Next Article in Journal
Investigation of Ifosfamide Toxicity Induces Common Upstream Regulator in Liver and Kidney
Next Article in Special Issue
Substituted Syndecan-2-Derived Mimetic Peptides Show Improved Antitumor Activity over the Parent Syndecan-2-Derived Peptide
Previous Article in Journal
A Possible Role for HMG-CoA Reductase Inhibitors and Its Association with HMGCR Genetic Variation in Parkinson’s Disease
Previous Article in Special Issue
Decreased Levels of Microfibril-Associated Glycoprotein (MAGP)-1 in Patients with Colon Cancer and Obesity Are Associated with Changes in Extracellular Matrix Remodelling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

3D Cell Culture Systems: Tumor Application, Advantages, and Disadvantages

by
Ola Habanjar
1,
Mona Diab-Assaf
2,
Florence Caldefie-Chezet
1 and
Laetitia Delort
1,*
1
Université Clermont-Auvergne, INRAE, UNH, Unité de Nutrition Humaine, CRNH-Auvergne, 63000 Clermont-Ferrand, France
2
Equipe Tumorigénèse Pharmacologie Moléculaire et Anticancéreuse, Faculté des Sciences II, Université Libanaise Fanar, Beyrouth 1500, Liban
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(22), 12200; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms222212200
Submission received: 27 October 2021 / Revised: 5 November 2021 / Accepted: 7 November 2021 / Published: 11 November 2021

Abstract

:
The traditional two-dimensional (2D) in vitro cell culture system (on a flat support) has long been used in cancer research. However, this system cannot be fully translated into clinical trials to ideally represent physiological conditions. This culture cannot mimic the natural tumor microenvironment due to the lack of cellular communication (cell-cell) and interaction (cell-cell and cell-matrix). To overcome these limitations, three-dimensional (3D) culture systems are increasingly developed in research and have become essential for tumor research, tissue engineering, and basic biology research. 3D culture has received much attention in the field of biomedicine due to its ability to mimic tissue structure and function. The 3D matrix presents a highly dynamic framework where its components are deposited, degraded, or modified to delineate functions and provide a platform where cells attach to perform their specific functions, including adhesion, proliferation, communication, and apoptosis. So far, various types of models belong to this culture: either the culture based on natural or synthetic adherent matrices used to design 3D scaffolds as biomaterials to form a 3D matrix or based on non-adherent and/or matrix-free matrices to form the spheroids. In this review, we first summarize a comparison between 2D and 3D cultures. Then, we focus on the different components of the natural extracellular matrix that can be used as supports in 3D culture. Then we detail different types of natural supports such as matrigel, hydrogels, hard supports, and different synthetic strategies of 3D matrices such as lyophilization, electrospiding, stereolithography, microfluid by citing the advantages and disadvantages of each of them. Finally, we summarize the different methods of generating normal and tumor spheroids, citing their respective advantages and disadvantages in order to obtain an ideal 3D model (matrix) that retains the following characteristics: better biocompatibility, good mechanical properties corresponding to the tumor tissue, degradability, controllable microstructure and chemical components like the tumor tissue, favorable nutrient exchange and easy separation of the cells from the matrix.

1. Introduction

Cell culture systems, applied in biology, have contributed to reducing laboratory animal use and ensured the progression of research, pharmaceutical discovery, and the evolution of medicine [1]. Initially, cells were grown in two dimensions and attached to polystyrene utensils or flat adherent surfaces (2D) but then researchers started to grow them with attachment proteins in a synthesized extracellular matrix (ECM) (3D) [2]. The two-dimensional (2D) in vitro cell culture system is a traditional application on flat support [3,4] for cell growth in a monolayer. Historically, this system has been applied in research since the early 1900s [5,6], specifically in the co-culturing of cellular heterogeneity [7,8], and in oncological research as a tool to evaluate the biological performance of bioactive molecules [7]. On the other hand, this type of culture has many limitations since it cannot ideally mimic physiological conditions and the natural microenvironment such as structure, physiology, biological signals of living tissues, and cell-matrix interactions [9,10,11,12,13,14]. Indeed, the communication of cells with their ECM, which is absent in 2D, controls the cell growth, proliferation, and function [15,16]. Cells cultured in 2D have been forced to modify various complex biological functions such as cell invasion, apoptosis, transcriptional regulation and receptor expression [2,14,17], cell proliferation and anti-apoptosis [18,19]. To overcome these limitations, researchers are currently providing and developing new in vitro 3D cell culture systems to boost research (healthy and tumor) [20,21]. The first 3D cell culture model was provided in 1992 by Petersen and Bissell, who described 3D organotypic structures to mimic breast structures in cancerous and non-cancerous cases [22]. This 3D system mimics the natural physiological properties and conditions, as well as enhances the development of new treatments at the preclinical stage in the future [23,24,25,26]. It overcomes the problems associated with traditional 2D in vitro culture, provides more valuable information about 3D cell-cell and cell-matrix interactions and presents a more clinically representative response to therapeutic agents [24,26,27]. Moreover, such a new type of culture will progressively evolve to have broader objectives to better assess the biological and molecular pathways during malignant transformation [28,29,30]. The main role of 3D culture is therefore to imitate the structure of the ECM of the tissue. ECM is a scaffolding of non-cellular fibrillary proteins, various structural macromolecules (accessory proteins), and adhesive molecules that provide structural and biochemical support to cells and are essential to many basic processes [31,32,33,34,35]. In addition, it forms cell-binding sites that control cell adhesion and migration [36]. From a structural point of view, it is composed of:
(i)
Interstitial ECM (stromal) contains biomolecules that can be organized into two main classes (proteins, glycoproteins) and proteoglycans (polysaccharides) [37]. It consists mainly of several protein molecules such as collagen I and III, self-arranged polysaccharides in fiber networks of glycosaminoglycans (GAG) such as hyaluronic acid (HA), proteoglycan (PG) and fibronectin [33,34,38,39,40];
(ii)
Basement membrane located at the basal side of epithelial or endothelial cells in normal tissues providing a physical barrier between epithelial cells and connective tissue (stroma) of the organ (always allow gas diffusion and transport of signaling molecules) [37,41,42].
These characteristics and ECM composition can be reached by 3D culture, and 3D culture can be achieved either via scaffold-free structures (i.e., spheroids) or scaffold-based structures (i.e., hydrogel-based supports and hard polymer material-based supports). Thus, in this review, we will be explaining the different techniques to reach 3D culture system, their advantages, and disadvantages.

