86
Towards an Engineering Perspective of Enzymatic Microreactors Inês Correia da Silva Thesis to obtain the Master of Science Degree in Biotechnology Supervisor Dr. Pedro Carlos de Barros Fernandes Examination Committee Chairperson: Prof. Luís Joaquim Pina da Fonseca Supervisor: Dr. Pedro Carlos de Barros Fernandes Member of the Committee: Prof. Maria Suzana Leitão Ferreira Dias Vicente October 2016

Towards an Engineering Perspective of Enzymatic Microreactors

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Towards an Engineering Perspective of Enzymatic

Microreactors

Inês Correia da Silva

Thesis to obtain the Master of Science Degree in

Biotechnology

Supervisor

Dr. Pedro Carlos de Barros Fernandes

Examination Committee

Chairperson: Prof. Luís Joaquim Pina da Fonseca

Supervisor: Dr. Pedro Carlos de Barros Fernandes

Member of the Committee: Prof. Maria Suzana Leitão Ferreira Dias Vicente

October 2016

i

Acknowledgments

I wish to express my sincere gratitude to all the people who contributed, one way or another, to

the development of this research work. First I would like to thank my supervisor Doctor Pedro Fernandes

for all the patience and guidance throughout the progress of this project.

I would also like to extend my gratitude to my friends and colleagues from the lab, especially to

Rita, Sónia, Rosarinho, Catarina, Sofia, Carolina and Jéssica for all the great moments and for providing

me a friendly work environment.

Finally, I would like to express my genuine appreciation to my parents, sister, grandfather and

boyfriend for all the absolute support during my academic life, for all the patience and love. Thank you

all!

ii

iii

Abstract

Continuous production operations using different enzymatic microreactor systems with packed-

bed configurations were performed in the present research project. The main goal of this work was to

evaluate the development and feasibility of microstructured reactors packed with different types of

biocatalysts. A synthetic polymer, polyvinyl alcohol (PVA), and sea sand silica particles extra pure were

used as support materials for the immobilization of inulinase and invertase enzymes from commercial

preparations. The silica-immobilized system was tested in batch mode for the characterization of the pH

and temperature profiles, and for the determination of the kinetic parameters.

Both inulinase and invertase were effectively immobilized in PVA beads through an extrusion

method using a peristaltic pump, and in silica particles in the presence of glutaraldehyde 10.0% (v/v).

Silica-biocatalysts displayed higher activity at 55ºC and pH 4.5 for inulinase and at 50ºC and pH 5.0 for

invertase. Kinetic studies performed for inulinase-silica system evidenced certain diffusion limitations as

result of the restricted movement of the molecules passing through the silica (micro)porous, and also

suggested the presence of some conformational changes of the enzyme as result of the immobilization

procedure.

Continuous production of reducing sugars by hydrolysis of inulin and sucrose 5.0% (w/v) was

performed using three different microreactor systems: a cylindrical glass column packed with different

sized activated PVA beads; a cylindrical steel microreactor and a PMMA-microfluidic reactor, with a

sandwich-like structure, packed with activated silica particles. Pressure drop studies and axial

dispersion studies were also performed for all the packed-bed systems used. Lastly, operational stability

studies were performed under continuous flow, during 10 days at 50ºC and pH 4.5, using the microfluidic

system packed with 200 and 500 mg of inulinase-silica particles. The reactor system highly packed with

500 mg of enzyme carriers presented higher operational stability, allowing the retention of approximately

83% of the initial enzymatic activity of the biocatalysts.

Keywords: Microreactors, Packed-bed configuration; Enzyme immobilization, Continuous operation,

Polyvinyl Alcohol beads, Silica particles.

iv

v

Resumo

O presente projecto de investigação focou-se no estudo de operações de produção em contínuo

usando diferentes sistemas de microreactores enzimáticos em leito empacotado. O objectivo principal

deste trabalho foi de avaliar o desenvolvimento e a viabilidade de reactores micro-estruturados

empacotados com diferentes tipos de biocatalisadores. Um polímero sintético, álcool polivinílico (PVA),

e partículas de sílica da areia do mar extra puras foram usados como materiais de suporte para a

imobilização dos enzimas inulinase e invertase de preparações comerciais. O sistema de imobilização

com sílica foi testado em modo descontínuo para a caracterização dos perfis de pH e temperatura, e

para a determinação dos parâmetros cinéticos.

Tanto a inulinase como a invertase foram eficientemente imobilizadas em partículas de PVA

através de um método de extrusão usando uma bomba peristáltica, e em partículas de sílica na

presença de glutaraldeído de 10,0% (v/v). Os biocatalisadores de sílica demonstraram maior actividade

a 55°C e a pH 4,5 para a inulinase e a 50°C e a pH 5,0 para a invertase. Os estudos cinéticos efectuados

para o sistema inulinase-sílica evidenciaram certas limitações de difusão, como resultado do

movimento condicionado das moléculas ao atravessarem os (micro)poros da sílica, e também

revelaram a presença de algumas alterações conformacionais do enzima como resultado do processo

de imobilização.

A produção contínua de açúcares redutores por hidrólise da inulina e sacarose de 5,0% (m/v) foi

efectuada usando três sistemas de microreactors diferentes: uma coluna cilíndrica de vidro empacotada

com partículas de PVA de diferentes tamanhos; um microreactor cilíndrico de aço e um reactor

microfluídico de PMMA, com uma estrutura tipo sandwich, empacotados com partículas activadas de

sílica. Estudos de queda de pressão e estudos sobre dispersão axial foram também efectuados para

todos os sistemas de leito empacotado usados. Por fim, estudos de estabilidade operacional foram

desenvolvidos sob fluxo contínuo, durante 10 dias a 50°C e a pH 4,5, usando o sistema microfluídico

empacotado com 200 e 500 mg de partículas de sílica com inulinase. O sistema reaccional densamente

empacotado com 500 mg de material enzimático apresentou uma maior estabilidade operacional,

permitindo a retenção de aproximadamente 83% da actividade enzimática inicial dos biocatalisadores.

Palavras-chave: Microreactores, Leito empacotado, Imobilização enzimática, Operação em contínuo,

Partículas de álcool polivinílico (PVA), Partículas de sílica.

vi

vii

Table of Contents

Acknowledgments ................................................................................................................ i

Abstract ...............................................................................................................................iii

Resumo ................................................................................................................................ v

Table of Contents ...............................................................................................................vii

Figure Index .........................................................................................................................ix

Table Index ..........................................................................................................................xi

List of Abbreviations ......................................................................................................... xiii

1. Introduction .................................................................................................................. 1

1.1. Enzymes as Biocatalysts ......................................................................................... 1

1.1.1. Historical Facts and Importance of Enzymes .................................................... 1

1.1.2. Enzyme Properties ........................................................................................... 2

1.1.3. Enzyme Immobilization .................................................................................... 2

1.1.3.1. Methods of Immobilization ......................................................................... 4

1.1.3.2. Advances and Applications of Enzyme Immobilization Technology ........... 7

1.1.3.3. Immobilization Supports: PVA and Silica ................................................... 9

1.2. Enzymatic model systems ......................................................................................12

1.2.1. Inulin hydrolysis to reducing sugars .................................................................12

1.2.2. Sucrose hydrolysis to reducing sugars ............................................................13

1.2.3. Inulinase and invertase immobilization procedures – review ...........................15

1.3. Immobilized Enzyme Reactors ...............................................................................15

1.3.1. Modes of Operation – Batch and Continuous mode ........................................16

1.3.2. Enzymatic Microreactors .................................................................................18

1.3.2.1. Principles of Enzymatic Microreactors ......................................................18

1.3.2.2. Microfluidic Reactor Technology in Industry .............................................19

1.3.2.3. Process Intensification .............................................................................20

1.3.2.4. Applications of Microreaction Technology ................................................20

2. Objectives ....................................................................................................................23

3. Materials and Methods ................................................................................................25

3.1. Materials ................................................................................................................25

3.2. Methods .................................................................................................................26

3.2.1. Inulinase and invertase Immobilization ............................................................26

viii

3.2.2. Scanning electron microscopy (SEM) ..............................................................28

3.2.3. Batch assays ...................................................................................................28

3.2.4. pH and temperature profiles of free and immobilized enzymes .......................28

3.2.5. Determination of the kinetic parameters ..........................................................28

3.2.6. Continuous Operations ....................................................................................29

3.2.6.1. Determination of the flow rates .................................................................29

3.2.6.2. Continuous Operation assays ..................................................................29

3.2.6.3. Operational stability under continuous flow ..............................................29

3.2.7. Analytical Methods ..........................................................................................30

i) Quantification of reducing sugars by dinitrosalicylic acid (DNS) method ..........30

ii) Quantification of proteins by Bradford protein assay .......................................30

iii) Determination of packed bed pressure drop ....................................................30

iv) Determination of the Péclet number ................................................................31

4. Results and Discussion ..............................................................................................33

4.1. Enzyme Immobilization Parameters .......................................................................33

4.2. SEM analysis of silica particles ..............................................................................36

4.3. Effect of temperature and pH on the enzymatic activity of free and immobilized

enzymes ...........................................................................................................................37

4.4. Free and immobilized enzymes kinetic parameters ................................................39

4.5. Continuous Flow Operations ..................................................................................41

4.6. Pressure drop evaluation .......................................................................................44

4.7. Axial dispersion evaluation .....................................................................................47

4.8. Operational stability under continuous flow ............................................................48

5. Conclusions and Future Work ....................................................................................51

References ..........................................................................................................................53

Annexes ..............................................................................................................................65

Annex I. Determination of the flow rates............................................................................65

Annex II. Quantification of reducing sugars by the DNS method .......................................66

Annex III. Protein quantification by the Bradford method ...................................................67

Annex IV. Mass transfer calculations ................................................................................67

Annex V. Average bed porosity………………………………………………………………….69

References – Annexes…………………………………………………………………………….70

ix

Figure Index

Figure 1. Enzyme Immobilization Methods15. ......................................................................................... 4

Figure 2. Schematic representation of irreversible and reversible enzyme immobilization methods. .... 4

Figure 3. Determinant factors involved in enzymatic immobilization and activity27. ............................... 8

Figure 4. Chemical structure of polyvinyl alcohol (partially hydrolyzed)48. ............................................. 9

Figure 5. Schematics of the functionalization and activation of inorganic support during covalent

immobilization, by action of polyethylenimine (PEI) and glutaraldehyde (GTA), respectively. ............. 11

Figure 6. Schematic representation of the action of endo-inulinase and exo-inulinase on inulin, for the

production of fructo-oligosaccharide syrup and ultra-high fructose syrup, respectively75. .................... 13

Figure 7. Schematic representation of the action of invertase on sucrose for the production of α-D-

glucose and β-D-fructose81. ................................................................................................................... 14

Figure 8. Examples of bioreactor systems used in continuous bioconversion processes involving free

or immobilized enzymes103,104. .............................................................................................................. 18

Figure 9. The three microreactor configurations used for continuous flow operations: cylindrical glass

column (A), cylindrical steel microreactor (B), and PMMA-microfluidic reactor (C). ............................. 26

Figure 10. Biocatalytic PVA beads with diameters ranging between 3 and 5 mm (A), and biocatalytic

silica particles with thickness 110 µm (B).............................................................................................. 27

Figure 11. SEM micrographs of the sea sand silica particles: sea sand silica particles (A); initial particle

surface topography (B); particles after immobilization procedure (C); particle surface topography after

immobilization procedure (D). The analysis after the immobilization procedure were performed for silica

particles with inulinase. .......................................................................................................................... 36

Figure 12. Temperature (A, C) and pH (B, D) activity profiles for inulinase and invertase in the free form

and immobilized in silica particles with 10.0% (v/v) glutaraldehyde. Bioconversion assays were carried

out in 5.0% (w/v) inulin and sucrose solutions for inulinase and invertase biocatalysts, respectively.

Standard deviation did not exceed 7.4%. .............................................................................................. 38

Figure 13. Michaelis-Menten plots for the free (A) and immobilized (B) enzyme, exhibiting the

experimental (vi observed) and predicted (vi expected) curves of the initial reaction rate (v) for inulin

hydrolysis in function of the substrate concentration (S). Bioconversion reactions were carried out at

55°C and pH 4.5 in 1 mL of substrate solution, with 1 µL of free enzyme solution 0.1% (w/v) and 70 mg

of immobilized enzyme. The data presented was obtained through Solver tool from Microsoft Excel

2013. ...................................................................................................................................................... 40

Figure 14. Continuous production assays for inulin and sucrose hydrolysis performed by inulinase and

invertase immobilized systems. A-B – Continuous operations using inulinase and invertase-PVA beads

packed in the cylindrical glass column. Inulin 5.0% (w/v) and sucrose 5.0% (w/v) hydrolysis were

performed at 50°C pH 4.5; C – Continuous operation using inulinase and invertase-silica particles

packed in the cylindrical steel microreactor; D – Continuous operation using inulinase and invertase-

silica particles packed in the PMMA-microfluidic reactor. Inulin 5.0% (w/v) hydrolysis performed by

x

inulinase-silica particles was carried out at 55°C pH 4.5, and sucrose 5.0% (w/v) hydrolysis performed

by invertase-silica particles was carried out at 50°C pH 5.0. ................................................................ 42

Figure 15. Pressure drop measurements of the different packed microreactors and the respective flow

resistances (RF): A - Pressure drop of the cylindrical glass column packed with different sized-PVA

beads; B - Pressure drop of the PMMA-microfluidic reactor and the cylindrical steel microreactor packed

with 200 mg of silica biocatalysts. ......................................................................................................... 45

Figure 16. Microchannel bed of the PMMA-microfluidic reactor packed with 200 mg of inulinase-silica

biocatalysts. ........................................................................................................................................... 49

Figure 17. Operational stability of the PMMA-microfluidic reactor with sandwich-like configuration for

inulin hydrolysis based on relative activity (%). 200 mg of inulinase-silica particles were used for the

hydrolysis of 5.0% (w/v) inulin solution in acetate buffer 0.1 M, pH 4.5 at 50°C. At day one, the

concentration of reducing sugars was 45.09±2.90 g.L-1, with a volumetric productivity of 75.63 g.L-1.h-1.

............................................................................................................................................................... 49

Figure 18. Microchannel bed of the PMMA-microfluidic reactor packed with 500 mg of inulinase-silica

biocatalysts. ........................................................................................................................................... 50

Figure 19. Operational stability of the PMMA-microfluidic reactor with sandwich-like configuration for

inulin hydrolysis based on relative activity (%). 500 mg of inulinase-silica particles were used for the

hydrolysis of 5.0% (w/v) inulin solution in acetate buffer 0.1 M, pH 4.5 at 50°C. At day one, the

concentration of reducing sugars was 55.10±0.63 g.L-1, with a volumetric productivity of 92.43 g.L-1.h-1.

............................................................................................................................................................... 50

Figure A1. Flow rate calibration curve for the cylindrical glass column packed with different sized PVA

beads (with diameters ranging between 3 and 5 mm). The assays were performed with distilled

water……………………………………………………………………………………………………………..65

Figure A2. Flow rate calibration curve for the cylindrical steel microreactor and PMMA-microfluidic

reactor packed with silica particles. The assays were performed with distilled water…………………..65

Figure A3. Schematic representation of the reduction of 3,5-dinitrosalicylic acid (DNS reagent) to 3-

amino,5-nitrosalicylic acid, in the presence of reducing sugars…………………………………………….66

Figure A4. DNS calibration curve for fructose concentrations ranging from 0 to 5 mg.mL-1…………..66

Figure A5. Calibration curve used for protein quantification analysis, obtained from BSA standards with

concentrations ranging from 0.0025 to 0.02 mg.mL-1……………………………………………………….67

Figure A6. Mass transfer coefficients (K), in function of the superficial velocity of the fluid, determined

for the cylindrical glass column reactor packed with three different sized PVA beads (with diameters

ranging between 3 and 5 mm)………………………………………………………………………………….68

Figure A7. Mass transfer coefficients (K), in function of the superficial velocity of the fluid, determined

for the cylindrical steel microreactor packed with 200 mg of silica particles biocatalysts…………………69

Figure A8. Mass transfer coefficients (K), in function of the superficial velocity of the fluid, determined

for the PMMA-microfluidic reactor packed with 200 mg of silica particles biocatalysts…………………...69

xi

Table Index

Table 1. Key effects caused by enzyme immobilization in biotransformation processes23. ................... 3

Table 2. Inulinase and invertase immobilization studies using different types of supports. ................. 15

Table 3. Summary of the main advantages of microreactor technology63,113,115. .................................. 19

Table 4. Model enzymatic systems involving microreactors. ................................................................ 21

Table 4 (cont.). Model enzymatic systems involving microreactors. .................................................... 22

Table 5. Immobilization parameters of inulinase and invertase from commercial preparations

immobilized in two different supports: PVA beads (with 5, 4 and 3 mm thickness) and silica particles

(110 µm thickness). ............................................................................................................................... 33

Table 6. Kinetic constants for inulin hydrolysis with free and immobilized inulinase. ........................... 41

Table 7. Dispersion coefficients (D) and Péclet number (Pe) values of the three different packed-bed

microreactor systems…………………………………………………………………………………………..48

Table A1. Average bed porosity values determined for the three different packed-bed reactor

systems…………………………………………………………………………………………………………..69

xii

xiii

List of Abbreviations

BSA Bovine Serum Albumin

CLEAs Cross-Linked Enzyme Aggregates

CLECs Cross-Linked Enzyme Crystals

CSTR Continuous Stirred Tank Reactor

DNS 3,5-Dinitrosalicylic Acid

FBR Fluidized-Bed Reactor

GTA Glutaraldehyde

IE Immobilization Efficiency

KM Michaelis constant

MR Membrane Reactor

MW Molecular Weight

PBR Packed-Bed Reactor

PEG Polyethylene Glycol

PEI Polyethylenimine

PMMA Polymethylmethacrylate

PVA Polyvinyl Alcohol

RF Flow Resistance

rpm rotations per minute

RT Room Temperature

SEM Scanning Electron Microscopy

STR Stirred Tank Reactor

Vmax Maximum Reaction Rate

xiv

1

1. Introduction

1.1. Enzymes as Biocatalysts

Enzymes are proteins with catalytic function, called biological catalysts that can promote the

conversion of chemical species in living systems1. Their tridimensional structure, which is, in general,

highly complex, results from the spontaneous folding of a few polypeptide chains, acquiring, this way, a

very high efficiency and specificity2. Enzymes present specificity to certain reagents, called substrates.

Many enzymes have much higher dimensions than the substrates that they act on, and only a small

portion of the enzyme, called the active site, interacts directly with the substrate, forming a complex,

while the role of the rest of the enzyme is to maintain the correct shape of the active site3.

1.1.1. Historical Facts and Importance of Enzymes

Historically, the performance of catalytic processes through enzymatic action was first reported

during the XIX century, when the production of wine and bread in the ancient Egypt was being studied.

In 1897, Eduard Büchner, a German biochemist, proposed the term “zymase” for the entity present in

the yeast extract which was responsible for the alcoholic fermentation process, being later awarded, in

1907, with the Nobel Price of Chemistry4. The enzymatic reactions involved in the process were only

studied and elucidated in the XX century by a group of biochemists, which were also able to study the

first metabolic pathway (the glycolytic pathway or glycolysis)2,5. Several other studies were performed

long before Büchner: In 1833, Payen and Persoz recognized, for the first time, the enzyme diastase,

which is, presently, known as amylase6; In 1857, Antoine Béchamp also reported an yeast zymase,

which was in reality an invertase, responsible for the formation of inverted sugar7.

Enzymes are essential components of animals, plants and microorganisms, being involved in

catalysis and directing the complex reactions of cellular metabolism. Throughout the years, thousands

of different enzymes have been isolated and described3. These biocatalysts are characterized by their

biocompatibility, biodegradability, and renewable resources. Enzymatic processes are performed under

mild conditions (close to ambient temperature, atmospheric pressure and physiological pH) in solution,

with high rates and selectivities8.

