Aggregation behavior of a temperature- and pH-responsive ... · Aggregation behavior of a...

106
Aggregation behavior of a temperature- and pH-responsive diblock copolymer in aqueous solution Micael Garrido Gouveia Thesis to obtain the Master of Science Degree in Bioengineering and Nanosystems Supervisor(s): Prof. Dr. Christine Maria Papadakis and Prof. Dr. Jos´ e Paulo Sequeira Farinha Examination Committee Chairperson: Prof. Dr. Lu´ ıs Joaquim Pina da Fonseca Supervisor: Prof. Dr. Christine Maria Papadakis Members of the Committe: Prof. Dr. Jos´ e Manuel Gaspar Martinho June 2016

Transcript of Aggregation behavior of a temperature- and pH-responsive ... · Aggregation behavior of a...

  • Aggregation behavior of a temperature- andpH-responsive diblock copolymer in aqueous solution

    Micael Garrido Gouveia

    Thesis to obtain the Master of Science Degree in

    Bioengineering and Nanosystems

    Supervisor(s): Prof. Dr. Christine Maria Papadakis andProf. Dr. José Paulo Sequeira Farinha

    Examination Committee

    Chairperson: Prof. Dr. Luı́s Joaquim Pina da FonsecaSupervisor: Prof. Dr. Christine Maria Papadakis

    Members of the Committe: Prof. Dr. José Manuel Gaspar Martinho

    June 2016

  • What I cannot create, I do not understand.Richard Feynman

  • Acknowledgments

    As opening courtesy, I would like to express my heartfelt gratitude to my supervisor Prof. Dr.

    Christine Papadakis who accompanied and guided my work since day one, in every stage of this

    research project, which she kindly accepted. Not only because she showed constant availability for

    fruitful and encouraging weekly discussions, but also because of the great level of patient and will

    power to help me to become a better scientist. As it is not enough, she gave me the opportunity to

    participate in activities such as scientific workshops in Grainau and Aachen, from which I’ve learned

    a lot and allowed me to explore the astonishing Germany as well. For that I will always be grateful.

    On equal grounds, I would like to thank Prof. Dr. José Paulo Farinha for the suggestion for this

    research project topic and for the opportunity to work in the CQFM group under supervising of Dr.

    Carina Crucho who I thank for the time spent. I would like to acknowledge as well all people that I

    contacted in during my work CQFM at that contributed for my learning and, of course for a cheerful

    environment.

    Furthermore, I wish to give my special thanks to Margarita Dyakonova, Bart-Jan Niebuur, Xiaohan

    Zhang and Natalya Vishnevetskaya, my office comrades that in one way or another guided me through

    my work, in a way they all were my unofficial supervisors. In almost every issue and problem or even

    just discussing topics that I was unclear of, they truly and readily made time to help me. And of course,

    thank you for the all nice and silly conversations that made the office a nice place to be and work.

    Also, one great share of gratitude goes to all people of the E13 group that are the living proof that

    a group can be fun, silly and productive at the same time. All the next office exploding laughter would

    cheer me up any day, any time. However, I should express a special thanks to my dear friends Ali

    Ozku, Richard Stockhausen, Edoardo Barabino and Chih-Chun Huang for turning the coffee time so

    cheerful, and for making me feel at home.

    Last but not least, a big thank you to my friends at home that kept in touch in a away that almost

    was like I never left Portugal. But to my family, I am most grateful. I would like to thank my parents for

    the unconditional love, support and for giving me the opportunity to go abroad and enjoy one in a life

    time experience. For that, and much much more, thank you.

    iii

  • Abstract

    Orthogonally switchable diblock copolymers show rich phase behavior in aqueous solution, in

    dependence on external stimuli. This behavior can be utilized for the preparation of ’smart’ drug de-

    livery systems, which are tuned to respond in accordance to a specific physiological environment.

    We investigate the aggregation behavior in water of a dual-stimuli responsive diblock copolymer with

    a pH-responsive block composed by (diisopropylamino) ethyl methacrylate (DPA), a basic tertiary

    amine-based monomer, and a temperature-responsive block composed by 2-(2-methoxyethoxy) ethyl

    methacrylate (MEO2MA), a polyethylene glycol (PEG) analogue monomer, featuring a lower criti-

    cal solution temperature (LCST). In the present work, the temperature- and pH-responsive PDPA-

    b-PMEO2MA diblock copolymer was successfully synthesized by reversible addition-fragmentation

    chain transfer (RAFT) polymerization, as confirmed by 1H NMR spectroscopy. The cloud-point tem-

    perature (Tcp) was determined through turbidimetry for different pH values, and the polymer hydro-

    dynamic radii dependence on pH at room temperature was evaluated by fluorescence correlation

    spectroscopy (FCS), and near Tcp by temperature-resolved fluorescence correlation spectroscopy (T-

    FCS). The characteristic length of the aggregates formed was also assessed below and above Tcp,

    under acidic and alkaline conditions, by small-angle X-ray scattering (SAXS). Below Tcp, our exper-

    iments have shown the formation of small aggregates with different densities depending on the pH

    in solution, mainly attributed to the presence/absence of electrostatic repulsions arising from proto-

    nated/deprotonated amine groups and to the increased hydrophilicity of the temperature-responsive

    block under neutral and alkaline conditions. Additional large aggregates were formed above Tcp, which

    was found to increase above acidic pH values.

    Keywords

    pH/temperature-responsive polymers, amphiphilic diblock copolymer, (diisopropyl amino)ethyl methacry-

    late (DPA), 2-(2-methoxy ethoxy)ethyl methacrylate (MEO2MA), fluorescence correlation spectroscopy

    (FCS), small-angle X-ray scattering (SAXS)

    v

  • Resumo

    Nesta tese foi investigado o mecanismo de agregação em água de um polı́mero de dibloco

    sensı́vel a dois estı́mulos simultaneamente, com um bloco pH-sensı́vel composto por (diisopropil

    amino) etil metacrilato (DPA), um monómero baseado numa amina terciária, e um bloco termos-

    sensı́vel composto por 2-(2-metoxietoxi) etil methacrilato (MEO2MA), um monómero análogo ao poli-

    etileno glicol (PEG), que exibe temperatura crı́tica em solução (LCST). O copolı́mero de dibloco

    PDPA-b-PMEO2MA sensı́vel ao pH e à temperatura em simultâneo foi sintetizado por polimerização

    de transferência reversı́vel de cadeia por adição-fragmentação (RAFT) e confirmado por análise de1H NMR. A temperatura do ponto de nuvem (Tcp) foi determinada através de turbidimetria, para difer-

    entes valores de pH, enquanto a dependência do raio hidrodinâmico do polı́mero em função do pH foi

    testado através de espectroscopia de correlação de fluorescência (FCS) à temperatura ambiente, e

    para valores próximos à Tcp por espectroscopia de correlação de fluorescência de temperatura con-

    trolada (T-FCS). A distância de correlação dos agregados foi também determinada para temperaturas

    acima e abaixo da Tcp, em condições acı́dicas e alcalinas, por espalhamento de raio-X a baixo ângulo

    (SAXS). Abaixo da Tcp, os resultados indicaram a formação de pequenos agregados que, consoante

    o pH em solução, exibiram densidades distintas dependendo da presença/ausência de repulsóes

    electrostáticas contralada pelo estado protonado/desprotonado dos grupos amina, e do aumento de

    hidrofilicidade do bloco termossensı́vel em solução alcalina. Adicionalmente, verificou-se a formação

    de agregados de maiores dimensões para temperaturas acima da Tcp, que por sua vez aumentou

    para valores de pH superiores ao correspondente a condições acı́dicas.

    Palavras Chave

    polı́mero pH/termossensı́vel, copolı́mero de dibloco anfifı́lico, (diisopropil amino)etil metacrilato

    (DPA), 2-(2-metoxietoxi) etil metacrilato (MEO2MA), espectroscopia de correlação de fluorescência

    (FCS), espalhamento de raio-X a baixo ângulo (SAXS)

    vii

  • Contents

    1 Introduction 1

    1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

    1.2 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

    1.2.1 Amphiphilic block copolymers: self-assembly . . . . . . . . . . . . . . . . . . . . 5

    1.2.2 Temperature-responsive polymers . . . . . . . . . . . . . . . . . . . . . . . . . . 7

    1.2.2.A Collapse mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

    1.2.3 pH-responsive polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

    1.3 Polymers under study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

    1.4 Rationale and objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

    2 Methods 15

    2.1 Controlled/living radical polymerization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

    2.1.1 RAFT principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

    2.1.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

    2.1.3 Data analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

    2.2 Fluorescence Correlation Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

    2.2.1 FCS setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

    2.2.1.A FCS for micellization and aggregation studies . . . . . . . . . . . . . . 24

    2.2.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

    2.2.3 Data analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

    2.2.3.A Triplet decay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

    2.2.3.B Multiple diffusion species . . . . . . . . . . . . . . . . . . . . . . . . . . 31

