Você está na página 1de 240

En vue de l'obtention du

DOCTORAT DE L'UNIVERSITÉ DE TOULOUSE


Délivré par :
Institut National Polytechnique de Toulouse (Toulouse INP)
Discipline ou spécialité :
Génie des Procédés et de l'Environnement

Présentée et soutenue par :


Mme ADRIANA BENTES CORREIA FERREIRA
le jeudi 18 novembre 2021

Titre :
Evaluation des potentiels industriels pour une méthode de séparation
innovante high-tech

Ecole doctorale :
Mécanique, Energétique, Génie civil, Procédés (MEGeP)

Unité de recherche :
Laboratoire de Génie Chimique ( LGC)
Directeur(s) de Thèse :
M. PATRICE BACCHIN
MME MICHELINE ABBAS

Rapporteurs :
M. FRÉDÉRIC PIGNON, UNIVERSITE GRENOBLE ALPES
MME CÉCILE LEMAITRE, UNIVERSITÉ LORRAINE

Membre(s) du jury :
MME GAËTANE LESPES, UNIVERSITE DE PAU ET DES PAYS DE L ADOUR, Président
M. EMMANUEL MIGNARD, UNIVERSITE DE BORDEAUX, Membre
MME ANA HIPOLITO, SOLVAY LYON, Invité(e)
MME MICHELINE ABBAS, TOULOUSE INP, Membre
M. PATRICE BACCHIN, UNIVERSITE TOULOUSE 3, Membre
M. PHILIPPE CARVIN, SOLVAY LYON, Membre
M. PIERRE AIMAR, UNIVERSITE TOULOUSE 3, Invité(e)
Assessment of the Industrial Potentials for an Innovative
High-tech Separation Method

Adriana Marisa Bentes Correia Ferreira

Thesis to obtain the Ph.D. Degree in

Chemical Process Engineering

Director: Patrice Bacchin, LGC Toulouse


Co-Director: Micheline Abbas, LGC Toulouse
Scientific Director: Philippe Carvin, Solvay RICL Lyon

Examination Committee
Rapporteur: Cécile Lemaitre
Rapporteur: Frédéric Pignon
Examiner: Gaëtane Lespes
Examiner: Emmanuel Mignard
Invited member: Ana Hipólito
Invited member: Pierre Aimar

November 18, 2021


ii
Agradecimentos

Em primeiro lugar, um muito obrigado aos meus supervisores de tese, Micheline Abbas, Patrice Bac-
chin e Philippe Carvin:
Obrigada Micheline por toda a tua dedicação ao projecto, pelo teu trabalho árduo e suporte ao
longo destes três anos. Obrigada por me teres mostrado o apaixonante mundo da CFD! Guardar-te-ei
como exemplo e inspiração.
Obrigada Patrice por toda a tua paciência, perseverança e motivação nos momentos difíceis.
Por sempre me incentivares e valorizares o meu trabalho, especialmente nos momentos em que eu
mesma não acreditava. A tua boa disposição, simpatia, paciência e benevolência foram pilares nestes
três anos. Obrigada por me abrires as portas ao mundo incrível da pressão osmótica, dos processos
membranares e pela sugestão do livro "Les Objets Fragiles"S.
Obrigada Philippe, por todas as nossas discussões relacionadas ao projecto e de orientação profis-
sional. Obrigada pela sua benevolência e supervisão. Guardarei as nossas conversas e tudo o que
aprendi consigo. Sei bem que teria um mundo de aprendizagem por descobrir...

Aos três, "Há gente que fica na história, na história da gente"

Um especial agradecimento ao Frédéric da Costa, por toda a sua motivação, empenho, paciência
e dedicação ao projecto. Dedico-te este poema de um heterónimo de Fernando Pessoa (não poderia
deixar de incluir as nossas raízes Portuguesas!)

"A arte é um esquivar-se a agir, ou a viver. A arte é a expressão intelectual da emoção, distinta da vida,
que é a expressão volitiva da emoção. O que não temos, ou não ousamos, ou não conseguimos,
podemos possuí-lo em sonho, e é com esse sonho que fazemos arte. Outras vezes a emoção é a tal ponto
forte que, embora reduzida a acção, a acção, a que se reduziu, não a satisfaz; com a emoção que sobra,
que ficou inexpressa na vida, se forma a obra de arte. Assim, há dois tipos de artista: o que exprime o
que não tem e o que exprime o que sobrou do que teve."

Obrigada Alice Douliez, pelo teu trabalho árduo durante o estágio. Foi uma ajuda valiosa nos
últimos meses da minha tese. Tenho a certeza que com a tua garra, curiosidade e empenho terás
uma carreira incrível!
Agradeço ao Pierre Aimar e à Ana Hipólito, pela oportunidade e pela sua colaboração no projecto.
Obrigada Pierre e a todo o staff do LGC pelo ambiente de trabalho incrível que é proporcionado aos
alunos de doutoramento.
Obrigada aos membros do júri Cécile Lemaitre, Gaëtane Lespes, Frédéric Pignon e Emmanuel
Mignard por terem aceitado avaliar o meu trabalho, pela discussão construtiva e pelas suggestões.
Obrigada a todos os membros do departamento GIMD, em especial à Clémence Coetsier por ter
partilhado comigo o laboratório e diversas vezes me ter deixado usar o seu material e ao Pierre Roblin
pelas diversas vezes que me auxiliou com o protótipo experimental.
Obrigada ao Vincent Bouvier, Sandrine Desclaux, Laure Latapie, Maria Escobar Munoz, Bruno
Boyer e Claudine Lorenzon que são essenciais ao desenvolvimento e progresso dos projectos.
Obrigada ao Mounir Bouaifi, Nicolas Perret e Michel Garrait da Solvay pela vossa colaboração no
projecto. Um especial obrigado ao Michel pela sua getileza e pela oportunidade que me deu.
Obrigada ao Rémi Couson e ao Pierre-François Calmon do LAAS pela colaboração no projecto.
Obrigada aos meus colegas e amigos do LGC, Michelle, Eduardo, Florent, Carlos, Yohann, Sér-
gio, Manuel, Thomas, Margot, Nouha, Marco, Nadhia, Robbie, Ali, Charaf, Nabiil, Chloé, Dihia e
Christophe, pelo três anos de companheirismo e suporte.
Um muito obrigado aos meus colegas de escritório Waldemir, Elise e Kalyani, os nossos anos de
Kadrimir serão sempre lembrados com muita saudade.
Kalyani, escrever sobre ti é quase tão difícil como começar um novo doutoramento. Desde o dia
que apareceste no nosso escritório que a minha vida é repleta de blocos de notas com desenhos de
patos, posts no instagram e twitter com patos, patos de borracha, patos de peluche... Resumindo,
patos! Obrigada pelo companheirismo, motivação, ajuda, gargalhadas, lágrimas, bolachas milka,
crème de marrons e tarte banane chocolat. Obrigada por TUDO. E para que isto não fique lamechas,
digo-te o resto quando tivermos em volta de uma tábua de queijos a beber o nosso copo de vinho
(que com certeza serás tu a escolher!)
Obrigada aos meus amigos Adriana Cunha, Rita, Filipa, Daniela e Bruno Fernandes, por serem
tão presentes. A distância física nunca foi uma barreira à nossa amizade e pude sempre contar com
os vossos conselhos, suporte e encorajamento.
Noire, obrigada por teres aparecido na entrevista e motivado a minha selecção (garantiste mais
uns aninhos de ração Royal Canin!), por me constrangeres com os teus roncos durante as minhas
reuniões em tempo de pandemia e por todas as vezes que poses-te baba nos meus documentos,
o focinho na frente do écran e me deste patadas para que eu largasse a tese e te desse atenção.
Preenches a minha vida de alegria e amor.
Guilherme, este trabalho é nosso. O teu suporte, ajuda, cuidado e amor têm sido a base do que
sou e faço.

iv
vi
Às minhas professoras Graça Filipe e Sílvia Domingues
Abstract

Currently, we are facing a general technological gap for the separation of mixtures containing col-
loidal particles (metallic nano-particles, macro-molecules, etc.). This project aims to extend the prin-
ciple of Hollow Fiber Flow Field-Flow Fractionation (HF5) toward continuous size-based separation
of colloidal suspensions. HF5 is an analytical technique used for particle fractionation (according to
their size) in view of their characterization. The principle of particle separation is based on consec-
utive focusing of particles of different sizes on different flow streamlines with the aid of a permeate
flow applied across the hollow-fiber wall and their differential elution by the main flow streamlines
at the fiber outlet.
In this Ph.D., we developed an automated sequential prototype based on the HF5 principles.
The accumulation and elution conditions were explored, leading to the prescription of a separa-
tion methodology, which applies for stable negatively charged Brownian particles (in order to limit
particle aggregation and adsorption on the membrane). A CFD model was developed to guide the
initial prototype design and the choice of the operating conditions. The numerical model was ex-
tended to include an original description of colloidal dynamics near phase transition and its impact
on near-membrane particle accumulation and relaxation processes.
We evidenced the existence of a permeate flow-rate (during elution) at which the Brownian
diffusion prevails over the drag force, and that allows controlling particle differential elution: this
reveals to be a crucial condition for successful separation. The developed prototype allows the pro-
cessing of higher sample amount than classical HF5, by using the total membrane surface during
particle focusing. We performed the separation of latex polystyrene nanoparticles with a size ratio
of 5, using 4.74 µg of particles per cm2 of membrane without compromising the peak resolution
(similar to AF4 configuration and superior to HF5). Moreover, the prototype allows running several
consecutive separation cycles (up to five cycles were tested), without increasing membrane fouling.
To allow prototype scale-up, further optimization of particle accumulation along the membrane and
permeate-flow control during elution are needed.
Finally, a system for continuous operation was developed. It is not included in this manuscript
for confidentiality reasons.

Keywords: size-based; separation, HF5, nanoparticles


x
Résumé

Un développement technologique est nécessaire pour la séparation de suspensions contenant diffé-


rentes populations de particules colloïdales (nanoparticules métalliques, macromolécules, etc.). Ce
projet vise à étendre le principe de la méthode analytique HF5 pour fractionner, en continu, les par-
ticules d’une suspension colloïdale en fonction de leur taille. Dans la HF5 la suspension colloïdale
est transportée dans une fibre creuse. Le principe de séparation est basé sur i) la focalisation de
particules de différentes tailles sur différentes lignes de courant en appliquant un flux de perméation
à travers la membrane, et ii) une étape d’élution.
Dans cette thèse, nous avons développé un prototype séquentiel automatisé basé sur les principes
de la HF5. Les conditions d’accumulation et d’élution ont été explorées, conduisant à la prescrip-
tion d’une méthodologie de séparation, qui s’applique aux particules browniennes stables chargées
négativement (pour limiter l’agrégation et l’adsorption des particules sur la membrane). Un modèle
CFD a guidé la conception initiale du prototype et le choix des conditions opératoires. L’équation de
transport y a été étendue pour inclure une description originale de la dynamique colloïdale près de la
transition de phase et de son impact sur les processus d’accumulation et de relaxation des particules
près de la membrane.
Nous avons mis en évidence l’existence d’un flux de perméation (lors de l’élution) auquel la
diffusion brownienne prévaut sur la force de traînée, permettant de contrôler l’élution différentielle
des particules : c’est une étape clé pour réussir le fractionnement. Le prototype développé permet
le traitement d’une quantité d’échantillon plus élevée que la HF5, grâce à l’utilisation de la surface
totale de la membrane lors de la phase d’accumulation. Nous avons effectué la séparation de 4,74 µg
de nanoparticules de latex (par cm2 de membrane) avec un rapport de taille 5 sans compromettre la
résolution du pic. Par ailleurs, le prototype permet d’effectuer plusieurs cycles de séparation (cinq
cycles consécutifs ont été testés), sans augmenter l’encrassement de la membrane. Pour permettre
un up-scaling du prototype, une optimisation supplémentaire de l’accumulation de particules le long
de la membrane et un contrôle du flux de perméat pendant l’élution sont nécessaires.
Enfin, un système de fonctionnement continu a été développé: il n’est pas inclus dans ce manus-
crit pour des raisons de confidentialité.

Mots Clés : granulométrie; séparation; HF5; nanoparticules


xii
Contents

List of Tables xvi

List of Figures xviii

Glossary xxvii

Acronyms xxxi

1 Introduction 1
1.1 Problem statement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Thesis structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2 Field-Flow Fractionation 7
2.1 The Field-Flow Fractionation Technology . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.1.1 Principle and mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.1.1.1 Accumulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.1.1.2 Elution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.1.2 Operating mode of the FFF technology . . . . . . . . . . . . . . . . . . . . . . . . 12
2.1.2.1 Detection techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.2.2 External fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 Flow Field-Flow Fractionation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2.1 Configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2.2 Operating steps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2.3 Impact of the operating conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3 Towards up-scaling and continuous operation . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3.1 Preventing the stop-flow step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3.2 Increase sample throughput . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.3.3 Continuous Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.4 Conclusion and thesis stance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3 From colloids transport to FFF simulations 31


3.1 Physics of colloidal dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.1.1 Definition of colloids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.1.2 Forces acting on colloidal dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . 32
h
3.1.2.1 Hydrodynamic force, F : . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
B
3.1.2.2 Brownian force, F : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
g
3.1.2.3 Gravity force, F : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
h B g
3.1.2.4 Comparison of the forces (F , F , and F ) magnitude . . . . . . . . . 34

xiii
3.1.3 Particle-particle Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.1.3.1 Electrostatic repulsive interactions . . . . . . . . . . . . . . . . . . . . . 35
3.1.3.2 van der Waals attractive interactions . . . . . . . . . . . . . . . . . . . . 37
3.1.3.3 DLVO theory for the description of interactions in diluted conditions 38
3.1.3.4 Osmotic pressure for the description of interactions in concentrated
conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.1.4 Phase transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.1.5 Transport dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.1.5.1 Diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.1.5.2 Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.1.6 Impact of the colloids physics in the Flow-FFF mechanism . . . . . . . . . . . . 43
3.2 Modelling of colloids transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.2.1 Hydrodynamic governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.2.2 Model for the prototype design: Simulation of the Flow FFF mechanism . . . 46
3.2.2.1 Numerical simulations setup . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.2.2.2 Effect of the operating conditions . . . . . . . . . . . . . . . . . . . . . . 49
3.2.2.3 Guidelines obtained for the prototype development . . . . . . . . . . 57
3.2.2.4 Model Improvement for high concentrations . . . . . . . . . . . . . . . 57
3.2.3 Second model: Modelling of near-phase transition and its impact in membrane
processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.2.3.1 Mathematical model for the osmotic pressure . . . . . . . . . . . . . . 60
3.2.3.2 Application of the mathematical model to study the mechanism of
gel relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.2.3.3 Fiber plugging caused by a gel phase relaxation . . . . . . . . . . . . . 64

4 Materials and methods 67


4.1 Prototype development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.1.1 Process-Flow Diagram (PFD) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.1.2 Equipment details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.1.2.1 Microfluidic flow control (MFCS-EZ) . . . . . . . . . . . . . . . . . . . . 71
4.1.2.2 Valve at the channel inlet (L-SWITCH) . . . . . . . . . . . . . . . . . . . 71
4.1.2.3 Valve at the channel outlet (M-SWITCH) . . . . . . . . . . . . . . . . . 72
4.1.2.4 Flow-meters (RD1 and RD2) . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.1.2.5 Channel (MHF5) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.1.2.6 Online fluorescence spectroscopy (OFS) system (FC, HDX, and DH-
2000) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.1.3 Operating Steps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.1.3.1 Injection Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.1.3.2 Accumulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
4.1.3.3 Elution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.1.4 Protocol development on Microfluidics Automation Tool (MAT) . . . . . . . . . 82
4.1.5 Photo of the developed prototype (SPFFF) . . . . . . . . . . . . . . . . . . . . . . 84
4.2 Sample properties and characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

xiv
4.2.1 Size and surface charge characterization . . . . . . . . . . . . . . . . . . . . . . . 85
4.2.1.1 Measurement of the Zeta potential (ζ) . . . . . . . . . . . . . . . . . . 86
4.2.1.2 Determination of the size (d p ) . . . . . . . . . . . . . . . . . . . . . . . . 86
4.2.2 Calibration curves of the online fluorescence spectroscopy (OFS) system . . . 87
4.3 Solvent properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.3.1 Sodium Dodecyl Sulfate (SDS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.3.2 Phosphate buffer saline solution (PBS) . . . . . . . . . . . . . . . . . . . . . . . . 94
4.4 Verification of the system hydrodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.4.1 Definitions for the experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.4.2 Measurement of the signal at three different parts of the prototype . . . . . . . 97
4.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

5 From exploration of operating conditions to method prescription 103


5.1 Exploration of the operating conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
5.1.1 Accumulation step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.1.1.1 Inlet pressure Pi =25 mbar: experiment A and B . . . . . . . . . . . . 106
5.1.1.2 Inlet pressure Pi =50 mbar: experiment C, D and E . . . . . . . . . . . 108
5.1.1.3 Inlet pressure Pi =100 mbar: experiment F and G . . . . . . . . . . . . 111
5.1.1.4 Conclusion of the tested accumulation operating conditions . . . . . 113
5.1.2 Elution step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.1.2.1 Radial flow-rate, Q̄ r e : experiments H and I . . . . . . . . . . . . . . . . 116
5.1.2.2 Axial flow-rate, Q̄ z : Experiment J . . . . . . . . . . . . . . . . . . . . . . 117
5.1.2.3 Conclusion of the elution operating conditions . . . . . . . . . . . . . . 118
5.2 Exploration of the effect of the particles size, d p . . . . . . . . . . . . . . . . . . . . . . . 119
5.3 Methodology prescription for a size-based separation using SPFFF . . . . . . . . . . . . 122

6 Application of the Separation Methodology 125


6.1 Application of the methodology for the separation of L100 and L500 latex particles
(size ratio 5) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
6.2 Optimization of the separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
6.2.1 Impact of the elution permeate flow-rate (Q r e ) . . . . . . . . . . . . . . . . . . . 128
6.2.2 Effect of the solvent properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
6.3 Towards more intensive separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
6.3.1 Increase of the injected quantity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
6.3.2 Cyclic operation: five sequences in a row . . . . . . . . . . . . . . . . . . . . . . . 136
6.4 Recommendation for smaller size ratio separation . . . . . . . . . . . . . . . . . . . . . . 139

7 Towards a continuous operation 141

8 Conclusion and perspectives 143


8.1 General Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
8.2 Perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
8.2.1 Technological optimization of the SPFFF . . . . . . . . . . . . . . . . . . . . . . . 144
8.2.2 Further exploration of the operating conditions in SPFFF . . . . . . . . . . . . . 145

xv
8.2.3 SPFFF scale-up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
8.2.4 Continuous operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

Bibliography 147

A Commercial channel structure of other FFF sub-techniques 155

B Illustration of the HF5 experimental steps 157

C Split channel configurations 159

D Submitted article: "Colloid dynamics near phase transition: simulation of the relaxation
of concentrated layers" 161

E Abstract submission: "Plugging of a hollow fiber capillary by gel relaxation during a


cleaning step" 173

F Equipment specification 175

G Online fluorescence spectroscopy 183

H Sample Characterization 185


H.0.1 Measurement of the maximum emission wavelength . . . . . . . . . . . . . . . . 185
H.0.2 Titration curves of zeta potential vs pH . . . . . . . . . . . . . . . . . . . . . . . . 186

I Residence Time Distribution modeling 189


I.1 Applied models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
I.1.1 Laminar Newtonian pipe flow model (LNPF) . . . . . . . . . . . . . . . . . . . . . 189
I.1.2 Cascade model reactores (CR) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
I.1.3 Dispersive plug flow model (DPF) . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
I.2 Signals Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
I.2.1 LV signal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
I.2.2 RD1 signal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
I.2.3 RTD signal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
I.3 Obtained parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191

J Measurement of the RD1 signal using ultra-pure water and SDS 193

K LV signal 195

L Repeatability of L100 + L500 separation 197

xvi
List of Tables

2.1 Summary of the Field-Flow Fractionation sub-techniques. . . . . . . . . . . . . . . . . . 14


2.2 Summary of the FFF progress for increasing sample quantity in batch process and to
obtain a continuous operation. Nomenclature: Q o - outlet flowrate; Q i - inlet flowrate;
Q y - permeate flowrate; L - channel length; w - channel thickness ; bo - maximum
channel breadth; b L - minimum channel breadth; Am - membrane surface area; and
H - height between steric-FFF channel top and bottom. . . . . . . . . . . . . . . . . . . . 30

3.1 Colloids classification according to the phase of the suspending medium and dispersed
phase. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.2 Summary of the forces applied to a colloidal dispersion. . . . . . . . . . . . . . . . . . . 32
3.3 Parameters used with the numerical model. . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.4 Retention time of particles with 10, 50, and 100 nm subjected to an elution stage with
different permeation conditions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.5 Efficiency of separation (η) of binary mixtures where one of the components repres-
ents up to 80% of the composition. In 10 nm + 50 nm and 10 nm + 100 nm is 80%
of 10 nm, and in 50 nm + 100 nm is 80% of 100 nm. . . . . . . . . . . . . . . . . . . . . 55

4.1 Dimensions of the prototype equipment and tubing. . . . . . . . . . . . . . . . . . . . . . 70


4.2 Reservoirs specifications. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.3 Summary of the channel properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.4 Properties of the model particles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.5 Samples composition for the investigation of the SDS effect in the fluorescence signal. 92
4.6 Measurement of the sample properties (size (d p ), zeta potential (ζ), and pH) at the
days 0 and 8 after preparation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.7 Hydrodynamic conditions and mass recovery of the signal after LV, the input signal
(RD1), and the output signal (RTD) experiments . . . . . . . . . . . . . . . . . . . . . . . 98

5.1 Operating conditions during the accumulation step of L100. . . . . . . . . . . . . . . . . 105


5.2 Designation of the experiments performed to investigate the accumulation operating
conditions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5.3 Hydrodynamic conditions of the experiments A and B. . . . . . . . . . . . . . . . . . . . 106
5.4 z-av g determined by DLS of the fractions (f1 to f5) collected during the experiments
A and B. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.5 Average flow conditions of the experiments C, D, and E. . . . . . . . . . . . . . . . . . . 109
5.6 z-av g determined by DLS of the fractions (f1 to f5) collected during the experiments
C, D, and E. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
5.7 Average flow conditions of the experiments F and G. . . . . . . . . . . . . . . . . . . . . 111

xvii
5.8 z-av g determined by DLS of the fractions (f1 to f5) collected during the experiments
F and G. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.9 Experiments to study the elution conditions. . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.10 Average flow conditions of the experiments H, I, and J. . . . . . . . . . . . . . . . . . . . 116
5.11 Summary of the limit permeation conditions for particles evacuation during the elu-
tion step. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

6.1 Imposed (Pi and Pr ) and measured average hydrodynamic conditions. . . . . . . . . . 126
6.2 Granulometry analysis (z-av g and count-rate values) of the collected fractions. . . . . 127
6.3 Imposed (Pi and Pr ) and measured average hydrodynamic conditions. . . . . . . . . . 129
6.4 Granulometry analysis (z-av g and count-rate values) of the collected fractions. . . . . 130
6.5 Granulometry analysis (z-av g and count-rate values) of the collected fractions. . . . . 133
6.6 Granulometry analysis (z-av g and count-rate values) of the collected fractions. . . . . 136
6.7 Granulometry analysis (z-av g and count-rate values) of the collected fractions. . . . . 138

A.1 Available AF4 Channel dimensions of Wyatt Technologies. . . . . . . . . . . . . . . . . . 156


A.2 Specifications of Postnova equipments. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

F.1 Specification sheet number 1 of the microfluidic flow controller (MFCS-EZ.) . . . . . . 175
F.2 Specification sheet number 2 of the inlet rotatory valve (LV). . . . . . . . . . . . . . . . 176
F.3 Specification sheet number 3 of the outlet rotatory valve (MV). . . . . . . . . . . . . . . 177
F.4 Specification sheet number 4 of the inlet flow-meter (RD1). . . . . . . . . . . . . . . . . 178
F.5 Specification sheet number 5 of the outlet flow-meter (RD2). . . . . . . . . . . . . . . . 179
F.6 Specification sheet number 6 of the fluorescence cell (FC). . . . . . . . . . . . . . . . . 180
F.7 Specification sheet of the spectrometer (HDX). . . . . . . . . . . . . . . . . . . . . . . . . 181
F.8 Specification sheet number 8 of the light source (DH-2000). . . . . . . . . . . . . . . . 181

H.1 Model particles specifications. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185


H.2 Supplied and measured maximum λemi of the L50, L100, and L500 nanoparticles. . . 186

I.1 Parameters obtained with the CR and PFR models. Nomenclature: t¯r - mean resid-
ence time, n - number of reactors determined bu the CR model, Pe - Peclet number
determined with the DPF model, and Dc - Dispersion coefficient calculated from the
Pe number obtained with the DPF model. . . . . . . . . . . . . . . . . . . . . . . . . . . . 191

J.1 Hydrodynamic conditions of the RD1 signal measurements. . . . . . . . . . . . . . . . . 193

K.1 Hydrodynamic conditions of the LV + tube 4 signal measurements. . . . . . . . . . . . 195

xviii
List of Figures

1.1 Particle separation processes in different size ranges. . . . . . . . . . . . . . . . . . . . . 2

2.1 Scheme showing the Field-Flow Fractionation mechanism of particles with two differ-
ent sizes. D1 and D2 represent the Brownian diffusion of larger and smaller particles,
respectively (D1 < D2 ). The magnitude of the external field (F y ) is the same for both
particles (F y = F y1 = F y2 ). The figure is merely illustrative and it is not at the right
scale. The particle layers are at O(∼ 10 µm) from the accumulation wall (1 − 10% of
total channel width). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2 Illustration of the layers formed during the accumulation of a dispersion with two
sizes. l1 and l2 correspond to the mean layer thickness of the large and small particles,
respectively. Notice that the figure is merely illustrative. In reality, the particle layers
have mixed regions due to inter-particle interactions. . . . . . . . . . . . . . . . . . . . . 9
2.3 Normalized concentration profiles for two particles with a size ratio of 10 as function
of the normalized distance from the wall. . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.4 Normalized sample retention time (t r /t 0 ) according to the particles size (character-
ized by their Pe y number). The calculations were performed considering a channel
with w = 0.7mm, an accumulation velocity of U y = 1 × 10−6 m/s, and particles with
7, 10, 50, 100, 200, and 500 nm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.5 Scheme illustrating the Field-Flow Fractionation (FFF) complexity. . . . . . . . . . . . 13
2.6 Scheme illustrating the edge view of the F1FFF configuration. The permeate flow
enters the channel from an external source and is represented by the orange arrows. 15
2.7 Scheme illustrating the edge view of the AF4 configuration. The permeate flow comes
from the mainstream flow due to the transmembrane pressure and is represented by
the blue arrows. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.8 Scheme illustrating an axial cut of the HF5 channel. The permeate flow comes from
the mainstream flow due to the transmembrane pressure and is represented by the
blue arrows. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.9 Scheme illustrating the AF4 commercial channel. Adapted from [32] . . . . . . . . . . 16
2.10 Scheme illustrating the HF5 commercial channel. . . . . . . . . . . . . . . . . . . . . . . 17
2.11 Scheme illustrating the injection step in an AF4 configuration. . . . . . . . . . . . . . . 17
2.12 Scheme illustrating the focusing/relaxation step in an AF4 configuration. . . . . . . . 18
2.13 Scheme illustrating the Elution step in an AF4 configuration. . . . . . . . . . . . . . . . 18
2.14 Diagram illustrating the operating conditions of the Flow-FFF techniques. . . . . . . . 21
2.15 Illustration of an AF4 channel with a frit inlet. Reproduced from [17]. . . . . . . . . . 23
2.16 Compartmentalized AF4 and HF5 channels. Reproduced from [41]. . . . . . . . . . . . 23
2.17 Representation of the circular configuration made of 12 AF4 trapezoidal channels [20]. 24
2.18 Continuous steric FFF channel configuration. From [42] . . . . . . . . . . . . . . . . . . 25

xix
2.19 Configuration of the Split-FFF device. Adapted from [13]. . . . . . . . . . . . . . . . . . 26
2.20 Scheme of the continuous 2D-ThFFF device. Reproduced from [21]. . . . . . . . . . . 27
2.21 Schematic representation of the 2D-AF4 channel. Reproduced from [22]. . . . . . . . 27

3.1 Calculation of the viscous (F h ), Brownian (F B ), and gravity (F g ) forces for polystyrene
(PS latex), silica oxide (SiO2 ) and cerium oxide (CeO2 ) nanoparticles within a radius
range of 1 nm to 1.5 µm. The PS latex particles density ρ p = 1050 [k g/m3 ], the CeO2
particle density ρ p = 7200 [kg/m3 ], the SiO2 particle density ρ p = 2000 [kg/m3 ], the
fluid density ρ= 1000 [kg/m3 ], the fluid dynamic viscosity µ= 1 × 10−3 [Pa.s], the
particles velocity U p = 1 × 10−6 [m/s], and the temperature T = 298.15 [K] . . . . . . 34
3.2 Normalized electrostatic potential (ψ scaled by ζ) of a spherical particle, calculated
with the Debye-Huckel approximation, as function of the distance from the particle
surface ds (ds = r − a). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.3 Interaction energy potential (W DLV O ) obtained by the addition of the electrostatic
(W ele ) and van der Waals (W vdw ) contributions for two approaching spherical particles
with d p = 100 nm, ζ= −40 mV , and Is = 1×10−5 M . Calculation parameters: ε= 78.3
(for water at 298.15 K - pg. 43 of [53]), kB = 1.381×10−23 J/K, T = 298.15 K, zi = 1,
e= 1.602 × 10−19 , and A= 1.3 × 10−20 J. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.4 Interaction energy potential (W DLV O ) for two approaching spherical particles with
d p = 100 nm, ζ= −40 mV , at different Is . Calculation parameters: ε r = 78.3, kB =
1.38064852 × 10−23 J/K, T = 298.15 K, zi = 1, e= 1.602176634 × 10−19 , and A=
1.3 × 10−20 J. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.5 Regions observed during the experimental measurement of the osmotic pressure, caused
by a dispersed-concentrated (disorder-order) phase transition. . . . . . . . . . . . . . . 41
3.6 Geometrical domain of the first implemented model. H - channel height, L - channel
length, L1 - length of the accumulation segment (permeable wall), and L2 - length of
the elution segment (solid wall). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.7 Boundary conditions applied during the (a) accumulation and (b) elution stages. . . 48
3.8 Signals of outlet flux of particles with 10 nm (blue line), 50 nm (orange line), and
100 nm (green line) for simulations using the elution permeate velocities of (a) U y =0
(v.u.), (b) U y = 1 × 10−3 (v.u.), and (c) U y = 1 × 10−2 (v.u.). . . . . . . . . . . . . . . 51

3.9 Retention time (t ) as function of particles sizes (d p ) for different permeation velocit-
ies U y . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.10 Signal of the outlet flux of particles with 10 nm (blue line), 50 nm (orange line), and
100 nm (green line) for simulations using an elution permeate velocity U y = 1 × 10−2
(v.u.) and the geometries (a) 1 (L1 /H = 5 and L2 /H = 5), (b) 2 (L1 /H = 10 and
L2 /H = 0), and (3) 3 (L1 /H = 5 and L2 /H = 10). . . . . . . . . . . . . . . . . . . . . . . 54

3.11 Retention time (t ) as function of particles sizes (d p ) for the different geometry con-
figurations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
3.12 Normalized concentration profiles (scaled by Cin ) at half of channel in the y-direction
(wall-normal direction) for U y equal to (a) 0.1 (v.u.), (b) 0.05 (v.u.), and (c) 2.5×10−4
(v.u.). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

xx
3.13 Representation of classical osmotic pressure curves determined experimentally, high-
lighting the different phase regions of colloidal suspensions. Π - osmotic pressure of
the dispersed region; Π2 - osmotic pressure of the transition region for φ<φcr ; Π3 -
osmotic pressure of the transition region for φ>φcr ; and Π4 - osmotic pressure of the
concentrated region. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.14 Dimensionless osmotic pressure Π∗ (scaled by the van’t Hoff coefficient A vH ) as func-
tion of φ for the isotherms Tc /T at 0.6, 0.7, 0.8, 0.9, 1, 1.2, 1.5, 2, 3, and 5, corres-
ponding to the blue, orange, green, red, purple, brown, pink, grey, yellow, and cyan
lines, respectively. Curves obtained for B1 = 1, B2 = 0, φ1 =0.3, φ2 =0.5, and k=0.5.
The dashed and solid blue lines correspond to the binodal, and spinodal decomposi-
tion, respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.15 (a) Experimental osmotic pressure Π(φ) curve measured by [60] for silica nano-
particles with d p = 10.2 nm. The curve was fitted with φ1 = 0.03, φ2 = 0.09, φcr =
0.06, φcp = 0.72, B1 = 4.0 × 104 [Pa], B2 = 15.0 × 105 [Pa], and k= 0.5, for different
Tc /T . (b) Diffusion coefficient (D∗ scaled by D0 ) derived from the osmotic pressure
curve and illustrated for the same Tc /T . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.16 Illustration of fiber clogging caused by the relaxation of a gel layer. Time t 0 corres-
ponds to the end of the accumulation stage and the times t 1 and t 2 represent different
relaxation times. The gel phase is the region delimited by φ ≥ 0.6 (φcr =0.6 repres-
ented by the black contour line). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

4.1 Process flow diagram of the experimental setup. . . . . . . . . . . . . . . . . . . . . . . . 69


4.2 Image with the side and top views of the L-SWITCH valve and visualization of the two
valve positions with the ESS T M C ont r ol software. . . . . . . . . . . . . . . . . . . . . . . 71
4.3 Image with the side and top views of the M-SWITCH valve and visualization of two
valve positions with the ESS T M C ont r ol software. . . . . . . . . . . . . . . . . . . . . . . 72
4.4 Image of the M-SWITCH rotor seal provided by Fluigent. . . . . . . . . . . . . . . . . . . 72
4.5 Image of the M-SWITCH structure provided by Fluigent. . . . . . . . . . . . . . . . . . . 73
4.6 Operating principle of the flow-meters. Adapted from [84]. . . . . . . . . . . . . . . . . 73
4.7 MHF5 channel composed by a regenerated cellulose hollow-fiber membrane placed
inside a plastic tube. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
−34
4.8 Simplified reproduction of the Jablonski diagram. h= 6.63 × 10 J/s is the Planck
8
constant and c= 3 × 10 m/s is the light speed. . . . . . . . . . . . . . . . . . . . . . . . . 76
4.9 Equipment for the assembling of the online fluorescence spectroscopy analytical method. 77
4.10 Intensity signal (I) as function of the wavelength (λ) of the light source. . . . . . . . . 78
4.11 Filling the experimental setup with the required solvent before sample injection. . . . 79
4.12 Illustration of the injection stage preparation with L-SWITCH in position 2. . . . . . . 79
4.13 Illustration of the injection and accumulation steps for a binary mixture. After accu-
mulation, the smaller particles (represented in orange) are closer to the center than
the larger particles (represented in yellow) which are closer to the membrane. . . . . 80
4.14 Scheme of the filtered volume. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
4.15 Illustration of the elution step with the collection of the more diffusive fraction (rep-
resented in orange). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

xxi
4.16 Illustration of the elution step with the collection of the less diffusive fraction (repres-
ented in yellow). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.17 Screenshot of a protocol developed on Microfluidics Automation Tool (MAT). . . . . . . 82
4.18 Instructions to perform a linear pressure increase and to change the valve position
(two actions in parallel with instructions designed in series). . . . . . . . . . . . . . . . 83
4.19 Setup of the sequential prototype. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
4.20 Calibration curve of Latex L100 measured at λ = 407 nm within a concentration range
of 1.194 × 10−4 − 2.068 × 10−1 g/L. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.21 Regions of the latex L100 calibration curve measured at λ = 407 nm. . . . . . . . . . . 89
4.22 Comparison of the L100 spectra at different concentrations with the reference spectra. 90
4.23 Spectra of the different latex samples (L50, L100, and L500), and comparison with
the reference spectra in the wavelengths range of (a) 200-800 nm and (b) 400-650 nm. 91
4.24 Average value of the the samples size (d p ) measured for day 1 and day 8. . . . . . . . 93
4.25 Scheme of the imposed and measured flow-rate conditions. . . . . . . . . . . . . . . . . 94
4.26 Flow conditions in the MHF5 module. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
4.27 Part of the PFD diagram for the representation of the signal measurement in three
different regions of the prototype. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
4.28 Dispersion regime for a tracer in a capillary tube. From [53] . . . . . . . . . . . . . . . 99
4.29 Representation of the ideal input and output signals. The figure is merely illustrative
and does not have the correct dimensions (the peaks have to present the same area
under the curves). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4.30 L100 signals detected in three regions of the prototype. . . . . . . . . . . . . . . . . . . 100
4.31 Illustration of the sample distribution across the FC cell during the measurement of the
calibration curves and the detection of the elution peaks of discontinuous experiments
(LV, RD1, RTD, and SPFFF). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

5.1 Scheme illustrating the operating conditions of the accumulation and elution steps.
Elution step can comprehend a backwash stage. The duration of the accumulation
step (t a ) is not represented but is also a parameter. Q i - measured inlet flow, Pi -
imposed inlet pressure, Q o - measured outlet flow, Po - imposed outlet pressure, Q r -
measured radial permeate flow, Pr - imposed radial permeate flow. The indexes a, e,
and b stand for the accumulation, elution, and backwash steps, respectively. . . . . . 104
5.2 Flow-rate conditions during accumulation, elution, and backwash of the experiments
A and B. The accumulation is longer for experiment B, while the operating conditions
for the elution and backwash stages are similar. . . . . . . . . . . . . . . . . . . . . . . . 106
5.3 Normalized intensity signal (I) detected at λ=407 nm (a) and flow-rate conditions
(b) measured during the elution stage of the experiments A and B. Six fraction f1 to
f5 were collected during the experiments for granulometry analysis. . . . . . . . . . . . 107
5.4 Size distribution obtained by DLS of the fractions collected during the experiments
(a) A, and (b) B. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.5 Flow-rate conditions during accumulation, elution, and backwash of the experiments
C, D, and E. The accumulation time (t a ) is the varied parameter. . . . . . . . . . . . . . 109

xxii
5.6 Normalized Intensity signal (I) detected at λ=407 nm (a) and flow-rate conditions
(b) measured during the elution stage of the experiments C, D, and E. Six fraction f1
to f5 were collected during the experiments for granulometry analysis. . . . . . . . . . 110
5.7 Size distribution obtained by DLS of the fractions collected during the experiments
(a) C, (b) D, and (c) E. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.8 Flow-rate conditions during accumulation, elution, and backwash of the experiments
F and G. The experiments were realized under the same elution and backwash condi-
tions, varying the accumulation time t a . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
5.9 Normalized intensity signal (I) detected at λ=407 nm (a) and flow-rate conditions
(b) measured during the elution stage of the experiments F and G. . . . . . . . . . . . 112
5.10 Size distribution obtained by DLS of the fractions collected during the experiments
(a) F, and (b) G. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.11 Flow-rate conditions used during the accumulation and elution stages of the experi-
ments H, I, and J. Experiments realized under the same accumulation conditions but
different elution flow-rates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5.12 Normalized intensity signal I detected at λ=407 nm (a) and permeate flow-rate (Q r e )
conditions (b) measured during the elution stage of the experiments B, H, and I. . . . 117
5.13 Normalized intensity signal I detected at λ=407 nm (a) and flow-rate conditions (b)
measured during the elution stage of the experiments B and J. . . . . . . . . . . . . . . 118
5.14 Scheme illustrating the effect of the Q lim
r e in the particles diffusion. If the permeate
radial flow-rate used during the elution step (Q r e ) surpasses the Q lim
r e , the particles
remain accumulated at the membrane. Otherwise, the particles diffusion (D) is pre-
dominant and allows their release. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
5.15 (a) Normalized intensity signal (I) for L50, L100, L200 and L500, being both peaks
presented for λ=407 nm, and (b) permeate flow-rate (Q r e ) conditions measured dur-
ing the elution step. I max =1115.133, 5233.845, 2114.777, and 2761.004 c.u., for
L50, L100, L200 and L500, respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
5.16 Radial permeate flux limit Q lim
r e as function of the (a) particle size d p and (b) particles
diffusion in diluted conditions D0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
5.17 (a) Normalized intensity signal (I)for L100, and S100, being both peaks presented for
λ=407 nm, and (b) permeate flow-rate (Q r e ) conditions measured during the elution
step. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
5.18 Methodology prescription for the size-based separation of nanoparticles using the se-
quential prototype SPFFF. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

6.1 Flow-rate conditions during the accumulation, elution, and backwash stages. . . . . . 126
6.2 Elution peaks detected at λ = 407 nm (blue line) and λ = 546 nm (orange line). . . 127
6.3 Size distribution obtained by DLS of the collected fractions. . . . . . . . . . . . . . . . . 128
6.4 Flow-rate conditions during the accumulation, elution, and backwash stages. . . . . . 129
6.5 Elution peaks detected at λ = 407 nm (blue line) and λ = 546 nm (orange line). . . 129
6.6 Comparison of the I vs λ spectra measured at (a) t e =20 minutes and (b) t e =45
minutes with the reference spectra. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
6.7 Size distribution obtained by DLS of the collected fractions. . . . . . . . . . . . . . . . . 131

xxiii
6.8 Comparison of the two experimental conditions used to separate the binary mixture
L100+L500. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
6.9 Experiments realized under the same accumulation and the elution conditions using
different solvents as carrier fluids. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
6.10 Size distribution obtained by DLS for the fractions collected during the experiments
that used (a) u.p. water, (b) PBS 1X , (c) SDS 0.1 g/L and (e) SDS 1 g/L, as carrier
fluid (solvent). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
6.11 Flow-rate conditions during the accumulation, elution, and backwash stages. . . . . . 135
6.12 Elution peaks detected at λ = 407 nm (blue line) and λ = 546 nm (orange line). . . 135
6.13 Size distribution obtained by DLS of the collected fractions. . . . . . . . . . . . . . . . . 136
6.14 Flow-rate conditions during the accumulation, elution, and backwash stages for the
five sequential cycles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
6.15 Elution peaks detected at λ = 407 nm (blue line) and λ = 546 nm (orange line). . . 138
6.16 Permeation flux (J p ) as function of the transmembrane pressure (T M P) measured
before and after the five cycles (the slope corresponds to the membrane permeability,
L p ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

A.1 Illustration of the centrifugal Field-Flow Fractionation mechanism. From CF2000


Series Postnova catalog. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
A.2 Illustration of the Thermal Field-Flow Fractionation mechanism. From TF2000 Series
Postnova catalog. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
A.3 Electrical Asymmetrical Flow FFF mechanism. From EAF2000 Electrical Flow FFF
Series catalog of Postnova. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

B.1 Experimental steps in a HF5 configuration. . . . . . . . . . . . . . . . . . . . . . . . . . . 157

C.1 (a) Centrifugal split channel and (b) Isoelectric split channel. From [44] . . . . . . . . 159
C.2 (a) Magnetic split channel. From [44]; (b) Gravitational split mechanism. From
GF2000 GS Series catalog of Postnova. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159

G.1 Visualization of I vs t at five λ (300, 407, 519, 520, and 546 nm) and I vs λ signals
at the right and left windows of the OceanView software, respectively. . . . . . . . . . . 183
G.2 Visualization of I vs λ signal in the OceanView software. The green line corresponds
to the reference spectrum and the blue line is the instantaneous spectrum measured
with the lamps turned off and (a) after a background and (b) without a background
subtraction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183

H.1 Spectrum of the L50, L100, and L500 determined by fluorescence spectroscopy by
exciting the photons at λex c . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
H.2 Titration curve of (a) L50 and (b) L100. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
H.3 Titration curve of (a) L200 and (b) L500. . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
H.4 Titration curve of S100. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

xxiv
I.1 (a) LV signal modeled with the laminar Newtonian pipe flow model (red curve), cas-
cade reactors model (orange), and dispersive plug flow reactor (green). (b) Figure
containing only the cascade reactors (orange curve) and dispersive plug flow reactor
(green curve) models (to visualize better the models fitting and the LV signal). . . . . 190
I.2 (a) RD1 signal modeled with the laminar Newtonian pipe flow model (red curve),
cascade reactors model (orange), and dispersive plug flow reactor (green). (b) Figure
containing only the cascade reactors (orange curve) and dispersive plug flow reactor
(green curve) models (to visualize better the models fitting and the RD1 signal). . . . 190
I.3 (a) RTD signal modeled with the laminar Newtonian pipe flow model (red curve),
cascade reactors model (orange), and dispersive plug flow reactor (green). (b) Figure
containing only the cascade reactors (orange curve) and dispersive plug flow reactor
(green curve) models (to visualize better the models fitting and the RTD signal). . . . 191

J.1 Signals measured at the outlet of the RD1 flow-meter. . . . . . . . . . . . . . . . . . . . 193

K.1 Signals measured at the after the LV + tube 4 circuit. . . . . . . . . . . . . . . . . . . . . 195

L.1 Experiments realized under the same accumulation and the elution conditions of Pi =
50 mbar and Pr e = 25 mbar. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
L.2 Experiments realized under the same accumulation and the elution conditions of Pi =
50 mbar and Pr e = 34 mbar. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
L.3 Experiments realized under the same accumulation and the elution conditions of Pi =
50 mbar and Pr e = 30 mbar. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198

xxv
xxvi
Glossary

AF4 Asymmetric Flow Field-Flow Fractionation. ix, 1, 3, 14–24, 27, 28, 44, 116, 134, 144

AIO All-in-One. 1, 68, 94

BC Boundary condition. 1, 47, 48

BSA Bovine Serum Albumin. 1, 12, 24, 26

CAD Computer-Aided Design. 1, 145

CE Capillary electrophoresis. 1, 2

CFD Computational Fluid Dynamics. ix, xi, 1, 3, 4, 45

DH-2000 Light source. 1, 68, 71, 75–77

DLS Dynamic Light Scattering. 1, 4, 35, 68, 85, 86, 92, 102, 105–108, 110, 111, 113, 114, 126,
127, 130–133, 136

DLVO Derjaguin-Landau-Verwey-Overbeek. 1, 4, 38, 39

E1FFF Electrical Flow Field-Flow Fractionation. 1, 14

EOS Equation of state. 1, 40

F1FFF Symmetrical Flow Field-Flow Fractionation. 1, 14–16, 18, 22

FC Cell for the measurement of the online fluorescence spectroscopy. 1, 68, 71, 75–78, 87, 91, 97,
98, 101, 145

FFF Field-Flow Fractionation. 1–3, 7, 8, 10–14, 18–22, 25, 26, 28, 34, 44, 67, 74, 91, 118

Flow-FFF Flow Field-Flow Fractionation. 1, 3, 4, 12, 14, 16–22, 24, 31, 34, 43, 44, 47, 49, 56, 104,
114, 116

HDX Spectrometer. 1, 68, 71, 75–77

HF5 Hollow Fiber Flow Field-Flow Fractionation. ix, xi, 1, 3, 4, 14–17, 19–21, 23, 24, 28, 44, 67,
74, 78, 101, 104, 116, 143

HPC High-performance computation. 1, 44

ICPMS Inductively Coupled Plasma-Mass-Spectrometry. 1, 13

xxvii
L100 Latex particles with 100 nm diameter. 1, 87–93, 97, 100, 104–107, 113, 116, 120–122, 125–
130, 133, 134, 136

L200 Latex particles with 200 nm diameter. 1, 90, 119, 120

L50 Latex particles with 50 nm diameter. 1, 90, 91, 120, 122

L500 Latex particles with 500 nm diameter. 1, 90, 91, 119, 120, 125–127, 129, 130, 133–137, 139

LV L-SWITCH valve. 1, 68, 70, 71, 79, 80, 82, 97, 98, 100, 101

M1 Hollow-fiber module 1. 1, 74, 97

M2 Hollow-fiber module 2. 1, 74

MALS Multi-Angle Light Scattering. 1, 13

MAT Microfluidic Automation Tool. 1, 19, 68, 71, 81, 82, 102, 117

MFCS-EZ Microfluidic Flow Controller. 1, 68, 70, 71, 82, 87, 94, 117

MHF5 Hollow fiber module where the HF5 mechanism occurs. 1, 68, 71, 74, 75, 77, 95, 97, 99,
144, 145

MV M-SWITCH valve. 1, 68, 70, 72, 80–82

OFS Online Fluorescence Spectroscopy. 1, 4, 67, 68, 71, 75–77, 87, 92, 93, 97, 101, 102, 105, 113,
114, 132, 146

PBS Phosphate buffer saline solution. 1, 91, 94, 131–134

PES Polyethersulphone. 1, 21

PFD Process Flow Diagram. 1, 67, 68, 98

PS Polystyrene. 1, 34

PTFE Polytetrafluoroethylene. 1, 16, 72

RC Regenerated cellulose. 1, 21, 74

RD Waste reservoir. 1, 68

RD1 Inlet flow-meter. 1, 68, 70, 73, 74, 82, 95, 97–101

RD2 Outlet flow-meter. 1, 68, 70, 73, 74, 82, 95, 100, 117, 145

RE Sample reservoir. 1, 68, 79

REFn Reservoirs for the collection of n fractions. 1, 68

RI Refractive Index. 1, 13

xxviii
RP Permeate reservoir. 1, 68

RS Solvent reservoir. 1, 68, 80

RTD Residence Time Distribution. 1, 68, 98–101, 107, 113, 114

S100 Silica oxide particles with 100 nm diameter. 1, 90, 121

SdFFF Sedimentation Flow Field-Flow Fractionation. 1, 14

SDK Software Development Kit. 1, 67

SDS Sodium Dodecyl Sulfate. 1, 91–94, 102, 104, 129, 131–134

SEC Size Exclusion Chromatography. 1, 2

SPFFF Sequentiel Prototype based on the FFF mechanism. 1, 4, 5, 57, 67, 74, 75, 85, 87, 89–91,
98, 101, 103, 114, 116, 118, 121, 122, 124, 125, 134, 143–146

ThFFF Thermal Flow Field-Flow Fractionation. 1, 14, 27, 28

Ultem polyetherimide. 1, 76

xxix
xxx
Acronyms

A Hamaker constant [J = (kg.m2 )/(s2 )]. 1, 32, 38, 39, 121

Am Membrane surface area [m2 ]. 1, 30, 53, 95

A vH Van’t Hoff coefficient [Pa = kg/(m.s2 )]. 1, 60, 62, 63

B1 First osmotic Virial coefficient [Pa = k g/(m.s2 )]. 1, 60, 62–64

B2 Second osmotic Virial coefficient [Pa = kg/(m.s2 )]. 1, 60, 62–64

C Concentration field [mol/m3 ]. 1, 9, 46, 87, 88, 96

C F L Courant number [dimensionless]. 1, 48

C p Heat capacity at constant pressure [J/(kg.K) = (m2 )/(s2 .K)]. 1, 74

C y Concentration at the accumulation wall (y-direction in cartesian coordinates) [mol/m3 ]. 1, 9

Cin j Injected sample concentration [g/L]. 1, 79

Cin Concentration of particles at the channel inlet [mol/m3 ]. 1, 50, 56

Cout Concentration field at the channel outlet [mol/m3 ]. 1, 50

D Brownian diffusion [m2 /s]. 1, 7, 118

D(φ) Collective diffusion coefficient (generalized Stokes-Einstein relation) [m2 /s]. 1, 42, 58, 59,
63

D∗ Normalized diffusion scaled by D0 [dimensionless]. 1, 64

D0 Diffusion coefficient in diluted conditions (Stokes-Einstein relation) [m2 /s]. 1, 9, 11, 12, 14, 42,
46, 48, 63, 64, 80, 120

D1 Diffusion for larger particles [m2 /s]. 1

D2 Diffusion for smaller particles [m2 /s]. 1

DT Thermophoretic mobility [m2 /s]. 1, 14

E(t) Residence time distribution [dimensionless]. 1, 97, 99

E f Applied electric field [V /m=(kg.m)/(A.s3 )]. 1, 86

F Faraday constant F = 96485.333 [C/mol = (A.s)/mol]. 1, 36

F B Brownian force [N =(k g.m)/s2 ]. 1, 33, 34

xxxi
F g Gravity force [N =(k g.m)/s2 ]. 1, 33, 34

F h Hydrodynamic force [N =(kg.m)/s2 ]. 1, 33, 34

F y Stokes drag force driving the particles to the accumulation wall [N =(kg.m)/s2 ]. 1, 10

