k m 2 /s Evaporation constant in d 2 law<br />

k gf W/(m K) Thermal conductivity in the film<br />

k l W/(m K) Thermal conductivity of the liquid<br />

k s W/(m K) Thermal conductivity of the solid layer<br />

L V (T s ) J/kg Latent heat of vaporization at T s<br />

Le - Lewis number<br />

m kg Droplet mass<br />

m p,k m Droplet mass in k th parcel<br />

ṁ kg/s Total evaporation rate<br />

M<br />

Ma - Mach number<br />

M k<br />

Moment set defined in QMOM<br />

k th moment<br />

M w kg/mol Water molecular weight<br />

¯M kg/mol Mean molecular weight in film<br />

n d (d) m −3 Number density function based on the droplet diameter d<br />

N m −3 Total number density<br />

N - Number of nodes in DQMOM<br />

Ñu - Modified Nusselt number<br />

Nu - Nusselt number<br />

Oh - Ohnesorge number<br />

p Pa Pressure<br />

P k,l<br />

Phase-space transform defined in DQMOM<br />

p m Pa Total pressure in the film<br />

p vap,i Pa Vapor pressure of component i<br />

Pr - Prandtl number<br />

Q L J/s Net heat transferred to the droplet<br />

Q f<br />

r m Radial coordinate<br />

Rate of change in f due to droplet coalescence<br />

r n m Approximated droplet radius of n th node in DQMOM<br />

r p,k m Droplet radius in k th parcel<br />

Re - Reynolds number<br />

Re d - Droplet Reynolds number<br />

R 0 m Initial droplet radius<br />

R i m Volume fraction based droplet radius<br />

R m Droplet radius<br />

R = dr<br />

dt<br />

m/s Rate of change in droplet radius due to evaporation


108 A. Nomenclature<br />

S n<br />

S g<br />

S l<br />

Vector of source terms<br />

Source term due to gas phase<br />

Source term due to liquid phase<br />

S α kg/s Chemical production rate of species α in mass<br />

Sc - Schmidt number<br />

˜Sh - Modified Sherwood number<br />

Sh - Sherwood number<br />

T K Droplet temperature<br />

T g K Gas temperature<br />

T s K Droplet surface temperature<br />

T ∞ K Temperature in the bulk of the gas<br />

T p,k K Droplet temperature in k th parcel<br />

t s Time<br />

t s s Time taken for initiation of solid layer<br />

u m/s Gas velocity in physical space<br />

V m/s Gas velocity in sample space<br />

v m/s Droplet velocity in physical space<br />

U g m/s Gas velocity<br />

u x m/s Axial component of gas velocity<br />

u r m/s Radial component of gas velocity<br />

v x m/s Axial component of droplet velocity<br />

v r m/s Radial component of droplet velocity<br />

V m 3 Droplet volume<br />

V i m 3 Volume of component i within the droplet<br />

We g - Gas Weber number<br />

〈<br />

x<br />

k 〉 Approximated k th moment<br />

X i - Mole fraction of species i<br />

X L,i - Mole fraction of species i in the liquid<br />

x m Geometrical coordinates<br />

Y i - Mass fraction of component i inside the droplet<br />

Y s - Mass fraction of vapor at the droplet surface<br />

Y ∞<br />

Mass fraction of vapor in the bulk of the gas<br />

Y i,s - Mass fraction of species i at the droplet surface<br />

Y i,∞ - Mass fraction of species i in the bulk of the gas<br />

γ i - Activity coefficient of component i<br />

Γ f<br />

Droplet coalescence function<br />

Γ h kg/(m s) Thermal diffusion coefficient<br />

Γ h,eff kg/(m s) Effective thermal diffusion coefficient<br />

Γ k,eff kg/(m s) Effective exchange coefficient for k


109<br />

Γ ɛ,eff kg/(m s) Effective exchange coefficient for ɛ<br />

Γ M,eff kg/(m s) Effective mean mass diffusion coefficient of the mixture<br />

∆t s Time step<br />

∆x m Spatial step<br />

δ k0 - Kronecker delta<br />

ɛ m 2 /s 3 Dissipation rate of turbulent kinetic energy<br />

λ W/(m K) Thermal conductivity<br />

µ kg/(m s) Dynamic viscosity<br />

µ eff kg/(m s) Effective viscosity coefficient<br />

µ f kg/(m s) Dynamic viscosity in the film<br />

µ l kg/(m s) Laminar viscosity coefficient<br />

µ t kg/(m s) Turbulent viscosity coefficient<br />

ξ - Set of internal coordinate<br />

π α - Variable in Wheeler algorithm<br />

ρ kg/m 3 Mass density<br />

ρ g kg/m 3 Gas density<br />

ρ l kg/m 3 Liquid density<br />

σ α,β<br />

Variable defined in Wheeler algorithm<br />

σ ɛ - Effective Schmidt number for ɛ<br />

φ - Relates the Spalding mass and heat transfer coefficients<br />

χ s −1 Dissipation rate of mixture fraction<br />

ψ - Evaporative flux<br />

Ω g s −1 Gas vorticity<br />

Subscripts and Superscripts<br />

Symbol Quantity<br />

d Droplet<br />

f Film<br />

g Gas<br />

l Liquid<br />

m Mixture<br />

n Index for the number of node<br />

p Parcel<br />

s Surface<br />

w Water<br />

〈 〉 Moment


110 A. Nomenclature<br />

List of Abbreviations<br />

Abbreviation<br />

Meaning<br />

CDM<br />

CFD<br />

CFL<br />

CFM<br />

CM<br />

CQMOM<br />

DDB<br />

DDM<br />

DNS<br />

DQMOM<br />

DSD<br />

ETAB<br />

LB<br />

LDA<br />

LES<br />

LHF<br />

ME<br />

MOM<br />

NDF<br />

PBE<br />

PD<br />

PDA<br />

PSIC<br />

PVP<br />

QBSM<br />

QMOM<br />

RANS<br />

RMM<br />

RTI<br />

SF<br />

TAB<br />

VOF<br />

WB<br />

continuous droplet model<br />

computational fluid dynamics<br />

Courant-Friedrichs-Lewy condition<br />

continuous formulation model<br />

method of classes<br />

conditional quadrature method of moments<br />

droplet deformation and breakup<br />

discrete droplet model<br />

direct numerical simulation<br />

direct quadrature method of moments<br />

droplet size distribution<br />

enhanced Taylor analogy breakup model<br />

lattice-Boltzmann method<br />

laser Doppler anemometry<br />

large eddy simulation<br />

local homogeneous flow model<br />

maximum entropy method<br />

method of moments<br />

number density function<br />

population balance equation<br />

product-difference algorithm<br />

phase Doppler anemometry<br />

particle-source-in-cell method<br />

polyvinylpyrrolidone<br />

quadrature based sectional method<br />

quadrature method of moments<br />

Reynolds-average Navier – Stokes equations<br />

rapid mixing model<br />

Rayleigh-Taylor instability<br />

separated flow model<br />

Taylor analogy breakup model<br />

volume of fluid method<br />

wave breakup model


B. Acknowledgements<br />

The research work was carried out at Interdisciplinary Center for Scientific Computing<br />

(IWR), University of Heidelberg, and this study is funded by German Science<br />