Microorganisms are capable of synthetizing large amount of enzymes, which are responsible for

numerous functions involved in growth, metabolism, lysis processes, etc. Most of these enzymes

function inside the cell and are integrated into specific subcellular compartments, in a highly arranged

environment, while other fraction of enzymes is secreted to the extracellular environment9. This ability

to produce enzymes in large quantities in a cost-effective manner makes microorganisms the preferred

source of such catalysts. Extracellular enzymes are preferred for industrial applications as their recovery

is easier and cheaper than that of intracellular enzymes. Such enzymes have thus been a widespread

research area in Biotechnology10.

2

1.1.2. Enzyme Properties

Biocatalysis can be classified as the application of enzymes and microbes in synthetic chemistry.

The main challenge of these applications is the limited stability of the biocatalyst11.

Enzymatic processes are described as more environmental friendly, more cost-effective and,

eventually, more sustainable. Accordingly, in the past recent years, biocatalysis has emerged as an

important technology, particularly in the production of pharmaceuticals, flavors and fragrances, vitamins

and other fine chemicals. Due to their catalytic activity at mild temperatures and pressure, enzymes

have a great commercial and industrial importance8,12,13.

Advances in biotechnology led to the improvement of certain features related to biological

catalysts that can be used in several research areas, such as textile, pharmaceutical, and chemical

industries. These biocatalysts can be used in different conditions: as individual entities in solution, in

aggregates along with other molecules, and attached to some surface. This last state has been a method

of particular interest for biotechnologists14.

1.1.3. Enzyme Immobilization

Immobilization of enzymes can be defined as any technique that allows the attachment of an

enzyme to an inert and insoluble material, with retention of their catalytic activity, making it possible for

the reutilization or continuous use of the biocatalyst for long periods of time15.

Since the second half of the previous century, several methods using insoluble immobilized

enzymes have been developed for a variety of applications8.

Enzyme immobilization has numerous advantages: the linkage of an enzyme to an inert insoluble

material offers the potential for improving the enzyme’s stability; enzymes can be retained and reused,

reducing overall costs; it allows for in-situ separation of the catalyst from the products; continuous

production processes are possible; enzymes can be stabilized against heat or solvent effects3,8.

In many industrial applications, enzymes, as well as cells, need to be immobilized, through

highly effective and simple procedures, in order to allow its continuous reutilization. Enzyme

immobilization can also consist in a powerful tool for improving certain enzyme properties.

The goal of any immobilization formulation is to maximize the stability, both enzymatic and

physical, of an enzyme in a form that best suits its application. Since the physical and chemical

properties of a support significantly influence the catalytic performance, studies regarding material

sciences is required for the successful development of an immobilized enzyme catalyst16. Relevant

effects in biotransformation processes caused by immobilization of enzymes are listed in Table 1.

3

Industrial applications represent more than 80% of the global market of enzymes17. Enzymes

display excellent functional properties (high activity, specificity and selectivity), being able to catalyze

reactions under very mild conditions. These are the main reasons why enzymes consist in excellent

industrial catalysts that can be used in innumerous areas of chemical industry18,19. In enzymatic

industrial applications, certain characteristics of enzymes have to be taken into account: they are soluble

catalysts, which lead to instability; they are susceptible to inhibition by substrates or products; they are

highly effective only in physiological conditions, etc. This way, in most cases, enzymes have to be

subjected to some processes in order to substantially improve their properties. The research field of

“Enzymatic Engineering” has become one of the most complex and exciting branches of

biotechnology20,21,22.

Table 1. Key effects caused by enzyme immobilization in biotransformation processes23.

Effects of Immobilization Description

Stabilization by Immobilization

Through the introduction of several covalent

bonds and noncovalent interactions between the

enzyme and the support, the enzyme can be

stabilized due to the decrease of its flexibility,

slowing down denaturing reactions.

Mass transfer effects

The reaction rate of a biochemical process is

determined by mass transfer and kinetic effects.

Under certain conditions, mass transfer is the

limiting factor of the overall reaction rate,

whereas under other conditions the kinetics is

the rate limiting.

pH shift

The support material can have an effect on the

microenvironment of an enzyme and thereby

interfere with its catalytic activity.

Substrate and Product Partitioning

The concentration of substrates and products

present in the microenvironment of the

immobilized enzyme can be very different from

the bulk concentrations, because of the mass

transfer limitations and because of partitioning of

the substances between the bulk liquid and the

immobilized material.

Inactivation during Immobilization procedure

Immobilization methods involving the formation

of chemical bonds can lead to the involvement

of essential groups (present in the active site) in

the reaction, which normally leads to the loss of

catalytic activity.

4

1.1.3.1. Methods of Immobilization

There are several methods for attaching enzymes to a support. The enzymes can be immobilized

to a material by interactions ranging from reversible physical adsorption and ionic connection to stable

and covalent bonds. All of the immobilization approaches can be divided into two main categories:

irreversible and reversible methods (Figures 1-2)15.

Each of the immobilized methods has its own advantages and disadvantages, and none of the

techniques is ideal for all immobilizing situations. However, some problems have been overcame by

the development of protocols that use combinations of the original methods24.

Figure 1. Enzyme Immobilization Methods15.

Irreversible Methods Reversible Methods

Covalent bonding Entrapment and

Encapsulation Crosslinking Adsorption

Figure 2. Schematic representation of irreversible and reversible enzyme immobilization methods.

En

zym

e I

mm

ob

iliz

ati

on

Meth

od

s

Irreversible Methods

Covalent Bonding

Entrapment and Encapsulation

Hydrogels

Fibers

CrossLinkingCLECs

CLEAs

Reversible Methods Adsorption

Physical

Ionic bondig

Affinity binding

carrier

5

i) Irreversible Methods

In irreversible immobilization, once the biocatalyst in attached to a support, it cannot be detached

without destabilizing the support or the biological activity of the enzyme. The most common techniques

of irreversible immobilization include covalent coupling, encapsulation/entrapment, and crosslinking.

Covalent coupling is currently the most used method of immobilization. This method has the

advantage of conferring high stabilization of the bonds formed between the enzyme and the support,

thus preventing enzyme leakage. In this type of attachment, particular care must be taken to ensure that

the covalent linkage to the support does not damage the amino acid residues involved in the catalytic

activity of the enzyme24. The most common functional groups present in the support carriers used for

establishing covalent bonds with the enzymes are epoxy groups, and amino and diol groups, which

react with nucleophilic groups and primary amines of enzymes, respectively25.

Covalent immobilization methods are usually required in processes where the final product must be

enzyme free. In industry, covalent binding of an enzyme to a support results in more robust biocatalysts,

in comparison with other physical interactions, that can be applied in aqueous, multi-phase and viscous

mixtures25.

Entrapment, or Encapsulation, method is the enclosure of enzymes within gels or fibers formed

by covalent or non-covalent bonding26. The main restrictions of this technique are related with mass

transfer limitations through membranes or gels. In this technique, since the enzyme is retained rather

than bound, there are very little changes in the intrinsic catalytic properties of the enzyme3. The greatest

advantage of entrapment methods is that they are very fast and easy to perform, they are cheap, and

they usually require very mild conditions.

In the conventional entrapping of enzymes, the enzyme molecules are incorporated in a matrix formed

by physical and/or chemical gelation of pre-gel components, like monomers such as acrylamide, or a

pre-gel polymer, such as gelatin, BSA (bovine serum albumin), PVA (Polyvinyl Alcohol), alginate,

chitosan, among others27,28.

Crosslinking is another irreversible immobilization technique. This method does not require any

support or carrier to link the enzyme to prevent its leakage into the substrate solution. Crosslinking is

performed by the development of intermolecular cross-linkages between the multiple enzyme

molecules, through the use of bi- or multifunctional reagents. One of the most frequently used

crosslinking reagents is glutaraldehyde, due to economical and availability reasons29,30.

Glutaraldehyde can react with several functional groups of proteins, such as amine, phenol, thiol, and

imidazole. This way, certain conditions have to be chosen carefully, in order to favor intermolecular

crosslinking between the enzyme molecules, instead of unwanted intramolecular links that can be

developed31,32.

The crosslinking method, also called carrier-free immobilization, has several advantages, and it

can be used to overcome certain drawbacks related with carrier-based immobilization. The

6

immobilization of an enzyme using a carrier often leads to loss of a large percentage of the enzyme’s

native activity. Thus, particular interest in crosslinking for enzyme immobilization has been increasing.

There are two types of carrier-free immobilized enzymes: cross-linked enzyme crystals (CLECs); and

cross-linked enzyme aggregates (CLEAs)33.

CLECs are suitable for biotransformations that occur in non-aqueous media or in organic-water

mixtures. The performance of the CLEC obtained is very dependent on the prearranged conformation

of the enzyme molecules in the crystal framework. This system, however, is not often used since it

requires a pure enzyme, which leads to higher production costs. CLEAs are based on a multipoint

attachment through intermolecular crosslinking between several enzyme molecules, and have greater

activity in organic solvents and in both aqueous and non-aqueous media34,35,36,37,38. A major advantage

of CLEAs systems, in addition to the fact that the enzyme extract only needs to be partially purified, is

that the non-catalytic portion of the immobilizate is minimal, contrary to what happens in immobilization

processes with solid supports, in which the majority of the formulation is of non-catalytic origin (support),

whereas the catalytic part (enzyme) represents approximately 1-10% of the entire system33.

ii) Reversible Methods

The adsorption-based enzyme immobilization was one of the first enzymatic immobilization

techniques. In 1916, one of the first protocols on enzymatic adsorption was performed by Nelson and

Griffin, when it was shown that the enzyme invertase was physically adsorbed by charcoal, maintaining

its catalytic properties39,35.

Adsorption techniques, throughout the years, have been intensively studied due to its intrinsic

advantages: reversibility, which enables the reutilization of the carriers; simplicity, which allows the

enzyme immobilization under mild conditions; and high retention of catalytic activity, since there’s no

chemical modifications. This technique, due to its reversibility, it is also very economical, since the

binding support can be regenerated and re-loaded with a new enzyme solution17.

In physical adsorption, it is usually required to soak the support into an enzyme solution, followed

by incubation during a certain period of time, in order to allow time for the physical adsorption to occur.

However, one of the drawbacks of these non-covalent interactions is related with the enzyme leakage

from the matrix, in cases where the interactions are relatively weak15.

The most simple adsorption method in enzyme immobilization is the nonspecific adsorption,

which is based on the physical adsorption or ionic binding. The physical adsorption of enzymes to a

matrix can be done through hydrogen bonding, hydrophobic interactions, or Van der Waals forces,

whereas the adsorption by ionic bonding can be done through salt linkages17.

7

1.1.3.2. Advances and Applications of Enzyme Immobilization Technology

Advances regarding characterization of immobilized enzymes have been made in order to

develop new methods that allow the physical-chemical characterization of both enzymes and support,

and thus allowing a more rational and adequate immobilization procedure. Bolivar et al. (2015)

demonstrated evidences related with the analysis of protein distribution on porous solid support using

microscopic imaging methods with spatiotemporal resolution capability; there are also presented

advances in the use of spectroscopic methods for the analysis of protein conformation on solid support.

It has been discussed that, while conventional characterization of immobilized enzymes continues to

rely mostly on the apparent parameters determined from measurements in the liquid phase, extended

understanding of the behavior of the biocatalytically active solid phase is indispensable for targeted

development40.

Other surface analysis technologies for enzyme immobilization have also been demonstrated,

such as thermal gravimetric analysis (TGA), field emission scanning electron microscopy

(FESEM)/scanning electron microscopy(SEM) and transmission electron microscopy (TEM), x-ray

photoelectron spectroscopy (XPS), surface plasmon resonance (SPR) by ultraviolet detection, circular

dichroism (CD) spectroscopy, and atomic force microscopy (AFM), among many others33. All of these

studies provide valuable insights into the effects of enzyme immobilization related with the stability and

activity after treatment with different immobilization methods. It is of paramount importance to help

developing methods that improve enzyme immobilization techniques and expand the use of immobilized

enzymes for various applications in numerous fields.

A major concern of modern biotechnology is centered in the industrial application of biocatalysts.

Currently, enzyme biocatalysis is used in the following areas: Food and Agriculture Industry;

Pharmaceutical Industry; Chemical Industry; Analytical Methods; and Medical Research41. In fine and

bulk chemical industries, enzymes are proven to produce, in a more productive way, enantiomerical

pure compounds, generally through chemoselectivity, regioselectivity, and stereoselectivity. In

pharmaceutical industry, due to the growing demand of low-cost, safer and greener technologies,

biocatalysts processes have been used as an alternative to traditional chemical procedures. In food

industry, biocatalysts have been used for several applications, such as (a) to produce raw materials and

fine products, (b) to control the quality of the products, (c) to modify certain properties of some foods,

and (d) to be used as food additives41,42.

Food applications continue to dominate the application of immobilized enzymes on a volume

basis, however, emerging applications for biodiesel production and carbon capture can be ultimately

applied on a vast scale. Also, the advent of nanotechnology and integrated systems engineering

promises to move immobilized enzyme products into other areas of application, such as those involving

biosensors and intelligent materials16.

8

When working with enzymes, the intended purpose can vary: There are cases in which the

enzymatic conversion of the raw material into the desired product is the key operation (e.g.: Production

6-aminopenicillanic acid from penicillin G with immobilized penicillin acylase); and there are cases in

which the enzyme is used as an additive for modification of certain functional properties of the product

(e.g.: Production of dough using fungal proteases)17. There are several immobilized enzymes used on

an industrial scale, among which: glucose isomerase, sucrose mutase, β-galactosidase, penicillin

acylase, D-amino acid oxidase, glutaryl amidase, thermolysin, nitrilase, aminoacylase and

hydantoinases43.

As biocatalysts are the key elements in several industrial processes, constant efforts are being

made in order to improve the enzyme activity, efficiency, stability and reproducibility during industrial

procedures. Certain factors that influence enzyme immobilization and the possible modifications for their

enhancement in activity are schematically represented in Figure 3.

Figure 3. Determinant factors involved in enzymatic immobilization and activity27.

The use of immobilized enzymes as biocatalyst will continue to be of special interest for industries

as the technique is highly efficient, as well as environmental friendly33. Commercialization of immobilized

enzymes is, however, limited by certain factors, such as costs and storage problems. Research should

be performed in order to overcome the current restrictions related to immobilization techniques, so as

to expand the branch of applications.

Enhancement of immobilized

enzyme's activity

Micro-environment

effect

Partition effect

Diffusion effect

Change in conformation

Binding method

Molecular orientation

Conformation Plasticity

9

1.1.3.3. Immobilization Supports: PVA and Silica

i) Polyvinyl-Alcohol (PVA)-based Hydrogel

Hydrogel consists of a polymeric material that possesses the ability to swell and retain a

substantial amount of water within its structure. Over the past 50 years, this type of material received

significant attention, due to its wide range of applications. Hydrogel-based materials present a high

degree of flexibility, very similar to natural tissue due to the high water content44.

A great variety of polymer compositions have been used to produce hydrogels. These

compositions can be divided into natural polymer hydrogels (e.g.: chitosan, collagen, dextran), synthetic

polymer hydrogels, e.g.: poly(viny alcohol), poly(vinyl acetate), poly(glucosylethyl methacrylate)) and

combinations of the two classes (e.g.: poly(ethylene glycol)-co-peptides, collagen-acrylate45.

Currently, poly(vinyl alcohol) (PVA) is one of the most frequently used synthetic polymer

hydrogels. PVA is a polymer of great interest due to its various desirable properties specifically for

several pharmaceuticals and biomedical applications46.

PVA is defined by a relatively simple chemical structure with a pendant hydroxyl group (Figure

4). This material is prepared from polyvinyl acetate by alcoholysis since vinyl alcohol cannot be

isolated46,47. The hydrolysis reaction is not entirely completed, resulting in polymers with a certain degree

of hydrolysis, which depends on the extent of the reaction. Thus, there are several commercially

polymers available that differ, for example, in the residual acetate content.

Figure 4. Chemical structure of polyvinyl alcohol (partially hydrolyzed)48.

PVA can be crosslinked in order to be used in a wide variety of applications, such as in medical

and pharmaceutical sciences. Some of the regular crosslinking agents that have already been used for

PVA hydrogel preparation include: glutaraldehyde, acetaldehyde, formaldehyde, among others.

However, to avoid these crosslinking procedures which can possibly lead to the release of toxic

substances, physical methods, such as gelation through multiple freeze-thaw processes which lead to

the formation of hydrogen bonds, have been demonstrated. One of the drawbacks of the freeze-thaw

method is the significant amount of energy required, reason why several other techniques for PVA

gelation have been developed, such as polymer formations with boric acid, and Lentikats®

technology46,49.

10

Lentikat® technology is the most suitable protocol performed at laboratory temperature, without

the use of any toxic and expensive materials. This protocol uses mild conditions of gelation to produce

hydrogel carriers with different shapes and sizes. The most typical carrier has a spherical shape, which

allows easy diffusional effects. Nevertheless, if the formed beads have a large diameter, there will be

diffusional limitation in the center of the biocatalyst, and, on the other hand, if the beads are too small,

some clogging problems may occur in the outlet of the reactor vessel. PVA particles contain a higher

internal porosity and thus high immobilization capacity. Besides, these carries present an excellent

physical and mechanical stability46,50,51,52.

Several experimental procedures have been developed using PVA as an immobilization support.

Fernandes et al. (2009) evaluated the enzyme immobilization of inulinase in PVA particles through

extrusion of LentiKat®Liquid directly in polyethylene glycol (PEG), for the production of suitable

biocatalysts for large-scale applications53. Poraj-Kobielska et al. (2015) demonstrated an effective

method of encapsulation of peroxygenase (UPO) in PVA/PEG-gel beads54. Bilal and Asgher (2015)

immobilized purified manganese peroxidases (MnPs) enzymes onto PVA-alginate beads to investigate

the enzyme’s potential for decolorization and detoxification of a new class of reactive dies and textile

wastewater55. Saallah et al. (2016) demonstrated the immobilization of cyclodextrin glucanotransferase

(CGTase) on nanofibers, in which CGTase solution and PVA solution were mixed followed by an

electrospinning step56. In Nunes et al. (2012), studies were carried out in order to evaluate and optimize

the immobilization of naringinase from P. decumbens in PVA networks, targeting for the hydrolysis of

naringin57. Nunes et al. (2016) developed PVA electrospun nanofibers using an electrospinning

technique58, and also reported an innovative and reproducible system methodology for the production

of customized 3D hollow microspheres of PVA using aromatic boronic acids (ABAs) as cross linkers59.

ii) Silica-based Support

Regarding covalent immobilization methods, several chemical functional groups can be inserted

into the surface of the support material (e.g.: –NH2, alcoholic –OH, –COOH, –SH) allowing the covalent

reaction with enzymes under appropriate conditions. Usually, adhesion of enzymes to a support by

covalent bonding requires a step of functionalization and activation. Conceptually, functionalization of

the support is the procedure by which a new chemical function is introduced in the carrier’s surface. In

turn, the activation step is the process by which the newly introduced chemical function is made reactive

towards the enzyme to be immobilized, by action of specific activating agents, such as glutaraldehyde60.

The scheme representing the covalent immobilization on inorganic support relevant for the present

review is summarized in Figure 5.

11

Figure 5. Schematics of the functionalization and activation of inorganic support during covalent immobilization,

by action of polyethylenimine (PEI) and glutaraldehyde (GTA), respectively.

Several silica-based materials are widely considered when working with inorganic supports. This

type of support presents all the crucial requirements of a support suitable for enzyme immobilization,

such as: is relatively cheap and environmentally harmless; presents excellent thermal, mechanical and

microbial resistance; and is totally inert under enzymatic operational conditions60,61. The three main

procedures used for immobilizing enzymes in porous silica supports are: physical adsorption, covalent

attachment onto the pore walls and cross-linking of enzyme molecules62.