F y1 Drag force for larger particles [N =(kg.m)/s2 ]. 1

F y2 Drag force for smaller particles [N =(k g.m)/s2 ]. 1

H Channel height [m]. 1, 30, 46, 47

H(φ) Happel function [dimensionless]. 1, 42

I Signal intensity [counts]. 1, 77, 78, 86–88, 94, 96, 100, 102, 107, 112, 117, 120, 121, 130

I D Inner diameter [m]. 1, 68, 74

Is Ionic strength [mol/m3 ]. 1, 36–41, 44

I max Maximum of the intensity signal [counts]. 1, 94, 102, 108, 120, 121, 126

J p Permeation flux [m3 /(s.m2 )]. 1, 138, 139

K(φ) Sedimentation coefficient [dimensionless]. 1, 42

L Lenght [m]. 1, 16, 19, 20, 30, 46, 47, 68, 74, 95

L1 Length of the accumulation region [m]. 1, 47, 52

L2 Length of the elution region [m]. 1, 47, 52

L p Membrane hydraulic permeability [m/(s.Pa) = (m2 .s)/k g]. 1, 59, 74, 138, 139

MW Molar mass [g/mol]. 1, 91

M r Mass recovery [%]. 1, 96

Np Number of particles [dimensionless]. 1, 35

N10 Total amount of particles with 10 nm collected at the channel outlet [mol]. 1, 53

Nin Total amount of injected particles [mol]. 1, 50, 53

O Order of magnitude. When a value a is given as O(∼ a) it means that a is an order of magnitude.
1, 7, 8, 12, 37, 49, 74, 75, 79, 89, 90, 97, 104–108, 111, 113, 114, 116, 117, 119, 121, 127,
128, 130, 132, 135, 136, 144

Pi Pressure imposed at the RS reservoir (inlet pressure) [mbar]. 1, 74, 94, 104, 105, 113, 114, 117,
126, 129

Po Atmosferic pressure imposed at the MHF5 outlet [mbar]. 1, 94, 104

Pr Pressure imposed at the RP reservoir (radial pressure) [mbar]. 1, 94, 104, 116, 126, 129

xxxii
Pr a Radial pressure during the accumulation stage [mbar]. 1, 117

Pr e Radial pressure during the elution stage [mbar]. 1, 117

Pe rlim
e Limit permeate Peclet number of the elution step [dimensionless]. 1, 116, 120

Pe r Permeate Peclet number in cylindric coordinates [dimensionless]. 1, 80, 96, 104

Pe x Axial Peclet number in cartesian coordinates [dimensionless]. 1, 48, 49

Pe y Peclet number in the wall-normal direction (y-direction for cartesian coordinates) [Dimension-
less]. 1, 9–12, 47, 51, 56

Pez Axial Peclet number in cylindric coordinates [dimensionless]. 1, 94, 96

Pe r a Permeate Peclet number of the accumulation step [dimensionless]. 1, 108, 109, 111–114

Pe r e Permeate Peclet number of the elution step [dimensionless]. 1, 116

Q lim
r e Limit elution radial permeate flow-rate [µL/min]. 1, 116, 118–120, 122, 125, 139, 143, 145

Q i Axial flow-rate at the channel inlet [m3 /s]. 1, 30, 94, 95, 99, 104, 106, 126

Q o Axial flow-rate at the channel outlet [m3 /s]. 1, 30, 94, 95, 100, 104, 107, 126

Q v Volumetric flow-rate [m3 /s]. 1, 73

Q r a Accumulation radial permeate flow-rate [µL/min]. 1, 80, 81, 103–106, 113, 114, 122, 143

Q r b Backwash radial permeate flow-rate [µL/min]. 1, 103

Q r e Elution radial permeate flow-rate [µL/min]. 1, 82, 103, 109, 114, 116–121, 125, 128, 129,
131

Q r Radial permeate flow-rate in cylindric coordinates [µL/min]. 1, 74, 81, 95, 104, 107, 116, 126,
129

Q y Wall-normal (in y-direction) permeate flow-rate in cartesian coordinates [µL/min]. 1, 14, 30,
57

Q z Elution axial flow-rate in cylindric coordinates [µL/min]. 1, 82, 114, 117, 145

R Constant of ideal gases R = 8.314 [J/(mol.K) = (k g.m2 )/(mol.K.s2 )]. 1, 36, 40

R L Retention level [dimensionless]. 1, 11

R f Fiber radius [m]. 1, 17, 19, 20, 80, 95

R m Membrane resistance [kg/(m2 .s)]. 1, 59

Re Reynolds number [dimensionless]. 1, 45, 48, 49, 94, 96, 98

Re p Particles Reynolds number [dimensionless]. 1, 33

xxxiii
S Cross section [m2 ]. 1, 50, 95

S0 Low energy level [e.u.]. 1, 75

S1 Excitation stage [e.u.]. 1, 75

S pt Total particles surface area [m2 ]. 1, 35

T Temperature [K]. 1, 8, 14, 32–34, 36, 38–40, 43

T M P Transmembrane pressure [mbar]. 1, 95, 103, 114, 138, 139

Tc Critical temperature [K]. 1, 62

Tc /T Gel irreversibility parameter [dimensionless]. 1, 61–64

U Velocity field [m/s]. 1, 45, 57

U f Fluid phase velocity [m/s]. 1, 33

Um Mixture velocity [m/s]. 1, 57

U p Particle velocity [m/s]. 1, 32–34, 57, 86

U x Axial velocity in cartesian coordinates [m/s]. 1, 47, 49, 50, 55

U y Wall-normal velocity (y-direction) in cartesian coordinates [m/s]. 1, 8, 11, 14, 33, 34, 47–56

Uin Velocity field at the channel inlet [m/s]. 1, 50, 53

U r Radial velocity in cylindric coordiantes [m/s]. 1, 58

Uz Axial velocity in cylindric coordiantes [m/s]. 1, 58

V Volume [m3 ]. 1, 40, 68

V0 Channel volume [m3 ]. 1, 14

V3 Volume of tube 3 [µL]. 1, 97, 98

V4 Volume of tube 4 [µL]. 1, 97, 98

V5 Volume of tube 5 [µL]. 1, 98

V6 Volume of tube 6 [µL]. 1, 97, 98

Vf Filtration volume [mL]. 1, 81

Vp Particles volume [m3 ]. 1, 14, 33

VF C Volume of the fluorescence cell (FC) [µL]. 1, 97, 98

VLV Volume of L-SWITCH (LV) valve [µL]. 1, 97, 98

VM H F 5 Volume of the Hollow-fiber (MHF5) [µL]. 1, 98

xxxiv
VRD1 Volume of the XL-unit flow-meter (RD1) [µL]. 1, 97, 98

Veq Volume of the prototype equipment placed from the LV valve until the FC cell [mL]. 1, 80, 100

Vin j Volume of the injection loop [µL]. 1, 79, 80

W DLV O Energy potential [J = (kg.m2 )/s2 ]. 1, 38, 39

W ele Repulsion energy potential [J = (kg.m2 )/s2 ]. 1, 37, 38

W vdw Attraction energy potential [J = (kg.m2 )/s2 ]. 1, 37, 38

Z(φ) Compressibility factor [dimensionless]. 1, 42

∆T Temperature difference [K]. 1, 14, 73

∆ψ Electrostatic potential difference [V = (k g.m2 )/(A.s3 )]. 1, 14

∆ρ Density difference [K g/m3 ]. 1, 32

∆t Simulations time step [s]. 1, 48

∆t in Duration of the injection step in Section 3.2.2 model [s]. 1, 50

∆ y Mesh cell size [m]. 1, 48

Π Osmotic pressure [Pa = k g/(m.s2 )]. 1, 39, 40, 59, 60, 62–64

Π∗ Osmotic pressure scaled by the van’t Hoff coefficient (A vH ) [dimensionless]. 1, 63

Πele (φ) Electrostatic contribution to the osmotic pressure [Pa = k g/(m.s2 )]. 1, 59

Πent (φ) Entropy contribution to the osmotic pressure [Pa = k g/(m.s2 )]. 1, 59

Π vdw (φ) van der Walls contribution to the osmotic pressure [Pa = kg/(m.s2 )]. 1, 59

Π1 Osmotic pressure in the dispersed phase for φ < φ1 [Pa = k g/(m.s2 )]. 1, 60

Π2 Osmotic pressure in the transition region for φ1 < φ < φcr [Pa = k g/(m.s2 )]. 1, 60

Π3 Osmotic pressure in the transition region for φ2 > φ > φcr [Pa = k g/(m.s2 )]. 1, 60

Π4 Osmotic pressure in the condensed phase for φ > φ2 [Pa = kg/(m.s2 )]. 1, 60, 62

Π t r Osmotic pressure in the transition region φ2 > φ > φ1 [Pa = k g/(m.s2 )]. 1, 61, 63

γ̇ Shear strain [1/s]. 1, 42

ε Medium permitivity [F /m=(A2 .s4 )/(kg.m3 )]. 1, 36, 38

ε0 Vacuo permitivity [F /m=(A2 .s4 )/(k g.m3 )]. 1, 32

ε r Relative permitivity or medium dielectric constant [dimensioness]. 1, 32, 36, 39, 86

η Efficiency of separation [%]. 1, 53, 55

xxxv
γ Dimensionless function of wall potential [dimensionless]. 1, 37

λ Wavelength [nm]. 1, 77, 78, 87, 88, 90, 130

λ D Debye length [m]. 1, 36, 37, 86, 134

λemi Maximum emission wavelength [nm]. 1, 75, 76, 85, 87, 88, 90, 92, 126

λe x c Maximum excitation wavelength [nm]. 1, 75, 76, 85, 90

µ Fluid dynamic viscosity [Pa.s=k g/(m.s)]. 1, 8, 32–34, 42, 43, 45, 48, 49, 86

µ E Electrophoretic mobility [m2 /(s.V )=(A.s2 )/kg]. 1, 14, 86

µs Solution viscosity [Pa.s=k g/(m.s)]. 1, 43

φ Volume fraction [dimensionless]. 1, 40–43, 57, 59, 60, 62, 63

φ1 Volume fraction corresponding to the beginning of the transition region [dimensionless]. 1, 60,
62–64

φ2 Volume fraction corresponding to the end of the transition region [dimensionless]. 1, 60, 62–64

φcp Maximum close packing volume fraction [dimensionless]. 1, 41, 43, 62–64

φcr Critical volume fraction for phase transition [dimensionless]. 1, 41, 59–64

ψ Electrostatic potential [V = (k g.m2 )/(A.s3 )]. 1, 35, 37

ρ Fluid density [k g/m3 ]. 1, 14, 32–34, 45, 74

ρ E Charge density [C/m3 = (A.s)/m3 ]. 1, 36

ρm Mixture density [kg/m3 ]. 1, 57

ρ p Particles density [kg/m3 ]. 1, 14, 33, 34, 57

τ Geometrical residence time [s]. 1, 49, 50, 53, 99, 100

τs Shear stress [N /m2 = kg/(m.s2 )]. 1, 42, 45

ζ Zeta potential [V = (kg.m2 )/(A.s3 )]. 1, 32, 35, 37–39, 44, 85, 86, 92, 93, 123

a Particle radius [m]. 1, 32, 33, 36, 37, 86

b L AF4 channel minimum breadth [m]. 1, 16, 19, 30

bo AF4 channel maximum breadth [m]. 1, 16, 19, 30

c Light speed c = 3 × 108 [m/s]. 1, 75, 76

c0 Ions concentration in the bulk [mol/m3 ]. 1, 36

ci Double-layer ions concentration [mol/m3 ]. 1, 36

xxxvi
d Inter-particles distance [m]. 1, 37

d p Particles diameter [m]. 1, 7, 8, 10, 11, 34, 38, 39, 44, 48, 52, 54, 56, 63, 85, 86, 92, 93, 119,
120

ds Distance from the particle surface [m]. 1, 36, 37

e Electron charge e = 1.602 × 10−19 [A.s]. 1, 36, 38, 39

f (t) Particles flux [mol/s]. 1, 49

f c (K D a) Correlation function [dimensionless]. 1, 86

f10 Flux of particles with 10 nm [mol/s]. 1, 52

f50 Flux of particles with 50 nm [mol/s]. 1, 52

g Acceleration of gravity at the earth surface g = 9.8 [m/s2 ]. 1, 14, 32, 33, 45

h Planck constant h = 6.63 × 10−34 [J/s = (kg.m2 )/s]. 1, 75, 76

h f Heat flow-rate [J/s]=[(k g.m2 )/s3 ]. 1, 73

k Compressive yield stress [dimensionless]. 1, 62–64

kB Boltzman constant kB = 1.381 × 10−23 [J/K=(k g.m2 )/(s2 K)]. 1, 8, 14, 32, 33, 36, 38, 39

k D Debye-Huckel parametert [1/m]. 1, 36, 86

l Mean layer thickness [m]. 1, 8, 10, 19

l ∗ Normalized mean layer thickness [dimensionless]. 1, 9

l1 Mean layer thickness for larger particles [m]. 1, 8

l2 Mean layer thickness for smaller particles [m]. 1, 8

mo Sample amount detected at the channel outlet [µg]. 1, 94, 96, 100

min j Injected sample quantity [µg]. 1, 79, 100

n Number of moles [mol]. 1, 40

ni Number density of ions in the bulk solution [1/m3 ]. 1, 37

p Pressure [Pa = kg/(m.s2 )]. 1, 40, 45, 58

q Heat [[J]=[(kg.m2 )/s2 ]]. 1, 73

r Radial direction in spherical and cylindric coordinates. 1, 36, 58, 79, 96

t Time [s]. 1, 73, 77, 78, 87

t ∗ Normalized retention time (scalled by τ) [dimensionless]. 1, 49–54

xxxvii
t 0 Void time [s]. 1, 11, 105, 107

t a Accumulation time [min]. 1, 10, 57, 80, 81, 103–109, 114, 122, 136, 143

t e Elution time [min]. 1, 126, 127, 129–131, 134, 135

t r Retention time [s]. 1, 11, 49–51, 100

t in j Injected time [s]. 1, 99, 100

t t r ansp Time to transport the particles to the membrane surface [min]. 1, 80, 81, 105

w Channel thickness [m]. 1, 9, 14, 16, 19, 30

x Axial direction in cartesian coordinates. 1, 46

x 10 Mol fraction of particles with 10 nm [dimensionless]. 1, 52, 53

y Wall-normal direction in cartesian coordinates. 1, 46, 56

z Axial direction in cylindric coordinates and third dimension in cartesian coordinates. 1, 46, 58,
79, 96

z-av g z-average of the DLS auto-correlation function [nm]. 1, 85–87, 105–107, 110, 112–114, 126,
127, 130, 132, 135

zi Ions charge number [dimensionless]. 1, 36, 38, 39

S1′ High energy level [e.u.]. 1, 75

Uzi Axial velocity at the fiber inlet [m/s]. 1

Uzo Axial velocity at the fiber outlet [m/s]. 1

[η] Intrinsic viscosity [dimensionless]. 1, 43

xxxviii
1.1 Problem statement
1
Introduction

Nanoparticles are present in numerous industrial applications, reaching a global market of 7.6 billion
euros in 20191 with a forecasted annual growth of 13.1% between 2020 and 2027 [1]. The nano-
particles properties (size, shape, density, charge, surface coating, and stability) provide for diverse
products and functions (specialized materials, paints, drug delivery, etc.)[2], nevertheless, slight
variations in the medium conditions (ionic strength, pH, and concentration increase) has strong
effects in their surface interactions and governing phenomena. Such changes have implications
on process performance and product specification [3]. In particular, particle size is an important
parameter that highly affects the physical and chemical characteristics of the suspension [4]. For
instance, cerium oxide and silica oxide nanoparticles are known by their abrasive properties and
used for chemical mechanical polishing processes in microelectronics [5; 6; 7]. In this type of ap-
plications, the particle size plays an important role since it can cause surface damage or affect the
process removal rate [8; 9].

Nowadays, the industrial separation of nanoparticles in liquid-solid suspensions is challenging.


The separation of particles with a diameter inferior to 10 nm occurs by adsorption (depth filtration) or
ultrafiltration processes. The techniques used for the separation of particles with diameters superior
to 10 µm are the classical filtration, sedimentation, or centrifugation techniques. Between these two
domains a technology gap exists (Figure 1). In this size range, the available separation methods are
microfiltration and hydrodynamic separation techniques, which present limitations in sample feed
and performance due to the phenomena related to membrane-particle interactions (concentration
polarization, gelation, fouling, and clogging) [10].

1
The value was not actualized with the Harmonized Index of Consumer Prices (HICP). The value was converted to
euro with an average exchange rate of 0.8931€ in 2019 (online accessed: https://www.exchangerates.org.uk/
USD-EUR-spot-exchange-rates-history-2019.html on 26/06/2021 at 17:42)

1
Figure 1.1: Particle separation processes in different size ranges.

The industrial fractionation of nanoparticles is still a problem to be addressed. Several research


groups developed microfluidic techniques for the continuous separation of microparticles and cells
[11; 12; 13; 14], although these techniques are not applicable at the sub-micrometer scale, where
the Brownian diffusion is the predominant phenomena. At the analytical and batch scale, the nan-
oparticles size-based separation occurs by Size Exclusion Chromatography (SEC), Capillary Electro-
phoresis (CE), and Field-Flow Fractionation (FFF) techniques [4]. The SEC mechanism, in common
with other chromatography techniques, occurs in a column filled with a porous stationary phase,
where the pore size controls the retention time of the particles. The major limitation of SEC for nan-
oparticle separation is the interaction between the particles and the stationary phase, which affects
particle retention and induces sample loss. The smaller particles can circulate through the porous
and are retained for longer times than the larger particles [15]. The CE is a separation technique
based on the particles electrophoretic mobility (size to charge ratio). The sample circulates in a cyl-
indrical micro-channel made of silica, where is applied an electrical field. The force generated by
the electrical field is counterbalanced by the hydrodynamic drag force (laminar flow transporting the
particles), and results in an inferior retention time for the smaller particles [4]. The FFF technique
has, analogous to CE, an open channel configuration. The FFF separation mechanism relies on the
combination of two essential ingredients, an external field ("Field") and a laminar flux ("Flow"), that
separates the particles based on their size through their diffusion property. Compared to SEC, the
FFF technique presents a higher recovery as the empty channel decreases sample interaction with
the equipment [15; 16; 4]. In comparison with the CE, the FFF allows a higher sample throughput
(10000 times superior) and exhibits better experimental repeatability [4]. Additionally, many re-
search groups confirmed the capability of FFF to process higher sample quantities [17; 18; 19; 20]
and to be adapted to a continuous operation [21; 22].

1.2 Objective
The purpose of this Ph.D. is the evaluation of the feasibility of an FFF-based process for the indus-
trial separation of liquid-solid nanoparticles dispersion within the size range of 1 nm - 1 µm, through
the design and production of a prototype based on the Hollow Fiber Flow Field-Flow Fractionation

2
technique (HF5), and its separation performance assessment. The objective is the production of a
separation device with a viable industrial scale-up strategy and a quasi-continuous or continuous
operation, the identification of the critical parameters, and the definition of optimal operating con-
ditions.

1.3 Thesis structure


This manuscript aims to answer the research question "Is a separation process based on the FFF
mechanism viable for the industrial separation of nanoparticles?", being divided into eight chapters,
where is presented a bibliographic review of the FFF technology, the physics governing the behavior
of the nanoparticles near membrane surfaces, the development of a sequential experimental proto-
type guided with a CFD model, the exploration of the prototype operating conditions and evaluation
of its separation performance, and the developments towards a continuous operation. The chapters
are organized as follows:

• Chapter 2: presents the FFF principle, mechanism, and sub-techniques. The FFF separation
mechanism requires the combination of two essential ingredients: an external field applied
to locate the particles at different distances from the wall-normal direction (accumulation
wall) and a laminar flow for the particles transport and increase of the axial distance between
different-sized populations. The emphasis is given to the Flow-FFF configurations, in par-
ticular the asymmetrical planar (Asymmetric Flow Field-Flow Fractionation - AF4) and the
symmetrical cylindrical (Hollow Fiber Flow Field-Flow Fractionation - HF5) configurations.
In AF4 and HF5, the external field applied is a wall-normal permeate flux, where the wall is
a semi-permeable membrane. The operating principle requires an injection, accumulation in
a confined zone (defined among the FFF community as "focusing" or "focusing + relaxation"
step), and an elution step. The particles are injected into the channel and accumulated in
a small area near the channel entry through the inversion of the outlet flow-rate (stop-flow
step). This step is responsible for the formation of particle layers, where the distance from the
accumulation wall (membrane) depends on the particle size (larger particles are closer to the
membrane than the smaller ones) and is behind the challenge of increasing sample processing
or operating in continuous. During elution, the particle layers occupying distinct wall-normal
locations are placed at different streamlines (laminar flow nature) and transported with dif-
ferent velocities. The smaller particles that reach higher speed zones have a shorter retention
time, thus a size-based separation is achieved. In Chapter 2 are discussed the operating con-
ditions affecting the AF4/HF5 separation performance and is presented the studies developed
by several research groups to improve its separation capacity and operating mode. At the end
of the chapter is given the thesis positioning comparing to the related work.

• Chapter 3 discusses the physics of colloids that are behind their complex phenomena and
presents the OpenFOAM CFD models developed to guide the initial conception of the experi-
mental prototype and later improved to account for the influence of the particles concentration
in the suspension motion. Nanoparticles are sol colloidal dispersions (solid phase dispersed
in a liquid medium) with complex interface phenomena. Depending on their density, they

3
are governed by the Brownian motion in a wide size range. The Brownian diffusion favors
particle proximity and collisions, which increases the osmotic pressure. The interaction po-
tential between two particles is described by the DLVO (Derjaguin-Landau-Verwey-Overbeek)
theory as the additional contribution of the repulsive electrostatic interactions and the van der
Waals attractive interactions. The DLVO shows that when the Brownian motion has kinetic
energy sufficient to overcome the electrostatic barrier, the particles suffer irreversible aggreg-
ation. Therefore, particle stabilization requires the electrostatic barrier increase, which can
be adjusted with the solvent properties, for instance, by regulating the ionic strength and the
pH. The augmentation of sample concentration leads to a phase transition from a dispersed
(disorder) to a condensed phase (order). The disorder-order phase transition is associated
with a discontinuity in the experimental curves of osmotic pressure as a function of the con-
centration. The osmotic pressure is the colloids equation of state (measures the resistance to
removing solvent molecules located between the approaching particles), being often used in
CFD models to account for multi-particle interactions and their effecr in the suspension dy-
namics. A 2D CFD model was initially developed to simulate the Flow-FFF mechanism and to
investigate the key parameters and geometrical configuration of the experimental prototype.
The model was latter improved by accounting for inter-particle interaction and using an axis-
symmetry geometry, which is ready for a future prototype optimization. An osmotic pressure
based model was developed to account for the colloids dynamics near a dispersed-condensed
phase transition, which was used to investigate the mechanisms of a gel layer relaxation (Art-
icle submitted to the Colloids and Surfaces A: Physicochemical and Engineering Aspects Journal:
"Colloid dynamics near phase transition: simulation of the relaxation of concentrated layers"),
and later used to investigate the threshold conditions that lead to the irreversible blockage of
a hollow-fiber capillary (Abstract submitted to an oral presentation at the EUROMEMBRANE
2021 conference: "Plugging of a hollow fiber capillary by gel relaxation during a cleaning step";
Ongoing article).

• Chapter 4 presents the development of a digital process based on the HF5 concept. The proto-
type is composed of micro/milli-fluidic equipment and works with a fixed-pressure operation.
The separation mechanism occurs in a hollow-fiber module. An online fluorescence spectro-
scopy system (OFS) was integrated into the prototype to allow the real-time visualization of
the elution peaks, which saves analysis time and permits a faster decision on the succeeding
operating conditions to test. The equipment operating mode and specification are given. The
prototype works with a sequential batch operation (SPFFF), composed of three operating steps
automatically controlled by software: injection, accumulation, and elution (which can include
a backwash step). The SPFFF accumulation step was simplified to avoid the flow inversion
required for the HF5 accumulation in a confined region. The SPFFF was tested using model
particles (latex polystyrene and silica oxide) marked with fluorochromes. The particle proper-
ties (optical properties, size, surface charge, and titration curves) are presented. The methodo-
logy for the determination of the particles calibration curve in the OFS is provided. The offline
analytical methods for the size (DLS) and surface charge (electrophoresis) determination are
described. The composition of the solvent properties and their impact on the detected signals
by the OFS system is discussed. The verification of the system hydrodynamics was performed

4
by measuring the sample signal at different locations of the prototype and calculating the mean
retention time and mass balance.

• Chapter 5 contains the exploration of the operating conditions using model latex polystyrene
particles with 100 nm. The tests performed revealed the existence of a permeate flow-rate
limit, below which the diffusion of the particles overcome the drag force and permits particles
re-suspension and evacuation from the fiber. The optimization of the accumulation conditions
for the more diffusive particles and the knowledge of the threshold conditions for the permeate
flow-rate permitted a methodology prescription for the size-base separation using the SPFFF
prototype. Additionally, we prove that the diffusion, and consequently the particles size, is the
discriminate property for the fractionation if the suspension has a negative charge, is stable,
and the interactions with the membrane are reduced (which are strongly dependent on the
membrane and solvent properties). An experiment applying the same accumulation and elu-
tion conditions for latex polystyrene and silica oxide nanoparticles with 100 nm resulted in the
same sample retention.

• Chapter 6 presents the application of the prescribed methodology for the separation of latex
polystyrene nanoparticles with 100 nm and 500 nm. The separation optimization was invest-
igated by varying the elution permeate flow-rate and the solvent composition. The process
intensification was tested by increasing the concentration of the injected sample and perform-
ing a five-cycle operation. The cyclic operation highligthed the need for the system to operate
with a fixed-flow operation that guarantees an accurate permeate flow-rate.

• Chapter 7 contains the progress towards a continuous operation inspired by the work de-
veloped by Marioli et al. [22]. This chapter has been removed for confidentiality reasons
(ongoing analysis for patent application), being converted into an internal report for Solvay.

• Chapter 8 gives the general conclusions and perspectives for the SPFFF optimization, and the
steps towards the scale-up and continuous operation.

5
6
2.1 The Field-Flow Fractionation Technology
Field-Flow Fractionation
2
In the 1960s, Giddings J. C. and his coworkers developed the Field-Flow Fractionation (FFF) tech-
nique, an analytical tool for characterization and separation in the submicrometer range. Initially,
its purpose was to overcome some of the limitations imposed by classical chromatography columns,
such as the shear degradation of high molecular weight polymers. Over the years, FFF became a
deeply investigated technique, and different sub-techniques and channel configurations emerged
[23; 16; 24]. Nowadays, FFF is a versatile technology applied to the various fields of biotechnology
and analytical chemistry, and its limitation is in the scalability of the analytical process for indus-
trial applications. The following sections cover the FFF system and its advances for the throughput
increase in batch processes and continuous operations.

2.1.1 Principle and mechanism


The FFF mechanism is based on two essential ingredients, a laminar flow and an external field per-
pendicular to the flow direction, causing a differential displacement. A classic FFF channel presents
a planar geometry, empty inside, and a thickness O(∼ 1000) times smaller than its length ("ribbon-
like channel"). Figure 2.1 illustrates the FFF mechanism applied for polydispersed colloidal particles
in normal elution mode (elution modes discussed in Section 2.1.1.2). Once the colloids with dif-
ferent diameters (d p ) are injected into the channel, they face an external field driving them to the
accumulation wall. Simultaneously, their Brownian diffusion (D) counteracts the drag force moving
the particles in the opposite direction. As the smaller particles present higher diffusion coefficients
(D ∼ 1/d p ), they occupy regions of the channel closer to the center than the larger ones. The lam-
inar flow profile is characterized by having a no-slip condition at the walls, a maximum velocity in
the channel center, and a continuous velocity increase between these two points (parabolic nature).
Therefore, the particles located closer to the wall are transported with smaller velocities than the
smaller particles, resulting in different retention times. The separation is obtained by collecting the
fractions in distinct moments of the elution. The smaller particles (more diffusive) leave the chan-
nel before the larger particles. The mechanism presented in Figure 2.1 illustrates the size-based
separation of colloidal particles however, FFF can be applied to separate other species with molecu-
lar weight ranges of 103 − 1018 g/mol and size ranges of 1 nm − 100 µm, depending on different
properties such as shape, surface composition, molecular weight, and density.

In sum, in Field-Flow Fractionation (FFF) the "Field" is responsible for sample accumulation at
the accumulation wall, which is essential to obtain a differential sample radial migration, and the

7
"Flow" is the parabolic profile of the carrier fluid (fluid that transports the sample towards the chan-
nel outlet) that as the role of increasing the distance between different sample components during
elution (resulting in distinct retention times between sample components). The combination of the
"Field-Flow" ingredients lies behind the fractionation. The next sections detail the accumulation and
elution steps of the classic FFF techniques.

Figure 2.1: Scheme showing the Field-Flow Fractionation mechanism of particles with two different
sizes. D1 and D2 represent the Brownian diffusion of larger and smaller particles, respectively (D1 <
D2 ). The magnitude of the external field (F y ) is the same for both particles (F y = F y1 = F y2 ). The
figure is merely illustrative and it is not at the right scale. The particle layers are at O(∼ 10 µm) from
the accumulation wall (1 − 10% of total channel width).

In the following sections, we consider the cartesian coordinates system where x stands for the
axial direction and y for the wall-normal direction.

2.1.1.1 Accumulation

During the accumulation step, the driving force originated from the external field drives the particles
towards the channel wall (designated as accumulation wall). According to the sample polydispersity,
the particles are located in layers with different distances from the accumulation wall (depending
on their size). Figure 2.2 represents an ideal accumulation of a population with two sizes. Note
that in reality there are regions where particles of different sizes are mixed together, as Figure 2.3
shows. The amount of smaller particles near the accumulation wall (black line) is smaller but not
negligible. The length dimension of the accumulation stage is defined as a mean layer thickness (l),
which represents the center of mass for a steady-state accumulated layer (above l is half of the sample
mass and below l is the other half). The larger particles are located closer to the accumulation wall
and have a mean layer thickness l1 . The smaller particles are located further way from the wall with
a mean layer thickness l2 . The mean layer thickness of the accumulated layer is a consequence of
the balance between the diffusion of the particles due to the thermal energy (KB T , where kB is the
Boltzmann constant and T is the temperature) and the energy associated to the Stokes drag force
driving them to the wall (F y = 3πµd p U y , where µ is the carrier fluid viscosity, d p is the particle

8
diameter, and U y is the velocity of the sample towards the accumulation wall), calculated according
to Equation 2.1 [16; 25].

KB T D0
l= = (2.1)
Fy w Uy

Where w is the channel thickness and D0 corresponds to the Brownian diffusion in diluted con-
ditions obtained from the Stokes-Einstein relationship (Equation 3.15 presented in Section 3.1.5.1).

Figure 2.2: Illustration of the layers formed during the accumulation of a dispersion with two sizes.
l1 and l2 correspond to the mean layer thickness of the large and small particles, respectively. Notice
that the figure is merely illustrative. In reality, the particle layers have mixed regions due to inter-
particle interactions.

Equation 2.2 defines the dimensionless form of the man layer thickness (l ∗ ) as:

l D0
l∗ = = (2.2)
w Uy w

This is analogous to represent l ∗ = 1/Pe y , being Pe y the accumulation Peclet number in the wall-
normal direction, defined according to Equation 2.3. The Peclet number is a dimensionless quantity
representing the ratio between the transport by convection and diffusion. For the accumulation
stage, both transport mechanisms occur in the wall-normal direction.

Pe y = U y w/D0 (2.3)

As a consequence of the sample accumulation, a steady-state concentration profile (C/C y , C the


concentration and C y the concentration at the wall) is developed at the accumulation wall, which is
described according to Equation 2.4 [16; 25].

−y
C/C y = e x p(− y/l) ⇔ C/C y = ex p( Pe y ) (2.4)
w

The wall-normal concentration profile (in the y-direction) in the channel is higher at the ac-
cumulation wall and decreases exponentially with the increase of the distance from the wall. The
exponential decrease with y is faster for higher Pe y values, meaning that the larger particles (less

9
diffusive - higher Pe y ) are more concentrated at the channel wall. In the other hand, the smaller
particles (more diffusive - smaller Pe y ) are located at higher distances from the wall. Figure 2.3
illustrates the concentration profiles for particles with a size ratio of 10.

Figure 2.3: Normalized concentration profiles for two particles with a size ratio of 10 as function of
the normalized distance from the wall.

Higher Pe y numbers result in layers more concentrated at the accumulation wall (smaller mean
layer thickness l). The Pe y depends on the sample size (d p ) and on the magnitude of the external
field force (F y ∼ U y ), thus the increase of the Pe y through F y originates layers with smaller l (closer
to the accumulation wall).

The operating conditions of the the accumulation stage are the accumulation time (t a ) that is
required to achieve a steady-state concentration profile and the Pe y through the imposition of the
magnitude of the external field force (F y ).

2.1.1.2 Elution

The elution step is responsible for the particles evacuation from the channel with distinct retention
times. In Field-Flow Fractionation (FFF), there are two main types of elution designated as normal
and inversed. The elution mode of FFF is associated with the wall-normal position of the laminar
flow occupied by the sample components that affect their retention time (according to the y-direction
location, the particles are placed at different streamlines and the particles closer to the channel
center occupy regions of higher velocity). In the normal elution mode, also called Brownian mode,
the species retention is related to their capacity to diffuse, resulting in shorter retention of the more
diffusive objects. In general, this is true for the size range of 1 nm−1 µm [16; 25]. For larger species,
the diffusion becomes negligible, and in the absence of a force to counteract the external field, the
differential displacement along the channel height between the smaller and larger components does
not occur. The sample components form thin layers at the wall, and the large objects reach higher
speed channel regions due to their dimensions, leaving the channel before the small ones. In this case
the elution is called steric mode [16; 26; 25]. When the object size exceeds 10 µm, the hydrodynamic
lift forces hinder the objects from touching the wall, and the elution is defined as hyperlayer mode
[13]. In short, in normal elution mode, smaller objects are transported faster than larger ones, and
the opposite occurs in inversed elution mode (steric or hyperlayer).

10
The elution mode is an important parameter to characterize a sample or to perform a separa-
tion. For the characterization of unknown sample properties, the measured retention time correl-
ates through the FFF retention theory (defined elsewhere [16] for both elution modes) to object
diffusion. The diffusion coefficient (D0 ) can predict the hydrodynamic diameter (d p ) through the
Stokes-Einstein relationship. The validity of the retention theory depends on the elution mode, thus
to correctly predict the sample properties, it is necessary to use the retention theory related to each
elution mode.

For a separation process, the coexistence of the two elution modes hinders the separation. To
illustrate, in an experiment for a polydispersed sample with a size range of 100 nm to 10 µm the
particles may experience the same retention time since the smaller particles are in normal elution
mode and the larger particles in the inversed mode. Thus, the separation may not occur, and the
characterization is not valid because the elution peaks overlap and the detection equipment inter-
polates the same properties for different species.

The scope of this work is the separation of colloidal particles in the size range of 10 nm - 1 µm,
therefore only normal elution mode is considered. In the retention theory of the normal elution
mode are defined two parameters to evaluate the separation potentiality or to obtain the sample
dimensional properties. These parameters are the retention time (t r ) and the retention level (R L ).
The t r is obtained from Equation 2.5, considering a steady-state concentration profile as starting
conditions, uniform external field along the entire channel length and for a parallel plates configur-
ation. The retention level, defined according to Equation 2.6, is the ratio between t r and the void
time t 0 (time required for unretained samples to leave the channel).

t0
tr = (2.5)
6l (coth(1/2l ∗ ) − 2l ∗ )

tr 1
RL = = ∗ (2.6)
t0 6l (coth(1/2l ∗ ) − 2l ∗ )

F y D0
Being U y = KB T , the t r can be expressed as:

t0 tr 1
tr = ¨ −wU y
«⇔ = ¦ −Pe y © (2.7)
t0 6 1+e
− 2
D 1+e D0 2D0 Pe y 1−e−Pe y Pe y
6 wU0y −wU y − wU y
1−e D0

From the equation 2.7, it is evident that the retention of the samples depends on their diffusion
coefficient (D0 ) and velocity imposed by the strength of the applied external field (U y ), thus the t r
depends on Pe y . For a characterization point of view, the experimental result of t r gives the D0 value
which can be correlated a stokes diameter (d p ), as mentioned previously.

Figure 2.4 demonstrates the impact of the sample size (characterized by Pe y ) on the retention
time (scaled by t 0 - t r /t 0 ), when the sample components are subjected to the same external field

11
strength. The larger particles have higher Pe y values because of their smaller D0 , being more retained
than the smaller objects.

Figure 2.4: Normalized sample retention time (t r /t 0 ) according to the particles size (characterized
by their Pe y number). The calculations were performed considering a channel with w = 0.7mm, an
accumulation velocity of U y = 1 × 10−6 m/s, and particles with 7, 10, 50, 100, 200, and 500 nm.

2.1.2 Operating mode of the FFF technology


In the Section2.1.1 was described the FFF physical phenomena that is behind the separation process.
From a technological point of view, FFF is a complex technique that is system-dependent, requiring
the selection of an appropriate external field that can drive the sample towards the accumulation
wall (accumulation stage), the use of adequate operating conditions (flow-rates, steps duration, and
physicochemical properties of the carrier fluid), the awareness of the elution mode (to properly
characterize or separate the samples), and the choice of the adequate online detectors. Figure 2.5
summarizes the complexity and technological options of the FFF separation method.

To elucidate the dependency of the FFF with the properties of the sample, the characterization
of polydispersed gold microparticles and the fractionation of Bovine Serum Albumin (BSA) protein
dimers (stokes diameter O(∼ 7 nm)) cannot be performed using the same FFF configuration. The
properties of the samples differ in size, shape, density, and surface composition resulting in different
physics (suspension stability and governing phenomena as discussed in Section 3.1.2). The gold
microparticles are affected by the gravity, and their retention time is governed by steric effects (in-
versed elution mode), while the retention of the BSA is merely governed by the Brownian diffusion.
Also, the carrier fluid properties have to stabilize sample components. For instance, a salt is required
to stabilize the structure of the BSA but the use of a salt with gold microparticles would promote
aggregation (destabilization) due to the increase of the ionic strength that reduces the Debye length
and favors attractive interactions as discussed in Section 3.1.3.1 to 3.1.3.3. Thus, the external field
and the operating conditions have to suit the physicochemical properties of the sample.

The sections 2.1.2.1 and 2.1.2.2 cover the coupling with online analytical methods, and the
possibilities for the external fields, respectively. Since the operating conditions depend on the sample
in analysis and on the applied external field, they are discussed in section 2.2.3 for the Flow Field-
Flow Fractionation (Flow-FFF) sub-techniques.

12
Figure 2.5: Scheme illustrating the Field-Flow Fractionation (FFF) complexity.

2.1.2.1 Detection techniques

For analytical characterization of the samples and quantification of the separation efficiency, it is pos-
sible to couple FFF with online analytical techniques, being commonly used the Multi-Angle Light
Scattering (MALS), Spectroscopy techniques (absorbance and fluorescence), Refractive Index (RI),
and ICPMS (Inductively Coupled Plasma-Mass-Spectrometry) [16; 27]. The choice of the more suit-
able analytical method depends on the sample properties and the purpose of the analysis. The MALS
is used to characterize macromolecules according to their molar mass, structure ratio, and size. The
RI is a concentration detector that measures the difference of the light speed in different mediums
and applies to all sample types. The absorbance and fluorescence spectroscopy are also concentration
detectors, but they depend on the optical sample characteristics. The ICPMS is a destructive analysis
that detects the concentration of atomic elements. The selection of the online detector depends on
the sample properties, the information required, and the need to preserve the sample.

2.1.2.2 External fields

The application of FFF with high efficiency depends on the sample properties and the right choice of
the external field. The samples can be of natural or synthetic origin, have different natures (macro-
molecules, living organisms, organic or inorganic solids), and present different shapes or structural
states (amorphous or crystalline). The separation of the sample components occurs due to their
distinct dimension parameters (size, shape, and molar mass) and chemical properties (elementary
composition, surface composition, and concentration), dictating the choice of the external field [16].

13
The nature and properties of the external field designate the sub-techniques of FFF, being the
most commonly used and with commercially available solutions the centrifugation or gravity, a tem-
perature gradient, an electric current, and a cross-flow (referred to as permeation flow hereafter).
When the applied field is centrifugation or gravity, the sub-technique is called Sedimentation-FFF
(SdFFF), and the sample needs to be sensitive to the gravity force [16]. Section 3.1.2.4 shows that
the impact of the gravity force depends on the particle size and density. In Thermal-FFF (ThFFF),
the channel is composed by a lower cold wall and an upper hot wall that induces a temperature
gradient. This system is mainly applied to organic-soluble samples because the property responsible
for the separation is the thermophoresis opposing the Brownian diffusion [16; 28]. In Electric-FFF
(E1FFF), the channel is composed of an upper wall positively charged (cathode) and a lower wall
negatively charged (anode), causing species migration according to the surface [16]. The use of this
technique implies different electrophoretic mobility between sample components. In the Flow-FFF
sub-techniques (F1FFF, AF4, and HF5) is a permeate flow that induces the species differential mi-
gration, counteracting the Brownian diffusion and driving the species to an accumulation wall made
of a semi-permeable membrane [16; 25; 27]. Table 2.1 summarizes the FFF sub-techniques in terms
of the field nature, applicability range, and separation parameters. The external field velocity (U y )
is presented for each sub-technique, so the retention time (for sub-techniques using parallel plate
configuration) can be deduced from replacing the respective U y in Equation 2.7.

Some other possible fields exist, such as magnetic, dielectric, photophoretic, concentration
gradient, and acoustic, although the experimental setup will have to be lab-made [16].
Table 2.1: Summary of the Field-Flow Fractionation sub-techniques.

Wall-normal velocity
FFF sub-techniques External Field Separation by Target samples range
(U y )
Vp D0 colloids and large particles
SdFFF Sedimentation or gravitation KB T (ρ − ρ p )g size and density differences
50 nm - 100 µm [16; 28]
particles and organic-soluble polymers
ThFFF Temperature gradient DT ∆T
w size and thermophoretic mobility
104 - 107 g/mol [16]
∆ψ
E1FFF Electrical Field |µ E | w size and electrophoretic mobility biological materials [16]
Qy macromolecules, colloids, and larger particles
F1FFF/ AF4/ HF5 Permeation Flow V0 w size
1 nm - 100 µm [16]
Nomenclature: Vp - particle volume; D0 - diffusion coefficient in diluted conditions; kB - Boltzmann constant; T - temperature; ρ - fluid
density; ρ p - particles density; g - acceleration of the gravity; DT thermophoretic mobility (Soret coefficient DT /D0 ); ∆T - temperature
difference between upper and lower plates; w - channel thickness; µ E - electrophoretic mobility; ∆ψ - potential difference between the
lower and upper plates; Q y - permeate flow-rate; V0 - channel volume.

The use of the Flow-FFF techniques for size-based separation of colloidal particles presents the
most simple channel configurations and an extensive range of applications, depending only on the
magnitude of the permeate flux (drag force) and the Brownian diffusion of species (property of all
colloids) [28]. In section 2.2, the Flow-FFF techniques are discussed in more detail since this project
relies on their separation mechanism. Appendix A presents the channel configuration for SdFFF,
ThFFF, and E1FFF and the commercially available solutions from the two principal suppliers Wyatt
and Postnova.

14
2.2 Flow Field-Flow Fractionation
In Flow Field-Flow Fractionation, the external field is a permeate flow, and the channel has three
possible configurations, resulting in the sub-techniques Symmetrical Flow-Field Flow Fractionation
(F1FFF), Asymmetric Flow Field-Flow Fractionation (AF4), and Hollow Fiber Flow Field-Flow Frac-
tionation (HF5). Their different designations are related to the channel geometry and the origin of
the permeate flow.

2.2.1 Configurations
In F1FFF, as shown in Figure 2.6, the channel has the classical parallel plates configuration, is com-
posed of a porous solid upper wall and a lower wall covered by a semi-permeable membrane. The
permeate flow comes from a secondary pump and enters the channel from the upper wall.

Figure 2.6: Scheme illustrating the edge view of the F1FFF configuration. The permeate flow enters
the channel from an external source and is represented by the orange arrows.

In AF4, the channel also presents the parallel plates configuration and a lower wall covered by
a membrane, differing from F1FFF by the existence of a non-porous upper wall made of plexiglass.
The permeate flow originates from the axial flow due to the transmembrane pressure (asymmetric
arrangement). Figure 2.7 illustrates the AF4 channel configuration.

Figure 2.7: Scheme illustrating the edge view of the AF4 configuration. The permeate flow comes
from the mainstream flow due to the transmembrane pressure and is represented by the blue arrows.

In HF5 the channel is a hollow fiber membrane (cylindrical geometry), and the permeate flow,
similar to AF4, comes from the mainstream flow rate due to the transmembrane pressure. Figure

15
2.8 illustrates the HF5 configuration.

Figure 2.8: Scheme illustrating an axial cut of the HF5 channel. The permeate flow comes from the
mainstream flow due to the transmembrane pressure and is represented by the blue arrows.

The F1FFF technique was the first Flow-FFF designed [23], and AF4 appeared one decade later
to simplify the F1FFF configuration [29]. As the permeate flow comes from the mainstream flow,
only one pump is required, reducing operation and maintenance costs [15]. Furthermore, AF4 has
the advantage of experiment visualization during run time through its plexiglass wall. The first
experiments using a hollow fiber channel emerged in 1974 [30], and in 1989 was published its
retention theory [31]. Among the Flow-FFF sub-techniques, the AF4 substituted the F1FFF and is
the most widely used and investigated [16; 25; 15].

The AF4 commercial channel is presented in Figure 2.9. The channel configuration consists of
upper and lower walls made of stainless steel having the upper wall a plexiglass window, two O-rings
providing a good fitting between the channel components, a spacer made of Polytetrafluoroethylene
(PTFE), an ultrafiltration membrane, a porous frit to support the membrane, and a temperature
regulator. The spacer is responsible for the channel geometry and dimensions and has a trapezoidal
shape to keep a constant axial channel velocity [27]. The classic channel dimension are: Length
L= 20 − 30 cm, maximum breadth bo = 2 − 2.5 cm, minimum breadth b L = 0.5 − 1.0 cm, and width
w= 75 − 500 µm.

Figure 2.9: Scheme illustrating the AF4 commercial channel. Adapted from [32]

16
For the HF5, Wyatt supplies a disposable cartridge illustrated in Figure 2.10. The channel sim-
plicity and low cost is the highest advantage of the HF5 configuration since it can be easily lab-made
by gluing the two extremities of a hollow fiber to a plastic module [28]. The typical dimensions of
the commercial channel are a length L = 20 cm and a radius (R f ) of 0.4-0.5 mm [33].

Figure 2.10: Scheme illustrating the HF5 commercial channel.

2.2.2 Operating steps


The experiments of Flow-FFF sub-techniques have three steps, injection, accumulation, and elution.
The symmetric and asymmetric configurations have some differences outlined when relevant, while
AF4 and HF5 have analogous steps. The steps are illustrated using the AF4 configuration to facilitate
following the discussion in Section 2.3. A representation of the experimental steps for HF5 is given
in Annex B.

During injection, the sample enters the channel from the injection port in the AF4 configuration
or from the inlet in the HF5 configuration. At the same time, there are two currents of the carrier
fluid entering the channel, one coming from the inlet and another through the outlet. These two
currents meet the sample in a thin region near the entrance. All the fluid entering the channel leaves
as a permeate flow through the membrane. Figure 2.11 shows the injection step.

Figure 2.11: Scheme illustrating the injection step in an AF4 configuration.

After injection, the sample is accumulated at the membrane surface as illustrates Figure 2.12).
In AF4, the accumulation occurs in a confined region (along the axial channel direction) near the

17
channel inlet and is designated, among FFF community, as a focusing/relaxation step. The accumu-
lation of the sample in a thin region is obtained through the inversion of the outlet flow (stop-flow
conditions), that start occurring during the sample injection. The inlet and outlet flows create the
streamlines that drive the sample to a confined region at the membrane surface. The purpose of
the accumulation step in a small region is to obtain an efficient sample fractionation by starting the
experiment from an equilibrium condition (it improves the repeatability once the starting conditions
are equivalent [16]) and having a large membrane surface available for the separation during the
elution step. However, the stop-flow condition can induce hydrodynamic instabilities (resulting in a
mixing of the accumulated layers), requires an extra pump, and is the main challenge in adapting
the Flow-FFF to a continuous operation.

In the original F1FFF technique, the accumulation occurred by stopping the axial flow and driv-
ing the sample into the membrane with the permeate flow. A split inlet and a frit inlet were imple-
mented in the F1FFF channel to achieve concentration equilibrium using a hydrodynamic relaxation
that did not require a stop-flow condition [34].

Figure 2.12: Scheme illustrating the focusing/relaxation step in an AF4 configuration.

The last step is the elution (Figure 2.13), where the samples are transported downstream by the
laminar flow of the carrier fluid. The permeation flux is maintained to keep species retained until the
end of the experiment, being essential for the separation and required for sample characterization
(otherwise the retention theory is not valid). Different permeate profiles1 can be used during the
elution step (linear or exponential decays), and its choice requires experimental investigation. In
section 2.2.3 is discussed the impact of the permeate in the separation efficiency.

Figure 2.13: Scheme illustrating the Elution step in an AF4 configuration.

1
Notice that the retention theory depends of the permeate flux profile [16].

18
2.2.3 Impact of the operating conditions
Several are the experimental conditions impacting the separation performance in Flow-FFF. This
section covers the most relevant ones.

1. Accumulation conditions: The accumulation step in AF4 has three parameters, the flow-
rates, the duration, and the focusing position. The flow-rate magnitude impacts the system
similarly to the permeate flow rate during elution, causing undesired attractive interactions
or incomplete sample accumulation at the membrane surface. The non-accumulated species
remain dispersed in the channel bulk and leave in the void peak (unretained sample) when the
elution starts, affecting separation performance. The duration of the accumulation step has
to be long enough to reach a radial steady-state concentration profile [15; 35]. Concerning
the accumulation position, it impacts the separation performance because its proximity to the
channel outlet reduces the available space for the separation [35].

2. Elution conditions: The elution step has two operating conditions, the permeate and axial
flow-rates. The permeate flow rate is one of the most critical parameters in Flow-FFF exper-
iments, being responsible for the differential retention of sample components. The strength
of the permeate flow during the elution step has two limits, an upper limit where it causes
sample-membrane absorption, sample aggregation, and peak broadening, and a lower limit
where the sample is not well retained, resulting in a premature elution in the void peak or a
poor fractionation between sample components [15; 36; 37]. Peak broadening is the sample
elution for a long time, resulting in broad, flat peaks, thus increasing dilution [37]. In an
FFF experiment, it is common among experimentalists to use a programmed decay of the per-
meation flux during the elution step [15]. The initial value has to retain the more diffusive
components to avoid elution in the void peak. A linear decay to a no-permeate flux condi-
tion allows the gradual release of the less diffusive species. The commercial equipment from
Wyatt and Postnova have the functionalities to adjust the permeate flow profile. For lab-made
prototypes, it is possible to program the permeate flow through software such as Microfluidic
Automation Tool (MAT) from Fluigent or LabVIEW.
The axial flow-rate impacts the shape of the velocity profile. High flowrates decrease the
possibility of sample-membrane interactions and reduce sample retention time [15].

3. Channel dimensions: The channel dimensions in HF5 configuration are the fiber radius R f
and length L. In the AF4 configuration are the channel breadths (bo and b L ), channel width
w, and channel length L. The increase of the HF5 radius or the AF4 width while keeping the
same flowrates leads the flow velocity to decrease. The transport velocity of sample layers is
reduced, increasing their retention time [15; 37]. Bria et al. [35] studied the impact of AF4
dimensions on separation efficiency by varying the channel breadth and increasing the sample
amount. They affirm that the increase of the channel breadth and width permits the fraction-
ation of higher sample amounts. It makes sense for the breadth because its increase results in
a higher membrane surface available during the accumulation step. However, the formation
of the layers only occupies a small percentage of the total channel width. K.G. Wahlund [25]
suggests that thin channels result in a smaller distance of the sample layers (l) from the ac-

19
cumulation wall having the disadvantage of sample interactions with the membrane. In this
case, it is reasonable to expect that increasing the sample amount would promote sample ad-
sorption on the membrane and that larger channel widths could be used to avoid it. For the
HF5 configuration, the radius can be a critical parameter since its increase results in a higher
membrane surface area available during accumulation. The role of the AF4 width and HF5
radius for the fractionation of higher sample amounts is not clear, and no publications show
experimental results to confirm it. Although, the retention theory suggests that the increase
of the channel width results in longer experiments and higher sample dilution and affects the
peaks shape [38; 37; 25].