Foundation (DFG) through the priority program ”SPP1423”.<br />

First and foremost I would to like to express my deep sense of gratitude to my<br />

supervisor, Prof. E. Gutheil, who constantly motivated, taught, encouraged, supported<br />

and patiently guided me throughout this work. She offered me this work while I was<br />

about to finish my Master studies, and I thank her for bestowing enormous confidence<br />

on me. She perseveringly corrected the research abstracts, papers, this thesis, and her<br />

style of correction given me an opportunity to be familiar with scientific writing. She<br />

helped me a lot in day to day activities as well. Her endless kindness and incomparable<br />

support cannot be thanked adequately here.<br />

I am thankful to my co-supervisor, PD Dr. N. Dahmen, for his guidance and support<br />

in my research, and for correcting this thesis.<br />

I would like to thank E. Wimmer, Prof. G. Brenn (TU Graz), Dr. R. Wengeler (BASF<br />

Ludwigshafen) for providing the experimental data. Special thanks to Prof. G. Brenn<br />

for extending his support and comments in un<strong>der</strong>standing and correction of experimental<br />

data. Special thanks to Prof. N. Urbanetz for the fruitful discussions on mannitol/water<br />

system. I would like thank all the members of the DFG ”SPP 1423” for<br />

their support.<br />

I express my sincere thanks to all my colleagues, L. Cao, H. Großhans, R. M. Humza,<br />

Y. Hu, H. Olguin, M. Trunk, X. G. Cui, D. Urzica and E. Vogel. Thanks to N. Wenzel<br />

for his help in experimental data handling. Special thanks to H. Großhans for proof<br />

reading this thesis and translating the abstract in German. I sincerely thank E. Vogel<br />

for her kind help in all administrative matters, which made my life easier in Germany,<br />

and thanks to her also for the corrections of German abstract.<br />

I thank my family members, especially, my mother (Sulochana), beloved brother<br />

(Suren<strong>der</strong>), and sister (Swarupa) for their constant support and endless love. I wish<br />

to express special thanks to my wife (Tanuja), who encouraged me in every hard hour,<br />

and corrected this thesis with humongous patience.<br />

Last but not least, the financial support of DFG through ”SPP1423” and HGS-<br />

MathComp is gratefully acknowledged.


Bibliography<br />

[1] K. Masters: Spray drying handbook: Longman Scientific & Technical, Harlow,<br />

England (1991)<br />

[2] R. Vehring: Pharmaceutical particle engineering via spray drying. Pharmaceutical<br />

Research, 25, 999–1022 (2008)<br />

[3] A. Mescher, E. M. Littringer, R. Paus, N. A. Urbanetz, P. Walzel: Homogene Produkteigenschaften<br />

in <strong>der</strong> Sprühtrocknung durch laminare Rotationszerstäubung.<br />

Chemie Ingenieur Technik, 84, 154–159 (2012)<br />

[4] F. G. Kieviet: Modelling quality in spray drying. PhD Thesis, Department of<br />

Chemical Engineering, Eindhoven University of Technology, (1997)<br />

[5] T. A. G. Langrish, D. F. Fletcher: Prospects for the modeling and design of<br />

spray dryers in the 21st century. Drying Technology, 21(2), 197–215 (2003)<br />

[6] R. Vieweg, M. Reiher, H. Scheurlen: Kunststoff-Handbuch XI : Carl-Hanser-<br />

Verlag, München (1971)<br />

[7] F. Ullmanns: Ullmanns Encyklopädie <strong>der</strong> technischen Chemie, 4. Auflage: Urban<br />

& Schwarzenberg (1980)<br />

[8] V. Bühler: Polyvinylpyrrolidone excipients for pharmaceuticals: Povidone,<br />

crospovidone and copovidone: Springer (2005)<br />

[9] F. Fischer, S. Bauer: Polyvinylpyrrolidon. Ein Tausendsassa in <strong>der</strong> Chemie.<br />

Chemie in unserer Zeit, 43, 376–383 (2009)<br />

[10] A. Göthlich, S. Koltzenburg, G. Schornick: Funktionale Polymere im Alltag:<br />

Vielseitig. Chemie in unserer Zeit, 39, 262–273 (2005)<br />

[11] A. Mescher, J. Kamplade, P. Walzel: Personal communication (2012)<br />

[12] Mannitol: http://en.wikipedia.org/wiki/Mannitol<br />

[13] M. W. Kearsley, R. C. Deis: Sorbitol and manntiol, in Sweeteners and sugar<br />

alternatives in food technology: Blackwell Publishing Ltd, Oxford (2006)


ii<br />

BIBLIOGRAPHY<br />

[14] A. J. Hickey: Lung deposition and clearance of pharmaceutical aerosols: What<br />

can be learned from inhalation toxicology and industrial hygiene?. Aerosol Science<br />

and Technology, 18(3), 290–304 (1993)<br />

[15] K. Bechtold-Peters, H. Luessen: Pulmonary Drug Delivery: Basics, Applications<br />

and Opportunities for Small Molecules and Biopharmaceutics: ECV Editio-<br />

Cantor-Verlag (2007)<br />

[16] S. K. Tee, C. Marriott, X. M. Zeng, G. P. Martin: The use of different sugars as<br />

fine and coarse carriers for aerosolised salbutamol sulphate. International Journal<br />

of Pharmaceutics, 208, 111–123 (2000)<br />

[17] S. G. Maas, G. Schaldach, P. Walzel, N. A. Urbanetz: Tailoring dry pow<strong>der</strong><br />

inhaler performance by modifying carrier surface topography by spray drying.<br />

Atomization and Sprays, 20(9), 763–774 (2010)<br />

[18] W. Kaialy, M. N. Momin, M. D. Ticehurst, J. Murphy, A. Nokhodchi: Engineered<br />

mannitol as an alternative carrier to enhance deep lung penetration of salbutamol<br />

sulphate from dry pow<strong>der</strong> inhaler. Colloids and Surfaces B: Biointerfaces, 79,<br />

345–356 (2010)<br />

[19] G. Pilcer, K. Amighi: Formulation strategy and use of excipients in pulmonary<br />

drug delivery. International Journal of Pharmaceutics, 392, 1–19 (2010)<br />

[20] D. E. Braun, S. G. Maas, N. Zencirci, C. Langes, N. A. Urbanetz, U. J. Griesser:<br />

Simultaneous quantitative analysis of ternary mixtures of d-mannitol polymorphs<br />

by FT-Raman spectroscopy and multivariate calibration models. International<br />

Journal of Pharmaceutics, 385, 29–36 (2010)<br />

[21] E. Littringer, A. Meshcer, S. Eckhard, H. Schröttner, C. Langes, M. Fries,<br />

U. Griesser, P. Walzel, N. Urbanetz: Spray drying of Mannitol as a drug carrier–<br />

the impact of process parameters on product properties. Drying Technology, 30,<br />

114–124 (2012)<br />

[22] S. G. Maas, G. Schaldach, E. M. Littringer, A. Mescher, U. J. Griesser, D. E.<br />

Braun, P. Walzel, N. A. Urbanetz: The impact of spray drying outlet temperature<br />

on the particle morphology of mannitol. Pow<strong>der</strong> Technology, 213, 27–35 (2011)<br />

[23] S. Sundaram, L. R. Collins: Collision statistics in an isotropic particle-laden<br />

turbulent suspension. Part 1: direct numerical simulations. Journal of Fluid Mechanics,<br />