    2.3 Small-Angle X-Ray Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

    2.3.1 SAXS principle and setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

    2.3.1.A SAXS setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

    2.3.2 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

    2.3.3 Data Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

    2.3.3.A Data Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

    3 Results and discussion 43

    3.1 Characterization of the diblock copolymers after synthesis . . . . . . . . . . . . . . . . . 44

    3.1.1 1H NMR spectra analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

    ix

  • 3.2 Diffusional studies of the self-assembled structures . . . . . . . . . . . . . . . . . . . . . 50

    3.2.1 pH dependence of the aggregation behavior at room temperature . . . . . . . . 52

    3.2.2 Temperature dependence of the aggregation behavior . . . . . . . . . . . . . . . 56

    3.2.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

    3.3 Structure of the aggregates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

    3.3.1 Structure dependence on temperature and pH . . . . . . . . . . . . . . . . . . . 62

    3.3.2 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

    4 Conclusions and Final remarks 67

    Bibliography 71

    Appendix A Appendix A-1

    x

  • List of Figures

    1.1 Schematic representation of plasma concentration profile of therapeutic drugs for dif-

    ferent delivery approaches. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

    1.2 Schematic illustration of a diblock copolymer self-assembly and dissociation in re-

    sponse to changes in pH or temperature. . . . . . . . . . . . . . . . . . . . . . . . . . . 5

    1.3 Formation of micelles in a selective solvent for one of the blocks and dependence of the

    unimer and micelles concentration as function of the polymer concentration in solution. 6

    1.5 Influence of the dimensionless packing parameter of polymer chains in the resulting

    self-assembly structures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

    1.6 Schematic phase diagrams associated with LCST (lower critical solution temperature)

    and UCST (upper critical solution temperature) behavior. . . . . . . . . . . . . . . . . . 8

    1.7 Schematic of the cooperative hydration of LCST type temperature-responsive polymers. 9

    1.8 Schematics of the effects of the solution pH in micelles composed by an hydrophobic

    core and a polybasic pH-responsive shell. . . . . . . . . . . . . . . . . . . . . . . . . . . 10

    1.10 Molecular structure of standard linear PEG and nonlinear PEG-analogs constructed

    with poly(ethylene glycol) monomers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

    1.12 Chemical structure of the PDPA-b-PMEO2MA diblock copolymer and the schematics

    of the expected behavior in aqueous solution under temperature and pH stimuli. . . . . 13

    2.1 Scheme for conventional radical polymerization. . . . . . . . . . . . . . . . . . . . . . . 16

    2.3 Schematics for reversible deactivation and reversible chain transfer. . . . . . . . . . . . 17

    2.4 Mechanism of RAFT polymerization and overall reaction. . . . . . . . . . . . . . . . . . 18

    2.5 Mechanism of block copolymer synthesis through the RAFT technique. . . . . . . . . . 19

    2.6 Schematic representation of a typical confocal FCS setup and its principle of operation. 23

    2.7 Scheme of the chemical structure of Rhodamine 6G . . . . . . . . . . . . . . . . . . . . 24

    2.8 Sketch of the temperature-resolved FCS frame setup . . . . . . . . . . . . . . . . . . . . 26

    2.9 Photographs of the custom-made temperature-resolved FCS sample holder. . . . . . . 27

    2.10 Custom-made system for gold deposition onto the edges of the ITO slips. . . . . . . . . 28

    2.11 Autocorrelation curves for different particle numbers (N) and diffusion time (τD). . . . . . 30

    2.12 Autocorrelation functions of two different species present in the detection volume with

    varying relative amplitude (ρi) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

    2.13 Scheme of the interaction of X-rays with matter. . . . . . . . . . . . . . . . . . . . . . . . 33

    xi

  • 2.14 Interference of detected scattered coherent waves and geometrical approach of point

    scatter. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

    2.15 Scheme of a SAXS instrument. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

    2.16 GANESHA 300XL SAXS system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

    2.18 Schematics of the origin of parasitic scattering, and an example of sample masking

    using Fit2D software. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

    2.19 SAXS capillary sample holder. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

    2.20 The power law describing large aggregates for different parameters. . . . . . . . . . . . 40

    2.21 The Ornsteins-Zernicke (OZ) contribution with different parameters describing density

    fluctuations in small aggregates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

    3.2 Chemical structure of the of the DPA and MEO2MA monomers and of the temperature

    and pH-responsive PDPA-b-PMEO2MA diblock copolymer. . . . . . . . . . . . . . . . . 45

    3.6 1H NMR spectra used for the characterization of the DPA monomer and the polymer. . . 48

    3.7 Normalized transmittance curves and photos of the cuvettes containing the samples

    taken before and right after the turbidimetry measurements. . . . . . . . . . . . . . . . . 51

    3.8 Normalized FCS autocorrelation functions from the samples with 0 mg/mL of polymer

    concentration, i.e. only Rh6G in aqueous solutions, for pH 3, 6, 7 and 10, at room

    temperature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

    3.10 Normalized FCS autocorrelation functions of aqueous solutions of PDA-b-PMEO2MA

    for concentrations ranging from 0.01 to 1.2 mg/mL and different pH values, at room

    temperature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

    3.11 Results from FCS measurements of PDPA-b-PMEO2MA aqueous solutions at various

    pH 3, 6,7 and 10 at room temperature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

    3.12 Temperature calibration of the temperature-resolved FCS system. . . . . . . . . . . . . 56

    3.14 Normalized temperature-resolved FCS autocorrelation functions for different pH condi-

    tions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

    3.15 Results from temperature-resolved FCS measurements of a 0.8 mg/mL polymer solu-

    tions at pH 3, 6, 7 and 10. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

    3.18 SAXS scattering curves of aqueous solution with a concentration of 50 mg/mL at pH 6

    (T=23°C and 28°C) and at pH 10 (T=23°C and 28°C) . . . . . . . . . . . . . . . . . . . 63

    xii

  • List of Tables

    3.1 Characteristics of DPA monomer synthesis and RAFT polymerization of the PDPA ho-

    mopolymer and PDPA-b-PMEO2MA obtained from the PDPA as the macro-CTA . . . . 49

    A.1 Hydrodynamic radii values from room temperature FCS measurements . . . . . . . . . A-2

    A.2 Fraction of the slow decay values from room temperature FCS measurements. . . . . . A-2

    A.3 SAXS model fitting parameters for pH = 6 sample . . . . . . . . . . . . . . . . . . . . . . A-3

    A.4 SAXS model fitting parameters for pH = 10 sample . . . . . . . . . . . . . . . . . . . . . A-3

    xiii

  • Glossary

    AIBN 2,2-azobis(2-methyl-propionitrile) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

    APD Avalanche photo diode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

    CAC Critical aggregation concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

    CMC Critical micelle concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

    CMT Critical micelle temperature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6

    CRP Controlled/living radical polymerization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

    CTA Chain transfer agent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

    DDS Drug delivery systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

    DMF Dimethylformamide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

    DP Degree of polymerization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .17

    DPA 2-(diisopropylamino)ethyl methacrylate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

    FCS Fluorescence correlation spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

    ITO Indium-tin-oxide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

    LCST Lower critical solution temperature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8

    PEG Poly(ethylene glycol) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

    RAFT Reversible addition-fragmentation chain-transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .17

    RTD Resistance temperature detector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

    SAXS Small angle x-ray scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

    SDD Sample-to-detector distance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

    SLD Scattering length density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

    TEA Triethylamine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

    TFA Trifluoroacetic acid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

    THF Tetrahydrofuran . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

    UCST Upper critical solution temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

    xv

  • List of Symbols

    p Packing parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

    a0 Surface area of the insoluble block at the interface soluble/insoluble blocks . . . . . 7

    lc Length of the insoluble block . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

    v Volume of the insoluble block . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

    Tcp Cloud-point temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

    Mn Number average molecular weight . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

    N Total number of fluorescent particles in the detection volume . . . . . . . . . . . . . 24

    n Number of different fluorescent species . . . . . . . . . . . . . . . . . . . . . . . . 24

    τD,i Diffusion time of the i th species . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

    ρi Amplitude of the ith species . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

    z0 Half-height of the detection volume . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

    w0 Half-width of the detection volume . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

    TT Triplet fraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

    τT Triplet time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

    τD,Rh6G Diffusion time of Rhodamine 6G . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

    DRh6G Diffusion coefficient of Rhodamine 6G . . . . . . . . . . . . . . . . . . . . . . . . . 25

    δF (t) Fluorescence intensity variance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

    F (t) Fluorescence intensity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

    〈F (t)〉 Average fluorescence intensity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

    G(τ) Normalized autocorrelation function . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

    〈C〉 Average concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    V eff Effective detection volume . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    z0 Half-height of 3D Gaussian profile . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    w0 Half-width of 3D Gaussian profile . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    D Diffusion coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    rH Hydrodynamic radius . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    kB Boltzmann’s constant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    T Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    η Solvent viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    Ds Diffusion coefficient of the sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