For those who desire to use Flow-FFF as an analytical tool for sample characterization, it is
necessary to know that to obtain correct sample properties from the retention theory, the ef-
fective channel thickness or radius needs to be determined according to the protocols defined
elsewhere [25; 37]. In AF4, the membrane suffers a compression by the spacer and intrudes
into it, resulting in smaller channel thickness [25; 37]. In HF5, the fiber can expand due to
the channel pressure, increasing the fiber radius (R f ) [38].

The channel length (L) impacts the space available for separation. It has to be long enough to
allow differential sample transport. If the separation is efficient with a given channel geometry,
increasing the length only affects the retention time.

4. Solvent (carrier fluid) and sample properties: The solvent properties (pH, ionic strength,
and surfactants addition) are essential to reduce surface attractive interaction between sample
components and between the sample and the membrane by adjusting its physicochemical
properties [37]. The only requirement for the solvent selection is that it is compatible with
the sample and the membrane [37], i.e, does not induce sample aggregation or damage the
membrane. Concerning the solvent temperature, the sample diffusion increases with the tem-
perature (D ∼ KB T ), allowing the species to reach higher velocity regions and reducing the
retention time. Thus, for the separation in the colloidal range, the temperature only affects
the retention time and can be used if it is interesting to reduce the experimental time. If the
colloids are near the inversion point of the elution mode, a temperature increase can be used
as a strategy to increase the sample diffusion and keep the normal elution mode [27].

5. Sample quantity: In analytical AF4, the maximum sample injection is O(∼ 100 µg) [35],
corresponding to O(∼ 5 µg/cm2 ) for a classical channel with a membrane surface of O(∼
20 cm2 ). The increase of sample quantity has a negative consequence in FFF designated as
overloading, which affects the sample retention time and induces peak broadening [37]. Bria
et al. [35] studied the effect of the injected quantity in the separation efficiency of silica
nanoparticles with 50 and 100 nm. The sample amount varied from 0.4 mg to 20 mg and its
increase resulted in a shorter retention time and worst peak resolution. S. Podzimek [27] says
that the overloading phenomena is caused by the viscosity increase during the accumulation
step, reducing the sample diffusion (the viscosity increase is a consequence of a disorder-
order phase transition that traps the particles in a condensed phase, reducing their motion as
discussed in the article presented in Annex D), affecting the radial concentration profile, and
resulting in higher retention times. The results presented in the latest concern polystyrene

20
polymers. These studies present opposite effects in the retention time caused by overloading
effects. A possible explanation is that different physical phenomena occurred such as distinct
surface interactions or polymers chains entanglement. Therefore, to learn the upper sample
load limit in an FFF prototype, it is necessary to perform experimental tests.

6. Membrane properties: There are many options for organic ultrafiltration membranes used in
Flow-FFF [16]. The selection requirements depend on the sample and solvent properties. The
membrane must retain the sample components without significant interactions and resist the
medium conditions (pH, temperature, solvent composition). The most commonly used mem-
branes are the regenerated cellulose (RC) and the polyethersulphone (PES). Ceramic fibers
were used for the HF5 configuration [39]. The limitations of organic membranes are the
sample absorption (more easily controlled by adjusting the solvent conditions than by chan-
ging the membrane composition [37]), the fouling, and membrane aging.

In sum, the operating conditions of Flow-FFF experiments, as all the FFF sub-techniques, are
several and system-dependent. Figure 2.14 illustrates the main operating conditions to optimize in
order to perform a separation using the Flow-FFF techniques. In [37] and [33] are given a range
of operating conditions of AF4 and HF5 analytical configurations, respectively. In [37] is given a
tutorial to start AF4 experiments. The conditions presented in the literature can be used as a starting
point to begin the experiments. Nevertheless, they must be optimized according to the response of
the system and the desired output.

Figure 2.14: Diagram illustrating the operating conditions of the Flow-FFF techniques.

21
2.3 Towards up-scaling and continuous operation
The adaptability of the analytical Flow-FFF process to an industrial scale presents several challenges.
The first one consists of the stop-flow condition to reach an equilibrium that brings instabilities and
induces system complexity. Still associated with the accumulation step is the limitation in sample
throughput. The injection of high sample amounts causes system overloading, decreasing the sep-
aration efficiency. The second one is associated with the membrane. Ultrafiltration membranes are
prone to fouling and aging phenomena requiring cleaning steps and membrane replacing. For that,
the system configuration needs to be simple to allow a easier membrane replacement and the mem-
branes lifetime has to be maximized to reduce process stoppages and membrane wastes.

The following sections present the studies performed to optimize the FFF batch systems and
to obtain continuous operations. Other techniques than Flow-FFF are presented since they share
the operation principle, and an approach that works for one technique may be a valid solution for
another.

2.3.1 Preventing the stop-flow step


The accumulation step occurring after sample injection (Figure 2.12) causes limitations in the batch
process because it can lead to sample adsorption at the membrane, sample aggregation, and hydro-
dynamic instabilities with the flow inversion [34].

Inspired by the work realized for the hydrodynamic relaxation in the F1FFF channel [34], Moon
et al. [17] developed the concept of stopless flow injection in an AF4 configuration. As illustrated
in Figure 2.15, the channel is composed of a frit inlet from which a secondary flow enters. The flow
has a higher velocity than the injected sample flow and rapidly drives the sample to the membrane
surface, occurring hydrodynamic relaxation (the term relaxation is used among FFF community to
refer to the layers formation during the accumulation step). The region where the accumulated
layers are formed is defined as the relaxation segment. The particles continue to be transported in
the axial direction and are fractionated in the separation segment. The advantages of the stopless
mainstream flow during relaxation are the reduction of sample-membrane interactions [34] and
the possibility of a continuous operation. The disadvantages are related to operation complexity in
terms of connections and an extra pump for the frit flow, the increase of sample dilution caused by
the additional flow, and band broadening [17; 40].

22
Figure 2.15: Illustration of an AF4 channel with a frit inlet. Reproduced from [17].

Wyatt [41] developed the concept of compartmentalized AF4 and HF5 channels. As shown in
Figure 2.16, the channels are divided into several compartments with the possibility of individual
flow rate control. With this type of configuration, "hydrodynamic relaxation" can occur without a
secondary flow source. The permeation flow in the first compartment can be controlled to reach
the concentration equilibrium, creating a relaxation zone segregated from the separation zone. The
system has a granted patent in the European Union (EP 2390660 B1) and United States (US 8163182
B2) with an expiration date of 23/12/2030 and 28/05/2030, respectively. In Japon, the patent had
a fee-related expiration (JP 5960948 B2), and in China is under publication (CN1 06501379 A) with
the status "IP right cessation". The claims protect all compartmentalized configurations that allow
the differential control of the flow, including membranes with different permeabilities.

Figure 2.16: Compartmentalized AF4 and HF5 channels. Reproduced from [41].

23
2.3.2 Increase sample throughput
Adapting Flow-FFF to industrial production lines, requires a higher throughput capacity, meaning
that the system must be capable of separating higher sample quantities. The strategies for process
scalability are operation parallelization and volume increase.

The first parallelization of AF4 emerged in 2005 [20], with the development of a circular system
composed of 12 channels, as illustrated in Figure 2.17. The channels have an individual inlet and
injection port and a shared outlet. The system separated 2.5 mg of polystyrene latex nanoparticles
with 50, 100, and 200 nm.

Figure 2.17: Representation of the circular configuration made of 12 AF4 trapezoidal channels [20].

Lee et al. [18] tested in 2009 the separation of 45 µg of a proteins mixture (BSA, Apoferritin, and
Thyroglobulin) using six HF5 parallel modules with a total surface area of 47.12 cm2 (0.95 µg/cm2 ).
The resolution between the Apoferritin and Thyroglobulin was worst when compared to a single HF5
module. The system required further improvement for the flow connection to avoid dead-volume
and non-uniform flow distribution in the fibers. The overloading effects were tested for BSA by
increasing the injection quantity from 6 − 200 µg. Above 50 µg, the BSA retention time started to
decrease.

In 2017, Bria et al. [35] used a semi-preparative commercial AF4 channel (from Postnova) with
a lab-made spacer having a channel surface area of 173.6 cm2 . This system allowed the separation
of 0.4 − 20 mg of silica nanoparticles with 50 and 100 nm (115 µg/cm2.). The resolution between
the peaks decreased with increasing sample amount being acceptable depending on the applications.

24
2.3.3 Continuous Operation
It exists several studies contributing to the progress of adapting FFF to a continuous operation. They
started to emerge one decade after its development and remain a challenge and a research topic in
the present days.

In 1979, Myers et al. [42] developed a continuous steric-FFF device for the continuous frac-
tionation of microparticles under the gravity effect. The channel device, as shown in Figure 2.18,
is inclined and has multiple outlets at the bottom. The particles are injected and transported along
the channel by a laminar flow. The larger particles, due to the higher effect of gravity, can reach the
regions of higher velocity and are transported faster to the bottom of the channel than the smaller
particles. Thus, different fractions are collected in separated outlets. The system was tested during
4 hours for the continuous separation of silica microparticles within a size range of 4 − 40 µm and a
feed concentration of 13 g/L. Microscope analysis of the collected fractions showed the viability of
the system.

Figure 2.18: Continuous steric FFF channel configuration. From [42]

In 1985 the split-FFF appeared for the continuous separation of particles within a size range of
300 nm - 15 µm [43]. The channel features an inlet and outlet split into two streams, as illustrated
in Figure 2.19. The solvent stream is introduced from the lower inlet at high flow rates, while the
sample is continuously injected from the upper inlet at lower velocities. The difference between the
solvent and sample flows creates imaginary planes defined as the inlet and outlet splitting planes
[44]. When the particles are injected into the channel they are confined to a narrow region of the
inlet splitting plane. The application of a force (gravitational, centrifugal, electrical, or magnetic [44;
45; 46; 47]) allow the more sensible particles to cross the inlet splitting plane, leaving the channel
through the lower outlet. In Annex C are presented some of the split-FFF channel configurations.
In [48], was tested a continuous split-FFF based on the diffusion mechanism. The splitting planes
originated by the different sample and solvent flow rates produce a thin liquid film that acts as a
"virtual dialysis membrane" separating species with different molecular weights. The more diffusive

25
components (smaller molecular weight) cross the film by diffusion and leave the channel from the
lower outlet. The species migration across the film can be controlled by adjusting the flowrates,
and consequently, adapt the film thickness. The constant film renew avoids fouling and undesirable
interaction phenomena that occur in systems using solid membranes. The apparatus was tested
with three proteins (γ-globulin, BSA, and cytochrome c) and sodium benzoate, and the results show
inefficient separation. The fractions collected are richer in different species, but more separation
stages would be necessary for further purification, which increases sample dilution.

Microfluidic devices have been designed for the application of magnetic, dielectrophoretic, grav-
itational, and acoustic forces using a split-channel configuration [13].

Figure 2.19: Configuration of the Split-FFF device. Adapted from [13].

A continuous 2D-Thermal FFF was developed in 2003 by Vastamaki et al. [49]. The device is
shown on Figure 2.20 and consists of a circular channel with a hot upper wall, a lower cold wall,
one inlet, one injection port, and 12 outlets at the edges of the circumference. The separation occurs
due to different sample displacements in two directions (2D). The first displacement is caused by
the temperature gradient moving the sample components with higher thermal diffusivity and lower
Brownian diffusion to regions close to the cold wall. The second displacement is caused by the
rotation of the cold wall. The rotatory movement affects the flow streamlines and has a higher
impact on the samples closer to the wall. The result of radial and tangential differential migration
allows the collection of the sample components in separated outlets. The separation of Polystyrene
polymers with 51, 520, and 1000 kDa molecular weight were tested. The analysis of the collected
fractions demonstrated the presence of fractions that are rich in each polymer. The use of higher
temperature gradients increased the separation efficiency [49; 21]. In [50], the authors present a
preliminary model to explore the 2D-ThFFF operating conditions and scalability limitations.

26
Figure 2.20: Scheme of the continuous 2D-ThFFF device. Reproduced from [21].

More recently, in 2019, M. Marioli et al. [22] developed a continuous 2D-AF4 system for mac-
romolecules and nanoparticles fractionation. The device was designed from the commercial AF4
channel with the frit inlet. The commercial spacer was replaced by two lab-made spacers allowing
the sample injection closer to the membrane and having six lateral outlets with non-uniform flow
rate distribution. The membrane was replaced by a microstructured membrane with a pattern in-
clination of 45◦ . The combination of the different device elements allows the stopless hydrodynamic
relaxation (frit inlet), a separation in two directions (along with and across the membrane groves),
and the collection of sample fractions in separated outlets. The separation of a binary mixture of
proteins was tested for a continuous operation and presented limited separation resolution. The
collected fractions had different compositions, being one fraction rich in the smaller protein and the
other fraction concentrated in the larger protein. Over time, there was a decay in the system per-
formance caused by membrane fouling. The author suggests improvement of the membrane pattern
and cleaning cycles for fouling reduction. Figure 2.21 illustrates the 2D-AF4 channel configuration.

Figure 2.21: Schematic representation of the 2D-AF4 channel. Reproduced from [22].

27
The concept of surface barriers applied for the microstructured membranes was introduced by
Giddings et al. [51], developing a ThFFF channel with a grooved lower cold wall. The arrangement
of the barriers perpendicular to the flow direction increased the sample retention time.

2.4 Conclusion and thesis stance


Several research groups have developed solutions to optimize the FFF techniques for industrial ap-
plications. For the size-based fractionation of colloids, the works that made the most progress are
the semi-preparative AF4 channel and the continuous 2D-AF4, from Bria et al. [35] and M. Marioli
et al. [22], respectively. The semi-preparative AF4 channel allowed a sample increase of 200 times
compared to the classic analytical AF4. The 2D-AF4 combines the different technological advances,
i.e., hydrodynamic relaxation, 2D-separation, multiple outlets, surface barriers, and continuous op-
eration, for an AF4 channel. Both systems present a limited separation efficiency, requiring optimiz-
ation and further development. Despite all the progress made towards a continuous operation or an
industrial batch process, the size-based separation of submicrometer species at the industrial level
continues to be a research topic. Table 2.2 summarizes the FFF developments for semi-preparative
purposes and continuous operations.

This project aims to progress towards a more industry adaptable solution using the HF5 concept.
The HF5 choice was motivated by several factors such as a valid industrial scalability concept, simple
membrane replacement, material availability, laboratory know-how in membrane processes, and ori-
ginality. While the AF4 configuration is widely investigated and presents limitations in terms of
scalability and membrane replacements, the HF5 configuration has more space for new develop-
ments, can be scaled using hollow-fiber modules, and membrane replacement is simple. Hollow
fiber modules are applied in industry for biological and chemical processes such as reverse osmosis,
containing several parallel fibers (tens of thousands) [52].

The project has two approaches, a sequential strategy for batch optimization and a continuous
design. The sequential prototype aims to achieve an efficient separation performance neglecting the
confined accumulation step. For that, an accumulation stage occurs along the entire fiber length.
A viable solution to obtain an accumulation region would be the existence of a compartmentalized
region or the use of a membrane with different permeabilities. However, these concepts have a
granted patent, meaning that they can be explored for research purposes but not used in the industry.
To obtain an original solution, an accumulation is performed by closing the channel outlet, resulting
in a sample spreading across the entire channel. The continuous prototype strategy consists on the
multiple outlets concept, as implemented in the steric-FFF, split-FFF, 2D-thermal FFF, and the 2D-AF4
channels.

Numerical simulations were performed to investigate the impact of the operating conditions
(namely the permeate flow during accumulation and elution and the channel length) on the separ-
ation performance, which are discussed in Chapter 3. The sequential prototype was developed on
Chapter 4 and the operating conditions of the accumulation and elution steps were investigated in
Chapter 5. The exploration of the operating conditions permitted the prescription of a methodology

28
for the size-based fractionation of colloidal particles, which was tested in Chapter 6 for the separ-
ation of model polystyrene latex nanoparticles with a size ratio of 5. In Chapter 7 is presented the
piece developed for adapting the prototype to a continuous operation, showing the geometry, the 3D
CFD model for future improvements, and a comparison of the tested 3D printing techniques.

29
Table 2.2: Summary of the FFF progress for increasing sample quantity in batch process and to obtain a continuous operation. Nomenclature: Q o - outlet
flowrate; Q i - inlet flowrate; Q y - permeate flowrate; L - channel length; w - channel thickness ; bo - maximum channel breadth; b L - minimum channel
breadth; Am - membrane surface area; and H - height between steric-FFF channel top and bottom.

Purpose Technique Configuration Details Operating Conditions Limitations Progress


Field: Permeation Flow
Channel Details: Cyclic device composed
by 12 AF4 trapezoidal channels Q o = 12 mL/min
Uniformity of flows Higher sample amount
in a quasi-parallel arrangement. Linear decay of Q y = 3 − 0 mL/min
Circular AF4 [20] and membrane properties in ∼25x analytical AF4
Total diameter of 20 cm 2.5 mg of PS Latex particles
the different channels capacity
and w of 500 µm. (50, 102, and 199 nm)
Regenerated cellulose membrane
with 10 kDa cut-off.
Field: Permeation Flow Q o = 0.72 mL/min
Semi-preparative
Channel Details: 6 HF5 modules in parallel Q y = 2.23 mL/min Uniformity of flows
Multiplexed HF5 [18] I.D.=1.0 mm, O.D.=1.4 mm, and L=25 cm ∼ 50 µg of proteins mixture and membrane properties in Higher sample amount
Polyacrylonitrile membrane (BSA-66 kDa, Apoferritin-444 kDa, the different channels
with 30 kDa cut-off. and thyroglobulin-670 kDa)
Field: Permeation Flow
Channel Details: Semi-preparative commercial
Q y /Q o = 2/4 Higher sample amount
channel with lab-made spacers
SP-AF4 [35] 20 mg of Silica Separation efficiency ∼200x analytical AF4
30

w=350 µm, b L = 2.5 cm, b0 = 10 cm, and L=23.6 cm


(50 and 100 nm) capacity
Regenerated cellulose membrane
with 30 kDa cut-off
Field: Gravity Q i = 0.4 mL/min
Channel Details: Inclined channel (23◦ ) Injected sample 0.04 mL/min Concept of multiple
Steric-FFF [42] Not applicable to colloids
with multiple outlets in the channel bottom 13 mg/mL of Silica particles outlets
L=136cm, H=635 cm, and w=0.127 mm (4-40 µm)
Field: Gravity, sedimentation, acoustic,
Q o1 = 17 mL/min Applicable to microparticles
and magnetic
Q o2 = 2 mL/min or particles with differential Concept of splitting
Splitt-FFF [43; 44] Channel Details: Channel with 2 splitted inlets
4-10 mg/mL of Palladium particles electrophoretic mobility inlet and outlets
and outlets
(1-53 µm) or magnetic properties
w=50-500 µm
Field: Temperature gradient
Continuous Operation Channel Details: Circular channel composed
Injected sample 0.417-2.5 µL/min
by a hot upper wall and a cold lower wall. Applicable to polymers
3.5-5.1 mg/mL of polymers
2D-ThFFF [49; 21] Lower wall with rotatory movement. or species with differential Concept of 2D separation
(19.8-1000 kDa)
One injection port, 1 inlet, 12 outlets. thermophoretic mobility
∆T = 37◦
Total diameter of 13.5 cm
and w of 125-250 µm
Field: Permeation Flow Q i = 2.8 mL/min
Combination of
Channel Details: AF4 commercial channel Q y = 2 mL/min
different technological
with a frit inlet, two lab-made Injected sample 0.05 mL/min Separation efficiency and
2D-AF4 [22] developments to
trapezoidal spacers, microstructured membrane, 0.25 mg/mL of proteins mixture membrane fouling
produce a continuous
and 6 lateral outlets. (Apoferritin-443 kDa,
AF4 system
L=13.3 cm, b0 =2.1 cm, w=315 µm, S=22.4 cm2 and thyroglobulin-670 kDa)
From colloids transport to FFF simulations

This chapter is dedicated to evaluating the physical phenomena of sol colloidal dispersion (solid
particles dispersed in a liquid medium) and its consequences in membranes processes. The first
3
section contains the predominant forces acting on colloids and their consequence in terms of inter-
particle interactions, phase transitions, and transport dynamics. The second section is dedicated to
the modeling of colloids transport with a continuum mechanical approach. The first implemen-
ted model was designed to study the operating conditions of the Flow Field-Flow Fractionation
(Flow-FFF) mechanism and to guide in the development of the experimental prototype (presen-
ted in Chapter 4). A second model was developed to describe the effect of inter-particle interactions
in disorder-order phase transitions and their impact on membrane processes.

3.1 Physics of colloidal dispersion

3.1.1 Definition of colloids


Colloids1 are species of different natures that present a size distribution in the range of 1 nm to 1 µm.
Their dispersion2 in suspending mediums of different states (gas, liquid, and solid) are behind their
distinct classifications as shows Table 3.1 [53; 54]. They are also designated as soft-matter because
small changes in the suspending medium (pH, temperature, composition, and ionic strength) or in
the properties of the particles (surface coating, size, and shape) can have strong consequences in
the processes (particles aggregation, adsorption to the equipment, low efficiency, etc.), as will be
discussed throughout this chapter.

1
IUPAC definition of colloids or colloidal: "The term refers to a state of subdivision, implying that the molecules or
polymolecular particles dispersed in a medium have at least in one direction a dimension roughly between 1 nm and 1
µm, or that in a system discontinuities are found at distances of that order."
2
IUPAC definition of colloidal dispersion: "A system in which particles of a colloidal size of any nature (solid, liquid or
gas) are dispersed in a continuous phase of a different composition (or state). The name dispersed phase for the particles
should be used only if they have essentially the properties of a bulk phase of the same composition."

31
Table 3.1: Colloids classification according to the phase of the suspending medium and dispersed
phase.

Suspending Medium Dispersed Phase Designation


Gas Liquid Aerosol
Gas Solid Aerosol
Liquid Gas Foam
Liquid Liquid Emulsion
Liquid Solid Sol
Solid Gas Foam
Solid Liquid Gel
Solid Solid Alloy

In the scope of this work, only the phenomena that govern spherical solid nanoparticles dis-
persed in a liquid medium (sol) are addressed.

3.1.2 Forces acting on colloidal dispersion


Colloidal particles dispersed in a fluid phase are subjected to different forces acting on their cen-
ter of mass (body forces) or related to their surface properties (surface forces), which results from
particle-particle and fluid-particle interactions (notice that in the presence of a surface, e.g. in the
accumulation of particles to a membrane there will be particle-membrane interactions) or from the
application of an external field (permeate flow, magnetic, electric, gravity, etc.). The forces that are
originated from particle-particle interactions are electrostatic repulsion and van der Waals attractive
interactions, discussed in Section 3.1.3. The forces resulting from fluid-particle interactions are the
Brownian thermal force and the hydrodynamic viscous and inercial forces. At the surface of the
earth and without the application of any additional external field, the particles are subjected to the
gravity force. Table 3.2 summarizes the forces applied to colloidal systems, including their order of
magnitude [55].
Table 3.2: Summary of the forces applied to a colloidal dispersion.

Origin Designation Type Magnitude


Electrostatic repulsions Surface force O(ε r ε0 ζ2 )
Particle-particle
van der Waals attractive interactions Surface force O(A/a)
Brownian force Body force O(kB T /a)
Fluid-particle Visocus force Body force O(µaU p )
Hydrodynamic force
Inertial force Body force O(a2 ρU p2 )
Gravity Body force O(a3 ∆ρ g)
External field
Others: Permeate flow, electric, magnetic, etc. Body force Depends on the field nature
Nomenclature: ε r - medium dielectric constant; ε0 - vacuum permittivity; ζ - zeta potential; A - Hamaker constant; a - particle radius
a= d p /2; kB - Boltzmann constant; T - temperature; µ - dynamic fluid viscosity; U p - particle velocity; ρ - fluid density; ∆ρ - difference
between particles and fluid density ∆ρ = ρ p − ρ; g - acceleration of the gravity.

The forces acting on a single particle (no particle/particle interactions) accumulated on a mem-
brane, with a permeate flow and the gravity force as the only applied external fields, are the hydro-

32
dynamic (F h ), Brownian (F B ), and gravitational (F g ) forces.

3.1.2.1 Hydrodynamic force, F h :

For a particle dispersed in a liquid medium, the hydrodynamic force (F h ) is divided into viscous and
inertial forces. The order of magnitude of the inertial forces when compared to the viscous forces
ρU p a
are given by the particles Reynolds number (Re p = µ ≪ 1). For colloidal dispersion, the viscous
effects are dominant. They lead to the Stokes drag, given by Equation 3.1 [54].

F h = 6πµa|U f − U p | (3.1)

Where µ is the fluid dynamic viscosity, a is the particle radius (a = d p /2), U f is the fluid velocity,
and U p is the particle velocity.

For a particle accumulated at a membrane, its velocity becomes negligible (U p ∼ 0), and the
hydrodynamic force is given by Equation 3.2.

F h = 6πµa|U y | (3.2)

In this case, U y represents the fluid permeation velocity in the y-direction (wall-normal direc-
tion) used during accumulation.

3.1.2.2 Brownian force, F B :

The Brownian force (F B ) is caused by the random thermal collisions of the fluid molecules with
the colloidal particles, supplying kinetic energy and resulting in a diffusive motion of the species.
The F B force is defined according to the Equations 3.3. The term kB T (kB = 1.38 × 10−23 J/K is
the Boltzmann constant and T is the absolute temperature) corresponds to the thermal motion and
represents the energy scale for colloidal interactions [54].

kB T
FB = (3.3)
a

3.1.2.3 Gravity force, F g :

The gravitational force (F g ) that acts on a particle is defined according to the Archimedes principle,
Equation 3.4.

4
F g = ∆ρVp g = (ρ p − ρ) πa3 g (3.4)
3

Where Vp is the volume of the particle, ρ p is the particle density, ρ the fluid density, and g is
the gravitational acceleration.

33
F B and F g are the only body forces permanently acting on a colloidal particle in the absence of
other external fields (Note: the gravity force is also external) [54].

3.1.2.4 Comparison of the forces (F h , F B , and F g ) magnitude

The magnitude of the viscous (F h ), Brownian (F B ), and gravity (F g ) forces was calculated for poly-
styrene latex (PS latex), silica oxide (SiO2 ) and cerium oxide (CeO2 ) nanoparticles dispersed into
water 3 . They are presented in Figure 3.1 for the radius size range of 1 nm to 1.5 µm. The cal-
culation was performed considering room temperature (T = 289.15 K) and a particle velocity of
U y = 1 × 10−6 [m/s], which is in the order of the permeate velocity used during the accumulation
stage of the experiments presented in Chapter 5 and 6.

Figure 3.1: Calculation of the viscous (F h ), Brownian (F B ), and gravity (F g ) forces for polystyrene
(PS latex), silica oxide (SiO2 ) and cerium oxide (CeO2 ) nanoparticles within a radius range of 1 nm
to 1.5 µm. The PS latex particles density ρ p = 1050 [k g/m3 ], the CeO2 particle density ρ p = 7200
[k g/m3 ], the SiO2 particle density ρ p = 2000 [kg/m3 ], the fluid density ρ= 1000 [k g/m3 ], the fluid
dynamic viscosity µ= 1×10−3 [Pa.s], the particles velocity U p = 1×10−6 [m/s], and the temperature
T = 298.15 [K]

For the PS latex nanoparticles, the Brownian forces are predominant when the radius of the
particles are below ∼ 450 nm. For larger sizes, the viscous forces (F h ) are the governing mechanism
(note that it depends on the velocity U y ). When the particle radius is superior to ∼ 1.2 µm, the gravity
force overcomes the Brownian diffusion. Performing the same calculation for spherical SiO2 and
CeO2 nanoparticles, the only force that varies is the F g since the particles density is ρ p = 2.0 g/cm3
and ρ p = 7.2 g/cm3 , respectively. In this case, the F g > F B at lower particles radius corresponding
to ∼ 560 nm and ∼ 350 nm, respectively. This illustrates that the physical phenomena impacting the
sized-base separation with the Flow-FFF mechanism depend not only on the particle size (d p ) but also
on its density (ρ p ). The normal elution mode may be inverted at the point where the gravity force
surpasses the Brownian diffusion. In this case, the sedimentation and centrifugal FFF4 , techniques
3
The choice of this type of particles is because we explored the prototype using model polystyrene latex nanoparticles,
and the future application is the separation of silica and cerium oxide nanoparticles produced by Solvay.
4
The steric-FFF and split-FFF, discussed in Chapter 2, permit a continuous separation of particles where the gravity

34
might be more appropriated for the separation.

3.1.3 Particle-particle Interactions


The force balance for single colloidal particles showed that the Brownian force is the predominant
mechanism for a wide size range (depending on the density of the particles). In colloidal dispersion
with several particles, the Brownian force affects the dynamics of the particles, leading to diffusion
(random motion) and promoting particle collisions. When the inter-particle distance decreases, the
particles interact with their neighbors according to their surface physicochemical composition, result-
ing in structural arrangements that can affect the dispersion stability and induce a phase transition.
The existence of surface interactions is behind the deviation from ideality and the reason why matter
has different physical states. The effects of the interactions depend on the number of available inter-
faces5 , the nature of the interactions and their range. The nature of the inter-particle interactions can
be divided into electrostatic repulsion and van der Waals attractive interactions (notice that this is a
simplified description since there are several forms of inter-particle interactions as the steric effects
and hydrogen bonds) [3; 55].

3.1.3.1 Electrostatic repulsive interactions

The electrostatic interactions are caused by the overlapping of the electronic double layer of equally
charged particles and, consequently, results in particles repulsion. The electronic double layer is
formed by the ions (counterions and coions) present in the suspending medium that surrounds the
particles to neutralize their surface charge (there is the neutralization of the surface charge but the
charge within the double layer is not neutral - the number of counterions is larger compared to the
coions). The model of the electrostatic double-layer considers the existence of a stern plane, where
the ions are immobilized at the particle surface and follow the particle movement, and a shear surface
at which the ions can move independently of the particle. From an experimental point of view, the
particle surface charge can not be measured due to the layers of ions, being measured the potential at
the shear surface. This potential is designated as zeta potential (ζ) and is experimentally determined
by DLS (according to the procedure presented in Section 4.2.1.1).

Electrostatic repulsion promotes colloids stability by avoiding aggregation between particles or


adsorption to surfaces with the same charges. The zeta potential (ζ) can be determined as function
of the medium pH (titration curve) to determine the isoelectric point, i.e. ζ = 0 mV . At the isoelec-
tric point, the colloids net charge (particles+immobilized ions) is neutralized and aggregation can
occur. The ζ gives information about the particles surface charge (positive or negative) at a specific
pH, about the system stability (the colloidal systems are considered stable for the empirical range:
−30 > ζ > +30 mV ), and the medium conditions to have stability (pH ranges to avoid to keep the
electrostatic repulsion).

The electrostatic potential (ψ) is the capacity to attract or repulse and is calculated recurring

force predominates the Brownian diffusion.


5
The small colloidal sizes result in a large quantity of species in solution and high surface areas available for inter-
particle interactions. In 10 mL of sample with a concentration of 1 g/L of spherical polystyrene latex with 100 nm
diameter, there is 1.8 × 1013 particles (Np ) and a total interface area (S pt ) of 0.6 m2 .

35
to Poisson equation (Equation 3.5) and Boltzmann equation (Equation 3.6) that combined result in
the Poisson-Boltzmann equation presented in Equation 3.7, which does not have a general analytical
solution. For small potentials (zi eψ ≪ kB T , i.e. for ψ <∼ 25 mV ) and spherical particles, Equation
3.7 has the analytical solution presented in Equation 3.8, known as the Debye-Huckel approximation
[53].

ρE
▽2 ψ = − (3.5)
ε

 ‹
ci −zi eψ
= ex p (3.6)
c0 kB T

 ‹  ‹
1 d 2 dψ −F X −zi eψ
r = zi ci e x p (3.7)
r2 d r dr ε0 kB T

ζae x p[−k D (r − a)]


ψ= (3.8)
r

In the equations, a is the particles radius, ρ E is the charge density, ε is the permittivity of the
medium, ci is the double layer ions concentration, c0 is the ions concentration in the bulk, zi is the
charge number, e is the electron charge (e = 1.602 × 10−19 A.s), r is the radial spherical coordinate,
F is the Faraday constant (F = 96485.333 A.s.mol −1 ), kB is the Boltzmann constant 1.380 × 10−23
J/K, T is the temperature, and k D is the Debye-Huckel parameter.

The thickness of the double layer responsible for the neutralization of the charged surface is
designated by screening or Debye length λ D (the inverse is the Debye-Huckel parameter k D ) and is
calculated according to Equation 3.9. The equation shows the influence of the ionic force (Is ) on
λ D . High ionic strength corresponds to a high concentration of ions in solution, thus a lower layer
thickness is necessary to neutralize the surface [53; 54]. Therefore, particles dispersed in a solvent
with high Is have a smaller screening length, which reduces the electrostatic repulsion and promotes
aggregation.

v
1 u ε RT
r 3.07 × 10−10
= λD = P 2 = (3.9)
t
p
kD 2F 2 zi ci I

1X 2
Is = zi ci (3.10)
2

Where ε r is the medium dielectric constant and R= 8.314 J/(mol.K) is the ideal gas constant.

Figure 3.2 shows the normalized electrostatic potential (scaled by zeta potential) given by the
Debye-Huckel approximation as a function of the distance from the particles surface (ds ) for a spher-

36
ical particle with 100 nm diameter dispersed in an electrolyte with the same number of cations and
anions (co = Is ). The electrostatic potential was calculated for ionic strengths6 varying from 1×10−5
M to 1 M . The profiles show that the electrostatic potential range decreases with the increase of Is ,
as expected since there are more available ions to neutralize the diffusive boundary layer and reduce
the Debye screening length (λ D ). At low Is , the Debye length is longer. This has a direct impact on
the inter-particle distance of two approaching particles: for low Is and high λ D , the overlapping of
the double layers occurs at higher inter-particle distances, preventing particles from aggregation.

Figure 3.2: Normalized electrostatic potential (ψ scaled by ζ) of a spherical particle, calculated with
the Debye-Huckel approximation, as function of the distance from the particle surface ds (ds = r −a).

The electrostatic potential energy (W ele ) between two colloidal particles that have the same size
and surface potential is described according to Equation 3.11.

d
W ele = 64πani kB T λ2D γ2 e x p(− ) (3.11)
λd

Where ni is the number density of ions in the bulk, d is the inter-particles distance, and γ2 =
t anh (zi ζe/4kB T ) is the dimensionless function of the wall potential.

3.1.3.2 van der Waals attractive interactions

The van der Waals attractive interactions between two spherical colloidal particles are caused by
instantaneous fluctuations of the electrons cloud of the atoms of one colloid that induce polarization
in the atoms of the other colloid (London interaction between instantaneous dipole-induced dipole).
The potential energy of attraction interaction (W vdw ) between two particles (negative by convention)
with the same radius a at distance d is calculated according to Equation 3.12 [53; 54].

6
To give reference values, the ionic strength of ultra-pure water is O(10−5 M ) and for seawater is O(10−3 M )

37
  
vdw A 2a2 2a2 2a2
W =− + + ln (3.12)
6 d 2 + 4ad d 2 + 4ad + 4a2 d 2 + 4ad + 4a2

Where A is the Hamaker constant that depends on the properties of the interacting particles and
the suspending medium [56].

3.1.3.3 DLVO theory for the description of interactions in diluted conditions

The Derjaguin-Landau-Verwey-Overbeek (DLVO) theory describes the interaction potential energy


necessary for the approaching of two colloidal particles7 , through the linear addition of the terms
for the van der Walls attraction potential and the electrostatic repulsion potential, as shows Equa-
tion 3.13.

W DLV O = W vdw + W ele (3.13)

Figure 3.3 represents the electrostatic, van der Waals, and DLVO potential energy of interaction
(scaled by the thermal energy kB T ) as a function of the distance between two approaching spherical
particles with d p = 100 nm, ζ=-40 mV , and I = 1 × 10−5 M . At long distances, the particles feel
van der Waals interactions and start to attract each other. In some cases, not visualized in this figure
because it depends on the particles charge and ionic strength conditions, a secondary minimum is
reached, leading to particles flocculation (reversible by shear or sonication). The electrostatic barrier
prevents further approach, however, if the thermal energy that causes Brownian motion provides
enough energy to overcome it, a primary minimum is achieved, resulting in particles aggregation
(irreversible). The higher is the electrostatic barrier, the higher is the thermal energy required for
the particles to aggregate, therefore the system has higher stability.

Figure 3.3: Interaction energy potential (W DLV O ) obtained by the addition of the electrostatic (W ele )
and van der Waals (W vdw ) contributions for two approaching spherical particles with d p = 100 nm,
ζ= −40 mV , and Is = 1 × 10−5 M . Calculation parameters: ε= 78.3 (for water at 298.15 K - pg. 43
of [53]), kB = 1.381 × 10−23 J/K, T = 298.15 K, zi = 1, e= 1.602 × 10−19 , and A= 1.3 × 10−20 J.

7
Notice that the DLVO theory, as the electrostatic and van der Waals interactions, is applied to any type of molecules,
being described in this manuscript for the prediction of colloidal stability

38
Figure 3.4 shows the interaction energy potential W DLV O (scaled by the thermal energy kB T )
for different ionic strengths (Is ). We observe that the lower Is results in a higher magnitude of the
electrostatic barrier, thus more thermal energy is required to achieve the primary minimum that
leads to irreversible particle aggregation. The increase of the ionic strength reduces this barrier and
facilitates aggregation.

Figure 3.4: Interaction energy potential (W DLV O ) for two approaching spherical particles with d p =
100 nm, ζ= −40 mV , at different Is . Calculation parameters: ε r = 78.3, kB = 1.38064852 × 10−23
J/K, T = 298.15 K, zi = 1, e= 1.602176634 × 10−19 , and A= 1.3 × 10−20 J.

The DLVO theory is valid only for diluted conditions (approaching of two bodies) and can not
be applied to evaluate the stability of concentrated systems. Although, it gives essential information
about the role of ionic strength in system stabilization. For concentrated suspensions, particle surface
interactions can be evaluated using the experimental osmotic pressure curves, as discussed in the next
section.

3.1.3.4 Osmotic pressure for the description of interactions in concentrated conditions

The osmotic phenomenon was discovered in 1748 by Abbé Nollet and consists of the movement
of water molecules from regions with lower to higher solute concentrations until thermodynamic
equilibrium is reached (equality of chemical potentials). In its turn, the osmotic pressure (Π) is the
pressure required to prevent the water molecules transfer [57]: the term is often used to refer to
aqueous solutions of molecules and salts. Similarly, it is common to describe the colloid dynamics
through the osmotic pressure phenomena8 [58]. As discussed in Section 3.1.3.1, the colloids dis-
persed in aqueous solutions are surrounded by the available ions, resulting in the formation of an
electrostatic double layer. The ions located in the diffusive part of the double layer (i.e. ions that are
not immobilized at the particle surface) can move to the bulk (at the edge of the electronic double
layer, the ions thermal motion gives the energy necessary for the ions to escape from the electrostatic
double-layer [53]). This results in a variation of the chemical potential that induces the osmotic flux
8
For reference values, the osmotic pressure of the seawater is 25×106 Pa and for 10 mL of a sample with a concentration
of 1 g/L of spherical polystyrene latex with 100 nm diameter is 7.4 × 10−3 Pa.

39
to reestablish the thermodynamic equilibrium.

In 1901, Van’t Hoff9 established an analogy between the colloids osmotic pressure (Π) in di-
luted conditions, named Van’t Hoff law, and the ideal gas law (pV = nRT , being p the pressure, V
the volume, n the number of moles, R the ideal gas constant, and T the temperature). A physical
description (expressed in terms of φ, the particle volume fraction in a colloidal suspension) of the
Van’t Hoff law is given in Equation 3.14 and can be deduced from the definition of chemical potential
and water activity in diluted conditions.

φ
Π= kB T (3.14)
Vp

The osmotic pressure (Π) is a colligative property (depends on the quantity of species in dis-
persion φ) and can be used to describe the colloids phenomena, representing the equation of state
(EOS). For concentrated dispersion, the presence of inter-particle interactions causes deviations from
ideality, and the van’t Hoff law is no longer valid. Therefore, it is necessary to include terms, as the
Virial expansion (also found in models for non-ideal gases) or develop mathematical models to cor-
rect the non-ideality. In Sections 3.2.2.4 and 3.2.3 we present two mathematical models used to
describe the osmotic pressure. The first model accounts for entropic effects and inter-particle in-
teractions for φ below a critical volume fraction that induces phase transition. The second model
describes disorder-order phase transitions and was used to study the relaxation dynamics of a con-
densed phase (also referred to as a gel) and its impact on membrane processes.

The experimental determination of osmotic pressure curves10 (Π vs φ) give an essential inform-


ation about the colloids stability according to the suspending medium properties as the pH and ionic
strength (Is ) [59], as described in the next section.

3.1.4 Phase transitions


The colloidal particles can face phase transitions from a dispersed to a condensed particles arrange-
ment caused by the increase of species concentration (φ) that lead to repulsive and attractive inter-
particle interactions.

Figure 3.5 represents a typical experimental curve of the osmotic pressure (Π) as a function of
the colloids volume fraction (φ) [60; 61]. For low volume fractions, the osmotic pressure follows the
van’t Hoff law (red dashed line): the particle motion essentially follows Brownian thermal diffusion
and the colloids behave as the atoms of an ideal gas. Increasing φ leads to an increase of the
osmotic pressure and a deviation from the ideal diluted conditions, caused by electrostatic repulsion
between colloidal particles due to the overlapping of electrostatic double layers (light-blue line).
The electronic repulsion reduces the effect of the Brownian diffusion and increases the system order.
The continuous increase of φ leads to a reduction of the inter-particles distance, which results in the
London van der Waals attractive interactions. The particles attraction can lead to a phase transition

9
For his research work on osmotic pressure, Van’t Hoff was awarded the Nobel Prize in Chemistry in 1901
10
For a review of the experimental methods to measure the osmotic pressure, the reader may refer to the Ph.D. thesis
of Benjamin Espinasse [59] pages 54-81.

40
from a dispersed (disordered) to a condensed (ordered) phase characterized by the appearance of
a plateau in the osmotic pressure curve (pink and purple lines). If the plateau is characterized by
∂ Π/∂ φ = 0, the system undergoes a spinodal decomposition, and the condensed phase is organized
in concentrated and diluted phases (unstable and irreversible phase transition). The phase transition
occurs at a volume fraction defined as a critical volume fraction (φcr ), which depends on the medium
properties, such as the ionic strength. For clay nanoparticles, the predominant phenomenon is the
attractive interactions that lead to the "house-of-card" structures formation, which are responsible
for the phase transition. High ionic strength benefits the attractive interactions, and permits the
occurrence of phase transition at lower φcr [62]. On the other hand, if the predominant phenomenon
is the electrostatic repulsion that results in particles arrangement in "cage" structures, as occurs for
latex polystyrene nanoparticles, the increase of ionic strength (Is ) shifts the phase transition to higher
φcr [63]. The systems where the attractive interactions are predominant are defined as colloidal
gels, and those governed by electrostatic repulsion are designated as colloidal glasses11 . Beyond
phase transition, the increase of φ induces a compression of the condensed phase (brown line), and
the increase of the osmotic pressure is related to the higher resistance of the solid phase to further
compression. The osmotic pressure curve diverges at the maximum close packing volume fraction
(φcp ).

Figure 3.5: Regions observed during the experimental measurement of the osmotic pressure, caused
by a dispersed-concentrated (disorder-order) phase transition.

3.1.5 Transport dynamics

3.1.5.1 Diffusion

The collisions of colloidal particles with the fluid molecules create a Brownian force (each particle
receives kinetic energy kB T ) that is responsible for a random motion of the particles in the fluid. This

11
There is no consensus among the physics community concerning the nomenclature of colloidal gels and colloidal
glasses, being presented in this manuscript the terms most frequently found in the literature.

41
motion is of diffusive nature. The colloid motion is self-diffusion if the suspension is homogeneous
and collective diffusion if concentration gradients are present, which leads to mass (colloid) transfer
[3].

The diffusion coefficient of a spherical particle in dilute conditions (D0 ) is given by the Stokes-
Einstein relation presented in Equation 3.15 [53].

kB T
D0 = (3.15)
6πµa

In the presence of gradients, a macroscopic flux of the colloids occurs, and the particles diffu-
sion is described by a gradient diffusion coefficient (D(φ)) as presented in Equation 3.16. This is the
generalized form of the Stokes-Einstein equation that takes into account hydrodynamic and thermo-
dynamic interactions by including the sedimentation coefficient K(φ) and the compressibility factor
Z(φ) respectively [55].
d [φ Z(φ)]
D(φ) = D0 K(φ) (3.16)

The sedimentation coefficient K(φ) represents the ratio between the particles settling velocity at
φ and their settling velocity in diluted conditions, i.e. it represents the particles motion hindered by
their neighbors. There are several theoretical correlations for K(φ), such as the Happel, Batchelor,
and Kozeny Carman correlations. The Happel function H(φ), given in Equation 3.17, is valid in a
wide range of φ and represents a resistance for the species motion [64]. Considering the Happel
function given by Equation 3.17, the compressibility factor calculated according to Z(φ)= Π/nkB T ,
and the mobility factor (Equation 3.18), the generalized Stokes-Einstein relation assumes the form
of Equation 3.19 [55].

1 6 + 4φ 5/3
= H(φ) = (3.17)
K(φ) 6 − 9φ 1/3 − 9φ 5/3 − 6φ 2

K(φ) 1
m(φ) = = (3.18)
6πµa 6πµaH(φ)


D(φ) = m(φ)Vp (3.19)

3.1.5.2 Viscosity

The viscosity is a colligative property of fluid flows. In Newtonian fluid flows, the shear stress (τs
[N /m2 ]) is linearly proportional to the shear strain (γ̇ [1/s] - velocity gradient), the constant of
proportionality being the viscosity (µ):

τs = µγ̇ (3.20)

42
The colloidal dispersion exhibit complex rheological behavior even if the suspending medium is
a Newtonian Fluid, depending on the quantity of particles (φ), the medium temperature (T ), and
the state of dispersion [55]. The Non-Newtonian behavior of colloidal suspensions can present shear-
thickening (an increase of viscosity with the increase of shear stress), shear-thinning (decrease of
viscosity with the increase of shear stress), and thixotropy (shear-thinning with time), some colloids
being able to present both behaviors [65].

The complex rheological behavior of the colloidal suspensions brings challenges to understand
and control chemical processes but is also responsible for the existence of diverse and essential
products (paintings, additives to improve material performance, moisturizing creams, etc.).

In the scope of this work, the change in viscosity (µ) caused by particles accumulation does
not affect the system performance. An increase in viscosity is a consequence of the inter-particle
interactions that can lead to a phase transition. This phase transition is to be avoided since we try
to keep electrostatic repulsion that maintains the particles dispersed (even during the accumulation
stage). The local impact of the suspension viscosity in the system hydrodynamics is not perceived
in the prototype performance. In our CFD models, we decided to use a constant viscosity to keep
the model simple. For the readers knowledge, there are viscosity models that can be used to bet-
ter predict the viscosity of colloidal suspensions. As an example, the Einstein viscosity law adds
viscosity dependency with the volume fraction φ, according to Equation 3.21, valid for hard spher-
ical particles, dispersed in an incompressible Newtonian fluid, transported in a laminar flow and in
diluted conditions (φ≤ 0.1) [53; 66].

µ(φ) = µ(1 + 2.5φ) (3.21)

To describe concentrated suspensions, Krieger-Dougherty derived Equation 3.22 from the Ein-
stein viscosity law. This equation takes into account the existence of a maximum volume fraction
(φcp ) that is physically reached. The Einstein coefficient of 2.5, presented in Equation 3.21, is re-
placed by the intrinsic viscosity [η]12 [53; 66].

 −[η]φcp
φ
µ(φ) = µ 1 − (3.22)
φcp

The two extreme limits correspond to the low and high shear stress. At the low shear stress,
the particles are randomly packed as the volume fraction φ increases, presenting a φcp = 0.64 and
a [η] = 2.5. For high shear stress, the particles are ordered in a hexagonal close packing achieving
a φcp = 0.71 and a [η] = 3.13 [66].

3.1.6 Impact of the colloids physics in the Flow-FFF mechanism


The Field-Flow Fractionation mechanism, as discussed in Chapter 2, is based on the particles diffu-
sion. From the colloids physical properties discussed in the previous sections, we realize the need
12 µs −µ
Particles contribution to the suspension viscosity: [η] = limφ→0 µφ , where µs is the suspension viscosity.

43
for adequated conditions for an efficient Flow-FFF mechanism. The medium conditions have to pro-
mote inter-particles and surface-particles (for AF4 and HF5 the surface is a membrane) electrostatic
interaction to have a stable dispersion (reduction of adsorption and aggregation) and an efficient
mechanism (high diffusion). The magnitude of the electrostatic repulsion and attractive interactions
can be controlled through the medium pH, ionic strength (Is ), and presence of surfactant molecules,
and evaluated by measuring the zeta potential (ζ) and particles size (d p ).

3.2 Modelling of colloids transport


Numerical models are powerful tools for the investigation of physical phenomena that can be com-
plex (or even impossible) to determine experimentally and to study process operating conditions
without having to perform several experiments that are time-consuming and expensive. With the de-
velopment of the computing resources (high-performance calculation HPC), it is possible to launch
several simulations in parallel and have the results within few days, which allows the study of differ-
ent parameters simultaneously. Numerical models can be strong allies for the experiments, guiding
in the design, exploring operating conditions, and used in process optimization.

Particularly, we used numerical modeling for the development of a new prototype to achieve
the size-based separation of colloidal particles inspired by the Flow-FFF mechanism, i.e. the inter-
play between permeate flowrate convection and particles diffusion resulting in different retention
times. From the Flow-FFF literature (discussed in Chapter 2), we noted that the system has to have
accumulation and elution steps, and the permeation flow rate (externally applied field) is the most
critical parameter. Our numerical model was developed to study the impact of the accumulation and
elution permeate flowrate in the separation performance and to test the effect of channel length on
the accumulation and elution steps to guide the first developments of the experimental prototype.
Section 3.2.2 presents the model configuration, parameterization and the obtained results.

The advantage of using numerical models in chemical engineering is that they can be simple
to give a direction for the operating conditions (direct the research efforts) and can be improved
with time (more sophisticated) to allow a more accurate physical description of the phenomena.
In this way, Section 3.2.2.4 presents a transport model that accounts for inter-particle interactions
through the contribution of entropy, electrostatic repulsion, and attractive interactions to the os-
motic pressure. This model is ready for the optimization of the prototype, in particular, to study the
accumulation conditions where the inter-particle interactions are not negligible and can affect the
process.

In parallel to the development of the prototype, we implemented a mathematical model that


describes the different regions of the experimental osmotic pressure curves (as the one presented
in Figure 3.5) for the investigation of the dynamics of a gel13 phase relaxation (Section 3.2.3) and
its impact in membrane processes (Section 3.2.3.3). This model aims to progress toward a better
understanding of the relaxation mechanism for colloids that undergo an accumulation stage, which
can occur in FFF, filtration [67], drying [68; 69], and pervaporation [70] processes and are experi-
13
The term gel is used to describe the condensed phase that is formed due to particles accumulation at a surface and
does not refer to any specific structural arrangement or predominant interaction phenomena.

44
mentally inaccessible or hard to observe.

The transport equations were implemented in OpenFOAM (Open Field Operation and Manipu-
lation), an open source CFD software developed14 in 1996 by Hrvoje Jasak at the Imperial College of
London [71]. The versions Foundation 6 and 7 were used. The results were treated using ParaView
(version 5.6.0 64-bit) and python scripts (version 2.7).

3.2.1 Hydrodynamic governing equations


The fluid equations are mass (the continuity equation) and momentum (the momentum equation)
conservation.

