152
Thesis Reference Intra-articular sustained release carriers for osteoarthritis management: formulation and bioactivity evaluation studies DE LACERDA SALGADO, Carlota Abstract Les différentes études présentées dans cette thèse ont permis de faire face et ont tenté de répondre à un certain nombre de défis rencontrés dans la recherche de l'IA DDS pour le traitement de l'arthrose. Tout d'abord, le potentiel des molécules thérapeutiques pour l'administration locale. Les produits naturels sont de riches sources de molécules thérapeutiques ; les médicaments contre l'arthrose potentiellement modificateurs de la maladie (DMOAD) et les voies moléculaires méritent d'être explorées plus avant en appliquant des modèles précliniques précis de l'arthrose. Ensuite, le développement de systèmes d'administration de médicaments par des vecteurs polymériques et enfin, l'établissement de modèles in vitro précis d'arthrose pour l'évaluation des résultats de ce type de formulations. Les microparticules polymères avec des nanoparticules de médicament broyées par wet milling et des taux de chargement de médicament réglables semblent répondre à plus d'une demande de traitement de l'arthrose par IA. Pour améliorer la conception et le développement de ces DDS IA, il est nécessaire [...] DE LACERDA SALGADO, Carlota. Intra-articular sustained release carriers for osteoarthritis management: formulation and bioactivity evaluation studies. Thèse de doctorat : Univ. Genève, 2021, no. Sc. 5546 DOI : 10.13097/archive-ouverte/unige:150608 URN : urn:nbn:ch:unige-1506080 Available at: http://archive-ouverte.unige.ch/unige:150608 Disclaimer: layout of this document may differ from the published version. 1 / 1

Thesis - archive-ouverte.unige.ch

  • Upload
    others

  • View
    22

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Thesis - archive-ouverte.unige.ch

Thesis

Reference

Intra-articular sustained release carriers for osteoarthritis

management: formulation and bioactivity evaluation studies

DE LACERDA SALGADO, Carlota

Abstract

Les différentes études présentées dans cette thèse ont permis de faire face et ont tenté de

répondre à un certain nombre de défis rencontrés dans la recherche de l'IA DDS pour le

traitement de l'arthrose. Tout d'abord, le potentiel des molécules thérapeutiques pour

l'administration locale. Les produits naturels sont de riches sources de molécules

thérapeutiques ; les médicaments contre l'arthrose potentiellement modificateurs de la

maladie (DMOAD) et les voies moléculaires méritent d'être explorées plus avant en

appliquant des modèles précliniques précis de l'arthrose. Ensuite, le développement de

systèmes d'administration de médicaments par des vecteurs polymériques et enfin,

l'établissement de modèles in vitro précis d'arthrose pour l'évaluation des résultats de ce type

de formulations. Les microparticules polymères avec des nanoparticules de médicament

broyées par wet milling et des taux de chargement de médicament réglables semblent

répondre à plus d'une demande de traitement de l'arthrose par IA. Pour améliorer la

conception et le développement de ces DDS IA, il est nécessaire [...]

DE LACERDA SALGADO, Carlota. Intra-articular sustained release carriers for

osteoarthritis management: formulation and bioactivity evaluation studies. Thèse de

doctorat : Univ. Genève, 2021, no. Sc. 5546

DOI : 10.13097/archive-ouverte/unige:150608

URN : urn:nbn:ch:unige-1506080

Available at:

http://archive-ouverte.unige.ch/unige:150608

Disclaimer: layout of this document may differ from the published version.

1 / 1

Page 2: Thesis - archive-ouverte.unige.ch

UNIVERSITÉ DE GENÈVE

Section des Sciences Pharmaceutiques

Laboratoire de Technologie Pharmaceutique

FACULTÉ DES SCIENCES

Prof. Eric Allémann

Dr. Olivier Jordan

Intra-articular sustained release carriers for

osteoarthritis management: formulation and

bioactivity evaluation studies

THÈSE

présentée à la Faculté des Sciences de l’Université de Genève

pour obtenir le grade de Docteur ès Sciences, mention Sciences Pharmaceutiques

par

Carlota SALGADO

de

Lisboa (Portugal)

Thèse N°: 5546

Genève

Atelier de reproduction Repromail

2021

 

 

 

Page 3: Thesis - archive-ouverte.unige.ch
Page 4: Thesis - archive-ouverte.unige.ch

 

 

 

 

 

 

 

 

 

 

 

 

 

Ao meu querido avô João, para sempre o meu maior fã.

Saudações benfiquistas!

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Page 5: Thesis - archive-ouverte.unige.ch

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Page 6: Thesis - archive-ouverte.unige.ch

Table of contents

 

 

 

 

 

 

Acronyms

1

Introduction

5

Chapter I Osteoarthritis in vitro models: applications and implications in development of intra-articular drug delivery systems

11

Chapter II In vitro anti-inflammatory activity in arthritic synoviocytes of A. brachypoda root extracts and its unusual dimeric flavonoids

43

Chapter III Nano wet milled celecoxib extended release microparticles for local management of chronic inflammation

69

Chapter IV Sustained release carriers for osteoarthritis: in vitro evaluation of an anti-inflammatory drug and a chondrogenic drug on a 3D chondrocyte OA model

103

Conclusions and perspectives 135

French summary 141

Acknowledgements 143

Page 7: Thesis - archive-ouverte.unige.ch

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Page 8: Thesis - archive-ouverte.unige.ch

1

Acronyms

 

ADAMTs Disintegrin and metalloproteinase with thrombospondin motifs

IL-6 Interleukin-6

COX-2 Cyclooxygenase 2 KGN Kartogenin CXB Celecoxib MMPs Matrix metalloproteinases

DAPI 4′,6-diamidino-2-phenylindole dihydrochloride

MPs Microparticles

DDS Drug delivery systems MPLC Medium pressure liquid chromatography

DLS Dynamic light scattering NMR Nuclear magnetic resonance

DMMB 1,9-dimethylmethylene blue

NSAIDs Nonsteroidal anti-inflammatory drugs

DMOADs Disease-modifying osteoarthritic drugs

OA Osteoarthritis

DMSO Dimethyl sulfoxide PDGF-BB Platelet-derived growth factor ECM Extracellular matrix PDLLA Poly D, L-lactic acid

ELISA Enzyme-linked immunosorbent assay

PGE2 Prostaglandin E2

FA Formic acid PLA Poly(lactic acid)

FGF-2 Basic fibroblast growth factor 2

PLGA Poly(lactic-co-glycolic acid)

GAG Glycosaminoglycans qPCR Quantitative polymerase chain

reaction

HA Hyaluronic acid

SEM Scanning electron microscopy

HBMSCs Human bone marrow mesenchymal stem cells

TFA Trifluoroacetic acid

HFLS Human fibroblast-like synoviocytes

TGFβ-1 Transforming growth factor beta 1

HPLCHigh performance liquid chromatography

TNFα Tumor necrosis factor alpha

HSCCC High-speed counter current chromatography

TPGS D-α-tocopherol polyethylene glycol 1000 succinate

IA Intra-articular UHPLC Ultra high performance liquid

chromatography IL-1β Interleukin-1 beta XRPD X-ray powder diffraction  

 

Page 9: Thesis - archive-ouverte.unige.ch

 

Page 10: Thesis - archive-ouverte.unige.ch

 

 

 

 

 

Introduction

Page 11: Thesis - archive-ouverte.unige.ch
Page 12: Thesis - archive-ouverte.unige.ch

Introduction

5

BACKGROUND

Osteoarthritis (OA) is a chronic, degenerative disease. This most common form of

arthritis has a high and ever-increasing prevalence in the aging population

worldwide (> 65 years). As a disease of the whole joint, it is characterized by

inflammation and pain, ultimately leading to movement impairment. As one of the

leading disability causes in the elderly, the socio-economic burden caused by OA is

substantial [1,2]. Presently, pharmacological treatment approaches are based on

symptom management by oral analgesic and anti-inflammatory drugs [3]. However,

there is a current unmet need for a cure: by better, more efficacious

formulations/administration routes or disease-modifying drugs (DMOADs) that help

slow and potentially revert the disease. In recent years, significant efforts have been

made in the research and development of novel molecules and improved drug

delivery systems [4,5]. Natural products have become appealing in the field of OA

due to the potential to tackle both inflammatory and cartilage degradation pathways

[6]. Despite common drug development hurdles, like physicochemical issues such

as solubility and development of clinical trials, the complexity of OA as a disease

proves an obstacle for the successful development of DMOADs. A great number of

target tissues (synovium, cartilage, bone) plus a myriad of molecular targets of pain

and inflammation is involved in OA mechanisms, making drug discovery a

challenge. Lastly, the variety of OA disease phenotypes, onset and clinical

manifestations hampers the development of accurate and predictive pre-clinical and

clinical trials [4,7]. Additionally, the joint as a target, with poor irrigation and blood

supply, decreases the efficacy of systemic oral therapy. Intra-articular (IA) local

delivery of therapeutic molecules has been explored as an alternative that bypasses

issues of systemic oral therapy, such as low joint retention and adverse effects. A

controlled, sustained release of molecules directly into the joint capsule has been

sought after with the development of drug delivery systems. IA drug delivery

systems like nano- and micro-carriers, hydrogels or liposomes have been

investigated in recent years [5,8]. In both DMOADs and IA drug delivery systems

research, pre-clinical trials are an essential and crucial step. In this phase of

development, it is crucial to consider accurate and highly predictive in vitro and in

vivo OA models for increased translation into humans. Better design, prediction, and

translation into humans are key in furthering OA research [9,10].

Page 13: Thesis - archive-ouverte.unige.ch

Introduction  

6

SCOPE AND AIMS

This thesis is based on three different axes: research of novel molecules with

potential as DMOADs; development of polymeric IA drug delivery systems for local

delivery of OA treatments and design and establishment of accurate in vitro OA

models. It follows previous projects led by Prof. Eric Allémann and Dr. Olivier

Jordan, thus continuing a 10-year research effort on novel OA treatments [11,12].

The main goal was the development of IA treatment options for OA by exploring

novel molecules and developing appropriate drug delivery systems. Throughout this

work, another aim was to accurately represent in vitro OA, with the establishment of

cellular models and adequate outcome evaluation.

GENERAL STRUCTURE

In an attempt to better design and formulate IA drug delivery systems for OA,

Chapter I reviews the literature on different in vitro models of OA. Models are

described and discussed regarding their specific application in the study and

development of IA drug delivery systems, such as polymeric micro and

nanoparticles. Chapter II explores the potential of natural plant extracts as local IA

treatment of OA. Root extracts of Arrabidaea brachypoda, a Brazilian shrub, are

known for their analgesic and anti-inflammatory properties in arthritics joints. In this

study, two different root extracts and three novel flavonoids isolated from one of the

extracts are evaluated in vitro, in human fibroblast-like synoviocytes, for their

potential as local OA therapeutic options. In Chapter III, approaches tackling the

inflammation component of OA are further explored. In this study, celecoxib, a non-

steroidal anti-inflammatory drug (NSAID) with established clinical use in OA, was

encapsulated into polymeric microparticles with the goal of developing a sustained,

controlled release IA drug delivery system. Using a spray-drying method, celecoxib

was incorporated, at three different drug loadings, as a drug solution or embedded

as nano-milled drug particles into PLA microparticles. Aiming at a long-lasting

therapeutic effect in OA, the main goal of this study was to assess the effects of the

incorporation of nano-milled drug, at high drug loadings, on the sustained prolonged

release of celecoxib. The same spray-drying method was applied to a different drug,

kartogenin, in Chapter IV. This chapter aimed to compare the performance of two

distinct molecules, incorporated in PLA microparticles as nano-milled drug at 10 %

drug loading: celecoxib and kartogenin. In the first part, both drug delivery systems

Page 14: Thesis - archive-ouverte.unige.ch

Introduction

7

were characterized. In the second part, human articular chondrocytes were

collected from three donors to form pellets of hyaline cartilage as a 3D in vitro model

of OA. The objective of this study was to determine the effects of a sustained

prolonged release of two molecules targeting two OA pathways (inflammation and

cartilage degradation) in a 3D in vitro model.

 A final section of Conclusions and perspectives serves as a summary of all

described studies and explores progress in IA drug delivery systems research for

OA treatment.

REFERENCES

[1] D.J. Hunter, S. Bierma-Zeinstra, Osteoarthritis, Lancet. 393 (2019) 1745–1759. https://doi.org/10.1016/S0140-6736(19)30417-9.

[2] J. Puig-Junoy, A. Ruiz Zamora, Socio-economic costs of osteoarthritis: A systematic review of cost-of-illness studies, Semin. Arthritis Rheum. 44 (2015) 531–541. https://doi.org/10.1016/j.semarthrit.2014.10.012.

[3] S.L. Kolasinski, T. Neogi, M.C. Hochberg, C. Oatis, G. Guyatt, J. Block, L. Callahan, C. Copenhaver, C. Dodge, D. Felson, K. Gellar, W.F. Harvey, G. Hawker, E. Herzig, C.K. Kwoh, A.E. Nelson, J. Samuels, C. Scanzello, D. White, B. Wise, R.D. Altman, D. DiRenzo, J. Fontanarosa, G. Giradi, M. Ishimori, D. Misra, A.A. Shah, A.K. Shmagel, L.M. Thoma, M. Turgunbaev, A.S. Turner, J. Reston, 2019 American College of Rheumatology/Arthritis Foundation Guideline for the Management of Osteoarthritis of the Hand, Hip, and Knee, Arthritis Rheumatol. 72 (2020) 220–233. https://doi.org/10.1002/art.41142.

[4] A. Latourte, M. Kloppenburg, P. Richette, Emerging pharmaceutical therapies for osteoarthritis, Nat. Rev. Rheumatol. (2020). https://doi.org/10.1038/s41584-020-00518-6.

[5] P. Maudens, O. Jordan, E. Allémann, Recent advances in intra-articular drug delivery systems for osteoarthritis therapy, Drug Discov. Today. 23 (2018) 1761–1775. https://doi.org/10.1016/j.drudis.2018.05.023.

[6] G.-E. Deligiannidou, R.-E. Papadopoulos, C. Kontogiorgis, A. Detsi, E. Bezirtzoglou, T. Constantinides, Unraveling Natural Products’ Role in Osteoarthritis Management—An Overview, Antioxidants. 9 (2020) 348. https://doi.org/10.3390/antiox9040348.

[7] M.A. Tryfonidou, G. de Vries, W.E. Hennink, L.B. Creemers, “Old Drugs, New Tricks” – Local controlled drug release systems for treatment of degenerative joint disease, Adv. Drug Deliv. Rev. 160 (2020) 170–185. https://doi.org/10.1016/j.addr.2020.10.012.

[8] L.M. Mancipe Castro, A.J. García, R.E. Guldberg, Biomaterial strategies for improved intra‐articular drug delivery, J. Biomed. Mater. Res. Part A. (2020) 1–11. https://doi.org/10.1002/jbm.a.37074.

[9] T.L. Vincent, Of mice and men: converging on a common molecular understanding of osteoarthritis, Lancet Rheumatol. 2 (2020) e633–e645. https://doi.org/10.1016/S2665-9913(20)30279-4.

[10] Y. He, Z. Li, P.G. Alexander, B.D. Ocasio-Nieves, L. Yocum, H. Lin, R.S. Tuan, Pathogenesis of osteoarthritis: Risk factors, regulatory pathways in chondrocytes, and experimental models, Biology (Basel). 9 (2020) 1–32. https://doi.org/10.3390/biology9080194.

[11] J. Pradal, M.-F. Zuluaga, P. Maudens, J.-M. Waldburger, C.A. Seemayer, E. Doelker, C. Gabay, O. Jordan, E. Allémann, Intra-articular bioactivity of a p38 MAPK inhibitor and development of an extended-release system, 93 (2015) 110–117. https://doi.org/10.1016/j.ejpb.2015.03.017.

Page 15: Thesis - archive-ouverte.unige.ch

Introduction  

8

[12] P. Maudens, C.A. Seemayer, C. Thauvin, C. Gabay, O. Jordan, E. Allémann, Nanocrystal-Polymer Particles: Extended Delivery Carriers for Osteoarthritis Treatment, Small. 14 (2018) 1703108. https://doi.org/10.1002/smll.201703108.

Page 16: Thesis - archive-ouverte.unige.ch

 

 

Chapter I  

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Page 17: Thesis - archive-ouverte.unige.ch

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Page 18: Thesis - archive-ouverte.unige.ch

 

11

Chapter I

Osteoarthritis in vitro models: applications and implications

in development of intra-articular drug delivery systems Carlota Salgado a,b, Olivier Jordan a,b and Eric Allémann a,b

a School of Pharmaceutical Sciences, University of Geneva, 1 Rue Michel-Servet 1211 Geneva, Switzerland b Institute of Pharmaceutical Sciences of Western Switzerland, University of Geneva, 1 Rue Michel-Servet 1211 Geneva, Switzerland

Published: Pharmaceutics, 2021, 13 (1), 60; https://doi.org/10.3390/pharmaceutics13010060

ABSTRACT

Osteoarthritis (OA) is a complex multi-target disease with an unmet medical need for the development of therapies that slow and potentially revert disease progression. Intra-articular (IA) delivery has seen a surge in osteoarthritis research in recent years. As local administration of molecules, this represents a way to circumvent systemic drug delivery struggles. When developing intra-articular formulations, the main goals are a sustained and controlled release of therapeutic drug doses, taking into account carrier choice, drug molecule, and articular joint tissue target. Therefore, the selection of models is critical when developing local administration formulation in terms of accurate outcome assessment, target and off-target effects and relevant translation to in vivo. The current review highlights the applications of OA in vitro models in the development of IA formulation by means of exploring their advantages and disadvantages. In vitro models are essential in studies of OA molecular pathways, understanding drug and target interactions, assessing cytotoxicity of carriers and drug molecules, and predicting in vivo behaviors. However, further understanding of molecular and tissue-specific intricacies of cellular models for 2D and 3D needs improvement to accurately portray in vivo conditions.

Keywords: osteoarthritis; intra-articular drug delivery systems; synovium; cartilage; in vitro cellular models; synoviocytes; chondrocytes

Page 19: Thesis - archive-ouverte.unige.ch

Chapter I  

12

1. INTRODUCTION

Osteoarthritis (OA) is a chronic disease with worldwide incidence in the population

aged 65 years and higher, representing a significant economic burden in terms of

global health [1,2]. As the most common form of arthritis and one of the leading

causes of disability in the elderly population, OA is characterized by chronic

inflammation, articular cartilage degeneration and structural changes of whole joints.

There is currently an unmet need for disease-modifying drugs (DMOADs) that slow

or even revert disease progression [3–5]. Pharmacological treatment options focus

on symptom management. Oral analgesic and nonsteroidal anti-inflammatory drugs

(NSAIDs) are first-line treatments for pain and inflammation. However, since OA

mainly affects the joint as a whole closed structure, systemic drugs result in less

than optimal efficacy rates [6,7]. A known alternative that circumvents most of the

drawbacks associated with systemic drug administration is the delivery of drugs

locally, by intra-articular (IA) injection. IA allows for higher drug doses and prolonged

delivery of drug molecules directly into affected joints. By this approach, more

effective relief of symptoms may be attained, while systemic adverse effects are

generally avoided. Different drug delivery systems (DDSs) have grown in the field

to improve the delivery of small molecules locally to joints. These include different

formulations such as polymeric nano and microparticles, hydrogels, liposomes and

micelles, which have been extensively reviewed [8–11]. Due to its local

administration, maintaining the selectivity of drug molecules and the carrier system

towards biological tissue targets in the joint while avoiding off-target effects is critical

when developing IA formulations. In this regard, the design of predictive in vitro OA

models is crucial in characterizing and understanding the studied drug delivery

systems for OA treatment. Different cellular models represent different tissues of the

joint: synoviocytes, the synovium, and chondrocytes are used to model articular

cartilage [12]. The different types of in vitro cellular models (i.e., monolayer, three-

dimensional or explant) have various applications according to the final goals of IA

formulation. Thus, a deep understanding of their intricacies is very important in this

field. The purpose of the present review is to discuss the relevance of the different

in vitro OA models in the development of IA formulations for OA treatment. At first,

an overview of the latest (5 years) intra-articular DDSs is presented, highlighting the

choice of in vitro model for each formulation. In this review, viscosupplementation

Page 20: Thesis - archive-ouverte.unige.ch

OA in vitro models  

13

formulations and delivery of cells (mesenchymal stem cells and platelet rich plasma)

have been excluded. The review focuses on nano and micro carriers, hydrogels and

liposomes containing drug molecules. Next, advantages and disadvantages, as well

as possible readable markers and targets of different in vitro OA models, are

discussed, based on their relevance for the development of intra-articular

formulations.

2. OSTEOARTHRITIS

Osteoarthritis is a chronic degenerative disease of the whole joint. It is characterized

by chronic inflammation, articular cartilage degeneration and structural changes in

several joint tissues. Age >65 years old, obesity, gender (double prevalence in

females), previous joint injuries and genetic predisposition to joint complications are

all considered risk factors in the development of mild to severe OA [2,13]. Other than

the economic burden it represents, OA is one of the primary causes of disability in

the elderly population. Considered the most common form of arthritis, its worldwide

incidence has repercussions on more than 100 million people [1,14]. The etiology of

OA is unknown (primary OA) in the majority of cases, with secondary OA (one that

follows joint injury) as an example of how trauma to the joint influences further

disease progression. Several biomechanical and molecular processes are known to

kick-start the pathology cycle. Tissue alterations of articular cartilage from increased

cell proliferation and microarchitectural changes to the structure of subchondral

bone are considered key events [15]. In early stages, degradation products of

proteoglycan and collagen are released into the joint cavity from hyaline cartilage.

This phenomenon stimulates immune cells from the synovial membrane to release

pro-inflammatory cytokines - mainly IL-1β, IL-6 and TNFα. This inflammatory state

induces catabolic mechanisms by the chondrocytes that produce matrix

metalloproteinase (MMPs) 1, 3 and 13 and aggrecanases 1 and 2 (disintegrin and

metalloproteinase with thrombospondin motifs - ADAMTS). Cartilage is further

degraded, and the inflammatory state is perpetuated. Due to its poor vascularization

and low cellular density, cartilage has a limited regeneration turnover. As disease

advances, catabolic mechanisms outweigh those of repair by the extracellular matrix

(ECM) [16–18]. As a result, there is a narrowing of the joint space due to cartilage

degradation, subchondral bone erosion with the formation of osteophytes and small

cysts, inflammation of the synovium (synovitis), and overall joint function loss (Figure

Page 21: Thesis - archive-ouverte.unige.ch

Chapter I  

14

1). Clinical manifestations of the disease with well-established symptoms, mainly

joint pain, stiffness and, consequently, a decrease in daily movement, appear

relatively late. When detected and adequately diagnosed by physical assessment

and bioimaging (X-ray, MRI), OA has often progressed to a stage where preventive

and possibly reverting measures are no longer efficacious, leaving symptom

management as the only option [19]. Currently, there is a substantial unmet need

for disease-modifying OA drugs (DMOADs) that actively slow disease progression

as no molecule of the sort has been approved or introduced in the market.

Throughout the management of OA, different non-pharmacological treatments are

adopted, like physical therapy, weight management and the use of different dietary

supplements. Pharmacological treatment regimens depend on disease stage (I to

IV, minimal to severe). Analgesics like paracetamol and nonsteroidal anti-

inflammatory drugs (NSAIDs) such as diclofenac are first-line treatments. However,

the drawbacks and adverse effects associated with the use of these systemic drugs

are limiting [20–22]. In further stages of the disease, local administration (intra-

articular) is an alternative that circumvents these issues. This local administration of

hyaluronic acid derivatives, known as viscosupplementation, combines pain relief

and improvement of joint motion from the greater cushioning effect provided by the

hydrogels. Other biological compounds, like injections of autologous platelet-rich

plasma have also been explored as local treatment of OA [23,24]. When symptom

management is no longer viable in later stages, full joint replacement surgery of hip,

knee or heel is an option [5,6].

Figure 1. Schematic representation of the structural composition of healthy and osteoarthritic knee.

Page 22: Thesis - archive-ouverte.unige.ch

OA in vitro models  

15

3. INTRA-ARTICULAR DRUG DELIVERY SYSTEMS AND

INTERACTIONS WITH OA JOINTS

In a clinical setting, despite progress in OA research and development of disease-

modifying drugs, joint anatomy and physiology still pose a challenge for effective

drug delivery. Systemic drug delivery is challenging due to the poor irrigation and

limited permeability of the synovial membrane and articular capsule of affected

joints. Local, intra-articular administration of small molecules and larger protein

products directly as solutions into joints is hindered by low retention times due to

fast clearance [25]. Therefore, IA administration is not used to deliver common

analgesic and anti-inflammatory drugs to the joint. As approved and in use, IA is

only applied to deliver glucocorticoids (GC) and hyaluronic acid (HA) for

viscosupplementation [26–28]. The lack of broader use of IA administration as a

drug delivery route for OA treatment might be due to some drawbacks such as

formulation issues and the invasiveness of the procedure, which limit the number of

yearly injections. Thus, attaining drug loadings high enough to release sufficient

therapeutic drug doses over extended periods represents a critical challenge [29].

Drug delivery systems with extended-release properties help circumvent these

issues and others, like potential low aqueous solubility of many molecules.

Comprehensive reviews on formulation aspects of IA DDSs have been published in

recent years [26,28,30,31]. Table 1 shows the drug delivery systems investigated in

the past five years to treat OA by IA administration. Different types of formulations—

micro- and nanoparticles, hydrogels, liposomes—allow for controlled and extended

release of drug and increased retention times in joints while avoiding systemic side

effects [10,32,33]. Various classes of molecules have been investigated as

DMOADs or for symptom management: analgesic/anti-inflammatory,

chondroprotective/regenerative and bone resorption inhibitors [28]. Each category

is linked to different target tissues in the joint. For example, anti-inflammatory drugs

like celecoxib target the synovium, and chondroprotective drugs like kartogenin

target articular cartilage tissue [34,35]. It is essential to consider tissue specificity

when formulating IA drug delivery systems as off-target effects may occur and

negatively impact OA progression. Understanding the structure of the different

tissues of joints is thus key for the design of DDSs. Human joints are complex

structures that connect bones, allowing the body’s movement. The main structures

Page 23: Thesis - archive-ouverte.unige.ch

Chapter I  

16

of synovial joints (diarthrosis, joints with movements) (Figure 1) are joint capsule,

synovial membrane, joint cavity with synovial fluid, cartilage, ligaments, muscles,

bursae, tendons, subchondral bone, nerves and vessels [36,37]. Synovial joints are

the most affected by OA, with two of the main features that make them unique being

key explored targets of therapeutic treatments: the synovial membrane and the

articular cartilage. Synovium or synovial membrane is the connective tissue that

lines the joint cavity. This heterogeneous tissue mainly comprises two types of

synoviocytes: type A macrophage-like synoviocytes, lesser in number and

increased in inflammatory conditions, have an important role in phagocytosis and

production of pro-inflammatory cytokines; and type B fibroblast-like synoviocytes,

the structural cells of the synovium (75% of cellular total), producing synovial fluid

and ECM components. Collagen fibers, fenestrated blood capillaries and lymph

vessels are other structures found in the inner layers of the membrane. The synovial

fluid, produced by ultrafiltration of plasma, nourishes the non-irrigated articular

cartilage, lubricates and absorbs shock [36,38]. Drug delivery systems with drugs

targeting the synovium, like TSG-6 (TNFα gene precursor) or VX-745 (p38 MAPK

inhibitor) are active on type B synoviocytes and macrophages, mostly through

inflammatory and pain pathways [39,40]. On the other hand, articular cartilage

(hyaline cartilage) is a connective tissue layer that lines the ends of the bones of the

joint, serving as a barrier to friction and shock between them. Contrary to the

synovium, this is an avascular, alymphatic and aneural tissue. It is composed of

chondrocytes (differentiated mature cells) and ECM, mainly collagen and elastin

fibers, aggrecan and proteoglycans. Its form and elasticity are determined by the

organization of the collagen fibers, proteoglycans and diffusion of water molecules

during movement. The lubricants and hyaluronic acid secreted both by the synovial

fluid and chondrocytes are shock-absorbing and provide a cushioning effect. This

cross-talk between tissues and synovial fluid is driven by mechanical load caused

by body movement. Cartilage is part of the osteochondral unit as it covers the sub-

chondral bone plate [41,42]. Examples of drugs targeting cartilage and bone

delivered by IA administration of delivery systems include kartogenin (a

chondrogenesis inductor from the RUNX-1 pathway) and doxycycline (an antibiotic

with MMP inhibitor functions) [43,44].

The successful development of an IA drug delivery system formulation greatly

depends on its interaction with the target tissue in the joint. Accurate choice and

Page 24: Thesis - archive-ouverte.unige.ch

OA in vitro models  

17

design of in vitro models of OA are crucial in understanding target interaction,

predicting in vivo outcomes, and developing effective IA formulations.

Page 25: Thesis - archive-ouverte.unige.ch

Chapter I  

18

Table 1. Intra-articular small molecule drug delivery systems for OA treatment, developed in the past 5 years. (Acronyms defined at the bottom of table)

Formulation Drug Carrier Type of study Main target tissue In vitro model In vivo model Authors;

Year; References

Microparticles

Doxycycline PCL Pre-clinical studies Cartilage3D rabbit

chondrocyte agarose model

Rabbits Aydin et al. 2015 [44]

Celecoxib PEA Pre-clinical studies Synovium Differentiated HI-

60 cells and lysates

Human synovium and synovial fluid (ex vivo);

rat ACLT model

Janssen et al. 2016 [32]

Etoricoxib PCL Pre-clinical studies Synovium, cartilage Not reported Rats Arunkumar et al.

2016 [45]

Lornoxicam Chitosan/TPP Pre-clinical studies Synovium, cartilage Not reported Rat MIA model Abd-Allah et al.

2016 [46]

Fluvastatin PLGA Pre-clinical studies CartilageHuman primary chondrocytes Rabbit ACLT model

Goto et al. 2017 [47]

Rhein (cassic acid) PLGA Pre-clinical studies Synovium THP-1

macrophages Not reported

Gomez-Gaete et al. 2017 [8]

Kartogenin PLA Pre-clinical studies CartilageHuman

synoviocytes Mice DMM model Maudens et al.

2018 [43] PH-797804,

Dexamethasone PLA Pre-clinical studies Synovium

Human synoviocytes

Mice AIA model Maudens et al.

2018 [48]

Triamcinolone acetonide (Zilretta™)

PLGA Phase II/III clinical trials in

OA patients 1 Synovium, cartilage Not reported Rat knee model 2

Kumar et al. 2015 2;

Kraus et al. 2018 1[49,50]

TSG-6 (tumor necrosis factor-alpha stimulated

gene-6) Heparin Pre-clinical studies Cartilage Not reported Rat MMT model

Tellier et al. 2018 [39]

Fluticasone propionate PVA Pre-clinical studies Synovium Not reported Beagle dogs Getgood et al.

2019 [51]

Celecoxib PLA Pre-clinical studies Synovium Human synoviocytes

Not reported Salgado et al. 2020 [52]

Rapamycin PLGA Pre-clinical studies Cartilage Human immortal chondrocytes

Mice Dhanabalan et al. 2020 [53]

Page 26: Thesis - archive-ouverte.unige.ch

OA in vitro models  

19

Nanoparticles

VX-745 (p38 MAPK inhibitor)

PLA and PLGA

Pre-clinical studies Synovium Human

synoviocytes Mice AIA model

Pradal et al. 2015 [40]

Dexamethasone Avidin/PEG Pre-clinical studies Synovium, cartilage Bovine knee

cartilage explants

Not reported Bajpayee et al.

2016 [54]

KAFAK (anti-inflammatory mitogen-

activated protein kinase-activated protein kinase 2

(MK2)-inhibiting cell-penetrating peptide)

pNiPAM-PEG Pre-clinical studies Synovium, cartilage Bovine knee

cartilage explants

Not reported Lin et al. 2016 [9]

Kartogenin; Diclofenac Chitosan/Plur

onic F127 Pre-clinical studies Synovium, cartilage

Human BMSCs (bone marrow mesenchymal

stem cells); Human primary chondrocytes

Rats Kang et al. 2016 [35]

Curcumin PLGA Pre-clinical studies Synovium, cartilage Not reported Rats Niazvand et al.

2017 [55]

Dexamethasone Avidin Pre-clinical studies Synovium Not reported Rabbit ACLT modelBajpayee et al.

2017 [56]

KAFAK pNiPAM-PEG Pre-clinical studies Synovium, cartilage

RAW 264.7 macrophages; Bovine knee

cartilage explants

Not reported McMasters et al.

2017 [57]

CAP (chondrocyte affinity peptide)

PEG-PAMAM Pre-clinical studies Cartilage Human primary chondrocytes

Rats Hu et al. 2018 [58]

Kartogenin Polyurethane Pre-clinical studies CartilageRat primary

chondrocytes Rat ACLT model

Fan et al. 2018 [59]

Adenosine PEG-b-PLA Pre-clinical studies Synovium, cartilage RAW 264.7

macrophages Rat ACLT model

Liu et al. 2019 [60]

Etoricoxib PLGA-PEG-

PLGA Pre-clinical studies Synovium, cartilage Human primary chondrocytes Rat ACLT model

Liu et al. 2019 [33]

Hyaluronic acid PLGA Pre-clinical studies CartilageRAW 264.7

macrophages Brine shrimp; Rats

Mota et al. 2019 [61]

Page 27: Thesis - archive-ouverte.unige.ch

Chapter I  

20

Hyaluronic acid and near-infrared dye

PLGA Pre-clinical studies CartilageHuman primary chondrocytes

Mice DMM model Zerrillo et al.

2019 [62]

Celastrol Mesoporous silica

Pre-clinical studies Cartilage Rat primary chondrocytes

Rat MIA model Jin et al. 2020 [63]

Diacerein PLGA Pre-clinical studies Synovium, cartilage Rat synoviocytes Rat MIA model Jung et al. 2020 [64]

Etoricoxib PLA/Chitosan Pre-clinical studies Synovium

MC3T3-E1 cells (mouse

osteoblast precursor)

Not reported Salama et al.

2020 [65]

MK2i (anti-inflammatory MK2-inhibiting peptide)

Linked and non-linked

NIPAm Pre-clinical studies Synovium, cartilage

Bovine primary chondrocytes

Rats Deloney et al.

2020 [66]

Oxaceprol PLGA Pre-clinical studies Synovium

Human primary LCLs

(lymphoblastoid cell lines)

Not reported Alarçin et al.

2020 [67]

Triamcinolone acetonide Dextran sulfate

conjugated Pre-clinical studies Synovium

RAW 264.7 macrophages;

L929 cells (mouse

fibroblast)

Mice MIA model She et al. 2020 [68]

Hydrogels

Amphotericin B Hyaluronic

acid/glyceryl monooleate

Pre-clinical studies Synovium, cartilage Not reported Rabbits Shan-Bin et al.

2015 [69]

Celecoxib PCLA-PEG-

PCLA Pre-clinical studies Synovium Not reported Horse

Petit et al. 2015 [34]

Methotrexate/dexamethasone/near-infrared dye

Hyaluronic acid + PLGA microcapsule

s

Pre-clinical studies Synovium RAW 264.7

macrophages Rat RA model

Son et al. 2015 [70]

Sinomenine hydrochloride

Phytantriol Formulation studies Not reported Not reported Not reported Chen et al. 2015 [71]

Dexamethasone Hyaluronic acid

Pre-clinical studies Synovium, cartilage Human primary chondrocytes

Rat ACLT model Zhang et al. 2016 [72]

Page 28: Thesis - archive-ouverte.unige.ch

OA in vitro models  

21

PEGylated Kartogenin Hyaluronic

acid Pre-clinical studies Cartilage

Human BMSCs; human primary chondrocytes

Rat ACLT model Kang et al. 2017 [73]

Celecoxib PCLA-PEG-

PCLA Pre-clinical studies Synovium Not reported Equine synovitis model

Cokeleare et al. 2018 [74]

Dexamethasone Hyaluronic acid/pNiPAM

Pre-clinical studies Synovium Human synoviocytes

Mice DMM model Maudens et al. 2018 [10]

Triamcinolone hexacetonide (Cingal®)

Hyaluronic acid

Phase II/III clinical trials in OA patients

Synovium, cartilage Not reported Not reported Hangody et al.