2. 2D versus 3D Cell Culture

Cell-cell and cell-matrix interactions cannot be studied in 2D models in contrast to 3D models that are able to mimic these conditions in vitro. So, 3D culture provides a pragmatic pathophysiological microenvironment [11,13,26] and plays a potential role in cancer drug discovery due to the lack of preclinical models relevant to 2D cultures [25,43,44,45]. Inserts can be made up of biomaterials with properties like the ECM, and cells can also produce certain ECM proteins like collagen. Differences in physical and physiological properties between 2D and 3D cultures make 2D cells more sensitive to the effects of drugs than 3D cells since 2D cells are unable to maintain normal morphology as 3D cells can, and due to the difference in the organization of surface receptors on the cell [46,47]. It should be noted that there is growing evidence suggesting that cells cultured in a 3D system behave differently from those cultured in a 2D system and retain important signals from the ECM [28,48,49,50,51,52,53,54,55]. Appropriate 3D culture thus provides a more physiologically relevant approach for the analysis of gene function and cell phenotype ex vivo [56]. In recent years, reconstructed 3D culture has become a method of choice for summarizing the tissue architecture of benign and malignant tumors [49,57]. Thus, 3D culture can provide an important tool for better understanding changes, interactions, and cellular and molecular signaling during malignant transformation [13,58,59]. Table 1 summarizes the differences between the characteristics of 2D and 3D culture models.

3. Extracellular Matrix Composition

The ECM forms the non-cellular physical support for the cellular constituents of all tissues and organs. The components of the ECM encompass cellular and biomechanical signals that maintain morphogenesis, differentiation, tissue homeostasis, integrity, and elasticity [60,61].
The ECM is divided into two main parts: the structural interstitial matrix surrounding the cells (collagen I and fibronectin) and the basement membrane (collagen IV, VIII and X, laminins, nidogen, perlecan, and integrin receptors) separating the epithelium from the surrounding stroma [62,63,64,65,66]. In mammals, ECM contains about 300 proteins (central matrisome), having 43 collagen subunits, 36 proteoglycans (PG), and about 200 complex glycoproteins [67]. The PG glycosaminoglycan (GAG) (proteoglycans that form the intercellular interstitial gel) [67,68,69] and fibrous proteins (collagen, elastin, fibronectins, laminins) form the essential macromolecules of ECM [60,66,70]. The shape and structure of PGs vary according to their functions in ECM. Three parameters allow the classification of PGs: central proteins, their localization, and their GAG composition (unbranched polysaccharide chains, sulfated or not sulfated) [69]. Normal glandular epithelial tissues, including the breast, are composed of a simple layer of epithelial cells that cover the internal cavity of the canal, their apical pole that is in contact with the light-filled with liquid, and their basal pole that rests on the basement membrane. Then, at the limit with the interstitial ECM (stroma), the layer of myoepithelial basal cells rests [71]. Then the homeostasis of this epithelial tissue depends on communication and reciprocal interactions with the stromal microenvironment [72]. Moreover, ECM is a rich reservoir of growth factors and other bioactive molecules (metabolic precursors). It is due to the signaling of this reserve that, on the one hand, cell proliferation, cell differentiation, and delay of apoptosis take place [73], and on the other hand, the reciprocal interaction of cells and the microenvironment is possible [74,75,76]. In addition, ECM plays a basic role in the development and maintenance of epithelial tissues. For example, the ECM of human breast tissue is composed of protein fibril complexes intertwined in a network of carbohydrate chains of GAGs.
From a structural point of view, protein components, including laminins, fibronectin, and collagens, resist tensile forces, while carbohydrates, composed mainly of hyaluronan chains, chelate water, and resist compressive forces. Thus, the ECM is a key regulator of normal homeostasis and tissue phenotype [77].
Collagen is the most abundant structural protein in human tissues and constitutes about 30% of the body’s total protein mass [42,66]. Generally, collagens, formed of 28 unique subtypes discovered [63,78,79], can be grouped either in fibrillar collagen (collagens I-III, V, and XI) or non-fibrillary [67]. They regulate adhesion, cell migration [42], and tensile strength to maintain homeostasis [80]. Basically, interstitial collagen is secreted by fibroblasts, being able by this to organize the alignment of collagen fibrils (I and III) [81] in the fibronectin, hyaluronic acid, metalloproteinases (MMP), growth factors promoting cell differentiation, growth, and migration [79,82,83].
Hyaluronic acid (HA) is a non-sulfated linear GAG polysaccharide [69] with hydrophilic characteristics and is resistant to high compressive forces. HA adopts very extensive conformations for the formation of hydrogels [60,84]. It is a natural component essential for ECM to maintain compliance with compression and ensure ideal homeostasis in combination with collagen [80]. In addition, its abnormal accumulation in ECM may promote tumor migration [85,86].
Fibronectin is a multidomain protein that interacts with different components of ECM to facilitate cell-ECM connection [66], thus forming a fibrillary network [87]. It is involved first in guiding the direction of the organization of interstitial ECM through its reciprocal assembly with type I collagen [87,88,89,90,91] and second in cell migration during development and tumor metastases [72].
Laminins are glycoproteins involved in adhesion, differentiation, migration, phenotype maintenance, and have resistance to apoptosis [92]. It ensures the assembly of the basement membrane as well as ECM-cell interactions [93,94,95,96].