335, 75–109 (1997)


BIBLIOGRAPHY<br />

iii<br />

[24] K. D. Squires, J. K. Eaton: Measurements of particle dispersion obtained from<br />

direct numerical simulations of isotropic turbulence. Journal of Fluid Mechanics,<br />

226, 1–35 (1991)<br />

[25] T. Baritaud, T. Poinsot, Baum, M.: Direct numerical simulation for turbulent<br />

reacting flows: Paris: Editions Technip (1996)<br />

[26] K. Luo, H. Pitsch, M. G. Pai, O. Desjardins: Direct numerical simulations and<br />

analysis of three dimensional n-heptane spray flames in a model swirl combustor.<br />

Proceedings of the Combustion Institute, 33, 2143–2152 (2011)<br />

[27] P. Moin, K. Mahesh: DNS: a tool for turbulence research. Annual Review of<br />

Fluid Mechanics, 30, 539–578 (1998)<br />

[28] D. L. Marchisio, R. O. Fox: Multiphase reacting flows: Modelling and simulation<br />

CISM International Centre for Mechanical Sciences, No. 492: Springer London,<br />

Limited (2007)<br />

[29] M. Boivin, O. Simonin, K. D. Squires: On the prediction of gas-solid flows with<br />

two-way coupling using large eddy simulation. Physics of Fluids, 12, 2080–2090<br />

(2000)<br />

[30] P. Givi: Quality assessment of the filtered density function for large eddy simulation.<br />

1st workshop on Quality Assessment of Unsteady Methods for Turbulent<br />

Combustion Prediction and Validation, June 16-17, Darmstadt (Seeheim-<br />

Jugenheim), Germany (2005)<br />

[31] J. Janicka, A. Sadiki: Large eddy simulation of turbulent combustion systems.<br />

Proceedings of the Combustion Institute, 30, 537–547 (2005)<br />

[32] J. A. Langford, R. D. Moser: Optimal LES formulations for isotropic turbulence.<br />

Journal of Fluid Mechanics, 398, 321–346 (1999)<br />

[33] J. Meyers, B. Geurts, P. Sagaut: Quality and reliability of large-eddy simulations.<br />

Vol. 12: ERCOFTAC series, Springer (2008)<br />

[34] Q. Z. Wang, K. D. Squires: Large eddy simulation of particle-laden turbulent<br />

channel flow. Physics of Fluids, 8(5), 1207–1223 (1996)<br />

[35] J. K. Duckowicz: A particle-fluid numerical model for liquid sprays. Journal of<br />

Computational Physics, 35(2), 229–253 (1980)


iv<br />

BIBLIOGRAPHY<br />

[36] M. A. Pakhomov: RANS simulation of effect of evaporating droplets on a turbulent<br />

heat transfer in a mist flow in a sudden pipe expansion. Computational<br />

Thermal Sciences, 2(4), 311–321 (2010)<br />

[37] S. B. Pope: Turbulent flows: Cambridge University Press (2000)<br />

[38] C. W. Hirt, B. D. Nichols: Volume of fluid method for the dynamics of free<br />

boundaries. Journal of Computational Physics, 39(1), 201–225 (1981)<br />

[39] S. Succi: The lattice Boltzmann equation: For fluid dynamics and beyond, Numerical<br />

Mathematics and Scientific Computation: Clarendon Press, Oxford (2001)<br />

[40] S. E. Elgobashi, T. W. Abou-Arab: A two-equation turbulence model for twophase<br />

flows. Physics of Fluids, 26, 931–938 (1983)<br />

[41] D. Ramakrishna: Population balances: Theory and applications to particulate<br />

systems in engineering: Academic Press, London (2000)<br />

[42] F. A. Williams: Spray combustion and atomization. Physics of Fluids, 1, 541–545<br />

(1958)<br />

[43] S. K. Freidlan<strong>der</strong>: Smoke, dust, and haze: Oxford University Press, Oxford<br />

(2000)<br />

[44] F. Gelbard, J. H. Seinfeld: Numerical solution of the dynamic equation for particulate<br />

systems. Journal of Computational Physics, 28, 357–375 (1978)<br />

[45] F. Laurent, M. Massot: Multi-fluid modeling of laminar poly-dispersed spray<br />

flames: origin, assumptions and comparison of sectional and sampling methods.<br />

Combustion Theory and Modelling, 5(4), 537–572 (2001)<br />

[46] F. Laurent, M. Massot, P. Villedieu: Eulerian multi-fluid modeling for the numerical<br />

simulation of coalescence in polydisperse dense liquid sprays. Journal of<br />

Computational Physics, 194(2), 505–543 (2004)<br />

[47] M. Massot, S. de Chaisemartin, L. Freret, D. Kah, F. Laurent: Modeling and<br />

computations of nanoparticles in fluid flows. Eulerian Multi-Fluid Models: Modeling<br />

and Numerical Methods, RTO - Lecture Series du von Karman Institute,<br />

1–86, (2009)<br />

[48] D. L. Marchisio, R. D. Vigil, R. O. Fox: Quadrature method of moments for<br />

aggregation–breakage processes. Journal of Colloid and Interface Science, 258,<br />

322–334 (2003)


BIBLIOGRAPHY<br />

v<br />

[49] D. L. Marchisio, J. T. Pikturna, R. O. Fox, R. D. Vigil, A. A. Barresi: Quadrature<br />

method of moments for population-balance equations. AIChE Journal, 49(5),<br />

1266–1276 (2003)<br />

[50] D. L. Marchisio, R. D. Vigil, R. O. Fox: Implementation of the quadrature<br />

method of moments in CFD codes for aggregation–breakage problems. Chemical<br />

Engineering Science, 58, 3337–3351 (2003)<br />

[51] R. McGraw: Description of aerosol dynamics by the quadrature method of moments.<br />

Aerosol Science and Technology, 27(2), 255–265 (1997)<br />

[52] A. Passalacqua, R. O. Fox, R. Garg, S. Subramaniam: A fully coupled<br />

quadrature-based moment method for dilute to mo<strong>der</strong>ately dilute fluid–particle<br />

flows. Chemical Engineering Science, 65, 2267–2283 (2010)<br />

[53] D. L. Marchisio, R. O. Fox: Solution of population balance equations using the<br />

direct quadrature method of moments. Journal of Aerosol Science, 36, 43–73<br />

(2005)<br />

[54] R. O. Fox: Bivariate direct quadrature method of moments for coagulation<br />

and sintering of particle populations. Journal of Aerosol Science, 37, 1562–1580<br />

(2010)<br />

[55] G. Dufour, P. Villedieu: A second-or<strong>der</strong> multi-fluid model for evaporating sprays.<br />