    τD,s Diffusion time of the sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

    xvii

  • rH Hydrodynamic radius . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

    φ Azimuth angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

    ~q Scattering vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

    q Momentum transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

    bi Scattering length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

    ρ Scattering length density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

    ρ(~r) Scattering length density of the monomer . . . . . . . . . . . . . . . . . . . . . . . 37

    ρs Scattering length density of the solvent . . . . . . . . . . . . . . . . . . . . . . . . . 37

    I(~q) Scattered intensity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

    f(~q) Scattering length amplitude . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

    P (~q) Form factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

    S(~q) Structure factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

    δ Chemical shift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

    xviii

  • 1Introduction

    Contents1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21.2 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51.3 Polymers under study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101.4 Rationale and objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

    1

  • 1.1 Introduction

    Over the paste few decades research activity in pharmaceuticals’ physico-chemical properties and

    physiological impact have been increasing significantly, becoming a subject of academic and indus-

    trial interest. With the emergence of novel drug therapies, there is an effort to redirect the attention

    to the delivering approach, in an attempt to improve therapy efficiency. Drawbacks related with the

    invasive character of intravenous drug administration (eg. infections) can be easily avoided by opt-

    ing for oral route, being the preferred way of pharmaceuticals agents intake [1]. Factors related with

    drug physico-chemical nature such as poor solubility, low membrane permeability and instability, as

    well as poor pharmacokinetics hinder bioavailability for proper drug absorption [2]. Conventional drug

    delivery systems (DDS) typically actuate in an uncontrollable way, leading to fluctuations of the local

    and/or systemic drug concentration in the bloodstream, resulting in short windows of optimal drug

    concentration, multiple drug administrations and potentially increase in its toxicity [3] (Figure 1.1).

    Consequently, despite being in continuous development, conventional DDS are known to be still far

    from featuring an ideal behavior, with unmet features concerning spatial and temporal release control.

    Nanoscaled oral DDS have shown the ability to encapsulate a variety of drugs, which (i) avoids drug

    degradation in the hostile environment of the gastrointestinal tract, (ii) enhances drug absorption by

    facilitating diffusing through the intestinal membrane, and (iii) improves the pharmacokinetic release

    profiles. Some of the current DDS are highlighted in the work of Mazzaferro et al. [4]. Additionally,

    besides a good performance, a robust DDS is also required to be compatible to the human body,

    which increases the need for the introduction of better and safer (bio)materials for encapsulation and

    targeting purposes, which can provide a more efficient drug loading and release [5]. Consequently,

    there is a need to investigate effective DDS, that allow therapeutic agents to be delivered in a timely

    and localized manner, providing a transportation vehicle that improves drug bioavailability, efficiency

    in drug absorption and ultimately patient compliance [6].

    Figure 1.1: Schematic representation of plasma concentration profile of therapeutic drugs for different deliveryapproaches - conventional release of two drug administrations (orange line), controlled release (green line).MaEC- maximum effective concentration, MiEC-minimum effective concentration.

    2

  • In order to tailor such pharmacological drugs in a safe and effective way, a compromise must

    be maintained between bioavailability, toxicity and disposition within the body. In last years, various

    modification techniques have been employed such as co-solvency, complexation, solid dispersion,

    permeation enhancers, surfactants and others have been reported to help overcoming some solubil-

    ity and permeability issues [7, 8]. Along possible therapeutic formulations, polymer colloidal systems

    have been increasingly used in drug delivery field, as comprehensibly exposed in [9], in great part

    due to the possibility of combining dynamic topological features of such systems and with the biocom-

    patibility that some polymers present. These conditions may be met with polymers able to change

    with its surrounding environment, in other words stimuli-responsive or smart-polymers, making pos-

    sible the design of active and tailor-made therapeutic solutions, reviewed elsewhere [10].The interest

    developed around stimuli-responsive polymers use is due to their ability to change abruptly theirs

    properties upon an small physical/chemical stimulus - as defined by Hoffman [11] - depending on the

    application, it could range from physical (temperature, magnetic field, light, etc), chemical (pH, spe-

    cific ion etc) to biochemical (enzymatic substrates or affinity ligands such as glucose) [12, 13]. These

    properties are given by the polymer nature, that could be synthetic, semi-natural or natural origin,

    and by the fabrication method, permitting combination of stimuli by which polymers can be sensitive,

    ideally resulting on a multifunctional system able to synchronize drug-release profiles with changing

    physiological conditions [14, 15].

    Temperature-responsive polymers are amongst the most widely investigated polymers due to their

    potential biomedical applications and the easy to control stimulus. A temperature response can be

    achieved via a number of mechanisms, including liquid crystal elastomer transitions [16], or solid

    state transitions in shape memory materials [17, 18]. However, the solubility transition of the poly-

    mer upon temperature change is the most commonly used response in drug delivery systems, as

    well as in tissue engineering, gene delivery and others [19, 20]. In these cases, the transition tem-

    perature of a polymer in solution is tuned to be within the biological range and can have an ex-

    ternal (e.g. microwaves) [21] or internal origin (raise of body temperature upon disease) [19, 20].

    Numerous polymers exhibit a temperature-responsive behavior, and they can be divided into two

    major classes according to their origin: naturally occurring polymers and synthetic materials. Cel-

    lulose, chitosan, xyloglucan, and gelatin, along with their derivatives, are some examples of nat-

    ural temperature-responsive polymers. Synthetic materials such as poly(N-alkyl methacrylamide)s

    are perhaps the most heavily studied temperature-responsive polymers, in particular, homopolymers

    of the gold standard poly(N-isopropylacrylamide). However, due to its questionable biocompatibility,

    other alternative polymers such as poly(N-(2-hydroxypropyl) methacrylamide) (PHPMA) [22], poly(2-

    alkyl-2-oxazoline)s and poly(2-oxazines)s and poly(2-alkyl-2-oxazoline)s, and lactam/pyrrolidone/pyrrolidine

    based temperature responsive polymer systems have been used, among others, and are extensively

    reviewed elsewhere [23]. Other relevant biocompatible materials are the poly(ethylene glycol) (PEG)

    (or poly(ethylene oxide) [23, 24]. The interest in such compounds arises from their versatility allowing

    for the design of novel macromolecular architectures, and the possibility to access high-molecular-

    weight PEG-based polymers using relatively mild synthesis conditions [25].

    3

  • On the other hand, pH-responsive polymeric materials have received considerable attention for

    enhanced diagnostics and therapeutic efficacy in the treatment of a wide range of diseases, due

    to their ability to convert environmental pH stimulus to an observable change, such as solubility,

    volume or chain conformation in aqueous solutions [26]. For instance, there has been much interest

    in developing pH-responsive drug delivery systems to target or be triggered in the extracellular matrix

    of solid tumor tissues [15]. Many of the reported pH-responsive polymers used as delivery systems

    have pKa values lower than 6.0, including the acid-labile pH-responsive polymer systems [27, 28],

    histidine-based systems [29], and 2-aminoethyl methacrylamide (AEMA)-based systems [30], which

    are specially suitable for intracellular targeting and delivery after cellular uptake in non-cancerous

    cells. However, the intracellular pH of many cancer cells is neutral or slightly alkaline relative to normal

    cells, which have intracellular endosomes or lysosomes with a pH of 4.5-6.5 [31, 32]. Considering the

    pH gradient difference between normal and cancerous tissue, pH-responsive polymer systems with a

    pKa of 6.5-7.2 [31, 32] need to be developed to specifically target the weakly acidic extracellular matrix

    microenvironment of solid tumors. Among the pH-responsive polymer systems with a pKa above 6.0,

    tertiary amine AEMA-based polymers have been recognized as one of the most promising groups,

    since their pKa can be tuned from approximately 4.5 to 8.5, depending on different substituent groups

    [33].

    When considering biomedical applications, it is often useful for more than one stimuli to be able

    to cause polymer conformational changes, simultaneously. In fact, through block copolymerization,

    where each type of monomer is arranged systematically in the form of blocks, it is possible to couple

    a wide variety of different stimuli responsiveness on the same molecule, by the simple incorporation

    of monomers with different properties. For instance, ionizable and temperature-responsive groups

    may be coupled (Figure 1.2). Because biological systems possess different physiological microenvi-

    roments, the possibility of joining virtually any combination stimuli-responsiveness represent an op-

    portunity for the formulation of suitable DDS. Temperature and pH dual-responsive polymers have

    been designed to improve drug release by the deformation and precipitation of core/shell nanoparti-

    cles in acidic tumor microenvironments [34, 35]. For example, it was reported that temperature- and

    pH-responsive block copolymer micelles, containing poly(ε-caprolactone) and aminoacid functional-

    ized poly(triethylene glycol) loaded with doxorubicin (DOX), exhibited greater anti-tumor activity and

    higher release efficacy in acidic environments than did free DOX in xenograft tumor models, mainly

    due to the pH-responsive phase transition of this system, which enhanced drug release in the acidic

    physiological environments typically present in tumors [36]. Other groups have constructed tempera-

    ture and pH dual-responsive nanoparticles by copolymerizing N-isopropylacrylamide with acrylic acid,

    methacrylic acid (MAA), and N,N-dimethylarylamide [37–39], indicating that the loading capacity and

    release kinetics of drugs can be modulated by varying the structural and physicochemical properties

    of the constituent block copolymers. Thus self-assembled structures based on dual-stimuli responsive

    block copolymers are promising candidates for effective DDS.