4. Three-Dimensional Cell Culture Scales

Various methods have been developed to address the growing demand for cell culture due to the lack of a single technology that fulfills the needs of all 3D cell cultures. 2D culture omits the effect of ECM molecules. Yet, its density and packaging contribute significantly to the creation of a 3D atmosphere. The 3D model is an in vitro reconstitution of the ECM after being inspired by the native microenvironment. It keeps the geometric, mechanical, and biochemical properties of the ECM [97]. It consists of different cells integrated into a specialized environment arranged in a way that forms 3D tissues similar to the natural tissue structure [98,99]. 3D culture models then make it possible to study the morphology and cellular organization shaped by ECM interactions, which are altered during oncogenic transformation. This makes the 3D models of in vitro tumors essential tools to study the mechanisms of cancer growth and metastasis [55,100,101]. They are most useful when they support tissue growth from primary human cells and include defined and physiologically relevant components [49,57,102,103]. Appropriate 3D culture thus provides a more physiologically relevant approach for the analysis of gene function and cell phenotype ex vivo [56]. The engineering of 3D culture is based on different main principles: the nature of the cells (the selected, isolated, appropriate strain cell line, primary cells, and tissue origin), 3D artificial microenvironment (ECM imitation) in which they are grown, scaffold-based of biomaterials (natural, synthetic, or hard), signaling molecules (proteins and growth factors) and bioreactors (cell culture) that support a cellular environment that is biologically active [104,105,106]. For these reasons, these parameters must be evaluated before choosing the most relevant technique and appropriate model. Culture systems can be either scaffold-based on natural or artificial solid scaffolding or scaffold-free, such as spheroids (non-scaffold based) [104]. The details of scaffold-based 3D techniques overview with these attributes are described in Table 2.

4.1. 3D Scaffolding Structures

In these techniques, the cells are grown in the presence of support which can be hydrogel-based supports and hard polymer material-based supports, natural or synthetic, of animal or vegetable origin. The polymer scaffolding offers a three-dimensionality favorable to cellular behavior in the microenvironment [124,125]. It is the most used model, especially collagen-based hydrogel, due to the major constituent elements of the basement membrane. Hydrogels are cross-linked networks formed of hydrophilic polymers attached through physical, ionic, or covalent interactions [126]. Their hydrophilic character allows them to absorb water that penetrates successively between the polymeric chains and causes swelling and thus the formation of the hydrogel [127]. Hydrogels can be natural (natural polymers), synthetic (synthetic polymers) or hybrid (natural and synthetic) depending on the biocompatibility advantage or the physico-chemical nature, respectively [128,129] (Table 2). These polymer-modified structures can be used as a matrix for cell culture in vitro or to make 3D spheroids [16]. The main advantage of hydrogels is that their physico-chemical properties are adjustable and could appropriately mimic the biochemical and mechanical properties of the true native ECM. Cells can be deeply seeded into a porous hydrogel and easily recapitulate nutrition and oxygen (by diffusion) [127].