ESAIM: Mathematical Modelling and Numerical Analysis (M2AN), 39(5), 931–<br />

963 (2005)<br />

[56] J. Gimbun, Z. K. Nagy, C. D. Rielly: Simultaneous quadrature method of moments<br />

for the solution of population balance equations, using a differential algebraic<br />

equation framework. Industrial and Engineering Chemistry Research, 48,<br />

7798–7812 (2009)<br />

[57] R. Grosch, H. Briesen, W. Marquardt, M. Wulkow: Generalization and numerical<br />

investigation of QMOM. AIChE Journal, 53(1), 207–227 (2007)<br />

[58] C. Laurent, G. Lavergne, P. Villedieu: Quadrature method of moments for modeling<br />

multi-component spray vaporization. International Journal of Multiphase<br />

Flow, 36, 51–59 (2010)<br />

[59] J. Akroyd, A. J. Smith, L. R. McGlashan, M. Kraft: Comparison of the stochastic<br />

fields method and DQMoM-IEM as turbulent reaction closures. Chemical<br />

Engineering Science, 65, 5429–5441 (2010)


vi<br />

BIBLIOGRAPHY<br />

[60] R. O. Fox, L. Laurent, M. Massot: Numerical simulation of polydisperse, dense<br />

liquid spray in an Eulerian framework: Direct quadrature method of moments<br />

and multi-fluid method. Journal of Computational Physics, 227(6), 3058–3088<br />

(2008)<br />

[61] J. Madsen: Computational and experimental study of sprays from the breakup<br />

of water sheets. PhD Thesis, The Faculty of Engineering and Science, Aalborg<br />

University, Denmark, (2006)<br />

[62] B. Abramzon, W. A. Sirignano: Droplet vaporization model for spray combustion<br />

calculations. International Journal of Heat and Mass Transfer, 32, 1605–1618<br />

(1989)<br />

[63] M. Friedrich, B. Weigand: Eulerian multi-fluid simulation of polydisperse dense<br />

liquid sprays by the direct quadrature method of moments. 10th Triennial International<br />

Conference on Liquid Atomization and Spray Systems (ICLASS), Kyoto,<br />

Japan (2006)<br />

[64] L. Schnei<strong>der</strong>: A concise moment method for unsteady polydisperse sprays. PhD<br />

Thesis, Fachbereich Maschinenbau, Technischen Universität Darmstadt, (2009)<br />

[65] D. Kah, F. Laurent, M. Massot, S. Jay: A high or<strong>der</strong> moment method simulating<br />

evaporation and advection of a polydisperse liquid spray. Journal of Computational<br />

Physics, 231, 394–422 (2012)<br />

[66] L. R. Mead, N. Papanicolaou: Maximum entropy in the problem of moments.<br />

Journal of Mathematical Physics, 25(8), 2404–2417 (1984)<br />

[67] M. R. Archambault: A maximum entropy moment closure approach to modeling<br />

the evolution of spray flows. PhD Thesis, Stanford University, USA, (1999)<br />

[68] R. M. Humza: Numerical investigation of non-reactive and reactive turbulent<br />

spray flows. PhD Thesis, University of Heidelberg, Germany, (2013)<br />

[69] H. Großhans: Large eddy simulation of atomizing sprays. PhD Thesis, Lund<br />

University, Sweden, (2013)<br />

[70] G. M. Faeth, L.-P. Hsiang, P.-K. Wu: Structure and breakup properties of sprays.<br />

International Journal of Multiphase Flow, 21, 99–127 (1995)<br />

[71] D. L. Marchisio, R. O. Fox: Computational models for polydisperse particulate<br />

and multiphase systems: Cambridge University Press, Cambridge (2013)


BIBLIOGRAPHY<br />

vii<br />

[72] G. M. Faeth: Evaporation and combustion of sprays. Progress in Energy and<br />

Combustion Science, 9, 1–76 (1983)<br />

[73] M. W. Thring, M. P. Newby: In: Fourth Symposium (International) on Combustion:<br />

Combustion and Detonation Waves: Williams & Wilkens Comp. (1953)<br />

[74] Khalil, Whitelaw: In: Sixteenth Symposium (International) on Combustion: The<br />

Combustion Institute, Pittsburg (1977)<br />

[75] C.-P. Mao, G. A. Szekely, G. M. Faeth: Evaluation of a locally homogeneous<br />

flow model of spray combustion Journal of Energy, 4(2), 78–87 (1980)<br />

[76] C. T. Crowe, M. P. Sharma, D. E. Stock: The particle-source-in cell (PSI-Cell)<br />

model for gas-droplet flows. Journal of Fluids Engineering, 99, 325–332 (1977)<br />

[77] J. T. Jurewicz, D. T. Stock, C. T. Crowe: The effect of turbulent diffusion on<br />

gas particle flow in an electric Field. First Symposium on Turbulent Shear Flows,<br />

12.27–12.33, April 18–20, University Park, PA (1977)<br />

[78] C. T. Crowe: A numerical model for the gas droplet flow field near an atomizer.<br />

First International Conference on Liquid Atomization and Spray Systems,<br />

August 27–31, Tokyo, Japan (1978)<br />

[79] F. A. Williams: Combustion theory, Addison-Wesley series in engineering science:<br />

Addison-Wesley Pub. Co. (1965)<br />

[80] A. D. Gosman, R. J. R. Johns: Computer analysis of fuel-air mixing in directinjection<br />

engines. SAE paper, 800091 (1980)<br />

[81] F. H. Harlow, A. A. Amsden: Numerical calculation of multiphase fluid flow.<br />

Journal of Computational Physics, 17(1), 19–52 (1975)<br />

[82] F. H. Harlow, A. A. Amsden: Flow of interpenetrating material phases. Journal<br />

of Computational Physics, 18(4), 440–464 (1975)<br />

[83] Z. Han, R. D. Reitz, P. J. Claybaker, C. J. Rutland, J. Yang, R. W. An<strong>der</strong>son:<br />

Modeling the effects of intake flow structures on fuel/air mixing in a directinjected<br />

spark-ignition engine. SAE paper, 961192 (1996)<br />

[84] S. Subramaniam, P. J. O’rourke: Numerical convergence of the KIVA-3 code for<br />

sprays and its implications for modeling. Atomization and Sprays, 8(4), 453–469<br />

(1998)


viii<br />

BIBLIOGRAPHY<br />

[85] M. Garcia, Y. Sommerer, T. Schonfeld, T. Poinsot: Assessment of Euler-Euler<br />

and Euler-Lagrange strategies for reactive large-eddy simulations. In 1st Workshop<br />

INCA–2005 , 65–74, SNECMA Villaroche, France (2005)<br />

[86] E. Riber, M. Garcia, V. Moureau, H. Pitsch, O. Simonin, T. Poinsot: Evaluation<br />

of numerical strategies for LES of two-phase reacting flows. Proceedings of<br />

the Summer Program–2006 , 197–211, Center for Turbulence Research, NASA<br />

Ames/Stanford Univ. (2006)<br />

[87] L. Huang, K. Kumar, A. S. Mujumdar: Simulation of a spray dryer fitted with<br />

a rotary disk atomizer using a three-dimensional computational fluid dynamic<br />

model. Drying Technology, 22, 1489–1515 (2004)<br />

[88] D. I. Kolaitis, M. A. Founti: A comparative study of numerical models for<br />

Eulerian-Lagrangian simulations of turbulent evaporating sprays. International<br />

Journal of Heat and Fluid Flow, 27(3), 424–435 (2006)<br />

[89] R. Kuriakose, C. Anandharamakrishnan: Computational fluid dynamics (CFD)<br />

applications in spray drying of food products. Trends in Food Science & Technology,<br />

21, 383–398 (2010)<br />

[90] J. J. Nijdam, B. Guo, D. F. Fletcher, T. A. G. Langrish: Lagrangian and Eulerian<br />

models for simulating turbulent dispersion and coalescence of droplets within a<br />

spray. Applied Mathematical Modelling, 30, 1196–1211 (2006)<br />

[91] A. A. Mostafa, H. C. Mongia: On the modeling of turbulent evaporating sprays<br />