2018 [75]

Simvastatin Gelatin Pre-clinical studies CartilageMouse primary chondrocytes

Mice Tanaka et al.

2019 [76]

Dexamethasone Agarose gel +

PLGA microspheres

Pre-clinical studies Synovium, cartilage

3D canine articular

chondrocyte construct

Canine osteochondral autograft model

Stefani et al. 2020 [77]

Diclofenac

Hyalomer (HA and

poloxamer 407)

Pre-clinical studies Synovium, cartilage Not reported Rat MIA model Hanafy et al.

2020 [78]

Diclofenac

Linked PAPE (2-

Pyridylamino substituted 1-phenylethanol

)

Formulation studies Not reported Not reported Not reported Kawanami et al.

2020 [79]

Eicosapentanoic acid Gelatin Pre-clinical studies Synovium Human primary chondrocytes

Mouse DMM model Tsubosaka et al.

2020 [80]

Hyaluronic acid/diclofenac sodium

Silica colloidal crystal beads-

pNiPAM Pre-clinical studies Synovium, cartilage

Human primary chondrocytes

Rat DMM model Yang et al. 2020 [81]

Liposomes Quercetin

mPEG-PA (Methoxy-

poly(ethylene glycol)-l-

poly(alanine))

Pre-clinical studies Synovium, cartilage Human primary chondrocytes

Rat ACLT model Mok et al. 2020 [82]

Page 29: Thesis - archive-ouverte.unige.ch

Chapter I  

22

Fish oil protein encapsulated in gold

nanoparticles DPPC Pre-clinical studies Synovium Not reported Rats

Sarkar et al. 2019 [83]

Glucosamine sulphate Distearoyl

phosphocholine

Pre-clinical studies CartilageMouse primary chondrocytes

Not reported Ji et al.

2019 [84]

Rapamycin

DSPC combined with low-intensity pulsed

ultrasound

Pre-clinical studies Cartilage Human primary chondrocytes

Guinea pigs Chen et al. 2020 [85]

Acronyms: PCL: polycaprolactone; PEA: polyetheramine; PLGA: poly(lactic-co-glycolic acid); PLA: polylactic acid; PVAL: poly(vinyl alcohol); PCLA: poly(ε-caprolactone-co-lactide); PEG: polyethylene glycol; pNiPAM: poly(N-isopropylacrylamide); PAMAM: poly(amidoamine); DPPC: dipalmitoylphosphatidylcholine; DSPC: 1,2-distearoyl-sn-glycero-3-phosphocholine; MIA: monoiodoacetate; AIA: antigen-induced arthritis; ACLT: anterior cruciate ligament transection; DMM: destabilization of medial meniscus.

Page 30: Thesis - archive-ouverte.unige.ch

OA in vitro models  

23

4. IN VITRO MODELS OF OA

Grasping the complexity of OA pathophysiological mechanisms remains a challenge

in OA research and negatively reflects on the successful development of DMOADs.

Many OA in vitro and in vivo models have been developed and refined over the

years. However, there is still no confirmed gold standard in vitro model to apply when

developing OA drug molecules and/or drug delivery systems [12]. The

establishment of accurate in vitro models is crucial as these influence choice of in

vivo OA models. Although there are relevant in vivo OA animal models, major gaps

in translation from animal to human OA conditions still prevail. Smart design and

choice of in vitro models could potentially help bridge these gaps by enhancing

predictability of OA models. Current OA in vivo models have additional limitations.

These models often actively portray either post-traumatic and/or late-stage (III/IV)

OA, leaving a large gap in understanding the spontaneous occurring disease and

its early stages, where slowing disease progression would be an attractive treatment

strategy. Sustainability and 3R initiatives (refinement, reduction and replacement)

have to be considered to assess the usefulness of these in vivo animal models,

where in vitro OA models can become the best alternative [86]. In this case, result

translation and predictability from in vitro to in vivo models still lack refinement and

accuracy. The processes of naturally occurring OA in certain animal species have

been proven similar to those of humans. Therefore, tissue collection from affected

animals is essential in the development of in vitro and ex vivo early-stage OA

models. Tissue collection in humans (articular cartilage or synovium) is complex due

to several ethical and regulatory issues, and retrieval at early stages of the pathology

is nearly impossible. Recovery of samples is restricted to patients undergoing total

joint replacement surgeries where OA is far evolved [87,88]. In this context, various

in vitro OA cellular models have been designed and explored: monolayer (2D), 3D

with or without scaffolds and tissue explants. Each model is adapted to a unique

target tissue of the joint and yields quantification of different markers (e.g.,

inflammatory cytokines, collagen type II, aggrecan or MMPs). In addition, each

model has its intricacies with relevant advantages and disadvantages, discussed in

Table 2. In the field of IA DDSs, these are important, especially in characterizing

release mechanisms and cytotoxicity of carrier systems. In Table 2, the different

applications of each model in IA DDS pre-clinical development are further described.

Page 31: Thesis - archive-ouverte.unige.ch

Chapter I  

24

Table 2. Overview of advantages and disadvantages of in vitro OA models and their application in IA DDS development

In Vitro OA Model Advantages Disadvantages

Models Applied in IA DDS

Development (as per Table 1)

Outcome Evaluation (as per Table 1)

2D cellular culture

Monolayer

High throughput, low cost. Homogenous cell exposition to nutrients. Allows for differences in cellular phenotype studies [12]

Furthest from natural in vivo tissue conditions. High variability (different passages). Better suited for synoviocytes than chondrocytes. 2D substrate induces de-differentiation and changes in morphology [12]

- Synoviocytes (human, mouse and rat)

- Chondrocytes (human, murine, rat and bovine)

- Macrophages (human and murine)

- BMSCs (human)

RAW 264.7 macrophages [33,57,60,68,70]: - Cytotoxicity assays - Quantification of NO. cAMP,

IL-6, IL-1β and TNF-α Synoviocytes [10,40,43,48,52,64]: - Cytotoxicity and proliferation - Quantification of IL-6, PGE2,

IL-1β, TNF-α, MMP-3, MMP-13, COX-2 and ADAMTS-5

Chondrocytes [47,53,59,62,63,66,72,73,81,84]: - Cytotoxicity, apoptosis and

proliferation assays - Quantification of IL-6, IL-1β,

TNF-α, GAG/DNA, Aggrecan, Collagen II, MMP-1, MMP-3, TAC-1, MMP-13, COX-2, PGE2, iNOS and ADAMTS-5

- Senescence assays after genotoxic and oxidative stress [53]

- Expression of inflammatory transcription factors: p-IKKα/β [80]

Co-culture

Important in studies of cell-to-cell interactions and studies of influence of different cellular phenotypes together [12]

Expensive and difficult to maintain. Lacks in three-dimensional characteristics of cartilage growth [87]

(examples not included in Table 1) - Synoviocytes-

chondrocytes - Chondrocytes-

osteoblasts [89,90]

3D cellular culture

Without Scaffold

High similarity with in vivo tissue conditions as it maintains structure from ECM growth. Cellular phenotype is preserved. Important in studies of intercellular and cell to ECM relationship and loading capacity assays [88,91]

Expensive and difficult to maintain. Restricted throughput (hard to propagate without compromising cell quality). Nature of scaffold plays role in cellular growth [92]

- Chondrocyte pellets

- Hanging drop BMSCs

- Quantification of: GAG/DNA, Collagen II, Aggrecan [35]

With Scaffold

- Hydrogels: biomaterial and synthetic

- Polymeric scaffolds (osteochondral plugs)

- Micro- and nanocarriers

- Fiber/Mesh scaffolds [88]

- GAG/DNA, MMP-13 and hydroxyproline quantification; proliferation in agarose assay by DNA quantification [44]

- GAG/DNA, Collagen II and Young’s/dynamic modulus (Eγ and G) [77]

- Proliferation in alginate beads - Quantification of IL-6, MMP-

13, Collagen II and Aggrecan [85]

Page 32: Thesis - archive-ouverte.unige.ch

OA in vitro models  

25

Explants

Easy to obtain and inexpensive. Allows for studies of intercellular and cell to ECM relationship because it maintains tissue as a whole [93]

High variability and limited amounts of replicates from source. Cell death at edge of extracted tissues [12]

- Articular cartilage and synovial membrane (human and bovine)

- Osteochondral plugs (human)

- Femoral chondyles (human, murine and equine)

- Cytotoxicity and cartilage penetration assays

- Quantification of IL-6 [9,57]

4.1. 2D cellular models

Two-dimensional cellular models can be described as monolayer culture (4.1.1.),

when a single cell line is cultured or co-culture (4.1.2.) when two or more cell lines

are cultured together, in a monolayer.

4.1.1. Monolayer culture

Culturing cells in monolayer is a well-established, cost-effective method to obtain

relatively fast, reliable, and high-throughput results. As OA in vitro models, these

can be immortal lines or harvested primary cells cultured adherent to plastic flat

surfaces. Their source can vary from murine, bovine to human. Table 2 lists the use

of the most common cell lines: RAW 264.7 macrophages, human primary

synoviocytes and human articular chondrocytes. To evaluate IA delivery systems,

2D models with cells like synoviocytes or chondrocytes that respond to cytokine

stimulation (typically IL-1β, mimicking the inflammatory catabolic environment of

OA) are thus ideal for screening of either anti-inflammatory or chondroprotective

molecules from DDSs by quantification of several inflammatory and cartilage

degradation markers: IL-6, TNF-α, PGE2, COX-2, NO, iNOS, MMP-1, MMP-3,

MMP-13, and ADAMTS-5. These monolayer models (especially human cell lines)

are also useful and largely explored for cytotoxicity and proliferation testing in local

administration cases since they correspond to the direct cellular target

[40,53,62,64,84]. Additionally, different signaling pathways can be explored from

these models, such as the inflammatory NF-κB pathway in human chondrocytes

[94]. Human bone marrow mesenchymal stem cells (HBMSCs) have the ability to

de-differentiate to mature articular chondrocytes; thus, these are also used to

quantify chondrogenesis through sulfated glycosaminoglycans (GAG), abundant in

Page 33: Thesis - archive-ouverte.unige.ch

Chapter I  

26

articular cartilage ECM content, gene expression of collagen II and aggrecan, in

addition to the cytotoxicity and screening of molecules [35,73]. However, problems

arise from culturing primary articular chondrocytes, as the actual cartilage tissue

would require a three-dimensional cell growth, interacting with the ECM, in contrast

to a flat surface (Table 2). Therefore, after a small number of passages, which limits

the number of experiments and length of studies, de-differentiation tends to occur

as cells change in phenotype and morphology from an orthogonal shape to an

elongated shape, resembling fibroblast-like chondrocytes. This phenotype is known

to produce collagen type I fibers instead of the collagen II fibers consistent with

articular cartilage, an issue when using this type of monolayer model to assess

cartilage growth from collagen II and aggrecan quantification. This lack of tissue-

mimicking properties prevents 2D in vitro models from accurately mimicking

intercellular and cell-to-ECM relationships. Not only this, but weight-bearing and

mechanical-loading experiments, crucial in the understanding of OA as a pathology,

are not easily explored using these models [12].

4.1.2. Co-culture

Monolayer culture of different joint cell lines is an alternative to improve intercellular

relationship studies. Differently from monolayers of a single cell line, in co-culture

where chondrocytes are incubated together with synoviocytes and stimulated by

pro-inflammatory cytokines, cross-talk between cells happens through intercellular

calcium and paracrine signaling, maintaining homeostasis of articular chondrocytes.

Evaluation of effects of anti-inflammatory or chondroprotective molecules in articular

cartilage is then higher in accuracy by the co-culturing of both cell types due to the

preservation of these intercellular signaling pathways [89]. Chondrocytes incubated

with osteoblasts help maintain cellular physiology and phenotype through paracrine

signaling. This is an useful model in investigating the effects of chondroprotection

(slowing of cartilage degradation) in bone remodeling [95]. Mesenchymal stem cells

are interesting in co-culture as pluripotency leads to specific de-differentiation,

allowing for different cellular pathways to be analyzed together with articular

chondrocytes from cellular secreted markers [96]. However, despite advantages

gained by culturing different types of cells together in terms of tissue-like

maintenance of homeostasis and phenotypes, this in vitro model is subject to some

of the same drawbacks as monolayer culture, notably, culturing in a flat surface and

Page 34: Thesis - archive-ouverte.unige.ch

OA in vitro models  

27

lack of growth structure. In addition, maintaining different cellular environments at

the same time is expensive.

4.2. 3D cellular models

Three dimensional cellular models can be classified into models without scaffold

(4.2.1.), where cells are grown in pellets and models with scaffold (4.2.2.) where

cellular growth happens in an external platform (biologic or synthetic polymer).

4.2.1. 3D cellular models without scaffold

Three-dimensional cellular pellets circumvent some of the disadvantages of

monolayer cultures, especially as they allow a structure, maintaining cellular growth

in all dimensions and synthesis of articular cartilage ECM. In this approach,

chondrocytes can be centrifuged together in conical bottom wells or tubes or

cultured under stirring using bioreactors. Inducing cell clustering forms cartilage

tissue-like pellets, after a specific incubation time, with sizes up to 5 mm [97,98].

These pellets can mimic articular tissue as a whole, providing insights into cell-to-

cell and cell-to-ECM relationships. Like in a monolayer culture, HBMSCs pellets can

replace 3D chondrocyte pellets. As an in vitro model for IA DDSs development, 3D

pellets have been applied in the evaluation of chondrogenesis and

chondroprotective effects after IL-1β stimulation by GAG content quantification and

gene expression of collagen II, aggrecan, and MMPs [35]. A primary reason as to

why pellets are not a standard in vitro OA model is linked to difficulties in maintaining

3D cellular cultures in terms of cost and quantity. 3D articular dedifferentiated

chondrocytes are not associated with high proliferation rates and derive from low

monolayer passages restricting cellular amounts. Culture media is supplemented

with a high amount of growth factors and chondrogenic stabilizers, representing

higher costs compared to monolayer culture [99]. Additionally, pellets have short

viability spans, where nutrients have difficulties in penetrating the pellet, inducing

cell death at its core. As a model for IA DDSs, interaction of formulations with the

tissue as a whole is essential in characterizing target specificity. The inability to fully

penetrate the pellets poses a limitation to the use of this model in the IA setting [92].

Bypassing these shortcomings is, however, made possible by establishing this type

of 3D cellular growth in external structures - scaffolds.

Page 35: Thesis - archive-ouverte.unige.ch

Chapter I  

28

4.2.2. 3D cellular models with scaffold

Cells can be cultured directly into external scaffolds, gaining three-dimensional

features. As an in vitro model for IA DDSs development, this alternative has great

potential for targeted delivery. Not only does it provide structural support for 3D

cellular growth by mimicking features of joint structure, making it a good model of

loading and weight-bearing in OA, as the nature of the scaffold (biologic or synthetic)

can play a role in cellular growth and maintenance. The most commonly used

scaffolds are hydrogels due to their high water content and the extensive ability to

tailor their mechanical and physicochemical properties. Biopolymers like agarose,

chitosan, alginate and hyaluronic acid have been applied to grow chondrocytes,

mimicking articular cartilage, and osteoblasts, aiming to model the osteochondral

plate. Combining the growth of both these types of cells has also been explored,

forming bilayer scaffolds, in an attempt to represent the whole articular joint [100].

As such, and after cytokine stimulation and exposure to therapeutic molecules,

different cartilage markers can be assessed by different assays: GAG content

(alcian blue assay), collagen II, aggrecan, MMPs (gene expression analysis) and

even pro-inflammatory cytokines (enzyme-linked immunosorbent assay ELISA)

[44,85]. Rheological measurements (elastic Young’s and G moduli) help investigate

the mechanical properties of chondrocytes in hydrogels (agarose) [77]. Synthetic

hydrogels and polymers can be applied as scaffolds, with advantages like

mechanical features and support. 3D printing has been applied in this field with

promising results in cartilage regeneration [101,102]. Compared to 2D models,

scaffold-based 3D culture is expensive, difficult to maintain and hard to standardize,

given the many options for scaffolds. Depending on their nature, problems may arise

with how these influence results. For example, biopolymer-based hydrogels may

themselves have a chondroprotective effect on cultured chondrocytes, skewing

effects of tested drugs. The nature of the scaffold may also translate into differences

between in vitro and in vivo models. For instance, hydrogels are rich in water, unlike

subchondral bones of joints; thus, the growth of osteoblasts in such scaffolds is not

an accurate representation of in vivo conditions [88].

4.3. Explants

Explants could be considered the most accurate in vitro model of OA as the whole

tissue is maintained in its form and function. Just like tissue where cells are

Page 36: Thesis - archive-ouverte.unige.ch

OA in vitro models  

29

harvested from, their source can be both animal and human. Explants of both

cartilage and synovial membrane are useful to investigate anti-

inflammatory/chondroprotective effects of DDSs or molecules. Bovine cartilage is

also commonly harvested to test the permeation and distribution of a drug or DDSs

into the cartilage and/or subchondral bone, using fluorescent-dye-labeled-

nanoparticles drug molecules. Femoral heads are attractive in loading and weight-

bearing studies, whereas osteochondral plugs are used to investigate the balance

between cartilage and bone regeneration when exposed to chondroprotective

drugs. By measuring DNA and cellular turnover, cell viability and proliferation can

also be assessed using explants after exposure to DDSs and/or free drug molecules

[9,57]. Despite clear advantages (Table 2) from using explants where intercellular

and cell-to-ECM relationships are preserved; extraction induces cell death on the

outer layers of the tissue, compromising the model. Accurate induction of OA may

pose another limitation, as often harvested tissues are healthy specimens and not

pathological as the ones collected in other cellular models (monolayer, for example).

Additionally, the maintenance of tissues in culture can be expensive and difficult to

control, with explants lasting up to 10 days. Conditions such as temperature, pH,

humidity, culture medium and supplements like insulin plus light exposure are crucial

in maintaining the viability of explants. Another substantial limitation of this model is

that viable replicates are very difficult to achieve, as tissue sources are finite and

not abundant [87].

4.4. Considerations on OA in vitro models for development of IA DDSs

Understanding the advantages and disadvantages of the different types of OA in

vitro models is crucial when developing intra-articular drug delivery systems. The

choice of in vitro model is influenced by how effects of the delivered drug can be

assessed, be it anti-inflammatory by quantification of released cytokines or

chondroprotection by evaluating GAG content and collagen II mRNA expression.

Different cell lines such as macrophages, synoviocytes or chondrocytes secrete

different factors and/or respond differently to cytokine stimulation. When evaluating

the anti-inflammatory effects of therapeutic molecules in DDSs, macrophages and

synoviocytes represent the most accurate cellular model. For chondroprotective

effects and/or subchondral bone protection, it is important to test these in accurate

representations of articular cartilage and subchondral bone. For this, 3D

Page 37: Thesis - archive-ouverte.unige.ch

Chapter I  

30

chondrocyte models or bilayer scaffolds for osteochondral defects are adequate

models. Furthermore, articular cartilage or subchondral bone cells/tissue do not

participate in inflammatory cascades directly, making these cellular models specific

for measuring cartilage degradation markers. When developing, for example, an IA

DDSs eluting a drug that has the synovium as a specific target, it is important to

measure not only off-target activity in the other joint tissues but also the response of

cartilage, for example, to the effects of the drug in the synovium. Monolayer models,

though abundantly investigated and easy to establish, fail in the evaluation of cross-

talk and molecular relationships within the different joint structures, especially

important when developing a local delivery system. However, 2D models are

relatively easy and accurate in assessing cytotoxicity and influence in cell

proliferation of both drug and carriers. In contrast, 3D models allow for a more

accurate and translational representation of joint tissues as phenotype and cellular

growth are preserved. 3D models are essential in evaluating cytokine stimulation,

cell-to-cell and cell-to-ECM relationships and, in the case of scaffold-based 3D

models, loading studies, as these have mechanical properties not found in

monolayer cultures. Nonetheless, their establishment requires highly specific

expertise and can be costly. In addition, the accuracy and reproducibility of the

outcomes, from cytotoxicity assays to gene expression analysis, can be high when

applying 3D models. Explants from specific tissues are good representations of in

vivo joints, as their intact features allow for loading and penetration studies of both

IA carriers and free drug molecules. However, representative experimental

replicates are not easily accessible, and molecular alterations may arise from the

extraction of the tissues. As mentioned previously, time and duration play an

important role in the development and application of in vitro models. OA is a slowly

progressing, chronic disease where molecular changes often only result in actual

physical symptoms very late. As such, tackling the effect over time on tissues is

crucial to understand disease mechanisms and potential therapeutic options.

However, experimentally, it is challenging to maintain cells and tissues viable for

long periods of time. Bioreactors or tissue-mimicking polymers could help

circumvent viability issues, maintaining OA conditions for slightly extended periods

[103].

Formulation aspects also influence the choice of in vitro OA model. The formulation

of DDSs (Table 1) implies the use of a carrier for a certain drug molecule. Carriers

Page 38: Thesis - archive-ouverte.unige.ch

OA in vitro models  

31

have an impact in terms of size and nature. In terms of size, the local administration

of nano-range carriers (nanoparticles) can induce phagocytosis and inflammatory

cascade from synoviocytes in the joint capsule [104]. Therefore, interactions at the

cellular level when testing these DDSs are important to consider if

macrophages/synoviocytes are the chosen in vitro cellular models. As previously

discussed, most cellular OA models are cultured on plastic surfaces, in well plates,

dishes or tubes. Micro-range carriers (microparticles or larger liposomes) are prone

to sedimentation in these cell culture set-tings, especially polymeric carriers, which

display high density when in a culture medium suspension. This sedimentation may

negatively affect experimental result, by uneven drug molecule distribution,

heterogeneous presentation to test cells and lower contact surfaces between the

carrier–drug complex and cells [105]. This issue can be bypassed by performing

experiments in orbital shakers. However, as described for 3D cellular models,

altering centrifuge force and balance induces changes in cellular growth and

phenotype [100]. When evaluating hydrogels (Table 1), either in monolayer or 3D

cellular models, even when using explants, it is important to consider the nature of

the polymer (synthetic or bio) and the viscosity of the gel. Like for nano-

/microcarriers, choice of polymer will have an impact on cellular response. Thus,

biocompatibility and innocuousness of polymers are important characteristics,

particularly when testing inflammation and anti-inflammatory effects, as further

induction of inflammatory cascades is undesired. The majority of hydrogels being

explored for OA treatment are HA-based, a natural component of articular cartilage

[10,69,70,72,73,75,78]. As such, it is important to assess their impact as stand-alone

carrier vs. carrier with drug, as it is expected that this type of gel will have an

influence on chondrocyte growth by inducing chondroprotection through CD44

receptor interaction. Lastly, rheological properties of hydrogels need to be

considered when applying in vitro cellular models. High viscosity may induce

occlusion effects in either cultured cells or tissues, generating hypoxia phenomena

and thus lowering viability scores [106]. Consideration of all different formulation

aspects does not exclude testing of drug-alone controls in these cellular OA models,

as these dictate why and how DDSs are better alternatives in IA administration.

Page 39: Thesis - archive-ouverte.unige.ch

Chapter I  

32

5. CONCLUSIONS AND FUTURE PERSPECTIVES

Improved design and development of efficacious IA DDSs relies on the use of

accurate, predictive in vitro and in vivo models. However, to date, there is no OA

gold standard in vitro model and few guidelines or models adapted specifically to IA

DDSs formulations. Presently, monolayer models, despite being easy to establish

and ideal for rapid screening of molecules, fail in representing ac-curate OA

conditions, such as cross-talk between different tissues. This could be bypassed by

co-culture of two types of cell lines, like synoviocytes and chondrocytes, but aspects

like cell de-differentiation and ECM growth are not negligible. Three-dimensional

models are considered better representations of in vivo OA, as in these models,

three-dimensional structures of tissues and cellular phenotype and growth are

preserved. However, with or without scaffold, 3D models are difficult to establish

and maintain, and outcomes vary greatly according to the source and nature of

scaffold. For studies in articular tissues, explants are considered best in correlation

to in vivo OA conditions. However, viable replicates and maintenance of tissues in

in vitro environments are important limitations. Recently, a bioengineering approach

combining 3D cell culture and microfluidics—organ-on-chip (OoC)—has been in the

field of OA. Cartilage-on-chip and osteochondral-tissue-on-chip have been

developed to perfectly mimic joint microenvironments, allowing for better

reproductions of in vivo conditions. Promising results have been described testing

the drug alone, making this a promising approach for the better development of IA

DDSs in the future [107–109]. In this context, considerations (Table 2) have to be

taken into account when designing and developing IA DDSs, especially when

deciding outcome readouts. To this extent, the type of formulation and mode of

action of drug molecules (Table 1) play a critical role. Monolayer models are better

suited for testing anti-inflammatory activity, whereas 3D chondrocyte models are

preferred to evaluate chondroprotection activities. When testing hydrogels, it is

important to assess the nature of the scaffold in 3D models and even occlusion in

explants. In the future, research advancements should focus on improving the

design and development of OA in vitro models for better prediction of in vivo and,

eventually, clinical results. This should be done while always considering the

tailoring of in vitro models to specific IA DDSs formulations, like maintaining cellular

viability conditions for testing of sustained prolonged drug release delivery systems.

Page 40: Thesis - archive-ouverte.unige.ch

OA in vitro models  

33

6. REFERENCES

[1] Hunter, D.J.; Schofield, D.; Callander, E. The individual and socioeconomic impact of osteoarthritis. Nat. Rev. Rheumatol. 2014, 10, 437–441, doi:10.1038/nrrheum.2014.44.

[2] Goldring, M.B.; Goldring, S.R. Osteoarthritis. J. Cell. Physiol. 2007, 213, 626–634, doi:10.1002/jcp.21258.

[3] Koszowska, A.; Hawranek, R.; Nowak, J. Osteoarthritis -a multifactorial issue. Reumatologia 2014, 52, 319–325, doi:10.5114/reum.2014.46670.

[4] Ratneswaran, A.; Rockel, J.S.; Kapoor, M. Understanding osteoarthritis pathogenesis: A multiomics system-based approach. Curr. Opin. Rheumatol. 2020, 32, 80–91, doi:10.1097/bor.0000000000000680.

[5] Abramoff, B.; Caldera, F.E. Osteoarthritis: Pathology, Diagnosis, and Treatment Options. Med. Clin. N. Am. 2020, 104, 293–211, doi:10.1016/j.mcna.2019.10.007.

[6] Kolasinski, S.L.; Neogi, T.; Hochberg, M.C.; Oatis, C.; Guyatt, G.; Block, J.; Callahan, L.; Copenhaver, C.; Dodge, C.; Felson, D.; et al. 2019 American College of Rheumatology/Arthritis Foundation Guideline for the Management of Osteoarthritis of the Hand, Hip, and Knee. Arthritis Rheumatol. 2020, 72, 220–233, doi:10.1002/art.41142.

[7] Nelson, A.E.; Allen, K.D.; Golightly, Y.M.; Goode, A.P.; Jordan, J.M. A systematic review of recommendations and guidelines for the management of osteoarthritis: The Chronic Osteoarthritis Management Initiative of the U.S. Bone and Joint Initiative. Semin. Arthritis Rheum. 2014, 43, 701–712, doi:10.1016/j.semarthrit.2013.11.012.

[8] Gómez-Gaete, C.; Retamal, M.; Chávez, C.; Bustos, P.; Godoy, R.; Torres-Vergara, P. Development, characterization and in vitro evaluation of biodegradable rhein-loaded microparticles for treatment of osteoarthritis. Eur. J. Pharm. Sci. 2017, 96, 390–397, doi:10.1016/j.ejps.2016.10.010.

[9] Lin, J.B.; Poh, S.; Panitch, A. Controlled release of anti-inflammatory peptides from reducible thermosensitive nanoparticles suppresses cartilage inflammation. Nanomed. Nanotechnol. Biol. Med. 2016, 12, 2095–2100, doi:10.1016/j.nano.2016.05.010.

[10] Maudens, P.; Meyer, S.; Seemayer, C.A.; Jordan, O.; Allémann, E. Self-assembled thermoresponsive nanostructures of hyaluronic acid conjugates for osteoarthritis therapy. Nanoscale 2018, 10, 1845, doi:10.1039/c7nr07614b.

[11] Dong, J.; Jiang, D.; Wang, Z.; Wu, G.; Miao, L.; Huang, L. Intra-articular delivery of liposomal celecoxib–hyaluronate combination for the treatment of osteoarthritis in rabbit model. Int. J. Pharm. 2013, 441, 285–290, doi:10.1016/j.ijpharm.2012.11.031.

[12] Johnson, C.I.; Argyle, D.J.; Clements, D.N. In vitro models for the study of osteoarthritis. Vet. J. 2016, 209, 40–49, doi:10.1016/j.tvjl.2015.07.011.

[13] Hunter, D.J.; Bierma-Zeinstra, S. Osteoarthritis. Lancet 2019, 393, 1745–1759, doi:10.1016/S0140-6736(19)30417-9.

[14] Kloppenburg, M.; Berenbaum, F. Osteoarthritis year in review 2019: Epidemiology and therapy. Osteoarthr. Cartil. 2020, 28, 242–248, doi:10.1016/j.joca.2020.01.002.

[15] Goldring, M.B.; Goldring, S.R. Articular cartilage and subchondral bone in the pathogenesis of osteoarthritis. Ann. N. Y. Acad. Sci. 2010, 1192, 230–237.

[16] Loeser, R.F.; Goldring, S.R.; Scanzello, C.R.; Goldring, M.B. Osteoarthritis: A disease of the joint as an organ. Arthritis Rheum. 2012, 64, 1697–1707, doi:10.1002/art.34453.

[17] Chen, D.; Shen, J.; Zhao, W.; Wang, T.; Han, L.; Hamilton, J.L.; Im, H.J. Osteoarthritis: Toward a comprehensive understanding of pathological mechanism. Bone Res. 2017, 5, 16044, doi:10.1038/boneres.2016.44.

[18] Mora, J.C.; Przkora, R.; Cruz-Almeida, Y. Knee osteoarthritis: Pathophysiology and current treatment modalities. J. Pain Res. 2018, 11, 2189–2196, doi:10.2147/JPR.S154002.

[19] Emery, C.A.; Whittaker, J.L.; Mahmoudian, A.; Lohmander, L.S.; Roos, E.M.; Bennell, K.L.; Toomey, C.M.; Reimer, R.A.; Thompson, D.; Ronsky, J.L.; et al. Establishing outcome measures in early knee osteoarthritis. Nat. Rev. Rheumatol. 2019, 15, 438–448, doi:10.1038/s41584-019-0237-3.

Page 41: Thesis - archive-ouverte.unige.ch

Chapter I  

34

[20] Brakke, R.; Singh, J.; Sullivan, W. Physical Therapy in Persons With Osteoarthritis. PM R 2012, 4, S53–S58, doi:10.1016/j.pmrj.2012.02.017.

[21] Allen, K.D.; Choong, P.F.; Davis, A.M.; Dowsey, M.M.; Dziedzic, K.S.; Emery, C.; Hunter, D.J.; Losina, E.; Page, A.E.; Roos, E.M.; et al. Osteoarthritis: Models for appropriate care across the disease continuum. Best Pract. Res. Clin. Rheumatol. 2016, 30, 503–535, doi:10.1016/j.berh.2016.09.003.

[22] Smith, S.R.; Deshpande, B.R.; Collins, J.E.; Katz, J.N.; Losina, E. Comparative pain reduction of oral non-steroidal anti-inflammatory drugs and opioids for knee osteoarthritis: Systematic analytic review. Osteoarthr. Cartil. 2016, 24, 962–972, doi:10.1016/j.joca.2016.01.135.

[23] Webb, D.; Naidoo, P. Viscosupplementation for knee osteoarthritis: A focus on hylan G-F 20. Orthop. Res. Rev. 2018, 10, 73–81, doi:10.2147/ORR.S174649.

[24] Southworth, T.M.; Naveen, N.B.; Tauro, T.M.; Leong, N.L.; Cole, B.J. The Use of Platelet-Rich Plasma in Symptomatic Knee Osteoarthritis. J. Knee Surg. 2019, 32, 37–45, doi:10.1055/s-0038-1675170.

[25] Geiger, B.C.; Grodzinsky, A.J.; Hammond, P. Designing drug delivery systems for articular joints. Chem. Eng. Prog. 2018, 114, 46–51.

[26] Kou, L.; Xiao, S.; Sun, R.; Bao, S.; Yao, Q.; Chen, R. Drug Delivery Biomaterial-engineered intra-articular drug delivery systems for osteoarthritis therapy Biomaterial-engineered intra-articular drug delivery systems for osteoarthritis therapy. Drug Deliv. 2019, doi:10.1080/10717544.2019.1660434.

[27] Kang, M.L.; Ko, J.Y.; Kim, J.E.; Im, G. Il Intra-articular delivery of kartogenin-conjugated chitosan nano/microparticles for cartilage regeneration. Biomaterials 2014, 35, 9984–9994, doi:10.1016/j.biomaterials.2014.08.042.

[28] Maudens, P.; Jordan, O.; Allémann, E. Recent advances in intra-articular drug delivery systems for osteoarthritis therapy. Drug Discov. Today 2018, 23, 1761–1775, doi:10.1016/j.drudis.2018.05.023.

[29] Rai, M.F.; Pham, C.T. Intra-articular drug delivery systems for joint diseases. Curr. Opin. Pharmacol. 2018, 40, 67–73, doi:10.1016/j.coph.2018.03.013.

[30] Brown, S.; Kumar, S.; Sharma, B. Intra-articular targeting of nanomaterials for the treatment of osteoarthritis. Acta Biomater. 2019, 93, 239–257, doi:10.1016/j.actbio.2019.03.010.

[31] Mancipe Castro, L.M.; García, A.J.; Guldberg, R.E. Biomaterial strategies for improved intra‐articular drug delivery. J. Biomed. Mater. Res. Part. A 2020, doi:10.1002/jbm.a.37074.

[32] Janssen, M.; Timur, U.T.; Woike, N.; Welting, T.J.M.; Draaisma, G.; Gijbels, M.; van Rhijn, L.W.; Mihov, G.; Thies, J.; Emans, P.J. Celecoxib-loaded PEA microspheres as an auto regulatory drug-delivery system after intra-articular injection. J. Control. Release 2016, 244, 30–40, doi:10.1016/j.jconrel.2016.11.003.

[33] Liu, P.; Gu, L.; Ren, L.; Chen, J.; Li, T.; Wang, X.; Yang, J.; Chen, C.; Sun, L. Intra-articular injection of etoricoxib-loaded PLGA-PEG-PLGA triblock copolymeric nanoparticles attenuates osteoarthritis progression. Am. J. Transl. Res. 2019, 11, 6775–6789.

[34] Petit, A.; Redout, E.M.; van de Lest, C.H.; de Grauw, J.C.; Müller, B.; Meyboom, R.; van Midwoud, P.; Vermonden, T.; Hennink, W.E.; René van Weeren, P. Sustained intra-articular release of celecoxib from in situ forming gels made of acetyl-capped PCLA-PEG-PCLA triblock copolymers in horses. Biomaterials 2015, 53, 426–436, doi:10.1016/j.biomaterials.2015.02.109.