    4

  • Figure 1.2: Schematic illustration of a diblock copolymer self-assembly and dissociation in response to changesin pH or temperature. (Adapted from [40]).

    1.2 Background

    1.2.1 Amphiphilic block copolymers: self-assembly

    Polymers permit the length scales to be greatly varied, the superstructure to be controlled, and

    specific functions to be performed. The simplest primary structure which allows these functions to

    be carried out are diblock copolymers, which are macromolecules consisting of one polymer chain

    made of two blocks, each with a different repeating unit covalently bound to each other. This covalent

    junction and the different chemical nature of the two blocks leads to the ability of self-assembling into

    a variety of morphologies. The length scale of these ordered structures matches the one of the poly-

    mer chains themselves, which generally range from 1-100 nm. These morphologies can be formed in

    the bulk state or in solution. For the purposes of the present work, only the latter will be addressed.

    The behavior of block copolymers in solution depends not only on the interaction between the two

    blocks, but also on their interaction with the solvent, usually water in biomedical applications. If the

    solvent is selective for one block, ordered structures will be formed with one block composing the

    water soluble exterior, and the other being located in the interior, protected from the solvent. In this

    scenario, where one block is hydrophobic and the other is hydrophilic, the copolymer is referred to as

    being amphiphilic. However, this definition should not be oversimplified, as self-assembled structures

    could arise, for example, from two water-soluble polymer blocks if they feature a sufficiently large

    difference in hydrophilicity ”degree”. Spontaneous self-assembly is driven by the natural tendency

    of the system to minimize its energy within a given solvent [41]. Thus, these systems have been

    proven to be of interest as they can create structures such as micelles - a sheltered hydrophobic core

    sterically stabilized by a hydrophilic corona, able to accommodate therapeutic poorly water-soluble

    compounds. Block copolymers self-organize into diverse micellar structures above a certain concen-

    tration, which is called the critical micelle concentration (CMC) (Figure 1.3), or critical aggregation

    concentration (CAC), if no defined structure is formed. Starting from low polymer concentration in

    aqueous solution, single chains are molecularly dissolved in solution and are called unimers. In this

    regime, the surface tension of an aqueous solution is high because the air/water interface is not

    fully populated by the polymer chains and therefore, there is no formation of micelle. As the polymer

    concentration increases, the surface tension at the water/air interface decreases until it reaches satu-

    ration, turning micelle formation energetically favorable when the CMC is crossed. Generally, the CMC

    depends on a variety of factors, such as the nature of the hydrophilic and the hydrophobic block, the

    5

  • presence or absence of charged groups, that ultimately will depend on the pH and the ionic strength

    of the solution, temperature, etc. As a general rule, an increase of either the hydrophilic or hydropho-

    bic content leads to lower CMC values. Temperature also plays an important effect in CMC, due to

    its influence on the water solubility of the two blocks. This feature is denominated critical micelle tem-

    perature (CMT) or Krafft point - defined as the minimum temperature at which micelles are formed

    [41]. From a thermodynamical perspective, micellization is well described by the closed association

    mechanism, which is defined by a dynamic equilibrium, above the CMC, between dissolved chains

    (i.e. unimers) and chains forming the micelles, where a constant exchange occurs (Figure 1.3).

    Figure 1.3: Formation of micelles in a selective solvent for one of the blocks (top) and dependence of the unimerand micelle concentration as function of the polymer concentration in solution (bottom). (Adapted from [42, 43]).

    Various micellar morphologies are accessible depending on the characteristics of the block copoly-

    mer (Figure 1.5 (a)). Generally, when the hydrophilic block is longer than the core block, the shape of

    the resulting micelles is spherical. Conversely, increasing the length of the core segment beyond that

    of the corona-forming chains may create other non-spherical structures in solution, including rods,

    vesicles and lamellae, among others, intensively reviewed elsewhere [44, 45]. The morphology of the

    resulting micelles is a result of the inherent molecular curvature arising from the relative sizes of the

    soluble and insoluble parts and of the way this influences the packing of the copolymer chains within

    the aggregates. The packing parameter p expresses the ratio of the molecular volume of the insoluble

    chain to the volume actually occupied by the copolymer blocks in the assembly, usually dictating the

    6

  • most likely self-assembled structure. The balance between hydrophobic and hydrophilic interactions

    gives rise to an optimal surface area, a0, of the insoluble block at the interface between the soluble

    and insoluble blocksṪhis area, together with the length, lc, and the volume, v, of the insoluble block

    contributes to the packing parameter, defined as:

    p =v

    a0lc(1.1)

    (a)

    (b)

    Figure 1.5: Influence of the dimensionless packing parameter of polymer chains in the resulting self-assemblystructures. (a) Schematics of the curvature of the molecule estimated through the dimensionless packing pa-rameter. (b) Various self-assembled structures formed by amphiphilic block copolymers with different packingparameters. (Adapted from [44, 46]).

    Regarding the biomedical use of block copolymers, the CMC value is also an indicator of the

    ability to form micelles and for their stability: the lower the CMC value, the easier micelles are formed,

    and the more stable they are. Consequently, to develop improved drug delivery systems, amphiphilic

    block copolymers that are able to form more stable micelles with low CMC values are appropriate

    candidates [47, 48].

    1.2.2 Temperature-responsive polymers

    As previously mentioned, the most commonly used mechanism in temperature-responsive poly-

    mers, for biomedical proposes, is the polymer volume phase transition in solution. When a certain

    7

  • temperature is crossed, it causes a sudden change in the solvation state of the polymer. Depending

    on the behavior upon heating, temperature-responsive polymers can be divided into two categories

    (Figure 1.6) - if phase separation occurs upon heating, the polymer is said to have a lower critical so-

    lution temperature lower critical solution temperature (LCST), whereas if it becomes soluble, i.e. just

    one single phase, with increasing temperature, the polymer presents a upper critical solution temper-

    ature upper critical solution temperature (UCST). Thus, this phase transition classification concerns

    the critical temperature points below and above which the polymer and solvent are completely mis-

    cible, respectively. Despite UCST and LCST behavior concepts apply for both organic and aqueous

    systems, only the latter will be considered as these are of interest for biomedical applications. It is

    noteworthy to stress that UCST and LCST terms should be used when associated with a phase dia-

    gram, being attributed to, respectively, the maximum or the minimum of that phase diagram, whereas

    the temperatures on the phase boundary line are called cloud-point temperature (Tcp).

    Figure 1.6: Schematic phase diagrams associated with LCST (lower critical solution temperature) and UCST(upper critical solution temperature) behavior. The blue line represents the phase separation boundary, whichresults in a cloud point. (From [49])

    The majority of water-soluble polymers that possess a LCST undergo phase separation upon

    heating, accompanied with a local structural transition involving water molecules surrounding cer-

    tain regions of the polymer chains in solution. The common characteristic of temperature-responsive

    hydrophilic homopolymers is the presence of hydrophobic groups such as methyl, ethyl, and propyl

    groups [50]. The interactions that take place in an aqueous polymer solution are: polymer-polymer,

    polymer-water, and water-water. For instance, there is a change in the hydration state of the polymer

    chains, which is a result of competing interactions, where intra- and intermolecular interactions of the

    polymer are favored compared to a solubilization by water. Phase transition then occurs as result of

    the polymer collapse to a globule-like conformation, leading to the formation of polymer-rich droplets

    and clouding of the solution [51].

    1.2.2.A Collapse mechanism

    As seen previously, temperature-responsive polymers with a LCST in water experience a collapse

    for temperatures higher than the Tcp, becoming less hydrophilic as they change conformation from an

    expanded (soluble) to a globular (insoluble) state, i.e. a coil-to-globule transition. More specifically, the

    8

  • LCST transition reflects local structural transitions involving water molecules surrounding the polymer

    monomers. At low temperatures, the polymer is hydrophilic as the water molecules are able to interact

    with its polar groups - and obliviously to themselves - via hydrogen bonds, which is the initial driving

    force for the polymer dissolution, thus the polymer chain features an extended coil conformation. At

    higher temperatures, water molecules are released into the bulk, allowing interactions between the

    newly exposed hydrophobic monomers, only slightly water-soluble, while trying to minimize the en-

    tropic loss of the system. However, the detailed reasons for the collapse are still unclear [52, 53].

    It has been theoretically predicted [54, 55] that a temperature increase gives rise to local collapsed

    structures within the polymer chain which, consequently, adopts an intermediate necklace-like con-

    formation (Figure 1.7) where hydrophilic groups are adjacent to single pearls of hydrophobic groups.