- Eulerian versus Lagrangian approach. International Journal of Heat and Mass<br />

Transfer, 30, 2583–2593 (1987)<br />

[92] M. Mezhericher, A. Levy, I. Borde: Probabilistic hard-sphere model of binary<br />

particle–particle interactions in multiphase flow of spray dryers. International<br />

Journal of Multiphase Flow, 43, 22–38 (2012)<br />

[93] M. Sommerfeld, G. Kohnen, H. H. Qiu: Spray evaporation in turbulent flow:<br />

Numerical calculations and detailed experiments by phase-Doppler anemometry.<br />

Oil & Gas Science and Technology-revue De L Institut Francais Du Petrole,<br />

48(6), 677–695 (1993)<br />

[94] L. Boltzmann: Weitere Studien über das Wärmegleichgewicht unter Gasmolekülen.<br />

Sitzungsberichte <strong>der</strong> Akademie <strong>der</strong> Wissenschaften, Wien, 66, 275–<br />

370 (1872)


BIBLIOGRAPHY<br />

ix<br />

[95] C. Cercignani, R. Illner, M. Pulvirenti: The mathematical theory of dilute gases:<br />

Springer, Berlin (1994)<br />

[96] F. Bouchet: Advances in kinetic theory and computing: Selected papers: World<br />

Scientific, Singapore (1994)<br />

[97] L. Fréret, S. de Chaisemartin, J. Reveillon, F. Laurent, M. Massot: Eulerian models<br />

and three-dimensional numerical simulation of polydisperse sprays. 7th International<br />

Conference on Multiphase Flow (ICMF), 1–7, May 30–June 4, Tampa,<br />

Florida (2010)<br />

[98] M. Massot, F. Laurent, D. Kah, S. de Chaisemartin: A robust moment method<br />

for evaluation of the disappearance rate of evaporating sprays. SIAM Journal on<br />

Applied Mathematics, 70(8), 3203–3234 (2010)<br />

[99] S. Kumar, D. Ramakrishna: On the solution of population balance equations by<br />

discretization I. A fixed pivot technique. Chemical Engineering Science, 51(8),<br />

1311–1332 (1996)<br />

[100] S. Kumar, D. Ramakrishna: On the solution of population balance equations<br />

by discretization II. A moving pivot technique. Chemical Engineering Science,<br />

51(8), 1333–1342 (1996)<br />

[101] S. Kumar, D. Ramakrishna: On the solution of population balance equations<br />

by discretization III. Nucleation, growth and aggregation of particles. Chemical<br />

Engineering Science, 52(24), 4659–4679 (1997)<br />

[102] M. J. Hounslow, R. L. Ryall, V. R. Marshall: A discretized population balance<br />

for nucleation, growth, and aggregation. AIChE Journal, 34, 1821–1832 (1988)<br />

[103] M. Vanni: Approximate population balance equations for aggregation-breakage<br />

processes. Journal of Colloid and Interface Science, 221, 143–160 (2000)<br />

[104] Y. Tambour: A sectional model for evaporation and combustion of sprays of<br />

liquid fuels. Israel Journal of Technology, 18, 47–56 (1980)<br />

[105] F. Laurent: Analyse numérique d’une méthode multi-fluide Eulérienne pour la<br />

description de sprays qui s’évaporent. C. R. Acad. Sci. Paris Ser. I , 334, 417–422<br />

(2002)<br />

[106] R. G. Gordon: Error bounds in equilibrium statistical mechanics. Journal of<br />

Mathematical Physics, 9, 655–663 (1968)


x<br />

BIBLIOGRAPHY<br />

[107] H. M. Hulbert, S. Katz: Some problems in particle technology: A statistical<br />

mechanical formulation. Chemical Engineering Science, 19, 555–574 (1964)<br />

[108] R. B. Diemer, J. H. Olson: A moment methodology for coagulation and breakage<br />

problems: Part 1 – analytical solution of the steady-state population balance.<br />

Chemical Engineering Science, 57, 2193–2209 (2002)<br />

[109] R. B. Diemer, J. H. Olson: A moment methodology for coagulation and breakage<br />

problems: Part 2 – moment models and distribution reconstruction. Chemical<br />

Engineering Science, 57, 2211–2228 (2002)<br />

[110] R. B. Diemer, J. H. Olson: A moment methodology for coagulation and breakage<br />

problems: Part 3 – generalized daughter distribution functions. Chemical<br />

Engineering Science, 57, 4187–4198 (2002)<br />

[111] R. A. LaViolette, R. A. Berry, R. McGraw: Homogeneous nucleation of metals in<br />

a plasma-quench reactor. Plasma Chemistry and Plasma Processing, 16, 249–264<br />

(1996)<br />

[112] R. McGraw, J. H. San<strong>der</strong>s: A condensation feedback mechanism for oscillatory<br />

nucleation and growth. Aerosol Science and Technology, 3, 367–380 (1984)<br />

[113] S. E. Pratsinis: Simultaneous nucleation, condensation, and coagulation in<br />

aerosol reactors. Journal of Colloid and Interface Science, 124, 416–427 (1988)<br />

[114] B. J. Jurcik, J. R. Brock: Numerical simulation of particle formation and growth<br />

in rapidly expanding axisymmetric flows. Journal of Physical Chemistry, 97,<br />

323–331 (1993)<br />

[115] R. O. Fox: A quadrature-based third-or<strong>der</strong> moment method for dilute gas-particle<br />

flows. Journal of Computational Physics, 227, 6313–6350 (2008)<br />

[116] C. Yoon, R. McGraw: Representation of generally mixed multivariate aerosols by<br />

the quadrature method of moments: I. Statistical foundation. Journal of Aerosol<br />

Science, 35, 561–576 (2004)<br />

[117] C. Yoon, R. McGraw: Representation of generally mixed multivariate aerosols<br />

by the quadrature method of moments: II. Aerosol dynamics. Journal of Aerosol<br />

Science, 35, 577–598 (2004)<br />

[118] C. Yuan, R. O. Fox: Conditional quadrature method of moments for kinetic<br />

equations. Journal of Computational Physics, 230(22), 8216–8246 (2011)


BIBLIOGRAPHY<br />

xi<br />

[119] R. O. Fox: Higher-or<strong>der</strong> quadrature-based moment methods for kinetic equations.<br />

Journal of Computational Physics, 228, 7771–7791 (2009)<br />

[120] H. Dette, W. J. Studden: The theory of canonical moments with applications in<br />

statistics, probability, and analysis: Wiley, New York (1997)<br />

[121] R. Fan, D. L. Marchisio, R. O. Fox: Application of the direct quadrature method<br />

of moments to polydisperse gas-solid fluidized beds. Pow<strong>der</strong> Technology, 139,<br />

7–20 (2004)<br />

[122] T. L. Chan, Y. H. Liu, C. K. Chan: Direct quadrature method of moments for<br />

the exhaust particle formation and evolution in the wake of the studied ground<br />

vehicle. Journal of Aerosol Science, 41, 553–568 (2010)<br />

[123] B. Selma, R. Bannari, P. Proulx: Simulation of bubbly flows: Comparison<br />

between direct quadrature method of moments (DQMOM) and method of<br />

classes (CM). Chemical Engineering Science, 65(6), 1925–1941 (2010)<br />

[124] L. Mazzei, D. L. Marchisio, P. Lettieri: Direct quadrature method of moments<br />

for the mixing of inert polydisperse fluidized pow<strong>der</strong>s and the role of numerical<br />

diffusion. Industrial and Engineering Chemistry Research, 49(11), 5141–5152<br />