[35] Kang, M.-L.; Kim, J.-E.; Im, G.-I. Thermoresponsive nanospheres with independent dual drug release profiles for the treatment of osteoarthritis. Acta Biomater. 2016, 39, 65–78, doi:10.1016/j.actbio.2016.05.005.

[36] Steven, R.; Goldring, M.B.G. Kelley’s Textbook of Rheumatology; Firestein, G.S., Budd, R.C., Gabriel, S.E., McInnes, I.B., O.J., Eds.; Saunders: Philadelphia, US, 2013; pp. 1–19, ISBN 9781455737673.

[37] Marieb, E.N.; Wilhelm, P.B.; Mallatt, J. Human Anatomy; Person: London, UK, 2017; ISBN 9780134243818.

Page 42: Thesis - archive-ouverte.unige.ch

OA in vitro models  

35

[38] Piluso, S.; Li, Y.; Abinzano, F.; Levato, R.; Teixeira, L.M.; Karperien, M.; Leijten, J.; Van Weeren, R.; Malda, J. Mimicking the Articular Joint with In Vitro Models. Trends Biotechnol. 2019, 37, 1063–1077, doi:10.1016/j.tibtech.2019.03.003.

[39] Tellier, L.E.; Treviño, E.A.; Brimeyer, A.L.; Reece, D.S.; Willett, N.J.; Guldberg, R.E.; Temenoff, J.S. Intra-articular TSG-6 delivery from heparin-based microparticles reduces cartilage damage in a rat model of osteoarthritis †. Biomater. Sci. 2018, 6, 1159, doi:10.1039/c8bm00010g.

[40] Pradal, J.; Zuluaga, M.-F.; Maudens, P.; Waldburger, J.-M.; Seemayer, C.A.; Doelker, E.; Gabay, C.; Jordan, O.; Allémann, E. Intra-articular bioactivity of a p38 MAPK inhibitor and development of an extended-release system. Eur. J. Pharm. Biopharm. 2015, 93, 110–117, doi:10.1016/j.ejpb.2015.03.017.

[41] Goldring, M.B. The role of the chondrocyte in osteoarthritis. Arthritis Rheum. 2000, 43, 1916–1926, doi:10.1002/1529-0131(200009)43:9<1916::aid-anr2>3.0.co;2-i.

[42] Goldring, S.R.; Goldring, M.B. Changes in the osteochondral unit during osteoarthritis: Structure, function and cartilage–bone crosstalk. Nat. Rev. Rheumatol. 2016, 12, 632–644, doi:10.1038/nrrheum.2016.148.

[43] Maudens, P.; Seemayer, C.A.; Thauvin, C.; Gabay, C.; Jordan, O.; Allémann, E. Nanocrystal-Polymer Particles: Extended Delivery Carriers for Osteoarthritis Treatment. Small 2018, 14, 1703108, doi:10.1002/smll.201703108.

[44] Aydin, O.; Korkusuz, F.; Korkusuz, P.; Tezcaner, A.; Bilgic, E.; Yaprakci, V.; Keskin, D. In vitro and in vivo evaluation of doxycycline-chondroitin sulfate/PCLmicrospheres for intraarticular treatment of osteoarthritis. J. Biomed. Mater. Res. Part B Appl. Biomater. 2015, 103, 1238–1248, doi:10.1002/jbm.b.33303.

[45] Arunkumar, P.; Indulekha, S.; Vijayalakshmi, S.; Srivastava, R. Synthesis, characterizations, in vitro and in vivo evaluation of Etoricoxib-loaded Poly (Caprolactone) microparticles-a potential Intra-articular drug delivery system for the treatment of Osteoarthritis. J. Biomater. Sci. Polym. Ed. 2016, 27, 303–316, doi:10.1080/09205063.2015.1125564.

[46] Abd-Allah, H.; Kamel, A.O.; Sammour, O.A. Injectable long acting chitosan/tripolyphosphate microspheres for the intra-articular delivery of lornoxicam: Optimization and in vivo evaluation. Carbohydr. Polym. 2016, 149, 263–273, doi:10.1016/j.carbpol.2016.04.096.

[47] Goto, N.; Okazaki, K.; Akasaki, Y.; Ishihara, K.; Murakami, K.; Koyano, K.; Ayukawa, Y.; Yasunami, N.; Masuzaki, T.; Nakashima, Y. Single intra-articular injection of fluvastatin-PLGA microspheres reduces cartilage degradation in rabbits with experimental osteoarthritis. J. Orthop. Res. 2017, 35, 2465–2475, doi:10.1002/jor.23562.

[48] Maudens, P.; Seemayer, C.A.; Pfefferlé, F.; Jordan, O.; Allémann, E. Nanocrystals of a potent p38 MAPK inhibitor embedded in microparticles: Therapeutic effects in inflammatory and mechanistic murine models of osteoarthritis. J. Control. Release 2018, 276, 102–112, doi:10.1016/j.jconrel.2018.03.007.

[49] Kumar, A.; Bendele, A.M.; Blanks, R.C.; Bodick, N. Sustained efficacy of a single intra-articular dose of FX006 in a rat model of repeated localized knee arthritis. Osteoarthr. Cartil. 2015, 23, 151–160, doi:10.1016/j.joca.2014.09.019.

[50] Kraus, V.B.; Conaghan, P.G.; Aazami, H.A.; Mehra, P.; Kivitz, A.J.; Lufkin, J.; Hauben, J.; Johnson, J.R.; Bodick, N. Synovial and systemic pharmacokinetics (PK) of triamcinolone acetonide (TA) following intra-articular (IA) injection of an extended-release microsphere-based formulation (FX006) or standard crystalline suspension in patients with knee osteoarthritis (OA). Osteoarthr. Cartil. 2018, 26, 34–42, doi:10.1016/j.joca.2017.10.003.

[51] Getgood, A.; Dhollander, A.; Malone, A.; Price, J.; Helliwell, J. Pharmacokinetic Profile of Intra-articular Fluticasone Propionate Microparticles in Beagle Dog Knees. Cartilage 2019, 10, 139–147, doi:10.1177/1947603517723687.

[52] Salgado, C.; Guénée, L.; Černý, R.; Allémann, E.; Jordan, O. Nano wet milled celecoxib extended release microparticles for local management of chronic inflammation. Int. J. Pharm. 2020, 589, doi:10.1016/j.ijpharm.2020.119783.

Page 43: Thesis - archive-ouverte.unige.ch

Chapter I  

36

[53] Dhanabalan, K.M.; Gupta, V.K.; Agarwal, R. Rapamycin-PLGA microparticles prevent senescence, sustain cartilage matrix production under stress and exhibit prolonged retention in mouse joints. Biomater. Sci. 2020, 8, 4308–4321, doi:10.1039/d0bm00596g.

[54] Bajpayee, A.G.; Quadir, M.A.; Hammond, P.T.; Grodzinsky, A.J. Charge based intra-cartilage delivery of single dose dexamethasone using Avidin nano-carriers suppresses cytokine-induced catabolism long term. Osteoarthr. Cartil. 2016, 24, 71–81, doi:10.1016/j.joca.2015.07.010.

[55] Niazvand, F.; Khorsandi, L.; Abbaspour, M.; Orazizadeh, M.; Varaa, N.; Maghzi, M.; Ahmadi, K. Curcumin-loaded poly lactic-co-glycolic acid nanoparticles effects on mono-iodoacetate-induced osteoarthritis in rats. Vet. Res. Forum 2017, 8, 155–161.

[56] Bajpayee, A.G.; De La Vega, R.E.; Scheu, M.; Varady, N.H.; Yannatos, I.A.; Brown, L.A.; Krishnan, Y.; Fitzsimons, T.J.; Bhattacharya, P.; Frank, E.H.; et al. Sustained intra-cartilage delivery of low dose dexamethasone using a cationic carrier for treatment of posttraumatic osteoarthritis. Eur. Cells Mater. 2017, 34, 341–364, doi:10.22203/eCM.v034a21.

[57] Mcmasters, J.; Poh, S.; Lin, J.B.; Panitch, A. Delivery of Anti-inflammatory Peptides from Hollow PEGylated Poly(NIPAM) Nanoparticles Reduces Inflammation in an Ex Vivo Osteoarthritis Model Graphical abstract HHS Public Access. J. Control. Release 2017, 258, 161–170, doi:10.1016/j.jconrel.2017.05.008.

[58] Hu, Q.; Chen, Q.; Yan, X.; Ding, B.; Chen, D.; Cheng, L. Chondrocyte affinity peptide modified PAMAM conjugate as a nanoplatform for targeting and retention in cartilage. Nanomedicine 2018, 13, 749–767, doi:10.2217/nnm-2017-0335.

[59] Fan, W.; Li, J.; Yuan, L.; Chen, J.; Wang, Z.; Wang, Y.; Guo, C.; Mo, X.; Yan, Z. Intra-articular injection of kartogenin-conjugated polyurethane nanoparticles attenuates the progression of osteoarthritis. Drug Deliv. 2018, 25, 1004–1012, doi:10.1080/10717544.2018.1461279.

[60] Liu, X.; Corciulo, C.; Arabagian, S.; Ulman, A.; Cronstein, B.N. Adenosine-Functionalized Biodegradable PLA-b-PEG Nanoparticles Ameliorate Osteoarthritis in Rats. Sci. Rep. 2019, 9, doi:10.1038/s41598-019-43834-y.

[61] Mota, A.H.; Direito, R.; Carrasco, M.P.; Rijo, P.; Ascensão, L.; Viana, A.S.; Rocha, J.; Eduardo-Figueira, M.; Rodrigues, M.J.; Custódio, L.; et al. Combination of hyaluronic acid and PLGA particles as hybrid systems for viscosupplementation in osteoarthritis. Int. J. Pharm. 2019, 559, 13–22, doi:10.1016/j.ijpharm.2019.01.017.

[62] Zerrillo, L.; Que, I.; Vepris, O.; Morgado, L.N.; Chan, A.; Bierau, K.; Li, Y.; Galli, F.; Bos, E.; Censi, R.; et al. pH-responsive poly(lactide-co-glycolide) nanoparticles containing near-infrared dye for visualization and hyaluronic acid for treatment of osteoarthritis. J. Control. Release 2019, 309, 265–276, doi:10.1016/j.jconrel.2019.07.031.

[63] Jin, T.; Wu, D.; Liu, X.M.; Xu, J.T.; Ma, B.J.; Ji, Y.; Jin, Y.Y.; Wu, S.Y.; Wu, T.; Ma, K. Intra-articular delivery of celastrol by hollow mesoporous silica nanoparticles for pH-sensitive anti-inflammatory therapy against knee osteoarthritis. J. Nanobiotechnol. 2020, 18, doi:10.1186/s12951-020-00651-0.

[64] Jung, J.H.; Kim, S.E.; Kim, H.J.; Park, K.; Song, G.G.; Choi, S.J. A comparative pilot study of oral diacerein and locally treated diacerein-loaded nanoparticles in a model of osteoarthritis. Int. J. Pharm. 2020, 581, doi:10.1016/j.ijpharm.2020.119249.

[65] Salama, A.H.; Abdelkhalek, A.A.; Elkasabgy, N.A. Etoricoxib-loaded bio-adhesive hybridized polylactic acid-based nanoparticles as an intra-articular injection for the treatment of osteoarthritis. Int. J. Pharm. 2020, 578, doi:10.1016/j.ijpharm.2020.119081.

[66] Deloney, M.; Smart, K.; Christiansen, B.A.; Panitch, A. Thermoresponsive, hollow, degradable core-shell nanoparticles for intra-articular delivery of anti-inflammatory peptide. J. Control. Release 2020, 323, 47–58, doi:10.1016/j.jconrel.2020.04.007.

[67] Alarçin, E.; Demirbağ, Ç.; Karsli-Ceppioglu, S.; Kerimoğlu, O.; Bal-Ozturk, A. Development and characterization of oxaceprol-loaded poly-lactide-co-glycolide nanoparticles for the treatment of osteoarthritis. Drug Dev. Res. 2020, 81, 501–510, doi:10.1002/ddr.21642.

Page 44: Thesis - archive-ouverte.unige.ch

OA in vitro models  

37

[68] She, P.; Bian, S.; Cheng, Y.; Dong, S.; Liu, J.; Liu, W.; Xiao, C. Dextran sulfate-triamcinolone acetonide conjugate nanoparticles for targeted treatment of osteoarthritis. Int. J. Biol. Macromol. 2020, 158, 1082–1089, doi:10.1016/j.ijbiomac.2020.05.013.

[69] Shan-Bin, G.; Yue, T.; Ling-Yan, J. Long-term sustained-released in situ gels of a water-insoluble drug amphotericin B for mycotic arthritis intra-articular administration: Preparation, in vitro and in vivo evaluation. Drug Dev. Ind. Pharm. 2015, 41, 573–582, doi:10.3109/03639045.2014.884129.

[70] Reum Son, A.; Kim, D.Y.; Hun Park, S.; Yong Jang, J.; Kim, K.; Ju Kim, B.; Yun Yin, X.; Ho Kim, J.; Hyun Min, B.; Keun Han, D.; et al. Direct chemotherapeutic dual drug delivery through intra-articular injection for synergistic enhancement of rheumatoid arthritis treatment. Sci. Rep. 2015, 5, 14713, doi:10.1038/srep14713.

[71] Chen, Y.; Liang, X.; Ma, P.; Tao, Y.; Wu, X.; Wu, X.; Chu, X.; Gui, S. Phytantriol-Based In Situ Liquid Crystals with Long-Term Release for Intra-articular Administration. AAPS PharmSciTech 2015, 16, doi:10.1208/s12249-014-0277-6.

[72] Zhang, Z.; Wei, X.; Gao, J.; Zhao, Y.; Zhao, Y.; Guo, L.; Chen, C.; Duan, Z.; Li, P.; Wei, L. Intra-articular injection of cross-linked hyaluronic acid-dexamethasone hydrogel attenuates osteoarthritis: An experimental study in a rat model of osteoarthritis. Int. J. Mol. Sci. 2016, 17, 411, doi:10.3390/ijms17040411.

[73] Kang, M.L.; Jeong, S.Y.; Im, G. Il Hyaluronic Acid Hydrogel Functionalized with Self-Assembled Micelles of Amphiphilic PEGylated Kartogenin for the Treatment of Osteoarthritis. Tissue Eng. Part A 2017, 23, 630–639, doi:10.1089/ten.tea.2016.0524.

[74] Cokelaere, S.M.; Plomp, S.G.M.; de Boef, E.; de Leeuw, M.; Bool, S.; van de Lest, C.H.A.; van Weeren, P.R.; Korthagen, N.M. Sustained intra-articular release of celecoxib in an equine repeated LPS synovitis model. Eur. J. Pharm. Biopharm. 2018, 128, 327–336, doi:10.1016/j.ejpb.2018.05.001.

[75] Hangody, L.; Szody, R.; Lukasik, P.; Zgadzaj, W.; Lénárt, E.; Dokoupilova, E.; Bichovsk, D.; Berta, A.; Vasarhelyi, G.; Ficzere, A.; et al. Intraarticular Injection of a Cross-Linked Sodium Hyaluronate Combined with Triamcinolone Hexacetonide (Cingal) to Provide Symptomatic Relief of Osteoarthritis of the Knee: A Randomized, Double-Blind, Placebo-Controlled Multicenter Clinical Trial. Cartilage 2018, 9, 276–283, doi:10.1177/1947603517703732.

[76] Tanaka, T.; Matsushita, T.; Nishida, K.; Takayama, K.; Nagai, K.; Araki, D.; Matsumoto, T.; Tabata, Y.; Kuroda, R. Attenuation of osteoarthritis progression in mice following intra-articular administration of simvastatin-conjugated gelatin hydrogel. J. Tissue Eng. Regen. Med. 2019, 13, 423–432, doi:10.1002/term.2804.

[77] Stefani, R.M.; Lee, A.J.; Tan, A.R.; Halder, S.S.; Hu, Y.; Guo, X.E.; Stoker, A.M.; Ateshian, G.A.; Marra, K.G.; Cook, J.L.; et al. Sustained low-dose dexamethasone delivery via a PLGA microsphere-embedded agarose implant for enhanced osteochondral repair. Acta Biomater. 2020, 102, 326–340, doi:10.1016/j.actbio.2019.11.052.

[78] Hanafy, A.S.; El-Ganainy, S.O. Thermoresponsive Hyalomer intra-articular hydrogels improve monoiodoacetate-induced osteoarthritis in rats. Int. J. Pharm. 2020, 573, 118859, doi:10.1016/j.ijpharm.2019.118859.

[79] Kawanami, T.; LaBonte, L.R.; Amin, J.; Thibodeaux, S.J.; Lee, C.C.; Argintaru, O.A.; Adams, C.M. A novel diclofenac-hydrogel conjugate system for intraarticular sustained release: Development of 2-pyridylamino-substituted 1-phenylethanol (PAPE) and its derivatives as tunable traceless linkers. Int. J. Pharm. 2020, 585, 119519, doi:10.1016/j.ijpharm.2020.119519.

[80] Tsubosaka, M.; Kihara, S.; Hayashi, S.; Nagata, J.; Kuwahara, T.; Fujita, M.; Kikuchi, K.; Takashima, Y.; Kamenaga, T.; Kuroda, Y.; et al. Gelatin hydrogels with eicosapentaenoic acid can prevent osteoarthritis progression in vivo in a mouse model. J. Orthop. Res. 2020, 38, 2157–2169, doi:10.1002/jor.24688.

Page 45: Thesis - archive-ouverte.unige.ch

Chapter I  

38

[81] Yang, L.; Liu, Y.; Shou, X.; Ni, D.; Kong, T.; Zhao, Y. Bio-inspired lubricant drug delivery particles for the treatment of osteoarthritis. Nanoscale 2020, 12, 17093–17102, doi:10.1039/d0nr04013d.

[82] Mok, S.W.; Fu, S.C.; Cheuk, Y.C.; Chu, I.M.; Chan, K.M.; Qin, L.; Yung, S.H.; Kevin Ho, K.W. Intra-Articular Delivery of Quercetin Using Thermosensitive Hydrogel Attenuate Cartilage Degradation in an Osteoarthritis Rat Model. Cartilage 2020, 11, 490–499, doi:10.1177/1947603518796550.

[83] Sarkar, A.; Carvalho, E.; D’Souza, A.A.; Banerjee, R. Liposome-encapsulated fish oil protein-tagged gold nanoparticles for intra-articular therapy in osteoarthritis. Nanomedicine 2019, 14, 871–887, doi:10.2217/nnm-2018-0221.

[84] Ji, X.; Yan, Y.; Sun, T.; Zhang, Q.; Wang, Y.; Zhang, M.; Zhang, H.; Zhao, X. Glucosamine sulphate-loaded distearoyl phosphocholine liposomes for osteoarthritis treatment: Combination of sustained drug release and improved lubrication. Biomater. Sci. 2019, 7, 2716–2728, doi:10.1039/c9bm00201d.

[85] Chen, C.-H.; Ming Kuo, S.; Tien, Y.-C.; Shen, P.-C.; Kuo, Y.-W.; Hsiang Huang, H. Steady Augmentation of Anti-Osteoarthritic Actions of Rapamycin by Liposome-Encapsulation in Collaboration with Low-Intensity Pulsed Ultrasound. Int. J. Nanomed. 2020, doi:10.2147/IJN.S252223.

[86] Verbost, P.M.; Van Der Valk, J.; Hendriksen, C.F.M. Effects of the introduction of in vitro assays on the use of experimental animals in pharmacological research. ATLA Altern. Lab. Anim. 2007, 35, 223–228, doi:10.1177/026119290703500211.

[87] Cope, P.J.; Ourradi, K.; Li, Y.; Sharif, M. Models of osteoarthritis: The good, the bad and the promising. Osteoarthr. Cartil. 2019, 27, 230–239, doi:10.1016/j.joca.2018.09.016.

[88] Samvelyan, H.J.; Hughes, D.; Stevens, C.; Staines, K.A. Models of Osteoarthritis: Relevance and New Insights. Calcif. Tissue Int. 2020, 1, 3, doi:10.1007/s00223-020-00670-x.

[89] Beekhuizen, M.; Bastiaansen-Jenniskens, Y.M.; Koevoet, W.; Saris, D.B.F.; Dhert, W.J.A.; Creemers, L.B.; Van Osch, G.J.V.M. Osteoarthritic synovial tissue inhibition of proteoglycan production in human osteoarthritic knee cartilage: Establishment and characterization of a long-term cartilage-synovium coculture. Arthritis Rheum. 2011, 63, 1918–1927, doi:10.1002/art.30364.

[90] Spalazzi, J.P.; Dionisio, K.L.; Jiang, J.; Lu, H.H. Osteoblast and Chondrocyte Interactions during Coculture on Scaffolds. IEEE Eng. Med. Biol. Mag. 2003, 22, 27–34, doi:10.1109/MEMB.2003.1256269.

[91] Edmondson, R.; Broglie, J.J.; Adcock, A.F.; Yang, L. Three-Dimensional Cell Culture Systems and Their Applications in Drug Discovery and Cell-Based Biosensors. Assay Drug Dev. Technol. 2014, 12, 207–218, doi:10.1089/adt.2014.573.

[92] Caron, M.M.J.; Emans, P.J.; Coolsen, M.M.E.; Voss, L.; Surtel, D.A.M.; Cremers, A.; van Rhijn, L.W.; Welting, T.J.M. Redifferentiation of dedifferentiated human articular chondrocytes: Comparison of 2D and 3D cultures. Osteoarthr. Cartil. 2012, 20, 1170–1178, doi:10.1016/j.joca.2012.06.016.

[93] Geurts, J.; Jurić, D.; Müller, M.; Schären, S.; Netzer, C. Novel Ex Vivo Human Osteochondral Explant Model of Knee and Spine Osteoarthritis Enables Assessment of Inflammatory and Drug Treatment Responses. Int. J. Mol. Sci. 2018, 19, 1314, doi:10.3390/ijms19051314.

[94] Al-Modawi, R.N.; Brinchmann, J.E.; Karlsen, T.A. Multi-pathway Protective Effects of MicroRNAs on Human Chondrocytes in an In Vitro Model of Osteoarthritis. Mol. Ther. Nucleic Acids 2019, 17, 776–790, doi:10.1016/J.OMTN.2019.07.011.

[95] Thysen, S.; Luyten, F.P.; Lories, R.J.U. Targets, models and challenges in osteoarthritis research. DMM Dis. Model. Mech. 2015, 8, 17–30, doi:10.1242/dmm.016881.

[96] Hendriks, J.; Riesle, J.; van Blitterswijk, C.A. Co-culture in cartilage tissue engineering. J. Tissue Eng. Regen. Med. 2007, 1, 170–178, doi:10.1002/term.19.

Page 46: Thesis - archive-ouverte.unige.ch

OA in vitro models  

39

[97] Ziadlou, R.; Barbero, A.; Stoddart, M.J.; Wirth, M.; Li, Z.; Martin, I.; Wang, X.; Qin, L.; Alini, M.; Grad, S. Regulation of Inflammatory Response in Human Osteoarthritic Chondrocytes by Novel Herbal Small Molecules. Int. J. Mol. Sci. 2019, 20, 5745, doi:10.3390/ijms20225745.

[98] Ziadlou, R.; Barbero, A.; Martin, I.; Wang, X.; Qin, L.; Alini, M.; Grad, S. Anti-Inflammatory and Chondroprotective Effects of Vanillic Acid and Epimedin C in Human Osteoarthritic Chondrocytes. Biomolecules 2020, 10, 932, doi:10.3390/biom10060932.

[99] Barbero, A.; Martin, I. Human articular chondrocytes culture. In Methods in Molecular Medicine; Hauser, H., Fussenegger, M., Eds.; Human Press: Totowa NJ, US, 2007; Volume 140, pp. 237–247.

[100] Erickson, A.E.; Sun, J.; Lan Levengood, S.K.; Swanson, S.; Chang, F.C.; Tsao, C.T.; Zhang, M. Chitosan-based composite bilayer scaffold as an in vitro osteochondral defect regeneration model. Biomed. Microdevices 2019, 21, doi:10.1007/s10544-019-0373-1.

[101] Di Bella, C.; Duchi, S.; O’Connell, C.D.; Blanchard, R.; Augustine, C.; Yue, Z.; Thompson, F.; Richards, C.; Beirne, S.; Onofrillo, C.; et al. In situ handheld three-dimensional bioprinting for cartilage regeneration. J. Tissue Eng. Regen. Med. 2018, 12, 611–621, doi:10.1002/term.2476.

[102] Deng, C.; Zhu, H.; Li, J.; Feng, C.; Yao, Q.; Wang, L.; Chang, J.; Wu, C. Bioactive scaffolds for regeneration of cartilage and subchondral bone interface. Theranostics 2018, 8, 1940–1955, doi:10.7150/thno.23674.

[103] Bicho, D.; Pina, S.; Oliveira, J.M.; Reis, R.L. In vitro mimetic models for the bone-cartilage interface regeneration. In Advances in Experimental Medicine and Biology; Springer New York LLC: New York, NY, US, 2018; Volume 1059, pp. 373–394.

[104] Pradal, J.; Maudens, P.; Gabay, C.; Seemayer, C.A.; Jordan, O.; Allémann, E. Effect of particle size on the biodistribution of nano- and microparticles following intra-articular injection in mice. Int. J. Pharm. 2016, 498, 119–129, doi:10.1016/j.ijpharm.2015.12.015.

[105] Smith, D.; Herman, C.; Razdan, S.; Abedin, M.R.; Van Stoecker, W.; Barua, S. Microparticles for Suspension Culture of Mammalian Cells. ACS Appl. Bio Mater. 2019, doi:10.1021/acsabm.9b00215.

[106] Cai, Z.; Zhang, H.; Wei, Y.; Wu, M.; Fu, A. Shear-thinning hyaluronan-based fluid hydrogels to modulate viscoelastic properties of osteoarthritis synovial fluids. Biomater. Sci. 2019, 7, 3143–3157, doi:10.1039/c9bm00298g.

[107] Lin, Z.; Li, Z.; Li, E.N.; Li, X.; Del Duke, C.J.; Shen, H.; Hao, T.; O’Donnell, B.; Bunnell, B.A.; Goodman, S.B.; et al. Osteochondral Tissue Chip Derived From iPSCs: Modeling OA Pathologies and Testing Drugs. Front. Bioeng. Biotechnol. 2019, 7, 411, doi:10.3389/fbioe.2019.00411.

[108] Occhetta, P.; Mainardi, A.; Votta, E.; Vallmajo-Martin, Q.; Ehrbar, M.; Martin, I.; Barbero, A.; Rasponi, M. Hyperphysiological compression of articular cartilage induces an osteoarthritic phenotype in a cartilage-on-a-chip model. Nat. Biomed. Eng. 2019, 3, 545–557, doi:10.1038/s41551-019-0406-3.

[109] Rosser, J.; Bachmann, B.; Jordan, C.; Ribitsch, I.; Haltmayer, E.; Gueltekin, S.; Junttila, S.; Galik, B.; Gyenesei, A.; Haddadi, B.; et al. Microfluidic nutrient gradient–based three-dimensional chondrocyte culture-on-a-chip as an in vitro equine arthritis model. Mater. Today Bio 2019, 4, 100023, doi:10.1016/j.mtbio.2019.100023.

Page 47: Thesis - archive-ouverte.unige.ch

 

40

Page 48: Thesis - archive-ouverte.unige.ch

 

 

Chapter II  

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Page 49: Thesis - archive-ouverte.unige.ch

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Page 50: Thesis - archive-ouverte.unige.ch

 

43

Chapter II

In vitro anti-inflammatory activity in arthritic synoviocytes of

A. brachypoda root extracts and its unusual dimeric

flavonoids

Carlota Salgado a,b, Hugo Morin a,b, Nayara Coriolano de Aquino c, Laurence Neff a,b, Cláudia Quintino da Rocha d, Wagner Vilegas e, Laurence Marcourt a,b, Jean-Luc Wolfender a,b, Olivier Jordan a,b, Emerson Ferreira Queiroz a,b and Eric Allémann a,b a School of Pharmaceutical Sciences, University of Geneva, 1 Rue Michel-Servet 1211 Geneva, Switzerland b Institute of Pharmaceutical Sciences of Western Switzerland, University of Geneva, 1 Rue Michel-Servet 1211 Geneva, Switzerland c Departamento de Química Orgânica e Inorgânica, Universidade Federal do Ceará, 60450-765 Fortaleza-CE, Brazil d Laboratório de Produtos Naturais, Centro de Ciência Exatas e Tecnologia, Departamento de Química, 65080-805 São Luís-MA, Brazil e Experimental Campus of the Paulista Coast, UNESP-São Paulo State University, 11330-900 São Vicente-SP, Brazil

Published: Molecules 2020, 25 (21), 5219; https://doi.org/10.3390/molecules25215219

ABSTRACT

Arrabidaea brachypoda is a plant commonly used for the treatment of kidney stones, arthritis and pain in traditional Brazilian medicine. Different in vitro and in vivo activities, ranging from antinociceptive to anti-Trypanosoma cruzi, have been reported for the dichloromethane root extract of Arrabidaea brachypoda (DCMAB) and isolated compounds. This work aimed to assess the in vitro anti-inflammatory activity in arthritic synoviocytes of the DCMAB, the hydroethanolic extract (HEAB) and three dimeric flavonoids isolated from the DCMAB. These compounds, brachydin A (1), B (2) and C (3), were isolated both by medium pressure liquid and high-speed counter current chromatography. Their quantification was performed by mass spectrometry on both DCMAB and HEAB. IL-1β activated human fibroblast-like synoviocytes were incubated with both extracts and isolated compounds to determine the levels of pro-inflammatory cytokine IL-6 by enzyme-linked immunosorbent assay (ELISA). DCMAB inhibited 30% of IL-6 release at 25 µg/mL, when compared with controls while HEAB was inactive. IC50 values determined for 2 and 3 were 3-fold higher than 1. The DCMAB activity seems to be linked to higher proportions of compounds 2 and 3 in this extract. These observations could thus explain the traditional use of A. brachypoda roots in the treatment of osteoarthritis. Keywords: Arrabidaea brachypoda; flavonoids; anti-inflammatory activity; osteoarthritis; high-speed counter current chromatography; mass-spectrometric quantification

Page 51: Thesis - archive-ouverte.unige.ch

Chapter II  

44

1. INTRODUCTION

Arrabidaea brachypoda (D.C.) is a shrub native to the Brazilian region of Cerrado

(neotropical savanna). It belongs to the Bignoniaceae family, which includes 120

genera and nearly 800 species of different plants scattered in tropical and

subtropical regions worldwide [1]. Plants of Arrabidaea genus are known sources of

C-glucosylxanthones, phenylpropanoids, flavonoids, anthocyanidins, allantoins,

and triterpenes [2–4]. These molecules are linked to the astringent, anti-

inflammatory, antimicrobial, antitumoral and wound healing properties that plants of

this genus are known for in traditional medicine [5]. In Brazil, Arrabidaea brachypoda

is known as “cervejinha do campo”, a decoction of its roots used to treat kidney

stones and arthritic joints [6,7]. Different molecules have been isolated and identified

from the hydroethanolic extract (HEAB) and dichloromethane extract (DCMAB).

From this last extract, three isolated aglycones - brachydin A (1), brachydin B (2),

and brachydin C (3) - have been identified. In the HEAB, among the 14 described

molecules, these three dimeric flavonoids exist in their glucoronated form [8,9].

Antinociceptive, anti-inflammatory, anti-Trypanosoma cruzi, gastroprotective,

antileishmanial, and antimicrobial activities have been recently reported in different

in vitro and in vivo assays from both root extracts and isolated compounds [8–15].

In these previous studies, different oral and/or topical administration setups were

explored. Overall, studies reporting on anti-inflammatory and antinociceptive

activities of A. brachypoda have focused on general mechanisms of pain and

inflammation [10,12]. In this study, we aimed to test the in vitro anti-inflammatory

activity of A. brachypoda root extracts and three isolated dimeric flavonoids in the

context of osteoarthritis (OA). This chronic disease has worldwide incidence, in

particular in the aging population and is considered the most common form of

arthritis. Characterized by chronic pain and inflammation, articular cartilage

degeneration, and structural changes of whole joints, OA represents the main cause

of physical disability and a great health economic burden [16–19]. Presently,

treatment options are primarily based on non-pharmacological and symptom

management approaches. There is a need for long-acting, targeted local anti-

inflammatory drug products, in order to tackle the main OA complications: pain and

inflammation [20,21]. Human fibroblast-like synoviocytes activated for OA-like

inflammation were used as the in vitro cellular model. Interleukin-6 (IL-6) is a

Page 52: Thesis - archive-ouverte.unige.ch

Anti-inflammatory activity of A. brachypoda extracts  

45

ubiquitous pro-inflammatory cytokine of acute inflammation, particularly involved in

the synovial hypertrophy, as well as an established therapeutic target for arthritis.

Therefore, IL-6 was selected as the molecular marker of inflammation in this study

[21]. Additionally, the isolated compounds were quantified in both HEAB and

DCMAB to establish a link between the chemical composition of the extracts and

anti-inflammatory properties of their individual constituents.

2. MATERIALS AND METHODS

2.1. Materials

2.1.1. General experimental procedures

All analytical HPLC-PDA analyses were performed using an Agilent Technologies

1260 Infinity system equipped with a photodiode array detector (Agilent

Technologies, Santa Clara, CA, USA). Preparative medium pressure liquid

chromatography (MPLC) was performed using a system equipped with a C-605

module pump, C-640 UV detector, and C-684 fraction collector all from Büchi (Flawil,

Switzerland). The system was controlled by the Sepacore Control software (Büchi

AG, Flawil, Switzerland). A coil connected to a resistance was used to control the

MPLC column temperature. The column (460 mm × 49 mm i.d.) was packed with

ZEOprep® C18 as the stationary phase (ZEOprep® C18, 15–25 µm; Zeochem,

Uetikon am See, Switzerland). Nuclear magnetic resonance (NMR) spectroscopic

data were recorded on a Bruker Avance III HD 600 MHz NMR spectrometer

equipped with a QCI 5 mm Cryoprobe and a SampleJet automated sample changer

(Bruker BioSpin, Rheinstetten, Germany). Chemical shifts were reported in parts per

million (δ) using the CD3OD residual signal (δH 3; δC 49) as internal standards for

1H and 13C NMR and coupling constants (J) were reported in hertz. High-speed

counter-current chromatography (HSCCC) coupled to UV was performed on a Tauto

TBE-300B instrument (Tauto Biotech, Shangai, China) equipped with two LC10AD

HPLC pumps (Shimadzu, Kyoto, Japan), a 20 mL injection loop, a Knauer K 2501

UV detector (Berlin, Germany) and a C-684 fraction collector from Büchi (Flawil,

Switzerland). Solvents used for extraction (methanol, ethyl acetate, hexane) were

all of analytical grade. Solvents used in the quantification of the isolated brachydins:

methanol, formic acid, acetonitrile and water were of LC-MS grade. All solvents were

Page 53: Thesis - archive-ouverte.unige.ch

Chapter II  

46

purchased from Sigma-Aldrich (St. Louis, MO, USA). Milli-Q® water from Merck

(Burlington, MA, USA) was used throughout this study. All other chemical products

were obtained from Sigma-Aldrich (St. Louis, MO, USA). For the in vitro anti-

inflammatory assays, all media solutions and ELISA kits were purchased from

Invitrogen (Carlsbad, CA, USA). Fetal bovine serum was purchased from Eurobio

(Les Ulis, France). Interleukin 1 beta (IL-1β) was obtained from R&D Systems (Bio-

Techne, Abingdon, UK).