Silica, also called silicon dioxide (SiO4), normally exists as a 3-D polymer, formed by units of

regular SiO4 tetrahedra. The entire structure, highly rigid, is defined by an infinite lattice constituted by

siloxane bridges (Si-O-Si). Silica is an abundant compound in nature. Various crystalline and

noncrystalline silica minerals of inorganic and biogenic origin are well known. Depending on the

environmental conditions, such as temperature and pressure, several different crystalline alterations

can be formed. Silicon dioxide can be found mainly as quartz, flint, sand, and several other

minerals60,63,64. Silica-based supports commercially available may have various pore diameters.

However, the most promising materials are mesoporous supports, which have pores ranging from 2 and

50 nm in size60. Mesoporous silica nanoparticles (MSNs) have been extensively used in therapeutics,

pharmaceuticals and diagnosis and, in addition, they have been proven to be favorable supports for

enzyme immobilization, enabling the stabilization of biocatalysts and the retention of its catalytic

activity61,65.

Iyer et al. (2003) performed an immobilization of cyclodextrin glucanotransferase (CGTase) from

a Bacillus circulans strain on purified sea sand, using glutaraldehyde as crosslinking agent66. The sea

sand used in enzyme immobilization procedures needs to be extra pure. Purified sea sand consists on

crude silica which is prepared by neutralization of aqueous alkali metal silicate with acid and further

calcined in order to remove trace metals66.

12

Hung et al. (2015) developed a carrier to be used in enzyme prodrug therapy by immobilizing

horseradish peroxidase (HRP) onto mesoporous silica nanoparticles, in which the functionalization of

the nanoparticle surfaces was made with 3-aminopropryltrimethoxysilane and the activation with

glutaraldehyde67. Du et al. (2015) and Gustafsson et al. (2015) performed co-immobilization procedures

using silica-based supports. In both works, it was possible to form cascade reactions involving two

immobilized enzymes, while maintaining the stability of the biocatalysts. In Du et al. (2015), the system

consisted of a silica microsphere core coated with two layers of individually immobilized enzymes (α-

amylase, AA; and glucoamylase, GluA), which were immobilized on carbon nanotubes (CNTs). In

Gustafsson et al. (2015), Aspergillus sp. glucose oxidase (GOD) and horseradish peroxidase (HRP)

were both immobilized on solid silica supports by using mesoporous silica nanoparticles and a

polycationic dendronized polymer. In general, the two different enzyme immobilization methodologies

allowed a spatially controlled co-immobilization using different types of enzymes on solid support68,69.

1.2. Enzymatic model systems

1.2.1. Inulin hydrolysis to reducing sugars

Inulin is a fructose polymer found in many species of Compositae and Gramineae, for example,

in tubers of Jerusalem artichoke, dahlia, roots of dandelion and chicory, that has a great content of D-

fructose (> 75%), being recognized as an ideal source for the production of high fructose and fructo-

oligosaccharide syrup70.

Fructose and fructo-oligosaccharides are highly important ingredients in the food, beverage and

pharmaceutical industry, due to their beneficial effects in diabetic patients, increase iron absorption in

children, low cariogenicity and higher sweetening capacity than sucrose (up to 1.5 times). The

sweetening capacity of fructo-oligosaccharides is, however, relatively low in comparison with fructose,

nonetheless it presents a strong prebiotic action. Additionally, these components enhance flavor and

color, as well as product stability in food and beverages. Regarding the therapeutic advantages, fructose

metabolism circumvents the well-known metabolic pathway of glucose, and thus, does not require

insulin71. Fructose has a very low GI (glycemic index), being therefore recommended that fructose might

be therapeutically useful as a dietary supplement for patients with type 2 diabetes mellitus72.

Inulin is hydrolyzed by enzymes known as inulinases. These enzymes can be classified into endo-

and exo-inulinases, depending on their mode of action. Endo-Inulinases (EC 3.2.1.7; 2,1-β-D-fructan

fructanohydrolase) are specific for inulin and can hydrolyze it by breaking the interior bonds of the

fructose units, leading to the production of oligosaccharides. Exo-inulinases (EC 3.2.1.80; β-D-

fructohydrolase) are capable of splitting terminal fructose units in inulin, sucrose and raffinose to release

fructose (Figure 6)73.

13

Inulinases can be found in plants and in many microorganisms (fungi: Aspergillus niger, A. ficuum,

Chrysosporium pannorum and Penicillium purpurogenum; Yeasts: Kluyveromyces marxianus, Candida

kefyr, Debaryomyces cantarellii and Pichia polymorpha)73.

The hydrolysis of inulin occurs in a single enzymatic step, yielding up to 95% fructose. This

process is quite simple and the substrate can be easily obtained. However, inulin has limited solubility

at ambient temperature, which increases the possibility of microbial contamination of substrate and

product. There are several methods that help increasing thermal stability of an enzyme, like through

enzyme immobilization, for example70.

Several methods regarding inulin immobilization, using different types of supports, have been

described: such as immobilization by covalent adsorption with chitosan beads70; by entrapment, using

polyvinyl alcohol (PVA) particles74 and other gel matrixes, and by crosslinking using activated

Amberlite71.

Figure 6. Schematic representation of the action of endo-inulinase and exo-inulinase on inulin, for the production

of fructo-oligosaccharide syrup and ultra-high fructose syrup, respectively75.

1.2.2. Sucrose hydrolysis to reducing sugars

Sucrose can be converted to fructose and glucose through hydrolysis reaction (Figure 7).

Invertase, or β-D-fructofuranosidase (EC 3.2.1.26) is an enzyme that can be found in Saccharomyces

cerevisiae, among other sources, and that presents major specificity for sucrose hydrolysis.

Invertase hydrolyzes sucrose by hydrolyzing the terminal non-reducing β-fructofuranoside

residues in β-fructofuranosides. This reaction results in the formation of an equimolar mixture of α-D-

glucose and β-D-fructose, known as invert syrup76. Invert sugar is also widely used in confectionery and

14

baking. This is due to its ability to retain moisture, hence reducing the trend of baked goods to dry and

stale. Additionally, invert sugar minimizes crystallization, contributing to retain softness of goods. Both

features contribute for enhanced shelf life77,78.

Immobilization of invertase has shown to have great biotechnological potential due to the

relatively inexpensive costs of its substrate, and also due to its importance in food and drink industry77,79.

Regarding the products that are generated in sucrose hydrolysis, glucose is a highly beneficial

product to human lives and environment, being also advantageous towards sustainable technology.

Glucose can be used for several purposes: such as feed stock for farm animals (cows, goats, and

sheep), artificial flavoring in food and confectionary; it can also be used, in a purified form, as a precursor

to make vitamin C (L-ascorbic acid), and as a substrate production of citric acid, gluconic acid,

bioethanol, sorbitol, and polylactic acid; Additionally, glucose can also be used in medication therapy,

as a component in intravenous drips80. Fructose is also derived from the digestion of sucrose, and is the

sweetest naturally occurring sugar estimated to be, nearly, twice as sweet as sucrose.

Figure 7. Schematic representation of the action of invertase on sucrose for the production of α-D-glucose and β-

D-fructose81.

The protocol to hydrolyze sucrose is usually carried out using acids at high temperature that leads

to dark coloration at the end of the process due to the formation of by-products called furfurals. These

by-products can be dangerous to the health when ingested with food or drinks. To overcome this

drawback, immobilization of invertase could be a good alternative, allowing the production, at mild

conditions, of invert sugars syrup free of furfurals77,82.

Several methods of invertase immobilization, using different types of supports, have been

described: immobilization with montmorillonite by covalent bonds; with resin by adsorption; with gelatin

hydro-gel by entrapment; with several types of biopolymers, such as porous cellulose beads and

carbohydrate moieties77,79.

15

1.2.3. Inulinase and invertase immobilization procedures – review

Both inulinase and invertase have been extensively used in several immobilization studies. Table

2 lists a few examples of studies using these two biocatalysts immobilized in different types of carriers.

Table 2. Inulinase and invertase immobilization studies using different types of supports.

Enzymes Type of enzyme Immobilization

support Reference

Inulinase

Commercial preparation of inulinases

(Fructozyme L) from A. niger Amberlite [71]

Commercial preparation of inulinases

(Novozym 230) from A. ficuum Porous glass beads [83]

Endo-inulinase preparation from A.

niger Chitin carrier [84]

Commercial preparation of inulinases

from A. niger

Alginate-chitosan

beads [53]

Commercial preparation of inulinases

from A. niger

Polyvinyl alcohol

capsules (LentiKats®) [85]

Invertase

Commercial preparation of invertase

(Maxinvert L 10000) from S.

cerevisiae

Silane-coated silica

carriers [86]

Invertase from baker’s yeast

Micro-porous acid-

activated

montmorillonite clay

[87]

Invertase from S. cerevisiae Celite and

polyacrylamide [88]

Invertase from S. cerevisiae Supermacroporous

polyacrylamide cryogel [89]

Invertase from baker’s yeast Polyvinyl alcohol

capsules (LentiKats®) [90]

1.3. Immobilized Enzyme Reactors

The major limiting factor of using enzymes in large-scale industrial applications is the process

economics, since the isolation and purification of enzymes for specific reactions is, usually, expensive.

Hence, the use of reaction engineering principles for the design and optimization of immobilized enzyme

reactors is essential91.

16

For their efficient utilization, the enzyme-carrier conjugates need to be properly characterized in

a suitable reactor environment. In order to predict the reactor performance, some factors need to be

considered, such as the activity of the fixed biocatalyst, the reactor configuration, mass transfer effects

adjacent to the solid phase, the electrostatic partitioning between the fluid and carrier phases, and the

residence time distribution within the reactor92.

1.3.1. Modes of Operation – Batch and Continuous mode

In process-scale operations, enzyme reactors can operate in batch mode or continuously, using

enzymes in their free or immobilized form.

In batch-mode operations, the enzyme preparation is loaded into the reactor system, generally a

stirred tank reactor (STR), along with the substrate, being the reaction carried out until a chosen degree

of conversion is reached. Batch reactors are usually preferred in cases in which free enzymes are used

as biocatalysts. In these cases, there is usually no recovery of the enzyme in the end of the operation.

Furthermore, when immobilized enzymes are used in batch-mode reactions, the immobilized

biocatalysts should be separated from the product stream though a subsequent purification step, such

as filtration or ultracentrifugation. However, most of the recovery procedures might lead to loss of

immobilized enzyme and it can cause, additionally, inactivation of the enzymes after repeated cycles of

purification, which leads to low productivities93.

In continuous mode operations, the reactants are continuously fed into the reactor system, and

the product stream continuously retrieved. Continuous reactions are more productive comparatively to

batch processes, mostly due to the lower frequency of unproductive time94, being used mainly for large-

scale production processes. In this type of reactors, immobilized enzymes are used as biocatalysts,

since the enzymes should be highly stable in order to perform continuous processes95.

There is no ideal reactor configuration for operating immobilized enzymes. However, the design

of a suitable and efficient bioreactor is the most critical factor in this technology. There are several

reactor configurations that have been proposed for developing immobilized enzyme reactions in

continuous processes. Among these configurations are (a) the continuous stirred tank reactor (CSTR),

(b) the packed-bed or plug flow reactor (PBR), (c) the fluidized bed reactor (FBR), and (d) the membrane

reactor (Figure 8). The choice of the most suitable reactor must be conditioned by the temperature

conditions, flow rates, pH control, ionic strength, substrate concentration, diffusion effects, etc96.

(a) Continuous Stirred-Tank Reactor (CSTR)

The CSTR is commonly used in industrial processes. This type of reactor is usually operated at

steady-state and is presumed to be perfectly mixed. Consequently, every variable (temperature,

concentration and reaction rate) is the same within the reactor vessel95. This type of reactor, besides

enabling an easy control of the temperature, it also provides a long residence time, due to the large

17

volume. The major drawbacks of the CSTR is the low bioconversion yields per unit volume and the low

heat transfer per unit area.

(b) Packed-bed Reactor (PBR)

Packed-bed reactors consist, typically, of a cylindrical column filled with a certain packing

material. This type of reactor is normally operated at steady-state, and the reactants are continually

consumed as they flow down the length of the reactor. The liquid passing through the column should be

distributed as uniformly as possible in order to be in contact with most part the packing material95,97. One

important factor to take into account in packed-bed systems, from a mechanical point of view, is that of

the pressure drop required for the liquid to flow through the column at a specific flow rate. The porosity

of the reactors’ bed can be affected by the packing mode, the ratio between column to particle diameter,

the particle shape, the particle size distribution and the roughness of the particle surface, and also by

the bed height. Concerning the pressure drop across a packed bed, the research carried out so far show

that pressure drop is dependent on: the fluid velocity, the physical properties of the fluid (viscosity and

density), the average porosity of the bed, the orientation of the packing, the size, the shape and the

surface of the particles, the ratio between the particle and the contained diameters (size wall effect), and

the ratio between the height of the packed bed and the diameter of the particles98.

(c) Fluidized-bed Reactor (FBR)

The fluidized-bed reactor can be defined as a hybrid of the CSTR and the PBR. This type of

reactor consists of a packed-bed cylindrical column through which fluid flows, at high velocities, making

the reactors’ bed loosened. In this case, the particles are held suspended by the fluid stream and, in

many cases, the fluidized-bed can behave like a liquid. The main advantages of the fluidized-bed reactor

system include (i) the uniform temperature distribution, (ii) the uniform production in batch-wise

processes due to the strong solid mixing, and (iii) the great solid-liquid/gas exchange area due to the

small solids grain size97,99.

(d) Membrane Reactor (MR)

The membrane reactor is another type of reactor capable of performing continuous operations.

The main characteristic of an enzymatic membrane reactor is the separation of enzymes from products

and/or substrates through a semi-permeable membrane. This selective separation can be achieved by

action of a driving force across the membrane (pressure, chemical potential, electric field) which

subsequently causes the movement of solutes100. In an enzymatic membrane device, the biocatalyst is

physically retained by the membrane, allowing the continuous reaction operation. The most common

type of MR is the hollow-fiber reactor, which consists of a module containing several thin tubular fibers

in a parallel array with inlet and outlet ports101. In MR systems, the main disadvantages are related with

enzyme leaking, polarization layer and induced limitation, and membrane fouling102.

18

Figure 8. Examples of bioreactor systems used in continuous bioconversion processes involving free or

immobilized enzymes103,104.

1.3.2. Enzymatic Microreactors

1.3.2.1. Principles of Enzymatic Microreactors

Currently, in chemical engineering, there is a wide diversity of reactors that can be used for gas-

liquid, liquid-solid and gas-liquid-solid reactions. The choice of reactor to be used depends on several

criteria, such as production output, reaction rate, residence time, volumetric ration of phases, heating-

cooling requirements, ease of scale-up, etc.105.

Throughout the years, the interest in performing chemical conversions using flow microreactors

has been growing, replacing traditional flasks or stirred vessels that operate in batch mode. These

microscale reactors require low input volumes, which is a major advantage since it allows the

development of rapid and simple assays, in a highly automated and continuous operation. Another

advantage of using flow microreactors is the easiness of transition from laboratory to industrial

production scale operation106,107.

Microreactors are normally classified as miniaturized reaction systems that are often fabricated

by methods using micro-technology and precision engineering108,109. Due to microreaction technology,

which allows the miniaturization of conventional macroscopic reactors, a new branch of reactors using

micro-engineered features has become available. Microreaction technology has been used in numerous

applications in chemical engineering, and has led to the development of different types of reactors, such

as thin-wall microreactor, membrane microreactor, packed-bed microreactor, single phase microreactor

and multiphase microreactor. Most of these microreactors are still in the research level, although a few

of them are already commercially available105,110.

Polymer-based materials can be used for the fabrication of enzyme microreactors, such as

polydimethylsiloxane (PDMS), polymethylmethacrylate (PMMA), polycarbonate and Teflon. This type of

(A) Continuous Stirred Tank Reactor

(CSTR)

(C) Fluidized-Bed Reactor (FBR)

(B) Packed-Bed Reactor (PBR)

(D) Membrane Reactor (MBR)

19

materials are highly stable in aqueous environment. Stainless steel is another type of material used for

building microreactor devices. This is a well-established material, resistant to organic solvents and to a

wide range of chemicals, and enables operations under pressure and high temperatures111,112.

1.3.2.2. Microfluidic Reactor Technology in Industry

Microfluidics has become a very attractive field to industry in the context of continuous flow

synthesis operation. The use of these microreactors in biocatalytic processing has several potential

advantages owing to their inherent properties, such as: cost reductions, in comparison to the other

technologies which are often performed in batch mode; efficient heat and mass transfer; rapid mixing,

greater control of reaction conditions; capability of continuous and integrated operations at high

pressures; and greener chemistry113,114. In the table below (Table 3), the main advantages of

microreaction technology are summarized.

Table 3. Summary of the main advantages of microreactor technology63,113,115.

Advantages Description

Cost reduction The operation costs are decreased by the low

energy demand and low material input.

Selectivity

It is possible to control the selectivity of the

reaction over a wide range of conditions with

high precision (e.g.: temperature, residence

time, etc.).

Green chemistry

Microfluidic reactor systems decrease the

energy requirements, increase throughput per

unit area, decrease reagent consumption, and

use less hazardous chemicals.

Rapid reactions Small reaction streams result in rapid diffusive

mixing.

Safe reactions Microfluidic reactors provide safer environments

in which to perform hazardous chemistry.

Continuous operation possible Continuous-flow microfluidic systems confer

great efficiency to different chemical reactions.

20

1.3.2.3. Process Intensification

Usually, when working with microreactors it is necessary to consider the process intensification,

which can be defined as the great reduction (≥100 fold) of the physical size of a reactor, allowing the

achievement of a given production objective, and the improvement of some key features related to

chemical processing (such as yield and selectivity)105,114,116. The decrease of physical size in

microreactors has a great impact on the intensification of mass and hear transfer, which allows the

improvement of flow patterns. Therefore, beneficial aspects concerning chemical engineering is the

main driving factor for microreactor based researches108.

In cases in which we consider process intensification, the microscale process has to be rigorously

compared with the conventional methods. One of the preferred types of reactor used in industrial

bioprocess production is the fixed-bed immobilized enzyme reactor43.

A distinctive feature of microreactor fluidic devices is the capability of having multiple basic units

operating in parallel, either fed separately, or using a common fed line, for production purposes. An

industrial plant design based on a large number of small reaction unit systems is capable, theoretically,

of performing a variety of reactions by changing the piping network, i.e., the pant can be adapted to

perform the synthesis of several substances using microreactor modules. The scale-up through

numbering up reactor units leading to a reduction of large investments, by removing costly reactor

redesign and pilot plant experiments, as well as shortening the process time from laboratory to

commercial production continues to be a wishful thinking105,108.

1.3.2.4. Applications of Microreaction Technology

Microreactors are widely characterized as important engineering tools for chemical process

research and development. The use of highly automated and fully instrumented microstructured reactors

simplifies the catalyst and/or reaction screening; the evaluation of immobilized catalyst; the analysis of

intrinsic reaction kinetics, in order to be used later for process optimization; the scale-up and process

control; the detailed characterization of transport effects on effective kinetic; and the determination of

enzyme operational stability117.

Miniaturized reactors enable, mainly, a faster transfer of research results into production, due to

their advantages of operating conditions, and due to the fact that they are capable of yielding more

precise laboratory data, in comparison with conventional reactors.

Different types of enzymatic microreactors can be used for different purposes. Successful

applications of these devices include analysis of chemical species, such as proteins and nucleic acids,

studies of model enzymatic systems, and kinetic studies118.

21

A variety of immobilization techniques and micro-fluidic strategies have been used to develop

enzymatic microreactors intended for the use of analysis of chemical species, determination of enzyme

kinetics, among other applications (Table 4).

Table 4. Model enzymatic systems involving microreactors.