(2010)<br />

[125] S. Ali, A. Vikhansky, T. Løvås: Direct quadrature conditional moment closure<br />

for modelling of turbulent combustion. Flow, Turbulence and Combustion, 87,<br />

493–509 (2011)<br />

[126] A. Bruyat, C. Laurent, O. Rouzaud: Direct quadrature method of moments<br />

for multicomponent droplet spray vaporization. 7th International Conference on<br />

Multiphase Flow (ICMF), May 30–June 4, Tampa, Florida (2010)<br />

[127] W. Gumprich, A. Sadiki, M. Sommerfeld: 3D simulation and analysis of dense<br />

turbulent sprays using an Eulerian approach coupled to the DQMOM. 13th International<br />

Conference on Numerical Combustion (ICNC), April 27–29, Corfu,<br />

Greece (2011)<br />

[128] W. Gumprich, A. Sadiki, M. Sommerfeld: Numerical investigation of dense turbulent<br />

sprays with an Eulerian approach and DQMOM. Proceedings in Applied<br />

Mathematics and Mechanics (PAMM), 11(1), 599–600 (2011)<br />

[129] W. Gumprich, A. Sadiki: Numerical study of dense turbulent sprays using a<br />

coupling of the direct quadrature method of moments with an Eulerian multi-size


xii<br />

BIBLIOGRAPHY<br />

moment model. 12th Triennial International Conference on Liquid Atomization<br />

and Spray Systems (ICLASS), 1–8, September 2–6, Heidelberg, Germany (2012)<br />

[130] C. Hollmann, E. Gutheil: Modeling of turbulent spray diffusion flames including<br />

detailed chemistry. Proceedings of Combustion Institute, 26(1), 1731–1738 (1996)<br />

[131] C. Hollmann, E. Gutheil: Flamelet-modeling of turbulent spray diffusion flames<br />

based on a laminar spray flame library. Combustion Science and Technology,<br />

135(1-6), 175–192 (1998)<br />

[132] E. Gutheil, F. A. Williams: A numerical and asymptotic investigation of structures<br />

of hydrogen-air diffusion flames at pressures and temperatures of high-speed<br />

combustion. Proceedings of Combustion Institute, 23(1), 513–521 (1991)<br />

[133] C. K. Law: Heat and mass transfer in combustion: Fundamental concepts and<br />

analytical techniques. Progress in Energy and Combustion Science, 10, 295–318<br />

(1984)<br />

[134] W. A. Sirignano: Fluid dynamics of sprays. Journal of Fluids Engineering,<br />

115(3), 345–378 (1992)<br />

[135] W. Feller: An introduction to probability theory and its applications, Wiley mathematical<br />

statistics series: Wiley (1966)<br />

[136] R. A. Sack, A. F. Donovan: An algorithm for Gaussian quadrature given modified<br />

moments. Numerische Mathematik, 18, 465–478 (1971)<br />

[137] R. O. Fox: Optimal moment sets for multivariate direct quadrature method of<br />

moments. Industrial and Engineering Chemistry Research, 48(21), 9686–9696<br />

(2009)<br />

[138] D. Choi, L. Schnei<strong>der</strong>, N. Spyrou, A. Sadiki, J. Janicka: Evaporation of tetralin<br />

spray with direct quadrature method of moments and Eulerian multi-fluid<br />

method. 7th International Conference on Multiphase Flow (ICMF), 1–8, May<br />

30–June 4, Tampa, Florida (2010)<br />

[139] I. Langmuir: The Evaporation of Small Spheres. Physical Review, 12(5), 368–<br />

370 (1918)<br />

[140] W. A. Sirignano: Fluid dynamics and transport of droplets and sprays: Cambridge<br />

University Press, New York (1999)<br />

[141] N. A. Chigier: Energy, combustion, and environment: McGraw-Hill, New York<br />

(1981)


BIBLIOGRAPHY<br />

xiii<br />

[142] R. Clift, J. R. Grace, M. E. Weber: Bubbles, drops, and particles: Academic<br />

Press, New York (1978)<br />

[143] I. Glassman: Combustion: Academic Press, New York, 2 nd Edition (1987)<br />

[144] Lefebvre: Atomization and sprays,: Hemisphere, Washington, DC (1989)<br />

[145] F. A. Williams: Combustion theory: The fundamental theory of chemically reacting<br />

flow systems: Benjamin/Cummings, Menlo Park, CA, 2 nd Edition (1985)<br />

[146] C. K. Law: Recent advances in droplet vaporization and combustion. Progress<br />

in Energy and Combustion Science, 8(3), 171–201 (1982)<br />

[147] S. S. Sazhin: Advanced models of fuel droplet heating and evaporation. Progress<br />

in Energy and Combustion Science, 32, 162–214 (2006)<br />

[148] P. Atkins, J. D. Paula: Atkins’ physical chemistry: Oxford Higher Education<br />

(2001)<br />

[149] D. H. Charlesworth, W. R. Marshall Jr: Evaporation from drops containing<br />

dissolved solids. AIChE Journal, 6, 9–23 (1960)<br />

[150] Y. Sano, R. B. Keey: The drying of a spherical particle containing colloidal<br />

material into a hollow sphere. Chemical Engineering Science, 37, 881–889 (1982)<br />

[151] S. Nesic, J. Vodnik: Kinetics of droplet evaporation. Chemical Engineering Science,<br />

46, 527–537 (1991)<br />

[152] G. Brenn: Concentration fields in drying droplets. Chemical Engineering & Technology,<br />

27, 1252–1258 (2004)<br />

[153] P. Seydel, J. Blömer, J. Bertling: Modeling particle formation at spray drying<br />

using population balances. Drying Technology, 24(2), 137–146 (2006)<br />

[154] S. El Golli, J. Bricard, P.-Y. Turpin, C. Treiner: The evaporation of saline<br />

droplets. Journal of Aerosol Science, 5, 273–292 (1974)<br />

[155] X. D. Chen: Heat-mass transfer and structure formation during drying of single<br />

food droplets. Drying Technology, 22, 179–190 (2004)<br />

[156] M. Farid: A new approach to modelling of single droplet drying. Chemical Engineering<br />

Science, 58, 2985–2993 (2003)<br />

[157] S. Nesic: The evaporation of single droplets: Experiments and modelling. Drying,<br />

89, 386–393 (1989)


xiv<br />

BIBLIOGRAPHY<br />

[158] C. S. Handscomb, M. Kraft, A. E. Bayly: A new model for the drying of droplets<br />

containing suspended solids. Chemical Engineering Science, 64, 628–637 (2009)<br />

[159] N. Tsapis, E. R. Dufresne, S. S. Sinha, C. S. Riera, J. W. Hutchinson, L. Mahadevan,<br />

D. A. Weitz: Onset of buckling in drying droplets of colloidal suspensions.<br />

Physical Review Letters, 94, 018302 (2005)<br />

[160] B. Golman: Modeling of the drying kinetics of slurry droplet in spray drying.<br />