The continuity equation obtained from the fluid balance in an infinitesimal control volume is
given in Equation 3.23:

∂ρ
+ (▽ · ρU) = 0 (3.23)
∂t

where U is the velocity field.

Equation 3.23 gives the variation of density with time at a fixed point in the space. For an
incompressible flow the density is constant and the continuity equations is simplified:

▽·U =0 (3.24)

The momentum conservation equation for a fluid with a constant density ρ is obtained doing
a balance at an infinitesimal element of volume (the total amount of forces acting on the control
volume is equal to the sum of the balance of convective, pressure, viscous, and body forces) following
Equation 3.25 [72; 73].

 ‹
∂U
ρ + U · ▽U = − ▽ p + ρ g − (▽ · τs ) (3.25)
∂t

Where p is the local pressure, τs is the viscous stress, and g is the gravity acceleration.

Considering a Newtonian fluid with a constant viscosity µ, the Equation 3.25 is simplified to the
Navier-Stokes equation [72]:

 ‹
∂U
ρ + U · ▽U = − ▽ p + ρ g + µ ▽2 U (3.26)
∂t

The Reynolds number of the fluid (Re) is given by Equation 4.19 and represents the ratio between

14
OpenFOAM has 32 official contributors (https://wiki.openfoam.com/Main_Page visited at 06/06/2021). Recently
was created the OpenFoam journal, which can be found in the link: https://journal.openfoam.com/index.php/ofj/
index

45
the inertia and viscous forces.

ρU L
Re = (3.27)
µ

Where L represents a length scale.

3.2.2 Model for the prototype design: Simulation of the Flow FFF mechanism
The first model solves the colloids as a scalar with an average concentration (C) transported by the
fluid. Therefore, the model solves the continuity and momentum equation presented in Equations
3.24 and 3.26, respectively, and the transport equation defined in Equation 3.28.

∂C
+ U · ▽C = D0 ▽2 C (3.28)
∂t

The diffusion coefficient (D0 ) is obtained from the Stokes-Einstein equation for diluted condi-
tions, not taking into account inter-particle interactions. This approach is valid only at low concen-
tration (C).

In OpenFOAM, the equations are solved in a discretized domain using the finite volume method.
The resolution of the momentum and transport equations was validated in steady-state by comparing
the OpenFoam output with analytical solutions for the hydrodynamics and mass transfer in channels
with porous walls [74; 75]. The transient state was validated by verifying the mass conservation and
comparing (qualitatively) the OpenFOAM output with Taylor dispersion theory [53]. The results are
not present in this manuscript.

3.2.2.1 Numerical simulations setup

The system domain is illustrated in Figure 3.6 and consists in a 2D rectangular geometry with a
height H and a total length L. The flow direction is along x-axis. The wall-normal direction is along
y-axis. The mesh is composed of 1500 cells in the x-direction, 250 cells in y-direction, and 1 cell in
z-direction15 . The mesh presents an expansion ratio of 0.01, which means that the ratio between the
size of a cell near the wall and the size of a cell at the channel center is 0.01. Mesh independence
tests (not presented in this manuscript) were performed to confirm that the solution does not depend
on the mesh resolution.

15
in OpenFOAM is required to build 3D meshes, and the simulation of 1D or 2D cases is defined by imposing "empty"
boundary conditions along the undefined dimensions

46
Figure 3.6: Geometrical domain of the first implemented model. H - channel height, L - channel
length, L1 - length of the accumulation segment (permeable wall), and L2 - length of the elution
segment (solid wall).

The domain is divided into two blocks (L = L1 + L2 ) to impose separated boundary conditions for
the particles flux: one region of length L1 is porous for the fluid and impermeable for the particles,
corresponding to the accumulation region. The remaining part of length L2 is a solid wall. The
membrane permeability is imposed by a boundary condition (BC) that sets a zero particles flux
across the membrane.

We studied the operating conditions: permeation velocity at the wall (U y ) during the accumu-
lation and elution steps, and dimension of the accumulation (L1 /H) and elution (L2 /H) regions.
The simulations were performed under the same conditions, being the varied parameter identified
in due course.

The simulations start with an accumulation stage, with the injection of a finite step of particles
(C = 1 (c.u.) during 0.1 (t.u.)). The injected particles are accumulated at the membrane dur-
ing 6H/U y (t.u.) (accumulation time relation obtained from the Flow-FFF accumulation duration),
with a constant permeation velocity (U y ) that corresponds to an accumulation Peclet number (Pe y )
defined as:

Uy H
Pe y = (3.29)
D0

The elution occurs immediately after the particles accumulation, having as parameters the ve-
locity components in the x-axial direction (U x ) and the y-direction (U y ). The U x was defined to be
equal to 1 (t.u.) (is a reference value, since the flow regime is imposed with the dimensionless Reyn-
olds and axial Peclet numbers defined later in this section) converges rapidly to a laminar parabolic
profile. The permeate velocity (U y ) is constant along the entire membrane and only exists in the
accumulation segment.

The boundary conditions are shown in Figure 3.7 for (a) the accumulation and (b) the elution
steps.

47
(a) Accumulation BC

(b) Elution BC

Figure 3.7: Boundary conditions applied during the (a) accumulation and (b) elution stages.

The hydrodynamics (Equations 3.24 and 3.26) is solved in steady-state decoupled from the
transport equation, with the SIMPLE algorithm, using the second-order scheme bounded Gauss
linear Upwind for the convection term, Gauss linear for the laplacian term and for the calculation
of gradients, and Euler for the time discretization. The discretization scheme used for the convection
term is the second-order scheme bounded Gauss limited Linear 1, while the remaining terms are
discretized with the same schemes as the momentum equation. The hydrodynamics and transport
calculation is decoupled to reduce simulation time. The minimum cell size (∆ y) is 9.3 × 10−5 (l.u.),
requiring a time step (∆t) smaller than 9.3 × 10−4 (t.u.) to respect the stability criteria for the
Courant number (C F L= U y ∆t/∆ y < 1). The decoupling of the equations is possible because the
viscosity (µ) and permeation velocity (U y ) are considered as constant (there is no dependency with
the concentration).

The experimental conditions are tested for three particle sizes (d p ) with 10, 50, and 100 nm.
We set the physical properties such that ρ = 1 (d.u.), U x = 1 (v.u.), D0 is imposed for the particles
to diffuse with the correct (experimental) Pe x number, and µ to have the Re number in the correct
range. Therefore, the particles size is introduced in OpenFOAM through the diffusion coefficient
(D0 ), according to the relation:

[U x H]OF
• ˜ex p • ˜OF
Ux H Ux H ex p
Pe ex p
= PeOF ⇔ = ⇔ D0OF = D0 (3.30)
x x
D0 D0 [U x H]e x p

48
Where Pe x is the Peclet number in the axial x-direction. The subscript ex p and OF correspond
to the experimental and simulation parameters, respectively. The Pe x of the Flow-FFF experiments
ex p
(Pe x ) for particles with 10, 50, and 100 nm is in the range of O(∼ 1 × 104 − 1 × 105 ). These Peclet
values result in thin layers of particles accumulated at the membrane, which require a high mesh
refinement to properly discretize the domain containing the boundary layer. We decided to use a
numerical Pe x (PeOF
x ) ten times smaller than the experimental one to prevent expensive simulations
3 3
that would be required to solve very refined meshes. The values of PeOF
x are 1 × 10 , 5 × 10 , and
1 × 104 for the particles with 10, 50, and 100 nm, respectively. This corresponds to a O(∼ 10)
ex p
times reduction of the experimental velocity (U x ), that does not have an impact in the Flow-
FFF mechanism. As discussed in Chapter 2 and 5, the axial velocity does not affect the balance
between the diffusion and the drag force, which is the phenomenon responsible for the separation
of the particles, and the mechanism aimed to simulate. Furthermore, the simulations and the Flow-
FFF experiments are both in the laminar (through the Reynolds number) and pure convection limit
(through the Pe x ) regimes. The objective of the model is not to simulate the exact conditions of the
Flow-FFF (even because there is an enormous range of operating conditions that can be applied) but
to have a tool to guide in the conception of the experimental prototype.

The Reynolds number (Re) is introduced in OpenFOAM through the viscosity (µ), according to
the relation:

[ρU x H]OF
• ˜e x p • ˜OF
ex p OF ρU x H ρU x H
Re = Re ⇔ = ⇔ µOF = (3.31)
µ µ Re ex p

being ρ OF =1. The ReOF of 0.04 was obtained considering a O(∼ 10) times reduction of the U x
ex p
comparing to the typical experimental values (U x ), to be in accordance with the PeOF
x .

The parametrization parameters obtained from the PeOF


x and Re
OF
are summarized in Table 3.3.
Table 3.3: Parameters used with the numerical model.

Simulation variables Experimental values OpenFOAM input


d p =10 nm Pe x =1000 D = 1.0 × 10−3 (l.u)2 /(t.u)
d p =50 nm Pe x =5000 D = 2.0 × 10−4 (l.u)2 /(t.u)
d p =100 nm Pe x =10000 D = 1.0 × 10−4 (l.u)2 /(t.u)
Re 0.04 µ=25.0 (m.u)/[(l.u)(t.u)]

3.2.2.2 Effect of the operating conditions

Study of the permeation velocity during the elution step: The influence of the permeate velocity
(U y ) in the peak retention for monodispersed particles populations with 10, 50, and 100 nm was
evaluated by analyzing the elution signal of particles obtained at the channel outlet and the corres-
ponding retention time (t r ). For this, we analyzed the signals of particle flux ( f (t)) as a function of
the normalized elution time (t ∗ = t e /τ, where τ is the geometric residence time defined in Equation
3.35, and t e is the simulation elution time).

49
Particle flux is calculated by doing a molar flux integration of the concentration at the channel
outlet (Cout ), weighted by the velocity field (U x ( y)) at the outlet cross-section S:

Z
f (t) = Cout ( y)U x ( y)dS (3.32)
S

The particle retention time (t r ) was calculated according to Equation 3.33:

R
t e f (t)d t
tr = (3.33)
Nin

Where Nin is the total amount of particles at inlet. Considering uniform particle distribution at
the inlet cross-section (Cin ) and uniform inlet velocity field (Uin ), Nin is obtained by multiplying Cin
and Uin by the duration of the injection step (∆t in ), and by the inlet cross section S, as shows the
Equation 3.34.

Nin = Cin × Uin × ∆t in × S (3.34)

The geometric residence time (τ) is the time required for unretained molecules (when U y = 0)
to be transported by convection, and is given according to Equation 3.35.

τ = L/Ui (3.35)

We performed three tests using the same geometry and accumulation conditions but varying
the permeation velocity (U y ) during the elution stage. The geometry dimensions were L1/H = 5
and L2/H = 5. The accumulation was performed during 40 (t.u)16 with a U y = 0.1 (v.u). Figure
3.8 shows the elution peaks obtained for the particles with 10 nm (blue line), 50 nm (orange line),
and 100 nm (green line), for the U y values of (a) 0 (v.u.), (b) 1 × 10−3 (v.u.), and (c) 1 × 10−2
(v.u.). Figure 3.8-(a) corresponds to an elution stage without sample retention (U y = 0). In Figure
3.8-(b) the permeate velocity increased, and the particles retention time is slightly higher than in
the previous case. The signals presented in Figure 3.8-(c) were obtained for stronger permeation
conditions and result in more retained peaks (t ∗ increased for all the species). The augmentation of
the retention time has consequences on the peak shape, which is broader and flatter.

16
Less than 6H/U y because at the time this simulation was performed the accumulation duration was not parameterized.
The accumulation time of 40 (t.u) is adequate to accumulate the particles at the membrane, and the fact that it is smaller
than 6H/U y impacts the maximum concentration at the membrane and the quantity of particles leaving in the void peak.
The void peak is the first peak observed in the fractograms that correspond to the particles dispersed in the channel
cross-section.

50
(a) (b)

(c)

Figure 3.8: Signals of outlet flux of particles with 10 nm (blue line), 50 nm (orange line), and
100 nm (green line) for simulations using the elution permeate velocities of (a) U y =0 (v.u.), (b)
U y = 1 × 10−3 (v.u.), and (c) U y = 1 × 10−2 (v.u.).

Table 3.4 summarizes the residence time (t r ) calculated for each particle size subjected to the
different U y conditions, which result in different Pe y . Figure 3.9 presents the retention time of each
particle according to the applied U y . We observe that the increase of U y does not have a big impact
on the retention time of the particles with 10 nm, which are the most diffusive. On the other hand,
the increase of U y has an important consequence on the retention time of the largest objects (less
diffusive). Furthermore, the higher is the U y , the bigger is the shift between the retention time of the
particles with 50 nm and 100 nm. Thus, the tests show that it is easier to separate the populations
with the average diameter of 10 and 100 nm, than populations with 10 and 50 nm. In this latter
case, there is a more significant overlapping between the fluxes of 10 and 50 nm exiting the channel.
The separation can be optimized by further tuning the permeation conditions. The consequence of
increasing U y , therefore increasing t ∗ , is the peak flattening and broadening, which in terms of real
experiments result in higher sample dilution (peak collected during a larger time interval).

51
Table 3.4: Retention time of particles with 10, 50, and 100 nm subjected to an elution stage with
different permeation conditions.

d p (nm) Elution U y (v.u.) Pe y t r (t.u.) t∗


0 0 15.4 1.54
10 0.001 1 15.8 1.58
0.01 10 19.1 1.91
0 0 22.2 2.22
50 0.001 5 23.6 2.36
0.01 50 39.9 3.99
0 0 26.0 2.60
100 0.001 10 28.8 2.88
0.01 100 65.8 6.58

Figure 3.9: Retention time (t ∗ ) as function of particles sizes (d p ) for different permeation velocities
Uy.

Study of the permeable wall length: Next, we studied the effect of length of the accumulation
and elution regions, L1 and L2 , respectively. The first idea to elaborate the prototype consisted of
creating two regions, one dedicated for the accumulation step and the other for the elution stage. The
accumulation segment of length L1 (permeable wall) is used for the crucial particles accumulation at
the membrane (as shown in the next section). The elution segment of length L2 (solid wall) is used
to increase the sample retention time by shifting, in the flow direction, the layers that are located at
different streamlines of the laminar flow. The tests carried in this section have as objective to reveal
the role of membrane permeability in separation efficiency. To evaluate and compare the simulation
results for the different geometrical configurations, we calculated the quantity of 10 nm particles
that can be obtained at the outlet with a fraction (x 10 ) of 0.8.

The fraction of 10 nm particles (x 10 ) in a binary mixture of 10 nm + 50 nm, for instance, is


calculated according to Equation 3.36. The variables f10 and f50 represent the outlet flux of particles

52
with 10 nm and 50 nm, respectively.

f10
x 10 = (3.36)
f10 + f50

The amount of particles with 10 nm collected at the channel outlet (N10 ) was calculated by
integrating the outlet flux such that x 10 > 0.8, as shown in Equation 3.37.

Z t x 10 >0.8
N10 = f (t)d t (3.37)
0

The efficiency of the separation (η) was estimated based on the ratio between the amount of
particles with 10 nm, obtained respecting the condition x 10 > 0.8, and the total amount of injected
10 nm particles (Nin ).

N10
η(%) = × 100 (3.38)
Nin

The geometries tested were: geometry 1 with equal length for the accumulation and elution
segments (L1 /H = L2 /H = 5), geometry 2 that only has an accumulation segment by having a
channel fully covered by a membrane (L1 /H = 10 and L2 /H = 0), and geometry 3 that has an
elution region two times the length of the accumulation segment (L2 /H = 5 and L1 /H = 10). The
simulation of the accumulation and elution steps were performed for the particles with 10 nm (blue
lines), 50 nm (orange lines), and 100 nm (green lines) using the three geometries. The accumulation
conditions were Uin = 1 (v.u.) (U y calculated according to the membrane surface Am - dead-end
filtration) during 6H/U y (t.u.). The particles flux collected at the channel outlet are presented in
Figure 3.10 for (a) geometry 1, (b) geometry 2, and (c) geometry 3.

Comparing the results for the different permeable wall lengths presented in Figure 3.10, we ob-
serve that the elution signal depends on the channel permeability. When the wall is fully permeable
(geometry 2), the signals of the larger particles are flattened and broader, starting evacuating imme-
diately at the beginning of the elution step. The smaller particles have a faster evacuation presenting
a narrower signal and inferior retention time. This seems to be favorable for the separation. In ad-
dition, membrane modules with uniform wall permeability are the easiest to fabricate. The peaks
of geometry 3 have the higher retention (t ∗ is not higher because it is scaled by τ which is superior
due to the longer channel) times.

53
(a) (b)

(c)

Figure 3.10: Signal of the outlet flux of particles with 10 nm (blue line), 50 nm (orange line), and
100 nm (green line) for simulations using an elution permeate velocity U y = 1 × 10−2 (v.u.) and the
geometries (a) 1 (L1 /H = 5 and L2 /H = 5), (b) 2 (L1 /H = 10 and L2 /H = 0), and (3) 3 (L1 /H = 5
and L2 /H = 10).

Note that Figure 3.10-(a) does not correspond to the Figure 3.8-(c) simulated for the same
geometry and permeation conditions because the accumulation time is higher. This also reveals the
effect of the accumulation time (that is observed with the experiments performed in Section 5.1.1)
on the quantity of particles leaving in the void peak. The reduction of the accumulation time results
in a higher quantity of particles dispersed across the channel height that are rapidly evacuated from
the channel when the elution starts (void peak). The observation of the mentioned figures shows a
reduction of the void peak and an increase of the maximum of the signal for higher accumulation
time.

Figure 3.11 shows the retention time (t ∗ ) as function of the particles size (d p ) for the three
simulated geometries. Comparing the retention time obtained for geometry 2 with the retention
time observed for geometry 1, geometry 2 decreases the retention time of the particles with 10 and
50 nm, and increases the retention time for the larger objects (100 nm), resulting in a higher t ∗
shift between the particles with 10 nm and 100 nm. Geometry 3 has the higher retention time (as
expected since the channel is longer) but comparing to geometry 1 and 2, it has a smaller decrease

54
in the retention time for the particles with 50 nm and decreases the retention time of particles by
100 nm. The shift between particles with 10 and 100 nm is inferior, which is less interesting for
separation.

Figure 3.11: Retention time (t ∗ ) as function of particles sizes (d p ) for the different geometry config-
urations.

The geometry that improves better the recovery of particles with 10 nm from binary mixtures
with 10 nm + 50 nm and 10 nm + 100 nm is the geometry 2, with recoveries of 82% and 94%, re-
spectively. The recoveries of particles with 10 nm are in the same order for geometry 1 and geometry
3, being the worst recovery results obtained with geometry 3. The recovery of particles with 50 nm
and 100 nm in the binary mixture 50 nm + 100 nm were also compared for both geometries. The
results evidence a hard separation with the maximum recovery of 45% for the particles with 100 nm
and using geometry 2. The recovery values are summarized in Table 3.5.
Table 3.5: Efficiency of separation (η) of binary mixtures where one of the components represents
up to 80% of the composition. In 10 nm + 50 nm and 10 nm + 100 nm is 80% of 10 nm, and in 50
nm + 100 nm is 80% of 100 nm.

Binary Mixtures Geometry 1 Geometry 2 Geometry 3


10 nm + 50 nm 69% 10 nm 82% 10 nm 66% 10 nm
50 nm + 100 nm 39% 100 nm| 0% 50 nm 45% 100 nm | 0% 50 nm 19% 100 nm | 9% 50 nm
10 nm + 100 nm 90% 10 nm 94% 10 nm 89% 10 nm

Influence of the accumulation permeation velocity: The importance of the accumulation step
was highlighted through the performance of simulations with different U y during the accumulation
stage. The elution was performed with U x = 1 (v.u.) and U y = 0.01 (v.u.). The geometry dimensions
were L1/H = 10 and L2/H = 0, corresponding to a channel fully permeable (entirely covered by a
membrane). Thus, the best conditions obtained, in the previous sections, for the permeate velocity
during elution and geometrical conditions were used. The U y values tested were 0.1, 0.05, and
2.5 × 10−4 (v.u.). The U x during accumulations were adjusted to respect the conservation of the
fluid flow rate (U x S = 2U y Am , where S = 0.1 (a.u.) is the inlet cross section and Am = 10 (a.u.)

55
the membrane surface). Figure 3.12 shows the concentration profiles for the U y values of (a) 0.1
(v.u.), (b) 0.05 (v.u.), and (c) 2.5 × 10−4 (v.u.). From the accumulated profiles we obtain three
information. For a value of U y , taking as example the Figure 3.12-(a), the higher are the particles
diameter (d p ) and, consequently, the Pe y (defined as Pe y = U y H/D0 ), the concentration profiles are
closer to the membrane surface, as expected from Figure 2.3 presented in Chapter 2. The reduction
of the permeate profile that is observed when comparing Figure 3.12-(a) and 3.12-(b), results in less
concentration of particles at the membrane and thicker accumulated layers. The reduction of U y to
a value 200 times smaller than the value used in 3.12-(b), results in particles spreading across the
entire cross-section, as shows 3.12-(c).

These results evidence the crucial role of the accumulation step for a size-based separation using
the Flow-FFF mechanism. If the accumulation conditions (U y or Pe y and accumulation time) are
not adequate, the particles remain mixed and dispersed across the entire channel cross-section, as
illustrates Figure 3.12-(c), meaning that particles with different sizes occupy the same streamlines,
leaving the channel in simultaneous (the separation can not be performed).

(a) (b)

(c)

Figure 3.12: Normalized concentration profiles (scaled by Cin ) at half of channel in the y-direction
(wall-normal direction) for U y equal to (a) 0.1 (v.u.), (b) 0.05 (v.u.), and (c) 2.5 × 10−4 (v.u.).

56
3.2.2.3 Guidelines obtained for the prototype development

The tests performed with the numerical model suggest that the accumulation stage is essential for
the separation performance. The accumulation conditions to be adjusted to properly perform sample
accumulation and reduction of the quantity of dispersed species evacuated in the void peak are the
permeate flow (Q y ) and the duration (t a ). Moreover, they showed that the permeate flow during
the elution stage has an important role in increasing the particles retention and shifting the peaks
for different particles sizes. Thus, the accumulation and elution operating conditions have to be
experimentally optimized to increase the separation efficiency of the prototype. Sections 5.1.1 and
5.1.2 present the exploration of the accumulation and elution conditions of our prototype (SPFFF),
respectively.

The simulation results using different geometries suggest that the separation between particles
with size ratios 5 and 10 has higher efficiency if the channel is entirely composed of a semi-permeable
membrane. This means that there is no need for distinct accumulation and elution segments, sug-
gesting that separation in a simple membrane with uniform permeability may occur. From a technical
point of view, it is favorable since hollow-fiber modules composed of membranes with uniform per-
meability are easy to fabricate. Additionally, the membrane uniform permeability avoids conflicts
with patented configurations.

3.2.2.4 Model Improvement for high concentrations

The model described above consisted of a simple tool to guide the initial design of the experimental
prototype. As discussed along Chapter 5 and 6, the prototype needs to be studied and optimized.
Further studies must be carried to understand the particles spreading along the membrane during
the accumulation step. Strategies to have accumulation and elution regions must be investigated
since the system may require an accumulation and elution segments (both permeable) to reduce
peak broadening and, consequently, sample dilution. Although the separation can be achieved using
a uniform permeable membrane, it results in higher sample dilution. For the prototype optimization
studies, it is necessary to have a model that takes into account the effect of particles concentration
φ in the transport dynamics, especially for the studies to investigate the accumulation conditions
(particles accumulation creates regions of high concentration near the membrane).

A new model was implemented for future developments with the experimental prototype. In this
second model, the colloids physics and motion is described through the resolution of the transport
and hydrodynamics equations following a continuum mechanical approach (simplified17 from the
Suspension Balance Model - SBM [76]). The continuity, momentum, and transport equations are
solved considering a volume-averaged suspension velocity (ρm Um = φρ p U p + (1 − φ)ρU), where
ρm is the mixture density, Um is the mixture mass average velocity, ρ p is the particles density, and
U p and U are the velocities of the particles and fluid phases, respectively [76]). Assuming that the

17
Simplified because the Suspension Balance Model (SBM) relates the suspension rheology with the motion of the
particle, while in our model we consider a constant viscosity

57
colloids do not interact with the membrane, the set of equations is:

▽ · Um = 0 (3.39)

 ‹
∂ Um
ρm + Um · ▽Um = − ▽ p + ▽(µ(φ) ▽ Um ) (3.40)
∂t

∂φ
+ Um · ▽φ = ▽(D(φ) ▽ φ) (3.41)
∂t

Notice that the Einstein diffusion coefficient presented in Equation 3.28 is replaced by the col-
lective diffusion (D(φ)) to account for hydrodynamic and inter-particle interactions as discussed
later in this section [77].

We neglected the viscosity dependence with the concentration as a first approximation and
considered that the suspension viscosity is equal to that of the solvent.

The model configuration consists of a 2D axisymmetric flow in a tube. The r-coordinate is for
the radial direction, and the z-coordinate for the axial direction. To reduce simulation costs, we set
the mixture viscosity to constant and, we independently solve the flow and mass transport equations.
The flow equations are solved assuming a quasi-steady-state and constant membrane permeability.

The continuity and momentum equations considering steady-state, in the r-radial direction, and
z-axial directions are given (in cylindrical coordinates) by the Equations 3.42 and 3.43, respectively:

1 ∂ ∂
(ρr U r ) = − (ρUz ) (3.42)
r∂r ∂z

 ‹
∂p µ ∂ ∂ Uz
= r (3.43)
∂z r ∂r ∂r

where p is the pressure field [Pa = k g.m−1 .s−2 ].

Applying the no-slip condition at the channel walls and maximum velocity at the channel center
(parabolic flow) to Equation 3.43, Uz is described as:

 
R2 ∂ p r2
Uz = −1 (3.44)
4µ ∂ z R2

The U r component is obtained by substituting the Equation 3.44 in Equation 3.42:

 
∂ 2 p a2 r3
Ur = r− 2 (3.45)
∂ z 8µ 2R

58
The filtration law (U r = L p (p − Π(φ))) describes the reduction of the permeate flow caused
by the concentration of particles at the membrane (concentration polarization phenomena), which
leads to an increase of the osmotic (Π(φ)) and trans-membrane pressures. The combination of the
filtration law with Equation 3.45 permits the description of the pressure field following a second-
order differential equation:

∂ 2p
 ‹
1
= 16µL p [p − Π(φ)] (3.46)
∂z r(2R2 − r 2 )

Where L p (p.u.) is the membrane hydraulic permeability that is the inverse of the membrane
resistance R m (L p = 1/(µR m )).

The osmotic pressure (Π) permits a physical description of multi-body interactions [78]. This is
important to describe the particles collective behavior near the membrane where they are accumu-
lated and relaxed [64]. The concentration polarization leads to a gradient of the chemical potential
that originates a flux of solvent molecules to dilute the particles and establish equilibrium (equal-
ity of chemical potentials). The flux of solvent increases the resistance of the accumulated layer to
compression and causes an increase in the osmotic pressure. The approach of Wigner Seitz [79]
considers the contribution of entropy (due to Brownian motion), electrostatic repulsion, and van der
Waals attractive interactions in the osmotic pressure. This method was used by Bacchin et al. [64]
to model the filtration phenomena of concentration polarization, gelation, and particles deposition
for colloidal suspensions. The osmotic pressure (Π(φ)) is then considered as the sum of entropic
(Πent (φ)), electrostatic (Πele (φ)) and van der Waals contributions (Π vdw (φ)), according to the re-
lation 3.47. The contributions of each term are given in the appendix of the article [64]. The author
gives a detailed discussion about the influence of particle size on the magnitude of the different
contributions.

Π(φ) = Πent (φ) + Πele (φ) + Π vdw (φ) (3.47)

The generalized form of the diffusion coefficient (D(φ)) relates the particles motion with the
osmotic pressure, as discussed in Section 3.1.5.1.

D0 Vp dΠ
D(φ) = (3.48)
kB T H(φ) dφ

The model achieves a maximum osmotic pressure at a critical particles volume fraction (φcr ) that
corresponds to a phase transition. The osmotic pressure decrease after this point is not physically
real (the experimental curves behave similarly to Figure 3.5, i.e., the compression of the species
after phase transition increases the osmotic pressure). Therefore, this model of Π(φ) is limited for
φ<φcr . A new mathematical model was developed to describe the entire osmotic pressure curve
and implemented to simulated the dynamics of colloids relaxation near a gel-phase transition, as
discussed in the next section.

59
3.2.3 Second model: Modelling of near-phase transition and its impact in membrane
processes

3.2.3.1 Mathematical model for the osmotic pressure

As previously discussed in Section 3.1.4, the experimental determination of the osmotic pressure
curve for colloidal suspensions reveals a discontinuity associated with a phase transition. As illus-
trated in Figure 3.13, the osmotic pressure curves permit the differentiation of three regions: dis-
persed, transition, and concentrated. We developed a mathematical model to describe the different
regions of the osmotic pressure curve represented in different colors. The cyan line represents the
osmotic pressure curve (Π1 ) in the dispersed region, where electrostatic repulsions are the predomin-
ant phenomena. The pink (Π2 ) and purple (Π3 ) lines represent the transition region reached through
van der Waals attractive interactions, which has two limits (φ1 < φ < φcr and φcr < φ < φ2 ) to
ensure the continuity of the curve between the dispersed and concentrated regions. The φ1 and φ2
φ1 +φ2
values delimit the transition region, and φcr corresponds to their arithmetic average (φcr = 2 ).
The concentrated region represented by the brown line (Π4 ) is associated with the compression of
the condensed phase that results in an increase of resistance to compression, therefore in the osmotic
pressure increase.

Figure 3.13: Representation of classical osmotic pressure curves determined experimentally, high-
lighting the different phase regions of colloidal suspensions. Π - osmotic pressure of the dispersed
region; Π2 - osmotic pressure of the transition region for φ<φcr ; Π3 - osmotic pressure of the trans-
ition region for φ>φcr ; and Π4 - osmotic pressure of the concentrated region.

• Dispersed region (φ < φ1 ): The dispersed region is described by the classical Virial expansion
(analogous to the correction of the ideal gas law [57]), that corrects the van’t Hoff law using
the osmotic Virial coefficient (positives due to electrostatic repulsion), Equation 3.49. The
kB T
parameter A vH is the van’t Hoff coefficient (A vH = Vp [Pa]), B1 ([Pa]) is the first osmotic
Virial coefficient that accounts for two-bodies repulsion, and B2 ([Pa]) is the second osmotic
Virial coefficient accounting for three-bodies electrostatic repulsion.

B1 2 B
Π1 = φ + φ + 2 φ3 (3.49)
A vH A vH

60
The derivation of Equation 3.49 leads to Equation 3.50:

dΠ1 B1 B
=1+2 φ + 3 2 φ2 (3.50)
dφ A vH A vH

• Transition region (φ1 < φ < φ2 ): The transition region has an additional contribution to
the osmotic pressure Π t r which was deduced from the double-well function of Gibbs free en-
ergy, analogous to the Cahn-Hilliard formulation [80]. The Π t r and its derivative are given in
Equation 3.51 and 3.52, respectively.

(φ − φ1 )2 (φ − φ2 )2 (φcr − φ)
 ‹
Tc dΠ1
Πt r = 16 (3.51)
T dφ φ=φcr (φ2 − φ1 )4

€ Š€ Š
(φ2 −φ1 ) (φ2 −φ1 )
dΠ t r
 ‹
Tc dΠ1 (φ − φ1 )(φ − φ2 ) φ − φcr + p
2 5
φ − φcr − p
2 5
= 80
dφ T dφ φ=φcr (φ2 − φ1 )4
(3.52)
The parameter Tc /T represents the irreversibility degree of the phase transition (Tc /T = 0 - no
phase transition and Tc /T ≥ 1 irreversible phase transition), which is further discussed later
in this section.
The main properties of Equation 3.51 are that the minimum of its derivative (Equation 3.52),
corresponding to the depth of diffusion well, is obtained at φcr and that the transition region
does not change the osmotic pressure values of φ1 and φ2 .
The transition region received a mathematical treatment (divided for concentration inferior
and superior to φcr ) to guaranty the continuity of Π and dΠ/dφ. The region below φcr is
described by Equations 3.53 and 3.54, while the region above φcr are represented by Equations
3.55 and 3.56.

– φ1 < φ < φcr :

Π2 = Π1 + Π t r (3.53)

dΠ2 dΠ1 dΠ t r
= + (3.54)
dφ dφ dφ
– φcr < φ < φ2 :

Π3 = Π4 + Π t r (3.55)

dΠ3 dΠ4 dΠ t r
= + (3.56)
dφ dφ dφ
• Concentrated region(φ > φ2 ): After the disorder-order phase transition occurs, the increase
of colloids concentration leads to a resistance to compression increase (higher is the number
of particles harder is the compression to remove water). Thus, the osmotic pressure curve

61
increases until the maximum close packing (φcp ). After reaching φcp , the increase of φ leads
to particles deposition above the compressed layer, followed by the divergence of the osmotic
pressure curve (Π(φcp ) → +∞). The osmotic pressure after phase transition (Π4 ) is obtained
from the theory of the compressed cake [81], as presented in Equation 3.57. The derivative is
given in Equation 3.58.

– 1/k ™
Π1 φcp − φcr
Π4 = Π1 +k (φcp − φcr ) −1 (3.57)
φ=φcr dφ φ=φcr φcp − φ

  1/k
dΠ4 dΠ1 1/k φcp − φcr
=k (φcp − φcr ) (3.58)
dφ dφ φ=φcr φcp − φ φcp − φ

The parameter k is the compressive yield stress that represents the resistance to compression.
Higher values of k result in more compressible layers. Therefore, the increase of φ leads to a
smaller increase of Π. This parameter depends on the suspension properties18 (particle shape,
cake porosity, and cake tortuosity).

The osmotic pressure curve is modelled with a mathematical description for the dispersed, trans-
ition, and concentrated phases, which receive the parameters φ1 , φ2 , φcr , φcp , B1 , B2 , and k. These
parameters are obtained by fitting the experimental osmotic pressure curves with the mathematical
model. The parameter (Tc /T ) is not an experimental condition but a parameter to simulate the
proximity to the phase transition.

• Irreversibly degree (Tc /T ) of the phase transition:

The Tc /T parameter is a ratio between the temperature and the critical temperature (Tc ) at
which a phase transition occurs. We call the readers attention to the fact that Tc /T is a para-
meter of the model to simulate the degree of irreversibility of the phase transition rather than
an experimental operating condition. The osmotic pressure curves are determined at a con-
stant temperature. However, the model was developed in analogy with the vapor-liquid dia-
grams (PV-diagram). Figure 3.14 illustrates the Π∗ (φ) (scaled by the van’t Hoff coefficient
A vH ) diagram at different isotherms. When Tc /T is inferior to 1, the concentrated phase is
monophasic. The increase of Tc /T leads the system to a near-phase transition. At Tc /T equal
1, the system passes an instability point associated to a dΠ/dφ = 0, which results in a neg-
ligible diffusion coefficient (D(φ) ∼ dΠ(φ)/dφ). For higher values of Tc /T , the system goes
inside the diphasic region (delimited by the binodal and spinodal lines) and undergoes a phase
separation creating regions rich and poor in colloids. This results in a negative derivative of
the osmotic pressure and in a "negative"19 diffusion. The mathematical model is implemen-
ted in OpenFOAM to solve the transport equation presented in Equation 3.41, and a negative
18
Analogous to the Kozeny constant of the Kozeny-Carman equation, which relates the pressure drop of a laminar flow
that traverses a compressed layer of solids.
19
The negative diffusion does not have a physical meaning. When the two-phase transition is reached, an interface
between the concentrated and diluted regions is created.

62
diffusion crashes the solver. Only the region Tc /T < 1 can be simulated in OpenFOAM with a
continuum mechanical approach20 .

Figure 3.14: Dimensionless osmotic pressure Π∗ (scaled by the van’t Hoff coefficient A vH ) as function
of φ for the isotherms Tc /T at 0.6, 0.7, 0.8, 0.9, 1, 1.2, 1.5, 2, 3, and 5, corresponding to the blue,
orange, green, red, purple, brown, pink, grey, yellow, and cyan lines, respectively. Curves obtained
for B1 = 1, B2 = 0, φ1 =0.3, φ2 =0.5, and k=0.5. The dashed and solid blue lines correspond to the
binodal, and spinodal decomposition, respectively.

• Experimental parameters:

The experimental osmotic pressure data was obtained from [60] for silica nanoparticles with
a particles size (d p ) of 10.2 nm. Figure 3.15 shows the (a) experimental points and the fit-
ted lines for Tc /T equal to 0, 0.4, 0.8, 0.95, 1, and 1.5, and the corresponding (b) diffusion
coefficients (D(φ) scaled by D0 ). The increase of Tc /T increases the depth of the Π(φ) double-
well function and results, as discussed previously, in a negative diffusion. Notice that the two
maximum peaks in the transition region of the diffusion curve are a consequence of the Π(φ)
double-well function from which the diffusion is obtained and not physical phenomena.

The parameters to fit the experimental data are: φ1 = 0.03, φ2 = 0.09, φcr = 0.06, φcp = 0.72,
B1 = 4.0 × 104 [Pa], B2 = 15.0 × 105 [Pa], and k= 0.5. Notice that the different Tc /T only
affects the Π(φ) curve in the transition region (as expected since it only appears in the Π t r
contribution).

20
To simulate the phase transition is necessary to develop a two-phase model, which can be performed in OpenFOAM.

63
(a) Osmotic pressure Π(φ) curve (b) Diffusion coefficient D∗ = D(φ)/D0

Figure 3.15: (a) Experimental osmotic pressure Π(φ) curve measured by [60] for silica nanoparticles
with d p = 10.2 nm. The curve was fitted with φ1 = 0.03, φ2 = 0.09, φcr = 0.06, φcp = 0.72, B1 =
4.0 × 104 [Pa], B2 = 15.0 × 105 [Pa], and k= 0.5, for different Tc /T . (b) Diffusion coefficient (D∗
scaled by D0 ) derived from the osmotic pressure curve and illustrated for the same Tc /T .

3.2.3.2 Application of the mathematical model to study the mechanism of gel relaxation

The mathematical model to describe the Π(φ) curve was implemented in OpenFOAM to simulate the
relaxation mechanism of silica nanoparticles (radius of 10.2 nm). The results unveil the existence of
an expansion and a removal phase. When the filtration stops, the accumulated gel layer suffers an
initial expansion that is followed by the gradual particles release from the condensed phase to the
dispersed phase (bulk). The readers must refer to the article in Annex D for a complete description
of the simulated cases and results.

3.2.3.3 Fiber plugging caused by a gel phase relaxation

In membrane processes, such as wastewater treatment [82], a gel layer formation during fouling
can lead to membrane clogging (blockage of the membrane porous) or the blockage of the fiber
capillary [83]. The particles gel layer formed during filtration are relaxed when the permeate flux is
ceased (no drag force keeping the gel at the membrane). During relaxation, the gel layer suffers an
expansion followed by particles removal to the bulk (as discussed in the article presented in Annex
D). When the gel phase reaches the fiber center, a blockage of the capillary lumen is observed.

The fiber plugging caused by the gel layer relaxation is the subject of an ongoing article and
an abstract (Annex E) submitted for oral and poster presentation at the conferences "The North
American Membrane Society" and "EUROMEMBRANES 2021", respectively. In this study, we apply
the model presented in Section 3.2.3.1 to investigate the effect of the filtration Peclet number, initial
accumulated concentration, and gel reversibility (Tc /T ) in the fiber plug. We aim to predict and
optimize the cleaning stage of filtration processes by giving the threshold conditions to prevent fiber
plugging.

Considering that the gel phase is the region delimited by φ ≥ 0.6 (φcr =0.6 represented by the

64
black contour line), Figure 3.16 shows an example for the evolution of an accumulated layer during
the relaxation step. For a relaxation time, t 2 the fiber diameter is fully occupied by the gel phase
(fiber plugging).

65
(a) End of accumulation stage (relaxation time = t 0 )

(b) Relaxation time t 1

(c) Relaxation time t 2

Figure 3.16: Illustration of fiber clogging caused by the relaxation of a gel layer. Time t 0 corresponds
to the end of the accumulation stage and the times t 1 and t 2 represent different relaxation times.
The gel phase is the region delimited by φ ≥ 0.6 (φcr =0.6 represented by the black contour line).

66
Materials and methods

4
The prototype was developed with the aim of adapting the analytical HF5 (described in Chapter
2) to a separation process that can be operated in production lines. The geometrical configuration
and initial operating conditions were explored with the numerical 2D model presented in Section
3.2.2. The simulations revealed the importance of two factors for separation efficiency, the existence
of an accumulation step (with the channel outlet closed) and an elution step with a permeation
profile adapted according to the particle size. Therefore, the prototype was designed considering the
simulation guidelines and the industrial specifications and requirements, resulting in a modified HF5,
denominated hereafter as SPFFF (Sequential Prototype FFF). The differences between the SPFFF and
the classic HF5 are:

1. Simplification of the classical HF5 accumulation step that made the process complex (it re-
quires the flow inversion from the outlet, which can induce hydrodynamic instabilities, and
requires extra equipment) and limits the increase of the sample quantity (the HF5 accumula-
tion is performed in a small area and high sample quantity causes undesired effects such as
the appearance of a gel phase).

2. Development of a fully automated sequential process that can operate with a reduced number
(to the minimum) of sequential steps.

3. Hyphenation with an online fluorescence spectroscopy system (OFS) for the detection of the
elution peaks and characterization of the separation performance. The assembling of online
detectors with analytical FFF is common, however, the OFS permits the development of a
"Smart-Process" (not integrated during this Ph.D. but this online analytical system was chosen
to allow further prototype developments, as discussed in Section 4.5) by integrating all the
equipment signals in one software (LabVIEW or L I N EU P T M Series Software Development Kit
(SDK) from Fluigent) resulting in an automatic adaptation of the equipment conditions/posi-
tion according to the outlet signal.

The term SPFFF is used in this manuscript to refer to the experiments performed with the de-
veloped sequential prototype that comprehend the accumulation and elution steps. The HF5 and
FFF terms are applied to refer to the Hollow-Fiber Flow Field-Flow Fractionation and the general
Field-Flow Fractionation techniques, respectively.

This chapter outlines in Section 4.1 the prototype development, presenting the Process-Flow
Diagram (PFD), equipment details, operating steps, and the software used to build the experimental

67
protocols. Section 4.2 contains the model particles used to test the prototype performance, the
sample characterization by online fluorescence spectroscopy (OFS), and off-line dynamic light scat-
tering (DLS), and the solvent properties. The last section, Section 4.4, describes the verification
of the particles transport in the prototype by measuring the Residence Time Distribution (RTD) at
three different locations of the prototype. This allows us to verify mass conservation and compare
the mean residence time with the geometrical residence time.

4.1 Prototype development

4.1.1 Process-Flow Diagram (PFD)


The Process-Flow Diagram (PFD) of the experimental setup is presented in Figure 4.1. The system
works as a fixed-pressure operation. Compressed air is fed into a microfluidic flow control (MFCS-
EZ) and is used to pressurize the recipients containing the solvent (RS and RP) and the sample (RE).
The pressurization induces a pressure-driven flow of the solvent, sample injection, and control of the
permeation flux (through the regulation of the transmembrane-pressure) if the pressurized recipients
are the RS, RE, and RP, respectively. The channel (MHF5) is a hollow-fiber module, having an L-
SWITCH valve (LV) and a flow-meter (RD1) placed at its entry. At the fiber outlet, a cell (FC) for
the fluorescence spectroscopy measurements, a flow-meter (RD2), and an M-SWITCH valve (MV)
are included. The FC is connected to a light source (DH-2000) and a spectrometer (HDX). The M-
SWITCH valve is connected to several reservoirs (REF1 to REFn, n=9 being the maximum number
of reservoirs) for fraction collection during the experimental run time. The flow-rates (measured
with RD1 and RD2) and pressure data (imposed with MFCS-EZ on RS and RP) are saved during
the experiments run time every 1s with the software All-In-One (AIO). The experimental protocol, to
impose the operating conditions and automatically perform the accumulation and elution stages, was
developed in the software Microfluidics-Automation-Tool (MAT), which is further discussed in Section
4.1.4. The optical measurements (detection of the peaks elution at the channel outlet by the online
fluorescence spectroscopy system (OFS)) are performed and saved using the software OceanView, as
detailed in Section 4.1.2.6. Table 4.1 contains the dimensions (internal diameter (I D), length (L),
and volume (V )) of the prototype equipment and tubing and Table 4.2 gives the specifications of the
reservoirs (RS, RE, RD, RP, and REFn).

68
69

Figure 4.1: Process flow diagram of the experimental setup.


Table 4.1: Dimensions of the prototype equipment and tubing.

Internal diameter, ID Length, L Volume, V Supplier


Tube / Equipment
(mm) (cm) (mm3 =µL) (reference)
Fluigent
1 Solvent feed 0.508 49.3 100.53
(FEP Tubing 1/16 OD x 0.020 ID)
Fluigent
2 Sample feed 0.508 32.3 65.47
(FEP Tubing 1/16 OD x 0.020 ID)
Fluigent
3 Injection Loop 0.508 16.3 33.04
(FEP Tubing 1/16 OD x 0.020 ID)
Fluigent
4 Tube before the flow-meter RD1 0.508 12.9 26.15
(FEP Tubing 1/16 OD x 0.020 ID)
Fluigent
5 Tube after the flow-meter RD1 0.508 3.6 7.30
(FEP Tubing 1/16 OD x 0.020 ID)
Fluigent
6 Tube before the fluo cell FC 0.508 4.5 9.12
(FEP Tubing 1/16 OD x 0.020 ID)
Fluigent
7 Tube after the fluo cell FC 0.508 7.8 15.81
(FEP Tubing 1/16 OD x 0.020 ID)
Fluigent
8 Tube after the flow-meter RD2 0.508 10.9 22.09
(FEP Tubing 1/16 OD x 0.020 ID)
Fluigent
LV - - 0.66
(L-SWITCH)
Fluigent
RD1 1.8 5 127.23
(XL-UNIT)
MHF5 1.4 26.5 407.94 Lab-made
Fluigent
RD2 1 5 39.27
(L-UNIT)
Fluigent
MV - - 11.6
(M-SWITCH)
Ocean Insight
FC 2 - 30
(79013)

Table 4.2: Reservoirs specifications.

Supplier
Reservoir Volume Material
(reference)
Standard lab bottles
RS 1L Cap from Fluigent
in borosilicate
Fluigent
RE 15 mL Polyoxymethylene
(FLUIWELL 1C-15 nr. 21015001)
RD 40 mL Polystyrene Common laboratory recipients
Fluigent
RP 50 mL Polyoxymethylene
(FLUIWELL 1C-50 nr. 21050001)
REFn 5 mL Polystyrene Common laboratory recipients

4.1.2 Equipment details


In this section, the details of the basic prototype equipment are presented. The microfluidic flow con-
trol (MFCS-EZ) is essential for the fixed pressure operation, i.e., to have sample and solvent flux. The
valves (LV and MV) are responsible for the performance of the automated sequences (accumulation
and elution). The flow-meters (RD1 and RD2) are important for the calculation of the flow-rates,

70
which allows the verification of the system hydrodynamics, and the calculation of the mass balance.
The channel (MHF5) is the key piece of the prototype, where the separation mechanism occurs. The
online fluorescence spectroscopy system (OFS) composed by the fluorescence cell (FC), light source
(DH-2000), and spectrometer (HDX), permits the detection in real-time of the elution peaks and,
along with the flow-rates data, allows the calculation of the mass balance. Appendix F contains the
specification sheets of the equipment, containing the more relevant equipment characteristics and
instructions for correct operation, the equipment prices, and the occurred malfunctions (including
the problems resolution and prevention measures).

4.1.2.1 Microfluidic flow control (MFCS-EZ)

The microfluidic flow control (MFCS-EZ) used in the prototype is the M F C S T M − E Z model from
Fluigent. It has a total of three channels (Fluigent have other M F C S T M − E Z systems with more
channels and a most modern system, the L I N EU P T M , that permits the assembly of several channels
with different pressure and vacuum ranges) where two channels have a pressure feed up to 1 bar
and one channel have a feed up to 345 mbar. The M F C S T M − E Z comes with an air-drier and a
manual pressure regulator used to connect the equipment to the local pressure network. For a proper
operation, the equipment needs an input pressure of 1.3 bar.

4.1.2.2 Valve at the channel inlet (L-SWITCH)

An L-SWITCH (LV) rotatory valve model from Fluigent having six ports and two positions is placed at
the channel inlet to allow the automatic switch between the solvent feed and the sample injection.
Figure 4.2 shows the side and top views of the valve and allows the visualization of the injection loop
(tube connecting two valve ports). During the preparation of the injection step, the valve changes to
position 2 to enable the sample to enter the feed loop. During injection, the valve returns to position
1, and the solvent transports the sample into the channel. The software ESS T M C ont r ol allows the
visualization and manual control of the valve positions, while the software Microfluidic Automation
Tool (MAT) permits automatic control of the valve through its integration in an experimental protocol.
The valve has three channels with a total internal volume of 660 nL.

Figure 4.2: Image with the side and top views of the L-SWITCH valve and visualization of the two
valve positions with the ESS T M C ont r ol software.

71
4.1.2.3 Valve at the channel outlet (M-SWITCH)

A Fluigent M-SWITCH (MV) rotatory valve, containing eleven ports and ten positions, is inserted at
the channel outlet. During the accumulation stage, the valve is in a closed position, while throughout
the elution stage, it alternates between opened positions. In this way, it is possible to collect samples
with different retention times in distinct recipients. Figure 4.3 presents the side and top views of
the M-SWITCH valve and the visualization of the valve in a closed and an opened position on the
software ESS T M C ont r ol. To have closed positions, an M-SWITCH plug is inserted in the valve port.

Figure 4.3: Image with the side and top views of the M-SWITCH valve and visualization of two valve
positions with the ESS T M C ont r ol software.

The valve is composed of a rotor seal (Figure 4.4) made of polytetrafluoroethylene (PTFE) to
reduce friction during rotation and a moving channel with 1 mm length, 11.6 µL of internal volume,
and no dead-volume (information given by the supplier Fluigent).

Figure 4.4: Image of the M-SWITCH rotor seal provided by Fluigent.

The rotor is pressed against a fixed stator to have proper sealing. The Figure 4.5 presents the
valve structure.

72
Figure 4.5: Image of the M-SWITCH structure provided by Fluigent.

According to the supplier information, the L-SWITCH valve has a similar structure, but the rotor
seal has three semi-circular channels instead of the one presented in Figure 4.4.

4.1.2.4 Flow-meters (RD1 and RD2)

The flow-meters are thermal sensors having a heater in the center and two temperature sensors
placed on both sides of the heater, as illustrated in Figure 4.6.

Figure 4.6: Operating principle of the flow-meters. Adapted from [84].

The heater gives a specific heat flow-rate (h f in [J/s]) to the fluid:

h f = q/t (4.1)

where q is the heat (J) and t is the duration (s) of heat transfer to the circulating fluid.

The temperature sensors measure the temperature difference (∆T [◦ C]), and the volumetric
flow-rate Q v [m3 /s] is calculated according to:

hf
Qv = (4.2)
ρC p ∆T

73
Where ρ is the fluid density (k g/m3 ) and C p is the fluid specific heat capacity at constant
pressure (J/(k g.K)), being 995.650 kg/m3 and 4.1774 kJ/(k g.K) for water at 20 ◦ C and 1 bar
(from [85] - Table 2.355 of page 2-311), respectively.

The inlet flow-meter RD1 is an XL-unit with a maximum flow-rate of 5 mL/min (measured
with water) and a sensor inner diameter of 1.8 mm. The outlet flow-meter RD2 is an L-unit with a
maximum flow-rate of 1 mL/min (measured with water), having 1.0 mm of sensor inner diameter.
Both flow-meters are from Fluigent and present accuracy of 5% of the measured value.