Immobilized Enzyme/Cells

System Application References

Trypsin

Porous Polymer Monoliths molded in

channels of microfluidic devices

Protein Mapping [119]

Trypsin

Reactor system based on organic-

inorganic hybrid silica monoliths

Measurement of enzymatic activity

[120]

Trypsin

Poly(methyl methacrylate)

(PMMA) microchannel surface

Protein Patterning [121]

Horseradish peroxidase;

β-Galactosidase

Microbeads packed into a chip-based

microreactor

Determination of enzyme kinetics

[122]

β-Galactosidase

Droplet-based microfluidic system with a concentration

gradient

Enzyme kinetics measurements

[123]

Aminoacylase

Polymeric membrane on the inner wall of a microchannel surface through cross-linking polymerization in a laminar flow using

poly-L-lysine

Simplification of a procedure of enzyme

immobilization in a microreactor

[30]

Urease

Glass beads coated with polyacrylonitrile

layer and Porous glass beads in a

microsystem using low temperature co-

fired ceramics (LTCCs) technology

Measurement of urea concentration with

high output signal and large scale

[124]

Lipase Continuous operated

pressure-driven microreactor

Isoamyl acetate synthesis

[125]

Angiotensin-converting enzyme

(ACE)

On-column enzyme microreactor created by an ionic binding

technique

Screening of enzyme inhibitors

[126]

22

Table 4 (cont.). Model enzymatic systems involving microreactors.

Immobilized Enzyme/Cells

System Application References

Glucose oxidase (GO)

On-chip enzyme immobilized monolith

microreactor that integrates a microfluidic

electrochemical cell

Rapid characterization of enzymatic kinetics

[127]

Glucose-, alcohol-, lactate-, galactose-

and l-amino acid oxidases (GO, AO,

LacO, GalO and LAAO)

Bioenzymatic analytical

microreactors integrated in a flow injection analysis

(FIA) system

Bioprocess monitoring

[128]

E. coli cells

Hydrogel micropatches

photopolymerized within a microfluidic

system

Intracellular enzyme reactions

[129]

23

2. Objectives

The main goal of the current research project is focused on the development and implementation

of miniaturized enzymatic reactors with different configurations, targeting for bioprocess intensification.

These devices are known for speeding up bioprocess development and, moreover, lead to significant

cost reductions given the low amount of consumables required. The identification of proper operational

conditions which maximize volumetric productivities is also intended.

In order to study the feasibility of the different microreactor configurations, the enzymatic

hydrolysis of inulin and sucrose to reducing sugars, performed by inulinase and invertase, respectively,

were chosen as model systems, since these are processes of industrial relevance. However, inulin

hydrolysis was considered to be the preferred model system. Both enzymes were immobilized onto two

different types of particles and used in microreactors with packed-bed configurations, which were

operated in continuous mode.

24

25

3. Materials and Methods

3.1. Materials

Fructanase Mixture (purified liquid), a commercial preparation of inulinase, from an Aspergillus

niger strain, was acquired from Megazyme (U.S.A.). Maxinvert® L 10,000 invertase enzyme, from a

selected strain of Saccharomyces cerevisiae, was a kind gift of DSM (Heerlen, Netherlands). The

substrates Inulin from chicory (Fibruline® S20), with an average polymerization degree of about 10, was

obtained from COSUCRA (Warcoing, Belgium), and sucrose (D(+) Saccharose) was obtained from

Fisher Scientific UK (Bishop Meadow Road, Loughborough). Poly(vinyl alcohol) (PVA) (Lentikat®Liquid

250) was obtained from GeniaLab® (Braunschweig, Germany). Poly(ethylene glycol), with average M.W.

600 (PEG 600) was acquired from Acros Organics (Geel, Belgium). Bradford reagent (for 1-1,400 µg/ml

protein) was obtained from Sigma-Aldrich (St Louis, U.S.A.). Polyethylenimine 50.0% aqueous solution

and glutaraldehyde 24.0% (w/w) were acquired from Sigma-Aldrich and Acros Organics, respectively.

The immobilization procedure using PVA hydrogel, as well as the continuous operation assays,

were performed using peristaltic pumps from Watson Marlow (200 series multi-channel cassette pumps)

with Tygon® Tubing. The control of environment’s temperature was promoted by a Thermomix® MM

thermostat. The heating of solutions was performed by IKA® HBR4 heating bath, 230V. For

quantification analysis, a SPECTROstar Nano spectrophotometer (Microplate reader), from BMG

LabTech was used. Analytical balance Mettler Toledo PJ300 was used for measuring purposes. A

Titramax 1000 Vibrating platform shaker from Heidolph was used for microplate shaking. For continuous

reaction operations, three microreactor systems, with packed-bed configurations, were used: a reactor

adapted from a cylindrical glass column with a bed height of 4.00 cm, inner diameter (I.D.) of 0.60 cm,

and wall thickness of 0.25 cm (Figure 9A); a cylindrical steel microreactor with a bed height of 2.60 cm,

I.D. 1.00 cm and wall thickness of 0.20 cm (Figure 9B); and a microreactor composed of a Teflon layer

with bed height of 0.10 cm, width 0.30 cm and length 9.30 cm, tightly squeezed with two poly(methyl

methacrylate) (PMMA) plates. This will briefly referred to as PMMA-microfluidic reactor (Figure 9C).

26

Figure 9. The three microreactor configurations used for continuous flow operations: cylindrical glass column (A),

cylindrical steel microreactor (B), and PMMA-microfluidic reactor (C).

3.2. Methods

3.2.1. Inulinase and invertase Immobilization

In the present work, inulinase and invertase enzymes were immobilized in a support through two

different immobilization procedures: enzyme entrapment into a hydrogel matrix and enzyme adsorption

to a solid support with crosslinking agent.

In the entrapment immobilization procedure, inulinase and invertase were immobilized in PVA.

The Lentikat®Liquid, due to its physical chemical properties, had to be heated at 95⁰C until it was

completely dissolved, and then cooled to 50⁰C. The enzyme solution (1 mL of 5-fold diluted enzyme

preparation in acetate buffer 0.1 M, pH 4.5) was added to 5 mL of Lentikat®Liquid, and mixed using

magnetic stirring. The enzyme preparation was then extruded, through a peristaltic pump, into a solution

of 50 mL of PEG 600, under mild magnetic stirring conditions. Through this procedure, enzyme beads

were produced, with diameters ranging from 3 and 5 mm (Figure 10A), which was possible by adjusting

the diameter of the outlet tubing system. After a 2 hour period, the formed capsules were harvested,

with the help of a drainer, and washed with acetate buffer 0.1 M, pH 4.5. The used PEG 600 solution

was stored in a 50 ml Falcon tube, at 4⁰C for further analysis. The capsules stayed in the buffer solution,

under mild stirring conditions, for about 30 minutes. After this, the beads were weighted and finally

stored soaked in acetate buffer at 4⁰C until further use. The used acetate buffer solution was stored in

a 50 ml Falcon tube, at 4⁰C for further analysis.

In the covalent immobilization procedure, inulinase and invertase were immobilized in silica sea

sand extra pure, purified by acid, from Carl Roth®. First, 10 g of the silica particles were washed in 100

mL of distilled water for approximately one hour and then vacuum filtered. This last procedure was

repeated using acetate buffer 0.1 M pH 4.5 instead. Following the wash of the support, this was

incubated, during a 2 hour-period at room temperature (RT), with 25 mL of polyethylenimine 10.0% (v/v).

The support particles were recovered through vacuum filtration and activated through incubation, with

A. B.

C.

27

50 mL of glutaraldehyde 10.0% (v/v) for approximately 18 hours at RT. After this period, the particles

were vacuum filtrated and washed, once again, with acetate buffer 0.1 M pH 4.5. The washed particles

were then incubated with 10 mL of diluted inulinase and invertase (10-fold in acetate buffer 0.1 M pH

4.5) at 4⁰C during 2 hours. Finally, the support containing the immobilized enzyme was vacuum filtered,

washed with acetate buffer 0.1 M pH 4.5 and stored at 4⁰C until further use. All the incubation steps

were performed under magnetic stirring.

In both enzyme immobilization procedures, the washing solutions used to wash the support with

the immobilized enzymes were stored at 4⁰C for further analysis. The efficiency of the immobilization

procedure was determined by the following equation (Equation 1).

𝐼𝑚𝑚𝑜𝑏𝑖𝑙𝑖𝑧𝑎𝑡𝑖𝑜𝑛 𝐸𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑐𝑦 (%) =𝐸𝑖𝑠 − 𝐸𝑤

𝐸𝑖𝑠

× 100

[Equation 1]

Where 𝐸𝑖𝑠 refers to the amount of enzyme present in the inorganic salt solution used to prepare the

enzyme preparation, and 𝐸𝑤 refers to the amount of enzyme present in the washing solutions.

In order to determine the enzymatic activity of the immobilized systems in terms of relative activity

(%), the following equation was used (Equation 2).

𝑅𝑒𝑙𝑎𝑡𝑖𝑣𝑒 𝐴𝑐𝑡𝑖𝑣𝑖𝑡𝑦 (%) =𝐴𝑐𝑡𝑖𝑣𝑖𝑡𝑦

𝑀𝑎𝑥𝑖𝑚𝑢𝑚 𝐴𝑐𝑡𝑖𝑣𝑖𝑡𝑦× 100

[Equation 2]

Figure 10. Biocatalytic PVA beads with diameters ranging between 3 and 5 mm (A), and biocatalytic silica particles with thickness 110 µm (B).

A1.

A2.

B.

28

3.2.2. Scanning electron microscopy (SEM)

For the analysis of the physical structure of the sea sand-silica particles, before and after the

immobilization procedure, these were placed on a double carbon tape and analyzed in a field emission

scanning electron microscope (Jeol JSM-7001F), as described elsewhere130.

3.2.3. Batch assays

In order to test the enzymatic activity of the immobilized systems, a chosen amount of enzyme

preparation was added to 1 mL of substrate solution – In each assay, in the case of the hydrogel enzyme

systems, 3 PVA beads (with the same diameter) were used, and in the case of the crosslinked

immobilized systems, roughly 70 mg of silica particles were used. The immobilized inulinase was added

to inulin solution 5.0% in 0.1 M pH 4.5 acetate buffer, and the immobilized invertase was added to

sucrose solution 5.0% in 0.1 M pH 4.5 acetate buffer. The biocatalysts were incubated, at 50⁰C, for

approximately 1 hour, under stirring (420 rpm) using micro-stirrer bars. The assays were performed in

triplicate, and the aliquots were collected and immediately analyzed by DNS method for quantification

of reducing sugars.

3.2.4. pH and temperature profiles of free and immobilized enzymes

In order to establish the effect of the pH and temperature in the enzymatic activity of free and

immobilized enzymes, batch runs were performed in a pH range of 3.5 to 5.5 and in a temperature range

of 40 to 60⁰C. Inulin and sucrose solutions 5.0% (w/v) in acetate buffer 0.1 M were used as substrates.

The runs were performed in 4 mL glass vials, using 1 mL of substrate solution. The agitation was

promoted by micro magnetic stirring (420 nm) and samples were collected after 10 minutes (free

enzymes) and 60 minutes (immobilized enzymes), being afterwards analyzed by the DNS method for

reducing sugars quantification. All trials were performed in triplicate.

3.2.5. Determination of the kinetic parameters

The kinetic parameters studies for the PVA-immobilization system were not performed in this

work, due to the existence of several published studies performed on the subject.

For the determination of the kinetic parameters of the free and immobilized inulinase in silica

particles, batch runs were performed in the presence of inulin solutions with concentrations ranging from

1.0 to 12.0% (w/v), and the corresponding initial reaction rates were evaluated. The solutions were

prepared in acetate buffer 0.1 M pH 4.5 and the assays were performed at 50°C. Each assay lasted 30

minutes, and samples of 100 µL were collected every 5 minutes, until a maximum of 12.0% substrate

conversion was reached. The enzyme activity for the free and immobilized forms was determined

29

through the initial reaction rates, and the respective kinetic parameters, Vmax and KM, were obtained by

the Solver tool from Microsoft Excel 2013.

3.2.6. Continuous Operations

3.2.6.1. Determination of the flow rates

Before initiating the continuous reaction assays, the flow rates imposed by the peristaltic pump

used to move the fluid thought the packed-bed microreactors were calibrated. For a given speed of

rotation (rpm), and once the steady state was reached, water samples were collected during a certain

time interval to an eppendorf tube. The amount of samples retrieved was weighted and the flow rates

were then calculated. This procedure was applied to several speed values (ranging between 0.5 and 10

rpm) and for all the microreactors used. A calibration curve relating the speed of rotation (rpm) and the

flow rate (mL.min-1) was established for the cylindrical glass column (Figure 9A) packed with different

sized PVA beads and for the cylindrical steel microreactor (Figure 9B) and PMMA-microfluidic reactor

(Figure 9C), both filled with equal-sized silica particles (Annex 1).

3.2.6.2. Continuous Operation assays

Continuous production of reducing sugars from inulin and sucrose was performed using the three

types of microreactors used in this work (Figure 9) packed with inulinase and invertase immobilized in

two different support materials: The cylindrical glass column was packed with 120, 78 and 45 mg of 5,

4 and 3 mm-diameter activated PVA beads, respectively; The cylindrical steel microreactor and the

PMMA-microfluidic reactor were packed with 200 mg of activated silica particles. The reactors were

immersed in a temperature controlled bath, at 50 or 55°C, and substrate solutions, prepared in acetate

buffer 0.1M pH 4.5 or 5.0, were fed to the inlet reactor system through the use of a peristaltic pump,

which was operated within a speed of rotation of 0.5 and 10 rpm. These continuous operation assays

were performed during 200 minutes, approximately, and product samples were retrieved every 30

minutes to eppendorf tubes. During the operations, it was verified that the flow rates of the systems were

maintained at the desired parameters.

3.2.6.3. Operational stability under continuous flow

Continuous production of reducing sugars from inulin was performed using the PMMA-microfluidic

reactor packed with 200 and 500 mg of inulinase immobilized in silica particles. The reactor was

immersed in a temperature controlled bath at 50ºC. The substrate solution, prepared in acetate buffer

0.1 M pH 4.5, was fed to the reactor system through the use of a peristaltic pump with a rotation speed

of 0.5 rpm. The continuous processes were performed during 10 days, samples were collected once a

day and immediately stored at 4ºC for later quantification of reducing sugars.

30

3.2.7. Analytical Methods

i) Quantification of reducing sugars by dinitrosalicylic acid (DNS) method

The produced reducing sugars were quantified by the 3,5-dinitrosalicylic acid (DNS) method131,

adapted to propylene 96-deep well format. In this procedure, 10 µL of sample were added to 90 µL of

distilled water, followed by the addition of 100 µL of DNS reagent, in a 96-well (deep well) plate. When

needed, samples were thawed prior to starting the procedure. The blanks were made using the substrate

solutions (inulin and sucrose 5.0% (w/v)). The solutions were incubated at 100⁰C, for 5 minutes, and the

plate was covered, in order to avoid evaporation of the samples. After the incubation, the solutions were

cooled at RT and then 500 µL of distilled water were added to the said solutions. After a quick

resuspension using a micropipette, 200 µL of the resulting solutions were transferred to 96-well

microplates, and the absorbance was measured in the microplate spectrophotometer at 540 nm.

ii) Quantification of proteins by Bradford protein assay

Protein quantification analysis were performed according to the Bradford protein method. The

process was performed according to the Instructions guide for a Microplate procedure, provided by

Coomassie (Bradford) Protein Assay Kit, from Thermo Scientific132. In this procedure, 150 µL of sample

were pipetted into a 96-well microplate, followed by the addition of 150 µL of Coomassie reagent to each

well. Afterwards, the microplate was mixed for 30 seconds, using a plate shaker, and then the plate was

incubated at RT for 10 minutes, in order to have more consistent results. The absorbance of the samples

was measured using a microplate spectrophotometer at 595 nm.

iii) Determination of packed bed pressure drop

The pressure drop along the length of the packed beds was determined using Ergun’s correlation

proposed for beds composed with spherical or non-spherical particles (Equation 3)98:

𝛥𝑃 = 150 µ (1 − 𝜀)2𝐿

𝜀3. 𝑑𝑝2

𝑣𝑠 + 1.75 (1 − 𝜀)𝐿𝜌

𝑑𝑝. 𝜀3𝑣𝑠

2

[Equation 3]

Where 𝛥𝑃 refers to the Pressure drop (Pa), µ to the Fluid viscosity (Pa.s), 𝜀 to the Average bed porosity

(inter-particle void fraction), 𝑑𝑝 to the Particle diameter (m), 𝐿 to the Bed height (m), 𝑣𝑠 to the Superficial

velocity of the fluid (m.s-1), and 𝜌 to the fluid density (Kg.m-3). The void fraction (𝜀) was determined

experimentally using liquid impregnation method which can be illustrated as follows (Equation 4)133. For

the case of PVA beads, this method consisted on filling the microreactor with the immobilized beads,

and then pouring a certain volume of distilled water in it till it covered all the beads. After that, the volume

31

of water was measured and the total volume of void was determined. For the case of the silica particles,

since these were very small, the average determination of the void volume was performed

mathematically.

𝜀 =𝑉𝑜𝑖𝑑 𝑣𝑜𝑙𝑢𝑚𝑒

𝑅𝑒𝑎𝑐𝑡𝑜𝑟 𝑏𝑒𝑑 𝑣𝑜𝑙𝑢𝑚𝑒

[Equation 4]

iv) Determination of the Péclet number

In order to verify the existence of axial dispersion of the fluid within the different reactor systems,

the dimensionless Péclet number (𝑃𝑒) was calculated, using equation 5134.

𝑃𝑒 =𝑅𝑎𝑡𝑒 𝑜𝑓 𝑡𝑟𝑎𝑛𝑠𝑝𝑜𝑟𝑡 𝑏𝑦 𝑎𝑑𝑣𝑒𝑐𝑡𝑖𝑜𝑛

𝑅𝑎𝑡𝑒 𝑜𝑓 𝑡𝑟𝑎𝑛𝑠𝑝𝑜𝑟𝑡 𝑏𝑦 𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛 𝑜𝑟 𝑑𝑖𝑠𝑝𝑒𝑟𝑠𝑖𝑜𝑛=

𝑢. 𝐿

𝐷

[Equation 5]

Where 𝑢 refers to the fluid velocity of the fluid passing through the reactor (m.s-1), 𝐿 is the characteristic

length of the reactor (m), and 𝐷 is the dispersion coefficient (m2.s-1).

32

33

4. Results and Discussion

4.1. Enzyme Immobilization Parameters

In the present work, two enzymatic immobilization methods were performed for inulinase and

invertase from A. niger and S. cerevisiae strains, respectively, obtained from commercial preparations:

entrapment in PVA hydrogel, using a peristaltic pump, and crosslinking adsorption in a silica-based

support – purified seasand particles. Some immobilization parameters, such as the immobilization

efficiency and the catalytic activity were evaluated (Table 5). The immobilization efficiencies were

determined taking into account the amount of enzyme lost during the immobilization procedures, which

was accessed by protein quantification assays, using the Bradford protein method. Subsequently, the

respective immobilization efficiencies (IE) were determined using Equation 1. Following this, the study

of the catalytic activities of the formed immobilized systems and of the free systems was performed. The

activity of inulinase and invertase fixed in both PVA beads and silica particles could be valued by

measuring the progress of hydrolysis of inulin and sucrose solution, respectively, under fixed standard

conditions, over a certain time interval.

Table 5. Immobilization parameters of inulinase and invertase from commercial preparations immobilized in two

different supports: PVA beads (with 5, 4 and 3 mm thickness) and silica particles (110 µm thickness).

Support Enzyme

Enzyme

Loading

(mgEnzyme/gsupport)*

IE

(%)

Initial Activity

(greducing sugars / h.mgEnzyme)

Activity

(greducing sugars

/ h.mgEnzyme)

Retained Activity

(%)

PVA

5 mm

Inulinase 0.051

95.9±1.6

40.7±1.3

6.5±0.2 15.95

4 mm 93.5±2.1 4.3±0.2 10.51

3 mm 90.9±4.0 6.1±0.8 15.07

5 mm

Invertase 0.144

98.9±0.8

14.7±0.5

2.8±0.1 18.79

4 mm 97.6±1.6 4.4±0.4 29.85

3 mm 98.7±0.2 6.4±0.3 43.47

Silica Inulinase 0.127 92.1±1.9 40.7±1.3 1.5±0.1 3.75

Invertase 0.360 89.6±0.2 14.7±0.5 2.1±0.0 14.33

*The standard deviation associated with the enzyme loading calculations can be neglected.