ASEAN Journal of Chemical Engineering, 11(2), 1–7 (2011)<br />

[161] H. W. Cheong, G. V. Jeffreys, C. J. Mumford: A receding interface model for<br />

the drying of slurry droplets. AIChE Journal, 32, 1334–1346 (1986)<br />

[162] M. Mezhericher, A. Levy, I. Borde: Theoretical models of single droplet drying<br />

kinetics: a review. Drying Technology, 28, 278–293 (2010)<br />

[163] B. Adhikari, T. Howes, B. R. Bhandari, V. Truong: Experimental studies and<br />

kinetics of single drop drying and their relevance in drying of sugar-rich foods:<br />

A review. International Journal of Food Properties,, 3, 323–351 (2000)<br />

[164] R. Vehring, W. R. Foss, D. Lechuga-Ballesteros: Particle formation in spray<br />

drying. Journal of Aerosol Science, 38, 728–746 (2007)<br />

[165] Y. Sugiyama, R. J. Larsen, J.-W. Kim, D. A. Weitz: Buckling and crumpling of<br />

drying droplets of colloid-polymer suspensions. Langmuir, 22, 6024–6030 (2006)<br />

[166] A. L. Yarin, G. Brenn, O. Kastner, C. Tropea: Drying of acoustically levitated<br />

droplets of liquid-solid suspensions: Evaporation and crust formation. Physics of<br />

Fluids, 14, 2289–2298 (2002)<br />

[167] K. Alexan<strong>der</strong>, C. J. King: Factors governing surface morphology of spray-dried<br />

amorphous substances. Drying Technology, 3, 321–348 (1985)<br />

[168] T. M. El-Sayed, D. A. Wallack, C. J. King: Changes in particle morphology<br />

during drying of drops of carbohydrate solutions and food liquids. 1. Effects of<br />

composition and drying conditions. Industrial and Engineering Chemistry Research,<br />

29, 2346–2354 (1990)<br />

[169] D. A. Wallack, T. M. El-Sayed, C. J. King: Changes in particle morphology<br />

during drying of drops of carbohydrate solutions and food liquids. 2. Effects<br />

on drying rate. Industrial and Engineering Chemistry Research, 29, 2354–2357<br />

(1990)


BIBLIOGRAPHY<br />

xv<br />

[170] G. Brenn, L. J. Deviprasath, F. Durst: Computations and experiments on the<br />

evaporation of multi-component droplets. 9th Triennial International Conference<br />

on Liquid Atomization and Spray Systems (ICLASS), Sorrento, Italy (2003)<br />

[171] O. Kastner: Theoretische und experimentelle Untersuchungen zum Stoffübergang<br />

von Einzeltropfen in einem akustischen Rohrlevitator. PhD Thesis, Friedrich-<br />

Alexan<strong>der</strong> University of Erlangen-Nürnberg, Germany, (2001)<br />

[172] S. R. Gopireddy, E. Gutheil: Numerical simulation of evaporation and drying of<br />

a bi-component droplet. International Journal of Heat and Mass Transfer, 66,<br />

404–411 (2013)<br />

[173] M. Mezhericher, A. Levy, I. Borde: Theoretical drying model of single droplets<br />

containing insoluble or dissolved solids. Drying Technology, 25, 1035–1042 (2007)<br />

[174] C. T. Crowe, M. Sommerfeld, Y. Tsuji: Multiphase flows with droplets and particles:<br />

CRC Press (1998)<br />

[175] G. K. Batchelor: An introduction to fluid mechanics: University Press London<br />

(1967)<br />

[176] L. Schiller, A. Z. Neumann: A drag coefficient correlation. VDI Zeitschrift, 77,<br />

318–320 (1933)<br />

[177] J. O. Hinze: Turbulence: McGraw-Hill, New York (1975)<br />

[178] M. Pilch, C. A. Erdman: Use of breakup time data and velocity history data to<br />

predict the maximum size of stable fragments for acceleration-induced breakup of<br />

a liquid drop. International Journal of Multiphase Flow, 13(6), 741–757 (1987)<br />

[179] R. D. Reitz: Modeling atomization processes in high-pressure vaporizing sprays.<br />

Atomisation and Spray Technology, 3, 309–337 (1987)<br />

[180] P. J. O’Rourke, A. A. Amsden: The TAB method for numerical calculation of<br />

spray droplet breakup. SAE Paper, 872089 (1987)<br />

[181] F. X. Tanner: Liquid jet atomization and droplet breakup modeling of nonevaporating<br />

diesel fuel sprays. SAE Tchnical Paper, 970050 (1997)<br />

[182] R. D. Reitz, Z. Liu, S. S. Hwang: Breakup mechanisms and drag coefficients of<br />

high-speed vaporizing liquid drops. Atomization and Sprays, 6, 353–376 (1996)<br />

[183] E. A. Ibrahim, H. Q. Yang, A. J. Przekwas: Modeling of spray droplets deformation<br />

and breakup. AIAA: Journal of Propulsion and Power, 9, 651–654 (1993)


xvi<br />

BIBLIOGRAPHY<br />

[184] G. H. Ko, H. S. Ryou: Modeling of droplet collision-induced breakup process.<br />

International Journal of Multiphase Flow, 31, 723–738 (2005)<br />

[185] C. Planchette, G. Brenn: Liquid encapsulation by binary collisions of immiscible<br />

liquid drops. 11th Triennial International Conference on Liquid Atomization and<br />

Spray Systems (ICLASS), Vail, Colorado USA (2009)<br />

[186] P. J. O’Rourke: Collective drop effects on vaporizing liquid sprays. PhD Thesis,<br />

Mechanical and Aerospace Engineering, Princeton University, (1981)<br />

[187] P. Villedieu, J. Hylkema: Une méthode particulaire aléatoire reposant sur une<br />

équation cinétique pour la simulation numérique des sprays denses de gouttelettes<br />

liquides. Mathematical Problems in Mechanics (Numerical Analysis), C. R. Acad.<br />

Sci. Paris, t. 328 , 1, 323–328 (1997)<br />

[188] R. Courant, K. Friedrichs, H. Lewy: Über die partiellen Differenzengleichungen<br />

<strong>der</strong> mathematischen Physik. Mathematische Annalen, 100, 32–74 (1928)<br />

[189] B. Perthame: Kinetic formulation of conservation laws: Oxford University Press,<br />

Oxford (2002)<br />

[190] O. Desjardins, R. O. Fox, P. Villedieu: A quadrature-based moment method for<br />

dilute fluid-particle flows. Journal of Computational Physics, 227, 2514–2539<br />

(2008)<br />

[191] S. R. Gopireddy, R. M. Humza, E. Gutheil: Modeling and simulation of evaporating<br />

spray flows with coalescence in an Eulerian framework. Chemie Ingenieur<br />

Technik, 84(3), 349–356 (2012)<br />

[192] J. An<strong>der</strong>son: Computational fluid dynamics: The basics with applications<br />

McGraw-Hill Series in Aeronautical and Aerospace Engineering: McGraw-Hill<br />

Education, New York (1995)<br />

[193] B. Fornberg: Generation of finite difference formulas on arbitrarily spaced grids.<br />

Mathematics of Computation, 51, 699–706 (1988)<br />

[194] W. Gautschi: Numerical analysis: An introduction: Birkhäuser, Boston (MA)<br />