    The water molecules bind to the polymer chain in sequences, a process called cooperative hydration

    [54]. As temperature approaches Tcp from below, the average length of the sequence is sharply re-

    duced. The thermodynamics of mixing and demixing of polymers in solution are thoroughly discussed

    elsewhere [53].

    Figure 1.7: Schematic of the cooperative hydration of LCST type temperature-responsive polymers. Sequentialhydrogen bonds form along the polymer chain due to the cooperative interaction between adjacent bound watermolecules. ξ: sequence length over which water molecules bind to the chain. The collapsed parts are marked bythin circles. (From [56]).

    In this context, it is noteworthy to mention that often, amphiphilic polymers which remain soluble

    in aqueous media after the transition to the globular state mimic proteins (eg. human hemoglobin),

    earning the designation protein-like polymers [57]. In both cases, hydrophobic units are enclosed in

    the core of the globule, while more hydrophilic ones form a shell around.

    1.2.3 pH-responsive polymers

    Along with temperature, pH is an important stimulus when considering biomedical applications,

    which can be used through pH-responsive polymers whose solubility, volume, configuration and

    conformation are able to be reversibly manipulated by changes in external pH [26]. Essentially, pH-

    responsive polymers are polyelectrolytes that possess weak acidic (anionic) or basic (cationic) groups

    9

  • that either accept or release protons in response to variations in the pH of the aqueous media. These

    alter the ionic interactions, hydrogen bonding, and hydrophobic interactions between the polymer and

    the surrounding water molecules resulting in a reversible microphase separation or self-organization

    phenomenon [58]. In this context, polymers bearing basic moieties such as amine groups, will accept

    protons and be positively charged (through protonation) under acidic conditions, thus the chains tend

    to expand due to the electrostatic repulsions. Whereas under alkaline conditions, the basic groups of

    the polymer release its protons, becoming uncharged (through deprotonation), leading to the collapse

    of the chain in solution (Figure 1.8). The opposite behavior is found in polymers containing acidic moi-

    eties, such as carboxyl acid groups, since the ionization of the acidic groups (through deprotonation)

    will occur under alkaline conditions.

    Figure 1.8: Schematics of the effects of the pH in solution in micelles composed by an hydrophobic core and apolybasic pH-responsive shell. (Adapted from [59]).

    The pH range in which a reversible phase transition occurs is tailorable, for example by selecting an

    ionizable moiety with a pKa matching the desired pH range. Moreover, the proper selection between

    acid or base should be considered for the envisaged application. For instance, the mildly acidic pH

    in tumor tissues (pH ~6.5-7.2) and in inflammatory tissues [31, 32], as well as in the endosomal

    intracellular compartments (pH ~4.5-6.5) [60, 61] could be potential environments to trigger drug

    release from pH-responsive DDS upon arrival at the targeted disease site.

    1.3 Polymers under study

    Taking into account the proven functionalities that temperature- and pH-responsive block copoly-

    mer systems possess for biomedical application, such as drug delivery systems, in the present work

    dual-responsive polymer sensitive to both temperature and pH are studied. Regarding tempera-

    ture stimulus, previous research have indicated the potential of an emerging class of temperature-

    responsive polymers that may compete with the well-studied pNiPAAm - polymers with PEG (poly(ethylene

    glycol)) short side chains [25, 57, 62–64]. Polymerization of these monomers provides direct access

    to previously inaccessible poly(ethylene glycol) (PEG) architectures, increasing the possibilities far

    beyond linear PEG (Figure 1.10).

    10

  • (a)

    (b)

    Figure 1.10: (a) Chemical structures of various oligo(ethylene glycol) methacrylates. Hydrophobic and hy-drophilic molecular regions are indicated in red and blue, respectively. Monomer unit used in the present work isindicated. (b) Molecular structure of standard linear PEG and nonlinear PEG-analogs (rights) from poly(ethyleneglycol)-containing monomers (left). (Adapted from [65]).

    This kind of monomers combines the biocompatibility of linear PEG, which has been widely applied

    in DDS, with the temperature-responsive properties of pNiPAAm with LCST behavior in aqueous

    media. For instance, PEG-methacrylates such as 2-(2-methoxyethoxy) ethyl methacrylate (MEO2MA)

    have been used, as a copolymer along with other PEG-methacrylate monomers, for drug delivery

    applications [23, 66, 67]. In aqueous solution, this monomer features a Tcp = 26-27°C with very weak

    hysteresis [68].

    On the other hand, various ionizable monomers could be the starting point for pH-responsive DDS,

    including sulfonates, carboxylic acids, and amines. The amine-containing 2-(diisopropylamino)ethyl

    methacrylate (DPA) monomer, is a biocompatible [69] pH-responsive material with a pKa around

    6.2-6.4 [33, 70], close to the values reported in literature on the pH in various cancerous microen-

    vironments [71, 72]. In aqueous solution, when the pH is lower than the polymer pKa, the tertiary

    amine groups are protonated (i.e. positively charged) and the monomers become soluble. When the

    solution pH is higher than the monomer pKa, the monomers become less hydrophilic (i.e. uncharged)

    due to deprotonation. Thus, DPA has been used as a comonomer to develop pH-responsive drug

    delivery systems [73–75]. Consequently, a diblock copolymer bearing these two monomers (DPA

    and MEO2MA) should feature different conformations depending on the conditions in solution. For

    11

  • instance, the temperature-responsive PMEO2MA block is expected to be, respectively, expanded (hy-

    drophilic) or collapsed (hydrophobic) in solution for temperatures below or above than the Tcp. The

    pH-responsive PDPA block is expected to be, respectively, expanded (hydrophilic) or collapsed (less

    hydrophilic) in solution, at pH values below or above than the pKa value of the PDPA block (Figure

    1.12 (b)). Thus, as exemplified in Figure 1.12 (c), the PDPA-b-PMEO2MA diblock copolymer has

    the potential to generate self-assembled structures in solution, such as core-shell micelles, under

    conditions that make the solvent selective for one of the two blocks, such as (i) low pH and high

    temperature: hydrophilic (positively charged) pH-responsive PDPA shell, surrounding a hydropho-

    bic core of collapsed temperature-responsive PMEO2MA blocks; and (ii) high pH and low tempera-

    ture: hydrophobic (uncharged) pH-responsive PDPA blocks form the core, being surrounded by the

    hydrophilic temperature-responsive PMEO2MA blocks. Moreover, (iii) low pH and low temperature

    should yield molecularly dissolved unimers, with both blocks being hydrophilic, whereas (iv) high pH

    and high temperature should make both blocks hydrophobic, hence causing formation of aggregates.

    In the first two cases ((i) and (ii)), the micelles formed could, ultimately, transport hydrophobic thera-

    peutic cargo which would be released in the latter conditions ((iii) and (iv)).

    (a)

    12

  • (b)

    (c)

    Figure 1.12: (a) Chemical structure of the PDPA-b-PMEO2MA diblock copolymer. (b) Schematics of the ex-pected collapse behavior of the PDPA-b-PMEO2MA diblock copolymer in aqueous solution. (c) Schematics of theself-assembled structures potentially formed as a result of the aggregation behavior of the PDPA-b-PMEO2MAdiblock copolymer in aqueous solution. (Adapted from [15]).

    13

  • 1.4 Rationale and objective

    Stimuli-responsive polymeric systems constitute an attractive option for enhancing diagnostics

    and therapeutic efficacy. This is achieved through their ability to convert environmental stimuli to

    an observable change, such as solubility, volume or chain conformation in aqueous solution. While

    research on temperature and pH-responsive block copolymers, based on PEG-methacrylates and

    amine-based methacrylates, has been increasingly developed, still few studies attempt to characterize

    the underlying mechanism of the stimuli-responsiveness of these kind of block copolymers. In face of

    the importance of the behavior of dual-stimuli responsive block copolymers in the creation of novel

    and effective drug delivery systems, this project aims the study of the aggregation behavior of the

    temperature- and pH-responsive PDPA-b-PMEO2MA diblock copolymer in aqueous solution, below

    and above of both the cloud-point temperature of the LCST type PMEO2MA block and the pKa of the

    basic PDPA block.

    The first part of our work envisages the synthesis of the PDPA-b-PMEO2MA diblock copolymer,

    achieved through reversible addition-fragmentation chain transfer (RAFT) polymerization, which pro-

    vides a powerful synthetic tool regarding the synthesis of block copolymers. By means of 1H NMR

    spectroscopy analysis, the resulting reaction products of each synthesis step are characterized. The

    second part aims at the investigation of the aggregation behavior of the PDPA-b-PMEO2MA diblock

    copolymer under different temperature and pH conditions in aqueous media. Fluorescence correla-

    tion spectroscopy FCS provides an insight on the aggregation behavior of our polymer for different pH

    conditions in solution at room temperature, in particular of the pH-responsive PDPA block, taking ad-

    vantage of the resolution power of this technique at low concentrations, where individually solubilized

    macromolecules are of interest. By adding a water-insoluble fluorescent dye (Rhodamine 6G) that

    attaches to the hydrophobic domains present in the polymer, FCS technique enables the detection

    of the formed structures and the determination of the correspondent hydrodynamic radii. Such infor-

    mation is also obtained with temperature-resolved FCS (T-FCS), whose measurements, performed at

    temperatures near the Tcp of the PMEO2MA block, allow for the evaluation of the influence of the tem-

    perature on the aggregation behavior of our polymer, under different pH conditions. This technique

    is utilized along with macroscopic methods such as turbidimetry, which helps to corroborate the as-

    sessment of the Tcp of the PMEO2MA block, for the same pH conditions. Complementary knowledge

    about the aggregation behavior in solution for different temperatures and pH values, is gained with

    small-angle X-ray scattering (SAXS) analysis, which allows to for the characterization of the structure

    properties of our polymer (such as size and conformation), at the local scale. Thereafter, with this

    experimental work we attempt to better understand the mechanism of the stimuli-responsiveness of a

    biomedical relevant diblock copolymer, able to respond simultaneously to temperature and pH.