Regarding the enzyme immobilization in PVA beads, the experimental set-up assembled for the

production of the immobilized biocatalysts, using a peristaltic pump, allowed the production of

homogenous particles (Figure 10A). This process, besides being totally automated, provided an easier

and less time consuming immobilization procedure, in comparison with the convention method of gel

extrusion using a syringe fitted with a fine needle74. Through this process, it was possible to control the

size of the beads formed, by adapting the diameter of the outlet tubing, and by adjusting the flow rate

imposed by the pump.

During the immobilization procedure, in which the enzyme molecules are being entrapped within

synthetic polymer network, leakage of the enzyme from the gel to the surrounding environment may

occur. Therefore, losses of enzyme during the washing steps of the immobilization process are frequent.

34

Through the analysis of the values obtained, reported in Table 5, it is possible to verify that, for both

cases, the immobilization procedure by entrapment/encapsulation using PVA hydrogel was highly

efficient, which indicates that no major losses of enzyme molecules (inulinase and invertase) occurred

during the immobilization process. One can establish that the overall efficiencies of encapsulation were

higher for the invertase case, in comparison with the inulinase case. This event might be explained by

the respective molecular mass of the enzymes. Even within the same yeast culture, invertase can exist

in more than one form, having different molecular weights. Usually, the intracellular invertase has a

molecular weight of about 150 kDa, whereas the extracellular variety has a molecular weight of nearly

270 kDa135. Relatively to the enzyme preparation from Maxinvert® L1000 and L200,000, it is known that

the MW is around 300 kDa136,137. Inulinases produced by Aspergillus strains can also have different

molecular weights138. Ricca et al. (2007) states that the exo-inulinases and endo-inulinases from a

commercial preparation of Fructozyme LTM display MWs of 74 kDa and 64 kDa, respectively139. Thus,

during the enzyme immobilization procedure, the molecules with lower molecular mass might leak more

easily from the polymeric matrix, thereby leading to a decrease of encapsulation efficiency.

After a first assessment of the results obtained of the enzymatic activity of the PVA immobilized

systems produced, it is possible to infer that the dimensions of the beads slightly influence the enzymatic

activity of the immobilizates. The higher activity values were accounted for the assays using beads with

5 and 3 mm-diameter, for the case of inulinase and invertase, respectively. In the case of invertase

beads, it is possible to verify a continuous increase of catalytic activity as the dimension of the beads

decrease. This can be explained by the increase surface area per unit weight of the support, leading to

a higher diffusion rate of the substrate solution into the polymeric network, due to the smaller particle

structure, which results in higher amount of product formed per milligram of immobilized enzyme. These

results are consisted with the ones reported by Karimi et al. (2014), but in this case the support was

non-porous silica nanoparticles140.

The enzymatic activity of the immobilizates, when compared with the free systems, presented

substantially lower values. When working with enzymes immobilized by entrapment in polymeric

networks, there are a few events that can lead to loss of enzymatic activity: when enzymes are

entrapped in a gel matrix, some of the enzyme molecules gradually diffuse towards the outer shell of

the gel and, eventually, leak into the surrounding medium, thus leading to apparent losses of enzyme

activity. This problem is more evident during the immobilization process, in continuous operation, or,

eventually, during the storage period of the immobilized systems. In the PVA case, taking into account

the dimensions of the pores and the dimensions of the enzyme molecules, the leakage problem would

be, mainly, associated with the encapsulation process52; another problem is associated with mass

transfer resistance to the substrate solution, i.e., in certain cases, the substrate cannot diffuse deep into

the gel matrix. At the same time, the diffusional resistance encountered by the product molecules can

sometimes lead to the accumulation of the product near the center of the gel to an undesirable high

level, causing product inhibition for some enzymes. This is why the gel particles should have a certain

size and configuration, in order to allow the substrate to easily move freely in the gel matrix.

35

Immobilization protocols using PVA hydrogel have been reported in several other works.

Fernandes et al. (2009) performed an efficient immobilization of inulinase from a commercial preparation

(Fructozyme L) in PVA particles through extrusion of LentiKat®Liquid in PEG 60053. Both Mohn Zain et

al. (2010) and Seker et al (2014) performed the conversion of sucrose, from liquid pineapple waste, to

glucose using Baker’s yeast invertase immobilized in PVA-alginate-sulfate beads80,141.

Relatively to the immobilization method using a silica-based support, the conventional

experimental procedure used, adapted from Rocha et al. (2006)71, allowed an efficient adsorption of the

enzyme molecules onto the support (Figure 10B). Still, leakage of enzyme from the particles during the

immobilization procedure was also accounted for. The efficiency of the crosslinking adsorption onto

silica particles was very similar for both enzymes (around 92 and 90% for inulinase and invertase,

respectively). Although the pore size is not established, the silica-based support used in this work is

semi-porous, as well as most silica materials, which reveals to be advantageous in enzyme

immobilization procedures, since porous structures enables a large quantity of enzyme molecules to be

immobilized and, furthermore, the entrapment in the pores may help protecting the enzymes from the

surrounding media. However, the enzymatic activity may be compromised when using covalent binding

to attach enzymes to silica carriers due to conformational changes that may disrupt part of the catalytic

site of the biocatalysts62.

Regarding the enzymatic activity upon the immobilization procedure, results show that the

biocatalysts activity was substantially decreased, for both enzymatic systems, as expected. The main

disadvantage of using covalent immobilization is the lower catalytic activity of the biocatalysts compared

to the free enzymes in solution62. It is possible that due to the immobilization, some active sites may

remain unexposed at high enzyme loading. It has been reported that the enzyme catalytic activity does

commonly decrease due to either modifications on the native three dimensional structure of the enzyme

or due to microenvironment effects caused by the interaction between the support and the enzyme-

substrate system142. When using crosslinking agents, a higher enzyme concentration is immobilized,

which may result in saturation of the binding sites with enzyme which further leads to diffusion limitation

phenomena.

The immobilization yield is highly dependent on a delicate balance of several factors such as the

nature of the enzyme, the concentration of both the enzyme and the reagents used, the pH and ionic

strength of the solution, the temperature, and the reaction time31. The catalytic activity of immobilized

enzymes prepared using multifunctional reagents, such as glutaraldehyde, can vary substantially, and

has been shown to be dependent on the amount of crosslinking agent used during immobilization

protocols, as well other factors like the reaction time. In the present work, the activation of the solid

support using glutaraldehyde solution 10.0% (v/v) was performed overnight (about 18 hours) at ambient

temperature, which might have led to an extensive activation of the support. Extensive crosslinking may

constitute a problem, since it can result in distortion of the enzyme structure, i.e., the active site

conformation. Due to these alterations, the accessibility and accommodation of the substrate can be

compromised, thus affecting the retention of biological activity.

36

An enzyme immobilization procedure using silica carriers from purified sea sand was reported by

Iyer et al. (2003). In this procedure, Cyclodextrin glucanotransferase from Bacillus circulans was

immobilized on the silica material, and optimum crosslinking with minimum denaturation of the enzyme

was achieved at 20°C using 2.0% (v/v) glutaraldehyde for 2 hours, yielding 98% of adsorption66. In

Carvalho and Fernandes (2015), invertase molecules were covalently bound to silane-coated silica

carriers, and the catalytic activity of the immobilized system was favored using glutaraldehyde solutions

of 0.6% (v/v) and pH 8.0 and overnight incubation time at RT86.

4.2. SEM analysis of silica particles

Figure 11. SEM micrographs of the sea sand silica particles: sea sand silica particles (A); initial particle surface

topography (B); particles after immobilization procedure (C); particle surface topography after immobilization

procedure (D). The analysis after the immobilization procedure were performed for silica particles with inulinase.

The physical structure of the silica carriers was analyzed by SEM before and after the incubation

with the immobilization reagents and the enzyme solutions (Figure 11). SEM analysis revealed a certain

heterogeneity in particle size and shape. The sea sand silica particles mostly present an irregular and

non-spherical shape with semi-porous rough surface (Figure 11A and 11B). After incubation of the

particles with the immobilization procedure reagents (polyethylenimine and glutaraldehyde) and the

enzyme solution, the created silica particles (Figure 11C) had a rougher appearance and presented

A. B.

C. D.

37

finer particle sizes, in comparison with the initial structure. This could be the outcome of the mechanical

stirring used during the immobilization process. One way to avoid this would be to use gentler mixing of

the preparations during the immobilization procedure. However, if the mechanical stirring is too low,

highly homogenous mixing of the components in solution would not be possible, and this could

compromise the uniform distribution of the enzyme onto the carriers. The surface of the carriers also

suffer some modifications after the immobilization process (Figure 11D), presenting higher degree of

roughness, in comparison with the more smooth structure of the initial support surface.

4.3. Effect of temperature and pH on the enzymatic activity of free and

immobilized enzymes

As a consequence of enzyme immobilization, certain properties of the enzyme molecule may

become altered, leading to changes of the most favorable reaction conditions of the biocatalyst. The

evaluation of the pH and temperature profiles upon enzyme immobilization is, therefore, of paramount

importance, since it allows the assessment of shifts in the optimum reactions conditions after the

immobilization procedure.

The effect of pH and temperature variations in the enzymatic activity of inulinase and invertase

immobilized in PVA particles has already been studied in other works. The highest enzymatic activity

reported by Fernandes et al. (2009)53 for sucrose hydrolysis performed by inulinase immobilized in

LentiKat®Liquid based-particles was at 55°C and pH 4.5. In Cattorini et al. (2009)143, both free and

immobilized forms of inulinase show an optimum pH and temperature for catalytic activity at 4.5 pH and

55°C, respectively. Studies related with the hydrolysis of sucrose by invertase entrapped in PVA

hydrogel capsules, carried out by Rebros and co-workers90, reported that the maximum activity for the

free and immobilized forms was achieved at 4.5 pH and 55°C. Long-term operations performed with

enzymes immobilized in LentiKat® particles at temperatures above 55°C, however, proved to be

unfeasible leading to loss of mechanical stability of the particles143,144.

Regarding the silica-immobilized systems, the influence of the temperature and pH on inulinase

and invertase activity was evaluated and is presented in Figure 12, in which the enzymatic activity values

are expressed in terms of relative activity (%).

The temperature profile assays were performed at pH 4.5 and the temperature effect was studied

within the 40-60⁰C range (Figure 12A and 12C). Regarding inulinase, the optimum temperature for the

immobilized enzyme was 55⁰C, which is in agreement with the free system and, therefore, no shift

phenomenon was verified in this case. Concerning invertase, the optimum temperature for the

immobilized enzyme was 55⁰C, which is slightly higher than the one obtained for the free enzyme (50⁰C),

since the immobilization procedure might have caused a certain damage to the conformation flexibility

of the enzyme, leading to the requirement of higher temperatures for the enzyme molecule to recognize

and achieve an appropriate conformation in order to keep its reactivity145.

38

It is also evident from the results that, within the range of temperatures examined, as the

temperature increases, the relative activity of the immobilized enzymes managed to be, in general,

slightly higher comparatively to the free enzymes. This suggests that the immobilization procedure

increased somewhat the tolerance of the biocatalysts towards temperature variations during the

reactions. The thermal stability of an enzyme can usually be increased by crosslinking due to the fact

that intra- and intermolecular crosslinks lead to a more rigid molecule which can be able to resist

conformational changes. The immobilized systems is, thus, capable of protecting the microenvironment

of the enzyme, limiting thermal movement and preventing denaturation.

Figure 12. Temperature (A, C) and pH (B, D) activity profiles for inulinase and invertase in the free form and

immobilized in silica particles with 10.0% (v/v) glutaraldehyde. Bioconversion assays were carried out in 5.0% (w/v)

inulin and sucrose solutions for inulinase and invertase biocatalysts, respectively.

The pH profile assays were carried out at 50⁰C and the pH influence was studied within the 3.5-

5.5 range (Figure 12B and 12D). The data obtained show that the maximum activity of free and

immobilized inulinase were at pH 4.0 and 4.5, respectively. The pH shift verified upon immobilization

are justified by the secondary interactions between the enzyme and the support, which might have led

to changes in inulinase conformation. The optimum pH value obtained for the free invertase was 5.0,

which is in agreement with the optimum conditions established for the free form of commercial invertase

from Maxinvert® L10,000 (pH 2.5-5.5), which was used in this work146. After immobilization, the enzyme

presented higher activity also at pH 5.0, which indicates that no shift of the optimum pH occurred.

A. B.

C. D.

39

In general, inulinase in the immobilized form, even if marginally, presented a higher tolerance,

i.e., higher relative activity, as the pH becomes less acidic, in comparison with the free system. In the

case of invertase, this event was not so pronounced. Nonetheless, for both cases, the immobilization

protocol roughly improved the activity of the biocatalysts at more acidic environments (3.5 pH), when

compared with the free system under such conditions.

The evaluation of the optimum enzymatic activity conditions of inulinase immobilized in solid

supports have been widely reported in several scientific papers. Rocha et al. (2006) immobilized

inulinase from a commercial preparation (Fructozyme L) in Amberlite using glutaraldehyde as

crosslinking agent, and the immobilized system displayed higher catalytic activity at 5.5 pH and 50°C,

whereas the optimum conditions for the free form were found at 4.5 pH and 55°C71. Karimi et al. (2014)

performed an immobilization of endo-inulinase from A. niger in non-porous silica nanoparticles,

producing crosslink immobilized biocatalysts which displayed higher activity at temperatures above

60°C, due to the increased rigidity of the enzyme structure140. In Yewale et al. (2013), inulinase, after

being immobilized on chitosan bead using a two-step covalent immobilization method, presented a pH

shift towards a more acidic environment (from 4.0 to 3.0 pH) and an optimum temperature of 60°C,

which was identical to that of the free system.

4.4. Free and immobilized enzymes kinetic parameters

The kinetic behavior of an immobilized enzyme can be significantly different from that of the same

enzyme in free solution. The properties of an enzyme can be altered upon immobilization, and these

changes may be caused by conformational modifications within the enzyme, which are dependent on

the immobilization procedure performed, as well as the nature of the immobilization support used43.

The kinetic constants, such as KM and Vmax, of enzymes may be altered after the process of

immobilization due to changes in the internal structure and restricted access to the active site of the

molecules. Thus, kinetic studies are imperative to clarify the effect that the immobilization procedure

has on the enzyme’s performance and, this way, evaluate possible mass transfer limitations and/or

structural changes. Kinetics of immobilized biocatalysts in the presence of modifying factors are called

the apparent kinetics.

The evaluation of the kinetic behavior was only performed for the case of inulin hydrolysis, using

free and immobilized inulinase as biocatalyst. The experimental kinetic evaluation of inulinase

immobilized in PVA beads was also not performed in the present work, since these studies have already

been reported by other works present in the literature: In Fernandes et al. (2009), in which the hydrolysis

of sucrose was performed using inulinase entrapped in PVA particles, an increase in the apparent KM

was verified from 43 to 76 g.L-1 and a 1.25-fold decrease in the apparent Vmax was reported upon

immobilization53. Cattorini et al. (2009) performed an immobilization of inulinase in PVA lenticular

40

particles (LentiKats®) and the apparent KM of the enzyme had a roughly 2-fold increase after the

immobilization procedure, suggesting diffusion limitation problems143.

The determination of the kinetic constants (KM and Vmax) of inulinase in the free form and

immobilized onto silica particles was performed at pH 4.5 and 55°C. Michaelis-Menten kinetics was

assumed for inulin hydrolysis performed by inulinase, as demonstrated in previous works71. The well-

known Michaelis-Menten equation (Equation 6) describes the enzymatic activity, i.e., rate of reaction

(V), as function of the substrate concentration ([S]) for an irreversible one-substrate reaction without

inhibition, where Vmax is the maximum rate reaction and KM is the Michaelis constant which expresses

the affinity between the enzyme and the substrate.

𝑉 =𝑉𝑚𝑎𝑥 . [𝑆]

𝐾𝑀 + [𝑆]

[Equation 6]

The Michaelis-Menten curves (Figure 13) and the kinetic parameters KM (g.L-1) and Vmax (mgreducing

sugars.mL-1.min-1) (Table 6) were determined through the Solver tool from Microsoft Excel 2013.

Figure 13. Michaelis-Menten plots for the free (A) and immobilized (B) enzyme, exhibiting the experimental (vi

observed) and predicted (vi expected) curves of the initial reaction rate (v) for inulin hydrolysis in function of the

substrate concentration (S). Bioconversion reactions were carried out at 55°C and pH 4.5 in 1 mL of substrate

solution, with 1 µL of free enzyme solution 0.1% (w/v) and 70 mg of immobilized enzyme. The data presented was

obtained through Solver tool from Microsoft Excel 2013.

A. B.

41

Table 6. Kinetic constants for inulin hydrolysis with free and immobilized inulinase.

System Kinetic Parameters

KM

(g.L-1) Vmax

(mgreducing sugars.mL-1.min-1)

Free inulinase 64.9 3.2

Immobilized Inulinase 83.7 0.97

The comparative analysis of the immobilized inulinase with the free system indicate that the

immobilization procedure, using glutaraldehyde as crosslinking agent, led to a 1.2-fold increase of the

KM value, and a decrease of the Vmax value of nearly 70% (3-fold decrease). The apparent increase in

KM value suggests loss of affinity between the enzyme and the substrate, inulin, which can be justified

by diffusion limitation problems caused by conformational changes of the enzyme upon attachment to

the silica-based support. When comparing these results with others reported in literature, similar

tendencies are verified. In Yewale et al. (2013), after the immobilization process, the KM value increased

nearly 1.19 times, while the Vmax value was roughly maintained70. In Karimi et al. (2014), KM value was

approximately 3.3-fold higher upon enzyme immobilization with crosslinking. In Trivedi et al. (2015),

after immobilization of inulinase on chitosan particles, KM value was increased 1.2 times, while the

reaction velocity was almost unaffected147. The apparent decrease in reaction velocity, Vmax, after

immobilization may be explained by loss of enzyme activity during the immobilization procedure, caused

by new interactions existent between the biocatalyst and the support, which may have compromised the

active site of the enzyme molecules.

4.5. Continuous Flow Operations

To study the performance and feasibility of the microreactors with packed-bed configurations

used in this work, continuous production operations were performed, at steady state, using immobilized

inulinase and invertase as biocatalysts. For this purpose, the cylindrical glass column reactor was

packed with activated PVA beads, with different particle sizes, and the PMMA-microfluidic reactor and

the cylindrical steel microreactor were filled with activated silica-based particles. All the reaction

processes were carried out in temperature controlled baths, at 50 or 55°C, and the substrate solutions

were fed to the reactors through the use of a peristaltic pump. In these operations, different reaction

flow rates were accessed and related with the amount of reducing sugars produced.

All the continuous reaction operations were performed under the optimum reaction conditions of

the immobilized biocatalysts. Operation stability was verified for all the reactive systems during roughly

200 minutes (data not shown), since it was possible to continuously produce reducing sugars by

hydrolysis on inulin and sucrose at steady-state conditions. The results presented herein (Figure 14)

show the relation between the concentration of product formed and the residence time of the substrate

42

solution inside the reactor bed. Through these experiments, it was thus possible to study the effect of

different flow velocities and residence times on product formation yield.

Figure 14. Continuous production assays for inulin and sucrose hydrolysis performed by inulinase and invertase

immobilized systems. A-B – Continuous operations using inulinase and invertase-PVA beads packed in the

cylindrical glass column. Inulin 5.0% (w/v) and sucrose 5.0% (w/v) hydrolysis were performed at 50°C pH 4.5; C –

Continuous operation using inulinase and invertase-silica particles packed in the cylindrical steel microreactor; D –

Continuous operation using inulinase and invertase-silica particles packed in the PMMA-microfluidic reactor. Inulin

5.0% (w/v) hydrolysis performed by inulinase-silica particles was carried out at 55°C pH 4.5, and sucrose 5.0%

(w/v) hydrolysis performed by invertase-silica particles was carried out at 50°C pH 5.0.