2.1.2. Plant material

Roots of Arrabidaea brachypoda were collected in April 2010 from the Sant’Ana da

Serra farm in João Pinheiro, Minas Gerais, Brazil. The plant was identified at the

ICEB of José Badine Herbarium of the Federal University of Ouro Preto by Prof.

Maria Cristina Teixeira Braga Messias. A voucher specimen (# 17935) was

deposited at the Herbarium of the Federal University of Ouro Preto, Brazil. The plant

was collected in accordance with Brazilian authorities (SISGEN # A451DE4).

2.2. Extraction and isolation of brachydins A (1), B (2) and C (3)

2.2.1. Plant extraction

Both HEAB and DCMAB of the roots of Arrabidaea brachypoda were obtained from

our previous study [9]. These extracts were stored and protected from light at -20

°C.

2.2.2. HPLC-PDA analysis

The DCMAB of A. brachypoda (Figure S1, Supplementary Material), the MPLC and

HSCCC fractions were analyzed by HPLC-PDA using a Waters X-Bridge C18

column (250 mm × 4.6 mm i.d., 5 μm; Waters, Milford, MA, USA) equipped with a

Waters C18 pre-column cartridge holder (10 mm × 2.1 mm i.d.). Water (A) and

methanol (B), both containing 0.1% of formic acid (FA) were used as the solvent

system. The column was equilibrated with 5% of B for 15 min. The separation was

performed in gradient mode, as follows: 5 to 100% of B in 60 min, and 100% of B

for 10 min. Flow rate 1 mL/min; injection volume 10 μL; sample concentration 10

mg/mL in methanol. The UV absorbance was measured at 254 nm and the UV-PDA

spectra were recorded between 190 and 600 nm (step 2 nm).

Page 54: Thesis - archive-ouverte.unige.ch

Anti-inflammatory activity of A. brachypoda extracts  

47

2.2.3. Isolation of brachydins A (1), B (2), and C (3)

DCMAB was fractioned by MPLC, following a previously published protocol [22].

Fractionation conditions were optimized on an HPLC column (see Sections 2.1.1

and 2.2.2) using the same stationary phase and an acidic (0.1% FA) water (A) and

methanol (B) gradient. Solvent system gradient went as follows: isocratic 5% B in

56 min, 5 to 38% B in 5.5 h, isocratic 38% B in 4.7 h, 38 to 62% B in 3.2 h, 62 to

100% B in 9 min and isocratic 100% B during 1.5 h. A dry load injection of the sample

was performed by mixing 7 g of the dichloromethane extract with 35 g of Zeoprep®

C18 (40–63 µm). The dry-load cell (11.5 cm × 2.7 cm i.d.) was subsequently

connected between the pumps and the MPLC column. The flow rate was set to 20

mL/min and UV absorbance was monitored at 280 nm. MPLC separation yielded 90

fractions of 250 mL, which were analyzed by ultrahigh pressure liquid

chromatography coupled to an UV detector (UHPLC-UV). This approach yielded

gram amounts of compound 1. Fractions containing pure compounds were

combined, dried, analyzed by nuclear magnetic resonance (NMR) and properly

stored. Other MPLC fractions containing mixtures of compounds 2 and 3 were

combined and subjected to purification by HSCCC coupled to a UV detector. The

ARIZONA solvent system approach was applied to perform the separation of

compounds 2 and 3. This approach is based on the use of 23 solvent mixture

compositions of methanol/ethyl acetate/hexane/water. The mixtures are labeled with

letters from the alphabet from A to Z (except E, I, and O) [23]. To determine the best-

suited ARIZONA solvent mixture, the partition coefficient (Kp) of each compound

was determined. As so, few milligrams of the fraction containing 2 and 3 were

solubilized in a specific mix of the ARIZONA system. Upper and lower phases of

each mixture of solvents were separated and analyzed by HPLC-UV. The UV area

peak of each compound was used to determine its coefficient of partition (Kp)

according to the equation in Figure S2 (Supplementary Material). The best results

for complete separation of compounds 2 and 3 were obtained by solvent mixture P-

methanol/ethyl acetate/hexane/water (6:5:6:5; v/v/v/v). The coil was first filled with

the two phases (upper and lower, 1:1) and rotation was set to 1000 rpm. Lower

phase was then pumped into the column at a flow rate of 3 mL/min using the head-

to-tail mode (mobile phase = lower phase; stationary phase = upper phase) with

rotation set at 8000 rpm. After equilibrium between the two phases, 500 mg of

sample in 20 mL of upper and lower phase solution (1:1) were injected. Four

Page 55: Thesis - archive-ouverte.unige.ch

Chapter II  

48

injections were performed. A total of 35 fractions (5 ml each) were obtained for each

injection. The fractions were combined according to the chemical composition

determined by HPLC-PDA analysis, dried, analyzed by NMR, and properly stored.

2.3. Quantification of compounds 1–3 from A. brachypoda root extracts by

uhplc-ms/ms

Sample preparation was carried out by dissolving the dried HEAB and DCMAB from

Arrabidaea brachypoda roots in 100% methanol. The final concentration of the

samples was 10 µg/mL, and three technical replicates were used for quantitative

analysis. The amounts of compounds 1, 2, and 3 in the extracts were quantified on

a QTRAP 4000 quadrupole linear ion-trap mass spectrometer (Sciex, Darmstadt,

Germany) with an ESI interface operating in positive ionization mode. UHPLC was

performed on an Acquity UPLC system (Waters, Milford, MA, USA) equipped with a

Waters UPLC BEH C18 column (50 mm × 2.1 mm ID, 1.7 µm). The solvent system

consisted of water: 0.1% FA for solvent A and acetonitrile: 0.1% FA for solvent B.

The elution was performed in gradient mode at 40 °C (400 µL/min flow rate) with the

following steps: 5 to 60% B in 4.2 min, 60 to 70% B in 1.3 min, 70 to 100% B in 1.5

min, maintaining B at 100% for 1 min and lastly, a reconditioning with 5% B for min.

The autosampler compartment was set at 10 °C throughout the analysis and the

injection volume was 5 µL. MS/MS quantitative analyses were performed in multiple

reaction monitoring (MRM) mode. Source and gas parameters were as follows:

curtain gas, 10 psi; collision gas, 10 psi; ionspray voltage, 3500 V; source

temperature, 400 °C; ion source gas 1, 40 psi; ion source gas 2, 50 psi. Nitrogen

was used as collision gas. MRM transitions and MRM parameters were all optimized

via flow injection analysis (FIA) technique with a mix containing 10 ng/mL of pure 1,

2, and 3 analytes (Table S1, Supplementary Material). Quantification was achieved

by constructing three separate external calibration curves from solutions of the

isolated pure compounds. The calibration curve concentrations for active

compounds were as follows: 500 ng/mL, 250 ng/mL, 125 ng/mL, 63 ng/mL, 31

ng/mL. As no analyte-free matrix was available, the displayed limits of quantification

(LOQ) were evaluated only on each pure standard. A signal-to-noise ratio of 10 was

considered as the approximate LOQ (Table S2, Supplementary Material).

Page 56: Thesis - archive-ouverte.unige.ch

Anti-inflammatory activity of A. brachypoda extracts  

49

2.4. In vitro anti-inflammatory bioactivity

2.4.1. Human fibroblast-like synoviocytes (HFLS) isolation and culture

Hip synovial membrane was collected from three male adult patients with clinical

osteoarthritis (OA) at the time of hip replacement surgery. This protocol was

conducted under the approval of the local Ethics Committee (CCER, Geneva,

Switzerland) (Authorization # 2017-02234) and with informed and consenting

patients. Once collected, tissue samples were processed, according to previously

described protocols [25]. Briefly, samples were finely minced and digested for 3 h

(37 °C, 5% CO2 incubation) in a 3 mg/mL collagenase IX-RPMI 1640 solution. After

centrifugation (200 × g) and supernatant removal, the resuspended pellet was

cultured (37 °C, 5% CO2) in medium containing RPMI 1640, M199 (1:1), 1%

penicillin/streptomycin (100 IU/mL:100 g/mL), 2 mM L-glutamine and 20% fetal

bovine serum. After overnight culture, non-adherent cells were removed. At

confluence, cells were trypsinized and passaged to 75 cm2 culture flasks (Corning,

NY, USA) in complete medium containing 10% fetal bovine serum. A double control

was performed to confirm presence of HFLS after isolation: visual morphology

evaluation (Figure S3) and flow cytometry with synoviocyte specific surface markers

(CD14+ and VCAM-1). A representative sample of all cell populations was found to

have above 90% HFLS markers. All cellular assays were performed from passage

3 to 9. Experiments were conducted twice per donor (n = 6).

2.4.2. In vitro cytotoxicity assay and bioactivity assay with the determination of il-6

release in HFLS

Confluent HFLS cells (25,000 cells/well) were treated in a 96-well plate with

compounds 1, 2, and 3 (3 µM, 6 µM, 12 µM, 25 µM, 50 µM, and 100 µM), HEAB and

DCMAB (increasing concentrations from 3 to 100 µg/mL), or control vehicle for 1 h.

All tested compounds were dissolved in a dimethyl sulfoxide (DMSO) stock solution

and further diluted in culture medium. Final concentration of DMSO in vehicle of

tested concentrations was 0.01%. After 1 h incubation, cells were activated by

addition of IL-1β (1 µg/mL) (R&D Systems, Bio-Techne, Abingdon, UK) with

subsequent incubation for 23 h. All supernatants were kept at −80 °C, before further

testing. Remaining adherent cells were tested for viability using the cell proliferation

reagent WST-1 (Roche, Basel, Switzerland), according to the supplier instructions.

Pro-inflammatory cytokine IL-6 release in cell supernatant was assessed by ELISA.

Page 57: Thesis - archive-ouverte.unige.ch

Chapter II  

50

For this, a human IL-6 ELISA kit from Invitrogen (ref # 88-7066-88; Thermo-Fisher

Scientific, Waltham, MA, USA) was used according to the manufacturer’s protocol.

All samples were diluted 90 times. Experiments were conducted twice per donor (n

= 6).

2.5. Half maximal inhibitory concentration (IC50) determination and

statistical analysis

In this study, data are presented as mean values ± standard deviation (s.d.). All

analyses were performed using GraphPad Prism 8.3 software. IC50 values were

determined using a nonlinear fitting of a log (inhibitor concentration) vs response

variable slope for each donor’s data set. Consequent values were then averaged.

After confirming the data normal distribution and homogeneity of variances (Brown-

Forsythe and Welch; Shapiro-Wilk, D’Agostino-Pearson and Anderson-Darling),

statistical analysis was performed using a two-way ANOVA with a Bonferroni

multiple comparisons test. Significance was determined at alpha level 0.05.

Significance values represented are * p < 0.05, ** p < 0.009 and ns for non-

significance.

3. RESULTS AND DISCUSSION

3.1. Isolation of compounds 1–3 from the DCMAB of A. brachypoda

The DCMAB of A. brachypoda was obtained according to the protocol described in

our previous study [8]. The high-performance liquid chromatography with

photodiode array detection (HPLC-PDA) analysis revealed three major compounds

(Figure 1) which were previously characterized as three unusual dimeric flavonoids

named brachydin A (1), brachydin B (2), and brachydin C (3) [8]. To study the anti-

inflammatory properties of these molecules, DCMAB was purified at large scale. As

a first step, separation conditions were optimized at HPLC-PDA analytical scale

(Figure S1, Supplementary Material) and transferred to medium pressure liquid

chromatography (MPLC) at a semi-preparative scale, using a gradient transfer

method. This yielded 4 g of brachydin A (compound 1; Figure 1A) in a single step.

Brachydin B (2) and brachydin C (3) were obtained in a mixture (4.9 g) (see Section

2.2.3), mostly due to a column overload, and were further purified by high-speed

counter current chromatography (HSCCC). The solvent system was determined

Page 58: Thesis - archive-ouverte.unige.ch

Anti-inflammatory activity of A. brachypoda extracts  

51

using the ARIZONA solvent system approach. The separation was carried out

successfully as a result of the different Kp determined (compounds 2 (Kp = 1.5) and

3 (Kp = 2.2) (see Section 2.2.3). A full recovery of pure compound 2 (1.4 g) and 3

(3.2 g) (Figure 1C) resulted, as HSCCC is a support-free liquid-liquid partition

chromatography technique [24]. According to the literature, this is the first time this

technique has been described for the isolation of pure compounds from A.

brachypoda root extracts.

Page 59: Thesis - archive-ouverte.unige.ch

Chapter II  

52

Figure 1. (A) Structures of brachydin A (1), brachydin B (2), and brachydin C (3) isolated from the dichloromethane extract (DCMAB) of A. brachypoda roots (B) Determination of coefficient of partition Kp by high-performance liquid chromatography with photodiode array detection (HPLC-PDA) of the solvent mixture P: methanol/ethyl acetate/hexane/water (6:5:6:5; v/v/v/v) of the ARIZONA system [23]. (C) High-speed counter current chromatography (HSCCC)-UV chromatogram at 280 nm, with clear separation of compounds 2 and 3.

Page 60: Thesis - archive-ouverte.unige.ch

Anti-inflammatory activity of A. brachypoda extracts  

53

3.2. Quantification of compounds 1–3 from A. brachypoda root extracts by

UHPLC-MS/MS

The amounts of compounds 1, 2, and 3 in the HEAB and DCMAB were determined

by ultrahigh pressure liquid chromatography coupled to mass spectrometer

(UHPLC-MS/MS) analysis, in the multiple reaction monitoring (MRM) mode. The

MRM parameters of each analyte were optimized to increase sensitivity for a specific

transition from mass spectrum2 (MS2) reported data (Table S1, Supplementary

Material) [8]. The range of the calibration curves was estimated for each compound

and was set to 31–500 ng/mL. This yielded five data point calibration curves with r2

> 0.99 (Table S2, Supplementary Material). The results obtained from the

quantitative analysis are displayed in Figure 2. The total content of compounds 1–3

was 3.35-fold higher in DCMAB than in HEAB. This was expected, since these

compounds are primarily occurring as glucoronated derivatives in HEAB [8], and

these are not detected by MRM. Compound 3 was the most abundant in both

DCMAB (123 mg/g (DW: dry weight) and HEAB (36 mg/g (DW)). Compounds 1 and

2 were present at 16 and 108 mg/g (DW) in DCMAB, respectively. Lower levels of

1 (5 mg/g) and 2 (33 mg/g (DW)) were found in HEAB. These results quantify, for

the first time, the amounts of the three isolated brachydins in both extracts.

Figure 2. Concentration (mg/g DW) of compounds 1–3 in hydroethanolic extract (HEAB) and dichloromethane extract (DCMAB), respectively. Bars correspond to mean values ± s.d.; n = 3.

Page 61: Thesis - archive-ouverte.unige.ch

Chapter II  

54

3.3. In vitro cytotoxicity and anti-inflammatory bioactivity of extracts and

isolated compounds

3.3.1. Arrabidaea bachypoda extracts (HEAB and DCMAB)

The cytotoxicity of both HEAB and DCMAB was assessed; results are shown in

Figure 3A (and corresponding scatter plot in Figure S4). Human fibroblast like

synoviocytes (HFLS) were viable after 24 h incubation, with tested concentrations

ranging from 3 to 100 µg/mL of HEAB. Conversely, at 50 μg/mL and 100 μg/mL of

DCMAB, HFLS presented viability values of 68% and 5%, respectively, compared

to controls (Figure 3A). Consequently, these two top concentrations were not

considered in the bioactivity assay, a sandwich enzyme-linked immunosorbent

assay (ELISA) (Figure 4A), due to toxicity. At the highest dose tested (100 μg/mL),

HEAB showed a 30% inhibition of IL-6 release. DCMAB yielded a similar inhibitory

effect at 25 μg/mL (31% inhibition; Table 1), the highest concentration inducing no

cytotoxicity. Both inhibition effects were non-significant (p value 0.27 and 0.19,

respectively) when compared to non-treated control: interleukin-1 beta (IL-1 β). In

line with these results, IC50 calculations (Table 1) show that an anti-inflammatory

effect occurs only above 100 µg/mL and 31 µg/mL for HEAB and DCMAB,

respectively.

3.3.2. Isolated brachydins from DCMAB (1–3)

The three isolated compounds 1, 2, and 3, were also tested for cytotoxicity (WST-

1) and anti-inflammatory activity (ELISA) and showed different activities (Figures

3B, S4 and 4B). Compound 1 was only cytotoxic at 100 µM, whereas compounds 2

and 3 induced a loss of viability at 50 µM, with no evident effect of dose increase,

unlike what was previously described for DCMAB in terms of cytotoxicity (Figure 3).

For this reason, the IL-6 ELISA assay was performed using supernatants recovered

after incubation with HFLS, up to 50 µM for 1 and only up to 25 µM for 2 and 3

(Figure 4B). At the highest concentration with remaining favorable viability (25 µM),

both 2 and 3 significantly decreased IL-6 release when compared to the activation

control IL-1β (80% and 94% inhibition, respectively (Table 1). Compound 1 showed

a non-significant inhibition of IL-6 release at 25 µM and 50 µM (10% and 24%,

respectively; Figure 4B and Figure S5, Supplementary Material). In Table 1, IC50

values (individual plots represented in Supplementary Material, Figure S6)

confirmed the increased anti-inflammatory effect of 2 (17 µM) and 3 (19 µM), in

Page 62: Thesis - archive-ouverte.unige.ch

Anti-inflammatory activity of A. brachypoda extracts  

55

comparison to 1 (62 µM). These different outcomes between compounds could be

linked to the fact that both compounds 2 and 3 are of higher structural similarity

between each other than compound 1, that exhibits higher polarity influencing

effects. Regarding cell viability of HFLS, HEAB and compound 1 presented the least

cytotoxicity and DCMAB, compounds 2 and 3, showed higher cytotoxicity. This is

expected, since these are the major two components of this extract, as above

mentioned (see Section 3.2). In order to compare the bioactivity results between the

tested compounds, IC50 results in µM (isolated compounds) were transformed into

µg/mL to match those of the extracts, that represent mixtures of several compounds

(Table 1). In terms of the anti-inflammatory effect, results are in line with those of

cytotoxicity. HEAB (IC50 > 100 µg/mL) was inactive in decreasing release of pro-

inflammatory cytokine IL-6 in HFLS. According to what was previously reported for

HEAB [9–11], this extract is comprised by the dimeric flavonoid aglycones (1–3) and

their various glucoronated derivatives [9]. The glucoronated derivatives appear in

higher amount than the corresponding aglycones. In an HFLS in vitro setting, the

metabolic pathways that hydrolyze the glucoronated forms into simple aglycones

(as it would in the acidic pH of the stomach) do not occur and the compounds are

absorbed as is [26]. Based on such considerations, the lack of activity of HEAB can

be thus correlated to the low amounts of 1, 2, and 3 (Figure 2) and the glucoronated

derivatives themselves, inactive in this experimental setup. This is also the

reasoning as to why all current studies and evaluations of this extract in vivo occur

after per os administration and not local [8–10,12]. DCMAB on the other hand, only

contains the three isolated brachydins A, B, and C in their non-glucoronated forms.

Such extract is thus more interesting to explore in this experimental setup and when

seeking local administration of anti-inflammatory treatments for OA. Albeit through

unknown pathways, this study is the first analyzing the anti-inflammatory activity in

the specific context of OA. Additionally, it compares effects and quantifies the

presence of isolated compounds in both A. brachypoda extracts in order to assess

their potential as OA therapeutic alternatives. Compound 1 exhibited similar anti-

inflammatory activity (IC50) when compared to DCMAB (33 µg/mL and 31 µg/mL,

respectively), whereas compounds 2 (9 µg/mL) and 3 (10 µg/mL) showed a 3-fold

higher activity. This suggests that the extracts activity results mainly from the

presence of the two latter compounds. Taking into account the amounts of each

compound 1-3 in the DCMAB (Table 1), the cumulative IC50 of each individual

Page 63: Thesis - archive-ouverte.unige.ch

Chapter II  

56

isolated compound (27 µg/mL) does not greatly differ from DCMAB alone - 31

µg/mL. Based on these results, no investigation on synergistic effect was pursued.

Differences between the anti-inflammatory activities between the three compounds

potentially stem from the higher polarity of compound 1, in comparison to 2 and 3.

Figure 3. The cellular viability of human fibroblast like synoviocytes (HFLS) incubated with all tested compounds at increasing concentrations, after 24 h. Root extracts (A) and isolated compounds from dichloromethane extract (DCMAB) (B). Bars correspond to mean values ± s.d.; n = 6; V = vehicle, 0.01% dimethyl sulfoxide (DMSO). Additional scatter plot with individual data in Supplementary Material, Figure S4.

Page 64: Thesis - archive-ouverte.unige.ch

Anti-inflammatory activity of A. brachypoda extracts  

57

Figure 4. Release of pro-inflammatory cytokine IL-6 from IL-1β activated HFLS incubated with root extracts (A) and isolated compounds 1–3 (B), after 24 h. Bars correspond to mean values ± s.d.; n = 6. *p < 0.05, **p < 0.009 and ns = no significance. Dotted lines represent cut-off of maximum IL-6 release by Il-1β stimulation.

Page 65: Thesis - archive-ouverte.unige.ch

Chapter II  

58

Table 1. IL-6 inhibitory activity and anti-inflammatory IC50 of all tested compounds.

Extracts IC50 (µg/mL) Inhibition of IL-6 release at highest non-toxic

concentration (%) [25 µg/ml] HEAB >100 2 ± 12

DCMAB 31 ± 1 31 ± 8

Compounds Molecular

weight (g/mol)

IC50

Inhibition of IL-6 release at

highest non-toxic

concentration (%)

[25 µM = 13 µg/ml *]

Compound in DCMAB

(% of DW **)

Compound concentration in IC50 of DCMAB

(µg/mL) (µg/mL) (µM)

1 524 33 ± 2 62 ± 2 10 ± 11 2 0.5 2 538 9 ± 3 17 ± 3 80 ± 9 10 3.4 3 508 10 ± 3 19 ± 3 94 ± 5 12 3.8

* Mean value for equivalence to 25 µM of Compound 1, 2 and 3. ** Dry Weight (DW).

4. CONCLUSIONS

Arrabidaea brachypoda is a plant commonly used in Brazilian traditional medicine

to treat pain and inflammation related to arthritic joints. The effects of HEAB,

DCMAB and three isolated dimeric flavonoids were assessed. In this study, specific

in vitro anti-inflammatory activity was tested using primary human arthritic

synoviocytes, in order to understand the mechanisms involved and potentially

identify molecules of interest for local drug delivery in OA. The present study is the

first comparing the anti-inflammatory effect and providing a quantification of the

three isolated compounds from the DCMAB. Despite high cellular viability scores at

tested concentrations, HEAB was not performant in terms of anti-inflammatory

effect. Studies suggest that the acidic hydrolysis in the stomach is responsible for

the releasing of aglycones deriving from glucoronated forms of not only brachydins

A, B, and C, but also other molecules of the extract [8,26]. In this study setup

however, hydrolysis does not take place, which explains the lack of activity of HEAB.

Conversely, DCMAB showed an effect against the release of pro-inflammatory

cytokine IL-6. Quantification of 1, 2, and 3 in this extract showed higher proportions

of 2 and 3 compared to 1. The bioactivity results are in line with these findings, where

1 alone is considerably less active than to 2 and 3 alone. Compound 3 presented

as both the most effective and cytotoxic compound, the fact that it is the main

Page 66: Thesis - archive-ouverte.unige.ch

Anti-inflammatory activity of A. brachypoda extracts  

59

component of DCMAB could account for the extract’s enhanced effect over both 1

and 2 and, partly, HEAB. Differences in polarity between compound 1 versus 2 and

3 can affect bioavailability and bioactivity. Compounds 2 and 3 seem to have a major

role in the anti-inflammatory effects of DCMAB and, potentially, other aglycones

present in HEAB deserve further exploring in terms of anti-inflammatory bioactivity.

Additionally, further research of in vivo studies and potential drug delivery systems

would enrich the understanding of the mechanisms underlying A. brachypoda roots

and its treatment of arthritic joints.

5. REFERENCES

[1] G. Lino Von Poser, J. Schripsema, A.T. Henriques, S. Rosendal Jensen, The distribution of iridoids in Bignoniaceae, Biochem. Syst. Ecol. 28 (2000) 351–366. https://doi.org/10.1016/S0305-1978(99)00076-9. [2] P. Mendonça Pauletti, I. Castro-Gamboa, D.H. Siqueira Silva, M.C. Marx Young, D.M. Tomazela, M. Nogueira Eberlin, V. Da Silva Bolzani, New Antioxidant C-Glucosylxanthones from the Stems of Arrabidaea samydoides, J. Nat. Prod. 66 (2003) 1384–1387. https://doi.org/10.1021/np030100t. [3] B. Gonzalez, H. Suarez-Roca, A. Bravo, R. Salas-Auvert, D. Avila, Chemical composition and biological activity of extracts from Arrabidaea bilabiata, Pharm. Biol. 38 (2000) 287–290. https://doi.org/10.1076/1388-0209(200009)3841-AFT287. [4] F. Martin, A.E. Hay, D. Cressend, M. Reist, L. Vivas, M.P. Gupta, P.A. Carrupt, K. Hostettmann, Antioxidant C-glucosylxanthones from the leaves of Arrabidaea patellifera, J. Nat. Prod. 71 (2008) 1887–1890. https://doi.org/10.1021/np800406q. [5] J.P. V. Leite, A.B. Oliveira, J.A. Lombardi, J.D.S. Filho, E. Chiari, Trypanocidal Activity of Triterpenes from Arrabidaea triplinervia and Derivatives, Biol. Pharm. Bull. 29 (2006) 2307–2309. https://doi.org/10.1248/bpb.29.2307. [6] T. Alcerito, F.E. Barbo, G. Negri, D.Y.A.C. Santos, C.I. Meda, M.C.M. Young, D. Chávez, C.T.T. Blatt, Foliar epicuticular wax of Arrabidaea brachypoda: Flavonoids and antifungal activity, Biochem. Syst. Ecol. 30 (2002) 677–683. https://doi.org/10.1016/S0305-1978(01)00149-1. [7] E. Rodrigues, F. Mendes, G. Negri, Plants indicated by Brazilian Indians to Central Nervous System disturbances: A bibliographical approach, Curr Med Chem. 6 (2005) 211–244. https://doi.org/10.2174/187152406778226725. [8] C.Q. Da Rocha, E.F. Queiroz, C.S. Meira, D.R.M. Moreira, M.B.P. Soares, L. Marcourt, W. Vilegas, J.L. Wolfender, Dimeric flavonoids from Arrabidaea brachypoda and assessment of their anti- Trypanosoma cruzi activity, J. Nat. Prod. 77 (2014) 1345–1350. https://doi.org/10.1021/np401060j. [9] C.Q. da Rocha, F.M. de-Faria, L. Marcourt, S.N. Ebrahimi, B.T. Kitano, A.F. Ghilardi, A. Luiz Ferreira, A.C.A. de Almeida, R.J. Dunder, A.R.M. Souza-Brito, M. Hamburger, W. Vilegas, E.F. Queiroz, J.L. Wolfender, Gastroprotective effects of hydroethanolic root extract of Arrabidaea brachypoda: Evidences of cytoprotection and isolation of unusual glycosylated polyphenols, Phytochemistry. 135 (2017) 93–105. https://doi.org/10.1016/j.phytochem.2016.12.002. [10] C.Q. Da Rocha, F.C. Vilela, G.P. Cavalcante, F. V. Santa-Cecília, L. Santos-E-Silva, M.H. Dos Santos, A. Giusti-Paiva, Anti-inflammatory and antinociceptive effects of Arrabidaea brachypoda (DC.) Bureau roots, J. Ethnopharmacol. 133 (2011) 396–401. https://doi.org/10.1016/j.jep.2010.10.009.

Page 67: Thesis - archive-ouverte.unige.ch

Chapter II  

60

[11] F.A. Resende, C.H. Nogueira, L.G. Espanha, P.K. Boldrin, A.P. Oliveira-Höhne, M. Santoro de Camargo, C. Quintino da Rocha, W. Vilegas, E.A. Varanda, In vitro toxicological assessment of Arrabidaea brachypoda (DC.) Bureau: Mutagenicity and estrogenicity studies, Regul. Toxicol. Pharmacol. 90 (2017) 29–35. https://doi.org/10.1016/j.yrtph.2017.08.010. [12] V. Rodrigues, C. Rocha, L. Périco, R. Santos, R. Ohara, C. Nishijima, E. Ferreira Queiroz, J.-L. Wolfender, L. Rocha, A. Santos, W. Vilegas, C. Hiruma-Lima, Involvement of Opioid System, TRPM8, and ASIC Receptors in Antinociceptive Effect of Arrabidaea brachypoda (DC) Bureau, Int. J. Mol. Sci. 18 (2017) 2304. https://doi.org/10.3390/ijms18112304. [13] V. Rocha, C. Quintino da Rocha, E. Ferreira Queiroz, L. Marcourt, W. Vilegas, G. Grimaldi, P. Furrer, É. Allémann, J.-L. Wolfender, M. Soares, Antileishmanial Activity of Dimeric Flavonoids Isolated from Arrabidaea brachypoda, Molecules. 24 (2018) 1. https://doi.org/10.3390/molecules24010001. [14] L.M. de Sousa Andrade, A.B.M. de Oliveira, A.L.A.B. Leal, F.A. de Alcântara Oliveira, A.L. Portela, J. de Sousa Lima Neto, J.P. de Siqueira-Júnior, G.W. Kaatz, C.Q. da Rocha, H.M. Barreto, Antimicrobial activity and inhibition of the NorA efflux pump of Staphylococcus aureus by extract and isolated compounds from Arrabidaea brachypoda, Microb. Pathog. 140 (2020) 103935. https://doi.org/10.1016/j.micpath.2019.103935. [15] C.Q. Da Rocha, F.C. Vilela, F. V Santa-Cecília, G.P. Cavalcante, W. Vilegas, A. Giusti-Paiva, M.H. Dos Santos, Oleanane-type triterpenoid: an anti-inflammatory compound of the roots Arrabidaea brachypoda, Rev. Bras. Farmacogn. 25 (2015) 228–232. https://doi.org/10.1016/j.bjp.2015.03.005. [16] M.B. Goldring, S.R. Goldring, Osteoarthritis, J. Cell. Physiol. 213 (2007) 626–634. https://doi.org/10.1002/jcp.21258. [17] A.E. Nelson, K.D. Allen, Y.M. Golightly, A.P. Goode, J.M. Jordan, A systematic review of recommendations and guidelines for the management of osteoarthritis: The Chronic Osteoarthritis Management Initiative of the U.S. Bone and Joint Initiative, Semin. Arthritis Rheum. 43 (2014) 701–712. https://doi.org/10.1016/j.semarthrit.2013.11.012. [18] A. Koszowska, R. Hawranek, J. Nowak, Osteoarthritis -a multifactorial issue, Reumatologia. 52 (2014) 319–325. https://doi.org/10.5114/reum.2014.46670. [19] D.J. Hunter, D. Schofield, E. Callander, The individual and socioeconomic impact of osteoarthritis, Nat. Rev. Rheumatol. 10 (2014) 437–441. https://doi.org/10.1038/nrrheum.2014.44. [20] S.L. Kolasinski, T. Neogi, M.C. Hochberg, C. Oatis, G. Guyatt, J. Block, L. Callahan, C. Copenhaver, C. Dodge, D. Felson, K. Gellar, W.F. Harvey, G. Hawker, E. Herzig, C.K. Kwoh, A.E. Nelson, J. Samuels, C. Scanzello, D. White, B. Wise, R.D. Altman, D. DiRenzo, J. Fontanarosa, G. Giradi, M. Ishimori, D. Misra, A.A. Shah, A.K. Shmagel, L.M. Thoma, M. Turgunbaev, A.S. Turner, J. Reston, 2019 American College of Rheumatology/Arthritis Foundation Guideline for the Management of Osteoarthritis of the Hand, Hip, and Knee, Arthritis Rheumatol. 72 (2020) 220–233. https://doi.org/10.1002/art.41142. [21] B. Abramoff, F.E. Caldera, Osteoarthritis: Pathology, Diagnosis, and Treatment Options, Med. Clin. North Am. 104 (2020) 293–211. https://doi.org/10.1016/j.mcna.2019.10.007. [22] S. Challal, E.F. Queiroz, B. Debrus, W. Kloeti, D. Guillarme, M.P. Gupta, J.L. Wolfender, Rational and Efficient Preparative Isolation of Natural Products by MPLC-UV-ELSD based on HPLC to MPLC Gradient Transfer, Planta Med. 81 (2015) 1636–1643. https://doi.org/10.1055/s-0035-1545912. [23] A.P. Foucault, Centrifugal partition chromatography - Chromatographic Science Series, Taylor & Francis, 1995. [24] K. Hostettmann, A. Marston, M. Hostettmann, Countercurrent Chromatography, in: Prep. Chromatogr. Tech., Springer Berlin Heidelberg, Berlin, Heidelberg, 1998: pp. 135–201. https://doi.org/10.1007/978-3-662-03631-0_7. [25] L. Neff, M. Zeisel, V. Druet, K. Takeda, J.P. Klein, J. Sibilia, D. Wachsmann, ERK 1/2- and JNKs-dependent synthesis of interleukins 6 and 8 by fibroblast-like synoviocytes stimulated with

Page 68: Thesis - archive-ouverte.unige.ch

Anti-inflammatory activity of A. brachypoda extracts  

61

protein I/II, a modulin from oral streptococci, requires focal adhesion kinase, J. Biol. Chem. 278 (2003) 27721–27728. https://doi.org/10.1074/jbc.M212065200. [26] R.M. Hill, H.B. Lewis, The hydrolysis of sucrose in the human stomach, Am. J. Physiol. Content. 59 (1922) 413–420. https://doi.org/10.1152/ajplegacy.1922.59.1.413.

 

 

Page 69: Thesis - archive-ouverte.unige.ch

Chapter II  

62

Supplementary material

Figure S1. Reverse phase HPLC-UV analysis at 254 nm - optimized conditions for MPLC separation. Brachydin A (1), Brachydin B (2) and Brachydin C (3).

Figure S2. Coefficient of partition Kp equation (ARIZONA system).

Figure S3. Human fibroblast like synoviocytes (optical microscope). Scale bar: 200 µm.

Page 70: Thesis - archive-ouverte.unige.ch

Anti-inflammatory activity of A. brachypoda extracts  

63

Figure S4. Scatter plot of Figure 3, showing cellular viability of human fibroblast like synoviocytes (HFLS) incubated with all tested compounds at increasing concentrations, after 24 h. Root extracts (A) and isolated compounds from dichloromethane extract (DCMAB) (B). Individual plotted values (n = 6) with mean values ± s.d.. V = vehicle, 0.01 % dimethyl sulfoxide (DMSO).

Page 71: Thesis - archive-ouverte.unige.ch

Chapter II  

64

Figure S5. Inhibition percentage of IL-6 release normalized to 100 % activation, as a function of each compound concentration. Bars correspond to mean values ± s.d.; n = 6.**** p <0.0001 and *** p = 0.0007 and ns = no significance.

Figure S6. IC50 calculations of hydroethanolic extract (HEAB) (A), dichloromethane extract (DCMAB) (B), 1 (C), 2 (D) and 3 (E). Points correspond to mean values ± s.d.; n = 6.

Page 72: Thesis - archive-ouverte.unige.ch

Anti-inflammatory activity of A. brachypoda extracts  

65

SRM, Selected reaction monitoring; DP, Declustering potential; EP, Entrance potential; CE, Collision energy; CXP, Collision cell exit.

r2, Correlation coefficient; LOD, Limit of Detection; LOQ, limit of quantification

Table S1. Optimized MRM parameters for the quantification of active compounds by UHPLC-MS/MS.