    14

  • 2Methods

    Contents2.1 Controlled/living radical polymerization . . . . . . . . . . . . . . . . . . . . . . . . 162.2 Fluorescence Correlation Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . 212.3 Small-Angle X-Ray Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

    15

  • 2.1 Controlled/living radical polymerization

    Polymers are constituted by repeating units (i.e. monomers), which can be assembled in a wide

    range of possibilities. Following the developments from early radical polymerization studies [76], pre-

    cise control of polymeric structural parameters prepared by radical polymerization techniques is now

    possible, giving rise to a virtually unlimited number of novel polymeric materials. Among others, stud-

    ies conducted by Rizzardo and coworkers have given tremendous contribution in this field [77–79]. To

    understand how the more recent forms of controlled/living radical polymerization (CRP) work, we first

    need to consider the mechanism of the conventional process. Briefly, conventional free radical poly-

    merization is a chain reaction. The chains are initiated by radicals (activated by an initiator) adding

    to monomer. Chain propagation then follows a sequential addition of monomer units to the radical

    so formed. Chain termination then occurs when propagating radicals react either by combination or

    disproportionation (Figure 2.1).

    Figure 2.1: Scheme for conventional radical polymerization (Pn - radical).

    Despite its wide spread use in industry, short life-time chain growth and fast irreversible termina-

    tion controlled by diffusion effects were translated in limitations, with respect to the control level of

    the polymer structure and molecular weight distribution. These bottlenecks are avoided by using a

    controlled/living radical polymerization. In an ideal living polymerization, all chains (i) are initiated at

    the beginning of the reaction, (ii) grow at a similar rate, and (iii) survive the polymerization: there

    is no irreversible chain transfer or termination. If initiation is fast compared to the propagation (i.e.

    growth) of the chains, the molecular weight distribution is very narrow and chains can be extended

    by further adding monomers into the reaction. Thus, living radical polymerization only becomes pos-

    sible if the species (X) reacts with the propagating radicals (P∗n) by reversible deactivation (Figure

    2.3 (a)) (only one propagating species) or reversible chain transfer (two different propagating species)

    (Figure 2.3 (b)), so that the majority of chains are maintained in a dormant form (Pn-X) and do not

    undergo radical-radical termination. A fast equilibrium change, between active and dormant forms,

    allows for an intermittently chain growth. Under these conditions, molecular weight increases linearly

    with conversion and the weight distribution is very sharp, which is not the case in the conventional

    processes.

    16

  • (a)

    (b)

    Figure 2.3: (a) Schematics for reversible deactivation and (b) reversible chain transfer.

    Thus, and according to Arslan [80], CRP techniques are expected to display (i) first-order kinetic

    behavior, (ii) pre-determinable degree of polymerization (DP), (iii) narrow molecular weight distribu-

    tion and (iv) long-lived polymer chains. With this, the synthesis of complex polymers with diverse

    topologies, and functional groups is achievable.

    2.1.1 RAFT principle

    Controlled/living radical polymerization (CRP) techniques such as nitroxide-mediated polymeriza-

    tion (NMP), atom-transfer radical polymerization (ATRP), and reversible addition-fragmentation chain-

    transfer (RAFT), had been worth of attention in current days [81]. Although these methods have

    different initiation systems, they all take advantage of mechanisms depending on activation and de-

    activation equilibrium [82]. Among the existing CRP techniques, RAFT polymerization is arguably the

    most versatile process, conceding the benefit of being able to synthesize well-defined polymers for

    a wider range of monomers under mild reaction conditions. Moreover, it is tolerant to a variety of

    reaction conditions and functionalities, and can be performed using existing conventional free-radical

    polymerization set-ups, such as solution, emulsion, and suspension polymerization [78, 79, 81, 83].

    The key feature of the RAFT polymerization technique is the sequence of addition-fragmentation

    equilibrium (Figure 2.4), that relies on the use of compounds employed as chain-transfer RAFT

    agents, or, more precisely, chain transfer agent (CTA) - organic compounds containing a thiocar-

    bonylthiol (ZC(=S)SR)) moieties. Commonly used RAFT agents include thiocarbonylthio compounds

    such as dithioesters, dithiocarbamates, and trithiocarbonates. Its purpose is to reversibly deactivate

    propagating radicals in a way that the majority of living chains are maintained in a dormant state and

    to create a rapid equilibrium between the active and dormant chains - the R group initiates the growth

    of polymeric chains, while the Z group activates the thiocarbonyl bond towards radical addition and

    stabilizes the intermediate radical. The structures of the R and Z groups are of critical importance to a

    successful RAFT polymerization. Also, the R group of a RAFT agent is important in the pre-equilibrium

    stage of the polymerization. The R group should feature faster kinetics than the propagating radical

    and must efficiently reinitiate monomer as an expelled radical [79, 84].

    17

  • Figure 2.4: Mechanism of RAFT polymerization and overall reaction. The first step of polymerization is theinitiation step, where a radical is created (step i). The oligomeric radicals produced in the initiation step react withthe RAFT agent (1) in the initial step (ii). The radical intermediate (2) can fragment back to the original RAFTagent (1) and an oligomeric radical or fragment to yield an oligomeric RAFT agent (3) and a reinitiating R radical.The structure of R should be such that it is a good reinitiating group. It should also fragment at least as quicklyas the initiator or polymer chains from the stabilized radical intermediate (2). Following re-initiation, polymerchains grow by adding monomer (step iii), and they rapidly exchange between existing growing radicals (as inthe propagation step) and the species capped with a thiocarbonylthio group (step iv). The rapid interchangein the chain-transfer step via the formationof intermediate 4 limits the termination reactions. Although limited,termination reactions still occur via combination or disproportionation mechanisms (step v). (Adapted from [85]).

    18

  • In practice, RAFT polymerization finds its use as a tool for synthesizing block copolymers and com-

    plex polymers with well-defined structures. Block copolymers may be formed by the chain extension

    of a homopolymer with a second monomer (Figure 2.5). Generally speaking, the product obtained by

    homopolymerization of the first monomer is typically isolated, and the chain extension of the resulting

    polymer is conducted afterwards with a second monomer in the presence of an appropriate initiator.

    In block copolymer synthesis by RAFT polymerization, the order of the blocks is of importance. The

    macro-CTA, S=C(Z)S - Pn, in which the Pn block corresponds to the initial homopolymer (poly(M1))

    and Z is the stabilizing group, should have a high transfer rate in the subsequent polymerization of

    the second monomer (M2) to give the poly(M2) block. This requires that the leaving ability of the

    propagating polymer radical (Pn) must be greater than, or at least comparable to, that of the second

    polymer radical (M2) under the reaction conditions. A fast conversion of the macro-CTA to a block

    copolymer is also required to achieve a low polydispersity, as it allows all the M2 blocks to be initiated

    at approximately the same time [85, 86]. When the appropriate parameters are achieved, RAFT poly-

    merization can be used for a broad range of applications, namely the synthesis of block copolymers

    able to self-assemble into organized structures such as micelles, suitable for drug delivery purposes

    [84, 87, 88].

    Figure 2.5: Mechanism of block copolymer synthesis through the RAFT technique. (i) Overall reaction and (ii)mechanism of RAFT polymerization for the synthesis of diblock copolymers - M1 and M2 are different monomers.(Adapted from [85]).

    2.1.2 Experimental

    Reagents: 2-(N,N-diisopropylamino)ethanol (Sigma-Aldrich, 98%), methacryloyl chloride (Fisher

    Scientific, 97%), triethylamine (TEA) , tetrahydrofuran (THF) (Scharlau, 99%), hydroquinone (Sigma-

    aldrich, 99%), 1,4-dioxane (Scharlau, 99.5%, stabilized with 2.5 ppm of 2,6-di-tert-butyl-4-methylphenol),

    dichloromethane (Sigma-Aldrich, 96%), 2,2-azobis(2-methyl-propionitrile) (AIBN) (Sigma-Aldrich, 99%),

    dimethylformamide (DMF)(Scharlau, 99.8%), trifluoroacetic acid (TFA) (Sigma-Aldrich, 99%), ethanol

    (Sigma-aldrich, 99.8%) and sodium sulfate (Panreac, 99%). The CTA/RAFT agent 3-(benzylsulfanylthiocarbonylsulfanyl)-

    propionic acid ) was synthesized according to the literature procedures [89].