The residence time refers to the time that the substrate solution spends inside the reactor’s bed

in contact with the immobilized biocatalysts. The farther the solution travel along the reactor before being

removed, the higher its residence time. A few factors may affect the residence time, mainly the packing

mode and the flow velocity. When a reactor is packed with a catalyst, the reacting fluid (substrate)

usually does not flow through the reactor in a uniform way, which means than there might be sections

in the packed bed that offer little resistance to flow, and as a result, major portion of the fluid may channel

through this pathway. Consequently, the solution following this pathway does not spend as much time

in the reactor as the one flowing through the regions of high resistance to flow. Thereby, there is a

distribution of times that the solution spend in the reactor in contact with the catalysts148. The different

residence times associated with each microreactor were determined using Equation 7.

𝑅𝑒𝑠𝑖𝑑𝑒𝑛𝑐𝑒 𝑡𝑖𝑚𝑒 (min) =𝑉𝑜𝑖𝑑 𝑣𝑜𝑙𝑢𝑚𝑒 𝑜𝑓 𝑡ℎ𝑒 𝑏𝑒𝑑 (𝑚𝐿)

𝐹𝑙𝑜𝑤 𝑟𝑎𝑡𝑒 𝑜𝑓 𝑡ℎ𝑒 𝑓𝑙𝑢𝑖𝑑 (𝑚𝐿. 𝑚𝑖𝑛−1)

[Equation 7]

B. A.

C. D.

43

Low flow rates lead to higher residence times, and vice-versa. The flow rate of the fluid had to be

determined for each packed-bed microreactor system, being the results presented in Annex 1. To do

that, first it was necessary to establish the steady-state time (i.e., time needed for the system to be in

equilibrium). For this experiment, it was assumed that the steady-state was achieved after four times

the residence time, which was previously validated with experimental results.

In the continuous flow process of hydrolysis of inulin and sucrose using biocatalysts immobilized

in 5, 4 and 3 mm-diameter PVA beads (Figures 14A and 14B), the bed of the microreactor was highly

packed with 18, 34 and 60 beads, with masses of 120, 78 and 45 mg, respectively. The substrate

solution, in this case, had a down-flow stream. For every case, as the residence time increased, i.e., as

the flow rate of the fluid decreased, the concentration of product (reducing sugars) increased likewise.

As the fluid flow rate decreases, the convection inside the reactor bed is less effective, however, this

does not have a negative impact on the global reaction rate. For lower flow rate values, the contact

between substrate and the biocatalysts immobilized onto the support particles will be higher, leading to

higher bioconversion yields. The relation between the different packing modes and the residence times

of the substrate solution is also evident from the results presented: for reactor beds packed with large

beads, the respective void volume will be higher, leading therefore to greater residence times, and vice-

versa.

Comparing the three types of PVA-biocatalysts for the inulinase case, it is possible to infer that

the smaller beads, with 3 mm-diameter, led to higher productivities at lower flow velocities:

concentrations of reducing sugars produced varied between 2.7 and 50.2 g.L-1, using flow rates between

134 and 7.1 µL.min-1, respectively. Concerning the different packing modes using invertase-PVA beads,

the variation of productivities with the alteration of particle size was not so evident, in comparison with

the previous case. In this case, the concentration of product formed did not vary so noticeably with the

decrease in size of the enzymatic beads: for the lowest flow rate analyzed (7.1 µL.min-1), it was possible,

for the three systems, to achieve total bioconversion of the substrate, yielding averagely 53.0 g.L-1 of

reducing sugars. The event previously verified, in which the productivity increased with the decrease of

the activated beads, is related with diffusional properties: with the decrease of particle diameter, the

reaction rate per unit volume of catalyst will increase due to an increase of the reactant concentration

towards the particles center. However, when using small particles in highly packed reactor systems, the

productivity rapidly decreases as the fluid velocity increases and the residence time is reduced.

Randomly packed beds containing small beads assembled in microreactors have also been reported

for biocatalytic conversions118,149. The available surface area is best used in cases which beads are

applied. However, the subsequent high pressure drop associated with continuous operations must be

considered in the case of micro-packed bed systems.

The cylindrical steel microreactor was packed with 200 mg of silica-biocatalysts. The biocatalyst

particles were suspended by a filter paper placed inside the reactor bed, in order to prevent clogging of

the outlet system. In this case, the reactor had a down-flow substrate stream, and the bed of the reactor

was not fully packed with the biocatalysts, since the silica particles were too small (with 110 µm

thickness) and that could lead to complete clogging of the system. This way, although the system in

44

question contained a large void volume, allowing, subsequently, a more unruffled fluidization of the fluid

across the reactor bed, the existent void volume between the particles clustered in the bottom of the

reactor was significantly small. By analysis of Figure 14C, it is possible to infer that, for both immobilized

systems, the concentration of product increased with the residence time of the substrate solution inside

the reactor bed, as expected. For inulin hydrolysis case, the highest concentration of reducing sugars

achieved was 24.0 g.L-1 and for sucrose hydrolysis case it was nearly 42.0 g.L-1, for a residence time of

almost 102 minutes and flow rates of approximately 11 µL.min-1.

In the case of the PMMA-microfluidic reactor, with a sandwich-like structure, 200 mg of silica-

biocatalysts were also packed in the microreactor bed. In this case, the enzyme particles were uniformly

attached, with adhesive tape, to the surface of the reactor’s single microchannel, which was used as the

reaction space. Results shown in Figure 14D demonstrate that the productivities of the hydrolysis

reactions were higher for lower flow velocities and higher residence times, as expected. It is also evident

from the results presented herein that, in this case, it was possible to yield higher product concentrations

at lower residence times, comparatively to the case of the cylindrical steel microreactor packed with the

same amount of silica particles. For inulin hydrolysis case, the highest concentration of reducing sugars

achieved was about 40.5 g.L-1 and for sucrose hydrolysis case it was nearly 55.0 g.L-1, for a residence

time of 20 minutes and at flow rates of 7.8 µL.min-1. This results indicate that the bioconversion of inulin

and sucrose into reducing sugars was higher using the microfluidic reactor, since this system offers high

surface-to-volume ratio and short diffusion paths, which dramatically increase the heat and mass

transfer. Accordingly, the more appropriate reactor device to use silica-biocatalysts is a microfluidic

device, in which the particles can be attached, as uniformly as possible, to a microchannel bed, having

low probability of clogging the system.

Studies regarding operation stability of microreactors packed with PVA particles have already

been performed in other works150,151, as well as studies relative to the use of particulate material, such

as Amberlite, in microreactors under continuous flow152.

4.6. Pressure drop evaluation

In reactors with packed-bed configurations it is possible to attain high volumetric-catalyst loads.

However, the main disadvantage of this type of systems is the high pressure drop. In order to study the

performance and optimization of reactions in microstructured devices, such as the ones used in the

present study, it is required information about the conditions inside the reactor, such as the pressure

drop over the bed and its porosity. The highest pressure drops usually occur in packed-bed

microreactors153. Regarding pressure drop across a packed-bed, the investigations carried out so far

indicate that pressure drop is dependent on the fluid velocity, the physical properties of the fluid

(viscosity and density), the average bed porosity of the reactor’s bed, the orientation of the packing, the

size and the shape of the particles, the ratio between the particle and the vessel diameter (side wall

effect), and the ration between the height of the bed and the diameter of the particles98. In the present

work, the respective pressure drop encountered in the three different microreactors were determined

45

through Ergun Equation (Equation 3), one of the most accepted correlations to perform pressure drop

measurements. This equation is valid for laminar and turbulent flows but, however, does not account for

the side wall effect98.

The results presented herein (Figure 15) demonstrate the effect of variations in the fluid (substrate

solution) flow rates on the pressure drop in the bed of the microreactors used in this work. To determine

the bed porosity of each microreactor, first it was necessary to determine the inter-particle void fraction

of the packed beds, as described in Equation 4. The average bed porosity values (ε) determined for the

different packed-bed reactor systems are presented in Annex V. This parameter was, however,

particularly difficult to determine for the case of the cylindrical steel microreactor system packed with

silica particles, being, therefore, estimated from the literature154. Other parameters required to determine

the pressure drop, such as the fluid viscosity and fluid density were established taking into account the

water parameters. The superficial velocity of the fluid was determined considering the flow rates

accounted for the different reactions performed and the area cross section of the microreactor.

Figure 15. Pressure drop measurements of the different packed microreactors and the respective flow resistances

(RF): A - Pressure drop of the cylindrical glass column packed with different sized-PVA beads; B - Pressure drop of

the PMMA-microfluidic reactor and the cylindrical steel microreactor packed with 200 mg of silica biocatalysts.

As the flow rate of a fluid passing through a packed-bed increases, the respective pressure drop

will also increase due to the increasing resistance of the fluid. The fluid resistance (RF) can be defined

as the ratio between the pressure drop and the flow rate, which corresponds to the slope of the graphs

(Figure 15).

Results presented in Figure 15A show the differences of the pressure drop existent in the catalytic

bed of the glass column reactor, with a volume of approximately 1.13 cm3, packed with PVA-beads of

different sizes. According to these results, it is clear that the catalytic bed that led to a higher pressure

drop is the one containing the 3 mm-diameter enzyme beads, exhibiting a flow resistance of about 28

times higher than the flow resistances of the two other catalytic beds (with 4 and 5 mm-diameter beads).

In this case, the reactor’s bed was densely packed with the small enzymatic beads, which led to a great

decrease in the average bed porosity. This event, consequently, caused the considerably increase of

the pressure. Regarding the packed-bed configurations using the larger enzyme beads, the pressure

drop obtained was similar, being slightly higher for the latter case, due to smaller inter-particle void

A. B.

46

volume. As verified, the pressure drop in the reactor can be reduced by using large catalyst particles,

which is at the expense of intra-particle diffusion limitations.

Results shown in Figure 15B demonstrate the differences between the pressure drop present in

the catalytic bed of two greatly different microreactors (cylindrical steel microreactor with bed volume of

nearly 1.20 cm3, and PMMA-microfluidic reactor with a sandwich-like structure and bed volume of 0.28

cm3) packed with the same amount of activated silica particles with 110 µm thickness. By analysis of

these results, it is evident that the pressure drop is considerably higher when silica-biocatalysts are

immobilized in the sandwich-like microfluidic reactor (RF=3152.8) comparatively with the cylindrical steel

microreactor (RF=302.1). In the latter case, the reactor bed was not fully packed with the particles, since

this would lead to complete clogging of the system as the particles were too small. Thus, as the liquid

solution passing through the catalytic bed had a down-flow stream, the majority of the particles were

huddled in the bottom of the reactor, leading to an increase of flow resistance in the outlet phase.

Reactors with larger areas have less resistance to flow that reactors with smaller areas. In the

case of the microfluidic reactor, as the surface to volume ratio increased, in comparison with the previous

case, the resistance to fluid flow also increased considerably. For packed-bed channels, the length

scales that mainly determine the pressure drop are the diameter of the particles, the cross-sectional

area of the flow and the channel length. For microreactor systems, in general, the engineering design

challenge is to balance the gains made in heat and mass transfer with the losses in pressure drop155.

47

4.7. Axial dispersion evaluation

Usually, the transport in porous media is evaluated by considering, simultaneously, the advection,

mechanical dispersion and diffusion. At low velocities, the transport of the fluid may be diffusion

dominant, whereas at high velocities, transport may be advection dominant156.

The fluid circulation in reactors can be evaluated by injecting a colored tracer into a flowing reactor

system. Through this technique, it is possible to verify that some of the dye will exit before the expected

time (short-circuiting), while some other will stay longer in the reactor (backmixing). In case of a plug-

flow system, the dispersion model takes into account the axial dispersion, which is the physical

parameter that assesses the axial backmixing and short-circuiting of fluid157. The phenomenon of axial

dispersion is indicated by the extent of residence times of the singular elements of a fluid stream passing

through any flow vessel158.

Péclet number is a dimensionless number that can relate the effectiveness of mass transport by

advection to the effectiveness of mass transport by either dispersion or diffusion. Generally, diffusion or

dispersion is considered to be the dominant transport mechanism for Péclet numbers smaller than 1156.

The dispersion coefficients, 𝐷, of the liquid-solid catalytic reactions taking place in the different

packed-bed microreactor systems used in this work were estimated taking into account the Reynolds

number, the fluid flow velocity, the particle diameter and the average bed porosity134. Subsequently, the

respective Péclet number values were determined using equation 5. By analysis of the values obtained

(Table 7), it is possible to verify that, for all the reactor systems in question, advection dominates

dispersion, since the Péclet number values are higher than 1.

Regarding the case of the reactor system packed with 5, 4 and 3 mm PVA beads, the Péclet

number values obtained were approximately the same (~1). In these cases, it is possible to assume

that the transport mechanism of the fluid passing through the interstices of the packed bed is roughly

balanced between advection and dispersion. Comparing the different dispersion coefficients

determined, one can conclude that the cylindrical glass column packed-bed system presents a greater

dispersion rate, exhibiting a better mixing within the packing voids.

Regarding the reactor systems packed with silica particles of 110 µm thickness, the Péclet

number calculated for the PMMA-microfluidic reactor is significantly higher, comparatively to the other

reactor systems, which indicates that the mechanical dispersion of the flowing solution is relatively low.

In this case, the average velocity of the fluid flow is significantly higher than the microscopic velocities

existent through the pores in the medium.

48

Table 7. Dispersion coefficients (𝐷) and Péclet number (𝑃𝑒) values of the three different packed-bed microreactor

systems.

Reactor system Packed material 𝑫 𝑷𝒆

Cylindrical glass column PVA beads

5mm 2.200E-07 1.681

4mm 1.728E-07 1.751

3mm 2.304E-07 1.168

Cylindrical steel microreactor Silica particles

8.635E-10 4.415

PMMA-microfluidic reactor 1.149E-08 226.6

4.8. Operational stability under continuous flow

Frequently, when an enzyme is immobilized onto a support, its operational stability is improved.

The concept of stabilization has therefore been an important factor for immobilizing enzymes. The

application of biocatalysts in industry, such as the ones used in the present study, requires the

operational stability of the immobilized systems to be high enough. However, one of the main problems

associated with the use of immobilized enzymes is the loss of catalytic activity. Due to the limited access

of the substrate to the active site of the enzyme, the activity is often compromised 159. To evaluate the

operational stability of the immobilized system using inulinase and silica carriers packed in the

microfluidic reactor device (Figure 9C), a continuous operation process was performed during 10 days

at 50ºC. The volumetric productivity (Equation 8) was also determined, according to Singh et al.

(2008)160.

𝑄𝑝 = 𝐶𝑝 ×𝐹

𝑉

[Equation 8]

Where 𝑄𝑝 refers to the volumetric productivity (g.L-1.h-1), 𝐶𝑝 is the concentration of reducing

sugars formed (g.L-1), 𝐹 is the flow rate of the fluid (L.h-1), and 𝑉 is the reactor volume (L).

The reactor system was first packed with 200 mg of inulinase biocatalyst immobilized in silica

particles, and the substrate solution was fed to the reactor system through a peristaltic pump with a

speed of rotation of 0.5 rpm, leading to a fluid flow rate of 7.8 µL.min-1. Although inulinase-silica

biocatalyst presents higher enzymatic activity at 55ºC, the continuous operation was performed at 50ºC

in order to prevent major decrease in activity due to thermal instability. Figure 16 shows the microreactor

channel bed packed with 200 mg of silica biocatalysts and Figure 17 reveals the data related with the

continuous production of reducing sugars in terms of relative activity of the biocatalyst.

49

Figure 16. Microchannel bed of the PMMA-microfluidic reactor packed with 200 mg of inulinase-silica biocatalysts.

Figure 17. Operational stability of the PMMA-microfluidic reactor with sandwich-like configuration for inulin

hydrolysis based on relative activity (%). 200 mg of inulinase-silica particles were used for the hydrolysis of 5.0%

(w/v) inulin solution in acetate buffer 0.1 M, pH 4.5 at 50°C. At day one, the concentration of reducing sugars was

45.09±2.90 g.L-1, with a volumetric productivity of 75.63 g.L-1.h-1. Through the Exponential Decay Model it was

possible to determine the deactivation constant (𝑘𝑑 = 7.10 days-1).

As it is possible to infer by the results presented in Figure 17, the operational stability at

continuous flow using 200 mg of silica biocatalysts packed in the microfluidic bed was not so efficient,

since it was verified a progressive decrease in activity along the time, leading to a loss of approximately

57% of activity at day ten. Afterwards, the microreactor was densely packed with 500 mg of silica

biocatalysts (Figure 18), and the operational stability of the reactor was greatly improved, as it can be

verified in Figure 19. This system allowed a retention of nearly 83% of the initial activity of the

biocatalysts. These results suggest that higher enzyme loads on the microreactor bed lead to a lower

decrease of enzyme activity along the time, allowing operational stability for longer time periods. The

comparison of the data gathered for the two different catalyst loads may also suggest that when the

higher payload is considered, there might be an excess of biocatalyst, which is gradually exhausted.

Hence, in the early stages of the process, only the particles closer to the inlet will be fully used. As their

activity is gradually exhausted, the particles downstream are increasingly used, until their activity is

exhausted and so forth. A parallelism can be made with a packed bed adsorption columns, where the

packing is gradually saturated until breakthrough occurs.

0

20

40

60

80

100

120

1 2 3 4 5 6 7 8 9 10

Pro

du

ct Y

ield

(%

)

Time (days)

50

Figure 18. Microchannel bed of the PMMA-microfluidic reactor packed with 500 mg of inulinase-silica biocatalysts.

Figure 19. Operational stability of the PMMA-microfluidic reactor with sandwich-like configuration for inulin

hydrolysis based on relative activity (%). 500 mg of inulinase-silica particles were used for the hydrolysis of 5.0%

(w/v) inulin solution in acetate buffer 0.1 M, pH 4.5 at 50°C. At day one, the concentration of reducing sugars was

55.10±0.63 g.L-1, with a volumetric productivity of 92.43 g.L-1.h-1.

0

20

40

60

80

100

120

1 2 3 4 5 6 7 8 9 10

Rel

ativ

e A

ctiv

ity

(%)

Time (days)

51

5. Conclusions and Future Work

Continuous systems for the production of reducing sugars in packed-bed microreactors with

immobilized inulinase and invertase were developed and characterized in the present research work.

Three different microreactor systems with packed-bed configurations were used: a cylindrical glass

column packed with different sized activated PVA beads, a cylindrical steel microreactor randomly

packed with activated sea sand silica particles, and a PMMA-microfluidic reactor, with a sandwich-like

structure, uniformly packed also with activated silica particles.

Both enzyme immobilization procedures were highly efficient, enabling high enzyme loadings

onto the carriers. The experimental set-up assembled for the production of activated PVA beads with

different sizes, using a peristaltic pump followed by incubation of the beads in a PEG 600 liquid solution

revealed to be a highly effective and easy immobilization procedure.

A decrease of enzymatic activity was verified in both immobilized systems in comparison with the

free systems, although this event was more prominent after the immobilization of the biocatalysts onto

silica particles using glutaraldehyde 10.0% (v/v) as crosslinking agent. This may be explained by the

conformational changes of the enzyme molecules that may have led to destabilization of the active site.

Short-period continuous flow operations were performed using the three model microreactor

systems at steady-state conditions, for approximately 200 minutes, and operational stability with no loss

of enzymatic activity was verified during that time. The glass column microreactor yielded higher product

concentrations when packed with PVA beads with smaller dimensions, which may be explained by

higher diffusional rates. The sandwich-like microreactor system packed with silica particles yielded

higher productivities, comparatively to the cylindrical steel microreactor, due to the higher surface-to-

volume ratio and lower diffusion paths.

Operational stability studies were carried out using the PMMA-microfluidic reactor packed with

two different enzyme loads, 200 and 500 mg, of inulinase-silica particles. When the reactor bed was

completely packed with 500 mg of silica particles, the relative activity of the biocatalysts was roughly

maintained during 10 days, suggesting that greater enzyme loads lead to an enhancement of the

operational stability of the system.