Compound Precursor

Ion SRM

transition DP (V)

EP (V)

CE (eV)

CXP (V)

Dwell time (ms)

1 525 525 ➝ 271 30

10 10 7

125 2 539 539 ➝ 285 3 509 509 ➝ 255 80 50 12

Table S2. Calibration curves parameters for active compounds and determination of the LOD and LOQ in ng/mL.

Compound Linear function r2 LOD (ng/mL) LOQ (ng/mL)

1 𝑦 = 816𝓍 + 6,06.103

0.998

0.21 0.64

2 𝑦 = 422𝓍 + 3,51.103 0.47 1.41

3 𝑦 = 122𝓍 + 639 1.42 4.25

Page 73: Thesis - archive-ouverte.unige.ch

66

Page 74: Thesis - archive-ouverte.unige.ch

 

 

Chapter III

Page 75: Thesis - archive-ouverte.unige.ch

 

Page 76: Thesis - archive-ouverte.unige.ch

 

69

Chapter III

Nano wet milled celecoxib extended release microparticles

for local management of chronic inflammation

Carlota Salgado a,b, Laure Guénée c, Radovan Černý c, Eric Allémann a,b and Olivier Jordan a,b a School of Pharmaceutical Sciences, University of Geneva, 1 Rue Michel-Servet 1211 Geneva, Switzerland b Institute of Pharmaceutical Sciences of Western Switzerland, University of Geneva, 1 Rue Michel-Servet 1211 Geneva, Switzerland c Department of Quantum Matter Physics, Laboratory of Crystallography, University of Geneva, 24 Quai Ernest Ansermet 1211 Geneva, Switzerland

Published: International Journal of Pharmaceutics 2020, 589, 119783; https://doi.org/10.1016/j.ijpharm.2020.119783

ABSTRACT

Osteoarthritis (OA), the most common form of arthritis, is characterized by chronic inflammation, degeneration of articular cartilage and whole joints. Local delivery by intra-articular (IA) injection of small molecules is an established treatment to relieve pain and improve joint motion, requiring month-lasting release of therapeutic drug doses. We incorporated anti-inflammatory drug celecoxib in poly (D, L-lactic acid) microparticles using two spray-drying approaches - either as a solid solution or embedded as milled nano drug. Differential scanning calorimetry, X-ray powder diffraction, electron microscopy and in vitro drug release allowed comparison of the microparticles. Both types resulted in spherical particles ranging from 20 to 40 μm mean size, with high drug loadings (10 % to 50 % w/w) and entrapment efficiencies > 80 %. However, after 90 days, in vitro celecoxib release from nano drug embedded microparticles presented a significantly slower release in comparison to drug in solution microparticles, attributed to the presence of stabilized amorphous drug. No cytotoxicity was observed in human articular synoviocytes and PGE2 release was fully suppressed at low doses of both microparticulate systems. This study provides techniques to release high drug loads over months in a tunable manner, providing valuable options for the IA management of osteoarthritis. Keywords: celecoxib; crystalline behavior; spray-dried microparticles; intra-articular injection; drug delivery systems; osteoarthritis

Page 77: Thesis - archive-ouverte.unige.ch

Chapter III

70

1. INTRODUCTION

Osteoarthritis (OA), the most common form of arthritis, is a chronic disease with

worldwide incidence in the population aged 65 years and higher [1,2]. It is

characterized by chronic inflammation, articular cartilage degeneration and

structural changes of whole joints. Symptoms range from mild to incapacitating,

mostly in terms of pain, inflammation and loss of movement. Risk factors include

age, gender (higher prevalence in females), overweight/obesity, previous joint

injuries and genetic predisposition to joint complications. Other than the economic

burden it represents, OA is one of the main causes of disability in the elderly

population [3–6]. Since this is a multifactorial, complex disease there is an unmet

need for disease modifying drug that target multiple tissues and mechanism, slowing

disease progression. Current treatment options are mainly based on non-

pharmacological and symptoms management approaches [7,8]. Joints are poorly

irrigated closed structures, making systemic drugs a liability. In this field, intra-

articular (IA) administration plays an important role. Local delivery of small

molecules not only circumvents systemic adverse effects but also allows for a more

effective relief of symptoms. However, drawbacks include fast joint space clearance

and limited number of yearly injections [9,10]. Polymeric microparticulate drug

delivery systems have grown in the field as a response to such disadvantages,

primarily by allowing higher drug loading and prolonged delivery of drug molecules

directly into the joint [11,12]. However, attaining drug loadings high enough for

release of effective therapeutic drug concentrations over extended periods remains

a challenge. In this study, to achieve high drug loadings required to deliver

therapeutic doses in a sustained manner, spray drying was selected as the

formulation technique. Spray drying is an established technique applied in the

formulation of drug encapsulating polymeric microparticles [13]. This continuous,

single step manufacturing process is an attractive technique due to its cost-

effectiveness, upscaling potential, versatility and high reproducibility [14,15]. The

process is based on the atomization of a liquid feed, forming droplets. Aided by a

drying gas and in a closed chamber of controlled environment, the droplets

transform to dry particles by solvent evaporation. The final product is separated and

collected according to size [16,17]. In this work, this technique was applied using

poly D,L-Lactic Acid (PDLLA) as the matrix polymer. A biodegradable material, this

Page 78: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

71

polymer was selected due to its high molecular weight and its known properties of

extended drug release [18,19]. The incorporated drug, celecoxib, is a crystalline

poorly soluble drug compound commonly prescribed in the symptomatic oral

treatment of OA. A strong anti-inflammatory nonsteroidal drug, the gastrointestinal

and cardiovascular side effects of this COX-2 specific inhibitor are well established

[20–23]. With the main goal of tailoring drug release to ensure therapeutic anti-

inflammatory activity over an extended period of time, in this study, celecoxib was

encapsulated into polymeric microparticles using two different approaches. Drug in

solution and nano wet milled drug particles were incorporated by spray drying into

microparticles sized 20 to 30 µm. Three different initial drug concentrations were

tested in order to evaluate the impact of drug loading in the delivery systems.

2. MATERIALS AND METHODS

2.1. Materials

Poly (D,L-Lactic Acid) (PDLLA - Resomer 205 S®) was purchased from Evonik

(Essen, Germany). Celecoxib was purchased from LC Laboratories® (Woburn, MA,

USA). Solvents dichloromethane, acetone, trifluoroacetic acid and acetonitrile were

all of analytical grade and purchased from Sigma-Aldrich (St. Louis, MO, USA). Milli-

Q® water from Merck (Burlington, MA, USA) was used throughout this study. All

other chemical products were commercially obtained from Sigma-Aldrich (St. Louis,

MO, USA). For the in vitro anti-inflammatory assays, all media solutions were

purchased from Invitrogen (Carlsbad, CA, USA). Fetal bovine serum was purchased

from Eurobio (Les Ulis, France). Interleukin 1 beta (IL-1β) and prostaglandin PGE2

quantification assay were purchased from R&D Systems (Bio-Techne, Abingdon,

UK).

2.2. Nano wet milled drug

Commercial celecoxib was first recrystallized by a non-solvent precipitation method

and then wet-milled to nano size. Briefly, 5 mg of drug were dissolved in 300 µL of

acetone in a 2 mL tube. Water was added dropwise in a 1:2 solvent/water volume

ratio. The organic solvent was evaporated at 4 °C from open tubes, overnight. Next,

100 µL of a 2 % w/v aqueous solution D-α-tocopherol polyethylene glycol 1000

succinate (TPGS) were incorporated in the tube, as a stabilizer. Lastly, 580 mg of

Page 79: Thesis - archive-ouverte.unige.ch

Chapter III

72

BeadBug™ 0.5 mm zirconium beads from the same supplier were added to each

tube. Size reduction was performed using a homogenizer (Precellys 24®) from Bertin

Instruments (Montigny-le-Bretonneux, France). Five cycles of 90 s at 10000 rpm

followed by a 5 min ice bath were completed twice. An extra 10 cycles of 60 s at

6600 rpm and ice bath succeeded. Once separated from the zirconium beads, all

nano drug suspensions were centrifuged and freeze-dried. Size determination was

done by dynamic light scattering (DLS) using a Nanosizer (Malvern Panalytical,

Malvern, England).

2.3. Spray dried polymeric microparticles

PDLLA microparticles were formulated using a 4M8-TriX spray dryer (ProCepT,

Zelzate, Belgium). After process optimization with unloaded particles, a set of

variable and fixed process parameters was selected. These parameters were:

polymer concentration 7 % w/v; feed rate 5 mL/min; bi-fluid nozzle 0.6 mm (internal

diameter); atomizing air force 6 L/min; inlet temperature 45 °C; outlet temperature

34 °C; cyclone size medium and inlet air flow 0.3 m3/min. Air was chosen as the

processing gas. Using dichloromethane as solvent, three different proportions of

celecoxib - 10, 25 and 50 % (weight percentage of total dry content) - were loaded

into two different types of microparticles (MPs): drug in solution microparticles and

nano drug particles embedded into microparticles. In the first case, both drug and

polymer were dissolved simultaneously in dichloromethane and the resulting

solution was sprayed. The second type of microparticles was obtained after spray

drying a suspension, using the same nominal drug loadings (10, 25 and 50 %).The

nano wet milled drug powder was extemporaneously dispersed into a

dichloromethane polymer solution, in an ice bath, just before spraying.

2.4. Drug and polymeric microparticles characterization

2.4.1. Size and size distribution

Particle size determination was conducted by dynamic light scattering (DLS) or laser

light diffraction using, a Nanosizer and a Mastersizer S from Malvern Instruments

Ltd (Malvern Pananalytical, Malvern, England), respectively. All measurements

were done in triplicate.

Page 80: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

73

2.4.2. Surface, morphology and drug incorporation

Morphology and surface analysis were conducted by optical microscopy using a

Carl Zeiss Axio microscope (Oberkochen, Germany) and scanning electron

microscopy (SEM) using a Jeol JSM-7001TA microscope (Tokyo, Japan). For SEM,

all samples were deposited onto carbon tape covered metal studs. Studs were

allowed to dry overnight under vacuum. Prior to analysis, studs were sputter coated

with gold. All measurements were conducted at a 5 kV voltage. To further

understand drug incorporation into the polymeric microparticles, drug release

mechanisms and water intake during in vitro release studies, a number of particles

were cross sectionned transversally. Briefly, a small amount of powder was

dispersed into a drop of OCT embedding matrix for frozen sections (CellPath,

Newtown, England) onto a metal stub. The mounted samples were frozen at -80 °C

for 24 h. Before cutting, using a CryoStar NX70 microtome (Thermo Fisher

Scientific; Waltham, MA, United States), all stubs were allowed to acclimate for 30

min in the cooling chamber of the apparatus, at -20 °C. All cuts were directly

mounted onto silicone wafers and analyzed by SEM using the same procedure

above described.

2.4.3. Drug loading and encapsulation efficiency

Drug content of all microparticulate systems formulated was quantified by ultra-high

performance liquid chromatography (UHPLC). All samples were analyzed using a

Waters Acquity™ Ultraperformance LC (Milford, MA, USA) equipped with an Acquity

UPLC® BEH C18 column (2.1 mm x 50 mm, 1.7 µm, Waters, USA) equilibrated at

40 °C. The mobile phase was composed of acetonitrile and water to which 0.1 %

trifluoroacetic acid (TFA) was added. A 2 min solvent gradient, at a constant flow

rate of 0.5 mL/min, was used. Detection was performed with a PDA detector set at

254 nm. The volume of injected sample was 2 µL. All samples were analyzed in

triplicate. Amount of encapsulated drug (drug loading, DL %) and the encapsulation

efficiency (EE %) were calculated by the following equations:

[1] DL %= Mass of CXB in MPs

Total mass of MPs ×100

[2] EE % = DL %

Theoretical DL %

Page 81: Thesis - archive-ouverte.unige.ch

Chapter III

74

2.4.4. Differential scanning calorimetry (DSC)

The different drug forms (commercial and nano wet milled) and all formulated

microparticles were analyzed using a Mettler Toledo DSC 3 Stare System differential

scanning calorimeter (Columbus, OH, USA). Sealed 40 µL aluminum crucibles were

used for all measurements. In a first instance, in order to delete humidity and thermal

history, all polymeric samples were heated from 25 to 100 °C. A cooling cycle from

100 to -10 °C followed. In a second cycle, samples were heated from -10 to 200 °C.

The same heating/cooling rate of 10 K/min was applied throughout and all

experiments were conducted under nitrogen flow (80 mL/min). Non-polymeric

samples, like commercial celecoxib, were only submitted to the cooling and second

heating cycles, under the same conditions. Glass transition temperature (Tg),

melting temperature (Tm), enthalpy (∆H) and crystallinity rates were calculated from

the last heating cycle’s resulting curve.

2.4.5. X-ray powder diffraction (XRPD)

Amorphous celecoxib was obtained from the commercial product by quench cooling.

Briefly, 25 mg of powder were melted in a tempered glass tube. Immediately after

reaching 192 °C, the tube was immersed into liquid nitrogen for 5 min. The resulting

glassy product was stored at -20 °C. X-ray powder diffraction data were collected

on an Empyrean diffractometer (Malvern Pananalytical, Malvern, England) equipped

with Cu K1,2 radiation and focusing mirror. Samples were enclosed in a glass

capillary (0.8 mm outer diameter). Throughout data collection temperature was

controlled between room temperature and 500 K using a 700+ Cryostream cooler

(OxfordCryosystems, Oxford, England). All collected data were analyzed using

TOPAS software [24].

2.5. In vitro drug release studies

In vitro drug release studies were carried over 3 months under sink conditions in

triplicates. Celecoxib solubility was determined to be 80 mg/L in phosphate buffer

saline (PBS) containing 0.5 % w/v of polysorbate 80 (Tween 80®) and 0.025 % w/v

of sodium azide. A weighed amount of microparticles was dispersed in 50 mL of

medium. Assay was performed in closed vials to avoid evaporation of water. At

alloted time points, 1 mL was withdrawn and replaced with fresh medium.

Quantification of released celecoxib was performed by UHPLC using the same

method described above.

Page 82: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

75

2.6. In vitro anti-inflammatory bioactivity assays

2.6.1. Human fibroblast-like synoviocytes (HFLS) isolation and culture

Hip synovial membrane was collected from three male adult patients with clinical

osteoarthritis (OA) at the time of hip replacement surgery. This protocol was

conducted under the approval of the local Ethics Committee (CCER, Geneva,

Switzerland, authorization # 2017-02234) with informed and consenting patients.

Once collected, tissue samples were finely minced and digested for 3 h (37 °C, 5 %

CO2 incubation) in a 3 mg/mL collagenase IX - RPMI 1640 solution. After

centrifugation (200 g), the resuspended pellet was cultured (37 °C, 5 % CO2) in

medium containing RPMI 1640, M199 (1:1), 1 % penicillin/streptomycin (100

IU/mL:100 g/mL), 2 mM L-glutamine and 20 % fetal bovine serum. After overnight

culture, non-adherent cells were removed. At confluence, cells were trypsinized and

passaged in 75 cm2 culture flasks (Corning, NY, USA), in complete medium

containing 10 % fetal bovine serum. All cellular assays were performed from

passage 3 until 9. Experiments were conducted twice per donor (n = 6).

2.6.2. In vitro cytotoxicity assay and determination of PGE2 levels in HFLS

supernatant by enzyme-linked immunosorbent assay (ELISA)

Confluent HFLS cells were treated in a 96-well plate for 1 h. Commercial celecoxib

in a dimethyl sulfoxide (DMSO) stock solution was further diluted in complete culture

medium and incubated at 1, 10 and 30 µM (DMSO final concentration at 0.1 %,

considered vehicle). Drug in solution and nano drug embedded microparticles at 10

% DL were resuspended in complete culture medium at drug equivalent

concentrations of1, 10 and 30 µM, based on microparticle drug loading. After 1 h

incubation, cells were activated by addition of IL-1β (1 µg/mL) (R&D Systems, Bio-

Techne, Abingdon, UK) with subsequent incubation for 23 h. All supernatants were

kept at -80 °C for further testing. Remaining adherent cells were tested for viability

using the cell proliferation reagent WST-1 (Roche, Basel, Switzerland), according to

the provider instructions. Prostaglandin E2 release in cell supernatant was assessed

by a sandwich enzyme-linked immunosorbent assay (ELISA). For this, a human

PGE2 parameter assay kit from R&D Systems (ref # KGE004B; Bio-Techne,

Abingdon, UK) was used according to the manufacturer’s protocol. All samples were

diluted 3 times. Experiments were conducted twice per donor (n = 6).

Page 83: Thesis - archive-ouverte.unige.ch

Chapter III

76

2.7. Statistical analysis

Data are shown in values of mean ± standard deviation (s.d.). Statistical analysis

was performed using a two-way variance analysis by a multiple groups ANOVA

(Bonferroni multiple comparisons; GraphPad Prism 8.3 software) after confirming

normal distribution and homogeneity of variances by Shapiro-Wilk, D’Agostino-

Pearson and Anderson-Darling tests. Significance was determined at alpha level

0.05. P values represented are ****p < 0.0001. A Student’s t-test - homoscedastic,

two tails – was performed in Microsoft Office Excel for the in vitro release studies

data (p < 0.05).

3. RESULTS AND DISCUSSION

3.1. Nano wet milled drug

Commercial celecoxib (Figure 1B) was recrystallized in water/acetone mixture prior

to wet milling in order to control the homogeneity and quality of the starting material.

Purity and crystallinity of the recrystallized compound were assessed by differential

scanning calorimetry and X-ray powder diffraction. As shown in Table 1 (Figure S1),

the differential scanning calorimetric analysis of both celecoxib forms presented

comparable values of Tm and enthalpy. Both values of 160 °C and 92 J/g have been

previously reported in the literature for pure celecoxib [25,26]. Displayed in Figure

1A are the XRPD patterns of commercial and recrystallized celecoxib. The same

diffraction pattern was observed for both samples with pronounced intensity at 2θ

values of 5.4°, 10.7°, 11.0°, 13.0°, 14.9°, 16.1°, 17.9°, 19.7°, 21.5°, 22.2°, 25.4° and

29.5°. Both crystalline compounds contained the triclinic (P1) polymorph III of

celecoxib, as verified by Rietveld refinement using the model from CSD database

[20,27]. The content of the amorphous fraction of the sample was lower in the

recrystallized celecoxib as observed by the lower background in this sample (Figure

S5 and S6). These results confirmed the quality and crystallinity of the drug

compound used for the wet milling procedure. The amorphous form of celecoxib

was analyzed in parallel for comparison. DSC analysis (Table 1, Figure S2) showed

an exothermic event (112 °C, ∆H 61.6 J/g) followed by an endothermic peak (161

°C, ∆H 72.5 J/g). This sequence corresponds to recrystallization of celecoxib above

its Tg, followed by melting, consistent with literature reports [25]. It is known that the

Page 84: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

77

phase which recrystallizes is the polymorph III [27]. Upon heating, the form III

transforms into the polymorph I, which also recrystallizes upon cooling of the melted

sample, as seen in XRPD patterns of the amorphous compound at different

temperatures (Figure S7). On the XRPD pattern (Figure 1A) the amorphous state of

the compound was confirmed with no distinct peaks observed. The resulting curve

had a wide broad band centered at around 2θ value of 20°, consistent with an

amorphous compound. After wet milling, the resulting product was also

characterized. As determined by dynamic light scattering, the newly produced drug

particles (Figure 1C) were of an average size of 593 nm (Z-average; PDI < 0.39). In

Table 1, the evaluation of the endothermic event of the DSC analysis showed a Tm

with an onset of 154 °C and 34.08 J/g enthalpy. The XRPD pattern of the wet milled

sample (Figure 1A) shows the presence of the polymorph III in the recrystallized

celecoxib (Figure S7). However, a decrease in intensity of all peaks was observed

at all 2θ values including a complete loss of the sharp intensity of values 10.7°, 11.0°

and 13.0°. This result is in line with the 63 % loss of the crystalline fraction in the

sample (∆H 93.8 J/g = 100 % crystalline CXB) observed in the DSC analysis results

(∆H 34.1 J/g = 36.3 %). Related to the wet milling process and contributing to these

results is the role of TPGS added to stabilize the particles. D-α-tocopherol

polyethylene glycol 1000 succinate is an amphiphilic, highly stable molecule. This

surfactant is commonly used in the drug delivery field as a solubilizer and

permeation enhancer of poorly soluble drugs [28–30]. In this case, due to

incorporation and coating of the compound in the nano wet milled drug particles, a

solubilization effect resulted in a loss of crystallinity. Another contribution to this is a

result of the high pressure, high entropy environment of the wet nano milling

technique. This induced the celecoxib size decrease and rearranged the crystal

formation to smaller, more fragile and brittle crystals (Figure 1C). These high impact

collision energies between zirconium beads during wet milling explain the zirconium

dioxide peaks (ZrO2) observed in the XRPD pattern of the nano wet milled drug

(Figure 1A) at 2θ values of 23.5°, 28.2° and 31.5° [31].

Page 85: Thesis - archive-ouverte.unige.ch

Chapter III

78

3.2. Polymeric microparticles: formulation and characterization

3.2.1. Size, morphology and drug encapsulation

After spray drying, all dry products were immediately characterized and stored at 4

°C. Residual solvent (DCM) was below 600 ppm in all the final dry products [data

not shown] [32]. Spray drying process yields were calculated based on the weight

of dry matter sprayed (Table 2). With overall values ranging from 7 to 35 %, drug in

Table 1. Differential scanning calorimetry endothermic events for all forms of celecoxib.

Endothermic thermal event

Celecoxib sample Onset (°C) ΔH (J/g)

Commercial 161.8 92.2

Recrystallized 160.2 93.8

Amorphous 161.0 72.5

Nano wet milled 154.3 34.1

Figure 1. X-ray powder diffraction patterns of all celecoxib forms and the nano wet milled drug (A); SEM micrographs of commercial (B; scale bar: 10 µm) and nano wet milled celecoxib (C; scale bar: 10 µm).

Page 86: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

79

solution microparticles batches exhibited higher production yields when compared

to nano drug embedded microparticles. High yields at increasing drug loadings and

different feed solutions, supported the selected spray drying process parameters,

choice of polymer, solvent and suggested the overall robustness of the method. The

lower yields can be explained by the small size of batches (few mL) prepared using

the ProCept spray drying equipment – pilot plant sized. Particle size and size

distribution were determined by laser light diffraction (Table 2). The three studied

loadings (10, 25 and 50 %) of drug in solution microparticles showed sizes (D50) of

21.3, 28.8 and 34.3 µm, respectively. In size distribution, span values were ≤ 3 for

this type of particles. In parallel, as shown in Table 2, nano drug embedded

microparticles presented sizes 31.3, 32.6 and 38.3 µm corresponding to the three

values of increasing drug loading. A higher polydispersity was observed where span

values ranged from 2.2 to 4.7. A trend of a slight increase in size of the embedded

particles was observed. This could relate to numerous phenomena such as feed

viscosity, solvent evaporation, shrinkage and surface tension. An increase in size

was also visible when increasing drug loading for drug in solution microparticles,

attributed to the increased viscosity of the consequent feed solution. Solubility of

celecoxib in the polymer was assessed [Data not shown] using a film method

described in the literature [33]. No visible precipitation of drug/ formation of particles

in the dried droplets, at any drug loading (10, 25 and 50 %) was observed under

optical microscopy. In comparison to dichloromethane alone (recrystallization after

drying), the polymer acted as a solubility enhancer for CXB (55 mg/mL = 7.5 % w/w,

as measured by UHPLC titration of a saturated solution). Thus, a full solubilization

in the polymer/dichloromethane solution as well as in the dried polymer was

confirmed. In addition, drug loadings of 10, 25 and 50 % were confirmed by UHPLC

(Table 2), values higher than those reported for celecoxib microparticles formulated

by spray drying [34] and those formulated by traditional methods [35]. High

entrapment efficiencies (> 83 %) were calculated for all studied microparticles. In

order to confirm these findings, SEM analysis (Figure 2) was performed in both dry

powder form and cross sectional transversal cuts of the particles at the time of

formulation (T=0). As shown in Figure 2 (a1 to a3), all drug in solution microparticles

were spherical. Despite no evidence of irregular features, different degrees of

surface smoothness, correlated to the increase in drug loading, were observed.

Polydispersity was also observed, where particles of approximately 10 µm and

Page 87: Thesis - archive-ouverte.unige.ch

Chapter III

80

smaller were found in all analyzed batches. In analyzing the inner core of drug in

solution microparticles (Figure 2a4, a5 and a6), no evident macrostructures (drug or

polymer) were found. Indeed, what was observed were homogeneous polymer

matrix cores. Porosity was identified at 10 % DL. At 50 % drug loading the inner core

of the particles presented as dense polymer-drug spheres. In the same analysis for

nano drug embedded microparticles (Figure 2B), the same spherical shape was

observed. Likewise, size polydispersity was confirmed (Figure 2b1). However,

observations of the outer layer of this type of microparticles revealed coarser and

rougher surfaces, unrelated to drug loading and observed in all batches (Figure

2b1b2b3). As shown in Figure 2b4 to b6, the inner cores of nano drug embedded

microparticles exhibited different features from those of drug in solution MPs. The

same degree of polymer matrix porosity was found throughout the three different

drug loadings. The outer layer (Figure 2b5) was also found to be more irregular than

those of drug in solution MPs (Figure 2a5) that presented as a continuous, even

line. The main difference, however, is the visible nano drug incorporation into these

particles. Increasing amounts of embedded nano celecoxib clusters were observed,

once again directly in relation to increasing drug loadings. As shown by the

micrographs in Figure 2B, at 10 % drug loading, sparse nano drug clusters were

found located near the inner surface of the particles. At 25 % DL (Figure 2b5), the

clusters appeared both near the surface and as well the center of the polymer

particle. At the highest loading (50 %; Figure 2b6) the observed clusters were found

to be embedded in all areas: surface, center and in the outer layer of the

microstructure.

Page 88: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

81

3.2.2. DSC and XRPD

Following size and morphology studies and as for the starting materials (Table 1)

the microparticles were evaluated by differential scanning calorimetry. In Table 3 (T

= 0), the evaluation of the resulting thermograms (Figure 3C and D) is shown. For

all six formulations, the first thermal event was the glass transition. All determined

values are in proximity with the Tg value reported for the chosen polymer – 52 °C.

This result illustrates no influence of the spray drying process in polymer integrity

since process temperature (45 °C) is lower than Tg. Glass transition temperature of

celecoxib is 53.8 °C, thus overlapping with the polymer Tg making it difficult to assert

the compounds presence by DSC.

3.2.2.1. Drug in solution microparticles

In the formulation of drug in solution microparticles, commercial celecoxib was

completely solubilized in the polymer solution prior to spray drying. The analysis of

the thermograms (Figure 3C) of drug in solution microparticles showed an evident

influence of the increase in drug loading (Table 3). At 10 % drug loading no other

thermal event than glass transition of the polymer was detected. Increasing the

amount of drug in solution had an impact on the solid-state of the polymeric matrix

incorporating drug molecules. For microparticles loaded 25 % and 50 %, at 110 °C,

an exothermic event was observed. This is consistent with a spontaneous

recrystallization of solubilized celecoxib induced by temperatures higher than the

polymer’s glass transition temperature and thus promoted by the consequent

Table 2. Size and loading characterization of all microparticulate systems formulated. Values

shown are mean values (n = 3).

Nominal

drug

loading

(%)

Microparticle sample Size D50

(µm) Span

Drug loading

(% w/w)

Entrapment

efficiency (%)

Process

yield

(%)

10 Drug in solution 21.3 2.9 10 100 35

Nano drug embedded 31.3 3.8 14 95 9

25 Drug in solution 28.8 3.0 22 87 13

Nano drug embedded 32.6 2.2 18 83 7

50 Drug in solution 34.3 3.0 50 100 15

Nano drug embedded 38.3 4.7 51 96 10

Page 89: Thesis - archive-ouverte.unige.ch

Chapter III

82

polymer relaxation, in agreement with values reported in literature [36]. A second

endothermic event consistent with melting of celecoxib was found at 118.5 °C for 25

% DL and 114.0 °C for 50 % DL. The respective enthalpies were 4.8 and 34.3 J/g.

In comparison to commercial celecoxib (Table 1), the observed crystallization

enthalpies were found to be consistent with the contents of drug inside these

microparticles. A normalization of the enthalpies to the corresponding drug amounts

resulted in values 21.7 J/g (25 % DL; 24 % crystallinity) and 68.5 J/g (50 % DL; 74

% crystallinity). The increase in crystallization enthalpy was observed in direct linear

correlation with increasing drug loading percentages (r2 = 0.99). In order to confirm

these results, the same drug in solution microparticles were analyzed by XRPD

(Figure 3A). No distinct diffraction pattern was found, in comparison to polymer

alone. In fact, a superimposition of the amorphous polymer’s background was

observed for all three tested drug loadings. The polymer curve showed an

amorphous broad band centered at 2θ value of 21° and the microparticles at 18°.

Amorphous state was also confirmed by comparison to commercial celecoxib

(Figure 1A), where a specific diffraction pattern would have been observed if the

drug presented in its crystalline form. Considering that this XRPD analysis was

performed at room temperature, these results are in line with those of DSC were

spontaneous drug recrystallization, and consequent crystalline behavior, occurred

only at temperatures above those of the polymer’s and celecoxib glass transition

temperatures (110 vs 52 and 54 °C, respectively). As well, the ability of the spray

drying process to preserve the physical state of drug was confirmed.

3.2.2.2. Nano drug embedded microparticles

The formulation of nano drug embedded microparticles was based on a different

process than that of drug in solution microparticles. In this case, celecoxib was first

wet milled. Then, the nano wet milled celecoxib was suspended in a polymer

solution, immediately prior to spraying. The DSC thermograms of this type of

microparticles were different from particles obtained with a solution of drug (Figure

3C and D). As shown in Table 3, no thermal event was detected at any drug loading.

However, the incorporated nano wet milled celecoxib, though not 100 % crystalline

exhibited a crystallinity of 37 % (Table 1). An effect of drug loading was expected to

be observable, in regards to the consequent increase of crystalline material

embedded inside the microparticles. Though polymer relaxation occurred, drug

Page 90: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

83

recrystallization and subsequent endothermic event were not visible in the

thermograms, in contrast to the drug in solution microparticles. In this case, the

presence of TPGS potentially acted as a recrystallization inhibitor. After size

reduction, newly formed nano drug particles were partly coated with, partly

entrapping the compound. An XRPD analysis was also performed in these

microparticles (Figure 3B). The resulting diffraction patterns resembled those of

drug in solution microparticles (Figure 3A). The same superimposition of the

amorphous polymer’s background was found for all drug loadings. In this case, the

resulting curves presented as broad bands, similar to those of polymer and other

type of microparticles. Incorporation of nano wet milled drug was confirmed by the

presence of the specific diffraction patterns of zirconium dioxide, also found in the

diffraction pattern of the starting material (Figure 1B). However, the diffraction

pattern of the starting nano wet milled celecoxib was not detected.

Figure 2. SEM micrographs, at t = 0, of formulated microparticles and cross sectional cuts demonstrating their inner core. Drug in solution microparticles (A) and nano drug embedded microparticles (B). Scale bars: 10 µm.

Page 91: Thesis - archive-ouverte.unige.ch

Chapter III

84

* No observable melting event from T0; solid-state solution (10 % loading). ** Nano drug embedded

microparticles: No observable event from T0 to T90. *** Remaining drug loading below 10 % - no observable

peak. Behavior relates to 10 % DL drug in solution microparticles.

Table 3. Differential scanning calorimetry thermograms evaluation for all formulated

microparticles, at different time points. Time points 45 and 90 days were samples taken from the in

vitro release study (Figure 4); these values are further discussed in section 3.3.

Time (days) 0 45 90

MPs sample

Glass

transition

Tg (°C)

Melting Glass

transition

Tg (°C)

Melting Glass

transition

Tg (°C)

Melting

Onset

(°C)

ΔH

(J/g)

Onset

(°C)

ΔH

(J/g)

Onset

(°C)

ΔH

(J/g)

10 % Drug in

solution

50.9 * 49.6 * 48.2 *

25 % 53.4 118.5 4.8 47.6 *** 46.1 ***

50 % 52.7 114.0 34.3 44.1 102.6 4.7 45.9 ***

10 % Nano drug

embedded

51.8

**

50.5

**

50.3

** 25 % 51.6 50.3 50.0

50 % 51.5 49.7 49.5

Figure 3. X-ray powder diffraction patterns of drug in solution microparticles (A) and nano drug embedded microparticles (B), compared to polymer PDLLA. Differential scanning calorimetry thermograms of drug in solution microparticles (solid lines, C) and nano drug embedded microparticles (dashed lines, D) at different drug loadings.

Page 92: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

85

3.3. In vitro drug release studies

Release of celecoxib from the different formulated microparticles was assessed in

sink conditions. Over 90 days, a quantification of the released drug in the PBS

medium was performed by UHPLC. Results are given in Figure 4 as percentage of

cumulative release normalized to the initial microparticle drug load. In parallel, at

four time points chosen to represent time zero, beginning, mid- and end of the

release study, an amount of particles was removed from the media; freeze-dried

and analyzed using SEM. As previously done in the characterization, the particles

were evaluated as both dry powder form and cross sectional transversal cuts (Figure

5).

3.3.1. Drug in solution microparticles

At lowest drug loading 10 %, a cumulative release of approximately 61 % was

attained after 3 months. Microparticles with higher drug loadings reached 64 and 74

%, respectively for 50 % and 25 % DL (Figure 4, solid lines). At this particular time

point (90 days), the value of cumulative release of these microparticles, 25 % DL, is

statistically different from the other two loadings by approximately 10 %. There was

no difference between the values of loadings 10 % and 50 % at the end of the study.

No apparent burst like effect was observed in the first days of the release study, at

any drug loading. Notwithstanding, at 50 % DL, a plateau in cumulative from 30 days

was detected, in comparison to the other two loadings. In order to further understand

their release mechanism, the resulting curve were fitted using different models: zero

and first order, Higuchi, Hixson and Korsmeyer-Peppas [37]. Until 30 days, before

plateauing, 50 % DL drug in solution microparticles demonstrated the strongest fit

with the Higuchi diffusion model (r2 = 0.98). Both 10 % and 25 % DL profiles, up to

90 days, also fitted with the Higuchi diffusion model (r2 = 0.97 and r2 = 0.98,

respectively). These outcomes indicate that celecoxib loading in the polymer matrix

may be a tunable parameter for controlling drug release. Analyzing the SEM

micrographs (Figure 5A) throughout the release study, it appeared evident the

impact of drug loading in the release of celecoxib from the microparticles. For all

drug loadings, morphological surface changes, tackiness and appearance of

crevices/pores were observed as early as the one-week time point (Figure S8). This

is consistent with water uptake and a plasticization of the polymer matrix. This

impact of water uptake and its effect on the polymer is well known and has been

Page 93: Thesis - archive-ouverte.unige.ch

Chapter III

86

extensively reported in the literature, for polymeric microparticles in similar release

media [38,39].The water uptake was supported by the decrease of Tg over time, with

greater impact at the higher drug loadings (Table 3). Induced changes in the

polymer’s conformation by possible plasticization may have translated into release

of drug by diffusion transport. The 10 % DL presented, at time zero, a Tg of 50.9 °C

that decreased to 48.2 °C after 90 days. Over the same period, the observed

decrease at 25 % DL went from 53.4 to 46.1 °C and from 52.7 to 45.9 °C at 50 %

DL. This suggests that the higher the drug content, the higher the polymer matrix

susceptibility to water uptake. Drug exited the particles facilitated by the erosion

provoked by the mentioned water intake and consequent polymer relaxation.