    DPA monomer was synthesized by a one-step reaction between 2-(N,N-diisopropylamino)ethanol

    19

  • and methacryloyl chloride. The synthesis was achieved as follows: hydroquinone (0.20 g) and an-

    hydrous THF (100 mL) were added into a two necked flask, closed with rubber septa to minimize

    oxygen and humidity intake. After the flask was put under argon atmosphere, a solution of 2-(N,N-

    diisopropylamino)ethanol (2.4 mL, 14 mmol, 1 eq) and anhydrous TEA (2 mL, 0.014 mmol, 1 eq) were

    added. After cooling down the reaction to 0°C, methacryloyl chloride (1.4 mL, 14 mmol, 1 eq) was ad-

    ministrated dropwise. The resulting solution was refluxed, under stirring, in an oil bath (80°C) for 2h,

    and filtered afterwards to remove the precipitated TEA-hydrochloric salt, which was rinsed with THF.

    The THF was removed from the filtrate using a rotary evaporator and the resulting residue was diluted

    with dichloromethane, washed with water, 10wt% potassium carbonate solution, saturated NaCl so-

    lution, and finally with water. The organic layer was collected and dried under agitation with sodium

    sulfate. After filtration of the salt, dichloromethane was removed using a rotary evaporator and the

    residue was distilled under reduced pressure to give pure DPA as an transparent oil (1.8g, 8.3 mmol).

    For the preparation of DPA homopolymer (PDPA), which served as a macro-CTA for the subse-

    quent block copolymerization, CTA (24 mg, 0.087 mmol), AIBN (1.45 mg, 0.009 mmol), DPA (0.82 g,

    3.77 mmol, 1eq), TFA (0.48 mg, 3.77 mmol) and ethanol (3 mL) were introduced into a Schlenk tube.

    The flask was sealed with a rubber septum, subsequently degassed through freeze pump cycles with

    liquid nitrogen, and left under argon atmosphere. Afterwards, the tube was heated (70°C) under stir-

    ring for 24h. The polymerization was terminated by quenching in liquid nitrogen. The resulting solution

    was concentrated. The product was precipitated into methanol, and dried under vacuum.

    Block copolymerization was prepared by adding 2-(2-methoxyethoxy) ethyl methacrylate (MEO2MA,

    Sigma-Aldrich, 95%) monomer to the PDPA homopolymer that served as macro-CTA . The synthesis

    of the diblock copolymer PDPA-b-PMEO2MA was proceeded within the work group (Lisbon, Portugal),

    similarly to the synthesis of the PDPA homopolymer, according to the literature procedures [70].

    The resulting products of DPA monomer synthesis, PDPA homopolymer, and PDPA-b-PMEO2MA

    diblock copolymer polymerization were analyzed by 1H nuclear magnetic resonance (1H NMR) spec-

    troscopy (Bruker Avance II 400 MHz, UltraShield Magnet - Lisbon, Portugal) in CDCl3 at room tem-

    perature, to determine the reaction conversions, the degree of polymerization (DP) and the number

    average molecular weight (Mn) through processing of the resulting 1H NMR spectra with MestReNova

    software [90]. The requisite peak areas were obtained by numerical integration of the peaks present

    in the 1H NMR spectra, after the assignment of each spectrum and identification of the correspondent

    peaks. In order to obtain a reliable baseline for the numerical integration, the signal of the isolated

    resonance of deutered chloroform (CDCl3 - 7.260 ppm) was chosen as a reference peak.

    2.1.3 Data analysis

    In order to confirm that the product of each step of the polymerization reaction performed is the

    envisaged one, 1H NMR analysis was performed. Regarding homopolymerization, whether of the

    PDPA or the PMEO2MA block, when the monomer is consumed during the polymerization, the vinyl

    protons from the methacrylate moiety, are used for the covalent bonds between adjacent monomers,

    and therefore the vinyl peaks become weaker, whereas protons from the functional side group remain

    20

  • as the conversion to polymer proceeds. From the 1H NMR spectra obtained, the degree of conversion

    was estimated by comparing integrated peak intensity of the vinyl protons of unreacted monomers, to

    the integrated peak intensity of protons from the side groups, present in both the monomer and the

    polymer [91].

    conversion(%) =(

    1− number of protons of resonancetotal integration area of monomer + polymer

    )× 100 (2.1)

    Another important characterization parameter is the DP, which gives the number of the repeat-

    ing units from each polymerization reaction. Here, the DP was estimated assuming (i) a CTA/initiator

    molar ratio high enough (in our case, [CTA]/[AIBN] = 10) to consider that every CTA molecule is in-

    corporated into each initiated polymer chain, and that (ii) every growing chain undergoes an efficient

    polymerization via the RAFT mechanism. As a result, the DP depends only the monomer and CTA ini-

    tial concentrations ([Monomer]0 and [CTA]0, respectively), and on the fraction of converted monomer,

    as follows:

    DP =[Monomer]0

    [CTA]0× conversion(%) (2.2)

    Consequently, the DP was used to estimate the number average molecular weight (Mn) (Equation

    2.3), by taking into account the molecular weight of the polymerized DPA monomer (M(Monomer))

    and the molecular weight of CTA (M(CTA)).

    Mn = DP ×M(Monomer) +M(CTA) (2.3)

    Moreover, because the diblock copolymer was synthesized using the PDPA homopolymer as

    macro-CTA, its molecular weight (M(PDPA)) was used instead of M (CTA), to give the total number

    average molecular weight of the PDPA-b-PMEO2MA diblock copolymer, as follows:

    Mn,total = DP ×M(Monomer) +M(macro-CTA) (2.4)

    2.2 Fluorescence Correlation Spectroscopy

    The underlying principle of fluorescence correlation spectroscopy (FCS) is based on the moni-

    toring of fluctuations of the fluorescence intensity originating from species diffusing through a very

    small detection volume. The fluctuations can be quantified in their strength and duration by tempo-

    rally autocorrelating the recorded intensity signal, a mathematical procedure that gave the technique

    its name. Essential information about the processes governing molecular dynamics of the diffusion

    species can be derived from the temporal pattern by which fluorescence fluctuations arise and decay.

    In general, all physical events that give rise to fluctuations in the fluorescence signal are accessible by

    FCS. Consequently, this technique is useful to study a broad range of systems through the detection

    of intensity variations arising from diffusional, rotational, and photochemical/physical processes of the

    21

  • fluorescently labeled particles within the detection volume. This yields information, for example, about

    the local concentrations, diffusion coefficients or characteristic of inter- or intramolecular interactions

    of those particles. The idea of fluorescence correlation spectroscopy was originally introduced in the

    early seventies by Weeb, Magde and Elson [92, 93] who used it to estimate the kinetics of DNA-dye

    interactions. Around the same time, Rigler et al. suggested the use of fluorescence intensity fluctua-

    tions to gain knowledge about rotational molecular motions and fluorescence life times [94]. Despite

    showing great promise since its early stages, the FCS technique did not became widespread readily,

    mainly due to issues concerning large detection volumes (leading to too high background noise), for

    which longer measurement times and higher dye concentrations were required to compensate for the

    low signal detected. Only later, with the introduction of extremely small detection volumes defined by

    diffraction-limited laser beams along with confocal observation, by Rigler et al. [95, 96], a substan-

    tially higher signal-to-noise ratio through better suppression of background radiation was achieved.

    Further improvements concerning the quality of confocal optics, the time-resolution and sensitivity

    of the detectors as well as the use of band-pass filters for better signal/noise discrimination lead to

    single-molecule detection sensitivity [97].

    2.2.1 FCS setup

    As mentioned above, FCS is a temporal correlation analysis of the intensity variations which arise,

    for instance, from the fluctuations of the concentration of fluorescent particles, entering or leaving

    the detection volume, or from changes in their fluorescence properties. In this context, to extract

    full potential from FCS technique, the concentrations and detection volume should be as small as

    possible, such that only few molecules are simultaneously detected and, at the same time, increase

    the fluorescence photon yield per single molecule [97]. A typical setup for fluorescence correlation

    spectroscopy closely resembles that of a classical confocal microscope (Figure 2.6). A high numerical

    aperture - ideally NA >0.9 - microscope objective (typically water or oil immersion) is used to focus

    a laser beam to a diffraction limited spot into the sample. The resulting fluorescence light is then

    collected by the same objective. After passing through a dichroic mirror and an emission filter, it is

    imaged onto a confocal pinhole, which improves vertical axis resolution by blocking the fluorescent

    light not originating from the focal plane. In this way a very small detection volume of less than 1 µm3

    is created. The detection is done by a fast and sensitive detector, typically a single photon counting

    avalanche photo diode (APD), that possess good quantum yield, and ultimately can reach the photon

    detection resolution down to picosecond range.