4.1.1. Hydrogels

Some used hydrogels such as those of collagen are expensive, present a lack of reproducibility, and require extensive handling and specific equipment but present the opportunity of cellular heterogeneity as well as spontaneous cell organization (can be heterogeneous) [8,130,131] (Table 2). Another option is to work with other scaffolding products, such as Hydrogel, which is a synthetic nanofiber peptide scaffolding. The stiffness of the 3D culture can be controlled by adjusting the hydrogel concentration. Above all, this system could be applied to study the interaction between any type of neoplastic cells. It may even be possible to design more complex systems using more than two different cell types [130,132]. Hydrogels are unique because of their ability to mimic ECM while allowing soluble factors such as cytokines and growth factors to travel through tissue-like gel [47]. There are different types of hydrogels: natural and synthetic. Natural gels (natural polymers) are, for example, fibrinogen, HA, collagen, Matrigel, gelatin, chitosan, and alginate [133,134,135]. The hydrogels are, by definition, networks composed of hydrophilic polymers that are not cross-linked, and this allows them to swell widely by covalent bonds or to be held together by intramolecular and intermolecular physical attractions [106,136,137], maintaining their 3D structure [138]. Due to their hydrophilic and hydrated character, they can absorb a large amount (thousands of percent) of water or biological fluids and inflater easily without dissolving, thus mimicking the structures and physical properties of soft tissue ECM. They are then soft and rubbery after swelling and resembling living tissues [106]. On the other hand, gels differ from hydrogels since they are semi-solid materials (can appear more solid than liquid) consisting of hydrophilic polymers comprising small amounts of solids, dispersed in relatively large quantities of liquid [139]. Depending on the nature of the polymer, hydrogels can be classified into different natural or synthetic categories and interconnected by physical and ionic interactions and even covalent bonds (hydrogels based on ECM proteins, natural hydrogels, and synthetic hydrogels) with distinct biochemical, physical and mechanical properties [111,112,113,114,115,138]. For this, they were also explored as 3D models for cancer research. The use of 3D scaffolding models based on scaffolding expands the range of options available to researchers [140,141,142,143,144,145,146,147,148].
  • Protein-based EMC
These naturally formed biomaterials from biological polymers have been used in the manufacture of 3D platforms for breast cancer culture, including Matrigel® [149,150,151], collagen [152,153,154,155,156,157,158,159], HA [160,161,162,163,164,165,166], alginate [1,167,168,169], and gelatin [169,170,171] (Table 2).
The Matrigel® is the gold standard. It is a tissue formed of a mixture of gelatinous proteins derived from the basement membrane isolated from the mouse sarcoma Engelberth–Holm–Swarm (EHS) and commercially available under the brand name Matrigel® (BD Biosciences) [107,108,109,110] (Table 2). This extract is liquid at 4 °C and turns into a gel at 37 °C under physiological pH and ionic strength. It is then a ready-to-use solution that allows user-defined use. Recently, it has been widely used in 3D experiments in cell biology to assess cell migration, cancer cell behavior, and to create organoids in vitro, as it produces a large amount of ECM rich in type I collagen, laminin-111, heparin sulfate proteoglycans (perlecan) and nidogen [172]. Matrigel® is considered the best product on the market used in the production of most 3D tests performed in cell biology [151,173,174,175,176]. The success of Matrigel® is also due to its biological activity, which allows under normal culture conditions the differentiation of several cell types and the formation of complex structures such as mammary glands of acinous structures [173,177]. In addition, the bioactivity of Matrigel® is due to the presence of soluble growth factors such as fibroblast growth factor (FGF), epidermal growth factor (EGF), transformative growth factor-β (TGF-β), and matrix metalloproteinases (MMP), including MMP-2 and -9. Although easy to use, the presence of these growth factors in unknown and uncontrollable quantities can have an impact on research (positively or negatively) which is why many researchers may prefer to manufacture their hydrogel systems using defined concentrations [24]. In addition, as Matrigel® is produced and purified from an animal, there is a lack of control over its exact composition and batch-to-batch variability in its contents [178]. Similarly, it presents difficulties of handling when it is in a refrigerated liquid state [104]. Although the drawbacks of Matrigel® are significant, Matrigel® is a widely available model for studying many fundamental questions in cell biology, cell adhesion, and cancer research as a versatile platform for in vitro 3D cell culture [151,173,174,175,176].
Most natural polymers are structural molecules derived from mammalian ECM. Many different materials were used to develop in vitro 3D breast cancer scaffolding. They were first used as a coating of tissue culture boxes to promote cell adhesion and spread after they are incorporated into 3D materials in different forms (hydrogels, freeze-dried materials, and surface coating of bulk inorganic materials such as elastin, collagens [152,153,154,179], fibronectin, laminin (mainly laminin-111) [152,161,162,179], GAGs (chondroitin sulfate and heparan sulfate, hyaluronan [162,165,166,179], alginate, and gelatin [167,168,169,170,171].
Classically, they are purified from ECM-rich animal tissues such as dermis and tendons (collagen type I, elastin), cartilage (collagen type II and GAGs), tumors (laminins and collagen IV), or directly from the blood (fibrin and fibronectin). On the other hand, with progress, it is now possible to obtain some of these molecules from recombinant DNA sources, which makes it possible to work with human ECM molecules, produced with a high degree of purity and free of many pathogens, but they are more expansive [180]. Non-mammalian ECM molecules are also widely used in the design of biomaterials, mainly for their ability to self-assemble in 3D structures, but due to their origin, they lack many aspects of cell adhesion on their structures and should most often be supplemented with adhesive molecules or peptides to obtain a biologically active material. These polymers are hydrogels based on chitin/chitosan (polysaccharide purified from the exoskeleton of fungi or arthropods), agarose, or alginate (both polysaccharides purified from algae), and fibroin (cocoon protein from silkworms and spiders) [181].
  • Natural hydrogels
Collagen-based hydrogels are natural hydrogels. Collagen, as the main component of ECM, plays a key role in the development and spread of cancer [40,106,182,183,184]. It affects the tumor microenvironment [185,186] especially in breast cancer signaling [182,183], differentiation and migration through cell-matrix interactions [187,188] (Table 2). Depending on the type of tissue, collagen fibrils organize themselves in a variety of ways to form collagen fibers suitable for the specific functions and properties of tissues [189]. The mechanical properties, architecture, and biodegradability of collagen hydrogels can be finely modulated by adjusting their concentrations [190,191] and preparation parameters [136,137,192,193]. Thanks to its specificity of spatial organization and self-assembly of collagen in acid solutions [194], the architecture of hydrogels can be controlled by the manipulation of ionic force, pH, and temperature during frost polymerization [174,195]. The concentration of collagen increases gradually with the regulated evaporation of the solvent, as well as modifies the organization of collagen molecules. The original collagen solution to a solid hydrogel structure (i.e., the so-called “soil/frost transition”) retains the tissue-like molecular organization of collagen molecules [196] (Figure 1a). In addition, an increase in collagen concentration (i.e., the ionic strength of gels) leads to an increase in fiber density and a reduction in pore size but has no effect on fiber diameter [195]. In contrast, increasing temperature and pH accelerates polymerization, reduces fiber diameter and pore size, and also increases the mechanical properties of hydrogel [174,195] (Table 3). As a result, many collagen-based 3D models of in vitro cancer culture have been developed [11,152,153,154,155,158,197,198,199,200,201,202].