(1997)<br />

[195] S. V. Patankar: Numerical heat transfer and fluid flow. Series in computational<br />

methods in mechanics and thermal sciences: Taylor & Francis (1980)<br />

[196] W. R. Marshall: Atomization and spray drying, Chemical engineering progress<br />

monograph series: American Institute of Chemical Engineers (1954)


BIBLIOGRAPHY<br />

xvii<br />

[197] M. Stieß: Mechanische Verfahrenstechnik-Partikeltechnologie 1 : Springer, New<br />

York, 3 rd Edition (2008)<br />

[198] S. R. Gopireddy, R. M. Humza, E. Gutheil: Numerical simulation of an evaporating<br />

spray using direct quadrature method of moments. 24th European Conference<br />

on Liquid Atomization and Spray Systems (ILASS), 1–8, September 5–7, Estoril,<br />

Portugal (2011)<br />

[199] S. R. L. Werner: Air suspension coating of dairy pow<strong>der</strong>s: A micro-level process<br />

approach. PhD Thesis, Massey University, Palmerston North, New Zealand,<br />

(2005)<br />

[200] S. R. Gopireddy, R. M. Humza, E. Gutheil: Application of direct quadrature<br />

method of moments in modeling of evaporating spray flows. 5. Symposium Produktgestaltung<br />

in <strong>der</strong> Partikeltechnologie, Band 5 , 275–284, May 19–20, Fraunhofer<br />

ICT, Pfinztal, Germany (2011)<br />

[201] T. Allen: Particle size measurement: Volume 1: Pow<strong>der</strong> sampling and particle<br />

size measurement, Pow<strong>der</strong> Technology Series: Springer (1996)<br />

[202] S. R. Gopireddy, R. M. Humza, E. Gutheil: Numerical simulation of evaporating<br />

sprays in a convective flow field. SPRAY - 2012 , 1–8, May 21–22, Berlin,<br />

Germany (2012)<br />

[203] S. R. Gopireddy, R. M. Humza, E. Gutheil: Numerical simulation of evaporating<br />

sprays in a convective flow field. 12th Triennial International Conference on Liquid<br />

Atomization and Spray Systems (ICLASS), 1–8, September 2–6, Heidelberg,<br />

Germany (2012)<br />

[204] A. Tratnig, G. Brenn: Drop size spectra in sprays from pressure-swirl atomizers.<br />

International Journal of Multiphase Flow, 36(5), 349–363 (2010)<br />

[205] S. R. Gopireddy, R. M. Humza, E. Gutheil: Euler–Euler and Euler–Lagrangian<br />

simulations of a water/air spray in a convective flow field. 8th International Conference<br />

on Multiphase Flow (ICMF), 1–8, May 26–31, Jeju, South Korea (2013)<br />

[206] E. Wimmer, G. Brenn: Personal communication (2011)<br />

[207] H. E. Albrecht, N. Damaschke, C. Tropea: Laser Doppler and phase Doppler<br />

measurement techniques: Springer-Verlag, Berlin (2003)<br />

[208] J. C. Wheeler: Modified moments and Gaussian quadratures. Rocky Mountain<br />

Journal of Mathematics, 4, 287–296 (1974)


xviii<br />

BIBLIOGRAPHY<br />

[209] S. R. Gopireddy, R. M. Humza, E. Gutheil: Modeling and simulation of water<br />

and PVP/water evaporating spray flows using the direct quadrature method of<br />

moments. Atomization and Sprays, submitted (2013)<br />

[210] Aa. Fredenslund, R. L. Jones, J. M. Prausnitz: Group-contribution estimation of<br />

activity coefficients in non-ideal liquid mixtures. AIChE Journal, 21, 1086–1099<br />

(1975)<br />

[211] T. Oishi, J. M. Prausnitz: Estimation of solvent activities in polymer solutions<br />

using a group-contribution method. Industrial & Engineering Chemistry Process<br />

Design and Development, 17(3), 333–339 (1978)<br />

[212] D. C. Kannan, J. L. Duda, R. P. Danner: A free-volume term based on the<br />

van <strong>der</strong> Waals partition function for the UNIFAC model. Fluid Phase Equilibria,<br />

228-229, 321–328 (2005)<br />

[213] A. Correa, J. F. Comesana, A. M. Sereno: Use of analytical solutions of<br />

groups (ASOG) contribution method to predict water activity in solutions of<br />

sugars, polyols and urea. International Journal of Food Science and Technology,<br />

29(3), 331–338 (1994)<br />

[214] A. Striolo, J. M. Prausnitz: Vapor-liquid equilibria for some concentrated aqueous<br />

polymer solutions. Polymer, 41, 1109–1117 (2000)<br />

[215] A. Bondi: Physical properties of molecular crystal, liquids and glasses., New York:<br />

Wiley (1968)<br />

[216] R. P. Danner, M. S. High: Handbook of polymer solution thermodynamics:<br />

DIPPR Project, AIChE, New York (1993)<br />

[217] T. E. Daubert, R. P. Danner: Data compilation of properties of pure compounds:<br />

DIPPR Project, AIChE, New York (1992)<br />

[218] A. Kleinhans: Personal communication (2012)<br />

[219] E. Littringer, M. Koester, N. Urbanetz: Personal communication (2012)<br />

[220] A. Z. Dakroury, M. B. S. Osman, A. W. A. El-Sharkawy: Properties of solutions<br />

of polyvinylpyrrolidone. International Journal Thermophysics, 11, 515–523<br />

(1990)<br />

[221] Z. A. Metaxiotou, S. G. Nychas: Experimental measurement of diffusion in aqueous<br />

polyvinylpyrrolidone solutions. AIChE Journal, 41, 812–818 (1995)


BIBLIOGRAPHY<br />

xix<br />

[222] I. S. Grigoriev, E. Z. Meylikhov: Physical values handbook: EnergoAtomIzdat,<br />

Moscow (1991)<br />

[223] G. L. Hubbard, V. E. Denny, A. F. Mills: Droplet evaporation: Effects of transients<br />

and variable properties. International Journal of Heat and Mass Transfer,<br />

18, 1003–1008 (1975)<br />

[224] R. Sedelmayer, M. Griesing, A. Halfar, H.-U. Moritz: Personal communication<br />

(2012)<br />

[225] S. R. Gopireddy, E. Gutheil: Numerical investigation of PVP/water spray flows<br />

in an Eulerian framework. 25th European Conference on Liquid Atomization and<br />

Spray Systems (ILASS), 1–7, September 1–4, Chania, Greece (2013)<br />

[226] S. R. Gopireddy, K. Martin, E. Littringer, A. Mescher, P. Walzel, N. Urbanetz,<br />

E. Gutheil: Numerical and experimental study mannitol/water droplet evaporation<br />

and drying. 25th European Conference on Liquid Atomization and Spray<br />

Systems (ILASS), 1–7, September 1–4, Chania, Greece (2013)<br />

[227] S. Ghosh, J. C. R. Hunt: Induced air velocity within droplet driven sprays.<br />

Proceedings of Royal Society London, A 444, 105–127 (1994)<br />

[228] A. C. S. B. Silva, J. A. B. Cunha Neto, R. Lamberts: Modelling spray vaporization<br />

for evaporative cooling of buildings. Building Services Engineering Research<br />

and Technology, 25(4), 351–361 (2004)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!