    22

  • Figure 2.6: (Left) Experimental setup for FCS. (a) A laser beam is first expanded by a telescope (L1 and L2),and then focused by a high-NA objective lens (OBJ) into a fluorescent sample (S). The fluorescence is collectedby the same objective, reflected by a dichroic mirror (DM), focused by a tube lens (TL), filtered (F), and passedthrough a confocal aperture (P) onto the detector (DET). (b) Magnified focal volume (green) within which thesample particles (black circles) are illuminated. (Right) (c) A typical fluorescence signal, as a function of time,measured for rhodamine green (RG) with a wavelength of 488 nm. (d) Portion of the same signal in panel c,binned, with an expanded time axis and average fluorescence F̄ . The signal is correlated with itself at a latertime (t + τ) to produce the autocorrelation G(τ). (e) Measured G(τ) describing the fluorescence fluctuation ofRG molecules due to diffusion only as observed by FCS. (Adapted from [98]).

    These experimental arrangements, combined with the high sensitivity of the modern detection

    systems result in several unique properties of FCS, namely, (i) the possibility of small concentrations

    of the fluorescent species to be well below 1 part/fL, (ii) the access to diffusion coefficients in the

    range 10-9 to 10-15 m2/s, (iii) the size of the fluorescent species can be as small as 1 nm, and (iv)

    the small confocal volume can be positioned on in a specific place in a heterogeneous system, e.g.

    inside a porous network. Despite of its high sensitivity and versatility, the FCS technique has also

    some intrinsic technical problems and limitations. For example, precise measurements of the diffu-

    sion coefficient and the concentration of the fluorescent species are possible only if the shape and

    the exact dimensions of the confocal detection volume are well known [99]. Typically its dimensions

    are determined by calibration measurements with a reference fluorescent dye with known diffusion

    coefficient, e.g. Rhodamine 6G (Rh6G) [100].

    23

  • Figure 2.7: Scheme of the chemical structure of Rhodamine 6G.

    2.2.1.A FCS for micellization and aggregation studies

    FCS is often used for the characterization of the micellization and aggregation of amphiphilic

    block copolymers in solution, taking advantage of the present hydrophobic domains, on which the

    highly hydrophobic dye Rh6G molecules associate. Consequently, this makes FCS suitable for the

    detection of critical micelle concentration [101–103] of such polymers. This is possible because of

    the difference in the diffusion time featured by Rh6G molecules below and above the CMC. In the

    former, R6hG molecules interact with molecularly dissolved chains, which diffuse rapidly, in contrast

    to the latter case, when they associate with the hydrophobic domains of larger self-assembled parti-

    cles which moves slower, thus featuring lower diffusion coefficient. Moreover, through the recorded

    diffusion time, the diffusion coefficient can be known, and therefore the hydrodynamic radius of the

    micelles/aggregates can be determined, using the Stokes-Einstein equation, as discussed in the next

    sections.

    2.2.2 Experimental

    FCS measurements were performed using a ConfoCor2 from Carl Zeiss Jena GmbH. An Ar+

    laser operated at 488 nm, a motorized pinhole of diameter 80 µm, a C-Apochromat 40x/1.2 water

    immersion objective, a BP 530-600 emission filter and an HFT 488 plate beam splitter. A Lab-tek

    8-well chambered coverglass from Nalge Nunc International was used as a sample chamber. Each

    measurement (run) time was 60 seconds, and the measurements were repeated 10-15 times, and the

    correlation functions were averaged. The laser power was attenuated to 0.1-1.0% of maximum power

    (200mW) to avoid bleaching of the dye probe. The count rate was maintained above 5 kcounts/s,

    usually 5-15 kcounts/s. The resulting autocorrelation curves were fitted using Equation 2.5 [104], and

    were normalized with respect to the lowest value of the fitted range (τ = 0.004 ms).

    G(τ) = 1 +1

    N

    [1 +

    TT1− TT

    exp

    τT

    )] n∑i=1

    ρi(1 +

    τ

    τD,i

    )(1 +

    1

    (z0/w0)2τ

    τD,i

    ) 12

    (2.5)

    In Equation 2.5, N is the total number of fluorescent particles in the detection volume, n the

    number of different fluorescent species, τD,i the diffusion time of the i th species, ρi the amplitude

    24

  • of the i th species, and z0 and w0 the half-height and half-width of the detection volume, respectively.

    TT and τT are the triplet fraction and time, respectively. From the fit of 1 and 2 difussing species (n

    = 1 and 2 in Equation 2.5), the triplet parameters were found to be TT = 0.1-0.2 and τT =1-3 µs. w0

    was determined by measuring the diffusion time of Rhodamine 6G - τD,Rh6G - (Sigma-Aldrich, DRh6G

    = 2.8×10-10 m2·s-1 [92]), and by using the relation:

    w0 =√

    4DRh6G · τD,Rh6G (2.6)

    A value w0 ∼= 0.2 µm was obtained. The aspect ratio of the confocal volume, z0/w0 (otherwise

    known as structure factor S, which describes the axial ratio of the detection) was determined from the

    fit of the Rh6G correlation function in water and was typically 6.

    To study the influence of the pH at room temperature, on the aggregation behavior of the polymer

    under study by means of FCS, a stock solution of the PDPA-b-PMEO2MA diblock copolymer (4.5

    mg/mL) was prepared in glass vials by dissolving the polymer in deionized water, for subsequent

    dilution upon preparation of 1.5 mL samples of concentration ranging from 0.01 to 1.2 mg/mL. An

    aliquot of the fluorescent dye Rh6G was added to the prepared polymer sample solutions and to the

    control sample (deionized water with Rh6G), in such way that the final concentration of Rh6G was

    4.79×10-7 mg/mL (10-9 M) to avoid signal saturation upon FCS analysis. After being briefly agitated,

    the pH of the solutions was adjusted dropwise, by adding a strong acidic/alkaline aqueous solution,

    so the sample concentration would not change significantly (10 µL of 0.1 M solution of either HCl

    or NaOH). The deionized water, HCl and NaOH solutions were filtered (Rotilabo syringe filter; pore

    diameter, 0.22µm) to remove dust particles, before the addition of the polymer and the Rh6G dye. The

    pH was measured with a pH meter carefully cleaned with precision tissues to avoid contamination of

    the sample by dust or other impurities. Using this sample preparation routine, 4 series of polymer

    solutions were prepared having concentrations between 0.01 to 1.2 mg/mL and pH 3, 6, 7 and 10,

    taking into account that the pKa of the pH-responsive block monomer DPA is ~6.4. The solutions were

    agitated overnight and their pH measured afterwards. Any substantial deviation from the envisaged

    pH value was corrected by adding more HCl or NaOH solution and the samples were agitated again

    overnight.

    To estimate the interesting temperature range, the polymer cloud-point temperature at different pH

    values was assessed through turbidimetry measurements, at a wavelength of 500 nm (Varian Cary

    50 UV-Vis photo spectrometer equipped with a temperature controller Varian Cary, Single Cell Peltier

    Acessory). The polymer was dissolved in dionized water at pH 3, 6, 7 and 10, adjusted dropwise

    using 0.03M HCl/NaOH solutions, resulting in four samples with a polymer concentration of 2 mg/mL.

    For each sample the temperature was increased in steps of 0.1°C at a rate of 0.1 °C.min-1 followed

    by a 10 min equilibration time. The resulting curves were normalized with respect to the transmittance

    values recorded for the initial temperature (20°C), for each pH condition. The cloud-point temperature

    was defined to be the highest temperature value matching a transmittance of 100%.

    In order to evaluate the influence of temperature and pH on the aggregation behavior of the di-

    25

  • block copolymer, temperature-resolved FCS experiments were performed at different pH values. The

    same technical parameters used in FCS measurements at room temperature (such as run time, laser

    power, etc). However, instead of the usual 8-well sample holder chamber used at room temperature,

    a custom made sample holder (Figure 2.8) able to accommodate a electrical conducting indium-tin-

    oxide (ITO) coated glass cover slips (8-12Ω, Spi Supplies) was used. The sample temperature was

    installed and controlled using a voltage generator coupled with a temperature control system (ther-

    mostat) [105]. Briefly, a voltage was imposed onto the ITO cover slip, forcing a current flow through

    it, causing an increase of its temperature. The temperature was measured by a Pt100 platinum resis-

    tance thermometer and controlled in the same fashion as a thermostat.

    Figure 2.8: Sketch of the temperature-resolved FCS frame setup. (a) Top view, (b) Side view. A potential dif-ference is imposed on the copper pads leading to current passage through the ITO glass which in turn heatsup. The temperature is controlled by negative feedback around a given value defined by the user. The sameapplies for the heated oil-immersed objective. A PVC box serves as a protection from both the laser and thermalinsulation when filled with foam. (Dimensions not to scale).

    The temperature is set by the user and fluctuates by ±0.2°C around the mean value. Several

    other steps were performed to optimize the process: (i) placing of thermopaste onto the contact point

    between the Pt100 sensor a