Polysaccharide-based hydrogels are natural hydrogels. Proteoglycans (PGs) have a protein that is covalently bound to GAG chains. Thanks to the polyanionic profile of GAGs (due to the sulfate groups), they attract water, thus causing the swelling of the GAG [220] (Figure 1b). This swelling can then open pathways of invasion and migration of cells that resemble the state of invasion and cancer metastases [221] (Table 2). These PGs are directly involved in cellular functions and the release of active molecules (growth factors, cytokines) [221]. Hyaluronic acid (HA) is a major category of structural macromolecular components of ECM. It is an unsulfated GAG. The polyanionic nature of GAG attracts water, causing GAG to swell. Concerning HA hydrogel, this natural polysaccharide can be chemically modified to similarly mimic native tumor tissue by adding acrylate or thiol groups that have cross-linked to form a network with a pore size of 70 to 100 nm [220]. Similarly, their sulfation pattern contributes to the binding of growth factors to GAGs [222,223,224]. Indeed, GAGs provide hydration and compressive strength by binding to water as well as are involved in different biological processes such as tumor progression, angiogenesis, and cell development [33,35,39]. It is not only a structural component of the tumor ECM but also a biologically active molecule that has been used extensively in the formation of 3D in vivo tumor models [225]. This use is because of its biodegradable, non-immunogenic, non-inflammatory [205] hydrodynamic and swelling characteristics to fill most of the extracellular interstitial spaces of tissues in the form of hydrated gels [33,35] and promote cancer progression [226] (Table 3). These properties make HA an ideal matrix for preparing 3D tumor models. HA gels are formed by covalent cross-linking (reaction with carboxylic acid groups) with hydrazide derivatives. They carry inherent biological properties such as protein grafting, but they are mechanically poor. HA is most often incorporated into the constituent materials of hydrogels [38,219,227,228,229]. Recent studies have demonstrated the usefulness of HA-based scaffolding for improving adipose tissue development in vivo [230,231] and in vitro [232,233]. Table 4 summarizes the differences between scaffolds made with pure collagen vs. collagen-HA-based ones.
Dextran is a bacterial polysaccharide, consisting mainly of a glucosidic-1,6 bond of D-glucopyranose residues, used for more than 60 years in the medical and biomedical fields [236]. Dextran is widely used for biomedical applications. Its advantages are its biocompatibility, low cost, good water solubility, ease of modification [237], antifouling properties [238,239]. Dextran glucose must be oxidized in duplicate and then followed by a freeze-drying step [240,241,242,243].
Chitosan is the second most abundant natural polymer. It is a linear polysaccharide derived from chitin in the form of a deacetylated derivative [244]. Its advantages are its structure (D-glucosamine bound in (1-4) + N-acetyl-glucosamine) that mimics the structural characteristic of the GAGs of ECM. It’s a biomaterial widely used in biomedical, biocompatible, biodegradable, and can be produced on a large-scale, easily transformed (simple freeze-drying) [245]. It shows a disadvantage of solubility in neutral solutions (it adds cysteine) [246]; no gel is formed without the grafting of cysteine, and poor mechanical properties of the gel are observed [247].
  • Synthetic hydrogels
Most synthetic hydrogels are synthesized by polymerization of synthetic polymers, which exhibit versatile biophysical, mechanical, and biological properties in 3D breast cultures to study the relationship between the microenvironment and malignant tumors in vitro [131,133,248]. They are formed of synthetic polymers comprising poly-ε-caprolactone (PCL) [249,250,251], polyethylene glycol oxide (PEG) [38,252,253,254], polyvinyl alcohol or in a mixed solution or a combination of copolymer with poly-lactide-co-glycolide (PLGA) (PLG) [249,250,251,255]. In addition, their use is more preponderant in 3D culture of many cell types including neural [206], bone [207,208], cartilaginous [209,210], muscle [211], and renal cells [212]. Synthetic organic polymers offer a wide range of creativity to produce 3D materials. Although they inherently lack basic biological activity, they have great flexibility in treatment and are easier to be produced [116] (Table 2). The diversity of synthetic polymers used in biomaterials is great as they can be transformed into 3D materials with many types of techniques (electro piping, foaming, hydrogel, and sheets), some of which are not bearable by biological polymers. Polyesters and polyhydroxy acids can be biodegradable thanks to the presence of hydrolyzable bonds in their skeleton, while polyacrylamides and polyacrylates are almost unbreakable in cell culture [117]. Synthetic polymers have the inherent properties to form a 3D scaffolding by which they contain active chemical groups (amine, acid, or alcohol functions) sensitive to chemical reactions, thus providing an ECM model with well-defined characteristics [204,217]. Physiologically, they are irrelevant and can release toxic degradation products to cells [216] (Table 3). Similarly, they are unable to provide the biochemical signals necessary to “communicate” with the cell. To overcome this limitation, synthetic polymers can be functionalized by adding signaling biomolecules, such as peptides, growth factors, and glycans [213,214,215].
Cells embedded in natural biopolymers take advantage of the signaling already present inside the matrix, while synthetic polymers lack signaling patterns capable of modulating the cellular outcome [203]. The stiffness of biomaterials is also very important for cell proliferation and behavior, but an increase in the stiffness of a matrix (PEG gels) acts as a physical barrier for 3D cells, preventing their proliferation and migration [212]. Similarly, studies have already shown that changes in the cross-linking density of PEG-based hydrogel cause changes in cell growth and morphology [218,256,257]. For all these reasons, their applications are limited in the engineering of tumors in vitro [216]. Levenberg et al. (2003) used PLGA and PLA to form porous scaffolding to create a 3D artificial microenvironment for human embryonic stem cell differentiation. While this was partly successful, they also demonstrated the difficulties in getting cells to infiltrate throughout the scaffolding [258]. Similarly, the biodegradation of scaffolding based on poly-lactic acid, poly-glycolic acid, and their copolymer PLGA can lead to the release of by-products such as lactic acid [259,260]. For these reasons, biodegradable materials are not practical for routine 3D cell culture where issues such as storage life, storage, and product consistency need to be taken into consideration [104] (Table 3).
Nanofibrous scaffolding is a thin sheet composed of synthetic polymer fibers (PLGA or PLGA-PEG) of nanometric size randomly aligned. It forms a fibrous support for ECM and provides topographic features necessary for cell adhesion and growth. These nanofibers can be manufactured by several techniques such as electrospiding, phase separation, and self-assembly with varying chemical properties, diameters, lengths, porosities, and mechanical properties [261,262,263,264,265]. In addition, the authors evaluated the usefulness of this model for drug testing by growing a tumor biopsy on 3D scaffolding and determining the effect of the drug on it. These nanofiber-based models have several key features; for example, providing topographic characteristics to cancer cells for 3D tumoroid development and reproducibility with the ease of tumoroid imaging could make it an elegant approach for drug testing.
Alvetex® hard-base scaffolding (Figure 2) provides a large internal volume and 3D space that cells can occupy and form a tissue. Cells find a more physiological shape because they are not seeded on a flat surface but in the presence of fibers or sponge-shaped structures with high consistency and reproducibility [119]. These scaffolds are composed of non-degradable inert materials (polystyrene or polycaprolactone PCL) to avoid the formation of by-products. They can be manufactured by electrospinning [120,121,122] or by gas foam technology [123] to improve their application in cell culture, but very few have been developed into a commercially successful process for 3D culture (Table 2). They are important tools in the study of cell-ECM interactions due to the ability of the scaffolding to reproduce the structure of the ECM and high porosity [266,267]. Moreover, it is less affected by cytotoxic compounds. Polystyrene is a familiar substrate to the user in 3D cell culture since it is inert and does not degrade during normal use [55,121,122,268,269] (Table 3). On the other hand, polystyrene has some disadvantages: it is rigid (its rigidity must be controlled), does not have the biomechanical properties found in soft tissues, can display cytotoxicity [270], and a lack of biochemical stimuli (for example, molecules dependent on cell anchorage). However, these can be solved by adding known ECM proteins. In this case, a balance must be found between the needs of the model and the objective of the study [104].
The advantages and disadvantages of the described hydrogel-based scaffolds are summarized in Table 3.