4.1.2.5 Channel (MHF5)

The MHF5 channel is laboratory-made, consisting of a hollow fiber whose extremities are glued
inside a plastic tube. The fiber is a regenerated cellulose membrane (RC) with a 30 kDa cut-off,
chosen because of its abundant and rapid availability in the laboratory and its classical use in the
FFF experiments (as revised in Chapter 2). Notice that the membrane properties might be an import-
ant parameter to explore. In particular, the membrane composition, surface charge, and roughness
should be considered for the SPFFF prototype future optimizations. The dimensions are 1.4 mm
of inner diameter (I D) and ∼ O(30) cm in length (L), having the ratio I D/L ∼ O(5 × 10−3 ) in the
range of the HF5 classic modules. The tests performed using the 2D model discussed in Section 3.2.2
suggested that the separation performance is improved by having a permeation flux along the entire
channel length. Therefore, there was no alteration to the membrane porosity (no glue was added
to create regions without permeation). The membrane permeability (L p ) to ultra-pure water is ∼
O(3×10−5 ) m.s−1 .bar −1 , and slightly decreases due to particle deposition in the membrane pores in
time, after several experimental runs, being often restored with backwash steps and cleaning solu-
tions (Ethanol 40% (v/v) and sodium hypochlorite (NaClO) 50 ppm). Before each experiment, the
L p should be verified to guarantee the reproducibility of the results. The verification is realized by
measuring the permeation flow-rate when the imposed pressure (Pi ) is 500 mbar (if the permeation
flow-rate (Q r ) is in the range 600 − 700 µL/min the membrane is considered proper). If there is
a significant decrease (not reverted by cleaning due to irreversible fouling) or increase (fiber per-
foration), it is necessary to replace the membrane. Since the modules are handmade, the length
of the membrane measured within the unglued region differs between different modules. The per-
meate contained inside the plastic tube can affect the separation performance when it has residues
of cleaning solutions or solvents from previous experiments. As prevention, the plastic tube needs to
be cleaned and the permeate replaced by the desired solvent. Table 4.3 summarizes the properties
of two modules used in the experiments. The M1 module was used to perform the verification of the
system hydrodynamics presented in Section 4.4, and the M2 was used in Chapter 5 and 6, for the
exploration of the operating conditions and test of the separation performance, respectively.

74
Figure 4.7: MHF5 channel composed by a regenerated cellulose hollow-fiber membrane placed inside
a plastic tube.

Table 4.3: Summary of the channel properties.

I D (mm) L (cm) L p (m.s−1 .bar−1 ) Cut-off (kDa) V (mL)


Fiber 1.4 27 2.9 × 10−5 30 0.416
M1
Plastic tube 15 29.6 - - 52.3
−5
Fiber 1.4 26.5 3.5 × 10 30 0.408
M2
Plastic tube 15 39 - - 68.9

4.1.2.6 Online fluorescence spectroscopy (OFS) system (FC, HDX, and DH-2000)

The evaluation of the experimental results was initially performed by collecting experimental frac-
tions at the channel outlet and posterior analysis in a UV spectrometer (UV/VIS lambda 365 PerkinElmer)
through the measurement of absorbed (or transmitted) light. During the SPFFF, the sample is ∼
O(100×) diluted compared to the injected concentration, and the volume of each fraction collected
was insufficient to fill the minimum volume of the analytical cell. Therefore, it was necessary to
add ultra-pure water to reach the required volume, resulting in a concentration below the detection
limit. To overcome this limitation, a more sensitive analytical system had to be used.

An online fluorescence spectroscopy (OFS) system was integrated into the prototype to enable
direct visualization of the experimental results and improvement of the detection at low concentra-
tions. In addition to overcoming the sensitivity limit of the offline absorbance detector, the online
fluorescence spectroscopy permits real-time visualization of the experimental results (analysis time
saving).

The principle of fluorescence spectroscopy is the emission of light from excited electronic singlet
states, as illustrated in Figure 4.8. When a sample is excited by the incidence of light, the photons
absorption (hc/λe x c , where h= 6.63 × 10−34 J/s is the Planck constant and c= 3 × 108 m/s is the
light speed) promotes an electronic transition from a low energy level (S0 ) to a higher energy level
(S1′ ). The electrons remain in this state for 10−15 − 10−9 seconds, being subjected to conformational
changes and interactions that causes, among other phenomena, an energy dissipation [86]. This
results in an excitation stage S1 with lower energy and, consequently, in emission wavelengths (λemi )
superior to the excitation ones (λe x c ). The electrons in the excited (S1′ and S1 ) and ground (S0 ) states
have opposite spins, thus the return from S1 to S0 (by the emission of light hc/λemi ) is spin allowed
and results in a rapid electronic transition (∼ O(10) ns) [87].

75
Fluorescence has higher sensitivity to low concentrations than absorbance because of the differ-
ence between λe x c and λemi , designated as Stokes shift. In absorbance, the quantity of light absorbed
by a sample is measured compared to the incident light, both at the same wavelength and in a straight
geometry. Diluted samples absorb small amounts of light, being in the range of the detection error. In
fluorescence spectroscopy, the measurement is realized at 90◦ allowing the detection of the emitted
photons isolated from the incident light (dark background).

Figure 4.8: Simplified reproduction of the Jablonski diagram. h= 6.63 × 10−34 J/s is the Planck
constant and c= 3 × 108 m/s is the light speed.

The components of the online fluorescence system (OFS) integrated into the prototype are
presented in Figure 4.9. It contains a UV-VIS spectrometer (model OCEAN-HDX-UV-VIS) with a
wavelength range of 200-800 nm, a light source (model DH-2000-S-DUV-TTL), a fluorescence cell
(model FIA-SMA-FL-ULT) made of Ultem (polyetherimide), and two optical fibers (model P600-1-
UV-VIS), both from Ocean Insight. The light source has two bulbs, one of deuterium (model DH-
2000-DUV-B) and the other of tungsten halogen (model DH-2000-DH), for light emission in the
wavelength ranges of 190-400 nm and 360-2500 nm, respectively. The optical fibers connect the
fluorescence cell (FC) to the light source (DH-2000) and the spectrometer (HDX), transporting light
with a wavelength range of 300-1100 nm. The emitted light is detected at 90◦ angle due to the FC
geometry. Both the light source and spectrometer are from Ocean Insight.

76
Figure 4.9: Equipment for the assembling of the online fluorescence spectroscopy analytical method.

It is possible to perform absorbance spectroscopy by simply replacing the FC cell by an ab-


sorbance cell (when the light source and detector are aligned (straight channel), light absorption
is measured, unlike when they are perpendicular (90◦ ), in which case, light emission is measured).
The absorbance cell (ref: FIA-Z-SMA-PEEK) was tested. We observed a loss of detection sensitivity:
the use of a slit connector (connection between the optical fiber and the spectrometer) with a smaller
opening to reduce light throughput is suggested for further developments with an absorbance cell.

The use of the OFS requires particles with optical properties (natural optical properties or
marked with fluorochromes), thus the prototype was developed and explored using model particles
marked with extrinsic fluorochromes, which are characterized by having excitation and emission
spectra (sample properties presented in Table 4.4 - Section 4.2). When the sample leaves the MHF5
channel and circulates inside the fluorescence cell (FC), the fluorochromes are excited by the incid-
ence of the UV-VIS light supplied by the source (DH-2000) and the photons emission are detected by
the spectrometer (HDX). The output of the signal received on the spectrometer is controlled using
the software OceanView. A protocol was built to present the data of signal intensity (I) at specific
wavelengths (λ) as function of time (t) and the signal intensity (I) as function of the wavelength (λ),
illustrated on Figure G.1 (Annex G). The fractograms presented in this chapter were determined by
using the deuterium light (halogen light was off) and using an acquisition frequency of 200 s−1 and
500 ms of integration time. At the beginning of the experiments, the background spectra (caused
by heating of the light source) must be eliminated by turning off the lamps. This procedure brings
the signal intensity to zero when no light source is on. With time, the background spectrum tends
to increase (Figure G.2 presented in the Annex G, shows the spectrum after a background measure-
ment and the intensity increase due to lamp heating, respectively). Therefore, it is necessary to wait
∼ 25 minutes for stabilization of the light source temperature (warm-up time) and then eliminate
the background spectra before each experiment.

To better understand the OFS signal, it is necessary to observe the emitted light spectra in the

77
absence of fluorochromes. Figure 4.10 presents the spectra of the light sources when ultra-pure
water is circulating in the cell (water molecules are not fluorescent). The blue line corresponds to the
deuterium lamp, having emission in the wavelength range of 200-800 nm (although the wavelength
range limit given by the supplier is 400 nm). The orange line corresponds to the halogen lamp with
an emission in a wavelength range of 500-800 nm. The green curve represents the spectrometer
signal when both lamps are on. Only the deuterium lamp is used because the model particles have
a maximum excitation wavelength at 546 nm, being in the range of the deuterium lamp. The use
of halogen would only increase the intensity signal in the region 500-800 nm, as observed with the
green line of Figure 4.10.

The deuterium spectrum is characteristic of the bulb, only varying if there is a species with
optical properties circulating or adsorbed at the fluorescence cell (FC) or bubbles that can reflect the
light. It is independent of the flow-rate, and it can serve as a reference spectrum for FC cleaning or
windows replacement. At the end of an experiment, no more particles are circulating in FC, thus
the I vs λ signal must overlap the deuterium spectrum. If a deviation is observed, the FC can have
particles adsorbed at the walls and windows or have air bubbles inside. During the experiments, the
deuterium spectrum could be eliminated to only observe the sample signal, however, the instabilities
mentioned above would be hard to detect. We decided to keep the deuterium spectrum, defined
hereafter as reference spectra, and subtract the baseline of the I vs t signals during the experimental
data treatment (python script).

Figure 4.10: Intensity signal (I) as function of the wavelength (λ) of the light source.

4.1.3 Operating Steps


The operating principle of the sequential prototype, as mentioned previously, consists of an accu-
mulation step followed by the sample elution. The accumulation stage was designed to simplify
the accumulation (in a confined region) of the classic HF5 experiments. During accumulation, the
channel outlet is closed, and all the inlet solvent is forced to leave through the membrane, transport-
ing the particles to the membrane surface. In this way, the differential radial transport of particles
with different diameters occurs without an outlet flow inversion. During elution, the particles are

78
transported by the solvent in the axial direction, and are collected at the channel outlet. This section
describes the equipment operation to perform the automatic switch between injection preparation,
accumulation and elution and presents the operating conditions of each step.

The cylindrical coordinates system is adopted in the next sections, standing r-axis for the radial
direction and z-axis for the axial direction.

4.1.3.1 Injection Preparation

At the beginning of an experiment, the system is filled with the desired solvent, as illustrated in
Figure 4.11.

Figure 4.11: Filling the experimental setup with the required solvent before sample injection.

During the injection step, there is no solvent flux in the channel, and the L-SWITCH (LV) valve
is in position 2 to connect the sample reservoir to the injection loop, as illustrated in Figure 4.12.
The duration of this step depends on the injection volume loop and the sample flow-rate. In the
experiments presented in this manuscript, the injection loop has ∼ O(30)µL and is filled in ∼ O(10)
seconds by the imposition of 200 mbar in the sample reservoir (RE). The choice of the injection loop
volume is justified in Annex K.

Figure 4.12: Illustration of the injection stage preparation with L-SWITCH in position 2.

The injection parameter is the sample quantity min j , which depends on the volume of the injec-
tion loop (Vin j and on the sample concentration (Cin j ), Equation 4.3. In the experiments presented
in Section 6.3.1, the injection quantity was varied by increasing the sample concentration.

min j = Vin j × Cin j (4.3)

79
4.1.3.2 Accumulation

The accumulation step occurs after the filling of the injection loop. The L-SWITCH (LV) valve moves
to position 1, connecting the injection loop to the solvent reservoir (RS) and the channel inlet. In
simultaneous, the M-SWITCH (MV) valve moves to a closed position. The solvent reservoir (RS) is
pressurized, transporting the sample to the channel and towards the membrane, thus injection and
accumulations occur simultaneously. Figure 4.13 illustrates the accumulation of a binary system: a
differential radial position between smaller and larger particles represented by the orange and yellow
colors, respectively. The accumulation flow-rate (Q r a ) and time (t a ) need to be set to avoid particle
aggregation, clogging, and/or adsorption to the fiber. The accumulation time has to be adjusted
accordingly to the flow-rate conditions, particle size, and injected quantity (controlled by increasing
the sample concentration) to promote an efficient accumulation at the membrane.

Figure 4.13: Illustration of the injection and accumulation steps for a binary mixture. After accumu-
lation, the smaller particles (represented in orange) are closer to the center than the larger particles
(represented in yellow) which are closer to the membrane.

A permeate Peclet number Pe r is defined as:

Ū r R f
Pe r = (4.4)
D0

Where Ū r is the arithmetic average of the radial velocity of the fluid transporting the particles
to the membrane surface, R f is the channel radius, and D0 is the particles diffusion coefficient in
diluted conditions.

The t a must be longer than the time required for the sample to be transported t t r ansp to the
membrane surface. The t t r ansp has to ensure that the volume of fluid filtered during accumulation is
greater than the volume of the sample to inject (Vin j ) and the total volume of the equipment placed
between the injection loop and the end of the channel Veq , as represented in Figure 4.14.

Figure 4.14: Scheme of the filtered volume.

80
The filtration volume (Vf ) is given by the arithmetic average of the permeate flow-rate (Q r ) and
filtration time (t t r ansp ) according to Equation 4.5.

Vin j + Veq
Vf = Vin j + Veq ⇔ Q̄ r × t t r ansp = Vin j + Veq ⇔ t t r ansp = (4.5)
Q̄ r

The t a can be estimated using Equation 4.6.

Vin j + Vp
t a ≥ t t r ansp ⇔ t a ≥ (4.6)
Q̄ r

The accumulation conditions (Q r a and t a ) are explored in Section 5.1.1.

4.1.3.3 Elution

At the end of the accumulation starts the elution step. The M-SWITCH (MV) valve moves to an
opened position, and the solvent flux transports the particle layers towards the channel outlet. The
M-SWITCH (MV) valve has 9 available opened positions, allowing the collection of fractions with
different size distributions, as illustrated in Figures 4.15 and 4.16. The switch in the M-SWITCH (MV)
positions is programmed in MAT according to the retention time of the different particles population.

Figure 4.15: Illustration of the elution step with the collection of the more diffusive fraction (rep-
resented in orange).

Figure 4.16: Illustration of the elution step with the collection of the less diffusive fraction (repres-
ented in yellow).

81
The parameters of the elution step are the axial flow-rate (Q z ), and the permeate flow-rate
(Q r e ), both investigated in Section 5.1.2.

4.1.4 Protocol development on Microfluidics Automation Tool (MAT)


The Microfluidics Automation Tool (MAT) is a software to build the experimental protocol to control
the F lui g ent equipment, microfluidic flow controller (MFCS-EZ), valves (LV and MV), and the flow-
meters (RD1 and RD2). An example of a protocol developed to perform the complete sequence of
injection, accumulation, and elution steps are presented in Figure 4.17. At the top-right of MAT there
is a list of the connected equipment, and, by selecting one equipment (in Figure 4.17 it is illustrated
the selection of the channel 1 of the microfluidic flow controller), appears in the top-left corner
the available function. The desired function is dragged and dropped to the protocol area. In the
protocol area, there is a toolbox with functions that allow grouping operations, repeat operations,
repeat operations under specific conditions, use the if statement, and keep an operation for a period
of time. The MAT can perform parallel instructions, for instance, it can increase pressure during one
hour and change the valve position every 10 minutes, but the list of actions has to be given in series.
For illustration, in the protocol presented in Figure 4.17 there is a linear increase of the pressure in
channel 1 (1 mbar during 60 min - P1 = 1 × t) that is realized with the actions contained in the
"Repeat n times" block. Inside this block, it is imposed a pressure step equal to 1 mbar ("MFCS 1
Channel #1 Pressure Step") during 1 min ("Wait"), which is repeated 60 times with the "Repeat n
times". To add the instruction to change the valve position every 10 minutes, it is necessary to create
6 blocks of "Repeat n times" with 10 repetitions instead of 60, and add the valve instruction below
each block, as illustrated in Figure 4.18.

Figure 4.17: Screenshot of a protocol developed on Microfluidics Automation Tool (MAT).

82
Figure 4.18: Instructions to perform a linear pressure increase and to change the valve position (two
actions in parallel with instructions designed in series).

83
4.1.5 Photo of the developed prototype (SPFFF)
Figure 4.19 shows the (a) side and (b) top view of the developed sequential prototype (SPFFF).

(a) Side view

(b) Top view

Figure 4.19: Setup of the sequential prototype.

4.2 Sample properties and characterization


The experimental prototype was tested using model particles. The model particles are fluorescent
spherical latex with 50, 100, 200 and 500 nm Fluoresbrite carboxylate polystyrene from Polyscience

84
(Biovalley distributor in France), silica oxide with 100 nm from micromod.

Table 4.4 presents the properties of the model particles (dispersed in water) in terms of fluoro-
chromes optical properties (maximum excitation-λe x c and emission wavelength-λemi ), size (d p ), sur-
face charge (zeta potential-ζ), and pH (which gives information about the concentration of H+ and
OH – ions in solution that affect the zeta potential value). The titration curves (ζ vs pH) reveal that
the particles are negatively charged over a pH range 2-10 (Annex H). Table H.1 of the Annex H gives
the samples price.
Table 4.4: Properties of the model particles.

Sample Supplier
Type of sample Size (d p ) λexc λemi z-average pH ζ
designation (reference)
Carboxylate polystyrene
L50 Biovalley (16661-10) 50 nm 441 nm 486 nm 62.0 nm 4.19 (24.9 ◦ C) -43.2 mV
nanoparticles −(CH2 −CH−Ph)n −
Carboxylate polystyrene
L100 Biovalley (19774-10) 100 nm 360 nm 407 nm 77.8 nm 4.31 (26.1 ◦ C) -35.5 mV
nanoparticles −(CH2 −CH−Ph)n −
Carboxylate polystyrene
L200 Biovalley (19391-10) 200 nm 529 nm 546 nm 110.9 nm 5.17 (24.4 ◦ C) -40.1 mV
nanoparticles −(CH2 −CH−Ph)n −
Carboxylate polystyrene
L500 Biovalley (19391-10) 500 nm 529 nm 546 nm 358.3 nm 5.50 (24.6◦ C) -33.8 mV
nanoparticles −(CH2 −CH−Ph)n −
Silica oxide
S100 micromod (42-00-102) 100 nm 485 nm 510 nm 98.3 nm 6.03 (21.4◦ C) -34.5 mV
nanoparticles SiO2

The size (z-av g) and surface charge (ζ) were determined by Dynamic Light Scattering (DLS),
respectively, being the method detailed in Section 4.2.1. The pH measurement was performed with
a pH-meter (model 315i from WTW).

4.2.1 Size and surface charge characterization


Light scattering is the natural phenomenon that occurs when light interacts with matter and is di-
vided into Static, Dynamic, and Raman scattering. The three types of light scattering are defined
accordingly with the information analyzed (Static Light Scattering analysis average intensity and
Dynamic Light Scattering measures intensity fluctuations) and the wavelength of the scattered and
incident light (in Static and Dynamic Light Scattering the scattered and incident light have the same
wavelength, while in Raman they are different). The Static Light Scattering is an important tool
for the determination of polymer molar mass, Raman spectroscopy gives information about molecu-
lar structure, and Dynamic Light Scattering is used to determine the hydrodynamic size (d p ) and
potential zeta (ζ) of colloids [15].

The samples used for the prototype tests are spherical model particles made of polystyrene
Latex or silica oxide (SiO2 ), thus their characterization relies on the size (d p ) and zeta potential (ζ)
determination and size distribution for the fractions collected during SPFFF experiments. Therefore,
only the Dynamic Light Scattering (DLS) technique is applied.

The Zetasizer nano ZS90 from Malvern was used to determine the zeta potential (ζ) and the
particles size (d p ), as describes the Sections 4.2.1.1 and 4.2.1.2, respectively.

85
4.2.1.1 Measurement of the Zeta potential (ζ)

The zeta potential (potential that exists at the slipping plane, as discussed in Section 3.1.3.1) is
determined by the electrophoresis mechanism, which consists of the motion of the particles in a
stationary fluid due to the application of an electric field. The particle velocity in an electric field,
also denominated as electrophoretic mobility, µ E is related to the zeta potential (ζ) through the
equation of Henry:

Up 2ε r ζ f c (k D a)
µE = = (4.7)
Ef 3µ

Where, U p is the particles velocity, E f is the applied electric field, ε r is the dielectric constant, µ
is the fluid viscosity, and f c (K D a) is the correction function that is related to the particle radius (a)
and to the Debye length (k D = 1/λ D ). For measurements in aqueous media with moderate electrolyte
concentration, the value of f c (K D a) used is 1.5 (Helmholtz-Smoluchowski approximation).

During measurement, a laser beam is divided into two beams, one being directed to the detector
(reference beam) and the other passes by an attenuator (which controls the light quantity according
to the sample concentration) and crosses the measurement cell in the center. The scattered signal
passes by compensation optics equipment, being received in the detector. The cell has two elec-
trodes, and the application of an electric field induces particle movement towards the electrode with
an opposite charge (electrophoresis). This movement causes a fluctuation of the incident beam light
that is scattered, and its frequency is proportional to the particle velocity. The two beams (refer-
ence and scattered) are compared, and the electrophoretic mobility is determined by Laser Doppler
Velocimetry (based on the Doppler effect) using the equipment software.

4.2.1.2 Determination of the size (d p )

The principle behind the colloids size measurement by DLS is based on the colloids Brownian motion
that causes fluctuations on the scattered light. The light illuminates the sample, and the scattered
signal detection occurs at 90◦ . The intensity fluctuation information is transformed into an auto-
correlation function (in the correlator), which expresses the probability of a particle being at the
same position as a function of time. Initially, the probability is high and constant, decreasing with
time until zero (for a short time, the position of the particle remains correlated to the initial position,
and with the time passing, the position becomes uncorrelated due to Brownian motion). Smaller
particles have higher diffusion coefficients and present a faster decay of the auto-correlation function
than the larger ones. The size distribution is obtained with the Stokes-Einstein relationship from the
analysis of the auto-correlation function. The software uses algorithms to predict a size distribution
that fits the measured signals.

In DLS, there are two methods for extracting the size distribution from the auto-correlation
function, namely the regulation method and the cumulant method. The Zetasizer nano ZS90 uses
the cumulant method (for a detailed description of the cumulant method, the reader should refer
to the article [88]), which allows determining the z-average parameter (z-av g) [89]. The z-av g is
calculated from the particles intensity (I) and size (d p ) according to Equation 4.8 [90] (where N is

86
the total particles number, and i is the particle number i), being a widely used parameter because of
its computational stability and insensitivity to noise.

PN
Ii
z-avg = PNi Ii
(4.8)
i d pi

In this manuscript, the analysis of the fractions collected during the SPFFF is explored in terms
of sample z-av g and intensity size distribution 1 . Count rate information is provided in the tables
presenting the z-av g value of each sample fraction. We intended to correlate the count rate data with
the amount of particles contained in each fraction, but the results exhibited considerable instability
and were beyond exploration. Nevertheless, this may provide information for comparison if the
experiments presented in this manuscript are replicated in the future.

4.2.2 Calibration curves of the online fluorescence spectroscopy (OFS) system


The fluorescence intensity signal at a given wavelength (I(λemi )) presents a linear relation to the
fluorescent sample concentration (C) under the conditions of low light absorption [86]. The calib-
ration curves of I(C) at a given λemi allow converting the light intensity into concentration. This
conversion is important, for instance, to verify the mass balance during the experiments.

For the determination of the calibration curves, 24 samples were prepared at the different con-
centrations for the Latex with 100 nm (L100) dispersed in ultra-pure water.

The followed procedure is presented below:

1. Connection of a reservoir to the MFCS-EZ and the fluorescence cell (FC) for continuous injec-
tion. The direct injection in FC using a manual syringe induces bubbles and signal instabilities.

2. Insertion of ultra-pure water in the reservoir and measurement of the reference spectrum (spec-
trum of the deuterium light source).

3. Insertion of the sample with smaller concentration for continuous sample injection. Registra-
tion of the desired number of spectra I vs λ once the signal I vs t is constant.

4. Insertion of water in the reservoir to clean the system from the previous solution. A reference
spectrum is saved after water circulation. Between each sample with different concentrations,
there is water circulation and registration of a reference spectrum.

5. Repetition of points 3 and 4 until all the concentrations are measured.

6. If there is a spectrum instability (strange signal or the circulation of water does not bring the
spectra to the reference spectra), cell cleaning is required, and the calibration curve meas-
urement can restart once the signal overlaps the reference spectrum. Otherwise, the points
1
The auto-correlation function is converted into intensity size distribution, from which are obtained the number (higher
sensibility for smaller particles) and volume (higher sensibility for larger particles) size distributions. The choice of ana-
lyzing the intensity size distribution relies on being closer to the raw data, having fewer calculation instabilities.

87
measured to build the calibration curves do not have the same starting conditions, resulting in
non-linear calibration curves.

7. Treatment of the I vs λ data saved with the OceanView software. According to the sample
type, obtain the intensity data at the desired λ (notice that each particle size is marked with
different fluorochromes, then the λemi is different for each particle: L50 - 520 nm, L100 - 407
nm, L200 - 546 nm, L500 - 546 nm, and SiO2 - 510 nm) and subtract the Intensity signal of
the reference spectra to the saved sample data.

Figure 4.20 shows the L100 calibration curve determined at λ=407 nm within a concentration
range of 1.194×10−4 −2.068×10−1 g/L. As observed, the increase of sample concentration starts to
flatten the emission intensity signal because higher sample quantity increases the light absorption. In
order to respect the linearity between I and C, the conversion of the intensity signal to concentration
is performed by dividing the calibration curve into four regions with different concentration ranges.
Figure 4.21 presents the calibration curves in the different concentration regions, and Expression 4.9
shows the calibration curves applied for the mass balance of L100 samples.

Figure 4.20: Calibration curve of Latex L100 measured at λ = 407 nm within a concentration range
of 1.194 × 10−4 − 2.068 × 10−1 g/L.

88
(a) Concentration range of (b) Concentration range of 4.151 × 10−2 − 9.177 × 10−2
9.177 × 10−2 − 2.068 × 10−1 g/L. g/L.

(c) Concentration range of (d) Concentration range of 1.194 × 10−4 − 9.975 × 10−3
9.975 × 10−3 − 4.151 × 10−2 g/L. g/L.

Figure 4.21: Regions of the latex L100 calibration curve measured at λ = 407 nm.


I−62.980
194361.518 , 0 < I(c.u.) < 2000







 I−548.438
 150951.097 , 2000 ≤ I(c.u.) < 7000


Cλ=407nm (g/L) = (4.9)
 I−2607.956
102813.404 , 7000 ≤ I(c.u.) < 12000









 I−7614.706
51239.701 , 12000 ≤ I(c.u.) ≤ 18000

Figure 4.22-(a) presents the reference emission spectrum and the L100 spectra at different con-
centrations. The Figure 4.22-(b) presents the L100 spectrum in the concentration range of an SPFFF
elution peak (when ∼ O(1) g/L is injected ).

89
(a) Reference and L100 spectra at different (b) Reference spectra and L100 spectra at a
concentrations. concentration equivalent to the detected signal of an
SPFFF experiment.

Figure 4.22: Comparison of the L100 spectra at different concentrations with the reference spectra.

The calibration curves of L50, L200, L500, and S100 are not presented since they are not used
in this manuscript.

The fluorochromes presented in the latex (L50, L100, L200, and L500) samples have a maximum
emission wavelength (λemi ) when exposed to the maximum excitation wavelength (λex c ). The con-
figuration of the used online fluorescence spectroscopy system transports light at λ=200-800 nm,
which results in excitation using a broad λ and not only at λe x c (the incident light is not mono-
chromatic). Thus, the emission signal does not present a narrow distribution around the maximum
λemi , instead emission at other λ can be observed. In Figure H.1 of the Annex H.0.1 we present the
emission spectra of the latex samples when excited only with the maximum excitation wavelength
(λe x c ). Figure 4.23 presents the reference spectrum and the L50, L100, and L500 (L200 has the
same fluorochrome as L500) spectra in the concentration range of the samples detected during an
SPFFF experiment (∼ O(10) mg/L when ∼ O(1) g/L is injected). From this figure, we observe that,
although the latex solution have a signal emission at a common range of wavelengths, they present
distinct spectra and a characteristic signal in the region of 460 − 650 nm, which can be used to
differentiate them.

90
(a) (b)

Figure 4.23: Spectra of the different latex samples (L50, L100, and L500), and comparison with
the reference spectra in the wavelengths range of (a) 200-800 nm and (b) 400-650 nm.

4.3 Solvent properties


The solvent, also known among FFF community as a carrier fluid, is responsible for the sample
transport towards (accumulation step) and along the membrane (elution step). Three different
solvents were used in this manuscript, depending on the type of the experiment. The verification
of the system hydrodynamics presented in Section 4.4 was performed using the L100 sample and
ultra-pure water (18.2 M Ωcm purified with an ELGA PURELAB purification system) as solvent. For
the experiments containing an accumulation and an elution step (experiments SPFFF presented in
Chapter 5 and 6) Sodium Dodecyl Sulfate (SDS) was used. The SDS was chosen as it is among the
list of surfactants suggested to stabilize Latex particles for FFF experiments (table 12.2 - page 194
of [16]) and was used in the study of Kato et al. [91] to allow the separation of silica particles
with 50 and 100 nm. In Section 6.2.2 the impact of different solvents (ultra-pure water, SDS 1g/L,
SDS 0.1 g/L, and PBS 1X) on the separation of a mixture containing L100 and L500 nanoparticles is
discussed. In this section, the properties of the SDS and PBS solutions are described.

4.3.1 Sodium Dodecyl Sulfate (SDS)


Sodium Dodecyl Sulfate (SDS) is an anionic surfactant with the chemical formula C12 H25 NaSO4 . The
SDS structure (presented bellow) consists of a sulfate polar head and a hydrophobic tail composed
of hydrocarbons. The critical micellar concentration of SDS in water at 25◦ C is 7.2-8.1 mM [92]
giving 2.08-2.34 g/L (molar mass MW =288.37 g/mol).

The SDS used in the experiments of this manuscript is from Si g ma (ref. L4509-500G). Annex J
discusses the impact of SDS (aqueous solution with 1 g/L and an ionic strength (Is ) of 3.47×10−3 M )
on the light signal detected in the fluorescence cell, in the tests carried with L100 particles. The SDS
molecules, similarly to water, are not fluorescent, and their circulation in the FC cell gives a spectrum
equal to the deuterium reference spectrum. Nevertheless, it can affect the fluorescence of fluoro-
chromes. The article [93] studied the effect of SDS on the fluorescence properties of two fluorescent
molecular sensors. The article [86] mentions the complex effect of the solvent physicochemical prop-
erties (viscosity, polarity, pH, and hydrogen bonding) on the fluorochromes fluorescence. This effect

91
can result, among others, in a change of the signal intensity and a shift in the emission wavelengths
(λemi ).

S
O O⊖ Na⊕
O
Sodium dodecyl sulfate

A study was performed to investigate the influence of the SDS solution on the signal intensity
emitted by the L100 fluorochrome. The objective was to understand if the signal intensity is similar
when ultra-pure water is used as a solvent and to verify if the signal is stable in time (the use of
calibration curves to perform mass balances requires their validity on each experiment, thus the effect
of SDS has to be stable). A total of nine L100 samples were prepared, having the concentrations
of 0.01, 0.1, and 1 g/L dispersed into SDS solution with 0.01, 0.1, and 1 g/L (Table 4.5). The
calibration curves of the nine samples were determined in the OFS system on two different days (the
same day that the samples were prepared and eight days after). The zeta potential (ζ), the size (d p ),
and the pH of each sample were also measured, at day 0 and day 8, being the results shown in Table
4.6. Figure 4.24 presents the average size values measured on day one and day eight.

Observing the results, it turns out that the samples pH, zeta potential (ζ), and size (d p ) remained
stable during the eight days of the stability test. We noticed that the d p of the samples is higher when
the L100 concentration is smaller (green diamond points), which can be justified by the limitation
of DLS to analyze diluted samples (higher particles skew the smaller particles scattered light). The
test shows that the ζ and d p are not considerably affected by the presence of SDS.
Table 4.5: Samples composition for the investigation of the SDS effect in the fluorescence signal.

Conc. L100
1 g/L 0.1 g/L 0.01 g/L
Conc. SDS
1 g/L a b c
0.1 g/L d e f
0.01 g/L g h i

92
Table 4.6: Measurement of the sample properties (size (d p ), zeta potential (ζ), and pH) at the days
0 and 8 after preparation.

Day 0 Day 8
Size Zeta potential Size Zeta potential
Samples pH pH
d p (nm) ζ (mV ) d p (nm) ζ (mV )
◦ ◦
a 3.63 (19.7 C) 81.3 -42.0 4.26 (20.7 C) 81.2 -43.6
b 4.60 (20.0◦ C) 1471.1 -42.5 5.51 (20.2 ◦ C) 174.3 -41.5
◦ ◦
c 5.19 (19.9 C) 242.2 -26.1 5.75 (20.5 C) 255.1 -40.5
◦ ◦
d 3.75 (19.9 C) 82.3 -43.8 3.90 (20.3 C) 80.8 -43.4
◦ ◦
e 4.64 (20.0 C) 146.8 -43.3 5.27 (20.0 C) 143.7 -24.9
f 5.22 (19.9◦ C) 241.2 -41.5 5.42 (19.8 ◦ C) 232.6 -27.4
◦ ◦
g 3.66 (24.8 C) 82.05 -39.5 4.04 (22.0 C) 85.55 -43.1
◦ ◦
h 4.75 (20.1 C) 119.4 -37.8 4.89 (19.8 C) 120.0 -31.8
◦ ◦
i 5.49 (21.2 C) 234.4 -22.0 5.37 (19.8 C) 220.9 -33.1

Figure 4.24: Average value of the the samples size (d p ) measured for day 1 and day 8.

The three points of the calibration curve (for the OFS system) corresponding to three differ-
ent concentrations of the L100 sample are presented in Figure 4.25. The points at different SDS
concentrations are in agreement with the points measured with water, showing that the SDS, in-
dependently of its distribution around the particles, does not interfere with the optical properties
of the fluorochromes. Therefore, the discrepancies of the magnitude of the signal observed when
the SDS (1 g/L) was used as a carrier fluid can not be justified by the influence of SDS in sample
fluorescence. A possible explanation can be the formation of small bubbles (SDS solutions present
foam) that reflect the light resulting in a higher peak magnitude.

93
Figure 4.25: Scheme of the imposed and measured flow-rate conditions.

Since the interference of SDS in the detected signals is not mastered, the mass balance is presen-
ted for the experiments performed to verify the hydrodynamics of the prototype (carrier fluid is ultra-
pure water) but is not applied to quantify the separation performance of the experiments where the
SDS with 1 g/L was used. In this case, the fractograms are presented in intensity (I) normalized by
the maximum intensity of the signal (I max ).

4.3.2 Phosphate buffer saline solution (PBS)


The PBS solution 1x was prepared from the PBS 10x purchased from Fisher BioReagents (ref. BP399-
4). The PBS 1x composition is 11.9 mM of phosphates (KH2 PO4 and Na2 HPO4 ), 137 mM of sodium
chloride (NaCl) and 2.7 mM of potassium chloride (KCl), having an ionic strength (Is ) of 0.4042 M
and a pH 0f 7.4.

4.4 Verification of the system hydrodynamics


The system hydrodynamics was verified by analyzing its residence time distribution (which gives
information on the mean residence time ( t̄ r ) and sample amount collected at the channel outlet
(mo )) and the flow regimes (defined according to the Reynolds (Re) and axial Peclet (Pez ) numbers).
To perform the required calculations, it is necessary to have the flow and concentration data of the
experiments. The flow information is saved during the experiments run time using the software
All-in-One (AIO). The fluorescence emission signal is obtained with the OceanView software and
converted to a concentration signal with the calibration curves. Both flow and fluorescence data are
treated using a python script (version 2.7).

4.4.1 Definitions for the experiments


Figure 4.26 illustrates the experimental imposed and measured operating conditions. The inlet (Pi )
and radial pressure (Pr ) are imposed with the microfluidic flow controller (MFCS-EZ), the outlet (Po )
when opened is at atmospheric pressure, and the inlet (Q i ) and outlet (Q o ) flow-rates are measured
2
Ionic strength was calculated considering a concentration of 5.95 mM for the monobasic and dibasic phosphates.

94
with the flow-meters RD1 and RD2. The pressure values are Gauge pressure (relative pressure),
which means that they are zero-referenced against atmospheric pressure. The channel length is
defined as L (z = L is the channel outlet), and the channel radius is R f (2R f = I D is the membrane
diameter).

Figure 4.26: Flow conditions in the MHF5 module.

The membrane cross-section (S) and membrane surface area (Am ) are obtained according to
Equations 4.10 and 4.11, respectively.

S = πR2f (4.10)

Am = 2πR f L (4.11)

The permeation flow-rate (Q r ) is calculated from the difference between the measured inlet
(Q i ) and outlet (Q o ) flow-rates, Equation 4.12.

Qr = Qi − Qo (4.12)

The average transmembrane pressure (T M P) is given by Equation 4.13.

P̄i + Po
TMP = − P¯r (4.13)
2

The arithmetic average of the axial velocity at the inlet (Ūzi ) and at the outlet (Ūzo ) are calcu-
lated by dividing the average inlet (Q̄ i ) or outlet (Q̄ o ) flow-rate by the channel cross sections (S),
Equation 4.14 and 4.15, respectively.

Q̄ i
Ūzi = (4.14)
S

95
Q̄ o
Ūzo = (4.15)
S

The arithmetic average of the axial velocity (Ūz ) is obtained by averaging mean inlet and outlet
axial velocities, Equation 4.16.
Ūzi + Ūzo
Ūz = (4.16)
2

The average of the radial velocity (Ū r ) is calculated by dividing the average permeate flow-rate
(Q̄ r ) by the membrane surface (Am ), Equation 4.17.

Q̄ r
Ū r = (4.17)
Am

The Peclet number, as previously discussed in Chapter 3, is a dimensionless quantity that rep-
resents the ratio between the transport by convection and the transport by diffusion. The Peclet
number in the axial z-direction (Pez ) is calculated according to equation 4.18 and the Peclet in the
radial r-direction (Pe r ) was previously given in Equation 4.4.

Ūz R f
Pez = (4.18)
Do

The Reynolds number (Re) represents the ratio between inertia and viscous forces, describing
the flow regime, Equation 4.19.

ρ Ūz R f
Re = (4.19)
µ

The collected sample mass (mo ) is estimated from the concentration signal (C) converted from
the intensity signal (I), according to Equation 4.20.

Z
mo = Qo C d t (4.20)

Note that if we knew the local concentration distribution C(r, θ ) at a given axial coordinate z,
then we could have calculated the mass from:

Z Z 
mo = U(r, θ )C(r, θ )dS d t (4.21)
T S

The mass recovery (M r ), that should tend to 100% in ideal measurement conditions, is defined

96
by equation 4.22.

M r (%) = mo /min j × 100 (4.22)

The residence time is obtained from the residence time distribution function E(t), which is a
probability density function that quantifies the external age distribution of the mixture in the channel.
The experimental determination of E(t) for the sample injection with a Dirac-pulse3 perturbation is
obtained from Equation 4.23:

C(t)
E(t) = R ∞ (4.23)
0
C(t)d t

The mean residence time ( t¯r ) corresponds to the weighted average residence times of the
volume fractions at the channel outlet, and is calculated according to Equation 4.24.

Z ∞
t¯r = t.E(t)d t (4.24)
0

4.4.2 Measurement of the signal at three different parts of the prototype


The aim of this section is to verify the hydrodynamics of the system at different parts of the prototype
(detection of major anomalies) and validate the measurement method (OFS) under flowing condi-
tions. For this, a sample was injected with a finite step perturbation and the signal corresponding to
the particle transport by the flow is analyzed in different sections of the prototype: after the injection
valve (LV), at the fiber inlet (MHF5) inlet and at the MHF5 outlet.

The experiments were performed, using the M1 module, and the concentration signals were
measured in three different circuits, as illustrated in Figure 4.27. For that, the FC cell was plugged
at the desired position in the circuits4 . In both experiments, the flow-rate conditions were O(∼ 150)
µL/min, being injected O(∼ 30) µL of L100 at 0.237 g/L (7.6µg). The circuit volumes are detailed
below:

1. The determination of the signal after LV was realized by placing the FC cell after tube 4, giving
a circuit volume of 74.04 µL (V3 + VLV + V4 + VF C /2 = 32.23 + 0.66 + 26.15 + 30/2).

2. The measurement of the input signal (RD1) was realized by placing the FC cell after tube 5
(and replacing tube 5 by tube 6 because the first was too small to allow the connection of RD1
to FC), resulting in a circuit volume of 210.39 µL (V3 + VLV + V4 + VRD1 + V6 + VF C /2 = 32.23
+ 0.66 + 26.15 + 127.23 + 9.12 + 30/2).
3
Actually in our experiments, the signal is not a Dirac-pulse as the injection takes ∼ 3 minutes – it rather has the form
of a finite time step.
4
It was necessary to change some of the connections to properly connect the FC cell to some equipment, namely, to the
RD1 flow-meter

97
3. The output signal measurement, called hereafter as RTD experiment, was performed by placing
the FC cell after tube 6 (original place for SPFFF experiments), having a circuit volume of
633.32 µL (V3 + VLV + V4 + VRD1 + V5 + VM H F 5 + V6 + VF C /2 = 32.23 + 0.66 + 26.15 +
127.23 + 7.30 + 415.63 + 9.12 +30/2).

Figure 4.27: Part of the PFD diagram for the representation of the signal measurement in three
different regions of the prototype.

The hydrodynamic conditions and the mass recovery of the different experiments are summar-
ized in Table 4.7. The Re number is low, indicating that the flow is laminar. The Peclet number
Pe = Uz a/D0 is large indicating the particles are weakly diffusive (note that in this section, a cor-
responds to the capillary radius rather than the particle radius to match the nomenclature in Figure
4.28, whereas in other section of the manuscript, a represents the particle radius). The Peclet num-
ber and the dimensionless length of the tubes corresponding to the different parts of the circuit are
added to Figure 4.28 (marked by LV, RD1, and RTD points). According to the Taylor dispersion
theory, which describes the transport of a tracer slug in a circular capillary tube by a fluid flow [53],
our experiments fall in the pure convection limit, i.e. the diffusion weakly affects particle transport
in the different parts of the experimental setup. This corresponds to a convection time scale much
smaller than the radial and axial diffusion time scales. Thus, the injected sample distortion is mainly
induced by the transport of the particles by the parabolic velocity profile of the carrier fluid (laminar
flow).
Table 4.7: Hydrodynamic conditions and mass recovery of the signal after LV, the input signal (RD1),
and the output signal (RTD) experiments

Q̄ i Q̄ o t̄ r τ mo
Experiment Re Pe M r (%)
(µL/min) (µL/min) (min) (min) (µg)
LV 159.580 161.862 1.533 1.755 × 105 0.983 0.458 11 147
5
RD1 153.128 153.123 1.612 1.845 × 10 1.489 1.380 9.6 126
5
RTD 155.608 153.808 1.621 1.855 × 10 4.715 4.123 9.2 121

98
Figure 4.28: Dispersion regime for a tracer in a capillary tube. From [53]

Figure 4.29 illustrates the ideal input (RD1) and output signals (RTD) for the MHF5 channel.
The ideal input signal has a step shape with the duration t in j , calculated from Equation 4.25 (which
corresponds to the geometrical residence time (τ) of the RD1 experiments). The ideal output signal
has the shape represented by the green curve (according to the E(t) model of laminar Newtonian
pipe flow [85]) and present the geometric retention time τ obtained from Equation 4.26.

Figure 4.29: Representation of the ideal input and output signals. The figure is merely illustrative
and does not have the correct dimensions (the peaks have to present the same area under the curves).

The injection time (t in j ) is given by:

Vinput
t in j = (4.25)
Qi
where Vinput = V3 + VLV + V4 + VRD1 + V6 + VF C /2 = 210.39µL, and Q i is the flow-rate measured in
RD1.

The geometric residence time (τ) is obtained from:

Veq
τ= (4.26)
Qo

99
where Veq = V3 + VLV + V4 + VRD1 + V5 + VM H F 5 + V6 + VF C /2 = 633.32 µL, and Q o is the flow-rate
measured in RD2.

The fluorescence signals measured after the valve LV, at the channel inlet (RD1) and at the
channel outlet (RTD) are presented in Figure 4.30. The peaks have the shape expected from the
Taylor dispersion theory, i.e. the transport of a tracer by a laminar pipe flow, in the limit of weak
tracer diffusion. Since all signals are in the pure convection limit, all peaks are expected to be
distorted.

Figure 4.30: L100 signals detected in three regions of the prototype.

There are two consistency tests to diagnose the system signal that consists of comparing the
injected (min j ) to the collected (mo ) sample mass, and the mean residence time (t r ) with the geo-
metric residence time (τ). The results presented in Table 4.7 show that the t̄ r > τ in all the detected
signals. In the LV signal, the sample takes ∼ 3.5 minutes to completely leave the circuit. In the RD1
signal, the sample is completely injected after ∼ 5 minutes (3.6 times longer than t in j ), not having
the ideal step shape illustrated in Figure 4.29 (as expected by Taylor dispersion since the RD1 peak
is in the pure convection limit). The RTD signal takes 2.9×τ to completely evacuate the particles
from the channel, where we recall that τ refers to the geometric residence time in the system. In
addition, the LV and RD1 signals have a second small peak maximum, revealing the possibility of
existing internal recirculation and stagnant regions in the equipment. Nevertheless, the time distri-
butions in Figure 4.30 give an idea about the time spent by the sample in different parts of the device
and the mean residence time t̄ r , in every part of the device, can be reasonably well estimated by the
geometric residence time τ. The largest difference between t̄ r and τ corresponds to the flow in the
valve, suggesting the presence of flow perturbation in this part.

Concerning the mass balance, Equation 4.20 allows the calculation of the mass through the
integration of the output signal (intensity I converted with the L100 calibration curve presented in

100
Equation 4.9). The mass calculation of the three experiments presents a sample recovery higher than
the injected quantity, as presented in Table 4.7. This may be explained by the signal dispersion in the
FC cell. During the measurement of the calibration curves, the sample is continuously injected in the
FC cell, resulting in a uniform distribution of sample along the FC cell cross-section, as illustrated
in Figure 4.31 (left panel). On the other hand, during a discontinuous experiment (injection of
a finite step), there is axial dispersion (explained by Taylor dispersion) that causes a non-uniform
distribution of the sample along the FC cross-section (right panel of Figure 4.31), which could lead
to producing an over-estimated mass recovery (comparing Equation 4.20 and 4.21, the access to the
local concentration and velocity values in the correct cross-section (dS) containing the signal would
permit an accurate mass balance, but, we can only measure average values). Therefore, the OFS
system over-estimates the calculated mass.

Figure 4.31: Illustration of the sample distribution across the FC cell during the measurement of
the calibration curves and the detection of the elution peaks of discontinuous experiments (LV, RD1,
RTD, and SPFFF).

From the above comparisons we mainly conclude that the fluorescence measurements make
sense and that there are no significant perturbations in the system that could impact the separation
experiments. For a complete hydrodynamic characterization, the LV, RD1, and RTD experiments
must be performed using inert tracers (our particles are a dispersion of colloidal polystyrene latex that
have particle-particle and particle-equipment electrostatic interactions due to their negative charge)
and the system must be described with models that include dead-volumes, internal recirculation, and
stagnant regions. We modeled the signals using the residence time distribution models for laminar
flow, dispersive plug flow, and cascade reactors (results presented in Annex I), however, these models
are simple, not providing relevant conclusions.

101
4.5 Conclusion
This project aims the development, from scratch, of a prototype to separate colloids using the HF5
mechanism to integrate it in production lines. The prototype was developed to perform the ac-
cumulation and elution steps necessary for the size-based separation of colloidal dispersions. To
conduct the experiments, micro/milli-fluidic equipment were controlled using the Microfluidic Auto-
mation Tool (MAT) software. An online fluorescence spectroscopy (OFS) system was integrated at
the channel outlet for the real-time detection of the elution peaks. The acquisition parameters and
the recording of the experimental results are controlled using the OceanView software.

The verification of the system hydrodynamics revealed a difficulty to perform the system mass
balance with our online detection method and a mean residence time superior to the geometrical
retention time (suggesting that there is equipment inducing perturbation in the flow). Although, no
major problem was detected that could affect the separation performance.

There are two challenges for the determination of sample recovery using OFS, the over-estimation
of the value (that is reproducible to all experiments when ultra-pure water is used as a carrier fluid)
and the non-reproducibility of the peak maximum when an SDS solution 1 g/L is used as the carrier
fluid. Due to the challenges in calculating the mass recovery with the OFS, the experimental results
presented in this manuscript are explored in terms of peaks retention time and granulometry analysis
by DLS. If no other information is given, the peaks detected with the OFS during the elution step
are presented in intensity (I) values (having the background spectra eliminated before the experi-
ment and a baseline subtraction during data treatment) normalized by the maximum intensity value
(I max ).

Although the OFS system does not permit accurate quantification of the separation performance
in terms of instantaneous concentration value, it captures the overall concentration distribution in
time. At this stage of prototype development, the OFS system allowed us to have a fast interpretation
of the experimental results and supported us in understanding the impact of operating conditions
on the response of the system. The prototype was designed considering future developments for
optimization of the sequential operation, adaptation for continuous functioning, and application
of an industrially valid scale-up strategy (multiple hollow fiber modules in parallel). The chosen
equipment will allow the development of a "Smart-Process", which is automatically regulated by the
signal detected at the fiber outlet (towards process digitalization).

102
From exploration of operating conditions to
method prescription
5
This chapter is divided into four sections, containing in Section 5.1 the exploration of the accumula-
tion and elution operating conditions, in Section 5.2 the effect of the particles size (diffusion) in the
retention time, and in Section 5.3 the prescription of a methodology to perform a separation using
our prototype.

5.1 Exploration of the operating conditions


The prototype operating steps comprehend an accumulation, elution, and backwash stages (Figure
5.1). The accumulation is responsible for the sample transport towards the membrane, which results
in the arrangement of the particles in layers with distinct location along the fiber radius (according
to the particle size as discussed in Chapter 2). The elution stage occurs after accumulation, for the
particles evacuation from the channel. The particles residence time depends on the radial position
of their layer. In the present section, we discuss the studies performed to understand the impact of
the operating conditions of the accumulation and elution steps in the SPFFF prototype response. The
exploration of the accumulation operating conditions was carried by investigating the variation of
the permeate flow-rates (Q r a ) and step duration (t a ) while performing an elution without retention
Q̄ +Q̄
(Q r e ∼ 0 µL/min). The best accumulation conditions were used to study the axial (Q¯z = i o ) and
2
radial permeate (Q r e ) flow-rates during elution. In this way, the effect of each experimental step
(accumulation and elution) is decoupled, providing an understanding of their effect in the particles
fractionation. After elution, there is a backwash step, where the transmembrane pressure (T M P) is
reduced, and the permeate flow enters the membrane. The objective of this last step is the release
of particles that remained aggregated or adsorbed at the membrane during the separation process.
This is essential to recover larger particles and to clean the membrane (which keeps the membrane
permeability for a longer time and allows the performance of multiple accumulation/elution/back-
wash sequences as discussed in Section 6.3.2). There is no need to explore the operating conditions
of the backwash step since it occurs after the separation, and the increase of the permeate back-flow
(Q r b ) reduces only the time required to remove the sample from the membrane surface. Figure 5.1
illustrates the operating conditions of the accumulation, elution, and backwash steps.

103
Figure 5.1: Scheme illustrating the operating conditions of the accumulation and elution steps. Elu-
tion step can comprehend a backwash stage. The duration of the accumulation step (t a ) is not
represented but is also a parameter. Q i - measured inlet flow, Pi - imposed inlet pressure, Q o - meas-
ured outlet flow, Po - imposed outlet pressure, Q r - measured radial permeate flow, Pr - imposed
radial permeate flow. The indexes a, e, and b stand for the accumulation, elution, and backwash
steps, respectively.

The experiments were performed using polystyrene latex model particles (L100) choose because
they are stable, negatively charged (negligible absorption to the equipment), have a proper emission
signal at λ=407 nm even at low concentration (∼ O(1 × 10−4 g/L)). Furthermore, such a size
allows to have a large diffusion coefficient: in Flow-FFF experiments, the Brownian diffusion is the
governing phenomenon. An L100 solution with a concentration of 0.957 g/L was prepared using
ultra-pure water (resistance of 18.2 mΩ supplied by MEDICA EDI 15/30). The solvent used as a
carrier fluid for the experiments was an aqueous solution of Sodium Dodecyl Sulfate (SDS) 0.1 %
w/v. The solvent composition was chosen in order to maximize the optimal operating conditions.
From previous experiments, we realized that the SDS allowed a better separation resolution (in
Section 6.2.2, we present the impact of the solvent composition in the separation efficiency). For
this reason, we decided to use SDS in all the experiments performed. The experiments presented in
this manuscript were all performed at room temperature.