The microreactors used in the present study could also constitute a useful tool for the evaluation

of several catalysis processes of industrial importance, since a wide range of conditions can be tested

over a considerably short period of time and at low costs. As possible future work, continuous flow

operations can be performed by pilling up a bundle of microreactors with microfluidic configurations,

followed by determination of the respective productivities and comparison with data from literature

involving conventional approaches.

52

53

References

1. Berg, J. M., Tymoczko, J. L., Stryer, L. Biochemistry. (W H Freeman, 2002).

2. Quintas, A., Freire, A. P. & Halpern, M. J. Bioquímica - Organização Molecular da Vida. (Lidel,

2008).

3. Teal, A. R. & Wymer, P. E. O. Enzymes and their role in biotechnology. (Biochemical Society,

1994).

4. Büchner, E. Alcoholic fermentation without yeast cells. Alkoholische Gihrung ohne Hefezellen.

Bericbte der Deurrcben Cl~ewzis-. Aen GeseUscbuft, Vol. 30, pages 117-124, (1897).

5. Cornish-Bawden, A. New Beer in an Old Bottle. Eduard Buchner and the Growth of Biochemical

Knowledge. (Universitat de València, 1997).

6. Payen, A. & Persoz, J. F. Mémoire sur la diastase, les principaux produits de ses reactions et

leur applications aux arts industriels. Annales de chimie et de physique 53:73–92 (1833).

7. Manchester, K. L. Antoine Béchamp: père de la biologie. Oui ou non? Endeavour 25, 68–73

(2001).

8. Sheldon, R. A. & van Pelt, S. Enzyme immobilisation in biocatalysis: why, what and how. Chem

Soc Rev 42, 6223–6235 (2013).

9. Wingard, L.B.; Katzir, E.; Goldstein, L. Enzyme technology. (Academic Press, 1979).

10. Fogarty, W. M. & Kelly, C. T. Microbial Enzymes and Biotechnology. (Springer Science &

Business Media, 2012).

11. Bornscheuer, U. T., Huisman, G. W., Kazlauskas, R. J., Lutz, S., Moore, J. C., Robins, K.

Engineering the third wave of biocatalysis. Nature 485, 185–194 (2012).

12. Gurung, N., Ray, S., Bose, S. & Rai, V. A Broader View: Microbial Enzymes and Their Relevance

in Industries, Medicine, and Beyond. BioMed Res. Int. 2013, 1–18 (2013).

13. Grunwald, P. Industrial Biocatalysis. (CRC Press, 2014).

14. Katchalski-Katzir, E. Immobilized enzymes — learning from past successes and failures. Trends

Biotechnol. 11, 471–478 (1993).

15. Guisan, J.M. Immobilization of Enzymes and Cells. 1051, (Humana Press, 2013).

16. DiCosimo, R., McAuliffe, J., Poulose, A. J. & Bohlmann, G. Industrial use of immobilized

enzymes. Chem. Soc. Rev. 42, 6437 (2013).

17. Illanes, A. Enzyme biocatalysis: principles and applications. (Springer, 2008).

18. Wong, C.-H. & Whitesides, G. M. Enzymes in Synthetic Organic Chemistry. (Academic Press,

1994).

54

19. Ghaffari-Moghaddam, M., Eslahi, H., Omay, D. & Zakipour-Rahimabadi, E. Industrial

applications of enzymes. Rev. J. Chem. 4, 341–361 (2014).

20. Robertson, D. E. & Steer, B. A. Recent progress in biocatalyst discovery and optimization. Curr.

Opin. Chem. Biol. 8, 141–149 (2004).

21. van den Burg, B. & Eijsink, V. G. H. Selection of mutations for increased protein stability. Curr.

Opin. Biotechnol. 13, 333–337 (2002).

22. Fernández-Lafuente, R., Hernández-Jústiz, O., Mateo, C., Terreni, M., Fernández-Lorente, G.,

Moreno, M. A., Alonso, J., García-López, J. L., Guisan, J. M. Biotransformations catalyzed by

multimeric enzymes: stabilization of tetrameric ampicillin acylase permits the optimization of

ampicillin synthesis under dissociation conditions. Biomacromolecules 2, 95–104 (2001).

23. Nagodawithana, T.; Reed, G. Enzymes in food processing. (Academic Press, 1993).

24. B. Katzbauer, M. N. Classification system for immobilization techniques. Bioprocess Eng. 12,

173–179 (1995).

25. Cantone, S. et al. Efficient immobilisation of industrial biocatalysts: criteria and constraints for

the selection of organic polymeric carriers and immobilisation methods. Chem. Soc. Rev. 42, 6262

(2013).

26. Singh, B. D. Biotechnology: Expanding horizons. (Kalyani Publishers, 2009).

27. Datta, S., Christena, L. R. & Rajaram, Y. R. S. Enzyme immobilization: an overview on techniques

and support materials. 3 Biotech 3, 1–9 (2013).

28. Veronese, F. M., Mammucari, C., Caliceti, P., Schiavon, O. & Lora, S. Influence of PEGylation on

the Release of Low and High Molecular-Weight Proteins from PVA Matrices. J. Bioact. Compat.

Polym. 14, 315–330 (1999).

29. Wang, A., Wang, H., Zhu, S., Zhou, C., Du, Z., Shen, S. An efficient immobilizing technique of

penicillin acylase with combining mesocellular silica foams support and p-benzoquinone cross

linker. Bioprocess Biosyst. Eng. 31, 509–517 (2008).

30. Honda, T., Miyazaki, M., Nakamura, H. & Maeda, H. Facile Preparation of an Enzyme-

Immobilized Microreactor using a Cross-Linking Enzyme Membrane on a Microchannel Surface.

Adv. Synth. Catal. 348, 2163–2171 (2006).

31. Migneault, I., Dartiguenave, C., Bertrand, M. J. & Waldron, K. C. Glutaraldehyde: behavior in

aqueous solution, reaction with proteins, and application to enzyme crosslinking. Biotechniques 37,

790–806 (2004).

32. Gupta, M. N. in Biocatalyst Design for Stability and Specificity 516, 307–324 (American

Chemical Society, 1993).

55

33. Mohamad, N. R., Marzuki, N. H. C., Buang, N. A., Huyop, F. & Wahab, R. A. An overview of

technologies for immobilization of enzymes and surface analysis techniques for immobilized

enzymes. Biotechnol. Biotechnol. Equip. 29, 205–220 (2015).

34. Cao, L., Langen, L. van & Sheldon, R. A. Immobilised enzymes: carrier-bound or carrier-free?

Curr. Opin. Biotechnol. 14, 387–394 (2003).

35. Cao, L. Carrier-bound immobilized enzymes: principles, applications and design. (Wiley-VCH,

2005).

36. Tran, D. N. & Balkus, K. J. Perspective of Recent Progress in Immobilization of Enzymes. ACS

Catal. 1, 956–968 (2011).

37. Velasco-Lozano, S., López-Gallego, F., Mateos-Díaz, J. C. & Favela-Torres, E. Cross-linked

enzyme aggregates (CLEA) in enzyme improvement – a review. Biocatalysis 1, (2016).

38. Sheldon, R. A. Characteristic features and biotechnological applications of cross-linked enzyme

aggregates (CLEAs). Appl. Microbiol. Biotechnol. 92, 467–477 (2011).

39. Aehle, W. Enzymes in industry: production and applications. (Wiley-VCH, 2007).

40. Bolivar, J. M., Eisl, I. & Nidetzky, B. Advanced characterization of immobilized enzymes as

heterogeneous biocatalysts. Catal. Today 259, 66–80 (2016).

41. Choi, J.-M., Han, S.-S. & Kim, H.-S. Industrial applications of enzyme biocatalysis: Current status

and future aspects. Biotechnol. Adv. 33, 1443–1454 (2015).

42. Patel, A. K., Singhania, R. R. & Pandey, A. Novel enzymatic processes applied to the food

industry. Curr. Opin. Food Sci. 7, 64–72 (2016).

43. Buchholz, K., Kasche, V. & Bornscheuer, U. T. Biocatalysts and Enzyme Technology. (John Wiley

& Sons, 2012).

44. Ahmed, E. M. Hydrogel: Preparation, characterization, and applications: A review. J. Adv. Res.

6, 105–121 (2015).

45. Hoffman, A. S. Hydrogels for biomedical applications. Adv. Drug Deliv. Rev. 64, 18–23 (2012).

46. Hassan, C. M. & Peppas, N. A. in Biopolymers· PVA Hydrogels, Anionic Polymerisation

Nanocomposites 37–65 (Springer, 2000).

47. Whelan, A. Polymer Technology Dictionary. (Springer Science & Business Media, 2012).

48. Saxena, S. Polyvinyl Alcohol (PVA) - Chemical and Technical Assessment. (2004).

49. Takei, T., Ikeda, K., Ijima, H. & Kawakami, K. Fabrication of poly(vinyl alcohol) hydrogel beads

crosslinked using sodium sulfate for microorganism immobilization. Process Biochem. 46, 566–571

(2011).

50. Parascandola, P., Branduardi, P. & de Alteriis, E. PVA-gel (Lentikats®) as an effective matrix for

yeast strain immobilization aimed at heterologous protein production. Enzyme Microb. Technol. 38,

184–189 (2006).

56

51. Krasňan, V., Stloukal, R., Rosenberg, M. & Rebroš, M. Immobilization of cells and enzymes to

LentiKats®. Appl. Microbiol. Biotechnol. 100, 2535–2553 (2016).

52. Schenkmayerová, A., Bucko, M., Gemeiner, P., Trelová, D., Lacik, I., Chorvát, D., Acai, P.,

Polakovic, M., Lipták, L., Rebros, M., Rosenberg, M., Stefuca, V., Nedela, V., Tihlariková, E. Physical

and Bioengineering Properties of Polyvinyl Alcohol Lens-Shaped Particles Versus Spherical

Polyelectrolyte Complex Microcapsules as Immobilisation Matrices for a Whole-Cell Baeyer–Villiger

Monooxygenase. Appl. Biochem. Biotechnol. 174, 1834–1849 (2014).

53. Fernandes, P., Marques, M. P. C., Carvalho, F. & Cabral, J. M. S. A simple method for biocatalyst

immobilization using PVA-based hydrogel particles. J. Chem. Technol. Biotechnol. 84, 561–564

(2009).

54. Poraj-Kobielska, M., Peter, S., Leonhardt, S., Ullrich, R., Scheibner, K., Hofrichter, M.

Immobilization of unspecific peroxygenases (EC 1.11.2.1) in PVA/PEG gel and hollow fiber modules.

Biochem. Eng. J. 98, 144–150 (2015).

55. Bilal, M. & Asgher, M. Sandal reactive dyes decolorization and cytotoxicity reduction using

manganese peroxidase immobilized onto polyvinyl alcohol-alginate beads. Chem. Cent. J. 9, (2015).

56. Saallah, S., Naim, M. N., Lenggoro, I. W., Mokhtar, M. N., Abu Bakar, N. F., Gen, M.

Immobilisation of cyclodextrin glucanotransferase into polyvinyl alcohol (PVA) nanofibres via

electrospinning. Biotechnol. Rep. 10, 44–48 (2016).

57. Nunes, M. A. P., Fernandes, P. C. B. & Ribeiro, M. H. L. High-affinity water-soluble system for

efficient naringinase immobilization in polyvinyl alcohol-dimethyl sulfoxide lens-shaped particles:

HIGH-AFFINITY WATER SYSTEM FOR EFFICIENT NARINGINASE IMMOBILIZATION. J. Mol. Recognit.

25, 580–594 (2012).

58. Nunes, M. A. P., Martins, S., Rosa, M. E., Gois, P. M. P., Fernandes, P. C. B., Ribeiro, M. H. L.

Improved thermostable polyvinyl alcohol electrospun nanofibers with entangled naringinase used

in a novel mini-packed bed reactor. Bioresour. Technol. 213, 208–215 (2016).

59. Nunes, M. A. P., Gois, P. M. P., Rosa, M. E., Martins, S., Fernandes, P. C. B., Ribeiro, M. H. L.

Boronic acids as efficient cross linkers for PVA: synthesis and application of tunable hollow

microspheres in biocatalysis. Tetrahedron (2016).

60. Zucca, P. & Sanjust, E. Inorganic Materials as Supports for Covalent Enzyme Immobilization:

Methods and Mechanisms. Molecules 19, 14139–14194 (2014).

61. Carlsson, N., Glustafsson, H., Thorn, C., Olsson, L., Holmberg, K., Akerman, B. Enzymes

immobilized in mesoporous silica: A physical–chemical perspective. Adv. Colloid Interface Sci. 205,

339–360 (2014).

62. Hartmann, M. & Kostrov, X. Immobilization of enzymes on porous silicas – benefits and

challenges. Chem. Soc. Rev. 42, 6277 (2013).

57

63. Flörke, O. W., Graetsch, H. A., Brunk, F., Benda, L., Paschen, S., Bergna, H. E., Roberts, W. O.,

Welsh, W. A., Libanati, C., Ettlinger, M. in Ullmann’s Encyclopedia of Industrial Chemistry (Wiley-

VCH Verlag GmbH & Co. KGaA, 2000).

64. Zelisko, P. Bio-Inspired Silicon-Based Materials. (Springer, 2014).

65. Popat, A., Budi Hartono, S., Stahr, F., Liu, J., Zhang Qiao, S., Lu, G. Q. Mesoporous silica

nanoparticles for bioadsorption, enzyme immobilisation, and delivery carriers. Nanoscale 3, 2801–

2818 (2011).

66. Iyer, J. L., Shetty, P. & Pai, J. S. Immobilisation of cyclodextrin glucanotransferase from Bacillus

circulans ATCC 21783 on purified seasand. J. Ind. Microbiol. Biotechnol. 30, 47–51 (2003).

67. Hung, B.-Y. et al. Utilization of Enzyme-Immobilized Mesoporous Silica Nanocontainers (IBN-4)

in Prodrug-Activated Cancer Theranostics. Nanomaterials 5, 2169–2191 (2015).

68. Gustafsson, H., Küchler, A., Holmberg, K. & Walde, P. Co-immobilization of enzymes with the

help of a dendronized polymer and mesoporous silica nanoparticles. J Mater Chem B 3, 6174–6184

(2015).

69. Du, K., Sun, J., Zhou, X., Feng, W., Jiang, X., Ji, P. A two-enzyme immobilization approach using

carbon nanotubes/silica as support. Biotechnol. Prog. 31, 42–47 (2015).

70. Yewale, T., Singhal, R. S. & Vaidya, A. A. Immobilization of inulinase from Aspergillus niger NCIM

945 on chitosan and its application in continuous inulin hydrolysis. Biocatal. Agric. Biotechnol. 2,

96–101 (2013).

71. Rocha, J. R., Catana, R., Ferreira, B. S., Cabral, J. M. S. & Fernandes, P. Design and

characterisation of an enzyme system for inulin hydrolysis. Food Chem. 95, 77–82 (2006).

72. Schaefer, E. J., Gleason, J. A. & Dansinger, M. L. Dietary Fructose and Glucose Differentially

Affect Lipid and Glucose Homeostasis. J. Nutr. 139, 1257S–1262S (2009).

73. Sirisansaneeyakul, S., Worawuthiyanan, N., Vanichsriratana, W., Srinophakun, P. & Chisti, Y.

Production of fructose from inulin using mixed inulinases from Aspergillus niger and Candida

guilliermondii. World J. Microbiol. Biotechnol. 23, 543–552 (2007).

74. Anes, J. & Fernandes, P. Towards the continuous production of fructose syrups from inulin

using inulinase entrapped in PVA-based particles. Biocatal. Agric. Biotechnol. 3, 296–302 (2014).

75. Fernandes, P. Marine enzymes and food industry: insight on existing and potential interactions.

Front. Mar. Sci. 1, (2014).

76. Valerio, S. G., Alves, J. S., Klein, M. P., Rodrigues, R. C. & Hertz, P. F. High operational stability

of invertase from Saccharomyces cerevisiae immobilized on chitosan nanoparticles. Carbohydr.

Polym. 92, 462–468 (2013).

77. Albertini, A. V. P., Silva, J. L., Freire, V. N., Santos, R. P., Martins, J. L., Cavada, B. S., Cadena, P.

G., Rolim Neto, P. J., Pimentel, M. C. B., Martínez, C. R., Porto, A. L. F., Lima Filho, J. L.

58

Immobilized invertase studies on glass–ceramic support from coal fly ashes. Chem. Eng. J. 214,

91–96 (2013).

78. Ranken, M.D.; Kill, R.C.; Baker, C. Food Industries Manual. (Springer US, 1997).

79. Kotwal, S. M. & Shankar, V. Immobilized invertase. Biotechnol. Adv. 27, 311–322 (2009).

80. Seker, D. C. & Mohd Zain, N. A. Response surface optimization of glucose production from

liquid pineapple waste using immobilized invertase in PVA–alginate–sulfate beads. Sep. Purif.

Technol. 133, 48–54 (2014).

81. Souza Guimaraes, L. H. in Carbohydrates - Comprehensive Studies on Glycobiology and

Glycotechnology (ed. Chang, C.-F.) (InTech, 2012).

82. Dziedzic, S. Z. & Kearsley, M. W. Handbook of Starch Hydrolysis Products and their Derivatives.

(Springer Science & Business Media, 2012).

83. Ettalibi, M. & Baratti, J. C. Sucrose hydrolysis by thermostable immobilized inulinases from

aspergillus ficuum. Enzyme Microb. Technol. 28, 596–601 (2001).

84. Nguyen, Q. D., Rezessy-Szabó, J. M., Czukor, B. & Hoschke, Á. Continuous production of

oligofructose syrup from Jerusalem artichoke juice by immobilized endo-inulinase. Process

Biochem. 46, 298–303 (2011).

85. Missau, J., Scheid, A. J., Foletto, E. L., Jahn, S. L., Mazutti, M. A., Kuhn, R. C. Immobilization of

commercial inulinase on alginate–chitosan beads. Sustain. Chem. Process. 2, 1–6 (2014).

86. Carvalho, F. & Fernandes, P. Packed bed enzyme microreactor: Application in sucrose

hydrolysis as proof-of-concept. Biochem. Eng. J. 104, 74–81 (2015).

87. Sanjay, G. & Sugunan, S. Enhanced pH and thermal stabilities of invertase immobilized on

montmorillonite K-10. Food Chem. 94, 573–579 (2006).

88. Mansour, E. H. & Dawoud, F. M. Immobilization of invertase on celite and on polyacrylamide

by an absorption procedure. J. Sci. Food Agric. 83, 446–450 (2003).

89. Olcer, Z., Ozmen, M. M., Sahin, Z. M., Yilmaz, F. & Tanriseven, A. Highly Efficient Method

Towards In Situ Immobilization of Invertase Using Cryogelation. Appl. Biochem. Biotechnol. 171,

2142–2152 (2013).

90. Rebroš, M., Rosenberg, M., Mlichová, Z. & Krištofíková, Ľ. Hydrolysis of sucrose by invertase

entrapped in polyvinyl alcohol hydrogel capsules. Food Chem. 102, 784–787 (2007).

91. Chang, T. M. S. Biomedical Applications of Immobilized Enzymes and Proteins. (Springer Science

& Business Media, 2013).

92. Wingard, L. B., Katchalski-Katzir, E. & Goldstein, L. Immobilized Enzyme Principles: Applied

Biochemistry and Bioengineering. (Elsevier, 2014).

93. Doble, M. & Gummadi, S. N. BIOCHEMICAL ENGINEERING. (PHI Learning Pvt. Ltd., 2007).

59

94. Illanes, A., Wilson, L. & Vera, C. Problem Solving in Enzyme Biocatalysis. 141–180 (John Wiley

and Sons Ltd, 2013).