4.1.2. Synthetic Strategies

  • Porous material
Despite the good biocompatibility of hydrogels due to their water content, they most often have low mechanical properties and high degradation rates. For this, porous materials have been designed with interconnected pore networks and surfaces or fibers to support cell adhesion. These materials are discriminated against 3D models of ECM of nano porous scaffolding where pore structures are in the range of cell diameter (about 10 μm) [130,271,272,273] (Figure 3a). Nevertheless, microporous structures allow more efficient cell penetration and migration into the material but with specific pore sizes since a micro-size can represent a barrier to cellular colonization of the material and limit cellular interactions at the edges of the material [274]. Like hydrogels, porous materials can be prepared with natural and synthetic polymers by different techniques, including electrospinning, phase separation, model creation, and vapor polymerization [274]. The porosity based is mainly used with synthetic polymers (PLA/glycolic acid and PCL) but is also suitable for natural polymers such as collagen and silk fibroin [275]. For example, porous collagen materials, due to their high porous structures, make so-called “collagen sponges”, by thermally induced phase separation. The phase separation of collagen molecules from the water-based solvent is due to the freezing of acidic collagen solution, which induces one that is then eliminated by freeze-drying. Pore size and interconnectivity can be altered by modulating collagen concentration and phase separation temperature or by mixing collagen solutions with other natural polymers such as GAG [276] or synthetic polymers such as PLA. The low mechanical properties of freeze-dried collagen materials are often enhanced by the addition of GAG by cross-linking them with chemical species (aldehydes) using dehydrothermal processes [191].
  • Hydrogel technology
The cross-linking of natural hydrogels such as agarose, fibrin, and especially collagen and HA with high water content is one of the popular options in 3D culture that ensures the encapsulation of cells in a hydrogel comprising a loose scaffolding structure [121,279,280,281]. Considerable progress has been made in the preparation of hydrogels called smart hydrogels to better mimic the artificial ECM protein microenvironment [282,283]. However, these hydrogels may undergo unusual changes in the structure, mechanical characteristics, and swelling behavior of the network (support, cell growth), as a result of variations in pH, temperature, light, ionic, and force or electric field or enzymes [284,285,286,287]. Regarding freeze-drying, its main purpose is to regulate porosity and form a collagen sponge. Changing the porosity of hydrogels (the number, size, shape, and interconnectivity of pores) will promote cell growth and homogeneous cell seeding in cell culture [288]. Pore size is also a very important parameter to avoid inhibition of cell penetration. However, the diameters of all cells are less than 200 μm (for nutrition and oxygenation) [289]. To do this, it is necessary to choose the optimal size suitable for each cell type [289,290,291]. A resume of this method, its advantages and disadvantages are summarized in Table 5.
  • Collagen Hydrogel by Freeze-Drying (Lyophilization)
Porosity is an amount of open pore volume in a scaffold that provides suitable support for cell colonization, ECM production, and subsequent spheroid formation. Highly porous scaffolding is then essential for the transport of nutrients and the disposal of waste [334]. Pore size (diameter of circular pores or the longest length for non-circular pores) is one of the major factors that affect cellular behavior in the porous matrix, such as migration, adhesion in pores, and interaction with neighboring cells [335]. Lyophilization is a dehydration technique that results in the formation of an interconnected circular/dry oval porous microstructure [276,292,293,294] (Figure 3b). The solution is frozen before undergoing a vacuum drying process leading to the sublimation of ice crystals [234]. This technique is currently the most used for the manufacture of collagen-based scaffolding due to the easy control of the architecture and mechanical properties. In addition, this architecture can be affected by collagen concentration, temperature, and freezing speed [276,295,296]. It has also been shown that a decrease in the final freezing temperature reduces the size of the ice crystal and consequently the size of the pores [297] even for the freezing speed, which must be slow and controlled to result in homogeneous scaffolding (pore shape and size) [298]. Variation in collagen concentration affects the stiffness [234], size and porosity of scaffolding pores since an increase in collagen concentration from 0.5% to 1% (w/v) increases pore size and reduces scaffold porosity [191,336]. To address the complexity of component alignment observed in the ECM, Campbell et al. used a polycarbonate mold, cylindrical wells, and sharp copper inserts (coated with PTFE). The inserts were thermally insulated from the freeze dryer shelf by a thin 1 mm rubber mat [337,338,339]. In a recent study, Hume et al. used 1% (w/v) of collagen (bovine Achilles’ tendon) solubilized in 0.05 M acetic acid followed by custom freeze-drying to produce reticulated anisotropic scaffolding (reticulated DAC/NHS). Samples of xenograft mammary tumors and fragment co-culture with 3T3-L1 pre-adipocyte cell line have been successfully cultured in collagen scaffolding (pore size 100 μm), highlighting the promise of ex-vivo application [235]. This is commonly used as a gel for adipose tissue engineering [340,341,342,343], noting that adipocytes increased the migration of tumor cells into scaffolding on day 10 [235]. A resume of this method, its advantages and disadvantages are summarized in Table 5.
  • Electrospinning Hydrogel
The electrospinning technique uses electrical forces to form a network of fibers that offer a large surface area from polymer solutions or melts [344] (Figure 3c). This technique is fast, efficient, relatively inexpensive, versatile, and produces microfibers with a diameter of less than 100 nm [345,346]. The diameter of the fibers increases with increasing concentration/viscosity of the polymer [347,348,349,350]. This creates uniform fibers and reduces the incidence of fiber defects such as beading (low concentration or surface tension problems) [348,351]. Fiber thickness generally increases pore size (space between fibers) [299,300], while fibers with a smaller diameter exhibit the opposite effect due to the higher density of the fiber network inside the scaffolding [352]. Noting that manufacturing parameters should be adjusted to avoid overly dense and tight fibrous networks as dense fibrous networks can prevent cellular infiltration into the scaffolding; high porosity is then essential [299,301,302]. In addition, its limited control of porosity and relatively poor mechanical properties reduces its use in a 3D culture based on hydrogel scaffolding [303,304,305,306] (Table 5).
Collagen-based scaffolding manufactured by electrospiding uses 1,1,1,3,3,3-hexafluoroisopropanol (HFIP) [307,353], trifluoroethanol (TFE) [308,309] solvents, although these nanofibrous scaffolds were wrung out using a more benign water/salt/alcohol solvent system [310]. The diameter of the fibers in these collagen-based scaffolds ranged from 100 to 900 nm, with differences obtained by changing the specific electrospiding parameters. On the other hand, crosslinking agents can be used to increase the mechanical properties of electrospirated scaffolding [307,309,310,311,312,353]. Highlighting the potential of collagen-based scaffolding in 3D in vitro culture model, excellent cell proliferation and viability has been observed in many manufactured models [307,311,312,354]. Szot et al. used this model to assess cellular behavior on electrospirated collagen/PCL fibers manufactured in terms of growth, proliferation, adhesion, and infiltration (fiber diameters ~400 nm (5%) to 2250 μm (15%) depending on concentration) [299]. Recently created (2021) by Malakpour-Permlid et al., a 3D culture model is based on a PCL fiber network [355] resembling the collagen network of ECM due to its involvement in the main tissues and organs of the body. And as collagen is one of the most widely used biopolymers in tissue engineering [81,356], this fiber network is applied in the culture of normal and cancer cells by mimicking the collagen structure that ensures the 3D fixation and growth of cells. Despite the advantageous bio-activity in vitro and in vivo of collagen-based hydrogels and biomaterials, the major problem of clinical translation and therapeutic use remains related to the animal origin of collagen (i.e., bovine collagen, of porcine origin type I), with potential pathogenic content (disease transmission) [357] and lack of growth factor (addition of TGF-B) [355].
  • 3D-Printing Scaffolding for 3D cell Culture via Stereolithography
Stereolithography is another method commonly used for imitation of complex structures in vitro of the intestine artificial 3DP models such as microvilli. It is an excellent candidate for studying homeostasis regeneration mechanisms in vitro. It is based on a construction of the different layers that harden to visible light or infrared [313,314] (Figure 3d). Each layer will be superimposed after the next layer has dried (generally by UV). These layers are printed by a specific thin sheet material: The Installation Multiresolution 3D Printer (Dilase 3D, Kloe France), layer by layer until the scaffolding is finished [313]; it is then placed under UV light where it is post-cured [315,316,317]. This support can be combined with a polymerizable photo hydrogel (PCL fibers, PCL/gelatin) that promotes cell line growth with 3D-printing stereolithography and produce different scaffolds size from mm to cm. However, the manufactured scaffolding is usually limited to a few tens of microns of resolution and needs specific end expensive equipment [277,316,317,318,319,358] (Table 5).
  • Micro Fluid
The microfluid support consists of silicon/elastomer-based devices. It creates a dynamic microenvironment presenting micro-channels with proportions from 1 to 1000 μm which are responsible for the exploitation of a small volume of fluids generally in the range of 10–9 to 10–18 (Figure 3e). In this system, the flow of the fluid is strictly laminar (in parallel layers) rather than turbulent (parallel and no strong) [320,321]. This support is normally applied in complex 3D structures or to synthesize matrices used in human transplantation. The generation of aggregates of different forms and the coculture of several cells reconstitute a more physiologically relevant tumor [322,323]. It simulates cell-cell contacts and biological signals controlled by spatial and temporal gradients of soluble biological factors [324,325,326], progression, invasion, angiogenesis as well as treatment efficacy [320,327] since the continuous flow of nutrients and therapeutic agents establishes a physiological profile such as that of blood circulation and intravenous injections. In fact, this system saves labor and the reagents used since it works automatically and consumes low cellular usage [278,328]. According to its manufacturing complexity, this system requires professional equipment and special design [322,323] and a higher budget than other strategies applied [329,330,331,332,333]. The different synthetic strategies are summarized in Table 5.