5.1.1 Accumulation step


The accumulation conditions of the prototype, as mentioned previously, are the permeate flow-rate
(Q r a ) and accumulation time (t a ). The Q r a is responsible for the particles driving to the membrane
surface and depends on the inlet conditions (Q i ). During accumulation, the outlet is closed, resulting
in Q i =Q r a . To place the experiments inside the classic HF5 conditions, input pressures (Pi ) of 25,
50, and 100 mbar were investigated. The correspondence of Pi with the Q̄ r , and the L100 Pe r are

104
displayed in Table 5.1. The time to drive the particles to the channel walls (t t r ansp ) was calculated
according to Equation 4.5. Preliminary experiments, not presented in this manuscript, showed that
t a =2t t r ansp produces appropriate accumulation conditions for the L100 particles, i.e., it reduces the
quantity of sample leaving in the void peak (peak of unretained sample with a residence time t 0 ).
The estimated t a according to the applied Pi are also presented in Table 5.1.
Table 5.1: Operating conditions during the accumulation step of L100.

Pi ฀r
Q ttransp ta = 2 × ttransp
Per (L100)
(mbar) (µL/min) (min) (min)
25 80 183 7.5 15.1
50 140 321 4.3 8.6
100 240 550 2.5 5.0

Table 5.2 presents the experiments performed to explore the impact of the accumulation con-
ditions in the sample elution peak. The three pressure conditions were tested with different accu-
mulation times through the realization of the experiments A to G. The grey-highlighted table entries
correspond to the accumulation time estimated from t a = 2 × t t r ansp , that is assumed as the "proper"
t a . The objective of these tests is to understand the impact of the operating conditions (Q r a and t a )
in sample loss (non-retained particles leaving in the void peak), adsorption to the membrane, and
aggregation.
Table 5.2: Designation of the experiments performed to investigate the accumulation operating con-
ditions.

ta
5 min 9 min 15 min
Pi
25 mbar - A B
50 mbar C D E
100 mbar F G -

The experiments A to G were realized by injecting 33 µL of the L100 solution with 0.957 g/L
in the channel (corresponding to 31.581 µg), performing an accumulation stage with the conditions
enunciated in Table 5.2, an elution step using the conditions Pi =50 mbar and Pr =30 mbar (to have
a T M P ∼ 0 mbar and, consequently, a Q̄ r e ∼ 0 µL/min) during 60 minutes, and a backwash step
by imposing Pi =50 mbar and Pr =45 mbar during 10 minutes. During the elution and backwash
stages, the signals were detected using the online fluorescence spectroscopy system (OFS), and signal
fractions were collected every 10 minutes for granulometry analysis by DLS. The collected fractions
present low concentration (dilution factor of the injected sample is O(> 100×)), resulting in a poor
quality of the measurement (the DLS equipment reported the warning "Count rate is out of range.
The sample concentration is too low. Insignificant signal collected."). The DLS analysis of L100 with
∼ 1 g/L gives a peak at 103.5 nm, with a z-av g of 79.68 nm, a polydispersity of 0.221, and a count
rate of 156.2 kcps. The same analysis for a L100 sample diluted to be in the same concentration
range of the fractions collected during the experiments (∼ 0.01 g/L) gives two peaks, one at 176.3

105
nm and another at 4.3 µm, a z-av g of 167.4 nm, a polydispersity of 0.337, and a count rate of 25.1
kcps. This shows the limitation of the DLS analysis, as the granulometry results for the L100 sample
with a dilution factor of 100 (the sample was not subjected to any type of phenomena only a dilution
was performed) gave higher size distribution. The DLS is more sensitive to higher objects (intensity
of the scattered light is proportional to d p6 ), and the dust or few aggregates presented in diluted
samples strongly affect the measured signal [4].

The sections 5.1.1.1 to 5.1.1.3 detail the experimental results obtained for the different accu-
mulation conditions.

5.1.1.1 Inlet pressure Pi =25 mbar: experiment A and B

The experiments A and B were performed under the same pressure conditions for different accumu-
lation times t a . The measured inlet (Q i ) and outlet (Q i ) flow-rates, and the calculated permeate
flow-rate (Q r a ) applied during the accumulation, elution (without retention, i.e. Q̄ r e = 0 µL/min
during elution), and backwash stages of the experiments A and B are shown in Figure 5.2. The
calculated mean flow conditions are presented in Table 5.3.

(a) Experiment A (b) Experiment B

Figure 5.2: Flow-rate conditions during accumulation, elution, and backwash of the experiments A
and B. The accumulation is longer for experiment B, while the operating conditions for the elution
and backwash stages are similar.

Table 5.3: Hydrodynamic conditions of the experiments A and B.

Parameters Experiment A Experiment B


Accumulation Elution Backwash Accumulation Elution Backwash
Q̄ i (µL/min) 95.679 177.012 158.752 92.049 179.217 160.968
Q̄ o (µL/min) 0.741 173.100 198,284 0.537 178.741 205.568
Q̄ r (µL/min) 94.938 3.913 -39.532 91.512 0.476 -44.600
5 5 5
Pez - 3.038 × 10 3.098 × 10 - 3.106 × 10 3.180 × 105
Pe r 217.579 8.967 -90.599 209.727 1.091 -102.215
Re - 2.653 2.706 - 2.713 2.778
Duration (min) 9 60 10 15 60 10

The accumulation was performed with a Pe r a ∼ O(200), and the elution was realized under the

106
laminar (Re∼ O(2.7)) and pure convection regimes (Pez ∼ O(3×105 )), for both A and B experiments.
Figure 5.3-(a) shows the normalized intensity signals (I/I max ) detected during the elution (0 < t e <
60) and backwash (60 < t e < 70) stages of the A and B experiments. Figure 5.3-(b) compares the
outlet (Q o ) and permeate (Q r ) flow-rate conditions to make easier the visualization that the elution
and backwash steps were performed under the same hydrodynamic conditions for both experiments,
and that the permeate flow-rate during elution was negligible (Q̄ r e ∼ 0). Thus, the magnitude
differences observed between the detected particles signals at the void peak (unretained) for the
experiments A (blue curve) and B (orange curve) are caused by the operating conditions of the
accumulation step, namely, the accumulation time (t a ) that was the only varied parameter.

In experiment A, an accumulation time of 9 minutes (inferior to the estimated value) was used,
which resulted in a higher quantity of unretained particles (higher void peak at t 0 ). The use of a
t a =2×t t r ansp , in experiment B, reduced the sample quantity leaving in the void peak. For both ex-
periments A and B, the maximum of the peak elution is located at t e ∼ 15 minutes and the peak
detection ends at t e ∼ 40 minutes. The results unveil that the existence of the accumulation stage
increases the sample retention O(∼ 10) times the mean residence time ( t̄ r ) determined for the RTD
experiment, presented in Section 4.4.2. In fact, experiments A and B differ only from the RTD experi-
ment because they have the accumulation stage before the elution without permeation (particles are
not retained during elution), highlighting the importance and influence of the accumulation stage in
the increase of the particles retention time. Signals were detected during the backwash step (back-
wash peaks), showing the presence of particles adsorbed at the membrane surface, which are only
released with the imposition of a back-flow (flow entering the channel with an opposite direction to
the permeate flux).

(a) (b)

Figure 5.3: Normalized intensity signal (I) detected at λ=407 nm (a) and flow-rate conditions (b)
measured during the elution stage of the experiments A and B. Six fraction f1 to f5 were collected
during the experiments for granulometry analysis.

The DLS results of granulometry distribution for the experiments A and B are presented in Figure
5.4, and the z-av g values are given in Table 5.4. The sizes obtained for both experiments are in the
range of O(150 − 330)µm, showing that the used t a resulted in similar size distribution. Comparing
the fractions granulometry results with the size distribution obtained for the L100 solution diluted
100x (sample in the same concentration range of the f2 collected fractions, without being submitted

107
to accumulation), we verify that there is not a significant difference and that the increase of the size
of the fractions may be related to the formation of aggregates or the DLS measurement limitation
for low sample concentration.

(a) Experiment A (b) Experiment B

Figure 5.4: Size distribution obtained by DLS of the fractions collected during the experiments (a)
A, and (b) B.

Table 5.4: z-av g determined by DLS of the fractions (f1 to f5) collected during the experiments A
and B.

Experiment z-av g in nm
f1 f2 f3 f4 f5
195.1 177.4 152.9 330.4 ∗
A
(34.3 kcps) (16.9 kcps) (9.7 kcps) (38.6 kcps)
204.0 182.0 203.6 ∗ ∗
B
(58.7 kcps) (30.8 kcps) (22.4 kcps)

Data not suitable for analysis. The measurement was aborted.

5.1.1.2 Inlet pressure Pi =50 mbar: experiment C, D and E

Three other tests were performed to verify the impact of increasing the accumulation Pe r a . The
experiments were realized testing three t a (experiments C, D, and E for t a of 5, 9, and 15 min,
respectively) to understand if at a Pe r a superior to that of experiments A and B, the t a result in
different particles size distribution. The experimental flow-rate conditions of experiments C, D, and E
are given in Figure 5.5 for the accumulation, elution, and backwash stages. The mean flow conditions
are summarized in Table 5.5.

The detected intensity signals (normalized by I max ) and flow-rate conditions measured during
the elution stage of the three experiments are shown in Figure 5.6. The analysis of Figure 5.6-(a)
shows that the increase of the accumulation time reduces the quantity of particles leaving in the void
peak (unretained sample), as observed for the experiments with Pe r a ∼ O(200) (experiments A and
B). In experiments C and E, the sample takes ∼40 minutes to leave the channel, while in experiment
D, the peak detection ceases at ∼30 minutes. The system works with a fixed-pressure operation,
and it can happen that the same pressure conditions do not reproduce the same flow conditions (a
small variation in the system resistance, as the increase of membrane fouling, can change the flow

108
conditions). The Figure 5.6-(b) and the Q̄ r values presented in Table 5.5 unveils the existence of
negative Q r e values in the experiments D and E. For experiment D, the negative Q̄ r e = −7.4 µL/min
results in a back-flow during the entire elution, which reduces the sample retention time and causes
a premature elution. If the Q r e is close to 0 µL/min, the peak retention is ∼40 minutes despite t a ,
being in agreement with the experiments A and B. Thus, the accumulation time has a consequence
in the reduction of the non-retained sample quantity. Once the proper conditions are applied, the
increase of Pe r a or t a does not affect the sample retention, resulting in a higher experimental time.

(a) Experiment C (b) Experiment D

(c) Experiment E

Figure 5.5: Flow-rate conditions during accumulation, elution, and backwash of the experiments C,
D, and E. The accumulation time (t a ) is the varied parameter.

Table 5.5: Average flow conditions of the experiments C, D, and E.

Parameters Experiment C Experiment D Experiment E


Accumulation Elution Backwash Accumulation Elution Backwash Accumulation Elution Backwash
Q̄ i (µL/min) 144.141 178.301 155.819 136.656 166.009 147.741 145.211 170.878 155.522
Q̄ o (µL/min) 1.539 177.983 194.218 0.981 173.450 195.339 0.537 174.850 203.253
Q̄ r (µL/min) 142.592 0.318 -38.398 135.675 -7.440 -47.548 144.674 -3.973 -47.731
Pez - 3.091 × 105 3.037 × 105 - 2.945 × 105 2.977 × 105 - 3.0 × 105 3.113 × 105
Pe r 326.793 0.728 -88.001 310.940 -17.052 -108.970 331.564 -9.105 -109.390
Re - 2.70 2.653 - 2.573 2.601 - 2.620 2.719
Duration (min) 5 60 10 9 60 10 15 60 10

109
(a) (b)

Figure 5.6: Normalized Intensity signal (I) detected at λ=407 nm (a) and flow-rate conditions (b)
measured during the elution stage of the experiments C, D, and E. Six fraction f1 to f5 were collected
during the experiments for granulometry analysis.

The size distribution of the fractions collected during the experiments C-E are presented in
Figure 5.7 and the z-av g are summarized in Table 5.6. The range of measured z-av g values is 140-
580 nm, being in the same order as those obtained for experiments A and B. The f4 and f5 fractions
suggest sample aggregation.

(a) Experiments C (b) Experiment D

(c) Experiment E

Figure 5.7: Size distribution obtained by DLS of the fractions collected during the experiments (a)
C, (b) D, and (c) E.

110
Table 5.6: z-av g determined by DLS of the fractions (f1 to f5) collected during the experiments C,
D, and E.

Experiment z-av g in nm
f1 f2 f3 f4 f5
210.2 185.5 143.0 237.4 ∗
C
(49.9 kcps) (35.5 kcps) (5.3 kcps) (24.8 kcps)
202.6 184.0 225.8 474.1 580.7
D
(59.9 kcps) (26.3 kcps) (60.1 kcps) (16.9 kcps) (48.5 kcps)
192.5 190.1 223.7 386.8 ∗
E
(99.6 kcps) (51.1 kcps) (31.9 kcps) (33.6 kcps)

Data not suitable for analysis. The measurement was aborted.

5.1.1.3 Inlet pressure Pi =100 mbar: experiment F and G

In the experiments F and G, the Pe r a value was increased to O(∼ 500). The objective is to understand
the maximum Pe r a that can be used during the accumulation step once the higher the flow-rate, the
shorter the accumulation time has to be. The flow-rate conditions applied during experiments F and
G are presented in Figure 5.8, and the mean values are summarized in Table 5.7.

(a) Experiment F (b) Experiment G

Figure 5.8: Flow-rate conditions during accumulation, elution, and backwash of the experiments F
and G. The experiments were realized under the same elution and backwash conditions, varying
the accumulation time t a .

Table 5.7: Average flow conditions of the experiments F and G.

Parameters Experiment F Experiment G


Accumulation Elution Backwash Accumulation Elution Backwash
Q̄ i (µL/min) 235.755 173.747 160.484 246.572 179.848 161.839
Q̄ o (µL/min) 1.481 173.930 199.806 0.968 177.396 202.307
Q̄ r (µL/min) 234.275 -0.183 -39.322 245.603 2.452 -40.468
5 5 5
Pez - 3.016 × 10 3.126 × 10 - 3.10 × 10 3.159 × 105
Pe r 536,910 -0.419 -90.117 562.873 5.619 -92.745
Re - 2.635 2.731 - 2.707 2.760
Duration (min) 5 60 10 9 60 10

111
The normalized intensity signal (I) detected during the sample evacuation from the channel,
and the flow-rate conditions, are presented in Figure 5.9. The analysis of Figure 5.9-(a) confirms
(as expected from the experiments A-E) that the use of an accumulation time higher than t a =2 t r ansp
does not impact the sample retention time during elution. The comparison of the backwash peaks
magnitude for the experiments with Pi =100 mbar (Pe r a ∼ 500) and the experiments with Pi =25
mbar (Pe r a ∼ 200) and Pi =50 mbar (Pe r a ∼ 300) suggests that the use of Pi =100 mbar during the
accumulation stage increases sample adsorption to the membrane, which can not be confirmed with
a mass balance as discussed previously. Also, it presents the smaller void peaks, showing that the
higher is the Pe r a , the faster is the sample transport to the membrane surface. The peak maximum
in both experiments occurs at ∼ 10 minutes, and the sample detection stops at ∼ 40 minutes. The
total peak retention time is in accordance with the results obtained for Pi =25 and Pi =50 mbar, and
the peak maximum occurs at the same time of the experiments with Pi =50 mbar.

(a) (b)

Figure 5.9: Normalized intensity signal (I) detected at λ=407 nm (a) and flow-rate conditions (b)
measured during the elution stage of the experiments F and G.

In Figure 5.10 is presented the size distribution of the collected fractions and the Table 5.8
contains the z-av g values. The experiments F to G present a size range of 180.0-1390.0 nm, being
the higher sizes associated to the backwash peaks (fractions f5 of the experiments F and G). The
comparison of the z-av g values for the f5 fractions of experiments D, F, and G reveals that the
increase of the Pe r a results in higher size distribution, thus suggesting higher sample aggregation.

112
(a) Experiments F (b) Experiment G

Figure 5.10: Size distribution obtained by DLS of the fractions collected during the experiments (a)
F, and (b) G.

Table 5.8: z-av g determined by DLS of the fractions (f1 to f5) collected during the experiments F
and G.

Experiment z-av g in nm
f1 f2 f3 f4 f5
197.6 193.7 188.3 180.0 1209.0
F
(48.5 kcps) (45.3 kcps) (23.2 kcps) (17.1 kcps) (151.7 kcps)
237.7 180.1 190.2 179.6 1390.0
G
(105.2 kcps) (21.6 kcps) (18.6 kcps) (38.8 kcps) (89.5 kcps)

5.1.1.4 Conclusion of the tested accumulation operating conditions

The investigation of the accumulation step contributed to the understanding of the accumulation
operating conditions in the prototype performance. The main conclusions retained are listed below:

1. The use of a t a < 2 × t t r ansp results in higher quantity of unretained L100 detected in the void
peak.

2. The accumulation time of t a = 2 × t t r ansp is appropriated to reduce L100 elution in the void
peak (reduction of unretained sample).

3. The elution peaks of the tested Pe r a (controlled with the imposed inlet pressure Pi ) presented
an equal retention time until 40 minutes, corresponding to a sample ∼10 times more retained
than if no accumulation occurs (as observed in the RTD experiment discussed in Section 4.4.2).
Higher Pi conditions result in higher Q r a and consequently higher Pe r a , increasing the size dis-
tribution measured by DLS (suggesting aggregation at the membrane that could be induced by
the stronger accumulation conditions). The fractions containing the backwash peaks presented
an increase in the z-av g value with the Pe r a increase.

4. The magnitude of the backwash peaks (detected with the online fluorescence spectroscopy
OFS) obtained when using a Pi =100 mbar (Pe r a ∼ O(500)) suggests higher sample aggrega-
tion and adsorption to the membrane.

113
5. The DLS results show the tendency of the z-av g values to increase with the increase of the
fractions retention time. Also, the z-av g values of the fractions collected before the backwash
stage (f1 to f4) are in the same order for both experiments A to G. This is in accordance with
the signals measured with the online fluorescence spectroscopy system (OFS), as expected.
The elution time is related to the particle size as discussed in Section 5.2.

The accumulation is an essential step for the sample size-based separation. The nonexistence of the
accumulation step corresponds to the performance of a RTD experiment (Residence Time Distribution
presented in Section 4.4.2), which results in similar sample retention once different particle sizes are
transported at the same velocity. Further experimental and numerical studies must be performed to
understand the sample distribution along the membrane and the impact of accumulation conditions
in the sample recovery and aggregation.

Comparing with the classical Flow-FFF techniques, our peaks are broader and collected during
∼ 40 minutes, which results in a considerable sample dilution. This occurs because the particles
are spread along the entire membrane during the accumulation step, while in the classic Flow-FFF
techniques, the accumulation is confined to a short region ("focusing"). We decided to perform an
accumulation along all the fiber length because it is easier to perform and is not in conflict with pat-
ented designs. Depending on the required product specifications (concentration and polydispersity)
for which the SPFFF prototype might be used, the creation of a shorter accumulation region may be
required to reduce the dilution factor.

For the results presented in the next sections and on Chapter 6, the accumulations conditions
are Pi =25 mbar (Pe r a ∼ O(200)) and t a =15 minutes because it results in proper sample retention
and, as we could not quantify the impact of the Pe r a in sample adsorption to the membrane and
particles aggregation, we decided to use the smoother accumulation conditions. For more diffusive
particles, the same Q r a can be used, but t a has to be increased (tests need to be performed to find
the appropriate accumulation time).

5.1.2 Elution step


The elution step has two operating parameters, the axial flow-rate (Q z ) imposed by Pi and the radial
flow-rate Q r e controlled with the transmembrane pressure (T M P). Three experiments (experimental
conditions summarized in Table 5.9) were performed to investigate the impact of the elution flow-
rates in the detected outlet signal. Both experiments were realized under the same accumulation
conditions of Pi =25 mbar and t a =15 minutes. The objective of the experiments H and I is the test of
two profiles for the permeate flow (Q r e ), and the finality of the experiment J is to analyze the impact
of the axial flow-rate by testing a Q z two times higher the one used in experiment B (experiment
realized under the same accumulation conditions, and with an elution without permeation).

114
Table 5.9: Experiments to study the elution conditions.

Experiment Accumulation Elution


Pi Duration Pi Pr Duration
(mbar) (min) (mbar) (mbar) (min)
H 50 1×t e 60
I 25 15 50 0.5×t e 120
J 100 50 60

Figure 5.11 shows the operating conditions of the experiments H-J during the accumulation and
elution stages, and Table 5.10 summarizes the mean hydrodynamic conditions.

(a) Experiment H (b) Experiment I

(c) Experiment J

Figure 5.11: Flow-rate conditions used during the accumulation and elution stages of the experi-
ments H, I, and J. Experiments realized under the same accumulation conditions but different elution
flow-rates.

115
Table 5.10: Average flow conditions of the experiments H, I, and J.

Parameters Experiment H Experiment I Experiment J


Accumulation Elution Accumulation Elution Accumulation Elution
Q̄ i (µL/min) 100.174 175.071 102.092 197.074 91.180 279.902
Q̄ o (µL/min) 0.487 178.045 0.474 148.525 0.396 283.644
Q̄ r (µL/min) 99.687 -2.973 101.618 48.549 90.785 -3.742
5 5
Pez - 3.064 × 10 - 2.998 × 10 - 4.889 × 105
Pe r 228.463 -6.814 232.888 111.264 208.060 -8.576
Re - 2.676 - 2.619 - 4.271
Duration (min) 15 60 15 120 15 60

5.1.2.1 Radial flow-rate, Q̄ r e : experiments H and I

In classic Flow-FFF (AF4 and HF5) is common to program a linear or exponential decay of the
permeate flow (as mentioned in Section 2.2.3) to optimize the separation between particles with
different sizes. Such procedure allows to have a more progressive release of the particles accumu-
lated at the membrane and can ensure a better peaks resolution. To investigate the impact of the
permeate flow-rate profile during the elution step of the SPFFF experiments, the experiments H and
I were realized. In the experiment H, the elution takes 60 minutes and the permeate pressure (Pr )
is set to increase 1 mbar per minute (Pr goes form 0 to 60 mbar in 60 minutes - Pr = 1 × t e ). In
the experiment I, Pr increases 0.5 mbar per minute (Pr goes form 0 to 60 mbar in 120 minutes -
Pr = 0.5 × t e ).

Figure 5.12 shows the detected outlet signal and flow-rate conditions determined during the
elution step of experiments H, I, and B. Experiment B is used as a reference because during its elution
stage, there was no permeation flow-rate. The observation of Figure 5.12-(a) shows that experiment
B (blue line) presents the smaller sample retention, while experiment I (green line) has the higher
sample retention. The increase of the retention time is related to the radial permeate flow-rate
(Q r ) values applied during the elution step. In experiment B, the Q̄ r e ∼ 0 µL/min, resulting in axial
sample transport without additional permeation keeping the accumulated particles at the membrane
surface during elution. In the experiments H and I, the Q r e decreased linearly with the elution time,
which resulted in a linear decrease of the Pe r e . The visualization of 5.12-(a) and 5.12-(b) permits to
understand that the L100 detection starts at ∼ t e =15 minutes for the experiment H (orange line) and
at ∼ t e =30 minutes in the experiment I, corresponding to the moment where the Q̄ r e ∼ 40 µL/min.
The Q̄ r e ∼ 40 µL/min corresponds to a Pe r e ∼ O(92), and can be defined as the permeation flow-rate
limit (Q lim lim lim
r e ) to retain the L100 particles that is associated to a Pe r e limit (Pe r e ). For Q r e values
inferior to 40 µL/min, the drag force is not sufficient to keep the particles at the membrane, and
the sample diffusion becomes predominant, thus leading to the release of the accumulated layers.
In experiment I, the decrease of Q r e has a smaller slope meaning that the flow-rate takes a higher
time to arrive at the Q lim
r e value, causing higher sample retention. This threshold radial permeate
flow-rate that allows the particles release during the elution stage is further discussed in Section 5.2
for particles with different sizes.

116
(a) (b)

Figure 5.12: Normalized intensity signal I detected at λ=407 nm (a) and permeate flow-rate (Q r e )
conditions (b) measured during the elution stage of the experiments B, H, and I.

5.1.2.2 Axial flow-rate, Q̄ z : Experiment J

Figure 5.13-(a) presents the detected intensity (I) signal at the channel outlet during the elution
step of experiments B and J. The experiment B was performed with the elution conditions of Pi =50
mbar and Pr e =30 mbar (Pez ∼ O(3 × 105 )), while the elution conditions of the experiments J were
Pi =100 mbar and Pr e =50 mbar (Pez ∼ O(5×105 )), being both experiments in the laminar and pure
convection regimes. As shows Figure 5.13-(b), the experiments were realized using a Q̄ r e ∼ 0 µL/min
and different Q z values. The use of higher Q z in experiment J resulted in shorter peak retention.
Instead of 40 minutes, the signal took ∼30 minutes to leave the channel. The use of higher Q z during
elution can be used to reduce peak broadening and sample dilution. The maximum Q z conditions that
can be tested with the current prototype equipment are Pi =450 mbar and Pr e =225 mbar because
the outlet flow-meter (RD2) has a measurement limit of 1.1 mL/min. These conditions were tested
(experiments not presented in this manuscript), but the challenges associated with the equipment
affected the experimental conditions. The accumulation step is performed with Pi =25 mbar and
Pr a =0 mbar and, when the elution starts, the Pr e value rapidly increases to 225 mbar, but the Pi
value takes some time to achieve 450 mbar, resulting in a strong back-flow at the first instants of the
elution. The majority of the sample left the channel in the void peak due to the back-flow conditions.
To further test and optimize the Q z values, the system has to work with a fixed-flow operation to
guarantee the Q̄ r e ∼ 0 µL/min condition. The microfluidic controller (MFCS-EZ) can be used with a
fixed-flow operation, which was tested and presented higher flow-rates fluctuations, thus the use of a
syringe pump is preferable for more accurate control of the permeation flow-rate. Nevertheless, the
experimental protocols developed in the Microfluidic Automation Tool (MAT) software would require
adjustments for a fixed-flow operation.

117
(a) (b)

Figure 5.13: Normalized intensity signal I detected at λ=407 nm (a) and flow-rate conditions (b)
measured during the elution stage of the experiments B and J.

5.1.2.3 Conclusion of the elution operating conditions

The elution step has a double-function in the classic FFF experiments that are the particles evacuation
from the channel with the simultaneous increase of the distance between different accumulated
layers. The accumulated layers, as mentioned previously, occupy distinct radial positions in the
fiber, being located in regions with different velocities due to the laminar nature of the parabolic
profile. In the classic FFF experiments, the particle layers are formed in a confined region close
to the channel inlet, resulting in almost the entire channel free from particles. During the axial
particles transport (along with the membrane) the space between particle layers with different sizes
starts to increase because of the differences in the transport velocities associated with the laminar
flow profile, resulting in layers with different retention times. In the sequential prototype (SPFFF)
experiments, the accumulation occurs along the entire channel length, therefore, there is no space
available between the particle layers. Thus, the difference in retention time of the particle layers
does not result from the laminar flow profile. Instead, the permeate flow-rate is controlled so that
the particles are intentionally retained at the membrane and progressively released by reducing the
drag force. The parameter Q lim
r e was defined and consists of the permeate flow-rate at which the
particles can no longer be retained (the drag force is not efficient to counterbalance the Brownian
diffusion D), as illustrates Figure 5.14.

Figure 5.14: Scheme illustrating the effect of the Q lim


r e in the particles diffusion. If the permeate radial
flow-rate used during the elution step (Q r e ) surpasses the Q lim
r e , the particles remain accumulated at
the membrane. Otherwise, the particles diffusion (D) is predominant and allows their release.

118
The axial flow-rate can be optimized to reduce the mean residence time ( t̄ r ), which reduces
peak broadening and sample dilution. For that, the experimental setup needs to be improved to
perform a fixed-flow operation. The permeate flow must be controlled with a syringe pump, and the
Fluigent equipment can operate in fixed-flow mode to regulate the axial flow-rate.

5.2 Exploration of the effect of the particles size, d p


The study of the elution conditions showed that there is a permeation flow-rate limit (Q lim
r e ) at which
the drag force does not predominate over the Brownian diffusion, and the sample is not retained
at the membrane surface. Once the Q r e value is inferior to the Q lim
r e associated with particle size,
the sample diffusion predominates, re-suspending the particles and allowing their evacuation. To
understand the impact of the particle size in the Q lim
r e value, experiments for latex particles with
different sizes were performed. The experimental conditions during the accumulation and elution
stages were the same as experiment H (Table 5.10). Figure 5.15 presents in (a) the detected signals
at λ = 407 nm 1 and (b) the permeate flow-rate (Q r e ) conditions used during the elution step. The
observation of Figure 5.15-(a) and Figure 5.15-(b), shows the difference in the retention time of the
samples with different sizes, when the same Q r e conditions are applied. The elution of latex particles
with 50 nm (L50) starts at t e ∼ 5 minutes, which is associated to a Q lim lim
r e =65 µL/min (Pe r e ∼ O(75)).
The evacuation of latex with 200 nm (L200) begins at t e ∼ 25 minutes, for a Q lim
r e =10 µL/min
(Pe rlim
e ∼ O(45)). The elution of latex particles with 500 nm (L500) occurs only by the imposition
of a back-flow (backwash). Even at Q r e =0 µL/min the L500 particles remained retained, and their
evacuation starts when Q r =-5 µL/min. This shows the importance of the diffusion coefficient. The
particles present different Brownian diffusion due to their distinct size, and larger is the diffusion
(smaller particle), higher needs to be the drag force to keep the particle retained at the membrane,
thus larger is the Q lim
r e . For larger particles, the small diffusion does not permit particle evacuation,
and a back-flow has to be applied.
1
Although the particles are marked with fluorochromes that present different λemi maximum, the peaks have a higher
signal at λ = 407 nm. For this reason, we decided to represent the signals of the different latex particles at λ = 407 nm,
as it permits better visualization of the beginning of particles detection.

119
(a) (b)

Figure 5.15: (a) Normalized intensity signal (I) for L50, L100, L200 and L500, being both peaks
presented for λ=407 nm, and (b) permeate flow-rate (Q r e ) conditions measured during the elution
step. I max =1115.133, 5233.845, 2114.777, and 2761.004 c.u., for L50, L100, L200 and L500,
respectively.

The effect of particle size, and consequently, particle diffusion, is shown in Figure 5.16. A linear
regression for the positive Q lim
r e , i.e. for d p equal to 50, 100, and 200 nm, is presented in Figure
5.16-(a). Table 5.11 summarizes, the Q lim lim
r e and Pe r e values obtained for the different particle sizes.
We expected to obtain a constant Pe rlim
e once the balance between the particles diffusion and the
drag force should be proportional. Although, the difficulty obtaining the exact value of Q lim
r e gives a
noisy range of Peclet values. The Q lim
r e value given is the radial permeate flux corresponding to the
moment of the elution where we start detecting the elution peaks. As the particles are marked with
different fluorochromes, and the fluorochromes have different intensity signals, the first evacuated
particles that are marked with fluorochromes with inferior emission signals may not be immediately
detected. The result is a shift of the Q lim
r e for higher values.

(a) (b)

Figure 5.16: Radial permeate flux limit Q lim


r e as function of the (a) particle size d p and (b) particles
diffusion in diluted conditions D0 .

120
Table 5.11: Summary of the limit permeation conditions for particles evacuation during the elution
step.

dp D0 Q lim
re
Pe rlim
e
(nm) (m2 /s) (µL/min)
50 8.74 × 10−12 65 75
−12
100 4.37 × 10 40 92
200 2.18 × 10−12 10 46
−13
500 8.74 × 10 -5 -

The same experiment was performed for silica oxide nanoparticles with 100 nm (S100), being
the elution fractogram presented in Figure 5.17-(a) and the permeate flow-rate conditions in Figure
5.17-(b). From the results, we observe that the elution peak of S100 overlaps the L100, showing
that both particles have the same retention time. The S100 fluorescence signal is inferior to the
L1002 , causing higher signal noise (the I max at λ = 407 nm is 5233.845 c.u. and 615.869 c.u.,
for the L100 and S100 particles, respectively). Both organic and inorganic materials presented the
same elution time, proving that the sequential prototype (SPFFF) uses the Brownian diffusion as a
discriminant property. If the Brownian diffusion is the governing phenomenon and the membrane-
particles interactions are reduced (depending on membrane properties, A between the membrane
and the particles, particles properties, and solvent composition), particles with the same size should
have the same retention time. The forces balance for silica particles showed that gravity becomes the
predominant force for a size of O(∼ 560 nm) (Figure 3.1 - Section 3.1.2), therefore, until d p ∼ 560
nm, latex and silica should have the same response to the prototype operating conditions (both
governed by diffusion).

(a) (b)

Figure 5.17: (a) Normalized intensity signal (I)for L100, and S100, being both peaks presented for
λ=407 nm, and (b) permeate flow-rate (Q r e ) conditions measured during the elution step.

Further tests must be carried to understand the impact of other operating conditions in the
permeate flow-rate limit (Q lim lim
r e ), namely the axial flow-rate (Q̄ z ). The Q r e is associated with the
2
This is a consequence of the fluorochrome quality or the procedure used to mark the particles, which depends on the
supplier.

121
radial position that the particles occupy and, consequently, with the velocity regions reached by
the particle layers. Therefore, the change of the axial velocity profile must have an impact on the
3
Q lim
r e . Additionally, we observed that in the experiments realized during summer (performed in a
room without climatization, having outside temperatures superior to 30◦ C), the particles presented
higher diffusion (elution at higher permeate flow-rates than the determined Q lim
r e and higher void
peak magnitude), probably caused by the decrease of viscosity, which affects the parabolic velocity
profile. To use the linear regression presented in Figure 5.16 the experiments must be performed in
a climatized room to guarantee the room temperature of T = 289.15 K.

5.3 Methodology prescription for a size-based separation using SPFFF


The exploration of the operating conditions revealed, beyond their impact on the L100 peak re-
tention, a methodology for the size-based fractionation of nanoparticles governed by the Brownian
force. The strategy for the separation of a binary mixture consists of:

1. Definition of an accumulation step that properly accumulates the more diffusive objects. If the
more diffusive components are accumulated, then the less diffusive are also properly accumu-
lated. To illustrate the consequence of the accumulation conditions, the peaks presented in
Figure 5.15-(a) show that the L50 has the void peak with the higher magnitude. As the L50
particles are twice as diffusive as L100, and the accumulation conditions that were applied
were optimized for L100, they were not appropriated to accumulate L50 (the t a or Q r a must
be increased until the L50 elution in the void peak is minimized). The starting conditions of
Q̄ r a = 100 µL/min and t a = 2 × t t r ansp can be used.

2. Determination of the Q lim


r e for the species present in the mixture to separate. An initial value can
be estimated from the regression presented in Figure 5.16-(a) for a temperature of T = 289.15
K: Q lim lim
r e = −0.357d p + 80.0. To confirm the Q r e value, an experiment can be performed
using the accumulation conditions determined in the previous point and testing the elution
conditions of Q̄ i = 150 µL/min and Pr e = 1 × t e during 60 minutes of elution (conditions of
experiment H).

3. Development of the separation strategy by tuning the elution conditions. Implementation of


an elution step by specifying the permeate flowrate conditions that retain the larger particles
(Q r e > Q lim
r e of larger objects) during the time required to evacuate the smaller ones. Op-
timization may be required to find the optimal balance between peak resolution and sample
dilution.

The combination of the ideal conditions for the accumulation of the most diffusive particles
with the value of the components of the sample Q lim
r e provides the necessary information to establish
the separation protocol. Figure 5.18 proposes a methodology to define the size-based separation
strategy of colloidal systems using the SPFFF prototype. Before applying the separation methodology,
the particles have to be stable and negatively charged. For the separation of positively charged
3
Experiments not presented in this manuscript.

122
particles, a membrane with a positive zeta potential and a cationic surfactant (to cover the remaining
equipment) might be used. Another strategy is to adjust the charge of the particle through the
solvent pH (recurring to the ζ vs pH titration curves), which can induced particles aggregation once
the particles cross the isoelectric point (observed with studies performed for cerium oxide CeO2
nanoparticles - internal report to Solvay).

123
124

Figure 5.18: Methodology prescription for the size-based separation of nanoparticles using the sequential prototype SPFFF.
Application of the Separation Methodology

In this chapter we test the separation performance of the sequential prototype (SPFFF) for the frac-
tionation of a binary mixture containing latex polystyrene nanoparticles with 100 (L100) and 500
6
(L500) nm. The separation methodology prescribed in Section 5.3 is applied in Section 6.1. In Sec-
tion 6.2, the separation is optimized by testing different radial permeate flow-rates during elution
and solvent compositions. After achieving a selective separation, we tested in Section 6.3 the separ-
ation productivity by increasing the number of cycles and the mixture quantity. In the last section,
Section 6.4, we give recommendations for the separation of binary mixture with smaller size ratio.
Repeatability tests were performed at different radial permeate flow-rates during elution, which are
presented in Annex L.

6.1 Application of the methodology for the separation of L100 and


L500 latex particles (size ratio 5)
The application of the prescribed methodology for the separation of a binary mixture containing the
same concentration of latex particles with 100 nm (L100) and 500 nm (L500) consists of analyzing
the following points:

1. Optimized accumulation conditions for the more diffusive particles: The most diffusive
object of the mixture containing L100+L500 is the L100 particles. From the experiments real-
ized in Section 5.1.1, we know that the accumulation conditions of Pi = 25 mbar and t a = 15
min are adequate to accumulate L100 particles, being then used for the accumulation step of
the L100+L500 mixture.

2. Verification of the permeate flow rate in which particles diffusion prevails over the drag
force (Q lim lim
r e ): The Q r e of the L100 and L500 particles was determined in Section 5.1.2.1
and 5.2, respectively, having the values of ∼ 40µL/min and ∼ −5µL/min, respectively. This
means that using an elution permeation flow rate (Q r e ) between 0-40 µL/min (Q r e < Q lim
r e of
L100) allows L100 to be collected while retaining L500, being the latter evacuated solely if a
back-flow is applied. A value of Q̄ r e ∼ 20µL/min was initially chosen. An experiment (not
presented in this manuscript) was performed where we verified that the total evacuation of
the L100 particles using Q̄ r e ∼ 20µL/min takes 60 minutes.

3. Separation strategy: The strategy used to separate the L100+L500 mixture consists of per-
forming an accumulation stage with Pi = 25 mbar, Pr a = 0 mbar, and t a = 151 minutes
1
In fact, for the experiment presented in Figures 6.2 and 6.5, the accumulation time was 16 minutes. These experiments

125
followed by an elution stage with Pi = 50 mbar and Pr e = 25 mbar during 60 minutes,
and Pi = 50 mbar and Pr e = 40 mbar during 10 minutes (backwash). In this way, the L100
particles are collected during the first 60 minutes (0 <t e < 60 min) and the L500 are evacuated
during the last 10 minutes (60 <t e < 70 min).

The initial injected mixture, containing L100 at 0.451 g/L and L500 at 0.507 g/L, was analyzed
by DLS. The granulometry results presented two peaks at 82.6 nm and 690.7 nm, a z-av g of 357.5
nm, a count-rate of 207.2 kcps, and polydispersity of 0.532. The injection volume was 33 µL. The
mixture composition and injection quantity are the same in all the experiments presented in this
chapter unless otherwise stated in due course. The inlet (Q i ), outlet (Q o ), and radial permeate flow-
rate (Q r ) conditions measured during the accumulation, elution, and backwash steps is presented in
Figure 6.1 and the average hydrodynamic conditions are summarized in Table 6.1.
Parameters Experiment Steps
Accumulation Elution Backwash
Pi 25 50 50
Pr 0 25 40
TMP 24.734 0.183 -14.965
Q̄ i 106.910 196.368 172.592
Q̄ o 0.841 179.985 201.068
Q̄ r 106.069 16.388 -28.476
5
Pez (L100) - 3.265 ×10 3.242 ×105
Pez (L500) - 1.633 ×106 1.621 ×106
Pe r (L100) 243.088 37.546 -65.261
Pe r (L500) 1215.441 187.724 -326.306
Re - 2.852 2.832
Duration 15 60 10
Figure 6.1: Flow-rate conditions during the
accumulation, elution, and backwash stages. Table 6.1: Imposed (Pi and Pr ) and measured
average hydrodynamic conditions.

The peaks evacuation during the elution (including the backwash) step were detected at the
wavelengths of λ = 407 nm and λ = 546 nm, corresponding to the maximum emission wavelength
of the L100 and L500 nanoparticles, respectively. We decided to present the intensity signal without
normalization (division by I max ) and without subtraction of the baseline to have a better visualization
of the signals detected at the two different wavelengths. The fractogram presented in Figure 6.2,
shows two peaks at λ = 407 nm for 0 <t e < 60 min and 60 <t e < 70 min, and one peak at λ = 546
nm for 60 <t e < 70 min. As explained in Section 4.2.2, the latex particles do not have an emission
spectrum exclusively at the maximum λemi . The L100 particles have a high signal emission at λ = 407
nm and do not emit at λ = 546 nm, on the other hand, the L500 particles have emission at both
wavelengths and the signal magnitude is even more important at λ = 407 nm than at λ = 546
nm. This means that the peak observed at 0 <t e < 60 min corresponds to the L100 particles, while
the peak at 60 <t e < 70 min belongs to the L500 particles. Different peak regions were collected
were performed before defining the separation strategy, and the accumulation time was empirically decided at the moment
that the experiments were realized. Note that the use of an accumulation time of 15 or 16 minutes does not impact the
detected signals resulting only in an increase of one minute of the experimental time.

126
during the experiments (fraction f1 to f5) for granulometry analysis by DLS. Table 6.2 contains
the measured z-av g values (and the measurement count-rate), and Figure 6.3 presents the size
distribution of each fraction. The granulometry results show the tendency of the z-av g to increase
for the fractions collected at higher elution times (t e ), as expected since the higher is the particle
size longer is the retention time. The size distribution of the fractions shows that, although the
presence of peaks at large particle sizes O(∼ 4µm), there are peaks around 100 and 500 nm. The
size distribution of the f2 fraction (yellow symbols), corresponding to the maximum elution peak
of L100, shows a peak at O(∼ 100 − 200nm). The fraction f5 (purple symbols), collected during
the elution of the L500 particles, shows a peak distribution around O(∼ 500 − 600nm). The higher
z-av g and peaks observed when comparing to the initial injected mixture are related to the formation
of aggregates and the DLS limitation for the analysis of low concentrated samples. The f2 and f5
fractions were observed at the optical microscope (results not presented in this manuscript) for a
qualitative analysis of the existence of the aggregates. We verified that there is the formation of
some aggregates although the majority of the sample is dispersed, which is also reinforced by the
appearance of peaks at O(∼ 100− 200 nm) and O(∼ 500− 600 nm) in the Figure 6.3, determined by
DLS. As the collected fractions present low concentrations, the aggregates scattered signal is strongly
accounted for in the z-av g values.

Figure 6.2: Elution peaks detected at λ = 407 nm (blue line) and λ = 546 nm (orange line).

Table 6.2: Granulometry analysis (z-av g and count-rate values) of the collected fractions.

z-av g nm
f1 f2 f3 f4 f5
232.9 253.4 471.8 1035 1959
(37 kcps) (55.7 kcps) (45.5 kcps) (63.8 kcps) (107.6 kcps)

127
Figure 6.3: Size distribution obtained by DLS of the collected fractions.

The application of the separation strategy resulted in a successful size-based separation of latex
particles with 100 and 500 nm. Optimization is required to reduce the dilution of the L100 particles.
The maximum of the peak presented a concentration O(∼ 0.005g/L) (signals not represented in
concentration as discussed in Chapter 4) collected during 60 minutes, which results in a dilution
factor superior to 100.

6.2 Optimization of the separation

6.2.1 Impact of the elution permeate flow-rate (Q r e )


The sample dilution can be reduced by decreasing the elution time of the L100 peak, which can
be achieved using an elution permeate flow-rate (Q r e ) closer to 0µL/min. An experiment was per-
formed using the same accumulation conditions as the previous experiment and an elution stage
with Pi = 50 mbar and Pr e = 30 mbar during the first 40 minutes, and Pi = 50 mbar and Pr e = 45
mbar during the last 10 minutes (backwash). The accumulation, elution, and backwash flow-rates
are presented in Figure 6.4, while the average hydrodynamic data is shown in Table 6.3.

128
Parameters Experimental Steps
Accumulation Elution Backwash
Pi 25 50 50
Pr 0 30 45
TMP 21.608 -5.01122 -20.013
Q̄ i 100.585 187.835 166.699
Q̄ o 0.410 185.946 210.190
Q̄ r 100.175 1.888 -43.492
Pez (L100) - 3.243 ×105 3.270 ×105
Pez (L500) - 1.621 ×106 1.635 ×106
Pe r (L100) 229.580 4.329 -99.674
Pe r (L500) 1147.900 21.646 -498.369
Re - 2.833 2.856
Duration 15 40 10
Figure 6.4: Flow-rate conditions during the
accumulation, elution, and backwash stages.
Table 6.3: Imposed (Pi and Pr ) and measured
average hydrodynamic conditions.

The signal detected during the elution and backwash stages is presented in Figure 6.5. The
decrease of Q r e permits to reduce the L100 retention time from 60 to 40 minutes. Although, the
permeate flow-rate oscillates and achieves the value required to release the L500 nanoparticles (ob-
serve the red dotted line and the Q r green line presented in Figure 6.4). For this reason the signal
detected at λ = 546 nm, associated to the elution of L500, is not as linear as the one presented in
Figure 6.2 for the elution time 0 <t e < 60 min. Also, the emission signal at λ = 407 nm is noisier,
which may be related to the release of particle aggregates that affects the emission signal or to the
SDS solvent that sometimes impacts the detected signal (in terms of intensity magnitude and peak
shape as observed and discussed in Section 4.3).

Figure 6.5: Elution peaks detected at λ = 407 nm (blue line) and λ = 546 nm (orange line).

129
Although the release of L500 particles during the elution of L100, the L100 particles are collected
at 0 <t e < 40 min and the L500 particles are collected in majority at 40 <t e < 50 min. Figures 6.6-(a)
and 6.6-(b) present the I vs λ spectra measured at the instants t e =20 minutes and t e =45 minutes,
corresponding to the maximum elution peak observed for the L100 and L500, respectively. The
spectra confirm that the first peak corresponds to L100 and the second peak to L500, due to the
characteristic emission I vs λ signals as discussed in Section 4.2.2.

(a) (b)

Figure 6.6: Comparison of the I vs λ spectra measured at (a) t e =20 minutes and (b) t e =45 minutes
with the reference spectra.

Fractions were collected every 10 minutes during the elution and backwash stages for DLS
granulometry analysis, being the z-av g values presented in Table 6.4, and the fractions size dis-
tribution shown in Figure 6.7. As observed in the previous experiment for the L100+L500 separ-
ation, the z-av g values increase for the fractions collected for longer elution times, and the size
distributions show, beyond the aggregation peaks, peaks detected around O(∼ 100 − 200nm) and
O(∼ 500 − 600nm).
Table 6.4: Granulometry analysis (z-av g and count-rate values) of the collected fractions.

z-av g in nm
f1 f2 f3 f4 f5
682.2 295 249.2 340.6 1757
(19.7 kcps) (10.7 kcps) (49.8 kcps) (44.2 kcps) (58.8 kcps)

130
Figure 6.7: Size distribution obtained by DLS of the collected fractions.

Figure 6.8 compare in (a) the elution peaks detected at λ = 407 nm and in (b) the permeate
flow-rates applied during the elution and backwash stages, for the two separation experiments. This
direct comparison permits us to visualize that a small variation in the permeate flow rate applied
during the elution step (Q r e ) strongly affects the peaks retention time. Furthermore, it elucidates
the difficulty to accurately control Q r e using a fixed-pressure operation.

(a) Elution peaks detected at λ = 407 nm. (b) Measured permeate flow-rates (Q r e ).

Figure 6.8: Comparison of the two experimental conditions used to separate the binary mixture
L100+L500.

6.2.2 Effect of the solvent properties


We performed tests to evaluate the possibility of changing the solvent composition or decreasing
the SDS surfactant quantity, which would result in more reproducible peaks (in terms of shape and
maximum magnitude). The experiments were performed under the same accumulation (Pi = 25
mbar, P r a = 0 mbar, and t a = 15 min) and elution (Pi = 50 mbar and P r e = 30 mbar for
0 <t e < 40 min and Pi = 50 mbar and P r e = 45 mbar for 40 <t e < 60 min) conditions, while
varying the carrier fluid composition. The solvents used were ultra-pure water (ionic strength of
1 × 10−5 M ), PBS 1x (ionic strength of 0.404 M ), and SDS 0.1 g/L (ionic strength of 3.47 × 10−4
M ), and the system response was compared to the experiment presented in Section 6.2.1, where an

131
aqueous solution of SDS 1 g/L (ionic strength of 3.47 × 10−3 M ) was used.

Figure 6.9 presents in (a) the peaks detected with the online fluorescence spectroscopy system
(OFS) at λ = 407nm, and in (b) the measured permeate flow-rates during the elution and backwash
steps of the different experiments. The red line corresponds to the previous experiment for the
L100+L500 separation using SDS 1 g/L and consists of the reference signal. The experiment realized
using ultra-pure water as carrier fluid (blue line) presents only one peak that is completely evacuated
in the first 30 minutes of elution. The orange line represents the peak detected while using PBS 1x
as a carrier fluid, observing a signal in the void peak, an elution peak that takes 30 minutes, and a
small backwash peak. The experiment realized with SDS at 0.1 g/L (green line) shows a behavior
similar to the experiment where ultra-pure water was used.

(a) Peaks detected during the elution step at λ = 407 (b) Measured permeate flow-rates (Q r e ).
nm.

Figure 6.9: Experiments realized under the same accumulation and the elution conditions using
different solvents as carrier fluids.

Fractions were collected every 10 minutes (fractions f1 to f6) for granulometry analysis by DLS.
Table 6.5 presents the z-av g values, and Figure 6.10 shows the size distribution of the fractions
detected peaks. The experiment realized with water presents z-av g values from 849.5-1743 nm. The
fractions f1 and f2 (containing the majority of the collected peak) present two peaks (apart from the
peaks corresponding to aggregates detection) around O(∼ 100nm) and O(∼ 500 − 600nm). Thus,
the cause of the particles premature elution is not sample aggregation, as there are important peaks
of dispersed objects. The same is observed for the fraction size distribution of the experiment that
uses SDS 0.1 g/L. For the experiment using PBS 1X, the f1 and f2 fraction present 2 peaks around
O(∼ 30nm) and O(∼ 500 − 600nm). The size distribution suggests sample aggregation, signals
not being observed around O(∼ 100nm). Also, the PBS 1x presents the higher z-av g values (from
1053-2675 nm).

132
Table 6.5: Granulometry analysis (z-av g and count-rate values) of the collected fractions.

z-av g in nm
Used Solvent
f1 f2 f3 f4 f5 f6
849.5 865.3 1504 1766 1743 ∗
U.p. water
(71.3 kcps) (18.1 kcps) (14.4 kcps) (63.7 kcps) (28.3 kcps)
1053 1399 2506 1781 2675 1823
PBS 1x
(22.2 kcps) (5.9 kcps) (75.3 kcps) (67.4 kcps) (97.8 kcps) (43.2 kcps)
525 839.4 1683 ∗ ∗ ∗
SDS 0.1 g/L
(22 kcps) (22.6 kcps) (16.1 kcps)
682.2 295 249.2 340.6 1757 ∗
SDS 1 g/L
(19.7 kcps) (10.7 kcps) (49.8 kcps) (44.2 kcps) (58.8 kcps)

Data not suitable for analysis. The measurement was aborted.

(a) U.p. water (b) PBS 1 X

(c) SDS 0.1 g/L (d) SDS 1 g/L

Figure 6.10: Size distribution obtained by DLS for the fractions collected during the experiments
that used (a) u.p. water, (b) PBS 1X , (c) SDS 0.1 g/L and (e) SDS 1 g/L, as carrier fluid (solvent).