Indeed, as time progressed, visible inner pores and tears were observed, particularly

from 45 days and at 50 % DL (Figure 5A). At 10 % drug loading, this effect was not

as noticeable. In fact, these microparticles, at 90 days resembled those of 25 and

50 % DL at 45 days. This erosion and consequent drug diffusion may happen at

later stages.

3.3.2. Nano drug embedded microparticles

Nano drug embedded microparticles showed drastically different cumulative release

profiles compared to drug in solution microparticles (Figure 4, dashed lines). After

90 days, a considerably slower release was observed. Additionally, no burst like

effect was found. At 10 % DL a 90 days cumulated drug release of 22.8 % was

reached, 27.8 and 19.2 % for 25 % and 50 % DL, respectively. A small but significant

difference was observed between 25 % DL and the other two drug loadings.

Nonetheless, after 30 days, the release from all three drug loadings remained very

slow, with a possible tendency to a plateau until the end of the study. For this reason,

only the points up to day 30 were considered for curve fitting calculations. All drug

loadings had the strongest fit with the Higuchi model, with similar slopes and r2

(0.038, 0.94; 0.043, 0.92; 0.033, 0.93 for 10, 25 and 50 % respectively). In this case,

no clear impact of drug loading was found. The fact that drug was not solubilized in

the polymer matrix, but embedded, had an influence in the release, suggesting that

its mechanism was drug particle-dependent and not relying on the matrix itself.

Indeed, water uptake was less pronounced in these microparticles. The relatively

constant values of the Tg of polymer over time (Table 3) corroborates these findings.

A decrease of approximately 1 °C was found in all drug loadings. As shown in Figure

Page 94: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

87

5B and Figure S8, the impact of water uptake in the particles morphology was also

visible in this type of microparticles. Despite visible pores and tears at the later

stages of the release, drug particles remain visible embedded in the polymer matrix.

Celecoxib is a poorly soluble drug (water solubility: 4.3 mg/L [40]) and as previously

mentioned, showed high solubility to the polymer solution (DCM). This, translates in

a higher affinity of the drug to the polymer matrix than to aqueous medium, slowing

down its release from the microparticles. Both size (nm) and the incorporation of

TPGS play a role in the reduced release rate. A similar slowing down of the release

rate upon drug nano wet milling and TPGS stabilization was observed using a p38

MAPK inhibitor [41]. As before mentioned, TPGS acts as a drug stabilizer, facilitating

crystalline formation and preserving this physical conformation thus altering the

effect of aqueous media in the drug particles. To our knowledge, the combination of

TPGS with a biodegradable polymer matrix to modulate and extend the release

kinetics has not yet been reported. Noteworthy, a long-term (6 months) complete

release (100 %) was observed for drug in solution microparticles whereas for nano

drug embedded microparticles only 55 % were attained, confirming the afore

mentioned ability to sustain the release of the later nano drug embedded

microparticles.

The overall slow release observed in both type of microparticles, could be possibly

explained by the mid to high molecular weight - 77kDa - of the polymer, as well as

its ester termination. The high solubility of drug in polymer solution could also play

a role. Celecoxib was incorporated in two different ways into the same polymer

matrix. In the presence of solubilized drug, the polymer may undergo conformation

changes making it more vulnerable to water uptake thus erosion over time that

promoted drug diffusion from the microparticles. This effect was directly correlated

to the increase of drug amounts. When drug particles were simply embedded, the

same effect was not noticeable, illustrating the polymer’s ability to sustain the

release of drug. The extended drug release observed in our microparticles herein

outperforms that of typical CXB microparticles, that release their drug content from

hours [42] up to a few weeks [43].

Page 95: Thesis - archive-ouverte.unige.ch

Chapter III

88

3.4. In vitro anti-inflammatory bioactivity

Different in vitro cellular assays were performed, in order to assess both the safety

and efficacy of the formulated drug eluting microparticulate systems. Cytotoxicity of

commercial celecoxib and both drug in solution and nano drug embedded

microparticles at three drug-equivalent concentrations (1, 10 and 30 µM) was

assessed using a WST-1 assay (Figure 6A). The primary human fibroblast-like

synoviocytes isolated from 3 donors were incubated with a suspension of

microparticles or drug directly diluted in culture medium for 24h. No cytotoxicity was

Figure 4. Celecoxib cumulative in vitro release from all formulated microparticles over 3 months. Solid lines: drug in solution microparticles; dashed lines: embedded nano drug microparticles. Symbols correspond to mean values ± S.D.; n = 3.

Figure 5. SEM micrographs of cross sectional cuts of all formulated microparticles. During the in vitro release study, microparticles were freeze-dried and cut at different time points. Drug in solution microparticles (A) and nano drug embedded microparticles (B). Scale bars: 10 µm.

Page 96: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

89

observed when compared to dead (DMSO) and alive (culture medium) controls. All

cellular viability values were > 96 %. Since no cytotoxicity was found, the three

products were tested at the same concentrations for their anti-inflammatory

bioactivity (Figure 6B). Prostaglandin E2 release was measured in the resulting

supernatants by ELISA. This prostaglandin was selected as the protein of interest

due to it being a direct product of COX-2 stimulation. Celecoxib, a specific COX-2

inhibitor, will suppress the release of this prostaglandin by blocking COX-2

expression [20,44]. Commercial celecoxib was effective in suppressing PGE2

release at all tested concentrations. Inhibitions of PGE2 release of 88, 89 and 92 %

(ns; > 0.999) were detected at 1, 10 and 30 µM, respectively (Figure S9). This

complete inhibition at all concentrations is consistent with IC50 values reported for

other human cell line, below the µM range [45]. Both drug in solution and nano drug

embedded microparticles were not effective in suppressing prostaglandin release at

1 µM (Figure 6B, Figure S9). Conversely, at 10 and 30 µM a dose response effect

was found. A 66 % statistically significant inhibition at 10 µM was calculated for both

type of particles. At 30 µM, drug in solution microparticles significantly inhibited 86

% of the PGE2 release and nano drug embedded microparticles 83 %. At this drug-

equivalent concentration, both types of microparticulate systems performed equally

to commercial drug alone in fully suppressing PGE2 release. An analysis of the in

vitro drug release study from the tested microparticles (Figure 4) might explain the

observed dose response effect. At 24 h, in the release medium, drug in solution and

nano drug embedded microparticles released 10.9 ± 0.5 % and 7.3 ± 2 % of total

amount of celecoxib, respectively. Therefore, no effect at a drug-equivalent

concentration of 1 µM is expected. Correspondingly, at 10 µM, the full amount of

drug is attained thus the effect of drug eluted from the microparticles should mimic

that of commercial celecoxib at this concentration. Indeed the inhibition of PGE2

release increased by 62 %. Nevertheless, no true correlation between the two

results is possible due to the differences of the PBS + surfactant release medium

and the complete culture medium, which may explain that a significant matching of

the effects of drug alone, at 24 h, is only observed (Figure 6) at a 3-fold concentration

increase (30 µM). This significantly comparable performance between the two types

of microparticles validates their drug loading (10 %) despite the different formulation

processes. Additionally, it confirms the release of therapeutic concentrations of

celecoxib from both types of systems, at very low amounts of microparticles in

Page 97: Thesis - archive-ouverte.unige.ch

Chapter III

90

suspension. Using these particles loaded at 10 %, in order to achieve 30 µM of drug-

equivalent concentration, a suspension of 110 µg/mL of microparticles was used. In

this case, the formulation would be performant at 0.011 % (w/v) of microparticles in

suspension. As per the literature [46,47], in a clinical setting, this is an acceptable –

even low - concentration of solid components of a formulation knee IA injection. At

higher drug loadings (25 and 50 %), this amount would further decrease, thus

augmenting the anti-inflammatory effect at low amounts of the proposed drug

delivery systems.

4. CONCLUSIONS

Local intra-articular administration of molecules is proving to be an answer to deliver

effective therapeutic doses to the joint space reducing/avoiding issues with systemic

adverse effects of said drug molecules. In this study, we have developed

microparticulate drug delivery systems encapsulating celecoxib into biocompatible

polymer by two different approaches – drug in solution and embedded nano-milled

Figure 6. In vitro cellular viability (WST- 1) of primary HFLS incubated with commercial celecoxib, drug in solution and nano drug embedded microparticles (10% DL) for 24 h. Vehicle – 0.1% DMSO (A). Release of PGE2 by IL- 1 β activated primary HFLS confirmed anti-inflammatory activity of commercial celecoxib and drug eluted from drug in solution and nano drug embedded microparticles (10% DL) (B). Bars correspond to mean values ± s.d.; n = 6; ****p < 0.0001.

Page 98: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

91

drug. An optimized spray drying process allowed for high drug loadings that did not

compromise the toxicity of both type of microparticles when in an in vitro culture with

primary human fibroblast-like synoviocytes. As so, and in the same cellular system,

effective anti-inflammatory bioactivity of eluted drug from 10 % DL microparticles

was proven by the suppression of PGE2 release. This efficacy at the lowest drug

loading tested, allows for minimal amounts of dry solid material in final injectable

formulations. Both microparticulate systems showed an extended controlled release

of celecoxib over three months. The embedding of nano wet milled drug

considerably slowed down the release of drug. As a potential OA therapeutic poorly

soluble drug delivery system, the described microparticles allow for a tailoring of

drug release profile by fine-tuning both drug loading percentages and type of

incorporation into the polymeric matrix.

5. REFERENCES

[1] M.B. Goldring, S.R. Goldring, Osteoarthritis, J. Cell. Physiol. 213 (2007) 626–634. https://doi.org/10.1002/jcp.21258. [2] A.E. Nelson, K.D. Allen, Y.M. Golightly, A.P. Goode, J.M. Jordan, A systematic review of recommendations and guidelines for the management of osteoarthritis: The Chronic Osteoarthritis Management Initiative of the U.S. Bone and Joint Initiative, Semin. Arthritis Rheum. 43 (2014) 701–712. https://doi.org/10.1016/j.semarthrit.2013.11.012. [3] A. Koszowska, R. Hawranek, J. Nowak, Osteoarthritis -a multifactorial issue, Reumatologia. 52 (2014) 319–325. https://doi.org/10.5114/reum.2014.46670. [4] M. Cross, E. Smith, D. Hoy, S. Nolte, I. Ackerman, M. Fransen, L. Bridgett, S. Williams, F. Guillemin, C.L. Hill, L.L. Laslett, G. Jones, F. Cicuttini, R. Osborne, T. Vos, R. Buchbinder, A. Woolf, L. March, The global burden of hip and knee osteoarthritis: estimates from the Global Burden of Disease 2010 study, (2014). https://doi.org/10.1136/annrheumdis-2013-204763. [5] D.J. Hunter, D. Schofield, E. Callander, The individual and socioeconomic impact of osteoarthritis, Nat. Rev. Rheumatol. 10 (2014) 437–441. https://doi.org/10.1038/nrrheum.2014.44. [6] A. Ratneswaran, J.S. Rockel, M. Kapoor, Understanding osteoarthritis pathogenesis: A multiomics system-based approach, Curr. Opin. Rheumatol. 32 (2020) 80–91. https://doi.org/10.1097/bor.0000000000000680. [7] S.L. Kolasinski, T. Neogi, M.C. Hochberg, C. Oatis, G. Guyatt, J. Block, L. Callahan, C. Copenhaver, C. Dodge, D. Felson, K. Gellar, W.F. Harvey, G. Hawker, E. Herzig, C.K. Kwoh, A.E. Nelson, J. Samuels, C. Scanzello, D. White, B. Wise, R.D. Altman, D. DiRenzo, J. Fontanarosa, G. Giradi, M. Ishimori, D. Misra, A.A. Shah, A.K. Shmagel, L.M. Thoma, M. Turgunbaev, A.S. Turner, J. Reston, 2019 American College of Rheumatology/Arthritis Foundation Guideline for the Management of Osteoarthritis of the Hand, Hip, and Knee, Arthritis Rheumatol. 72 (2020) 220–233. https://doi.org/10.1002/art.41142. [8] B. Abramoff, F.E. Caldera, Osteoarthritis: Pathology, Diagnosis, and Treatment Options, Med. Clin. North Am. 104 (2020) 293–311. https://doi.org/10.1016/j.mcna.2019.10.007. [9] L. Kou, S. Xiao, R. Sun, S. Bao, Q. Yao, R. Chen, Drug Delivery Biomaterial-engineered intra-articular drug delivery systems for osteoarthritis therapy Biomaterial-engineered intra-articular drug delivery systems for osteoarthritis therapy, (2019). https://doi.org/10.1080/10717544.2019.1660434.

Page 99: Thesis - archive-ouverte.unige.ch

Chapter III

92

[10] P. Maudens, O. Jordan, E. Allémann, Recent advances in intra-articular drug delivery systems for osteoarthritis therapy, Drug Discov. Today. 23 (2018) 1761–1775. https://doi.org/10.1016/j.drudis.2018.05.023. [11] M.L. Kang, J.Y. Ko, J.E. Kim, G. Il Im, Intra-articular delivery of kartogenin-conjugated chitosan nano/microparticles for cartilage regeneration, Biomaterials. 35 (2014) 9984–9994. https://doi.org/10.1016/j.biomaterials.2014.08.042. [12] C. Gómez-Gaete, M. Retamal, C. Chávez, P. Bustos, R. Godoy, P. Torres-Vergara, Development, characterization and in vitro evaluation of biodegradable rhein-loaded microparticles for treatment of osteoarthritis, Eur. J. Pharm. Sci. 96 (2017) 390–397. https://doi.org/10.1016/j.ejps.2016.10.010. [13] A. Paudel, Z.A. Worku, J. Meeus, S. Guns, G. Van Den Mooter, Manufacturing of solid dispersions of poorly water soluble drugs by spray drying: Formulation and process considerations, Int. J. Pharm. 453 (2013) 253–284. https://doi.org/10.1016/j.ijpharm.2012.07.015. [14] A.H. Salama, Spray drying as an advantageous strategy for enhancing pharmaceuticals bioavailability, Drug Deliv. Transl. Res. 10 (2020) 1–12. https://doi.org/10.1007/s13346-019-00648-9. [15] M. Davis, G. Walker, Recent strategies in spray drying for the enhanced bioavailability of poorly water-soluble drugs, J. Control. Release. 269 (2018) 110–127. https://doi.org/10.1016/j.jconrel.2017.11.005. [16] A. Sosnik, K.P. Seremeta, Advantages and challenges of the spray-drying technology for the production of pure drug particles and drug-loaded polymeric carriers, Adv. Colloid Interface Sci. 223 (2015) 40–54. https://doi.org/10.1016/j.cis.2015.05.003. [17] T.W. Wong, P. John, Advances in spray drying technology for nanoparticle formation, in: Handb. Nanoparticles, 2015: pp. 329–346. https://doi.org/10.1007/978-3-319-15338-4_18. [18] Y. Zhang, S. Fei, M. Yu, Y. Guo, H. He, Y. Zhang, T. Yin, H. Xu, X. Tang, Drug Development and Industrial Pharmacy Injectable sustained release PLA microparticles prepared by solvent evaporation-media milling technology Injectable sustained release PLA microparticles prepared by solvent evaporation-media milling technology, (n.d.). https://doi.org/10.1080/03639045.2018.1483382. [19] A. Soad, S. Yehia, A.-H. Adel, M.Y. Aziz, Drug Development and Industrial Pharmacy Formulation and evaluation of injectable in situ forming microparticles for sustained delivery of lornoxicam Formulation and evaluation of injectable in situ forming microparticles for sustained delivery of lornoxicam, (n.d.). https://doi.org/10.1080/03639045.2016.1241259. [20] R.V. Dev, K.S. Rekha, K. Vyas, S.B. Mohanti, P.R. Kumar, G.O. Reddy, Celecoxib, a COX-II inhibitor, CIF Access Acta Cryst. 55 (1999) 9900161. https://doi.org/10.1107/S0108270199098200. [21] D. Clemett, K.L. Goa, Celecoxib: A review of its use in osteoarthritis, rheumatoid arthritis and acute pain, Drugs. 59 (2000) 957–980. https://doi.org/10.2165/00003495-200059040-00017. [22] R. Wan, P. Li, H. Jiang, The efficacy of celecoxib for pain management of arthroscopy: A meta-analysis of randomized controlled trials, Med. (United States). 98 (2019). https://doi.org/10.1097/MD.0000000000017808. [23] S. Shin, Safety of celecoxib versus traditional nonsteroidal anti-inflammatory drugs in older patients with arthritis, J. Pain Res. 11 (2018) 3211–3219. https://doi.org/10.2147/JPR.S186000. [24] A.A. Coelho, Whole-profile structure solution from powder diffraction data using simulated annealing, J. Appl. Cryst. (2000). [25] G. Chawla, P. Gupta, R. Thilagavathi, A.K. Chakraborti, A.K. Bansal, Characterization of solid-state forms of celecoxib, Eur. J. Pharm. Sci. 20 (2003) 305–317. https://doi.org/10.1016/S0928-0987(03)00201-X. [26] V.K. Kakumanu, A.K. Bansal, Enthalpy Relaxation Studies of Celecoxib Amorphous Mixtures, 2002. [27] G.W. Lu, M. Hawley, M. Smith, B.M. Geiger, W. Pfund, Characterization of a novel polymorphic form of celecoxib, J. Pharm. Sci. 95 (2006) 305–317. https://doi.org/10.1002/jps.20522. [28] C. Yang, T. Wu, Y. Qi, Z. Zhang, Recent Advances in the Application of Vitamin E TPGS for Drug Delivery, Theranostics. 8 (2018) 464–485. https://doi.org/10.7150/thno.22711.

Page 100: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

93

[29] A. Tuomela, J. Hirvonen, L. Peltonen, A.K. Bansal, pharmaceutics Stabilizing Agents for Drug Nanocrystals: Effect on Bioavailability, Pharmaceutics. 8 (2016) 16. https://doi.org/10.3390/pharmaceutics8020016. [30] Y. Guo, J. Luo, S. Tan, B.O. Otieno, Z. Zhang, The applications of Vitamin e TPGS in drug delivery, Eur. J. Pharm. Sci. 49 (2013) 175–186. https://doi.org/10.1016/j.ejps.2013.02.006. [31] C.J. Howard, R.J. Hill, B.E. Reichert, Structures of ZrO2 polymorphs at room temperature by high‐resolution neutron powder diffraction, Acta Crystallogr. Sect. B. 44 (1988) 116–120. https://doi.org/10.1107/S0108768187010279. [32] Residual Solvents Under USP 467 (ICH Q3C) Guidelines, (n.d.). http://www.apertuspharma.com/analytical-services/residual-solvents-usp-467-ich-q3c/ (accessed March 23, 2020). [33] J. Panyam, D. William, A. Dash, D. Leslie-Pelecky, V. Labhasetwar, Solid-state solubility influences encapsulation and release of hydrophobic drugs from PLGA/PLA nanoparticles, J. Pharm. Sci. 93 (2004) 1804–1814. https://doi.org/10.1002/jps.20094. [34] F. Wan, A. Bohr, M.J. Maltesen, S. Bjerregaard, C. Foged, J. Rantanen, M. Yang, Critical solvent properties affecting the particle formation process and characteristics of celecoxib-loaded PLGA microparticles via spray-drying, Pharm. Res. 30 (2013) 1065–1076. https://doi.org/10.1007/s11095-012-0943-x. [35] M. Homar, N. Ubrich, F. El Ghazouani, J. Kristl, J. Kerč, P. Maincent, Influence of polymers on the bioavailability of microencapsulated celecoxib, J. Microencapsul. 24 (2007) 621–633. https://doi.org/10.1080/09637480701497360. [36] K. Grzybowska, M. Paluch, A. Grzybowski, Z. Wojnarowska, L. Hawelek, K. Kolodziejczyk, K.L. Ngai, Molecular Dynamics and Physical Stability of Amorphous Anti-Inflammatory Drug: Celecoxib, J. Phys. Chem. B. 114 (2010) 12792–12801. https://doi.org/10.1021/jp1040212. [37] N.A. Peppas, Formes pharmaceutiques nouvelles: Aspects technologique, biopharmaceutique et médical (New pharmaceutical formulations: Technological, biopharmaceutical and medical aspects), J. Control. Release. 4 (1986) 143–144. https://doi.org/10.1016/0168-3659(86)90049-0. [38] B. Gu, X. Sun, F. Papadimitrakopoulos, D.J. Burgess, Seeing is believing, PLGA microsphere degradation revealed in PLGA microsphere/PVA hydrogel composites, J. Control. Release. 228 (2016) 170–178. https://doi.org/10.1016/j.jconrel.2016.03.011. [39] H. Gasmi, F. Danede, J. Siepmann, F. Siepmann, Does PLGA microparticle swelling control drug release? New insight based on single particle swelling studies, J. Control. Release. 213 (2015)120–127. https://doi.org/10.1016/j.jconrel.2015.06.039. [40] US EPA; Estimation Program Interface (EPI) Suite, (2012). [41] P. Maudens, C.A. Seemayer, F. Pfefferlé, O. Jordan, E. Allémann, Nanocrystals of a potent p38 MAPK inhibitor embedded in microparticles: Therapeutic effects in inflammatory and mechanistic murine models of osteoarthritis, J. Control. Release. 276 (2018) 102–112. https://doi.org/10.1016/j.jconrel.2018.03.007. [42] D.M. Ghorab, M. Mohamed Amin, O.M. Khowessah, M. Ibrahim Tadros, Drug DeliveryColon-targeted celecoxib-loaded Eudragit ® S100-coated poly-#-caprolactone microparticles: Preparation, characterization and in vivo evaluation in rats, (2011). https://doi.org/10.3109/10717544.2011.595841. [43] S.P. Ayalasomayajula, U.B. Kompella, Subconjunctivally administered celecoxib-PLGA microparticles sustain retinal drug levels and alleviate diabetes-induced oxidative stress in a rat model, (2005). https://doi.org/10.1016/j.ejphar.2005.02.019. [44] J.Y. Park, M.H. Pillinger, S.B. Abramson, Prostaglandin E2 synthesis and secretion: The role of PGE2 synthases, Clin. Immunol. 119 (2006) 229–240. https://doi.org/10.1016/j.clim.2006.01.016. [45] T. Yoshino, A. Kimoto, S. Kobayashi, M. Noguchi, M. Fukunaga, A. Hayashi, K. Miyata, M. Sasamata, Pharmacological profile of celecoxib, a specific cyclooxygenase-2 inhibitor., Arzneimittelforschung. 55 (2005) 394–402. https://doi.org/10.1055/s-0031-1296878. [46] H.R. Schumacher, L.X. Chen, Injectable corticosteroids in treatment of arthritis of the knee, Am. J. Med. 118 (2005) 1208–1214. https://doi.org/10.1016/j.amjmed.2005.05.003.

Page 101: Thesis - archive-ouverte.unige.ch

Chapter III

94

[47] D.H. Neustadt, Intra-articular injections for osteoarthritis of the knee., Cleve. Clin. J. Med. 73 (2006) 897-898,901-904,906-911. https://doi.org/10.3949/ccjm.73.10.897.

Page 102: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

95

Supplementary material

Figure S1. Differential scanning calorimetry analysis of commercial celecoxib.

Figure S2. Differential scanning calorimetry analysis of amorphous celecoxib.

Page 103: Thesis - archive-ouverte.unige.ch

Chapter III

96

Figure S3. Differential scanning calorimetry analysis of polymer Resomer 205 S.

Figure S4. Differential scanning calorimetry analysis of the physical mixture of celecoxib (50 % DL) and polymer Resomer 205 S.

Page 104: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

97

Figure S5. Rietveld plot of commercial celecoxib.

Figure S6. Rietveld plot of celecoxib recrystallized in acetone.

Page 105: Thesis - archive-ouverte.unige.ch

Chapter III

98

Figure S7. XRPD patterns of amorphous celecoxib at different temperatures. Heating from room temperature until 200 °C followed by cooling back to room temperature.

Figure S8. SEM micrographs for surface morphology assessments. Dry powder of the sectioned (Figure 5) drug in solution microparticles (A) and nano drug embedded microparticles (B). Each time point corresponds to a sample taken from the in vitro release study (Figure 4); these results are further discussed in section 3.3.1. and 3.3.2.. Scale bar: 10 µm.

Page 106: Thesis - archive-ouverte.unige.ch

  Nano wet milled celecoxib microparticles

99

Figure S9. Inhibition percentage of PGE2 release normalized to 100 % activation, as a function of celecoxib concentration. Bars correspond to mean values ± S.D.; n = 6.

Page 107: Thesis - archive-ouverte.unige.ch

100

Page 108: Thesis - archive-ouverte.unige.ch

 

 

Chapter IV

Page 109: Thesis - archive-ouverte.unige.ch
Page 110: Thesis - archive-ouverte.unige.ch

 

103

Chapter IV

Sustained release carriers for osteoarthritis: in vitro

evaluation of an anti-inflammatory drug and a chondrogenic

drug on a 3D chondrocyte OA model

Carlota Salgado a,b, Olivier Jordan a,b and Eric Allémann a,b a School of Pharmaceutical Sciences, University of Geneva, 1 Rue Michel-Servet 1211 Geneva, Switzerland b Institute of Pharmaceutical Sciences of Western Switzerland, University of Geneva, 1 Rue Michel-Servet 1211 Geneva, Switzerland

ABSTRACT

Osteoarthritis (OA), the most common form of arthritis, causes the narrowing of the joint space and it is considered the number one cause of disability worldwide in the elder population. With no known cure, local administration of drug compounds by intra-articular (IA) injection represents an alternative to conventional oral OA treatment. In previous studies, the incorporation of nano drug particles in polymeric microparticles has proven to allow a sustained controlled release of therapeutic drug compounds for IA delivery. This chapter, divided in two sections, aimed at characterizing and comparing the incorporation as nano wet milled particles of two drug compounds – celecoxib and kartogenin. Secondly, it aimed at assessing the potential of an in vitro 3D chondrocyte OA model in the evaluation of polymeric microparticulate carriers incorporating these compounds. Both drug compounds were wet milled and incorporated with comparable results, despite different in vitro drug release profiles. Human chondrocyte pellets were successfully cultured and established as an accurate in vitro model. However, due to the sustained release nature of the evaluated carriers, clear optimization steps are imperative to fully characterize the bioactivity of the eluted drug compounds over long periods of time. Keywords: osteoarthritis; intra-articular drug delivery systems; cartilage; celecoxib; kartogenin; in vitro 3D model

Page 111: Thesis - archive-ouverte.unige.ch

Chapter IV

104

INTRODUCTION

Osteoarthritis (OA) is a chronic disease that affects whole joints. Considered the

most common form of arthritis, it is the number one cause of disability worldwide for

people above 65 years old. Often described as a “wear and tear” disease, OA

causes the narrowing of the joint space due to chronic inflammation of the synovium

and synovial fluid, leading to cartilage degradation of articular joints. In more

advanced stages, it leads to structural changes and damage like subchondral bone

remodeling and erosion with formation of osteophytes and overall loss of joint

function. Clinically, it is generally associated with pain, inflammation, and loss in

amplitude of movement [1–3]. Current treatment approaches are predominantly

based on symptom management - acetaminophen and nonsteroidal anti-

inflammatory drugs (NSAIDs) are common treatment options for pain and

inflammation – rely on several non-pharmacological measures (physical therapy)

and often lead to joint replacement surgeries. Yet, due to the complex and

multifactorial nature of OA, there is a great need for disease modifying therapies

that effectively slow disease progression [4,5]. However, the closed features of the

joint structure result in low distribution of systemic drugs thus a suboptimal efficacy

of treatment by this delivery route. Local administration, by intra-articular injection

(IA), represents an alternative to the shortcomings of oral drug administration. Not

only this is an interesting approach in terms of circumventing systemic adverse

effects commonly associated to pain and inflammation treatment options, but due to

the fact that it can be directly administered to diseased joints, IA can provide a more

sustained and prolonged delivery of drug compounds [6–8]. In this field, multiple

types of formulations like polymeric nano and microparticles, liposomes, micelles

and hydrogels, have been explored in order to answer to the different critical IA

predicaments [9]. Once in the joint, different structures and tissues can represent

the targets of these drug delivery systems. The two main components of the articular

joint are the synovium and the cartilage itself. The first is associated to the

inflammatory component of OA, and is formed by synoviocytes and different lymph

and blood vessels that produce the synovial fluid, which irrigates cartilage. This is a

connective tissue formed by extracellular matrix and chondrocytes, responsible for

absorbing friction and shock between bones during movement [10,11]. In order to

develop safer and more efficacious IA treatment options for OA, adequate in vitro,

Page 112: Thesis - archive-ouverte.unige.ch

In vitro evaluation of sustained release carriers for OA

105

ex vivo and in vivo OA models are required. Established preclinical models

encompass mouse, rat, rabbit and larger animals such as dogs and horses [12].

Different in vitro established OA models mimic partial joint structures (ex.: bovine

cartilage explants and monolayer cultured synoviocytes) making it challenging to

assess effects of IA drug delivery systems in the joint, as a whole. Currently, main

options for in vitro testing of OA include 2D and 3D cellular cultures. 2D monolayer

cultures of synoviocytes are applied to evaluate cytokine release and synovium

characterization; co-culturing different cells of different articular tissues such as

synoviocytes, chondrocytes and osteoblast is helpful in understanding crosstalk

between different tissues in response to therapies. Explants (mostly bovine and

equine) of whole joints or articular cartilage are helpful in permeation and drug

delivery studies. Cellular 3D cultures maintain cellular phenotype and cell to cell

interactions, preserving tissue connectors. As so, these are considered more

accurate and actual cell models of articular cartilage in OA than monolayer cultures

[13–15]. In this chapter, a scaffold free 3D human chondrocyte model (pellets) was

developed and applied to evaluate the anti-inflammatory and/or chondroprotective

effects of polymeric (poly-lactic acid) microparticulate drug delivery systems. The

potential of this model in assessing combination of drugs and/or different aspects of

OA is also discussed. Two different drug compounds were embedded in the

microparticles. The first drug compound, celecoxib (CXB), is a NSAID COX-2

inhibitor that to our knowledge is firstly tested here using the 3D human chondrocyte

pellet model [16]. Second drug compound, is the small molecule kartogenin (KGN),

a drug linked to articular cartilage regeneration by activation of collagen type II,

aggrecan and TIMPs. The increased expression of these important proteins stems

from the CBFβ-RUNX1 complex – formed by the translocation of CBFβ to the

nucleus, once freed from filamin A by kartogenin [17,18]. Kartogenin has been

extensively tested with positive outcomes in 3D cellular cultures using human

mesenchymal stem cells pellets but no accounts have been found describing effects

in human articular chondrocyte pellets [19].

This chapter comprises two parts: 1 – that elaborates on the spray dried polymeric

microparticles with embedded nano wet milled celecoxib and kartogenin; and 2 –

where a 3D human chondrocyte model is described and the effects of the two types

of microparticles are discussed.

Page 113: Thesis - archive-ouverte.unige.ch

Chapter IV

106

1. NANO WET MILLING AND SPRAY DRIED POLYMERIC

MICROPARTICLES

This first part describes the development of spray dried polymeric microparticles

loaded with nano wet milled celecoxib and kartogenin. Characterization and in vitro

drug release are explaining and a comparison between the two types of formulated

microparticles is provided.

1.1. Materials and methods

1.1.1. Materials

The base polymer Poly (D,L-Lactic Acid) (Resomer 205 S®) was purchased from

Evonik (Essen, Germany). Celecoxib was acquired from LC Laboratories® (Woburn,

MA, USA) and Kartogenin from Merck (Buchs, Switzerland). Analytical grade

working solvents dichloromethane, acetone, tetrahydrofuran, trifluoroacetic acid and

acetonitrile were all purchased from Sigma-Aldrich (St. Louis, MO, USA), as well as

all other chemical products. In all experimental setups, Milli-Q® water was employed

(Merck, Burlington, MA, USA).

1.1.2. Nano wet milling

In a first step, both commercial drug compounds were recrystallized using a non-

solvent precipitation method described in detail in Chapter 3 [20]. Kartogenin was

processed in a similar method, based on previous works in this lab [21]. For that, 5

mg of drug were dissolved in 300 µL of organic solvent (CXB: acetone; KGN:

tetrahydrofuran) in a 2 mL tube. Using a 1:2 solvent/water volume ratio, water was

added dropwise (600 µL). Open tubes were left overnight at 4 °C, for the evaporation

of organic solvent. Immediately before wet milling, 100 µL of D-α-tocopherol

polyethylene glycol 1000 succinate (TPGS) aqueous solution (CXB: 2 % w/v; KGN:

4 % w/v) were added to the 2 mL tube. Then, 580 mg of BeadBug™ 0.5 mm

zirconium beads from Sigma-Aldrich (St. Louis, MO, USA) were added. Wet milling

was carried out using the Precellys 24® homogenizer from Bertin Instruments

(Montigny-le-Bretonneux, France). For CXB: two times five cycles of 90 s at 10000

rpm were completed, followed by a 5 min ice bath. At these time and speed settings,

only 5 cycle repetitions are allowed. An ice bath was added in order to control

temperature increase inside the tubes. This was succeeded by 10 cycles of 60 s at

Page 114: Thesis - archive-ouverte.unige.ch

In vitro evaluation of sustained release carriers for OA

107

6600 rpm and another 5 min ice bath. For KGN: two times five cycles of 90 s at

10000 rpm were performed, with 5 min ice baths in between cycles. Then, zirconium

beads were removed and nano drug suspensions were washed three times with

water, and separated by centrifugation (10000 x g; 15 min; 4 °C). Resulting

suspensions were frozen at -80 °C, then freeze-dried overnight and stored at 4 °C

for further analysis and incorporation in polymeric microparticles.

1.1.3. Spray drying

Polymeric microparticles were formulated using a 4M8-TriX spray dryer (ProCepT,

Zelzate, Belgium). Settings for spray drying were optimized in a previous study [20]

and included: 7 % w/v polymer concentration; 5 mL/min feed rate; 0.6 mm (internal

diameter) bi-fluid nozzle; 6 L/min atomizing air force; air as the processing gas; 45

°C inlet temperature; 34 °C outlet temperature; medium cyclone and 0.3 m3/min inlet

air flow. Dichloromethane was selected as solvent for polymer solution. Immediately

before spray drying, nano wet milled drug powder amounts corresponding to 10 %

nominal drug loading (weight percentage of total dry content) were dispersed into

the polymer solution, in an ice bath. Collected dry products were further

characterized and stored at 4 °C.

1.1.4. Characterization of drug products and polymeric microparticles

1.1.4.1. Size and distribution

Particle size determination of nano milled drug products was performed by dynamic

light scattering (DLS) using a Nanosizer. Laser light diffraction was used to

determine size and distribution of formulated microparticles in a Mastersizer S. Both

equipment from Malvern Instruments Ltd (Malvern Pananalytical, Malvern,

England). All measurements were taken in triplicate.

1.1.4.2. Surface, morphology and drug incorporation

Morphology and surface analysis were analyzed by optical microscopy with a Carl

Zeiss Axio microscope (Oberkochen, Germany) and scanning electron microscopy

(SEM) using a Jeol JSM-7001TA microscope (Tokyo, Japan). In this technique, dry

samples were deposited onto carbon tape covered metal studs. Studs were allowed

to dry overnight under vacuum and were sputter coated with gold, prior to analysis

at a 5 kV voltage. To observe drug incorporation into the microparticles, a small

amount of particle powder was transversally cross-sectioned using a CryoStar NX70

Page 115: Thesis - archive-ouverte.unige.ch

Chapter IV

108

microtome (Thermo Fisher Scientific; Waltham, MA, United States). For this,

samples were prepared by dispersing a small amount of powder into a drop of OCT

embedding matrix for frozen sections (CellPath, Newtown, England) onto a metal

stub. Mounted samples were frozen at - 80 °C for 24 h. Before cutting, all stubs were

allowed to acclimate for 30 min in the cooling chamber of the apparatus, at - 20 °C.