95. Fogler, H. S. Essentials of chemical reaction engineering. (Prentice-Hall, 2011).

96. Messing, R. Immobilized enzymes for industrial reactors. (Elsevier, 2012).

97. Subramanian, R. S. Flow through packed beds and fluidized beds. Clarkson Univ. (2004).

98. Ribeiro AM, Neto P, Pinho C. Mean Porosity and Pressure Drop Measurements in Packed Beds

of Monosized Spheres: Side Wall Effects. International Review of Chemical Engineering, Vol. 2, N.1.

(2010).

99. Werther, J. in Ullmann’s Encyclopedia of Industrial Chemistry (Wiley-VCH Verlag GmbH & Co.

KGaA, 2000).

100. Basile, A. Handbook of Membrane Reactors: Reactor Types and Industrial Applications.

(Elsevier, 2013).

101. Vicaria, J. M. Hollow-fiber Bioreactors: Present and Future in the Biotechnology Industry.

(2013).

102. Rios, G. M., Belleville, M. P., Paolucci, D. & Sanchez, J. Progress in enzymatic membrane

reactors – a review. J. Membr. Sci. 242, 189–196 (2004).

103. Kunii, D. & Levenspiel, O. Fluidization Engineering. (Butterworth-Heinemann, 1991).

104. Ancheyta Juárez, J. Modeling and simulation of catalytic reactors for petroleum refining.

(Wiley, 2011).

105. Iliuta, I., Hamidipour, M., Schweich, D. & Larachi, F. Two-phase flow in packed-bed

microreactors: Experiments, model and simulations. Chem. Eng. Sci. 73, 299–313 (2012).

106. Wiles, C. & Watts, P. Recent advances in micro reaction technology. Chem. Commun. 47, 6512

(2011).

107. Hartman, R. L. & Jensen, K. F. Microchemical systems for continuous-flow synthesis. Lab. Chip

9, 2495–2507 (2009).

108. Ehrfeld, W., Hessel, V. & Löwe, H. in Microreactors 1–14 (Wiley-VCH Verlag GmbH & Co. KGaA,

2000).

109. Bolivar, J. M., Wiesbauer, J. & Nidetzky, B. Biotransformations in microstructured reactors:

more than flowing with the stream? Trends Biotechnol. 29, 333–342 (2011).

110. Barrow, D., Taylor, S., Morgan, A. & Giles, L. in Microreactors in Organic Chemistry and

Catalysis (ed. Wirth, T.) 1–33 (Wiley-VCH Verlag GmbH & Co. KGaA, 2013).

111. Miyazaki, M., Honda, T., Yamaguchi, H., Briones, M. P. P. & Maeda, H. Enzymatic Processing in

Microfluidic Reactors. Biotechnol. Genet. Eng. Rev. 25, 405–428 (2008).

112. Fernandes, P. Miniaturization in Biocatalysis. Int. J. Mol. Sci. 11, 858–879 (2010).

60

113. Elvira, K. S., i Solvas, X. C., Wootton, R. C. R. & deMello, A. J. The past, present and potential

for microfluidic reactor technology in chemical synthesis. Nat. Chem. 5, 905–915 (2013).

114. Becht, S., Franke, R., Geißelmann, A. & Hahn, H. Micro Process Technology as a Means of

Process Intensification. Chem. Eng. Technol. 30, 295–299 (2007).

115. Lin, W.-Y., Wang, Y., Wang, S. & Tseng, H.-R. Integrated microfluidic reactors. Nano Today 4,

470–481 (2009).

116. Charpentier, J.-C. Process Intensification by Miniaturization. Chem. Eng. Technol. 28, 255–258

(2005).

117. Bolivar, J. M. & Nidetzky, B. Multiphase biotransformations in microstructured reactors:

opportunities for biocatalytic process intensification and smart flow processing. Green Process.

Synth. 2, 541–559 (2013).

118. Urban, P. L., Goodall, D. M. & Bruce, N. C. Enzymatic microreactors in chemical analysis and

kinetic studies. Biotechnol. Adv. 24, 42–57 (2006).

119. Peterson, D. S., Rohr, T., Svec, F. & Fréchet, J. M. J. Enzymatic Microreactor-on-a-Chip:  Protein

Mapping Using Trypsin Immobilized on Porous Polymer Monoliths Molded in Channels of

Microfluidic Devices. Anal. Chem. 74, 4081–4088 (2002).

120. Ma, J., Liang, Z., Qiao, X., Deng, Q., Tao, D., Zhang, L., Zhang, Y. Organic−Inorganic Hybrid Silica

Monolith Based Immobilized Trypsin Reactor with High Enzymatic Activity. Anal. Chem. 80, 2949–

2956 (2008).

121. Qu, H., Wang, H., Huang, Y., Zhong, W., Lu, H., Kong, J., Yang, P., Liu, B. Stable Microstructured

Network for Protein Patterning on a Plastic Microfluidic Channel:  Strategy and Characterization of

On-Chip Enzyme Microreactors. Anal. Chem. 76, 6426–6433 (2004).

122. Seong, G. H., Heo, J. & Crooks, R. M. Measurement of Enzyme Kinetics Using a Continuous-

Flow Microfluidic System. Anal. Chem. 75, 3161–3167 (2003).

123. Bui, M. P. N., Li, C. A., Han, K. N., Choo, J., Seong, G. H. Enzyme Kinetic Measurements Using a

Droplet-Based Microfluidic System with a Concentration Gradient. Anal. Chem. 83, 1603–1608

(2011).

124. Malecha, K., Pijanowska, D. G., Golonka, L. J. & Torbicz, W. LTCC microreactor for urea

determination in biological fluids. Sens. Actuators B Chem. 141, 301–308 (2009).

125. Žnidaršič-Plazl, P. & Plazl, I. Modelling and experimental studies on lipase-catalyzed isoamyl

acetate synthesis in a microreactor. Process Biochem. 44, 1115–1121 (2009).

126. Tang, Z. & Kang, J. Enzyme Inhibitor Screening by Capillary Electrophoresis with an on-Column

Immobilized Enzyme Microreactor Created by an Ionic Binding Technique. Anal. Chem. 78, 2514–

2520 (2006).

61

127. He, P., Greenway, G. & Haswell, S. J. Development of enzyme immobilized monolith micro-

reactors integrated with microfluidic electrochemical cell for the evaluation of enzyme kinetics.

Microfluid. Nanofluidics 8, 565–573 (2009).

128. Vojinović, V., Esteves, F. M. F., Cabral, J. M. S. & Fonseca, L. P. Bienzymatic analytical

microreactors for glucose, lactate, ethanol, galactose and l-amino acid monitoring in cell culture

media. Anal. Chim. Acta 565, 240–249 (2006).

129. Heo, J., Thomas, K. J., Seong, G. H. & Crooks, R. M. A Microfluidic Bioreactor Based on Hydrogel-

Entrapped E. coli:  Cell Viability, Lysis, and Intracellular Enzyme Reactions. Anal. Chem. 75, 22–26

(2003).

130. Bernardino, S. M. S. A., Fernandes, P. & Fonseca, L. P. Improved specific productivity in

cephalexin synthesis by immobilized PGA in silica magnetic micro-particles. Biotechnol. Bioeng. 107,

753–762 (2010).

131. Miller, G. L. Use of Dinitrosalicylic Acid Reagent for Determination of Reducing Sugar. Anal.

Chem. 31, 426–428 (1959).

132. ‘Instructions – Coomassie (Bradford) Protein Assay Kit – Thermo Scientific, USA.’ [Online].

Available: https://www.thermofisher.com/order/catalog/product/23200 [Accessed: 08-Jan-2016].

133. Al-Mayah, A. M. R. & Al-Mayah, A. M. R. Simulation of Enzyme Catalysis in Calcium Alginate

Beads, Simulation of Enzyme Catalysis in Calcium Alginate Beads. Enzyme Res. Enzyme Res. 2012,

2012, e459190 (2012).

134. Fogler, H. S. Elements of Chemical Reaction Engineering (Pearson Education, Inc., 2006).

135. Reed, G. Yeast technology. (Springer Science & Business Media, 2012).

136. Risk and technical assessment report – Application A1055; Short Chain Fructo-

oligosaccharides.

137. MERCIER-BONIN, M. & Fonade, C. Enzyme transmission during crossflow filtration of yeast

suspensions using gas/liquid two-phase flows. Ann. N. Y. Acad. Sci. 984, 480–491 (2003).

138. BRENDA - Information on EC 3.2.1.7 - inulinase. Available at: http://www.brenda-

enzymes.org/enzyme.php?ecno=3.2.1.7#MOLECULAR%20WEIGHT. (Accessed: 15th January 2016)

139. Ricca, E., Calabrò, V., Curcio, S. & Iorio, G. The State of the Art in the Production of Fructose

from Inulin Enzymatic Hydrolysis. Crit. Rev. Biotechnol. 27, 129–145 (2007).

140. Karimi, M., Chaudhury, I., Jianjun, C., Safari, M., Sadeghi, R., Habibi-Rezaei, M., Kokini, J.

Immobilization of endo-inulinase on non-porous amino functionalized silica nanoparticles. J. Mol.

Catal. B Enzym. 104, 48–55 (2014).

141. Mohd Zain, N. A., Mohd Suardi, S. & Idris, A. Hydrolysis of liquid pineapple waste by invertase

immobilized in PVA–alginate matrix. Biochem. Eng. J. 50, 83–89 (2010).

62

142. Curcio, S., Ricca, E., Saraceno, A., Iorio, G. & Calabrò, V. A mass transport/kinetic model for the

description of inulin hydrolysis by immobilized inulinase: Modeling of inulin hydrolysis by

immobilized inulinase. J. Chem. Technol. Biotechnol. 90, 1782–1792 (2015).

143. Cattorini, S., Marques, M. P. C., Carvalho, F., Chheub, V., Cabral, J. M. S., Fernandes, P.

Lentikat®-based biocatalysts: effective tools for inulin hydrolysis. Chem. Biochem. Eng. Q. 23, 429–

434 (2009).

144. Rebroš, M., Pilniková, A., Slmciková, D., Weignerová, L., Stloukal, R., Kren, V., Rosenberg, M.

Recombinant α-L-rhamnosidase of Aspergillus terreus immobilization in polyvinylalcohol hydrogel

and its application in rutin derhamnosylation. Biocatal. Biotransformation 31, 329–334 (2013).

145. Keerti, Gupta, A., Kumar, V., Dubey, A. & Verma, A. K. Kinetic Characterization and Effect of

Immobilized Thermostable β -Glucosidase in Alginate Gel Beads on Sugarcane Juice. ISRN Biochem.

2014, 1–8 (2014).

146. MAXINVERT L10000 - specification sheet maxinvert. Available at:

http://courses.ecolechocolat.com/lobjects/pdf/specification_sheet_maxinvert.pdf. (Accessed: 1st

July 2016)

147. Trivedi, S., Divecha, J., Shah, T. & Shah, A. Rapid and efficient bioconversion of chicory inulin

to fructose by immobilized thermostable inulinase from Aspergillus tubingensis CR16. Bioresour.

Bioprocess. 2, (2015).

148. Danckwerts, P. V. Continuous flow systems. Distribution of residence times. Chem. Eng. Sci. 50,

3855 (1995).

149. Bogdan, A. R., Mason, B. P., Sylvester, K. T. & McQuade, D. T. Improving Solid-Supported

Catalyst Productivity by Using Simplified Packed-Bed Microreactors. Angew. Chem. Int. Ed. 46,

1698–1701 (2007).

150. Anes, J. P. F. Development of an immobilized enzyme system aiming at inulin hydrolysis under

continuous flow. (Instituto Superior Técnico, 2012).

151. Nunes, M. A. P., Rosa, M. E., Fernandes, P. C. B. & Ribeiro, M. H. L. Operational stability of

naringinase PVA lens-shaped microparticles in batch stirred reactors and mini packed bed reactors-

one step closer to industry. Bioresour. Technol. 164, 362–370 (2014).

152. Pinto, D. S., Carvalho, F., Neves, A. & Fernandes, P. Continuous flow miniature bioreactors for

enzymatic inulin hydrolysis, 11th International Meeting of the Portuguese Carbohydrate Group -

Glupor 11 and 6th Iberian Carbohydrate Meeting (2015).

153. Hessel, V., Schouten, J. C. & Renken, A. Micro Process Engineering: A Comprehensive

Handbook. (John Wiley & Sons, 2009).

63

154. Jourak, A., Hellström, J. G. I., Lundström, T. S. & Frishfelds, V. Numerical derivation of

dispersion coefficients for flow through three-dimensional randomly packed beds of monodisperse

spheres. AIChE J. 60, 749–761 (2014).

155. Losey, M. W., Schmidt, M. A. & Jensen, K. F. A Micro Packed-Bed Reactor for Chemical Synthesis

(Springer, 1999).

156. Huysmans, M. & Dassargues, A. Review of the use of Péclet numbers to determine the relative

importance of advection and diffusion in low permeability environments. Hydrogeol. J. 13, 895–904

(2005).

157. Bird, R. B., Stewart, W. E. & Lightfoot, E. N. Transport phenomena. (J. Wiley, 2002).

158. Khanna, R. T. Axial Dispersion of Liquid in Packed Beds. (McGill University, 1966).

159. Brena, B., González-Pombo, P. & Batista-Viera, F. Immobilization of enzymes: a literature

survey. Methods Mol. Biol. Clifton NJ 1051, 15–31 (2013).

160. Singh, R. S., Dhaliwal, R. & Puri, M. Development of a stable continuous flow immobilized

enzyme reactor for the hydrolysis of inulin. J. Ind. Microbiol. Biotechnol. 35, 777–782 (2008).

64

65

Annexes

Annex I. Determination of the flow rates

Figure A1. Flow rate calibration curve for the cylindrical glass column packed with different sized PVA beads

(with diameters ranging between 3 and 5 mm). The assays were performed with distilled water.

Figure A2. Flow rate calibration curve for the cylindrical steel microreactor and the PMMA-microfluidic reactor

packed with silica particles. The assays were performed with distilled water.

y = 0.0121x + 0.0049R² = 0.9973

y = 0.0149x + 0.0005R² = 0.9964

y = 0.0134x + 0.0004R² = 0.9915

0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0 2 4 6 8 10 12

Flo

w r

ate

(mL.

min

-1)

Speed of rotation (rpm)

5 mm beads 4 mm beads 3 mm beads

y = 0.0148x + 0.0032R² = 0.994

y = 0.0132x + 0.0012R² = 0.9929

0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0 2 4 6 8 10 12

Flo

w R

ate

(mL.

min

-1)

Speed of rotation (rpm)

Cylindrical steel microreactor PMMA-microfluidic reactor

66

Annex II. Quantification of reducing sugars by the DNS method

The reagent 3,5-Dinitrosalicylic acid (DNS) is an aromatic compound that reacts with reducing

molecules to form 3-amino-5-nitrosalicylic acid (Figure A3), which strongly absorbs light at 540 nm. DNS

method is one of the most used methods for the determination of reducing sugars, such as fructose and

glucose.

Figure A3. Schematic representation of the reduction of 3,5-dinitrosalicylic acid (DNS reagent) to 3-amino,5-

nitrosalicylic acid, in the presence of reducing sugars.

In the present work, the DNS method was the preferred procedure to quantify the

biotransformation of inulin and sucrose to reducing sugars. In this sense, the calibration curve was

prepared using different fructose concentrations ranging from 0 to 5 mg.mL-1 (Figure A4).

Figure A4. DNS calibration curve for fructose concentrations ranging from 0 to 5 mg.mL-1.

y = 0.5858x + 0.056R² = 0.9982

0

0.5

1

1.5

2

2.5

3

3.5

0 1 2 3 4 5 6

Ab

sorb

ance

(5

40

nm

)

Fructose (mg.mL-1)

Reduction

3,5-dinitrosalicylic acid (Yellow)

3-amino,5-nitrosalicylic acid (Orange-red)

67

Annex III. Protein quantification by the Bradford method

Total protein quantifications were performed by the Bradford method. Thus, a calibration curve

was prepared. Bovine serum albumin (BSA) was used as standard for protein calibration, and the

calibration curve was performed using a set standards with concentrations ranging from 0.0025 to 0.02

mg.mL-1 (Figure A5). Inulinase and invertase concentrations of 1.27 and 3.60 mg.mL-1, respectively,

were achieved. These concentrations were used in further determinations.

Figure A5. Calibration curve used for protein quantification analysis, obtained from BSA standards with

concentrations ranging from 0.0025 to 0.02 mg.mL-1.

Annex IV. Mass transfer calculations

One of the most important parameters needed in the design of packed-bed systems is the particle-

to-fluid mass transfer coefficient. Mass transfer coefficients (𝐾) were calculated for all the packed-bed

systems using the three different microreactor types. The calculations were performed according to

Tibhe et al. (2013)1, and the following equations were used:

𝑅𝑒 =𝜌𝑉𝐷

µ(1 − 𝜀)

[Equation 9]

Where 𝑅𝑒 refers to the dimensionless Reynolds number, 𝜌 (Kg.m-3) is the fluid density, 𝑉 (m.s-1) is the

superficial velocity of the fluid, 𝐷 (m) is the diameter of the bed, µ (Pa.s) is the viscosity of the fluid, and

𝜀 is the reactor’s void fraction.

y = 21.638x + 0.0735R² = 0.998

0

0.1

0.2

0.3

0.4

0.5

0.6

0 0.005 0.01 0.015 0.02

Ab

sorb

ance

(5

95

nm

)

[BSA] (mg.mL-1)

68

𝑆ℎ = 𝑅𝑒0.5. 𝑆𝑐13

[Equation 10]

Where 𝑆ℎ refers to the dimensionless Sherwood number and 𝑆𝑐 to the dimensionless Schmidt number.

𝐾 =𝑆ℎ. 𝐷

𝐿

[Equation 11]

Where 𝐾 (m.s-1) refers to the mass transfer coefficient, 𝐷 (m2.s-1) is the mass diffusivity, and 𝐿 (m) is the

characteristic length of the reactor bed (i.e. internal diameter or equivalent).

These determinations were performed considering the system of inulin hydrolysis by inulinase

biocatalysts. Certain parameters, like the fluid viscosity and fluid density, were established taking into

account the water parameters. The mass diffusivity of inulin in PVA and in the aqueous pores in silica

particles were considered to be 2.5 × 10−10 and 2.8 × 10−10 m2.s-1, respectively2,3. In the following

graphic representations (Figures A6-8), the mass transfer coefficients are represented in function of the

superficial velocity of the fluid (i.e. substrate solution).

Figure A6. Mass transfer coefficients (K), in function of the superficial velocity of the fluid, determined for the

cylindrical glass column reactor packed with three different sized PVA beads (with diameters ranging between 3

and 5 mm).

69

Figure A7. Mass transfer coefficients (K), in function of the superficial velocity of the fluid, determined for the

cylindrical steel microreactor packed with 200 mg of silica particles biocatalysts.

Figure A8. Mass transfer coefficients (K), in function of the superficial velocity of the fluid, determined for the PMMA-

microfluidic reactor packed with 200 mg of silica particles biocatalysts.

Annex V. Average bed porosity

Table A1. Average bed porosity values determined for the three different packed-bed reactor systems.

Reactor System Particle diameter ε

Cylindrical glass column

5 mm 0.265

4 mm 0.221

3 mm 0.111

Cylindrical steel microreactor 110 µm 0.360

PMMA-microfluidic reactor 110 µm 0.563

70

References – Annexes

1. Tibhe, J. D., Fu, H., Noel, T., Wang, Q., Meuldijk, J., Hessel, V. Flow synthesis of phenylserine

using threonine aldolase immobilized on Eupergit support. Beilstein J. Org. Chem. 9, 2168–2179 (2013).

2. Barranco-Florido, E., Garcı́a-Garibay, M., Gómez-Ruiz, L. & Azaola, A. Immobilization system

of Kluyveromyces marxianus cells in barium alginate for inulin hydrolysis. Process Biochem. 37, 513–

519 (2001).

3. Cochlear Fluids Simulator. Available at: http://oto2.wustl.edu/cochlea/model/diffcoef.htm.

(Accessed: 30th July 2016)