Experiments reveal the need to use SDS 1 g/L to perform fractionation of the mixture contain-
ing the L100 and L500 particles. Using an SDS solution with 0.1 g/L or ultra-pure water did not
separate the sample components. The reason behind this unsuccessful separation is not related to
sample aggregation that could invert the elution mode (as discussed in Section 2.1.1.2) since there
are L100 and L500 particles dispersed, detected in the DLS signals. In [91], the author used SDS to

133
improve the separation of SiO2 nanoparticles with 50 and 100 nm using the AF4 technique, suggest-
ing that the SDS do not change the charge of the particle but decrease the Debye length (λ D ), which
reduces the collective diffusion. Despite the possible existence of a collective diffusion that induces
the simultaneous mass flux of L100 and L500 particles, the separation cannot only be justified by
the Debye length decrease. Otherwise, salt addition would improve the separation performance, not
observed in the experiment performed with PBS 1X. The covering of a particle surface with surfact-
ants changes the particle interactions by adding electro-steric effects and particles slipping, altering
the nature of particle collisions. The addition of PBS 1x caused sample aggregation due to the elev-
ated ions concentration in solution, available to neutralize the particles surface charge (reduction of
the Debye screening length). In this way, the inter-particle distance is reduced, promoting sample
aggregation.

The separation of the L100+L500 particles using the SPFFF prototype requires SDS at high
concentrations. The minimum SDS concentration was not determined. The separation occurs for an
SDS concentration of 1 g/L, failing for an SDS concentration of 0.1 g/L. The physical phenomenon
behind the stabilization of the particles is complex, and the separation might be permitted by the
electro-steric effects or particles slipping introduced with SDS. Other surfactants can also be tested as
the FL-70, suggested in [16]. For the separation of another type of suspension, the reader must refer
to Chapter 12 ("Sample preparation and choice of the carrier liquid in Field-Flow Fractionation") of
[16].

6.3 Towards more intensive separation

6.3.1 Increase of the injected quantity


The quantity of the injected mixture was doubled by increasing the concentration of the binary mix-
ture. The injected sample consisted in 0.949 g/L of L100 and 0.952 g/L of L500, resulting in 4.74
µg/cm2 (mass of particles per membrane surface area). The accumulation and elution conditions
were the same as the experiment presented in Section 6.2.1. Figure 6.11 shows the flow-rates meas-
ured during accumulation, elution, and backwash steps. Figure 6.12 presents the peaks detected
during the elution and backwash steps at λ = 407 nm and λ = 546 nm. As observed in the previous
experiments, there is a peak detected during the elution step (0 <t e < 40 min) and a second peak
observed during the backwash step (40 <t e < 50 min). The fluorescence results suggest that the first
peak corresponds to L100 and the second peak to L500.

134
Figure 6.11: Flow-rate conditions during the accumulation, elution, and backwash stages.

Figure 6.12: Elution peaks detected at λ = 407 nm (blue line) and λ = 546 nm (orange line).

Fractions were collected every 10 minutes (f1 to f5) for granulometry analysis. The z-av g values
are presented in Table 6.6 and the the fractions size distribution is shown in Figure 6.13. The z-av g
values are higher than expected, as they are superior to the z-av g values of the experiment presented
in Section 6.2.1. The visualization of the size distribution permits us to verify that the fractions f1
to f4 have peaks around 100 nm and O(∼ 500 − 600nm), revealing the evacuation of L500 in the
first 40 minutes of elution. Although the signal detected at λ = 546 nm (orange line of Figure 6.11)
for < 0t e < 40 min is negligible (the signal is practically linear), the results show that there is L500

135
elution in simultaneous with the L100 elution peak. A small amount of L500 is released prematurely
since no signal is visualized at λ = 546nm during the first 40 minutes of elution (the particles were
not detected by fluorescence spectroscopy). Fraction f5, containing the second elution peak, presents
a size distribution with two peaks around O(∼ 500 − 1000nm), showing that aggregation occurs.
Note that, in the discussion of the DLS size distribution results for the different collected fractions, the
peaks detected at O(∼ 4µm) are not mentioned. As already explained, the existence of aggregates
in samples with low concentration yields the appearance of large particle sizes in the DLS results.
Table 6.6: Granulometry analysis (z-av g and count-rate values) of the collected fractions.

z-av g in nm
f1 f2 f3 f4 f5
485.9 651.6 694.1 589.2 1058
(17.8 kcps) (118.4 kcps) (31.9 kcps) (11.8 kcps) (69 kcps)

Figure 6.13: Size distribution obtained by DLS of the collected fractions.

The separation of a higher quantity of L100+L500 mixture was achieved. Although, the DLS
result reveals the simultaneous elution of L100 and L500 particles in the first 40 minutes of elution
and higher L500 aggregation (comparing with the experiment presented in 6.2.1). The increase
of sample quantity may require an optimization of the accumulation conditions. The accumulation
time t a must be increased to permit the complete formation of the particle layers (L500 closer to
the membrane and L100 more distant from the membrane). As there are more particles available,
attractive interactions are favored.

6.3.2 Cyclic operation: five sequences in a row


The performance of the sequential operation was tested by doing five cycles in a row. The experi-
mental conditions were the same as the ones presented in Section 6.2.1, in terms of accumulation,
elution, and backwash conditions, injected quantity, and a carrier fluid. The flow-rates measured
during the accumulation, elution, and backwash steps of the five cycles are presented in Figure 6.14.
The peaks detected at λ = 407 nm and λ = 546 nm are shown in Figure 6.15. The fractograms
reveal a poor resolution between the L100 and L500 peaks since the first run, which gets worst in

136
the last two cycles. The separation performance was affected by the permeate flow-rate that is diffi-
cult to control accurately. Observing the red dotted line of Figure 6.6, we observe that the permeate
flow-rate is smaller than zero almost in the entire experimental sequence, getting more negative in
the last two cycles. As the L500 particles are released with the application of a back-flow Q̄ r e − 5
µL/min, the oscillation of the permeate flow permitted L500 evacuation during all the elution time
(signal detected at λ = 546 nm). Two options exist for the improvement of the cyclic separation
performance: the first option is the realization of the cyclic operation for the experimental condi-
tions presented in Section 6.1, as the Q̄ r e =∼ 20 µL/min, the flow-rate oscillation will not allow the
evacuation of L500, and the second option is to control the permeate flow-rate with a syringe pump
instead of working with a fixed-pressure operation.

Figure 6.14: Flow-rate conditions during the accumulation, elution, and backwash stages for the five
sequential cycles.

137
Figure 6.15: Elution peaks detected at λ = 407 nm (blue line) and λ = 546 nm (orange line).

Table 6.7: Granulometry analysis (z-av g and count-rate values) of the collected fractions.

z-av g in nm
Cycle
f1 f2 f3 f4 f5
462.6 443.1 1031 1652 1708
1
(29.6 kcps) (16.2 kcps) (130.1 kcps) (74.8 kcps) (72.3 kcps)
571.8 587 789.6 971.7 1291
2
(41.5 kcps) (101 kcps) (81.5 kcps) (11.8 kcps) (22.4 kcps)
548.7 745 779.8 1380 2439
3
(9.5 kcps) (45 kcps) (20.6 kcps) (25.5 kcps) (67.1 kcps)
710.6 685.8 1213 1667 1858
4
(21.9 kcps) (29 kcps) (31.1 kcps) (46.6 kcps) (94.7 kcps)
696 792 1401 1648 2447
5
(60.1 kcps) (52.8 kcps) (30.1 kcps) (87.2 kcps) (21.3 kcps)

The limitation of the continuous solution presented by M. Marioli [22] was associated with the
poor sample resolution caused by the increase of membrane fouling. To find out if we would have
the same problem, we measured the permeation flux (J p ) as a function of transmembrane pressure
(T M P) to check the membrane permeability (L p ) and the values are shown in Figure 6.16. The
points measured before and after the cycles are represented by the blue squares and the orange
circles, respectively. As observed, there was no significant change in membrane permeability com-
pared to the value determined one month earlier. The backwash step permits the recovery of the
membrane permeability after each experiment. Therefore, the poor separation performance of five
cycles is not related to the decrease of the membrane permeability but with the oscillation of the
permeation flow rate, which is positive since it is easier to change the operation mode than to play

138
with membrane properties to reduce fouling.

Figure 6.16: Permeation flux (J p ) as function of the transmembrane pressure (T M P) measured


before and after the five cycles (the slope corresponds to the membrane permeability, L p ).

6.4 Recommendation for smaller size ratio separation


Along this chapter was observed that the separation of particles with size ratio 5 is possible, al-
though it becomes challenging when the flow-rate approached the Q lim
r e of L500. For the separation
of particles with a smaller size ratio, it is necessary to accurately control the permeation flow rate
as the difference between the Q lim
r e value of the different components is inferior. The closer the per-
meate flow rate is to the Q lim
r e of the larger particles, the lower the retention of the smaller particles
and hence the lower the sample dilution. Therefore, the prototype has to work with a fixed-flow
operation to allow the optimized separation (maximum possible reduction in sample dilution) of
particles with a lower size ratio.

139
140
Towards a continuous operation
7
This chapter was transformed into an internal report for Solvay to consider a patent
application.

141
142
8.1 General Conclusion
Conclusion and perspectives
8
This research aimed to assess the potential of a new separation process based on the Hollow Fiber
Flow Field-Flow Fractionation (HF5) mechanism for the industrial separation of liquid-solid nan-
oparticles suspensions. An automatized sequential prototype (SPFFF) was developed, interlinking
the three operating steps of injection, accumulation, and elution. In the beginning, one of our main
concerns was the accumulation step used in the classical HF5 configurations. The accumulation is
an essential part of the HF5 mechanism, although it requires the outlet flow inversion to confine
the sample near the inlet port. The issues are the limited sample injection, since the accumulation
surface area is minimal, and the flow inversion that can induce hydrodynamic instabilities, resulting
in sample re-mixing. Also, it brings higher operation complexity, requiring the use of an extra pump.
This step was simplified in the SPFFF by placing a downstream valve that closes the outlet during
accumulation. The particles are accumulated along the entire membrane, allowing a higher sample
injection and facilitating the operation mode.

The exploration of the accumulation and elution operating conditions guided the prescription
of a separation methodology using the SPFFF. The accumulation conditions (t a and Q r a ) need to be
optimized for the more diffusive species present in the separation mixture, otherwise, the sample
loss in the elution void peak is observed. A threshold condition (Q lim
r e ), at which the Brownian
diffusion dominates the drag force of the radial permeate flow-rate, was unveiled and is an essential
parameter to achieve a separation on the SPFFF. Accordingly, the critical operating parameters are
the accumulation and elution radial permeate flow-rate and the accumulation time.

The application of the separation methodology for size-based fractionation of polystyrene latex
nanoparticles with a size ratio of 5 was successfully achieved. The importance of choosing a suitable
solvent was demonstrated, requiring the use of a surfactant with a concentration greater than 0.1
g/L. Increasing the injected sample concentration was also tested, showing that separation of twice
the amount of particles is achievable.

The SPFFF is system-dependent, being impossible to give a unique recipe to all types of suspen-
sions. Depending on the suspension nature, the operating conditions have to be adjusted. In general,
if the Brownian diffusion is the governing phenomenon, the nanoparticle composition does not affect
the separation (unless the membrane properties favor van der Waals attractive interaction, increas-
ing particles retention time [4]). Particles positively charged interact with the equipment surface,
affecting the equipment performance and increasing sample retention. In this case, two strategies

143
are possible: cover the equipment surface with a cationic surfactant or stabilize the particles with
a negative charge. Hence, the solvent composition and pH conditions must be investigated. The
prescribed separation methodology can be used to orient the operating conditions.

A five-cycle batch operation was performed without significant reduction of the membrane per-
meability. One of the main limitations to use the device developed by M. Marioli et al. [22], where
the Asymmetric Flow Field-Flow Fractionation (AF4) mechanism is applied with a continuous op-
eration, is the scalability limitations (channel increase? how many parallel channels?), resolution
loss due to membrane fouling, and membrane replacement that requires the laborious channel dis-
assemble. Our prototype requires improvement and optimizations, which are recommended in the
subsequent sections, although its design permits a viable and industrial proved scale-up strategy
(multiple hollow fiber modules), and the cyclic operation does not significantly reduce membrane
permeability. Also, the membrane replacement is easier than in the AF4 channels, and the fact that
the authors use a micro-structured membrane can also impose difficulties in terms of membrane re-
placement, as it requires complex membrane fabrication methods, which may be long and result in
small differences in the pattern location that can affect the reproducibility of the results. Although
the device of [22] uses a 2D separation method that addresses the technological gap of continuous
particle separation, it may be a complicated solution for industrial lines.

A continuous solution may be required to integrate our prototype upstream or downstream of an


industrial process. In this regard, we have developed a continuous separation strategy not included
in this manuscript for confidentiality reasons (patent application is under review).

8.2 Perspectives

8.2.1 Technological optimization of the SPFFF


The technological optimization of the SPFFF relies on the use of a fixed-flow operation, creation of
an accumulation region, design of a non-disposable MHF5 channel, and development of a "Smart-
Process".

1. The fixed-pressure mode does not accurately control the permeate flow-rate imposed during
the elution step (key parameter), compromising the performance of operating in several cycles.
The adequate control of the permeate flow-rate is essential for the size-based separation, and
its optimization can reduce sample dilution. Hence, the use of a syringe pump for the permeate
flow control is strongly advised for future developments.

2. The samples fractionated using the SPFFF have dilution factors superior to O(∼ 100) com-
pared to the initial injected concentration. This is caused by the particles accumulation along
the entire membrane length, which does not take advantage of the laminar flow nature of
the elution step, requiring the use of strong permeate flow-rates to discriminate the retained
sample. As a consequence, the particles are evacuated for long periods. The creation of an
accumulation region compatible with the scaling strategy must be investigated. The compart-
mentalized channel could be an efficient solution, however, this configuration has a granted

144
patent [41].

3. The fibers used along this project had high durability and, depending on the solvents and
particles tested, a significant decrease of permeability was not observed or was restored with
cleaning stages. Some fibers were used for ∼ 5 months with weekly usage. Although, a re-
placement is often required to test new conditions (operating conditions, solvent properties,
and other suspensions) and guarantee that the results are not affected by the existence of ad-
sorbed or aggregated particles from previous experiments. The MHF5 channel is a lab-made
hollow-fiber module made of an organic membrane, which can be optimized to facilitate the
replacing process and to increase the membrane lifetime. Currently, to replace the channel,
we need to build the fiber module, print (by 3D printing) the piece that allows the permeate
control, glue this piece to the module, and connect the microfluidic flow controller to the piece.
The module replacement is time-consuming because it needs time to assemble a new module
and properly fit all the equipment without fluid leaks and to obtain a good membrane per-
meability since the new membrane has fabrication products that need to be removed (glycerin
and others). An important improvement of the prototype is the build of a channel (using CAD
software and fabrication by 3D printing) that allows the easiest replacement of the fiber keep-
ing the module cartridge. In parallel, the test of inorganic membranes can be an important
improvement to reduce the need for membrane replacement, since they are more resistant
to strong cleaning conditions than organic membranes. The objective must be to produce a
channel that requires the minimum replacement possible and reduces wastes.

4. A "Smart-Process" can be developed that consists in the automatic switch of the L-SWITCH and
M-SWITCH valves, or the imposed operating conditions (at the moment we impose pressure),
according to the detected signal in the FC cell instead of programming it depending on the
sample retention time, which has to be obtained from previous experiments.

8.2.2 Further exploration of the operating conditions in SPFFF


1. The accumulation step is essential to achieve the size-based separation of colloidal particles,
and more investigation is required to understand the impact of the operating conditions in the
axial particles spreading (along the membrane z-direction). Experimental and numerical (us-
ing the OpenFOAM model that accounts for inter-particle interactions) tests can be performed.
Additionally, the impact of the accumulation operating conditions in particles aggregation may
be favorable to find the optimal conditions (balance between the less experimental time and
the less sample aggregation).

2. The axial flow-rate applied during the elution step (Q z ) can be optimized to reduce the sample
retention time, thus decrease the dilution factor. The impact of the axial flow-rate in the
permeate flow-rate limit (Q lim
r e ) must be investigated. A flow-meter with higher capacity may
be required at the fiber outlet to increase the measured flow-rate capacity (replace L-unit (RD2)
by an XL-unit).

3. To better understand the impact of the solvent composition, future studies could investigate
other types of surfactants and verify the minimum required surfactant concentration. Also, it

145
may be advantageous to find a solvent that reproduces the same peak magnitude detected by
the online fluorescence spectroscopy system (OFS). Even if the mass balance is overestimated,
we could have a quantification of the separation performance. A list of possible solvents to test
according to the nature of the suspending medium is given in table 12.2 - page 194 of [16].

4. The maximum injection limit of the SPFFF prototype must be investigated. We verified that
injecting the double of particles quantity the separation was still feasible, requiring an adjust-
ment of the accumulation conditions. However, the maximum injection limit was not explored.

8.2.3 SPFFF scale-up


After optimization of the SPFFF with one fiber, a module with more fibers could be tested. The first
difficulties to face would be the connections, which will probably create dead volumes and non-
uniform distribution of the flow along the different fibers as was observed in [18].

8.2.4 Continuous operation


Included in the internal report for Solvay.

146
Bibliography

[1] Grand View Research. Nanomaterials market size, share & trends analysis report by product
(carbon nanotubes, titanium dioxide), by application (medical, electronics, paints & coat-
ings), by region, and segment forecasts, 2020 - 2027. https://www.grandviewresearch.
com/industry-analysis/nanotechnology-and-nanomaterials-market, 2021. Online;
accessed 26 June 2021.

[2] Ibrahim Khan, Khalid Saeed, and Idrees Khan. Nanoparticles: Properties, applications and
toxicities. Arabian journal of chemistry, 12(7):908–931, 2019.

[3] Patrice Bacchin. Génie des interactions physico chimiques: Applications à la transformation de la
matière molle. PhD thesis, Université Paul Sabatier-Toulouse III, 2006.

[4] Thilak K Mudalige, Haiou Qu, Desiree Van Haute, Siyam M Ansar, and Sean W Linder. Ca-
pillary electrophoresis and asymmetric flow field-flow fractionation for size-based separation
of engineered metallic nanoparticles: A critical comparative review. TrAC Trends in Analytical
Chemistry, 106:202–212, 2018.

[5] Zefang Zhang, Lei Yu, Weili Liu, and Zhitang Song. Surface modification of ceria nanoparticles
and their chemical mechanical polishing behavior on glass substrate. Applied Surface Science,
256(12):3856–3861, 2010.

[6] Pavel Janoš, Jakub Ederer, Věra Pilařová, Jiří Henych, Jakub Tolasz, David Milde, and Tomáš
Opletal. Chemical mechanical glass polishing with cerium oxide: Effect of selected physico-
chemical characteristics on polishing efficiency. Wear, 362:114–120, 2016.

[7] MJ Kao, FC Hsu, and DX Peng. Synthesis and characterization of sio2 nanoparticles and their
efficacy in chemical mechanical polishing steel substrate. Advances in Materials Science and
Engineering, 2014, 2014.

[8] GB Basim, JJ Adler, U Mahajan, RK Singh, and BM Moudgil. Effect of particle size of chem-
ical mechanical polishing slurries for enhanced polishing with minimal defects. Journal of the
Electrochemical Society, 147(9):3523, 2000.

[9] Taofang Zeng and Thomas Sun. Size effect of nanoparticles in chemical mechanical polishing-a
transient model. IEEE transactions on semiconductor manufacturing, 18(4):655–663, 2005.

[10] Jimmy L Humphrey and George E Keller. Separation process technology. McGraw-Hill New York,
1997.

[11] Nicole Pamme. Continuous flow separations in microfluidic devices. Lab on a Chip, 7(12):1644–
1659, 2007.

147
[12] P Sajeesh and Ashis Kumar Sen. Particle separation and sorting in microfluidic devices: a
review. Microfluidics and nanofluidics, 17(1):1–52, 2014.

[13] Andreas Lenshof and Thomas Laurell. Continuous separation of cells and particles in micro-
fluidic systems. Chemical Society Reviews, 39(3):1203–1217, 2010.

[14] Thoriq Salafi, Kerwin Kwek Zeming, and Yong Zhang. Advancements in microfluidics for nan-
oparticle separation. Lab on a Chip, 17(1):11–33, 2017.

[15] Stepan Podzimek. Light scattering, size exclusion chromatography and asymmetric flow field flow
fractionation: powerful tools for the characterization of polymers, proteins and nanoparticles.
John Wiley & Sons, 2011.

[16] Martin E Schimpf, Karin Caldwell, and J Calvin Giddings. Field-flow fractionation handbook.
John Wiley & Sons, 2000.

[17] Myeong Hee Moon, Hansun Kwon, and Ilyong Park. Stopless flow injection in asymmetrical
flow field-flow fractionation using a frit inlet. Analytical chemistry, 69(7):1436–1440, 1997.

[18] Ju Yong Lee, Ki Hun Kim, and Myeong Hee Moon. Evaluation of multiplexed hollow fiber
flow field-flow fractionation for semi-preparative purposes. Journal of Chromatography A,
1216(37):6539–6542, 2009.

[19] Carmen RM Bria, Patrick W Skelly, James R Morse, Raymond E Schaak, and S Kim Ratanath-
anawongs Williams. Semi-preparative asymmetrical flow field-flow fractionation: A closer look
at channel dimensions and separation performance. Journal of Chromatography A, 1499:149–
157, 2017.

[20] Michael Maskos and Wolfgang Schupp. Circular asymmetrical flow field-flow fractionation for
the semipreparative separation of particles. Analytical chemistry, 75(22):6105–6108, 2003.

[21] Pertti Vastamäki, Matti Jussila, and Marja-Liisa Riekkola. Continuous two-dimensional field-
flow fractionation: a novel technique for continuous separation and collection of macromolec-
ules and particles. Analyst, 130(4):427–432, 2005.

[22] Maria Marioli and Wim Th Kok. Continuous asymmetrical flow field-flow fractionation for the
purification of proteins and nanoparticles. Separation and Purification Technology, 242:116744,
2020.

[23] J Calvin Giddings, FJ Yang, and Marcus N Myers. Flow-field-flow fractionation: a versatile new
separation method. Science, 193(4259):1244–1245, 1976.

[24] J Calvin Giddings, Gwo-Chung Lin, and Marcus N Myers. Fractionation and size analysis of
colloidal silica by flow fieldflow fractionation. Journal of Colloid and Interface Science, 65(1):67–
78, 1978.

[25] Karl-Gustav Wahlund. Flow field-flow fractionation: critical overview. Journal of Chromato-
graphy A, 1287:97–112, 2013.

148
[26] Marcus N Myers and J Calvin Giddings. Properties of the transition from normal to steric
field-flow fractionation. Analytical Chemistry, 54(13):2284–2289, 1982.

[27] Stepan Podzimek. Asymmetric flow field flow fractionation. Encyclopedia of Analytical Chem-
istry: Applications, Theory and Instrumentation, 2006.

[28] Judith EGJ Wijnhoven, Jan Paul Koorn, Hans Poppe, and Wim Th Kok. Hollow-fibre flow field-
flow fractionation of polystyrene sulphonates. Journal of Chromatography A, 699(1-2):119–
129, 1995.

[29] Karl Gustav Wahlund and J Calvin Giddings. Properties of an asymmetrical flow field-flow
fractionation channel having one permeable wall. Analytical Chemistry, 59(9):1332–1339,
1987.

[30] HL Lee and EN Lightfoot. Preliminary report on ultrafiltration-induced polarization chromato-


graphyan analog of field-flow fractionation. Separation Science and Technology, 11(5):417–440,
1976.

[31] Jan Aake Joensson and Alf Carlshaf. Flow field flow fractionation in hollow cylindrical fibers.
Analytical Chemistry, 61(1):11–18, 1989.

[32] WYATT Technology. AF4. https://www.wyatt.com/products/instruments/


eclipse-field-flow-fractionation-system.html#eclipseneon-5. Online; accessed
08 March (2021).

[33] Andrea Zattoni, Sonia Casolari, Diana C Rambaldi, and Pierluigi Reschiglian. Hollow-fiber flow
field-flow fractionation. Current Analytical Chemistry, 3(4):310–323, 2007.

[34] Min Kuang Liu, P Stephen Williams, Marcus N Myers, and J Calvin Giddings. Hydrodynamic
relaxation in flow field-flow fractionation using both split and frit inlets. Analytical chemistry,
63(19):2115–2122, 1991.

[35] Carmen R. M. Bria. Development of asymmetrical flow field-flow fractionation for the characteriz-
ation of proteins, protein aggregation, and nanoparticles. PhD thesis, Colorado School of Mines,
2016.

[36] Katri Eskelin, Minna M Poranen, and Hanna M Oksanen. Asymmetrical flow field-flow frac-
tionation on virus and virus-like particle applications. Microorganisms, 7(11):555, 2019.

[37] Julien C Gigault, John M Pettibone, Charlene E Schmitt, and Vincent A Hackley. A rational
strategy for characterization of nanoscale particles by asymmetric flow field-flow fractionation.
2014.

[38] Won Ju Lee, Byoung-Ryul Min, and Myeong Hee Moon. Improvement in particle separation
by hollow fiber flow field-flow fractionation and the potential use in obtaining particle size
distribution. Analytical Chemistry, 71(16):3446–3452, 1999.

[39] M Van Bruijnsvoort, R Tijssen, and W Th Kok. Assessment of the diffusional behavior of poly-
styrene sulfonates in the dilute regime by hollow-fiber flow field flow fractionation. Journal of
Polymer Science Part B: Polymer Physics, 39(15):1756–1765, 2001.

149
[40] Myeong Hee Moon. Frit-inlet asymmetrical flow field-flow fractionation (fi-afifff): A stopless
separation technique for macromolecules and nanoparticles. BULLETIN-KOREAN CHEMICAL
SOCIETY, 22(4):333–348, 2001.

[41] Philip J Wyatt. Compartmentalized field flow fractionation, April 24 2012. US Patent
8,163,182.

[42] Marcus N Myers and J Calvin Giddings. A continuous steric fff device for the size separation of
particles. Powder Technology, 23(1):15–20, 1979.

[43] J Calvin Giddings. A system based on split-flow lateral-transport thin (splitt) separation cells for
rapid and continuous particle fractionation. Separation Science and Technology, 20(9-10):749–
768, 1985.

[44] C Bor Fuh. Peer reviewed: Split-flow thin fractionation., 2000.

[45] Bhajendra N Barman, P Stephen Williams, Marcus N Myers, and J Calvin Giddings. Split-flow
thin (splitt) cell separations operating under sink-float mode using centrifugal and gravitational
fields. Industrial & Engineering Chemistry Research, 57(6):2267–2276, 2018.

[46] C Bor Fuh and JC Giddings. Isoelectric split-flow thin (splitt) fractionation of proteins. Separ-
ation science and technology, 32(18):2945–2967, 1997.

[47] Myeong Hee Moon, Hyun-Joo Kim, So-Yeon Kwon, Se-Jin Lee, Yoon-Seok Chang, and Heungbin
Lim. Pinched inlet split flow thin fractionation for continuous particle fractionation: application
to marine sediments for size-dependent analysis of pcdd/fs and metals. Analytical chemistry,
76(11):3236–3243, 2004.

[48] P Stephen Williams, Shulamit Levin, Timothy Lenczycki, and J Calvin Giddings. Continuous
splitt fractionation based on a diffusion mechanism. Industrial & engineering chemistry research,
31(9):2172–2181, 1992.

[49] Pertti Vastamäki, Matti Jussila, and Marja-Liisa Riekkola. Study of continuous two-dimensional
thermal field-flow fractionation of polymers. Analyst, 128(10):1243–1248, 2003.

[50] Pertti Vastamäki, P Stephen Williams, Matti Jussila, Michel Martin, and Marja-Liisa Riekkola.
Retention in continuous two-dimensional thermal field-flow fractionation: comparison of ex-
perimental results with theory. Analyst, 139(1):116–127, 2014.

[51] J Calvin Giddings, Larell K Smith, and Marcus N Myers. Surface barriers for retention enhance-
ment in field-flow fractionation. Separation Science and Technology, 13(4):367–385, 1978.

[52] FG Glavis and JF Woodman. Kirk-othmer encyclopedia of chemical technology, vol. 13, fourth
edition. 1995.

[53] Ronald F Probstein. Physicochemical hydrodynamics: an introduction. John Wiley & Sons, 2005.

[54] Jan Mewis and Norman J Wagner. Colloidal suspension rheology. Cambridge University Press,
2012.

150
[55] William Bailey Russel, WB Russel, Dudley A Saville, and William Raymond Schowalter. Col-
loidal dispersions. Cambridge university press, 1991.

[56] Erle C Donaldson and Waqi Alam. Chapter 2surface forces. Wettability, pages 57–119, 2008.

[57] E.G. De Azevedo. Termodinamica Aplicada. ESCOLAR Editora, 2011.

[58] A Scotti, M Pelaez-Fernandez, U Gasser, and A Fernandez-Nieves. Osmotic pressure of suspen-


sions comprised of charged microgels. Physical Review E, 103(1):012609, 2021.

[59] Benjamin Espinasse. Approche théorique et expérimentale de la filtration tangentielle de colloïdes:


flux critique et colmatage. PhD thesis, Université Paul Sabatier-Toulouse III, 2003.

[60] Jeanne Chang, Pierre Lesieur, Michel Delsanti, Luc Belloni, Cecile Bonnet-Gonnet, and Bernard
Cabane. Structural and thermodynamic properties of charged silica dispersions. The Journal
of Physical Chemistry, 99(43):15993–16001, 1995.

[61] C. Pasquier, S. Beaufils, A. Bouchoux, S. Rigault, B. Cabane, M. Lund, V. Lechevalier, C. Le Floch-


Fouéré, M. Pasco, G. Pabœuf, et al. Osmotic pressures of lysozyme solutions from gas-like to
crystal states. Phys. Chem. Chem. Phys., 18(41):28458–28465, 2016.

[62] A Mourchid, A Delville, J Lambard, E Lecolier, and P Levitz. Phase diagram of colloidal dis-
persions of anisotropic charged particles: equilibrium properties, structure, and rheology of
laponite suspensions. Langmuir, 11(6):1942–1950, 1995.

[63] Bernard Cabane and Sylvie Hénon. Liquides. Solutions, dispersions, émulsions, gels. Belin, 2015.

[64] Patrice Bacchin, Djaffar Si-Hassen, Victor Starov, Michael J Clifton, and Pierre Aimar. A unifying
model for concentration polarization, gel-layer formation and particle deposition in cross-flow
membrane filtration of colloidal suspensions. Chemical Engineering Science, 57(1):77–91, 2002.

[65] Vincent Labalette. Structure and rheology of anisotropic colloids. PhD thesis, Toulouse, INPT,
2020.

[66] Paul C Hiemenz. Principles of colloid and surface chemistry, volume 188. M. Dekker New York,
1986.

[67] Patrice Bacchin, Aurélie Marty, Paul Duru, Martine Meireles, and Pierre Aimar. Colloidal sur-
face interactions and membrane fouling: Investigations at pore scale. Advances in colloid and
interface science, 164(1-2):2–11, 2011.

[68] Alexander F Routh. Drying of thin colloidal films. Reports on Progress in Physics, 76(4):046603,
2013.

[69] Patrice Bacchin, David Brutin, Anne Davaille, Erika Di Giuseppe, Xiao Dong Chen, Ioannis
Gergianakis, Frédérique Giorgiutti-Dauphiné, Lucas Goehring, Yannick Hallez, Rodolphe Heyd,
et al. Drying colloidal systems: Laboratory models for a wide range of applications. The
European Physical Journal E, 41(8):1–34, 2018.

151
[70] Nadia Ziane and Jean-Baptiste Salmon. Solidification of a charged colloidal dispersion invest-
igated using microfluidic pervaporation. Langmuir, 31(29):7943–7952, 2015.

[71] Hrvoje Jasak. Error analysis and estimation for the finite volume method with applications to
fluid flows. 1996.

[72] R Byron Bird, Warren E Stewart, and Edwin N Lightfoot. Transport phenomena, volume 1. John
Wiley & Sons, 2006.

[73] Roland Clift, John R Grace, and Martin E Weber. Bubbles, drops, and particles. Courier Corpor-
ation, 2005.

[74] Abraham S Berman. Laminar flow in channels with porous walls. Journal of Applied physics,
24(9):1232–1235, 1953.

[75] Pierre Haldenwang, Pierrette Guichardon, Guillaume Chiavassa, and N Ibaseta. Exact solution
to mass transfer in berman flow: Application to concentration polarization combined with
osmosis in crossflow membrane filtration. International Journal of Heat and Mass Transfer,
53(19-20):3898–3904, 2010.

[76] Prabhu R Nott, Elisabeth Guazzelli, and Olivier Pouliquen. The suspension balance model
revisited. Physics of Fluids, 23(4):043304, 2011.

[77] DN Petsev, VM Starov, and IB Ivanov. Concentrated dispersions of charged colloidal particles:
sedimentation, ultrafiltration and diffusion. Colloids and Surfaces A: Physicochemical and En-
gineering Aspects, 81:65–81, 1993.

[78] Lilian C Johnson and Roseanna N Zia. Phase mechanics of colloidal gels: osmotic pressure
drives non-equilibrium phase separation. Soft Matter, 17(14):3784–3797, 2021.

[79] W Richard Bowen and Frank Jenner. Dynamic ultrafiltration model for charged colloidal dis-
persions: a wigner-seitz cell approach. Chemical Engineering Science, 50(11):1707–1736, 1995.

[80] Amy Novick-Cohen and Lee A Segel. Nonlinear aspects of the cahn-hilliard equation. Physica
D: Nonlinear Phenomena, 10(3):277–298, 1984.

[81] JD Sherwood. Initial and final stages of compressible filtercake compaction. AIChE Journal,
43(6):1488–1493, 1997.

[82] T. Zsirai, P. Buzatu, P. Aerts, and S. Judd. Efficacy of relaxation, backflushing, chemical cleaning
and clogging removal for an immersed hollow fibre membrane bioreactor. Water Research,
46(14):4499 – 4507, 2012.

[83] WJC Van de Ven, Kv vant Sant, IGM Pünt, A Zwijnenburg, AJB Kemperman, WGJ Van der
Meer, and M Wessling. Hollow fiber dead-end ultrafiltration: Influence of ionic environment
on filtration of alginates. Journal of Membrane Science, 308(1-2):218–229, 2008.

[84] Fluigent. Flow unit technology. https://www.fluigent.com/product/


microfluidic-components-3/frp-flow-rate-platform/, 2021. Online; accessed
03 May 2021.

152
[85] D. W. Green R.H. Perry and J. O. Maloney. Perry’s chemical engineers’ handbook. McGraw-Hill
Education, 1997.

[86] Jana Sádecká, Veronika Uríčková, and Michaela Jakubíková. Fluorescence spectroscopy. Spec-
troscopic methods in food analysis, pages 189–223, 2018.

[87] Joseph R Lakowicz. Principles of fluorescence spectroscopy. Springer science & business media,
2013.

[88] Barbara J Frisken. Revisiting the method of cumulants for the analysis of dynamic light-
scattering data. Applied optics, 40(24):4087–4091, 2001.

[89] Dennis E Koppel. Analysis of macromolecular polydispersity in intensity correlation spectro-


scopy: the method of cumulants. The Journal of Chemical Physics, 57(11):4814–4820, 1972.

[90] John C Thomas. The determination of log normal particle size distributions by dynamic light
scattering. Journal of colloid and interface science, 117(1):187–192, 1987.

[91] Haruhisa Kato, Ayako Nakamura, Hidekuni Banno, and Mikiko Shimizu. Separation of
different-sized silica nanoparticles using asymmetric flow field-flow fractionation by control
of the debye length of the particles with the addition of electrolyte molecules. Colloids and
Surfaces A: Physicochemical and Engineering Aspects, 538:678–685, 2018.

[92] Akihiko Yamagishi. Transient electric dichroism studies on interaction of a cationic dye with
sodium dodecylsulfate. Journal of Colloid and Interface Science, 81(2):511–518, 1981.

[93] Junhong Qian, Yufang Xu, Xuhong Qian, Jiaobing Wang, and Shenyi Zhang. Effects of anionic
surfactant sds on the photophysical properties of two fluorescent molecular sensors. Journal
of Photochemistry and Photobiology A: Chemistry, 200(2-3):402–409, 2008.

[94] Francisco Lemos, José Madeira Lopes, and Fernando Ramôa Ribeiro. Reactores químicos. 2002.

153
154
Commercial channel structure of other FFF
sub-techniques
A
Figure A.1: Illustration of the centrifugal Field-Flow Fractionation mechanism. From CF2000 Series
Postnova catalog.

Figure A.2: Illustration of the Thermal Field-Flow Fractionation mechanism. From TF2000 Series
Postnova catalog.

Figure A.3: Electrical Asymmetrical Flow FFF mechanism. From EAF2000 Electrical Flow FFF Series
catalog of Postnova.

155
Table A.1: Available AF4 Channel dimensions of Wyatt Technologies.

Channel Type Length (mm) Width (mm) Height (mm)


a
AF4 Mini Channel (MC) 158 56 41
a
AF4 Short Channel (SC) 201 56 41
AF4 Large Channel (LC) a 291 56 41
b
Semi-preparative AF4 Channel (SP2) 315 90 47
c
Frit-Inlet AF4 Channel (FI) 291 70 -
a
From Separation Channel User Manual MC (Mini channel), SC (Short channel) and LC (Long channel)
catalog of Wyatt Technologies.
b
From The Semi Prepatrative Channel catalog of Wyatt Technologies.
c
From Frit Inlet Channel User Manual catalog of Wyatt Technologies.

Table A.2: Specifications of Postnova equipments.

Instrumentation Measurement Range Channel Dimensions Injection volume Operating Conditions


Volume: 0,5 - 2,5 mL
Particles: 1 nm - 100 µm 1 - 1000 µL Axial Flow: 0 - 10 mL/min
AF2000a Size: 335 x 60 x 40 mm
Polymers: 500 Da - 1012 Da standart 20 µL Crossflow: 0 - 8,5 mL/min
Thickness: 250 µm
Volume: 1,15 or 2,21 mL
1 - 1000 µL
TF2000b Particles: 10 nm - 1 µm Size: 456 x 20 x H mm Main Flow: 0,01 - 2,0 mL/min
standart 20 µL
Thickness: 130 or 250 µm
Size: 200 x 40 x H mm
GF2000c Particles: 1 - 50 µm - Main Flow: 1 - 100 mL/min
Thickness: 318 - 1150 µm
Main Flow 0,3 - 4,0 mL/min
Inner Volume:∼2,7 mL Centrifugal acceleration:
CF2000d Particles: 10 nm - 20 µm Inner Thickness: 250 µm 10 - 100 µL 11 - 26369 m/s2
Channel diameter: 200 nm Maximum rotation speed:
4900 rpm
Equipment Dimensions Current: +/- 75 mA
EAF2000e - -
430 x 270 x 90 mm Max. Voltage: +/- 22,5 V
a
From AF2000 MultiFlow Series catalog of Postnova.
b
From TF2000 Series catalog of Postnova.
c
From GF2000 GS Series catalog of Postnova.
d
From CF2000 Series catalog of Postnova.
e
From EAF2000 Electrical Flow FFF Series catalog of Postnova.

156
Illustration of the HF5 experimental steps
B

Figure B.1: Experimental steps in a HF5 configuration.

157
158
Split channel configurations
C
(a) (b)

Figure C.1: (a) Centrifugal split channel and (b) Isoelectric split channel. From [44]

(a) (b)

Figure C.2: (a) Magnetic split channel. From [44]; (b) Gravitational split mechanism. From GF2000
GS Series catalog of Postnova.

159
160
D
Submitted article: "Colloid dynamics near phase
transition: simulation of the relaxation of
concentrated layers"

161
Colloid dynamics near phase transition : simulation of
the relaxation of concentrated layers†
Adriana Ferreiraa,b† , Micheline Abbas∗a , Philippe Carvinb , and Patrice Bacchina

The dynamics of concentrated colloidal dispersion close to phase transition is challenging to predict.
The presence of attractive interactions leads to hindered motions through the change in transport
properties like collective diffusion. We propose a continuum mechanical model to describe the
dynamics of the relaxation of concentrated colloids near phase transition. The model relies on
a specific description of the equation of state of colloids, the osmotic pressure deriving from a
free energy double-well function, that induces a decrease in collective diffusion close to the phase
transition. The implementation of such transport properties in Computational Fluid Dynamics (CFD)
codes allowed to describe a fast expansion of accumulated gels near an interface (induced for instance
by evaporation or filtration processes) followed by a relaxation stage where particles are progressively
released from the layer toward the dilute bulk region. The relaxation kinetics is slower when the
transition is approached.

1 Introduction
predominant. It is widely known that strongly attractive particles
The accumulation and relaxation of a concentrated colloidal form gels at relatively low critical concentration 6 (resulting from
phase is of interest in many applications, like drying and the formation of fractal clusters 7 ), whereas weakly attractive
filtration, where out-of-equilibrium transport processes are particles form glass at high critical concentration (permanent
encountered. A significant amount of experimental and theo- trapping of particles within cages formed by nearest neighbor
retical works has been focused on the accumulation of colloidal particles) 8 . Nevertheless, colloidal dispersions with long-range
particles near an interface (a free surface and a solid membrane electrostatic interactions can exhibit disorder-to-order transition,
in the drying 1,2 , filtration 3 and pervaporation 4 configurations, even at volume fractions as low as φ = O(0.1) 9,10 . Once the
respectively). However there is evidence that concentration concentration is sufficiently large, networks of large number of
profiles, after the filtration is ceased, differ from those measured particles are consolidated, leading the colloidal suspension to
directly during filtration 5 . The dynamics and the mechanisms take the properties of a solid 11 .
controlling the colloid relaxation (when colloidal compression is
released) are far from being fully understood.
It is not unusual to observe the equation of state of colloidal
suspensions to exhibit discontinuity near phase transition 9,12 ,
Colloidal particles can remain stable in a dispersed state for
leading to dramatic slowing of the relaxation time-scales which
long duration, often due to repulsive electrostatic interactions
manifests as a divergence in the local viscosity 13,14 as hallmarks
or to the presence of some stabilizing agents. When particle
of liquid-solid transition 15 . These concentrated suspensions
concentration increases in time, due for instance to settling,
will be called gels hereafter. The properties of gels and their
drying or filtration, particles are compressed leading the osmotic
ability to relax is strongly related to inter-particle interaction
pressure to increase. The suspension can display non-equilibrium
(physico-chemical properties) and to the temperature 16,17 .
transition, like gelation or vitrification which are processes
To the authors’ knowledge, a general description of the gel
resulting from the failure of liquid-gas and liquid-solid crystal
relaxation in a dilute medium has never been considered yet.
transitions, respectively. The volume fraction at which phase
The objective of this work is to undertake a step forward toward
transition occurs depends on the interactions between particles.
the description of the relaxation process of a concentrated sus-
When the concentration increases, particles are forced to get
pension, where the motion of particles is hindered by collective
close to each other and attractive interaction forces become
interactions close to phase transition. The description will be
tackled with the aid of simple and realistic numerical simulations.
a
Laboratoire de Génie Chimique, Université de Toulouse, CNRS, INPT, UPS, Toulouse,
France; E-mail: micheline.abbas@ensiacet.fr;adriana.bentescorreia@ensiacet.fr; Numerical modeling allows studying the effect of physico-
patrice.bacchin@univ-tlse3.fr.
b chemical colloidal interactions on the suspension dynamics, with
Solvay,Research and Innovation Center of Lyon, Saint-Fons, France.
† Electronic Supplementary Information (ESI) available: github link for the Open- the possibility to uncouple different effects when relevant 18 .
FOAM code (solver and configurations). This is typically the case for nanometric particles (like macro-

1–10 | 1
molecules or small particles) which exhibit gelation phenomena. 2 Model based on hydrodynamic and ther-
The characteristic time and length scales corresponding to gelly
modynamic concepts
material relaxation can be very large, compared to the scales
corresponding to the diffusive motion of individual colloids, in 2.1 Colloid diffusion down a concentration-gradient
a way that depends on physico-chemical properties 13 . Hence, The concentration of the colloidal particle phase will be described
the development of Eulerian numerical tools allowing to capture by the volume concentration φ that varies in space x and in time
these phenomena at macroscopic scale, including first order t. The transport of colloids in the solvent is based on continuous
transitions (as it has been attempted for gelation of colloids near description:
a membrane 19 ) are relevant for practical applications. ∂φ
= −∇ · J (1)
∂t
In order to study the gel layer relaxation at large length and where J denotes the colloid flux. In a suspension at rest, the col-
time scales (compared to the colloid size and the associated loid flux J from higher to lower concentration regions originates
diffusion time scale), we developed a tool that achieves a contin- from a thermodynamic force associated with the spatial gradient
uum description of the suspension concentration and motion. It of chemical potential, and is counterbalanced by a solvent flux
is based on concepts from statistical mechanics (the Suspension in the opposite direction, i.e. J = −D(φ )∇(φ ). The diffusion of
Balance Model) 20 . In addition, the transport equations relies colloidal particles down a concentration-gradient accounting for
on the modeling of the colloidal properties accounting for hydrodynamic interactions can be expressed in the frame of the
surface interactions: the cohesion of the matter that impacts its generalized Stokes-Einstein relation 25,26 :
relaxation is related to inter-particle interactions. It is now well  
admitted that osmotic pressure is a good way to account for local K(φ ) φ ∂µ
D(φ ) = (2)
multi-body interactions between particles 21 . The advantage of 6πη a 1 − φ ∂φ P,T
using this property is twofold. First it can be easily accounted for K(φ ) is the mobility function of a particle hindered by the
in the frame of the continuous description, allowing to capture presence of neighbouring particles. The inverse of this mobility,
the effect of local interactions at larger scales. Second, its i.e. the resistance function, valid across a wide concentration
5
dependence on the concentration (that represents the Equation 1 6+4φ 3
range is from Happel and Brenner H(φ ) = K(φ )
= 1 5 .
Of State in the phase diagram) is experimentally accessible from 6−9φ +9φ 3 −6φ 2
3

the measurement of the suspension concentration at equilibrium


The thermodynamic force responsible of particle diffusion is
in dialysis bags or more recently in microfluidic devices 22 .
due to the chemical potential per particle µ . The chemical po-
Osmotic pressure based simulations have already proved that
tential variation with the concentration leads to a change in the
this approach allows to detect, how concentration polarisation
osmotic pressure. This relation can be written following for in-
can lead to gel or deposit formation during filtration 23 or to
stance Schurr 26 , while neglecting the mixture compressibility:
the formation of a skin layer during drying 24 . In that context a
critical osmotic pressure was introduced, without the description
   
∂µ 1−φ ∂Π
= Vp (3)
of the equation of state during phase transition. ∂ φ P,T φ ∂ φ µ ,T

In this work, we develop an osmotic pressure model containing where Π denotes the osmotic pressure and Vp refers to the par-
the essential ingredients to describe continuously the impact of ticle volume. Using eq. 3 in and scaling the collective diffusion
interparticle interactions on dispersion phase transition. The cor- coefficient by the particle diffusion coefficient in the dilute limit
kT
responding diffusion coefficient, deduced from Stokes-Einstein D0 = 6πη a (where k and T correspond respectively to the Boltz-
generalized law, exhibits low values close to the irreversibility mann constant and temperature), leads to:
point. The diffusion model is used to perform simulations of
∂ Π∗
 
D(φ ) 1
concentrated colloids dynamics near phase separation i.e. for D∗ (φ ) = = (4)
D0 H(φ ) ∂φ µ ,T
temperature, T , slightly above the critical temperature, Tc for
phase separation ( TTc < 1) in order to prevent phase separation V
where D∗ and Π∗ = kTp Π denote the dimensionless diffusion
from taking place. Simulations show that these properties can coefficient and osmotic pressure, respectively.
explain different relaxation kinetics of the accumulated layers ac-
cording to the ratio TTc that reflects the proximity to the instability In a more general situation, the flux J can be written as the
zone and thus the irreversibility degree of the phase transition. sum of a contribution induced by the mixture motion of the form
In the following, section 2 describes the equation for colloid φ um and the contribution resulting from collective diffusion. Eq.
transport used for colloid accumulation and relaxation, while 1 takes then the form:
section 3 explains the background of the osmotic pressure model.
The results displayed in section 4 demonstrate that such model of ∂φ
= −∇ · [φ um − D(φ )∇φ ] (5)
phase transition by osmotic pressure can describe the kinetics of ∂t
relaxation of gel layers with a rapid gel expansion followed by a The mixture velocity field um can vary in space x and time t, and
slow relaxation, this kinetics being controlled by the TTc parameter. should obey local momentum conservation. In a general frame-
work, this transport equation can be written in a dimensionless

2| 1–10
form, using the Péclet number that compares the transport by motic pressure is defined as the sum of two contributions defined
convection and diffusion, i.e. Pe = U δ /D0 , where U is a char- in equations 8, 9 and 10, as following:
acteristic mixture velocity and δ is a characteristic length scale

for the concentration gradient. Since unidirectional mass transfer ∗
Π∗ = Π + Πtr (7)
considered in this work lacks an imposed geometrical length scale
(as explained later in section 4), the equations are maintained in Below the phase transition φ ≤ φc : the osmotic pressure is de-
their dimensional form. We draw the reader’s attention that in the scribed with classical virial expansion :
frame of the suspension balance model, the equation of particle ∗
Π |φ ≤φc = φ +C2 φ 2 +C3 φ 3 (8)
transport can be written in another form, that evidences parti-
cle migration under the effect of mechanical forces arising from The first term of the relationship represents the Van’t Hoff law
local variation in the osmotic pressure. Indeed, in the absence for ideal colloids (without specific interactions) and the others
of external forces, the suspension balance model reflects the fact terms are Virial coefficients relative to the non ideality induced
that spatial variation of osmotic pressure is counter-balanced by by colloidal interactions.
a drag force, i.e. φ Fdrag = Vp dΠ
d φ ∇φ , that leads to particle slip with
respect to the local mixture. This can be formulated by: Above the phase transition φ > φc : the osmotic pressure is de-
  fined with a relationship that diverges near the close packing vol-
∂φ K(φ )
= −∇ · φ um − φ Fdrag (6) ume fraction φcp :
∂t 6πη a

" 1 #
The dual representation of the diffusive colloidal flux in eqs. ∗ ∗ dΠ  φcp − φc κ
Π |φ >φc = Π (φc ) + κ φcp − φc −1 (9)
1 and 6 is very helpful to interpret dynamics of concentrated dφ φc φcp − φ
dispersed systems, whether colloidal or not 27–29 . This approach
will allow us to further discuss the connection between the The derivative of the main osmotic pressure contribu-
particle slip, with respect to the local mixture, and the diffusion tion evaluated at the critical concentration is noted as


strength. dφ = 1 + 2C2 φc + 3C3 φc2 . The expressions of the main
φc
osmotic pressure contribution Π ensure its continuity and
that of its first derivative at the critical concentration φc . The
2.2 Osmotic pressure near phase transition parameter κ in this equation represents the isothermal inverse
Our main idea is to develop a model of the osmotic pressure Π(φ ) compressibility of the condensed phase: a high compressibility
property that include eventual effect of phase transition. First we leads to a weak increase of the osmotic pressure when increasing
shall briefly discuss the osmotic pressure and its dependence on the volume fraction. Note that attributing some compressibility
the concentration. The equation of state of a colloidal system to dense layers of particles to account for established networks
can be expressed in terms of osmotic pressure, which is defined, of connected particles can be commonly found in modeling of
from the thermodynamic viewpoint, as the negative derivative flocculated suspensions dewatering, although this is usually done
of Helmholtz free energy with respect to volume at constant through compressive yield stress 31,32 .
temperature. In dilute systems, it arises from the thermal fluc-
tuation of colloids. As the concentration increases, inter-particle Close to the phase transition φ1 < φ < φ2 : an additional contribu-
interactions (of interfacial and hydrodynamic origin) induce tion to the osmotic pressure Πtr originates from a double well
mechanical forces and consequently the suspension mechanical free energy, this approach being commonly adopted within the
pressure increases. The dependence of osmotic pressure on context of phenomenological Cahn Hilliard equation 33 . The ad-
concentration is similar to that of water activity: the larger ditional osmotic pressure can be written as :
the colloid concentration, the higher is the energy required ∗
(φ − φ1 )2 (φ − φ2 )2 (φc − φ )
 
∗ Tc dΠ
to extract additional water from the concentrated suspension. Πtr = 16 (10)
T dφ φc (φ2 − φ1 )4
Available osmotic pressure-concentration isotherms accessible
through direct or indirect methods (based on scattering), exhibit This expression is built in such a way to satisfy the continuity of
disorder-to-order phase transition signature, as it has been shown the equation of state and its first derivative along the entire range
in several experiments 9,12,30 . In order to obtain continuous of accessible volume fractions. The transition does not change the
solutions of practical mass transfer problems (by Eulerian value of the osmotic pressure, neither at φ1 nor at φ2 , in order to
simulations tools) in the frame of Stokes-Einstein generalized satisfy the Maxwell equal area rule (equilibrium between the dif-
law, we define a continuous and differentiable relationship for ferent phase as seen in Fig.1). The main parameter of this model
Π(φ ) in a wide range of concentration including the transition is TTc that compares the temperature to the critical one at phase
range φ1 < φ < φ2 , . transition. This parameter allows to scale the depth of the well
energy. We stress here that the temperature should be understood
We consider here that the phase transition occurs within a con- as a concept that describes the system proximity to phase transi-
centration range φ ∈ [φ1 , φ2 ]. We define the "critical concentration" tion, and that eases the analogy with first order gas-liquid phase
as the arithmetic average of φ1 and φ2 , i.e. φc = φ1 +2 φ2 . The os- transition. This will be explained further in section 3. Never-

1–10 | 3
theless, we note here that our model in inspired by the universal 3 Experiment-based osmotic pressure
character of gel transition and its analogy with the vapor-liquid
Fig. 2 shows the evolution of the osmotic pressure as a func-
transition, as suggested by Tanaka 34 and modeled by Manning 35 .
tion of the concentration for particles of diameter 10.2 nm, as ob-
At high temperature (T ≫ Tc ), TTc tends toward 0, the contribu-
tained from the experimental work of Chang et al. 9 . The model
tion of Πtr to the global osmotic pressure is negligible. Said dif-
explained in section 2 is applied here to describe the osmotic pres-
ferently, there is no irreversibility in the system when far from the
sure of this experimental data, with the following set of parame-
spinodal phase separation. If T = Tc , the equation of state passes
ters to fit the data: C2 = 4 × 104 , C3 = 15 × 105 , κ = 0.5, φ1 = 0.03
through the critical point and is tangent to the instability zone.
and φ2 = 0.09. The different curves are obtained for different
At this point the osmotic pressure derivative is zero, leading to a
values of TTc to illustrate the description of the transition zone be-
negligible diffusion coefficient, i.e. irreversible transition. At low
tween φ1 and φ2 . We pause here to explain the meaning of this
temperature (T < Tc ), the system looses its time reversal symme-
parameter. The osmotic pressure measurements are realized at
try. An instability takes place, leading to phase separation of the
a given temperature and ionic strength (salt concentration). As
suspension into highly concentrated regions (with aggregated or
the particle concentration is increased, the particles get closer to
ordered particles) and regions depleted from particles. This situa-
each other, leading the average potential energy to increase. Be-
tion corresponds to a negative derivative of the osmotic pressure,
low φ1 , the average interparticle distance remains greater than
and thus gives rise to an apparent negative diffusion coefficient,
the repulsive barrier, however above φ1 , the distance can become
as a hallmark of spinodal decomposition. Fig. 1 illustrates how
smaller than that barrier. When the concentration falls in the
the model (Eq. 7-10) can represent the equation of state with
transition range, the suspension undergoes metastable states, re-
a typical variation near and inside the phase separation domain,
sulting in a change of the variation of the potential energy per
based on different TTc ratios.
particle with the concentration 36 . This change in the potential
energy depends on the value of the temperature compared to the
critical temperature, which cannot be unambiguously measured
from macroscopic measurements, as those realized by Chang et
al. 9 . Therefore, TTc is left as parameter that represents the re-
versibility degree of the concentrated layers.