All cuts were directly mounted onto silicon wafers and analyzed by SEM using the

same procedure above described.

1.1.4.3. Drug loading and encapsulation efficiency

Ultra-high performance liquid chromatography (UHPLC) was applied to quantify

drug content of all formulated microparticles. Analysis were performed in a Waters

Acquity™ Ultraperformance LC (Milford, MA, USA) equipped with an Acquity UPLC®

BEH C18 column (2.1 mm x 50 mm, 1.7 µm, Waters, USA) equilibrated at 40 °C.

The selected mobile phase: acetonitrile and water with added 0.1 % formic acid

(FA), was run in a 2 min gradient at 0.5 mL/min constant flow rate. Detection by a

PDA detector was set at 254 nm for samples containing CXB and 280 nm for

samples containing KGN. The volume of injected sample was 2 µL. All samples

were analyzed in triplicate. Amount of encapsulated drug (drug loading, DL %) was

determined by a ratio between the mass of drug determined in the microparticles

and the total mass of particles, transformed into percentage. Encapsulation

efficiency (EE %) was calculated by dividing calculated DL % values by the

theoretical drug loading, in this case 10 %.

1.1.5. In vitro drug release

In vitro drug release studies were carried over 2 months under sink conditions, in

triplicates.

1.1.5.1. Celecoxib (CXB)

Prior to establishing the release assay, the solubility of celecoxib was determined to

be 80 mg/L in the release medium - phosphate buffer saline (PBS) containing 0.5 %

w/v of polysorbate 80 (Tween 80®) - to ensure solubility in sink conditions - and

0.025 % w/v of sodium azide – to preserve medium over extended periods of time -

at pH 7.4. A weighed amount of microparticles with equivalent CXB amounts based

on drug loading (10 %) was dispersed in 50 mL of medium. At allotted time points,

1 mL was withdrawn and replaced with fresh medium.

Page 116: Thesis - archive-ouverte.unige.ch

In vitro evaluation of sustained release carriers for OA

109

1.1.5.2. Kartogenin (KGN)

Solubility of kartogenin (850 mg/L) was determined in the release medium -

phosphate buffer saline (PBS) at pH 7.4 containing 0.1 % w/v of sodium dodecyl

sulfate (SDS) by a saturated solution. SDS was found to be a more suitable

surfactant to ensure sink conditions of KGN throughout the study. As for celecoxib,

an amount of microparticles with equivalent KGN loading based on DL 10 %, was

dispersed in 2 mL of the buffer. At each time point, total volume was removed,

centrifuged prior to analysis and replaced by fresh conditioned medium.

In an orbital shaker (80 rpm), both assays were performed at 37 °C, in closed, amber

vials to avoid evaporation of water and to protect samples from light degradation.

Once collected at each given time point, the different aliquots were centrifuged

(10000 x g; 15 min; 4 °C) in order to ensure no polymeric particles were present.

Quantification of released drug from microparticulate systems was carried out by

UHPLC detection, using the methods above described.

1.2. Results and discussion

1.2.1. Nano wet milling

Prior to wet milling, both CXB and KGN were recrystallized by non-solvent

precipitation to ensure quality and homogeneity. Recrystallization of commercial

celecoxib in acetone resulted in large (approx. 30 µm) brittle, clear crystals, with an

acicular habit (Figure 1A). Once wet milled, these crystals were reduced to 593 nm

and gained a rod like appearance (Figure 1a2; Table 1). Commercial kartogenin

crystals had a polyhedric habit, as a thin white powder. Once wet milled, although

the shape resembled the celecoxib rod-shaped particles (Figure 1B), kartogenin had

lower mean size (350 nm) (Table 1). The wet milling processes were adapted to the

recrystallized drug products. Process optimization was performed based on crystal

appearance and general physical chemical properties of the drug compounds, such

as melting point. Precipitation solvent, TPGS concentration, temperature, time and

speed settings of tissue homogenizer were all adjusted accordingly. Due to the fact

that the recrystallized CXB presented as brittle large crystals, a 2 % (w/v) TPGS

solution was sufficient to stabilize crystal properties [22,23] thus size of the drug

particles considerably decreased by wet milling. However, with a low melting point

at 160 °C, it was important to control the temperature of the wet milling process,

Page 117: Thesis - archive-ouverte.unige.ch

Chapter IV

110

hence why ice bath dippings were introduced between cycles. Due to the lower

amount of TPGS (for KGN a 4 % (w/v) was used), which acts as a crystal stabilizer,

an extra round of 10 cycles at lower speed (6600 rpm) was added to the two main

rounds of long and fast cycles (10000 rpm). For KGN, that recrystallized in THF as

hard, white crystals, the higher melting temperature (272 °C) and the higher

concentration of TPGS solution (4 % w/v) facilitated the wet milling. In this case,

only two rounds of the high energy/high speed cycles were required to decrease

drug particles size to the nano range. Considering that individual drug properties

(solubility, melting point, and shape) are taken into account when optimizing the wet

milling process, using a tissue homogenizer for wet milling of powder products can

be widely applied to different drug compounds. As previously reported by our group

[20], despite size reduction, no changes were observed in other drug properties like

crystallinity and polymorphism. Additionally, it has convenient advantages such as

controlled speed (4000 - 10000 rpm), temperature (possibility to incorporate an ice

cooler or cool the rack externally) and duration (cycles from 5 to 90 s). The volume

of wet milled material can also be adjusted if different tubes and tube holders are

selected, our equipment allows for 24x 2 mL tubes yielding approximately 100 mg

of milled drug product in roughly 30 min of process. When compared to techniques

developed previously in our lab namely, vortexing [21,24], and other techniques

described in the literature (for CXB) like grinding mills, dispersing and high pressure

homogenizers [25,26], our method shows comparable results in less steps, with an

easier and considerably faster and cost-effective process. These advantages make

for high reproducibility between batches (Table 1) and thus, an ideal technique for

scale-up in drug product manufacturing using wet media mills, cooling settings and

zirconium beads [27].

Page 118: Thesis - archive-ouverte.unige.ch

In vitro evaluation of sustained release carriers for OA

111

Table 1. Properties of nano wet milled drug products and spray dried microparticles.

Experiments were carried out in triplicates.

Characterization

Drug compound

CXB KGN

Nano wet milled drug Size Z-average [nm] 593 ± 120 350 ± 89

PDI < 0.39

Microparticles 10 % DL

Size D50 [µm] 31.3 ± 12 30.6 ± 14.2

Span 3.8 2.2

Drug loading [% w/w] 12 ± 10 8.3 ± 7.2

Entrapment efficiency [%] 95 ± 5 83 ± 6

Cumulative drug release at 2 months [%] 32.8 ± 0.95 67.2 ± 0.13

Figure 1. Scanning electron micrographs of commercial celecoxib (a1), kartogenin (b1) and the corresponding wet milled drug products (a2 and b2). Scale bars: 10 µm (a1) and 5 µm (a2, b1 and b2).

Page 119: Thesis - archive-ouverte.unige.ch

Chapter IV

112

1.2.2. Polymeric microparticles: Celecoxib and Kartogenin

1.2.2.1. Characterization

Microparticles incorporating nano milled CXB have been extensively characterized

in Chapter 3 [20]. At 10 % drug loading, these microparticles presented as spherical

with a porous, rough outer surface when compared to unloaded microparticles (Fig.

2 a1,b1), that present a smooth surface. This difference can be linked to the

presence of drug particles in suspension, at the moment of spray drying that

changes the physical properties of the feed solution. Incorporation of drug was

confirmed visually by the presence of embedded nano drug clusters at the outer

areas of the polymer matrix (Figure 2b2). The polymer matrix of the unloaded

microparticles was porous and homogeneous, as can be observed in Fig. 2b1 in the

cross sectional cut micrograph. Drug loading was confirmed by UPLC with

entrapment efficiency of 95 % (Table 1). Polydispersity was as well observed in

these microparticles, sized 31.3 µm (Table 1). Similarly, microparticles incorporating

nano milled KGN presented as spherical with mean size around 30 µm and a slightly

porous outer surface (Table 1; Figure 2a3). The nano milled drug incorporation was

confirmed by UHPLC at 8.3 %, with EE % comparable to those of CXB

microparticles (Table 1). Additionally, embedded nano milled kartogenin was visible

(Figure 2b3). The distribution of drug particles in these microparticles was different

than those of embedded CXB where clear clusters with tendency to outer layers

were observed. In the KGN microparticles, no agglomeration tendency was

observable, with nano drug particles distributed throughout the polymer matrix. This

could be linked to the different solubility of the two drugs in the polymer solution (7

% w/v PDLLA in dichloromethane) that has an impact in droplet drying process.

Favorable drug precipitation or polymer drying will translate in different distribution

of nano drug particles within the microparticles. The nano size of the drug particles

is another potential factor influencing the distribution inside the dried microparticles.

As mentioned before, nano wet celecoxib has a difference of approximately 200 nm

in size from nano wet milled kartogenin (Table 1). The greater size of the nano drug

particles could have an impact on the dispersion process (magnetic stirring) when

suspending the particles prior to spray drying. Nonetheless, the optimized spray

drying process is once again validated by the fact that comparable microparticles -

Page 120: Thesis - archive-ouverte.unige.ch

In vitro evaluation of sustained release carriers for OA

113

in terms of size, drug incorporation and morphology – were formulated incorporating

two different drug products.

1.2.2.2. In vitro drug release

Drug release from both formulated 10 % DL microparticles was assessed in sink

conditions, in two different PBS-based media. The cumulative release from

embedded nano wet milled celecoxib has been previously described in Chapter 3,

here presented in Fig. 3, are additional time points within the 2 months’ time frame

that match those withdrawn for the KGN microparticles. CXB microparticles with 10

% drug loading showed a slow release with no burst like effect and a steady plateau

release from 30 days. At 2 months, 33 % of celecoxib was released from these

Figure 2. SEM micrographs of unloaded and drug incorporating microparticles. Microparticles with 10 % drug loading of nano wet milled CXB, KGN or unloaded (A) and cross sectional cuts (B). Scale bars: 10 µm.

Page 121: Thesis - archive-ouverte.unige.ch

Chapter IV

114

microparticles. At a later stage (data not shown), this cumulative percentage doubles

and 66 % of drug were released after 6 months. This slow and steady release was

expected and explained by the incorporation of drug particles as a suspension

(embedded), were drug solubility in the aqueous media is dependent on the drug

particle alone, and their reduced, TPGS stabilized nano size. Additionally, it was

previously demonstrated that celecoxib has high solubility in the polymer solution

[20] making the fact that the drug particles are suspended and not dissolved in the

polymer matrix, a plus in slowing down release. The same phenomena was not

observed regarding the KGN microparticles. In Figure 3, a burst release during the

first seven days can be observed where kartogenin release reaches 66 % of

cumulative release. From this point onwards, the drug release steadies and a

plateau is reached at 67 % until 2 months. Presently in the literature, there are no

extensive reports of polymeric microparticles incorporating kartogenin apart from the

ones previously developed in our laboratory [21] and kartogenin-chitosan

conjugated MPs [28]. If comparing the formulated 10 % DL kartogenin microparticles

to these systems, different hypothesis could explain this disparity in outcomes. In

the first case, the microparticles were also formulated using an ester terminated

PLA, albeit with a lower molecular weight. However, these microparticles were spray

dried using a less concentrated polymer solution in dichloromethane. Spraying of a

different concentration of polymer in an organic solvent will inherently change the

feed solution properties in terms of viscosity, which consequently alters the final dry

product in terms of drug incorporation, yield and size. Process inlet temperatures

were also different (80 °C vs 45 °C) giving different outcomes of evaporation from

sprayed droplets and shrinkage of polymer matrix. Lastly, these microparticles were

loaded at approximately 30 %, fact that could also influence drug release. In our

study, different increasing drug loadings would need to be explored in order to

comment on an effect of drug loading, drug incorporation and drug release. A

parallel between our formulated microparticles and the chitosan conjugated ones is

challenging given the fact that these were not formulated by spray drying, drug was

not incorporated as a suspension and the matrix polymer is entirely different.

Nonetheless, we can acknowledge that incorporation of commercial kartogenin with

chitosan results in a slower drug release (55 % after 50 days). Fundamentally, it is

important to adjust the polymeric drug delivery system to its desired outcome in

terms of drug delivery. In this case, it will mainly depend on the drug compounds

Page 122: Thesis - archive-ouverte.unige.ch

In vitro evaluation of sustained release carriers for OA

115

activity. In OA both a fast and a slow therapeutic onset of anti-inflammatory drugs

(celecoxib) are important, whereas a slow, steady and prolonged therapeutic

window is ideal for disease modifying chondroprotective/chondrogenic drugs (such

as kartogenin) where cartilage regeneration can take up to months.

Figure 3. Cumulative in vitro drug release from both formulated microparticles over 2 months. CXB microparticles: squares, green line; KGN microparticles: circles, blue line. Symbols correspond to mean values ± S.D.; n = 3.

Page 123: Thesis - archive-ouverte.unige.ch

Chapter IV

116

2. 3D IN VITRO MODEL OF OA: CHONDROCYTE PELLETS

The second part of this chapter focuses on the optimization of a 3D human

chondrocyte pellet model of OA and its role in the screening for anti-inflammatory

and chondroprotective effects of the spray dried microparticulate systems

incorporating CXB and KGN, described in the first part of the chapter.

2.1. Materials and Methods

2.1.1. Materials

For the cellular in vitro assays, base DMEM culture medium (#10938-025), HEPES

buffer, sodium pyruvate, DPBS-, trypsin (0.25 %) and penicillin-streptomycin

(10,000 U/mL) were purchased from Invitrogen (Gibco Life Technologies, Carlsbad,

CA, USA). Fetal bovine serum was purchased from Eurobio (Les Ulis, France) and

collagenase type II from Bioconcept (Allschwil, Switzerland). L-ascorbic acid 2

phosphate, human serum albumin, dexamethasone, ITS+ (insulin, transferrin and

selenic acid), 1,9-dimethylmethylene blue chloride, chondroitin-4-sulfate, 4′,6-

diamidino-2-phenylindole dihydrochloride and calf thymus DNA were acquired from

Sigma-Aldrich (St. Louis, MO, USA). Interleukin 1 beta (IL-1β) and prostaglandin

PGE2 quantification assay were purchased from R&D Systems (Bio-Techne,

Abingdon, UK), as well as all growth factors: transforming growth factor beta-1

(TGFβ-1), fibroblast growth factor-2 (FGF-2) and platelet-derived growth factor-BB

(PDGF-BB). The DNA/RNA extraction kit Nucleospin™ Triprep was purchased from

Macherey-Nagel (Düren, Germany) and SuperScript™ II Reverse Transcriptase

from Invitrogen (Carlsbad, CA, USA).

2.1.2. Human articular chondrocytes

Methods of isolation, monolayer and 3D culture of human articular chondrocytes

were adapted from the literature [29,30]. All experiments were conducted under

sterile environment.

2.1.2.1. Isolation and monolayer expansion

Following a protocol approved by the local Ethics Committee (CCER, Geneva,

Switzerland, authorization number 2017-02234) with informed and consenting

patients, knee cartilage was collected from the femoral and tibial condyles of three

male adults (aged between 65 and 72 years old) with clinical osteoarthritis (OA) at

Page 124: Thesis - archive-ouverte.unige.ch

In vitro evaluation of sustained release carriers for OA

117

the time of knee replacement surgery. After collection, pieces of cartilage were

scraped and minced into small pieces using scalpel blades. Then, the cartilage was

transferred to a T75 culture flask where it was digested in a type II collagenase

solution (1 mL/100 mg sample) for 24 h in an orbital shaker (37 °C, 5 % CO2

incubation). After, the solution was filtered (100 µm) and washed twice by

centrifugation (200 x g/4 min) with 20 mL PBS- with 5 % FCS. Resulting pellet was

suspended in 10 mL of complete medium: DMEM, 1 % penicillin-streptomycin (PS),

10 % FCS, 10mM HEPES buffer and 1mM sodium pyruvate. Cells were counted

and frozen at passage 0 in 10 % dimethyl sulfoxide (DMSO). For monolayer

expansion, once removed the freezing medium, cells were seeded at 150000

cells/mL in T75 culture flasks (15 mL) in growth factor medium (complete medium

with 1ng/mL, 5ng/mL and 10 ng/mL of TGFβ-1, FGF-2 and PDGF-BB, respectively)

at 37 °C/5 % CO2. Cells were expanded to 90 % confluence with twice/week media

changes and passaged at 37500 cells/mL. No assays were performed in monolayer

culture as it was exclusively performed for cellular expansion; pellets were cultured

with cells at passage 2 and 3.

2.1.2.2. 3D pellet chondrocyte culture

Using 96 V-bottom well plates, passage 2 or 3 chondrocytes were seeded at 250000

cells/well (100 µL) in chondrogenic medium (DMEM, 10 mM HEPES buffer, 1mM

sodium pyruvate, 1 % PS, 1 % ITS+, 0.1 nM L-ascorbic acid 2-phosphate, 1.25

mg/mL human serum albumin, 10-7 M dexamethasone and 10 ng/mL TGFβ-1). In

order to form pellets, the plates were centrifuged at 400 x g for 5 min and then

incubated for 2 weeks at 37°C/ 5% CO2, changing medium twice/week.

2.1.3. Cytotoxicity assay

Cytotoxicity of tested conditions was assessed, prior to biochemical and gene

expression analysis. Grown pellets were incubated for 8 days with supernatants

containing 3 concentrations of the two drug compounds - CXB and KGN - and the

corresponding 10 % DL microparticles (Table 3). IL-1β (1 ng/mL) was added to all

wells in order to induce the inflammatory environment of OA. DMSO was used as

dead control, media alone as negative control and pellets were also incubated with

vehicle (medium with 0.01% DMSO). All conditions were tested in duplicates. After

8 days, supernatants were removed and stored at - 80 °C and remaining pellets

were incubated with cell proliferation reagent WST-1 (Roche, Basel, Switzerland),

Page 125: Thesis - archive-ouverte.unige.ch

Chapter IV

118

according to the provider instructions. Experiments were conducted on the 3 donor

samples twice (n = 6) using pellets grown from two passages (3 and 4).

2.1.4. Quantification of glycosaminoglycans (GAG)

Pellets (n = 2) were digested in 500 µL of 1 mg/mL proteinase K/Tris-EDTA buffer

solution (50 and 1 mM, respectively) for 16 h, at 56 °C. Sulfated glycosaminoglycan

(GAG) quantification was performed by 1,9-dimethylmethylene blue (DMMB), a

thiazine chromotrope that induces metachromasia when bonded with GAGs [31].

Briefly, in a 24-well plate, 40 µL of sample or standard (chondroitin-4-sulfate) were

mixed with 1 mL of dye solution (35 µM of 1,9-dimethylmethylene blue chloride in a

pH 3.5 sodium formate buffer) and absorbance was read at 535 nm. Concentration

of GAG was determined from a standard curve (µg/mL). From the same digested

sample (n=2), DNA quantification was performed using a 4′,6-diamidino-2-

phenylindole dihydrochloride (DAPI) assay [32]. Using a 96-well plate, the

fluorescence of 200 µL of DAPI buffer (10 mM Tris, 10 mM EDTA, 100 mM NaCl

and 100 ng/mL DAPI at pH 7) was read at 360 nm excitation and 450 nm of

emission. To each appropriate well, 1 µL of sample or standard (calf thymus DNA)

was added, a total of 5 times, with fluorescence readings in between. From the

consequent increased fluorescence values, a fluorescence curve was created for

every sample. From these curves, the slope was retrieved and a ratio between the

standard slopes was calculated. Concentration of DNA for each sample were

determined by multiplying the resulting ratio by the standard’s concentration (20

µg/mL). Lastly, a ratio was calculated between GAG (µg) and DNA (µg) values

(GAG/DNA). Experiments were conducted twice per donor (n =6).

2.1.5. Gene expression analysis

RNA was extracted from pellets (n = 2) using an extraction kit (Nucleospin™

Triprep), following the recommended protocol. Pellets were vigorously disrupted in

the lysis buffer using a syringe equipped with a 21G needle, tris(2-carboxyethyl)

phosphine hydrochloride (TCEP) was added as a disulfide bonds reducing agent.

Reverse transcription was accomplished using a SuperScript™ II kit. Quantitative

PCR (qPCR) was performed using SYBR Green Power UP master mix and the

QuantStudio6Flex instrument and software (Applied Biosystems, Foster City, CA,

USA). Analyzed gene sequences can be found in Table 2. The relative gene

expression was calculated by the 2-ΔΔCT quantification method with GAPDH as

Page 126: Thesis - archive-ouverte.unige.ch

In vitro evaluation of sustained release carriers for OA

119

control (housekeeping gene). Experiments were conducted once per donor (n =3),

by pooling both passages.

GAPDH: Glyceraldehyde 3-phosphate dehydrogenase; COL2A1: collagen type II ; ACAN: aggrecan; COX2: cyclooxygenase-2

2.1.6. Anti-inflammatory activity assay

Collected supernatants (pooled per tested condition, per passage) from pellet

incubation were tested for prostaglandin E2 release by a sandwich enzyme-linked

immunosorbent assay (ELISA). For this, a human PGE2 parameter assay kit from

R&D Systems was used according to the manufacturer’s protocol. All samples were

diluted 3 times. Experiments were conducted once per donor (n = 3), by pooling two

passages. Pellets in media alone were used as negative control, IL-1β (1 ng/mL)

was used as positive control and vehicle (0.01% DMSO) was also tested for PGE2

release.

2.1.7. Imaging and histological analysis

Monolayer and histological cuts images were acquired using a Zeiss Axioscope

(Oberkochen, Germany) optical microscope equipped at x40 magnification. Pellets

were analyzed using a Cytation 3 cell imaging multi-mode reader with the Gen5™

data analysis software (Biotek, Winooski, VT, USA). Prior to sectioning, pellets were

mounted in cryoembedding compound MCC and frozen using the PrestoCHILL

system (Milestone Medical, Sorisole, Italy). Using a CryoStar NX70 microtome

(Thermo Fisher Scientific; Waltham, MA, United States), pellets were sectioned in 6

µm cuts. These were separately stained using hematoxylin/eosin and safranin-

O/fast green dyes.

Table 2. Human primers used in real time PCR. Gene Primer Sequence

GAPDH Forward 5’-ATGGGGAAGGTGAAGGTCG-3’ Reverse 5’-TAAAAGCAGCCCTGGTGACC-3’

COL2A1 Forward 5’-GGCAATAGCAGGTTCACGTACA-3’ Reverse 5’-CGATAACAGTCTTGCCCCACTT-3’

ACAN Forward 5’-TCGAGGACAGCGAGGCC-3’ Reverse 5’-TCGAGGGTGTAGCGTGTAGAGA-3’

COX2 Forward 5’-CCCTTGGGTGTCAAAGGTAA-3’ Reverse 5’-GCCCTCGCTTATGATCTGTC-3’

Page 127: Thesis - archive-ouverte.unige.ch

Chapter IV

120

2.1.8. Statistical analysis

Data are shown in mean values ± standard deviation (s.d.). Statistical analysis was

performed using a two-way variance analysis by a multiple groups ANOVA Tukey

multiple comparisons (GraphPad Prism 8.3 software). Significance was determined

at alpha level 0.05 and p values represented are ****p < 0.0001.

2.2. Results and discussion

2.2.1. Characterization of 3D chondrocyte model

Both the isolation and propagation of cultured articular chondrocytes from human

donors were optimized for our collected samples taking into consideration previously

published methods [29,30]. Steps like freezing conditions and seeding

concentrations were adjusted. Although from the same gender, the advanced age

of the three donors (mean: 69 ± 3.6 years) affects results in terms of yield and

chondrogenic ability, where younger donors are more prone to yield higher

quantities [33]. Additionally, heterogeneity between donors was observed in terms

of quantity and quality of cartilage sourced from the knee samples. Both knee size

and OA grade (OARSI score [34]) from mild to severe, played a role in the amount

of sample collected from condyles, which influenced the cellular yield after isolation

[35]. The more severe the OA grade and the smaller the obtained condyles the least

amount of cartilage was retrieved. In order to efficiently complete all the desired

experimental plan, reaching a balance between a high enough cellular yield (higher

passages) while maintaining the adequate phenotype is crucial. However,

continuous culture and passaging in plastic vessels leads to dedifferentiation,

characterized by morphological changes, from polygonal shaped chondrocytes to

fibroblast like chondrocytes (fibrocartilage) that are typically spindle like shaped [36].

This phenotype is linked to lower productions of collagen type II and aggrecan,

decreased growth yields and high synthesis of collagen type I fibers. While

redifferentiation into polygonal morphology is possible, the full reversal of this

phenotype is not [37,38]. In our study, monolayer chondrocytes were grown until the

second and third passage (approximately 7 days until confluence) in monolayer, in

order to avoid loss of quality and the phenotype of interest. Figure 4A and B shows

differences between first passage and second passage chondrocytes. In Figure 4B

a distinct increase of spindle like chondrocytes is visible, comparing to passage 1

Page 128: Thesis - archive-ouverte.unige.ch

In vitro evaluation of sustained release carriers for OA

121

chondrocytes in Figure 4A, more polygonal in shape. Additionally, the media

supplementation with growth factors (TGFβ-1, FGF-2 and PDGF-BB) increases

proliferation rates and redifferentiation ability of cultured cells [29]. A supplementary

gene expression analysis, to quantify collagen type II, aggrecan, collagen type I and

determine chondrocyte phenotype would have been advantageous, however, due

to low sample amounts this was not performed. After reaching confluence on the

second or third passage, chondrocytes were cultured in a non-adherent V-bottom

96 well plate in order to form 3D pellets (P3 and P4) aiming to accurately reproduce

cartilaginous tissue, scaffold free. In Figure 4C and D, the 1-2 mm spherical pellets

can be observed. In addition to growth factors (TGFβ-1), specific supplements

(ITS+) were added to keep physical integrity of the pellets. A clear increase in

cellular growth on the borders of the pellets in comparison with to matrix production

in the core region can be observed in Figure 4E, stained by hematoxylin/eosin. This

could be explained by the limited contact with supplemented media inside the core

of the pellets. Despite this uneven distribution, articular ECM can be observed and

classified as healthy by the blue hues, characteristic of high proteoglycan content

[39]. In Figure 4F, the cartilage specific safranin-O/fast green staining was applied

to sections of pellets. Here, in untreated pellets, intense red coloration (staining

proteoglycan content) representing healthy cartilage can be observed. Non-collagen

fibers and components would be stained green, not found in any section analyzed

of the different sections [39]. The grown 3D chondrocyte pellets were classified as

representations of healthy hyaline cartilage, where no inflammation induced loss of

cartilage is visible [30], thus as accurate models for the screening of anti-

inflammatory/chrondroprotective effects of compounds.

Page 129: Thesis - archive-ouverte.unige.ch

Chapter IV

122

2.2.2. Evaluation of anti-inflammatory and/or chondroprotective effects of CXB and

KGN polymeric microparticles

The anti-inflammatory and chondroprotective ability of the two types of polymeric

microparticles was assessed by incubating for 8 days both drug alone at three

different concentrations and microparticles (10 % DL) at drug-equivalent

concentrations (Table 3). Set concentrations were selected based on previous

literature reports, using 10 µM of CXB and 100 nM of KGN as working

concentrations. Three-fold higher and lower concentrations were used to test on a

range [17,40–43]. Simultaneously, IL-1β was added to all treated wells in order to

induce the inflammatory environment that triggers OA processes in cartilaginous

tissue [44]. The assessment of the OA environment of the tested pellets was

performed by gene expression quantification – Figure 5. When compared to non-

Figure 4. Characterization of 3D chondrocyte pellets. Monolayer pellets in passage 1 (A) and passage 2 (B), prior to pellet formation. 3D chondrocyte pellet (C and D). Histological staining: hematoxylin/eosin (E) and safranin-O/fast green (F) on cross sections of pellets (represented in whole at the superior left corner). Scale bars: 100 µm.

Page 130: Thesis - archive-ouverte.unige.ch

In vitro evaluation of sustained release carriers for OA

123

treated unstimulated pellets (control -), the IL-1β stimulated pellets showed a

decrease in collagen type II and aggrecan, healthy cartilage markers, and an

increase of inflammation marker, COX-2 thus validating the 3D model for

assessment of anti-inflammatory and chondroprotective activities.

*assuming similar in vitro drug release in both release buffer and chondrocyte culture medium.

Cytotoxicity of drug compounds alone and corresponding microparticles was

assessed by WST-1, in the 3D human chondrocyte pellets (Figure 6). Assessing

cellular viability in 3D pellets was deemed more pertinent than in monolayer cultured

chondrocytes given that all other assays were performed based on the three-

dimensional aggregates. After 8 days incubation, all tested conditions exhibited a

cellular viability higher than 90 % (versus dead and alive controls). Yet, and due to

Table 3. Tested drug concentrations and final drug amount after 8 days incubation.

Drug

concentration

Corresponding MPs (10 % DL)

concentration

Drug released from MPs (8 d) [Figure 3]

Estimated* released drug concentration

[µM] [mg/ml] % [µM]

CXB 3 1.1*10-2

16 0.5

10 3.8*10-2 1.6 30 11.4*10-2 4.8

[nM] [nM]

KGN 30 9.6*10-5

66 19.9

100 3.2*10-4 66.2 300 9.5*10-4 198.6

Figure 5. Gene expression levels of collagen type II, aggrecan and cyclooxygenase 2 in IL-1β stimulated pellets versus control – (unstimulated, non-treated) pellets. Values are normalized to internal controls and represented as mean values ± s.d.; n = 3, each donor with pooled samples from passages 3 and 4 pellets.

Page 131: Thesis - archive-ouverte.unige.ch

Chapter IV

124

the practicality of the assay being performed directly in the V-bottom shape wells

and this assay being mainly used to quantify mitochondrial activity of attached

monolayer cells, this test isn’t optimized to these conditions as the reagent does

easily penetrate the entire pellets. Furthermore, commercial drug compounds were

tested in solution at working concentrations, whereas microparticulate systems were

tested as suspension in the culture medium. This fact, and the afore-mentioned

enclosed plate bottom, may have contributed to accumulation of microparticles that

sediment over time - despite media changes accompanied by gentle shaking and

pipetting of the liquid contents. In an ideal setting, a transwell setup would have been

an advantage in avoiding this phenomenon.

Figure 6. In vitro cellular viability (WST-1) of 3D human chondrocyte pellets after 8 days incubation with three different concentrations of drug alone (A: celecoxib; B: kartogenin) and drug-equivalent microparticles (10 % DL). Vehicle = culture medium with 0.01% DMSO. Bars correspond to mean values ± s.d.; n = 6, each donor (3) tested at two different passages (2).

Page 132: Thesis - archive-ouverte.unige.ch

In vitro evaluation of sustained release carriers for OA

125

The anti-inflammatory activity of the CXB microparticles was assessed by

quantification of prostanglandin E2 (PGE2) release by the chondrocytes into the

culture media from the recovered supernatants (Figure 7). Celecoxib is a specific

COX-2 inhibitor, of which PGE2 is a byproduct. Albeit non-inflammatory cells, the 3D

human chondrocyte pellets are, in this study, stimulated by IL-1β and representative

of the inflammatory conditions clinically observed in OA. Moreover, different pro-

inflammatory cytokines are secreted into the cultured medium and have been

quantified previously [45]. Tested concentrations (3, 10 and 30 µM) of drug alone

exhibited significant high inhibition of prostaglandin release (70, 72 and 80 %,

respectively – Figure S1). CXB microparticles comparably and significantly inhibited

the release of PGE2 from stimulated chondrocytes (70, 73 and 81 % of inhibition at

3, 10 and 30 µM, respectively). A slight dose effect with a 10-fold increase from 3 to

30 µM was observed. The same dose effect can be found in Table 3, where the

supposed concentrations after 8 days drug release were calculated from in vitro

release results (Figure 3). Though the release medium and culture medium are not

comparable, we can infer the behavior of the microparticulate systems in aqueous

media. The concentrations – 0.5, 1.6 and 4.8 µM - correspond to 16 % drug release

from microparticles at 3, 10 and 30 µM and take into account their 10 % drug loading

and confirm the anti-inflammatory effect of CXB in this model, even at low dose.

Figure 7. Quantification of PGE2 release in culture medium by ELISA from drug alone vs CXB 10 % DL microparticles. Controls: CTRL- refers to non-stimulated, non-treated pellets; IL-1β refers to stimulated, non-treated pellets. Bars correspond to mean values ± s.d.; n=3, each donor with two passages pooled. **** p < 0.0001 (treatment vs stimulated non treated - IL-1β).

Page 133: Thesis - archive-ouverte.unige.ch

Chapter IV

126

In order to further assess the anti-inflammatory and chondroprotective activities of

the microparticles, the anabolic effects were assessed by quantification of

glycosaminoglycan (GAG) content (Figure 8). GAGs are polysaccharides related to

protein (proteoglycans) binding functions in cartilage. The three main GAGs found

are keratan sulphate, chondroitin sulphate and hyaluronic acid with healthy adult

cartilage being mainly composed of chondroitin sulphate [46,47]. Assessing GAG

concentration gives an insight on cartilage matrix production. In our study, a DMMB

assay was performed. Compared to control (stimulated, non-treated pellets) results

suggest no effect of CXB alone (Figure 8A), however, a clear increase in GAG levels

can be observed at 30 µM drug-equivalent microparticles. For KGN (Figure 8B),

drug alone presented an increase in GAG production at 100 nM yet no effect at 30

and 300 nM. Drug eluted from the microparticles, once again, in no correlation to

drug alone, showed an effect at the highest drug-equivalent concentration – 300 nM.

Additionally, a slight effect could be observed at 30 nM, however none was present

at 100 nM so dose effect cannot be estimated. Colorimetric assays are commonly

used to quantify these structures but are considered crude and unspecific, often

requiring high amounts of sample. Immunohistochemistry and mass quantification

assays can give higher quality results but require specific equipment, are costly and

time consuming [48]. The GAG increase at highest drug-equivalent concentrations

of both microparticulate systems but not the same concentration of drug alone

deserves further investigation, looking into cell/particles interactions or developing

an effective transwell model to avoid any interactions. Issues in quantification using

the DMMB method could also explain the disparity of results. As discussed above,

sample amounts play an important role in colorimetric assays of GAG detection, and

in this study the 3D pellets are of approximately 1-2 mm (Figure 4) which translates

to a minimal amount of chondrocyte sample to quantify.