4.2. Scaffold-Free Spheroids

Self-assembly is a natural phenomenon that occurs during morphogenesis and organogenesis. In culture, these techniques are considered the least complicated to apply because of the absence of a fastening surface or scaffolding that allows cell colonies to self-assemble and form aggregates of non-adherent 3D microtissues called spheroids [359,360,361,362,363,364,365,366]. According to a general definition, spheroids are aggregates of cells growing in a 3D way in suspension; they can be mono or multicellular (homo or heterotypic). The cells form hard spherical structures with a well-balanced morphology of variable size (50 to 150 μm), formed by a necrotic nucleus and a peripheral layer [10,367,368,369]. Historically, Holt Freter was the first to use spheroids as a morphogenic model in his investigation of skin behavior during development in 1944 [370]. Then this mode was considered a powerful tool in research and clinical applications and is the best in vitro cellular model for high-throughput screening [138,371,372,373]. Recent advances in tissue engineering and regeneration have provided new techniques for the generation of tumor 3D spheroids for a variety of cancer types in vitro. This model of multicellular tumor spheroid was initially created in the early 1970s by radiobiologists, and then it was managed to be used for a wide variety of cancer cell lines. Nowadays, a multitude of techniques are used for the production of spheroids because cells are unable to adhere to the support [7,363,364,365,374,375], either in the adhesion-free gel of the micro-well with superposition of cellular suspensions (e.g., agarose gel or alginate gel) [376,377,378,379,380,381] or by other techniques as pellet culture, suspended goutte and filature culture [382,383,384].
Spheroids have been widely used by a simple, high-yield, inexpensive application protocol that allows the production of more spheroids to mimic the architectural and functional characteristics of native tissues [138,371,372,373,385,386], and to assemble models of different types of cancer in vitro such as breast cancer (spherical shape of the breast canal) [363,380,381,387]. On the other hand, this culture requires intense and precise work and presents a risk of the formation of spheroids of uniform size and shape [376,377,378,379,380,381]. Noting that this variation may be at the origin of the gel used but, in general, the gels (agarose, alginate, chitosan) used in the spheroid formations have benign characteristics such as plant origin that do not present a risk of animal contamination, significant stability at room temperature (but are biodegradable) and non-toxicity [167,168,169,369,388]. In contrast, agarose gels were formed by heating (near the boiling temperature) the solution that freezes with cooling. For these reasons, different porous architectures and mechanical properties can be constructed according to the modulation of agarose concentration [38]. For alginate citing a commercial product AlgiMatrix™ ready to use, it looks like a highly porous sponge (>90%), ready to use (freeze-dried alginate), stable at room temperature with long-term viability, non-toxic, biodegradable, and can easily be degraded by a dissolution buffer in a few minutes leaving cell aggregates intact for analysis (spheroids 50 to 150 μm) [10,367,368,369]. Similarly for chitosan (derived from crustaceans) has no binding domain to human cells therefore favorable to spheroids [245,377,389,390]. Table 6 summarizes the advantages and disadvantages of spheroid culture.