Fig. 1 Equation of state – osmotic pressure - volume fraction isotherms


for different TTc ratios of 0.6, 0.8, 0.95, 1, 1.2, 1.5, 2, 3, and 5 correspond-
ing to the blue, orange, green, red, purple, brown, pink, grey, and yellow
lines, respectively. Spinodal decomposition (dashed line) occurs when the
derivative of EOS is negative. C2 = 1,C3 = 0, φ1 = 0.3, φ2 = 0.5, κ = 0.5.
Fig. 2 Modeling of osmotic pressure experimental data (blue and green
Here φ1 and φ2 are assumed constant for simplicity.
symbols below and above the transition respectively) for silica 9 . The full
line represent the fitting with the model for different TTc ratio of 0, 0.4,
0.8, 0.95, 1 and 1.5 corresponding to the blue, orange, green, red, purple,
and brown lines, respectively. The dashed lines represents the limit of
In this paper, we investigate the effect of an abrupt decrease Van’t Hoff equation and the close packed volume fraction for dilute and
of the diffusion coefficient on the suspension relaxation when the concentrated dispersions respectively.
system is close to phase transition without being subject to spin-
odal phase separation, i.e. only in the region where TTc < 1 in Fig. The collective diffusion coefficient obtained from eq. 4, which
1. In practice, this corresponds to situations where particles be- is based on the derivative of the osmotic pressure (Fig.2), is dis-
come ordered due to interactions with their neighbors, without played in Fig.3 as a function of the concentration, for different
being subject to irreversible aggregation, so that the mass trans- values of TTc . The diffusion coefficient is independent of Tc /T for
port equation (eq. 1) can be solved appropriately with a positive φ > φ2 and φ < φ1 . Between φ1 and φ2 , a minimum of diffusion
diffusion coefficient. Otherwise, if phase separation takes place, appears when the dispersion approaches the phase transition, this
other classes of transport equations should be used, accounting minimum being deeper and closer to zero when T is close to Tc .
for separated particle-rich and particle-poor regions. A very low diffusion coefficient (related to a plateau in osmotic

4| 1–10
pressure with volume fraction) is a hallmark of nearly arrested
dynamics, where phase transition is approached: this has been
demonstrated experimentally and theoretically 36,37 . This behav-
ior means that the mobility between concentrated and diluted
phases close to the transition zone is negligible, i.e. the relative
velocity of the phases associated with the drag force in Eq. 6 is
zero, thus leading to slow dynamics 38 . The low diffusion coeffi-
cient can be considered as a way to describe the dynamical arrest,
which is a characteristic feature of colloidal gels or glasses 8 . Note
that this minimum value of the diffusion coefficient occurs along
with two peaks below and a above the critical volume fraction.
This variation is the consequence of the osmotic pressure associ-
ated with a double-well energy function (Eq. 10) from which the
diffusion coefficient is derived (through Eq. 4).

Fig. 4 Illustration of the simulation procedures, with the plot of typ-


ical concentration profiles versus the wall-normal position y. Boundary
conditions are given at y = 0 and y → ∞. Two different situations are
considered: a) Relaxation of an initially uniform film at φ0 = φ2 – b1)
Colloid accumulation near a wall driven by a negative mixture velocity
um (the different concentration profiles corresponding to different times),
and b2) Relaxation of an accumulated film in a quiescent fluid. In the
figures corresponding to the relaxation of colloidal layers, the starting
concentration profiles are shown in red solid lines, while the transient 1D
simulations are shown with dashed lines. The flux J of colloids leaving
the concentrated layers is also sketched with red arrows.

trated on one side (at low y → 0), while the suspension is dilute
at y → ∞ . Zero flux is imposed at the boundaries, i.e. ∂∂φy = 0 at
Fig. 3 Variation of the diffusion coefficient as a function of the volume
fraction for different TTcr of 0, 0.4, 0.8, 0.95, 1 and 1.5 corresponding to y = 0 and y → ∞. The discretization of the divergence operator is
the blue, orange, green, red, purple, and brown lines, respectively. The performed using the "Gauss" discretization approach, and "Gauss
minimum of the diffusion coefficient approaches zero when the temper- linear" for the interpolation of the gradients. The Crank–Nicolson
ature becomes close to the critical temperature. Vertical dotted, dot- transient scheme is used for the discretization in time of the diffu-
dashed and dashed lines indicate the diffusion at φ1 , φc and φ2 , respec-
sion term. A fixed one-dimensional spatial grid with two regions
tively.
is used. The first region (0 < y < 1) where the relaxation occurs
contains 3200 cells with an expansion ratio of 100. Here l.u. de-
4 Simulation of colloid relaxation notes a numerical length unit in the simulation box. The second
Simulations have been realised with the model presented in region (1 < y < 20 l.u.), used to mimic semi-infinite conditions,
the previous section in order to determine the relaxation of has 2000 cells with an expansion ratio of 20. The time step re-
concentrated colloidal layers. Transient relaxation of colloidal quired to respect the Fourier stability criteria should be less than
suspensions displaying non-equilibrium phase transition are ∆y2min /(2D0 ), where the smallest mesh size is denoted as ∆ymin (in
obtained from two initial conditions (Fig. 4). The first idealized our simulations the time step is 1 × 10−5 ).
situation corresponds to a colloidal film of uniform concentration
φ2 (Fig. 4a). This allows to discuss the time required by a com- 4.1 Relaxation of an initially uniform film φ0 ≈ φ2
pacted suspensions to relax near phase transition, as a function We consider the relaxation of a uniform layer of initial con-
of the system reversibility characterized by the parameter Tc /T . centration φ0 and thickness δ0 . Outside this layer, the bulk
The second situation corresponds to a film accumulated near a concentration is set to a small value φbulk (here we assume
membrane-like wall in order to mimic practical situations like gel φbulk = 0 without loss of generality). The colloid transfer from
formation during filtration or drying processes (Fig. 4b). the layer to the bulk is thus set by (φ0 − φbulk ). In this case, the
macroscopic length and time scales characterizing the relaxation
The Equations 1, 4, and 7 are solved numerically using the of this layer are δ0 and δ02 /D0 , where D0 is the particle diffusion
open source software OpenFOAM (Open Field Operation and Ma- coefficient in the dilute limit. As an example, if we consider the
nipulation) Foundation 7. In the present paper, the transport of relaxation of a concentrated layer of thickness 100µ m containing
colloids is restricted to one spatial direction (the y direction), as- particles of diameter equal to 10nm in water at room temperature,
suming symmetry in other directions. The colloids are concen- the corresponding time scale is ≈ 500s.

1–10 | 5
Fig. 6 Evolution in time of the gel thickness δg , starting from a layer
Fig. 5 Relaxation of a uniform colloidal layer: wall-normal concentration
of thickness δ0 of uniform concentration φ0 = φ2 . The blue crosses, or-
profiles in time for Tc /T = 0.95. The profiles with red line, blue squares,
ange circles, green squares, red triangles, purple plus, brown diamonds,
orange triangles, green crosses, purple plus, brown diamonds, pink left
pink pentagons, grey hexagons, and yellow thin diamonds correspond to
triangles, and grey pentagons are obtained at dimensionless relaxation
Tc /T = 0, 0.4, 0.6, 0.8, 0.9, 0.95, 0.96 and 0.97, respectively. The gel
times (scaled by δ02 /D0 ) equal to 0, 0.01, 0.02, 0.05, 0.1, 0.14, 0.2, and
thickness is scaled by the initial film thickness δ0 and the time by δ02 /D0 .
1 respectively. Horizontal dotted, dot-dashed and dashed lines indicate
φ1 , φc and φ2 , respectively.

To study the colloid flow from the concentrated layer into the of that layer is called δg , in reference to gel. In one dimension,
R
dilute bulk region, we have set φ0 = φ2 . This is especially useful this thickness is defined as δg = φ >φc dy, with the corresponding
R
to describe the system relaxation when close to phase transition. mass of particles per unit area being mg = ρ p φ >φc φ dy. In
In the absence of phase transition, the colloid diffusion down order to characterize mass transfer of colloids from the dense
the concentration gradient progresses smoothly, like in a classic layer toward the bulk, we examine the decrease of δg in time,
diffusion problem. Hence we focus on the dynamics close to displayed in figure 6, where the film thickness is scaled by δ0 and
the critical point. The evolution in time of the wall-normal the time by δ02 /D0 .
concentration profiles is shown in figure 5, for Tc /T = 0.95. At
short times, one can clearly identify on this figure two different During the early stages of film relaxation, the film thickness

concentration gradients: a smooth one inside the suspension decreases like t and the curves collapse, for any irreversibility
layer for φ < φ2 , and a steep gradient close to φc , where the parameter. This decrease corresponds to the particle diffusion
particle release from the concentrated layer is damped as the from a layer where the diffusion coefficient is almost constant
collective diffusion coefficient approaches small values. This (D(φ2 )) to a semi-infinite medium, initially at constant con-
extends the lifetime of the quasi-uniform layer near the wall. centration (equal to 0 in this case). After a short time scale,
The flux of particles leaving the layer toward the dilute region as the concentration varies across the film and the diffusion
leads to a reduction of the layer thickness. At very large times, coefficient gradually decreases with y, it is not straightforward
once the layer concentration becomes less than the critical to predict theoretically the rate at which δg decreases. From the
concentration, the evolution of the concentration profiles follows simulations, the decrease of film thickness δg in time seems to
classic diffusion in a semi-infinite medium. follow a power law t n , with a power n > 1 that depends on the
irreversibility parameter. It is remarkable that the power tends
We stress here that, the closer the suspension to phase tran- to 1 when Tc /T approaches 1, or equivalently that the flux of
sition, i.e. Tc /T → 1, the steeper is the concentration variation particles leaving the dense layer is constant in time. Indeed,
at the layer front. A rigorous numerical solution of the problem this is expected to be a consequence of the concentration jump
would have required the use of an adaptative mesh, to finely re- at the interface between the dense layer and the bulk. The
solve the transient steep-gradient region while the colloid layer concentration jump is particularly steeper when the parameter
shrinks. As our numerical scheme does not follow this proce- Tc /T is closer to 1. The particle flux at the so-called interface
dure, and to avoid situations where numerical diffusion prevails yc can be obtained from D(yc ) ddyφ |yc . As Tc /T → 1, while the
over the physical one, we limit our study to systems where the diffusion coefficient D(yc ) is small and the concentration gradient

minimum diffusion coefficient at φc is small compared to D(φ2 ) dy |yc is large, the resulting product seems to remain finite and
without approaching zero. constant during this second relaxation stage.
From the concentration profiles, we extracted a quantity of
interest for practical applications, i.e. the amount of material From the plot of the gel thickness δg in time, we defined the
jammed above the critical concentration that can potentially form film relaxation time scale tr as the time it takes for the maximum
a gel or a glass, and studied its evolution in time. The thickness concentration in the layer to become less than φc . Said differ-

6| 1–10
ently, this corresponds to the instance at which δg cancels. Figure in figure 8 result from the competition between particle ac-
7 shows tr as a function of the irreversibility parameter Tc /T (here cumulation by convection and opposite diffusion down the
again, the time is scaled by δ02 /D0 ). It can be clearly noticed that concentration gradient. D0 /|um | and D0 /u2m are the unique length
the relaxation time increases with the irreversibility parameter. and time scales during the accumulation: they will be called La
When Tc /T approaches 1, we expect to observe the divergence of and τa , respectively. La represents the characteristic thickness
this time scale, since the particle release from the concentrated of the interface between the concentrated layer and the dilute
layer will be significantly hindered by the motion arrest near φc suspension and τa characterizes the speed at which this front
(at the layer front). Indeed, from eqs. 4  and 10it can be in- moves forward along the y direction. If we consider the filtration
ferred that the diffusion coefficient D(φc ) ∝ 1 − TTc dΠ d φ |φc , drops of particles of diameter 10nm, at a typical filtration speed of
to small values especially when the suspension approaches the 10−5 m/s, La = O(1)µ m and τa = O(0.1)s. Our model with the
critical point. As the relaxation time of the particle layer by dif- diffusion coefficient based on the osmotic pressure assumes
fusive mass transfer is inversely proportional to the diffusion co- that the time required for the colloids to reach thermodynamic
efficient at the front of the layer, we expect the relaxation time to equilibrium at a given concentration is smaller than the accu-
diverge like a power of 1Tc . mulation time scale τa . For a given irreversibility parameter, a
1− T
unique representation of the concentration
 profiles during the
accumulation stage takes the form φφ0 Lya , τta .

Fig. 7 Relaxation time tr∗ (scaled by δ02 /D0 ) of a uniform particle layer
versus Tc /T .

Fig. 8 Concentration profiles obtained during particle accumulation at


a wall, for different accumulation times, t/τa = 1, 6, 10, and 15 corre-
4.2 Relaxation of an accumulated film sponding to the blue, orange, green, and red lines, respectively. The
concentration φ is scaled by the bulk concentration φ0 , while the posi-
To simulate colloid accumulation near an interface (called wall
tion with respect to the wall is scaled by the characteristic accumulation
hereafter) corresponding to evaporation or filtration processes, length scale La = D0 /|um |. The irreversibility parameter is Tc /T = 0.95.
we have solved the transport equation 5, with a constant Horizontal dotted, dot-dashed and dashed lines indicate φ1 , φc and φ2 ,
(negative) wall-normal mixture velocity um (first step in Fig.4b). respectively.
Far from the wall, the suspension is assumed to be sufficiently
dilute, with a concentration φbulk < φc . During the accumulation During the colloid accumulation at the wall the concentration
stage, the pure fluid phase is allowed to evacuate through the progressively increases as y → 0. The exact shape of the concen-
wall located at y = 0 but not the particles. The zero particle flux tration profile depends on the irreversibility parameter. Figure 8
J is ensured at the wall by setting the concentration gradient displays the concentration profiles at three accumulation times,
dφ um
dy |y=0 = D(φw ) φw , where φw denotes the particle concentration in a case where the suspension is close to phase transition,
at the wall. Note that the concentration gradient at the wall i.e. Tc /T = 0.95. As the diffusion coefficient drops significantly
is negative as the mixture velocity um is negative, and that in near phase transition at φc , a barrier is established near that
this accumulation region, the particle velocity (of increasing concentration. The closer the system to the critical point (large
concentration) is smaller than the fluid phase velocity, but the Tc /T ), the steeper is the concentration jump. The jump in the
mixture velocity remains constant along the y direction satisfying concentration profile near the transition region is associated with
the mass balance. This behavior is consistent in the frame of the a weak slip that the particles experience in that region. Indeed,
suspension balance formulation. the weak slip between the particle phase and the local mixture
results from the small diffusion coefficient as φ approaches φc (as
The particle accumulation near the wall is an infinitely shown in eqs. 5 and 6). Once the particles cross that barrier, they
transient process, where the thickness of colloidal layer grows fall in the dense region where the diffusion coefficient increases
continuously in time. Typical concentration profiles displayed continuously with the concentration. When the concentration

1–10 | 7
at the wall φw exceeds the transition range, the concentration the concentration exceeds φc . δg versus time is displayed for
profiles become self-similar. They exhibit a sharp variation in different irreversibility parameter Tc /T . Note that the initial
the transition region, and smoother variation in the concentrated state corresponding to all these curves is obtained separately,
and dilute regions, respectively. However, the concentration from the simulation of particle accumulation with different Tc /T
gradient is larger in the concentrated region, where the diffusion parameters.
coefficient is larger than in the dilute region (to give the reader
a better idea, D(φ2 )/D(φ1 ) ≈ 2.5 for the particles considered in
figure 3).

Next we examine the relaxation of the layers accumulated


near the wall (second step in Fig.4b). At a given instant, the
wall-normal velocity um is switched off, and the colloids are
allowed to relax through the diffusion process. In that case, the
characteristic length scale in the system is set by the thickness
of the accumulated layer. While the shape and thickness of
the film depend on the irreversibility parameter, we define
a characteristic film thickness δc where the concentration is
equal to the critical concentration φc . A sharp concentration
gradient is observed near δc as Tc /T → 1 (as observed in figure
8), while this jump is absent as Tc /T → 0. At the onset of the
relaxation, the position of the film where φ = φc will be called δc0 .

Fig. 10 Evolution of (a) the mass of particles per unit area (scaled by the
accumulated mass m0 at the onset of relaxation) and (b) the gel thick-
ness, during the relaxation stage for different irreversibility parameters
Tc /T = 0, 0.4, 0.6, 0.8, 0.85, 0.9, 0.95, 0.96, and 0.97 corresponding to
the blue, orange, green, red, purple, brown, pink, grey, and yellow lines,
respectively. The time is scaled by δc0 2 /D . The concentration profile at
0
the onset of the relaxation stage is taken after an accumulation time of
t/τa = 10.
Fig. 9 Relaxation of the accumulated colloidal film: wall-normal con-
centration profile in time for a Tc /T = 0.95 at different relaxation times
t = 0, 4.2 × 10−4 , 2.1 × 10−3 , 4.2 × 10−3 , 2.1 × 10−2 , 4.2 × 10−2 , 8.5 × 10−2 ,
From figures 9 and 10, it can be observed that upon canceling
and 4.2 × 10−1 , corresponding to the blue, orange, green, red, purple,
brown, pink, and grey lines respectively. The time is scaled by δc0 2 /D . the wall-normal (filtration or evaporation) velocity, the film
0
Horizontal dotted, dot-dashed and dashed lines indicate φ1 , φc and φ2 , relaxation exhibits two main stages, both of them being faster
respectively. when the suspension is fully reversible, i.e. Tc /T = 0, as there is
no real barrier against particle diffusion from the concentrated
We consider for instance the relaxation of the green profile to the dilute region. During the first stage, in case the suspension
(in figure 8), where φw /φ0 ≈ 40, obtained at t/τa = 10 for is close to phase transition, the region of high concentration near
Tc /T = 0.95. At the onset of particle relaxation (resetting the the wall, subtended by the sharp front, quickly expands under
initial time to t = 0) the film resulting from accumulation is the effect of high colloid diffusion or osmotic pressure in there.
geometrically separated in two phases. At y > δc0 , the suspension During this stage, the concentration gradient of the layer between
behaves like a fluid phase where particles diffuse smoothly the wall and the sharp front decreases significantly, without

toward y → ∞. However in the region falling near y ≈ δc0 , the vanishing however (thus the decrease of film thickness as t is
colloids are arrested as they exhibit very low diffusion coefficient not observed at the end of this stage). During a second stage,
in that region. The evolution of the concentration profiles in the colloid release from the near-wall layer is relatively slower,
time is displayed in figure 9. The corresponding film thickness the kinetics being controlled by the irreversibility parameter. In
δg as a function of time is displayed in figure 10. Like in case Tc /T → 1, the dynamics during this stage is similar to that
figure 6, δg represents the thickness of the particle layer where observed in the previous section, corresponding to the relaxation

8| 1–10
of a uniform particle layer (with φ = φ2 at t0 ). Indeed, the is important to study if this law holds true, when the colloidal
characteristic relaxation time scales are similar, and δg (t) tends suspensions is fully out of equilibrium as it is the case in this
to become linear as Tc /T → 1. work. Third, the relaxation time scale computed in this work is
qualitative. Indeed, we kept the spatial grid fixed during particle
Furthermore, the time tr necessary to complete the relaxation accumulation and relaxation. Nevertheless, as the concentration
of the accumulated layers is displayed in figure 11 as a function jump near the singularity associated with colloidal arrest near
of the irreversibility parameter Tc /T . Here tr corresponds to φc is very steep, quantitative validation should be tested against
the time at which δg and mg tend to 0. As the gel first expands advanced CFD tools that allow considering adaptative spatial
before relaxing, the appropriate scaling for the time during the discretization near the region where the singularity is located.
2
gel relaxation will be δc,max /D0 , where δc,max corresponds to the These issues are open for future investigation.
characteristic film thickness of the gel at the end of the expansion
stage (the maximum in figure 10b). This allows roughly to
rationalize the curve of the relaxation time as a function of the
5 Conclusion
irreversibility parameter, when relaxation is considered starting This paper explores a model that can be used to study the
from different accumulation profiles: the blue and orange sym- dynamics of colloidal suspensions near phase transition. The
bols in figure 11 correspond to the relaxation of the green and model for colloid transport relies on the relationship between
red curves in figure 8, respectively. Indeed, the non-dimensional the colloid diffusion and osmotic pressure in the frame of the
time for relaxation are very close confirming a good scaling law generalized Stokes-Einstein relation, and can be implemented
for the relaxation time. in the frame of any Eulerian CFD simulation tool. The osmotic
pressure model accounts for first order transition near the critical
point; the associated particle diffusion coefficient becomes very
small when this parameter approaches 1. The model was applied
to study the relaxation of a dense colloidal layer accumulated
near an interface, subsequent to normal filtration or evaporation
process. The suspension proximity to the critical point has been
the main parameter of study, by means of a parameter Tc /T
that describes the irreversibility degree of the phase transition.
The simulations have shown that the film relaxation undergoes
first an expansion stage where the concentration in the layer
decreases down to the value φ2 corresponding to the upper
boundary of phase transition. The kinetics of particle relaxation
is not dependent on the irreversibility parameter in this stage.
During a second stage, a concentration jump subsists at the
interface between the dense and dilute regions, especially when
the local suspension state approaches the critical point. In that
case, the particle release from the film is slowed down by the
Fig. 11 Relaxation time tr∗ as a function of the irreversibility parameter
2
Tc /T of the non-uniform layers. The relaxation time is scaled by δc,max /D0 , particle arrest near the critical point, leading to longer relaxation
where δc,max corresponds to the maximum gel thickness at the end of time scales.
the gel expansion stage. The blue crosses and orange circle are obtained
from the relaxation of concentrated layers obtained at accumulation times The numerical simulations have shown that the relaxation time
t/τa = 10 and t/τa = 15, respectively.
of a colloidal layer depends on both the irreversibility degree
(associated with physico-chemical properties of transition) and
By comparing figures 11 and 7, we observe a similar trend the accumulated mass near an interface. In practice, one can
for the evolution of the relaxation time required as a function estimate based on our results, the relaxation time from the gel
of the irreversibility parameter. The fact that colloidal particles mass as following. After the gel expansion, the concentration
are arrested near phase transition retards the relaxation of a is relatively equal to φ2 , the corresponding thickness of the gel
dense colloidal layer near a wall, and the relaxation time tends phase δc,max can be estimated as the ratio mg,0 /φ2 , where mg,0
to diverge as the critical point is approached. corresponds to the mass of the gel per unit area, accumulated
near an interface, at the onset of relaxation. The relaxation
Notes on the results: We end this section by three remarks. First, time, in seconds, would be obtained from the dimensionless time
we only considered in this work situations where the diffusion displayed in figure 11 as [tr∗ × δc,max
2 /D0 ]. To conclude, the model
of colloidal particles is significantly reduced near the critical and simulations shown herein constitute a first step toward the
point, without allowing the suspension to undergo irreversible analysis of the variety of gel time relaxation data obtained with
spinodal decomposition. Otherwise, this would require solving different experimental conditions as discussed in the introduction
the interfacial dynamics between separated phases. Second, the of this paper, and to characterize the relaxation dynamics with
osmotic pressure law assumes thermodynamic equilibrium. It the Tc /T parameter.

1–10 | 9
1032–1038.
17 M. Doi, Journal of the Physical Society of Japan, 2009, 78,
Conflicts of interest 052001.
There are no conflicts to declare. 18 G. C. Agbangla, P. Bacchin and E. Climent, Soft Matter, 2014,
10, 6303–6315.
Notes and references
19 P. Bacchin, B. Espinasse, Y. Bessiere, D. F. Fletcher and
1 A. F. Routh, Reports on Progress in Physics, 2013, 76, 046603.
P. Aimar, Desalination, 2006, 192, 74–81.
2 P. Bacchin, D. Brutin, A. Davaille, E. Di Giuseppe, X. D. Chen,
20 P. R. Nott and J. F. Brady, J. Fluid Mech., 1994, 275, 157–199.
I. Gergianakis, F. Giorgiutti-Dauphiné, L. Goehring, Y. Hallez,
21 L. C. Johnson and R. N. Zia, Soft Matter, 2021, 17, 3784–
R. Heyd et al., The European Physical Journal E, 2018, 41,
3797.
1–34.
22 H.-T. Nguyen, M. Massino, C. Keita and J.-B. Salmon, Lab on
3 P. Bacchin, A. Marty, P. Duru, M. Meireles and P. Aimar, Ad-
a Chip, 2020, 20, 2383–2393.
vances in colloid and interface science, 2011, 164, 2–11.
23 P. Bacchin, D. Si-Hassen, V. Starov, M. J. Clifton and P. Aimar,
4 N. Ziane and J.-B. Salmon, Langmuir, 2015, 31, 7943–7952.
Chem. Eng. Sci., 2002, 57, 77–91.
5 J. D. Sherwood, AIChE J., 1997, 43, 1488–1493.
24 K. Ozawa, T. Okuzono and M. Doi, Japanese journal of applied
6 J. Bergenholtz, W. C. K. Poon and M. Fuchs, Langmuir, 2003,
physics, 2006, 45, 8817.
19, 4493–4503.
25 G. Batchelor, J. Fluid Mech., 1976, 74, 1–29.
7 A. Krall and D. Weitz, Phys. Rev. Lett., 1998, 80, 778. 26 J. M. Schurr, Chem. Phys., 1982, 65, 217–223.
8 V. Trappe and P. Sandkühler, Curr. Opin. Colloid Interface Sci., 27 M. Frank, D. Anderson, E. R. Weeks and J. F. Morris, J. Fluid
2004, 8, 494 – 500. Mech., 2003, 493, 363–378.
9 J. Chang, P. Lesieur, M. Delsanti, L. Belloni, C. Bonnet-Gonnet 28 P. Bacchin, J. Phys - Condens. Mat., 2018, 30, 294001.
and B. Cabane, J. Phys. Chem., 1995, 99, 15993–16001.
29 M. Abbas, A. Pouplin, O. Masbernat, A. Liné and S. Décarre,
10 H. Tanaka, J. Meunier and D. Bonn, Phys. Rev. E, 2004, 69, AIChE J., 2017, 63, 5182–5195.
031404.
30 A. Boire, P. Menut, M. H. Morel and C. Sanchez, J. Phys. Chem.
11 R. Buscall and L. R. White, J. Chem. Soc., Faraday Trans. I, B, 2015, 119, 5412–5421.
1987, 83, 873–891.
31 K. A. Landman, C. Sirakoff and L. R. White, Phys. Fluids A:
12 C. Pasquier, S. Beaufils, A. Bouchoux, S. Rigault, B. Ca- Fluid Dynamics, 1991, 3, 1495–1509.
bane, M. Lund, V. Lechevalier, C. Le Floch-Fouéré, M. Pasco,
32 G. H. Meeten, Colloid Surface A, 1994, 82, 77–83.
G. Pabœuf et al., Phys. Chem. Chem. Phys., 2016, 18, 28458–
33 A. Novick-Cohen and L. A. Segel, Physica D, 1984, 10, 277–
28465.
298.
13 T. R. Kirkpatrick and D. Thirumalai, Rev. Mod. Phys., 2015,
34 T. Tanaka, in Phase Transitions of Gels, 1992, ch. 1, pp. 1–21.
87, 183–209.
35 G. S. Manning, ACS Omega, 2018, 3, 18857–18866.
14 J. G. Wang, Q. Li, X. Peng, G. B. McKenna and R. N. Zia, Soft
36 J. C. F. Toledano, F. Sciortino and E. Zaccarelli, Soft Matter,
Matter, 2020, 16, 7370–7389.
2009, 5, 2390–2398.
15 G. L. Hunter and E. R. Weeks, Reports on Progress in Physics,
37 Z. Varga and J. Swan, Soft Matter, 2016, 12, 7670–7681.
2012, 75, 066501.
38 L. Cipelletti and L. Ramos, Journal of Physics: Condensed Mat-
16 E. Del Gado and W. Kob, Europhysics Letters (EPL), 2005, 72,
ter, 2005, 17, R253.

10 | 1–10
172
E
Abstract submission: "Plugging of a hollow fiber
capillary by gel relaxation during a cleaning step"

173
Plugging of a hollow fiber capillary by gel relaxation during a
cleaning step
A. Ferreira1,2 , M. Abbas1 , P. Aimar1 , P. Bacchin1 , P.Carvin2 , and A.Hipólito2
1
Laboratoire de Génie Chimique, Université de Toulouse, CNRS, INPT, UPS, Toulouse, France
2 Solvay,Research and Innovation Center of Lyon, Saint-Fons, France

In membrane processes, the cleaning of gel or layers of accumulated particles on membranes is of impor-
tance. Gel formation during fouling can lead to membrane clogging during wastewater treatment [1] or to
the formation of plugs in hollow fibers [2]. The gel removal procedure is often prohibited by particle agglom-
eration, qualified here as an irreversible phase transition. Our study aims to contribute to the understanding
of the relationship between operating conditions (hydrodynamics and physico-chemical properties) and gel
reversibility, with the aim of predicting and optimizing the cleaning stages in filtration processes.

For this purpose, we developed a model that solves the colloid transport in the frame of continuum
fluid mechanics, including a description of the osmotic pressure near sol/gel phase transition, applied to
nanoparticles. The colloid transport is solved using the software OpenFOAM. We first considered the
accumulation of particles during tangential filtration in a hollow fiber. Second, the flux was stopped, and
we investigated the relaxation of particle layers, searching for the eventual occurrence of fiber plugging
during the expansion stage. We studied the impact of the filtration Peclet (P e) leading to a given initial
accumulated concentration (φo ) and of the gel reversibility on the gel relaxation time scale. We observed
that, under certain conditions, the gel (defined as the region where the concentration φc > 0.06) fills the
entire cross-section leading to an irreversible blockage, as shown in Figure 1.

Figure 1: Contour plot of the concentrated layers inside a hollow fiber a) after filtration (t0 ), b) during the
relaxation (t1 ) and after a plug is formed (t2 ). The fiber length is scaled by a factor 1/20 in the x-direction
to ease the illustration of the concentration field. The domain width in the y-direction represents the 1-mm
radius of the hollow (cylindrical) fiber. In each snapshot, the boundaries of each contour plot represent the
membrane (top), the fiber center (bottom) the fiber inlet (left), and the fiber outlet (right). The black line
denotes the limit of the gel phase, where the concentration is larger than φc = 0.06.

References
[1] T. Zsirai et al. “Efficacy of relaxation, backflushing, chemical cleaning and clogging removal for an
immersed hollow fibre membrane bioreactor”. In: Water Research 46.14 (2012), pp. 4499–4507. issn:
0043-1354.
[2] WJC Van de Ven et al. “Hollow fiber dead-end ultrafiltration: Influence of ionic environment on filtration
of alginates”. In: Journal of Membrane Science 308.1-2 (2008), pp. 218–229.

1
Equipment specification

Table F.1: Specification sheet number 1 of the microfluidic flow controller (MFCS-EZ.)
F
Specification Sheet 1
Equipment: Microfluidic flow controller

Designation MFCS-EZ

Function Generation of a pressure-driven flow

Model MAESFLO-EZ

Specifications
Number of channels 3
channel 1 - 345 mbar
Pressure range channel 2 - 1000 mbar
channel 3 - 1000 mbar
Input pressure 1300 mbar
Pressuring gaz Non corrosive or explosive gas (air, N2 , O2 , ...)
Connections Female luerlock with 4mm OD tube
Price
Equipment already available in the lab.
Observations
TM
Set with M AESF LO software a stabilization period of 10 minutes at the beginning of the day (with the
feed pressure tubes disconnected). Disconnect all the tubing at the end of the day to prevent fluid entering
in the equipment (which would damage the electrical part).
Occurred malfunctions: the green light started to turn off (what stops the pressure supply) during the
experiments or when touching the channels to connect and disconnect the reservoirs.
The problem was solved by replacing the channels (Fluigent intervention).

175
Table F.2: Specification sheet number 2 of the inlet rotatory valve (LV).

Specification Sheet 2
Equipment: Rotatory valve

Designation LV

Function Perform sequential sample injection and solvent circulation

Model L-SWITCH

Specifications
Number of ports 6
Number of positions 2
Internal volume 660 nL
Maximum applied pressure 7 bar
Connections FEP Tubing 1/16 OD
Price at 03/05/2019
Switch board (to connect all the valves - only one is required): 510 €
L-SWITCH: 2395 €
L-SWITCH connectors and tubing kit: 130 €
Observations
Verification of the valve flow-rate before each experiment. The injection loop tends to clog between
experiments and needs to be cleaned (with u.p. water and air). At the end of the day, clean the valve with
u.p. water. Use a support to tilt the valve down and prevent fluid leaks from reaching the electrical part.

176
Table F.3: Specification sheet number 3 of the outlet rotatory valve (MV).

Specification Sheet 3
Equipment: Rotatory valve

Designation MV

Function Differential collection of sample fractions

Model M-SWITCH

Specifications
Number of ports 11
Number of positions 10
Internal volume 11.6 µL
Maximum applied pressure 7 bar
Connections FEP Tubing 1/16 OD
Price at 03/05/2019
M-SWITCH: 3010 €
M-SWITCH connectors and tubing kit: 145 €
Observations
At the end of the day, clean the valve with u.p. water. Use a support to tilt the valve down and prevent fluid
leaks from reaching the electrical part.
Occurred malfunctions: The valve stopped working due to fluid leaks that reached the motor.
Solved (Fluigent intervention) by replacing the electrical part of the valve and prevented using a support.

177
Table F.4: Specification sheet number 4 of the inlet flow-meter (RD1).

Specification Sheet 4
Equipment: Flow-meter

Designation RD1

Function Measurement of the inlet flow-rate

Model XL-unit

Specifications
Range 5 mL/min
Accuracy 5% of the measured value
Maximum input pressure 15 bar
Temperature range 10-50 ◦ C
Internal volume 127.23 µL (ID=1.8 mm)
Material PEEK (Polyether ether ketone) and Borosilicate Glass
Connections FEP Tubbing 1/16 OD inserted into a nut and a ferrule
Price at 03/05/2019
FLOWBOARD (to connect all the flow-meters - only one is required): already available in the lab
XL-unit: 1530 €
Flow Unit XL connectors and tubing kit: 79 €
Operating principle

Thermal sensors.
The heater transfers a heat flow-rate to the fluid
(h = q/t [kJ/s]), the temperature sensors measure the
temperature difference (∆T [◦ C]), and the volumetric
flow-rate is calculated according to Q v = h/(ρC p ∆T ) [m3 /s]

178
Table F.5: Specification sheet number 5 of the outlet flow-meter (RD2).

Specification Sheet 5
Equipment: Flow-meter

Designation RD2

Function Measurement of the outlet flow-rate

Model L-unit

Specifications
Range 1 mL/min
Accuracy 5% of the measured value
Maximum input pressure 15 bar
Temperature range 10-50 ◦ C
Internal volume 39.27 µL (ID=1.0 mm)
Material PEEK (Polyether ether ketone) and Borosilicate Glass
Connections FEP Tubbing 1/16 OD inserted into a nut and a ferrule
Price at 03/05/2019
Equipment and connectors already available in the lab.
Operating principle

Thermal sensors.
The heater transfers a heat flow-rate to the fluid
(h = q/t [kJ/s]), the temperature sensors measure the
temperature difference (∆T [◦ C]), and the volumetric
flow-rate is calculated according to Q v = h/(ρC p ∆T ) [m3 /s]

179
Table F.6: Specification sheet number 6 of the fluorescence cell (FC).

Specification Sheet 6
Equipment: Fluorecence cell

Designation FC

Function Detection of the elution peaks

FIA-SMA-FL-ULT
Model
ref. 79013
Specifications
Inner diameter, ID = 2 mm
Dimensions
Internal volume, V = 30 µL
Windows material = Fused silica
Material
Cell material = Ultem (polyetherimide)
Optical fibber (SMA 905) connection: 2 stainless steel connectors
Connections
Tubing connectors: connector 1/4-28 flat-bottom for 1/16 OD tubing 1/4 − 28
Price
FC: 700 €(at 24/10/2019)
Cleaning kit with 2 silica windows (ref: FIA-SMA-WIN): 81 €(at 13/02/2021)
Cleaning kit with 2 silica windows and 2 stainless steel connectors (ref: FIA-SMA-WIN-SS): 144 €
(at 13/02/2021)
Optical fibbers: 384 €/unit (at 24/10/2019)
Observations

Figure illustrating the cell geometry.


The cell windows need to be frequently replaced (link showing
the windows replacement:
https://www.flowinjection.com/hardware/flow-cells/sma-z-series)
Occurred malfunctions: Use of positively charged CeO2 caused strong
adsorption to the cell walls and windows. Cell was damaged in the
cleaning process. The FC had to be replaced by a new one (all the
calibration curves had to be redone)
To prevent: verify the sample charge/stabilization before measurement.

180
Table F.7: Specification sheet of the spectrometer (HDX).

Specification Sheet 7
Equipment: Spectrometer

Designation HDX

Function Reception of the emitted light

Model OCEAN-HDX-UV-VIS

Specifications
Range 200-800 nm
Connections Slit (to connect to the optical fiber SMA 905) with a dimension of 200 µm
Price (at 24/10/2019)
HDX: 5005 €
Observations
The slit can have different sizes from 10-200 µm, acting as an entrance aperture through which the light passes.
Smaller slit sizes provide best optical resolution and larger slits allow higher light throughput.

Table F.8: Specification sheet number 8 of the light source (DH-2000).

Specification Sheet 8
Equipment: Light source

Designation DH-2000

Function Light supply

Model DH-2000-S-DUV-TTL

Specifications
Deuterium DH-2000-DUV-B (lifetime 1000 hrs)
Light sources
Halogen DH-2000-DH (lifetime 900 hrs)
Connections Optical fiber SMA 905
Warm-up time 25 min at 23◦ C
Price
DH-2000: 2007 €(at 24/10/2019)
Deep UV Deuterium Bulb (ref: DH-2000-DUV-B): 749 €(at 24/03/2021)
Observations

The light source spectra is the light emitted by the lamps


and reflected by the cell, being specific for each cell.
A different cell presents the same spectra shape but different intensities.
For a different cell is necessary to measure the light source spectra
and redo the calibration curves.
There are three factors that may affect the spectra intensity:
the need to clean the cell and replace the windows, lamps lifetime,
and unstable electrical supply.

181
182
Online fluorescence spectroscopy
G
Figure G.1: Visualization of I vs t at five λ (300, 407, 519, 520, and 546 nm) and I vs λ signals at
the right and left windows of the OceanView software, respectively.

(a) (b)

Figure G.2: Visualization of I vs λ signal in the OceanView software. The green line corresponds to
the reference spectrum and the blue line is the instantaneous spectrum measured with the lamps
turned off and (a) after a background and (b) without a background subtraction.

183
184
Sample Characterization

Table H.1: Model particles specifications.


H
Sample Supplier
Type of sample Size / MW λexc (nm) λemi (nm) Price
designation (reference)
Carboxylate polystyrene
L50 Biovalley (16661-10) 50 nm 441 486 425.60€ at 24/07/2020
nanoparticles −(CH2 −CH−Ph)n −
Carboxylate polystyrene
L100 Biovalley (19774-10) 100 nm 360 407 289.75€ at 24/06/2020
nanoparticles −(CH2 −CH−Ph)n −
Carboxylate polystyrene
L200 Biovalley (19391-10) 200 nm 529 546 302.10€ at 02/04/2021
nanoparticles −(CH2 −CH−Ph)n −
Carboxylate polystyrene
L500 Biovalley (19391-10) 500 nm 529 546 295.75€ at 24/07/2020
nanoparticles −(CH2 −CH−Ph)n −
Silica oxide
S100 Micromod (42-00-102) 100 nm 485 510 150€ at 03/02/2021
nanoparticles SiO2

H.0.1 Measurement of the maximum emission wavelength


The emission spectra (λemi ) of the L50, L100, and L500 model particles were determined by fluor-
escence spectroscopy at Laboratoire des IMRCP, being the analysis performed by Dr. Charles-Louis
Serpentini. The results show that there are slight differences between the spectra announced by the
supplier and the measured spectra. Figure H.1 shows the spectra of L50, L100, and L500, and Table
H.2 presents the measured and supplied maximum λemi .

Figure H.1: Spectrum of the L50, L100, and L500 determined by fluorescence spectroscopy by ex-
citing the photons at λex c .

185
Table H.2: Supplied and measured maximum λemi of the L50, L100, and L500 nanoparticles.

Concentration Supplied Supplied Measured


Sample
(mg/L) λexc (nm) λemi (nm) λemi (nm)
L50 9.5 441 486 487 + 510
L100 8.8 360 407 411
L500 9.9 529 546 555
S100 2.3 485 510 514

H.0.2 Titration curves of zeta potential vs pH

(a) (b)

Figure H.2: Titration curve of (a) L50 and (b) L100.

(a) (b)

Figure H.3: Titration curve of (a) L200 and (b) L500.

186
Figure H.4: Titration curve of S100.

187
188
I.1 Applied models
Residence Time Distribution modeling
I
I.1.1 Laminar Newtonian pipe flow model (LNPF)

0 τ
if t ≤ 2
E(t) = 2 (I.1)
 τ3 if t > τ
2t 2

Where τ is the geometric residence time.

I.1.2 Cascade model reactores (CR)


t n−1 t
E(t) = e− τ (I.2)
τn (n − 1)!

Where τ is the geometric residence time and n is the number of reactors.

I.1.3 Dispersive plug flow model (DPF)


v
2
 
1 t Pe −Pe (τ−t)
4tτ
E(t) = e (I.3)
2 πtτ

Where τ is the geometric residence time and Pe is the Peclet number defined as Pe = Uz L/Dc
(being U Z the axial velocity, L the circuit length, and Dc the dispersion coefficient).

189
I.2 Signals Modeling

I.2.1 LV signal

(a) (b)

Figure I.1: (a) LV signal modeled with the laminar Newtonian pipe flow model (red curve), cascade
reactors model (orange), and dispersive plug flow reactor (green). (b) Figure containing only the
cascade reactors (orange curve) and dispersive plug flow reactor (green curve) models (to visualize
better the models fitting and the LV signal).

I.2.2 RD1 signal

(a) (b)

Figure I.2: (a) RD1 signal modeled with the laminar Newtonian pipe flow model (red curve), cascade
reactors model (orange), and dispersive plug flow reactor (green). (b) Figure containing only the
cascade reactors (orange curve) and dispersive plug flow reactor (green curve) models (to visualize
better the models fitting and the RD1 signal).

190
I.2.3 RTD signal

(a) (b)

Figure I.3: (a) RTD signal modeled with the laminar Newtonian pipe flow model (red curve), cascade
reactors model (orange), and dispersive plug flow reactor (green). (b) Figure containing only the
cascade reactors (orange curve) and dispersive plug flow reactor (green curve) models (to visualize
better the models fitting and the RTD signal).

I.3 Obtained parameters


Table I.1: Parameters obtained with the CR and PFR models. Nomenclature: t¯r - mean residence
time, n - number of reactors determined bu the CR model, Pe - Peclet number determined with the
DPF model, and Dc - Dispersion coefficient calculated from the Pe number obtained with the DPF
model.

Signals
LV RD1 RTD
Models
t¯r = 0.90 min t¯r = 1.10 min t¯r = 4.02 min
CR
n =4.88 n =12.42 n =7.37
t¯r = 0.76 min t¯r = 1.08 min t¯r = 3.68 min
DPF
Pe = 7.38 ⇔ Dc = 0.0286 m2 /s Pe = 12.23 ⇔ Dc = 0.0384 m2 /s Pe = 11.03 ⇔ Dc = 0.131 m2 /s

191
192
Measurement of the RD1 signal using ultra-pure
water and SDS
J
The experiments consisted of the injection of 33µL of L100 at 0.518 g/L and measurement of the
signal at the outlet of the RD1 flow-meter. Two types of solvents were used to push the sample
towards the system (carrier fluid), ultra-pure water and SDS (1g/L). Each experiment was repeated,
being the hydrodynamic conditions presented in Table J.1.
Table J.1: Hydrodynamic conditions of the RD1 signal measurements.

Q̄ i t̄ r τ
Experiment Re Pez
(µL/min) (min) (min)
5
RD1 with u.p. water 157.488 2.406 2.754 × 10 1.547 1.234
5
RD1 repetition with u.p. water 153.860 2.340 2.682 × 10 1.559 1.273
RD1 with SDS 1 g/L 163.655 2.472 2.829 × 105 1.702 1.239
5
RD1 repetition with SDS 1 g/L 161.784 2.460 2.817 × 10 1.475 1.247

Figure J.1 presents the measured intensity signals at the RD1 outlet. It is visible that the peaks
shape is reproducible when the experiments are performed with water (blue and orange lines) but
the same is not observed using SDS (green and red lines). The SDS peaks have higher magnitude
than the water signals and the two repetitions are not reproducible in terms of peak maximum and
peak shape.

Figure J.1: Signals measured at the outlet of the RD1 flow-meter.

193
This phenomenon was observed with another type of experiment, namely during the SPFFF
experiments presented in Chapters 5 and 6. The choice of showing the difference in the usage of
SDS and u.p.water in the signals determined after the flow-meter RD1 relies on the fact that only the
SDS can be the cause of the signal discrepancies (there is no membrane, thus the existence of fouling
that gets released during the experiments can not be the cause of signals with higher intensity). The
SDS is affecting the detected signal, thus interfere in the calculation of the mass balance for the
SPFFF experiments, however, it is required to stabilize the sample when an accumulation step is
performed. Although it affects the signal magnitude in the SPFFF experiments, the peaks retention
time is reproducible (the repetition of SPFFF experiments using the same operating conditions and
SDS as carrier fluid results in peaks with different maximum intensities but the same retention times).

194
K
LV signal

Experiments were performed to investigate the impact of the LV valve in the injected signal disper-
sion. The tests were performed using u.p. water as carrier fluid and using two injection volume
loops of 30 and 60 µL (repeated experiments). The hydrodynamic of the experiments is presented
in Table K.1 and the detected signals are shown in Figure K.1.
Table K.1: Hydrodynamic conditions of the LV + tube 4 signal measurements.

Q̄ i t̄ r τ
Experiment Re Pez
(µL/min) (min) (min)
5
Vin j = 30µL 159.580 2.436 2.789 × 10 0.984 0.458
5
Repetition Vin j = 30µL 151.534 2.304 2.638 × 10 1.050 0.496
5
Vin j = 60µL 156.217 2.342 2.681 × 10 1.120 0.711
5
Repetition Vin j = 60µL 155.913 2.335 2.673 × 10 1.111 0.714

Figure K.1: Signals measured at the after the LV + tube 4 circuit.

The mean residence time of the peaks ( t̄ r ) is higher than the geometric residence time (τ), and
the sample takes ∼ O(3.5) minutes to completely leave the LV + tube 4 circuit. This means that the
signal is retarded due to possible interactions between the sample (within the sample itself) and the
material of the circuit [94].

There is no significant difference in the total injection time using volume loops with 30 and

195
60 µL, however, the use of a volume loop with 1 mL resulted in a total injection time of 8 minutes
(experiment not shown in this manuscript). The problem of long injection times is that an important
part of the sample is not submitted to the necessary accumulation time, resulting in a high sample
elution in the void peak. For this reason, and because higher volume loop results in long tubes
connected to the valve, the use of an injection loop volume of ∼ O(30) µL (in the order of the classic
HF5) was chosen, and the injection amount is increased through sample concentration.

196
Repeatability of L100 + L500 separation
L
The separation of the binary mixture containing L100+L500 (each component with a concentration
of O(∼ 0.5g/L)) is reproducible. Despite the peaks maximum value and shape are not identical in
all experiments (differences caused by the SDS effect and permeate flow-rate oscillation), we master
the peaks retention time. In Figures L.1 to L.3 are presented the experimental results realized under
the same accumulation conditions (Pi = 25 mbar, Pr a = 0 mbar, and t a = 15minut es) and different
elution conditions (enunciated in the Figures legend), containing in each Figure the repetition of the
same elution conditions. The results show that the peak retention time is reproducible and responds
as expected.

(a) Peaks detected during the elution step at λ = 407 (b) Elution permeate flow-rates Q r e .
nm.

Figure L.1: Experiments realized under the same accumulation and the elution conditions of P i = 50
mbar and Pr e = 25 mbar.

197
(a) Peaks detected during the elution step at λ = 407 (b) Elution permeate flow-rates Q r e .
nm.

Figure L.2: Experiments realized under the same accumulation and the elution conditions of P i = 50
mbar and Pr e = 34 mbar.

(a) Peaks detected during the elution step at λ = 407 (b) Elution permeate flow-rates Q r e .
nm.

Figure L.3: Experiments realized under the same accumulation and the elution conditions of P i = 50
mbar and Pr e = 30 mbar.

198
200

Você também pode gostar