Page 134: Thesis - archive-ouverte.unige.ch

In vitro evaluation of sustained release carriers for OA

127

Further analysis to assess cartilage matrix production either from an anti-

inflammatory or a chondroprotective/restorative effect of the studied drug

compounds was needed. For that, a gene expression analysis of collagen type II,

aggrecan and cyclooxygenase 2 was performed by real time PCR. Results are

detailed in Supplementary Material - Figure S2. Although the establishment of the

3D human chondrocyte pellets model was successful (Figure 5), outcomes of

cartilage production were not in line with the expected results. Indeed, collagen II

and aggrecan does not reflect any increase thus no cartilage matrix production and

COX-2 gene expression shows inexplicable results in regards to CXB alone and

corresponding microparticles - for KGN, no effect was expected as no anti-

inflammatory effect is expected from this compound. In general, no effect in blocking

gene expression of COX-2 was observed though a clear anti-inflammatory effect

was determined by immunoassay in the quantification of a product of COX-2

inflammation - PGE2 (Figure 7). This questions the assay itself, rather than the 3D

chondrocyte pellet model. Effectively, in an in vivo mice OA model both these drug

compounds have shown positive outcomes in decreasing inflammation environment

and increasing cartilaginous matrix [51]. In the present study, gene expression

analysis was performed once for each donor, by pooling the two tested passages

(P3 and P4). A higher number, of at least 3 repetitions would be needed in order to

confirm the obtained results. Additionally, though this model has been established

in other studies [30], sample size had an influence in our assay battery, where larger

pellets would have translated into higher genetic material making both biochemical

Figure 8. GAG production of 3D human chondrocyte pellets after 8 days incubation with (A) CXB and (B) KGN (drug alone and drug equivalent 10 % DL microparticles). Results are normalized to DNA content in samples. Symbols correspond to mean values ± s.d.; n=6, each donor (3) tested at two different passages (2).

Page 135: Thesis - archive-ouverte.unige.ch

Chapter IV

128

and gene expression analysis more robust. As mentioned before, the earlier and the

less chondrocytes are passaged, the lower the risk of dedifferentiation into

undesired phenotypes. In our study, pellets were cultured at both passage 3 and

passage 4, which might have negatively influenced the results in terms of phenotype

change and thus different responses to treatment. Lastly, incubation time may play

a role in the outcomes of certain treatments. Although inflammation is nearly

immediate after IL-1β stimulation, production of hyaline cartilage is a long process.

In order to avoid further dedifferentiation and cellular hypoxia at the core of the

pellets, this model in this current setup cannot withstand incubation times longer

than two weeks, which could hinder the development of quantifiable cartilage

markers and thus effect of treatments.

CONCLUSIONS

OA is a complex multifactorial chronic disease. Currently, there is an unmet need

for both disease modifying drugs to slow progression and more effective treatments

in terms of administration route. IA seems to answers these demands as it can

deliver directly to targets a high drug payload in a controlled and sustained manner.

To this extent, in this study, spray dried microparticles of approximately 30 µm have

been developed, with comparable features, incorporating both nano wet milled anti-

inflammatory drug - CXB - and a disease modifying drug - KGN. In an attempt to

accurately assess treatments, different clinically relevant in vitro models of OA have

been developed. In the current study, three-dimensional human chondrocyte pellets

have been successfully established to evaluate the anti-inflammatory and/or

chondroprotective activities of both CXB and KGN polymeric microparticles versus

drug compound alone (in solution). In our findings, a significant anti-inflammatory

effect from CXB eluting microparticles was observed in an immunoassay, at all

tested concentrations. However, this effect was not confirmed by GAG quantification

and gene expression analysis. The biological effects of KGN microparticles were not

quantifiable using the chosen setup and assays although these have been reported

in the literature for other type of 3D cellular structures. In the future, a clear

optimization and further development of the quantification assays as well as the 3D

pellet model – lower passages and increased pellet size - is imperative to adequately

infer effects of disease modifying drugs that require long incubation periods.

Page 136: Thesis - archive-ouverte.unige.ch

In vitro evaluation of sustained release carriers for OA

129

REFERENCES

[1] C.A. Emery, J.L. Whittaker, A. Mahmoudian, L.S. Lohmander, E.M. Roos, K.L. Bennell, C.M. Toomey, R.A. Reimer, D. Thompson, J.L. Ronsky, G. Kuntze, D.G. Lloyd, T. Andriacchi, M. Englund, V.B. Kraus, E. Losina, S. Bierma-Zeinstra, J. Runhaar, G. Peat, F.P. Luyten, L. Snyder-Mackler, M.A. Risberg, A. Mobasheri, A. Guermazi, D.J. Hunter, N.K. Arden, Establishing outcome measures in early knee osteoarthritis, Nat. Rev. Rheumatol. 15 (2019) 438–448. https://doi.org/10.1038/s41584-019-0237-3. [2] D.J. Hunter, S. Bierma-Zeinstra, Osteoarthritis, Lancet. 393 (2019) 1745–1759. https://doi.org/10.1016/S0140-6736(19)30417-9. [3] B. Abramoff, F.E. Caldera, Osteoarthritis: Pathology, Diagnosis, and Treatment Options, Med. Clin. North Am. 104 (2020) 293–311. https://doi.org/10.1016/j.mcna.2019.10.007. [4] R. McLaughlin, Management of chronic osteoarthritic pain, Vet. Clin. North Am. - Small Anim. Pract. 30 (2000) 933–949. https://doi.org/10.1016/S0195-5616(08)70016-0. [5] N.J. Manek, N.E. Lane, Osteoarthritis: Current Concepts in Diagnosis and Management, Am. Fam. Physician. 61 (2000) 1795–1804. [6] E. Rivera-Delgado, A. Djuhadi, C. Danda, J. Kenyon, J. Maia, A.I. Caplan, H.A. von Recum, Injectable liquid polymers extend the delivery of corticosteroids for the treatment of osteoarthritis, J. Control. Release. 284 (2018) 112–121. https://doi.org/10.1016/j.jconrel.2018.05.037. [7] I.A. Jones, R. Togashi, M.L. Wilson, N. Heckmann, C.T. Vangsness, Intra-articular treatment options for knee osteoarthritis, Nat. Rev. Rheumatol. 15 (2019) 77–90. https://doi.org/10.1038/s41584-018-0123-4. [8] X. Yang, H. Du, G. Zhai, Progress in Intra-Articular Drug Delivery Systems for Osteoarthritis, Curr. Drug Targets. 15 (2014) 888–900. https://doi.org/10.2174/1389450115666140804155830. [9] P. Maudens, O. Jordan, E. Allémann, Recent advances in intra-articular drug delivery systems for osteoarthritis therapy, Drug Discov. Today. 23 (2018) 1761–1775. https://doi.org/10.1016/j.drudis.2018.05.023. [10] S. Piluso, Y. Li, F. Abinzano, R. Levato, L.M. Teixeira, M. Karperien, J. Leijten, R. Van Weeren, J. Malda, Mimicking the Articular Joint with In Vitro Models, Trends Biotechnol. 37 (2019) 1063–1077. https://doi.org/10.1016/j.tibtech.2019.03.003. [11] S.R. Goldring, M.B. Goldring, Changes in the osteochondral unit during osteoarthritis: structure, function and cartilage–bone crosstalk, Nat. Publ. Gr. 12 (2016) 632–644. https://doi.org/10.1038/nrrheum.2016.148. [12] E.L. Kuyinu, G. Narayanan, L.S. Nair, C.T. Laurencin, Animal models of osteoarthritis: classification, update, and measurement of outcomes, J. Orthop. Surg. Res. 11 (2016) 19. https://doi.org/10.1186/s13018-016-0346-5. [13] H.J. Samvelyan, D. Hughes, C. Stevens, K.A. Staines, Models of Osteoarthritis: Relevance and New Insights, Calcif. Tissue Int. 1 (2020) 3. https://doi.org/10.1007/s00223-020-00670-x. [14] C.I. Johnson, D.J. Argyle, D.N. Clements, In vitro models for the study of osteoarthritis, Vet. J. 209 (2016) 40–49. https://doi.org/10.1016/j.tvjl.2015.07.011. [15] P.J. Cope, K. Ourradi, Y. Li, M. Sharif, Models of osteoarthritis: the good, the bad and the promising, Osteoarthr. Cartil. 27 (2019) 230–239. https://doi.org/10.1016/j.joca.2018.09.016. [16] M. Krasselt, C. Baerwald, Celecoxib for the treatment of musculoskeletal arthritis, Expert Opin. Pharmacother. 20 (2019) 1689–1702. https://doi.org/10.1080/14656566.2019.1645123. [17] K. Johnson, S. Zhu, M.S. Tremblay, J.N. Payette, J. Wang, L.C. Bouchez, S. Meeusen, A. Althage, C.Y. Cho, X. Wu, P.G. Schultz, A stem cell-based approach to cartilage repair., Science. 336 (2012) 717–21. https://doi.org/10.1126/science.1215157. [18] G. Cai, W. Liu, Y. He, J. Huang, L. Duan, J. Xiong, L. Liu, D. Wang, Recent advances in kartogenin for cartilage regeneration, J. Drug Target. 27 (2019) 28–32. https://doi.org/10.1080/1061186X.2018.1464011. [19] S. Zhang, P. Hu, T. Liu, Z. Li, Y. Huang, J. Liao, M. Rana Hamid, L. Wen, T. Wang, C. Mo, M. Alini, S. Grad, T. Wang, D. Chen, G. Zhou, Kartogenin hydrolysis product 4-aminobiphenyl

Page 137: Thesis - archive-ouverte.unige.ch

Chapter IV

130

distributes to cartilage and mediates cartilage regeneration, Theranostics. 9 (2019) 7108–7121. https://doi.org/10.7150/thno.38182. [20] C. Salgado, L. Guénée, R. Černý, E. Allémann, O. Jordan, Nano wet milled celecoxib extended release microparticles for local management of chronic inflammation, Int. J. Pharm. 589 (2020). https://doi.org/10.1016/j.ijpharm.2020.119783. [21] P. Maudens, C.A. Seemayer, C. Thauvin, C. Gabay, O. Jordan, E. Allémann, Nanocrystal-Polymer Particles: Extended Delivery Carriers for Osteoarthritis Treatment, Small. 14 (2018) 1703108. https://doi.org/10.1002/smll.201703108. [22] A. Tuomela, J. Hirvonen, L. Peltonen, A.K. Bansal, pharmaceutics Stabilizing Agents for Drug Nanocrystals: Effect on Bioavailability, Pharmaceutics. 8 (2016) 16. https://doi.org/10.3390/pharmaceutics8020016. [23] C. Yang, T. Wu, Y. Qi, Z. Zhang, Recent Advances in the Application of Vitamin E TPGS for Drug Delivery, Theranostics. 8 (2018) 464–485. https://doi.org/10.7150/thno.22711. [24] P. Maudens, C.A. Seemayer, F. Pfefferlé, O. Jordan, E. Allémann, Nanocrystals of a potent p38 MAPK inhibitor embedded in microparticles: Therapeutic effects in inflammatory and mechanistic murine models of osteoarthritis, J. Control. Release. 276 (2018) 102–112. https://doi.org/10.1016/j.jconrel.2018.03.007. [25] Z. Ding, L. Wang, Y. Xing, Y. Zhao, Z. Wang, J. Han, Enhanced Oral Bioavailability of Celecoxib Nanocrystalline Solid Dispersion based on Wet Media Milling Technique: Formulation, Optimization and In Vitro/In Vivo Evaluation, Pharmaceutics. 11 (2019) 328. https://doi.org/10.3390/pharmaceutics11070328. [26] J. He, Y. Han, G. Xu, L. Yin, M.N. Neubi, J. Zhou, Y. Ding, Preparation and evaluation of celecoxib nanosuspensions for bioavailability enhancement, RSC Adv. 7 (2017) 13053. https://doi.org/10.1039/c6ra28676c. [27] M. Malamatari, K.M.G. Taylor, S. Malamataris, D. Douroumis, K. Kachrimanis, Pharmaceutical nanocrystals: production by wet milling and applications, Drug Discov. Today. 23 (2018) 534–547. https://doi.org/10.1016/j.drudis.2018.01.016. [28] M.L. Kang, J.Y. Ko, J.E. Kim, G. Il Im, Intra-articular delivery of kartogenin-conjugated chitosan nano/microparticles for cartilage regeneration, Biomaterials. 35 (2014) 9984–9994. https://doi.org/10.1016/j.biomaterials.2014.08.042. [29] A. Barbero, I. Martin, Human articular chondrocytes culture., in: H. Hauser, M. Fussenegger (Eds.), Methods Mol. Med., 2nd ed, Human Press, 2007: pp. 237–47. http://www.ncbi.nlm.nih.gov/pubmed/18085212 (accessed November 29, 2017). [30] R. Ziadlou, A. Barbero, M. j. Stoddart, M. Wirth, Z. Li, I. Martin, X. Wang, L. Qin, M. Alini, S. Grad, Regulation of Inflammatory Response in Human Osteoarthritic Chondrocytes by Novel Herbal Small Molecules, Int. J. Mol. Sci. 20 (2019) 5745. https://doi.org/10.3390/ijms20225745. [31] C.B. Whitley, M.D. Ridnour, K.A. Draper, C.M. Dutton, J.P. Neglia, Diagnostic test for mucopolysaccharidosis. I. Direct method for quantifying excessive urinary glycosaminoglycan excretion., Clin. Chem. 35 (1989) 374–379. https://doi.org/https://doi.org/10.1093/clinchem/35.3.374. [32] C.F. Brunk, K.C. Jones, T.W. James, Assay for nanogram quantities of DNA in cellular homogenates, Anal. Biochem. 92 (1979) 497–500. https://doi.org/10.1016/0003-2697(79)90690-0. [33] A. Barbero, S. Grogan, D. Schäfer, M. Heberer, P. Mainil-Varlet, I. Martin, Age related changes in human articular chondrocyte yield, proliferation and post-expansion chondrogenic capacity, Osteoarthr. Cartil. 12 (2004) 476–484. https://doi.org/10.1016/j.joca.2004.02.010. [34] W. Waldstein, G. Perino, S.L. Gilbert, S.A. Maher, R. Windhager, F. Boettner, OARSI osteoarthritis cartilage histopathology assessment system: A biomechanical evaluation in the human knee, J. Orthop. Res. 34 (2016) 135–140. https://doi.org/10.1002/jor.23010. [35] M. Laganà, C. Arrigoni, S. Lopa, V. Sansone, L. Zagra, M. Moretti, M.T. Raimondi, Characterization of articular chondrocytes isolated from 211 osteoarthritic patients, Cell Tissue Bank. 15 (2014) 59–66. https://doi.org/10.1007/s10561-013-9371-3. [36] S.P. Grogan, X. Chen, S. Sovani, N. Taniguchi, C.W. Colwell, M.K. Lotz, D.D. D’Lima, Influence of cartilage extracellular matrix molecules on cell phenotype and neocartilage formation, Tissue Eng. - Part A. 20 (2014) 264–274. https://doi.org/10.1089/ten.tea.2012.0618.

Page 138: Thesis - archive-ouverte.unige.ch

In vitro evaluation of sustained release carriers for OA

131

[37] J. Bonaventure, N. Kadhom, L. Cohen-Solal, K.H. Ng, J. Bourguignon, C. Lasselin, P. Freisinger, Reexpression of cartilage-specific genes by dedifferentiated human articular chondrocytes cultured in alginate beads, Exp. Cell Res. 212 (1994) 97–104. https://doi.org/10.1006/excr.1994.1123. [38] K. Von Der Mark, V. Gauss, H. Von Der Mark, P. Müller, Relationship between cell shape and type of collagen synthesised as chondrocytes lose their cartilage phenotype in culture, Nature. 267 (1977) 531–532. https://doi.org/10.1038/267531a0. [39] N. Schmitz, S. Laverty, V.B. Kraus, T. Aigner, Basic methods in histopathology of joint tissues, Osteoarthr. Cartil. 18 (2010) S113–S116. https://doi.org/10.1016/j.joca.2010.05.026. [40] S.C. Mastbergen, N.W.D. Jansen, J.W.J. Bijlsma, F.P.J.G. Lafeber, Differential direct effects of cyclo-oxygenase-1/2 inhibition on proteoglycan turnover of human osteoarthritic cartilage: An in vitro study, Arthritis Res. Ther. 8 (2005) R2. https://doi.org/10.1186/ar1846. [41] S.C. Su, K. Tanimoto, Y. Tanne, R. Kunimatsu, N. Hirose, T. Mitsuyoshi, Y. Okamoto, K. Tanne, Celecoxib exerts protective effects on extracellular matrix metabolism of mandibular condylar chondrocytes under excessive mechanical stress, Osteoarthr. Cartil. 22 (2014) 845–851. https://doi.org/10.1016/j.joca.2014.03.011. [42] H. Cho, A. Walker, J. Williams, K.A. Hasty, Study of Osteoarthritis Treatment with Anti-Inflammatory Drugs: Cyclooxygenase-2 Inhibitor and Steroids, (2015). https://doi.org/10.1155/2015/595273. [43] W. Fan, J. Li, L. Yuan, J. Chen, Z. Wang, Y. Wang, C. Guo, X. Mo, Z. Yan, Intra-articular injection of kartogenin-conjugated polyurethane nanoparticles attenuates the progression of osteoarthritis, Drug Deliv. 25 (2018) 1004–1012. https://doi.org/10.1080/10717544.2018.1461279. [44] Y. Jiang, C. Hu, S. Yu, J. Yan, H. Peng, H.W. Ouyang, R.S. Tuan, Cartilage stem/progenitor cells are activated in osteoarthritis via interleukin-1β/nerve growth factor signaling, Arthritis Res. Ther. 17 (2015). https://doi.org/10.1186/S13075-015-0840-X. [45] R. Ziadlou, A. Barbero, I. Martin, X. Wang, L. Qin, M. Alini, S. Grad, Anti-Inflammatory and Chondroprotective Effects of Vanillic Acid and Epimedin C in Human Osteoarthritic Chondrocytes, Biomolecules. 10 (2020) 932. https://doi.org/10.3390/biom10060932. [46] A. Sharma, L.D. Wood, J.B. Richardson, S. Roberts, N.J. Kuiper, Glycosaminoglycan profiles of repair tissue formed following autologous chondrocyte implantation differ from control cartilage, Arthritis Res. Ther. 9 (2007) R79. https://doi.org/10.1186/ar2278. [47] H.J. Mankin, L. Lippiello, The glycosaminoglycans of normal and arthritic cartilage., J. Clin. Invest. 50 (1971) 1712–1719. https://doi.org/10.1172/JCI106660. [48] F. Kubaski, H. Osago, R.W. Mason, S. Yamaguchi, H. Kobayashi, M. Tsuchiya, T. Orii, S. Tomatsu, S. Tomatsu, Glycosaminoglycans detection methods: Applications of mass spectrometry, Mol Genet Metab. 120 (2017) 67–77. https://doi.org/10.1016/j.ymgme.2016.09.005. [49] J.H. Kim, J.E. Huh, Y.H. Baek, J.D. Lee, D.Y. Choi, D.S. Park, Effect of Phellodendron amurense in protecting human osteoarthritic cartilage and chondrocytes, J. Ethnopharmacol. 134 (2011) 234–242. https://doi.org/10.1016/j.jep.2010.12.005. [50] S.J. Wang, J.Z. Qin, T.E. Zhang, C. Xia, Intra-articular Injection of Kartogenin-Incorporated Thermogel Enhancing Osteoarthritis Treatment, Front. Chem. 7 (2019) 677. https://doi.org/10.3389/fchem.2019.00677. [51] J.Y. Kwon, S.H. Lee, H.S. Na, K.A. Jung, J.W. Choi, K.H. Cho, C.Y. Lee, S.J. Kim, S.H. Park, D.Y. Shin, M. La Cho, Kartogenin inhibits pain behavior, chondrocyte inflammation, and attenuates osteoarthritis progression in mice through induction of IL-10, Sci. Rep. 8 (2018). https://doi.org/10.1038/s41598-018-32206-7.

Page 139: Thesis - archive-ouverte.unige.ch

Chapter IV

132

Supplementary material

Figure S1. Inhibition percentage of PGE2 release normalized to 100 % stimulation with IL-1β, as a function of celecoxib concentration. Bars correspond to mean values ± s.d.; n = 3.

Figure S2. Gene expression levels of collagen type II (COLL2), aggrecan (ACAN) and cyclooxygenase 2 (COX2) in IL-1β stimulated pellets versus control (stimulated, non-treated) pellets. Values are normalized to internal controls and represented as mean values ± s.d.; n = 3, each donor with pooled samples from passages 3 and 4 pellets.

Page 140: Thesis - archive-ouverte.unige.ch

 

Conclusions and perspectives

Page 141: Thesis - archive-ouverte.unige.ch

 

Page 142: Thesis - archive-ouverte.unige.ch

Conclusions and perspectives

135

CONCLUSIONS AND PERSPECTIVES

As a leading cause of disability worldwide, osteoarthritis (OA) is a chronic disease

that poses a severe threat both for healthcare and socio-economic scenarios.

Several approaches have been explored in the search for a cure, a current unmet

need. Local delivery, intra-articular (IA) of therapeutic molecules, has risen in recent

years as an alternative to oral systemic treatment by circumventing drawbacks like

low bioavailability of molecules in joint capsules and off-target adverse effects.

This thesis aimed at taking a more in-depth look at the development of intra-articular

drug delivery systems (IA DDS) for OA treatment, in an attempt to push this research

further.

A deep understanding of OA in vitro models was deemed crucial for better design

and formulation of IA DDS for OA treatment and, as one of the aims of this thesis.

As so, in Chapter I, current OA in vitro models were extensively reviewed.

Advantages and drawbacks of 2D, 3D cellular models, and explants were explored

regarding outcome testing for different IA DDS like hydrogels, polymeric

microparticles, and nanoparticles. To this extent, IA DDS developed in the past five

years for OA treatment were listed and reviewed regarding type of formulation,

tissue target, and in vitro model applied. The different models - cell lines, type of

scaffold or explant tissue - were compiled and linked to the outcomes evaluated in

these studies. Monolayer cellular models, typically human synoviocytes or murine

macrophages, are more frequently used in screening of viability and/or bioactivity of

IA DDS carrying anti-inflammatory molecules. Conversely, for IA DDS eluting

chondroprotective molecules, studies in cartilage-mimicking models or actual tissue

like chondrocyte 3D cellular models or explants are preferred. Still, as IA DDS

represent an appealing option for OA treatment, and no gold standard is set for OA

in vitro models, further research efforts should focus on better tailoring models to

suit outcome evaluation of this type of formulation.

Knowledge gathered in in vitro OA models allowed the establishment of model and

outcome evaluation applied in Chapter II. In this study, the viability and anti-

inflammatory activity of A. brachypoda root extracts, a shrub used in traditional

Page 143: Thesis - archive-ouverte.unige.ch

Conclusions and perspectives

136

Brazilian medicine, were evaluated in human fibroblast-like synoviocytes (HFLS).

The main goal of this first study was to assess the potential of two root extracts, one

hydroethanolic and the other in dichloromethane, and three compounds isolated

from the latter, as IA deliverable molecules. Another goal was to deepen the

understanding of the use of this plant in OA settings. The last task was to isolate

and quantify the three compounds in the dichloromethane extract. The

hydroethanolic extract was found to be safe but not active in terms of suppressing

inflammation in HFLS. Conversely, and due to the presence of the three

compounds, the dichloromethane extract showed increased bioactivity against OA

HFLS. This anti-inflammatory activity, however, was accompanied by an increase in

cytotoxicity. It was determined that the most abundant in the dichloromethane

extract was Compound 3, followed by Compound 2 and Compound 1. Their anti-

inflammatory strength and cytotoxicity were found to be directly correlated to their

amounts in the extract. Overall, the therapeutic window of the isolated compounds

was considered too narrow for a local administration in OA treatment. Nonetheless,

and seemingly due to the protective effects of the mixture of these isolated

compounds, the dichloromethane extract remains attractive for the development of

IA formulations.

After the investigation of potential IA drug delivery molecules in studies of natural

products, research efforts were shifted in the direction of known drug molecules

used in OA treatment. Celecoxib is a non-steroidal anti-inflammatory drug (NSAID),

frequently taken prescribed for symptomatic management of pain and inflammation

in OA. Chapter III describes the formulation and characterization of celecoxib

polymeric microparticles for IA treatment of OA. In this study, celecoxib was

incorporated either as drug in solution or as nano wet-milled drug particles, using a

spray-drying method. The main goal was to assess the effects on drug release

profiles of drug incorporation as nano wet-milled particles vs. drug in solution, and

of increasing drug loadings. Resulting microparticles were characterized in terms of

size, morphology, drug incorporation, crystallinity, and their 90 days celecoxib

release profiles were compared. The microparticles were tested for viability and anti-

inflammatory activity against free drug, using the same in vitro OA model described

in the previous chapter. The combination of high drug loading and incorporation of

celecoxib as nano-milled drug in polylactic acid (PLA) translated into 30 µm

Page 144: Thesis - archive-ouverte.unige.ch

Conclusions and perspectives

137

microparticles with a long-lasting steady drug release profile. In the local treatment

of OA (IA), it is advantageous for a drug delivery system to have a sustained,

controlled drug release profile to ensure a therapeutic window over long periods of

time, thus avoiding repeated IA injections. However, it is also viewed as favorable

to have an early burst-like release of drug from the delivery systems, to achieve an

almost instant relief of symptoms like pain and inflammation. The developed

celecoxib microparticles cover both these desirable features of an IA drug delivery

system as their tunable drug loading, both in type and concentration, enables

tailoring of the drug release profile. This sustained release, confirmed by the HFLS

in vitro OA model, could also be interesting to explore using other OA molecules

with chondroprotective effects. This developed IA DDS is highly versatile as different

microparticles, with different drug loadings, of molecules in different forms and even

different molecules, with different drug release profiles could be combined in a single

formulation.

The potential of this IA DDS in incorporating other molecules and as a combination

treatment of OA is further reported in Chapter IV. In this part of the thesis, the wet-

milling and spray-drying methods from the previous chapter were used to

incorporate kartogenin. This molecule induces chondrogenesis, thus cartilage

repair, making it an attractive therapeutic molecule for OA. PLA microparticles with

embedded nano-milled kartogenin were characterized and compared to celecoxib

microparticles with the same drug loading (10 %). Comparable results in terms of

size, morphology and drug incorporation, between the tested microparticles

contributed to the versatility, robustness, and reproducibility of the developed

formulation methods. Considering the mode of action of kartogenin and knowledge

gathered in Chapter I, in this study, the drug delivery systems were tested for

viability and bioactivity in a 3D chondrocyte OA model. Pellet characterization

allowed for confirmation of the selected model as an accurate OA in vitro model.

Based on this belief, cytotoxicity was assessed over eight days, as the

microparticles exhibited a sustained controlled release over a long period of time.

During the same one-week incubation period, the bioactivity of both drugs eluted

from the microparticles systems was evaluated. The anti-inflammatory effects of

celecoxib and chondrogenic activity of kartogenin, were assessed by ELISA, gene

expression analysis, and glycosaminoglycan (GAG) content. However, the chosen

Page 145: Thesis - archive-ouverte.unige.ch

Conclusions and perspectives

138

model was not capable to fully characterize the bioactivity of the established drug

delivery systems.

The different studies presented in this thesis allowed to face and tried to reply to a

number of challenges encountered in research of IA DDS for OA treatment. Firstly,

the potential of therapeutic molecules for local delivery. Natural products are rich

sources of therapeutic molecules; potentially disease-modifying osteoarthritis drugs

(DMOADs) and molecular pathways deserve further exploration by applying

accurate OA pre-clinical models. Then, the development of polymeric carrier drug

delivery systems and lastly, the establishment of accurate OA in vitro models for

outcome evaluation of this type of formulations. Polymeric microparticles with nano

wet-milled drug particles and tailorable drug loadings seem to answer to more than

one demand of IA OA treatment. For improved design and development of these IA

DDS, better and more predictable OA in vitro models need to be explored, with

particular regards to features, like long-lasting drug release, of these formulations.

All the explored angles of IA DDS research deserve, and will benefit, from further

improvement with the ultimate goal of better OA treatment options.

Page 146: Thesis - archive-ouverte.unige.ch

 

French summary

Page 147: Thesis - archive-ouverte.unige.ch
Page 148: Thesis - archive-ouverte.unige.ch

French summary

141

FRENCH SUMMARY

L'arthrose est une maladie chronique et dégénérative. Cette forme d'arthrite la plus

courante a une prévalence élevée et toujours croissante dans la population

mondiale vieillissante (> 65 ans). Étant l'une des principales causes d'invalidité chez

les personnes âgées, la charge socio-économique causée par l'arthrose est

considérable Actuellement, les traitements pharmacologiques sont basés sur la

gestion des symptômes par des analgésiques et des anti-inflammatoires oraux.

Cependant, il existe toujours un besoin pour un traitement curatif, soit utilisant des

formulations ou voies d'administration plus efficaces, soit par des médicaments

modificateurs de la maladie (DMOAD) qui aident à ralentir sa progression ou même

inverser son cours. Ces dernières années, des efforts importants ont été déployés

dans la recherche et le développement de nouvelles molécules et de meilleurs

systèmes d'administration des médicaments. Parmi ces études, l'administration

locale intra-articulaire (IA) de molécules thérapeutiques a été explorée comme une

alternative pour contourner les problèmes de la thérapie orale systémique, tels que

la faible biodisponibilité articulaire et les effets indésirables. Une libération contrôlée

et prolongée des molécules directement dans la capsule articulaire a été envisagée

avec le développement de systèmes de délivrance de médicaments. Dans la

recherche sur les DMOADs et les systèmes d'administration de médicaments par

voie IA, les essais précliniques constituent une étape essentielle et cruciale. Dans

cette phase de développement, il est crucial d'envisager des modèles d'arthrose in

vitro et in vivo subtiles et hautement prédictifs pour une meilleure translation vers

une application clinique. De meilleures conception et prédiction pour une application

chez l'homme sont essentielles pour faire progresser la recherche sur l'arthrose.

Cette thèse s'articule autour de trois axes différents : recherche de nouvelles

molécules ayant un potentiel en tant que DMOADs; développement de systèmes

d'administration de médicaments IA (IA DDS) pour la délivrance locale de

traitements contre l'arthrose et finalement, conception et établissement de modèles

in vitro précis d'arthrose. L'objectif principal était le développement d'options de

traitement de l'arthrose par l'exploration de nouvelles molécules et la mise au point

de systèmes d'administration de médicaments appropriés. Tout au long de ces

Page 149: Thesis - archive-ouverte.unige.ch

French summary

142

travaux, un autre objectif était de modéliser précisément l'arthrose in vitro, avec

l'établissement de modèles cellulaires et une évaluation adéquate des résultats.

Les différentes études présentées dans cette thèse ont tenté de répondre à un

certain nombre de défis rencontrés dans la recherche des systèmes de libération

intra-articulaires pour le traitement de l'arthrose. Tout d'abord, concernant le

potentiel de molécules thérapeutiques pour l'administration locale. Les produits

naturels sont de riches sources de molécules thérapeutiques ; les médicaments

contre l'arthrose potentiellement modificateurs de la maladie (DMOAD) et les voies

moléculaires méritent d'être explorées plus avant en appliquant des modèles

précliniques précis de l'arthrose. Les études ont porté dans un deuxième temps sur

le développement de systèmes d'administration de médicaments par des vecteurs

polymériques et enfin, sur l'établissement de modèles in vitro de l'arthrose

pertinents, pour l'évaluation des résultats de ce type de formulations. Les

microparticules polymères avec des nanoparticules de médicament broyées par

broyage en milieu humide (wet-milling) et des taux de chargement de médicament

élevé, semblent répondre à plus d'une demande de traitement de l'arthrose par IA.

Pour améliorer la conception et le développement de ces systèmes de libération

intra-articulaires, il est nécessaire d'explorer des modèles in vitro d'arthrose

améliorés et plus prédictifs, en accordant une attention particulière aux

caractéristiques de ces formulations, comme la libération prolongée du

médicament.

Tous ces aspects de la recherche sur les systèmes de libération intra-articulaires

méritent et bénéficieront d'améliorations futures dans le but ultime d'offrir des

options efficaces de traitement permettant de modifier de façon significative

l’évolution des patients souffrant de l'arthrose.

Page 150: Thesis - archive-ouverte.unige.ch

143

Acknowledgements

Firstly, I would like to thank Prof. Eric Allémann for allowing me to join his lab and

mostly, the osteoarthritis group. Thank you for let me engage and follow this PhD

journey. Thank you for your help, guidance and endless discussions. For putting up

with my very opinionated and, at times, stubborn self and my never-ending

comments on all lab matters. A big thank you to Dr. Olivier Jordan, for the same

reasons. For your always poignant and important comments and reviews of

manuscripts, endless hours of discussion, help and guidance. Mostly for putting up

with my non-mathematical brain and solving all my statistical dilemmas and the

goiabada cascão, a must! A special thanks to Profs. Norbert Lange and Gerrit

Borchard for always being available to lend me an ear and discussions on my future

and career prospects.

My warmest thank you to everyone with whom I have collaborated in the making of

this thesis. The first and most important, Dr. Laurence Neff. Thank you Laurence for

your dedication, patience, time and endless help with all matters of biology and

cellular work. This thesis would have not been possible without your valuable help.

Thank you to Dr. Emerson Ferreira Queiroz for always being available to help,

pushing our research efforts further and our always interesting discussions. Thank

you to Nayara, Hugo and Thanise for their valuable contributions, help and good

spirits. To Prof. Jean-Luc Wolfender for the nice discussions and Dr. Laurence

Marcourt for all the help and reviews of the manuscript. A very big thank you to Prof.

Radovan Černý and Dr. Laure Guénée for their valuable contributions, many great

discussions and never ending availability. Thank you to Luca, for his assistance. It

has been a true pleasure working with you all! In this category, my biggest thank

you to my two master students, Alexandre and Benjamin. You have taught me more

than you could imagine and I am thankful to have you as good friends. And a big

thank you to Archie, without whom this manuscript would look less colorful and oh

so drab.

They say it takes a village, well in this case, it very well did. Thank you to the FATEC-

FABIO teams for all the great moments. For allowing my party planner, costume

obsessed persona to soar, always saying yes to my crazy ideas. Nathalie and

Page 151: Thesis - archive-ouverte.unige.ch

144

Marco, two people who have been crucial and the pillars of this village. Thank you

Nathalie for all your help and endless hours of therapy sessions at the SEM. For

always being available to listen to my many many dilemmas both scientific and

personal. And mostly, for your good mood and positive energies. Marco Paulo, thank

you. For always, without fail, being there for me. In and out of lab matters, you made

it possible. My biggest thank you to past lab members who have guided and helped

me in this long hard PhD journey: Imène, Yanna, Jordan, Viorica, Floriane and

Tiziana. Thank you to my dear friends, Karolina and Vassily. My day ones, here’s to

many more memories together! An immense, deep thank you to all my work wives,

who have literally carried me through these past 5 years and to whom I own

everything: Thais, Bettina, Sara, Ester, Kenza and Cíntia. Thank you to some very

important people who have been more than a support system in this lab: Souad,

Joel, Allegra, Adèle, Martin (my successor, big hopes!) and Franck. May all our

futures shine bright! Thank you to some other elements of this village, always ready

to lend a helping hand: Aymeric, Micaela, Julia, Marta, Hisham and Barbara. Thank

you to George and Ghali, you keep me forever young and always on my toes.

Thank you to my friends, both in Geneva and back home. You have kept me sane

and motivated to always carry on: Sara, Jasmin, Karin, Arjan, Vanessa, Ana, Pedro,

André, Zé Diogo, Diogo, Inês, Joana, Rebekah, Teresa, Filipa, Francisca and many

more not named here.

My biggest thank you goes out to my family. To my GVA family, without whom none

of this would have been possible, Tia Matu and Tio Zé, my cousins João, Mariana,

Zeca and Matilde. To all my cousins, may we always be this united. To my

grandparents, thank you for keeping me rooted and always showing me the way.

Especially to Avô João, to whom this thesis is dedicated to. Last but never ever

least, the other 5 pieces of my heart. Thank you to my brothers Zé Maria, Henrique

and Francisco who are my pride and joy, keep me on my toes, are there for me

through thick and thin and until the end of times. To my parents, my biggest

examples in life. Thank you for everything, no words can express how thankful I am

for all you have given me and just wish one day to be able to give back. This thesis

is as much yours as it is mine. Obrigada!

Page 152: Thesis - archive-ouverte.unige.ch