4.2.1. Technical Methods of Spheroid Formation

  • Pellet Culture
In this system, cells are concentrated at the conical bottom of a tube by centrifugal force (500 g, 5 min) to maximize cell-to-cell adhesions [359,391] (Figure 4a). After that, the supernatants are removed, and cell pellets are resuspended in a spheroid formation cell culture medium. We noted that shear stress due to centrifugation could damage cells, so to optimize results, the suspension can be incubated on an agitator for one hour before centrifugation [361,392,393,394,395,396] (Table 7).
  • Hanging drop
Hanging drop is a spheroid culture technique that uses surface tension and gravitational force to form definite size spheroids in the form of droplets that rely on gravity self-disassembly. It allows single cells to aggregate and fabricate spheroids in the form of droplets (Figure 4b) [300,301]. By controlling the volume of the drop or density of cell suspension, it is possible to control the spheroid size. We prepare a cell suspension at desired density distribution in the wells of a mini-plateau, then it will be placed on the mini-tray, and the entire mini-tray is overturned upside down [397,398]. The drop remains fixed on the mini tray on the inverted surface (surface tension). Therefore, this technique is mostly used due to the defined and controlled size of the spheroid (drop volume and suspension density) with inexpensive equipment. Despite high amount production (384 spheroids in a single network), heterotypic spheroids can appear with narrow size distribution [359,361,383,397,398,399,400] (Table 7).
  • Cultivation of Molded Lozenges and Liquid Overlay (Static Suspension)
The cultivation of molded lozenges and liquid overlay (static suspension) are culture technique that forms spheroids by interrupting the adhesion of cells on non-adherent culture plates or gel with non-adherent properties such as agarose with micro-well (agar) gel or pHEMA with superposition of cellular suspensions (Figure 4c) [359,361,401,402,403,404]. It is a simple method to monitor the formation and growth of spheroids. Since the cell binding to the support is inhibited, cells spontaneously form spheroids. It is forced to aggregate by continuous agitation with/without centrifugation. Despite the excellent non-adherent properties of agarose, this biomaterial has drawbacks in terms of producing heterogeneous spheroids in size and shape (Table 7) [359,361,376,399,403,404,405,406].
  • Spinner Culture Technique
It refers to the technique wherein the cell suspension in spinner centrifugal flask bioreactor, generated by a magnetic stirring wheel container, and which is continuously mixed by convection force stirring (Figure 4d). It adds the uniform and well-mixed single-celled suspension with constant continuous stirring to form the spheroid (may not be useful for cells with low cohesion; they have the risk of apoptosis). The stirring rate must be constant because a high stirring rate induces damage to the spheroid cells and a slow speed makes the cells sink to the bottom of the container (blocks the spheroids). In addition, it is difficult to follow the spheroids during formation (Table 7) [26,359,361,407,408,409,410,411].
Table 7 summarizes the technical methods of spheroid formation.

4.2.2. Technical Methods of Tumor Spheroid Formation

Due to their particular interest, spheroids are the most applied 3D models in oncogenic research. They form an effective tool capable of studying the variation in morphology, topography, size, cell organization, protein expression, and genes in the invasive and metastatic potential of cancer cells [363,380,381,382,383,384,387]. These tumor spheroids have a heterogeneous distribution with active cells proliferating on the surface of spheroid cells (oxygen and nutrients) and resting cells in the center [104,366]. Tumor spheroids may be homotypic formed only of cancer cells or heterotypic consisting of cancer cells with other cell types [375]. Spherical cancers can be classified into four groups:
  • Multicellular tumor spheroids are obtained after aggregation and compaction of the cultured cell suspension (1–7 days) under non-adherent conditions (well plates, vials or boxes + agar gel, agarose or polyH + a traditional culture medium depending on the cell line) [399,412,413].
  • Tumorospheres (floating sphere): Tumors are formed from a single cell capable of giving rise to a sphere by clonal expansion (5–7 days up to 1–2 months) under conditions of low adhesion (plastic with low adhesion) and with a stem cell medium (depending on the type of cancer, growth factors may be preferentially added) [414,415,416,417,418].
  • Tissue-derived tumor spheres (endoscopic biopsy): Tumor spheres derived from cut (scalpel blade) and minced partially dissociated cancerous tissues are generated by partial dissociation of tumor tissue and compaction/remodeling (2–5 days up to 12–18 days) in conventional FBS-supplemented medium [416,419,420].
  • Organotypic multicellular spheroids are formed from the cutting of partially dissociated tumor tissue (mechanically or enzymatically) under non-adherent conditions (plastic treated in culture and then non-adherent conditions) that have rounded during culture (1–3 days) [419,421,422,423].

5. Conclusions

Standard cell culture studies are widely used to delineate biological, chemical, and molecular pathways, first by traditional 2D culture and then by enhanced 3D culture. According to 2D cell limitations in some practices, recent advances in tissue engineering and regeneration then provided new techniques for a variety of 3D in vitro models. Cells develop in an organized three-dimensional (3D) matrix, and their behavior depends on interactions with immediate neighbors and ECM. The 3D culture can provide an important tool for better understanding changes, interactions, and cellular and molecular signaling during malignant transformation and metastasis. Three-dimensional (3D) cellular scaffolding is then essential for tissue engineering. So far, various natural and synthetic polymer hydrogels have been used to design 3D scaffolding as biomaterials. This is a barrier to mimicking the native ECM microenvironment, and therefore synthetic scaffolds may be more useful for investigations of specific tumorigenic steps. We provide here characteristics, advantages, and disadvantages of 3D cell culture compared to 2D types, different types of 3D matrices such as natural, synthetics hydrogel, and spheroids—the best rational classification of the most used 3D strategies models in cancer research. Finally, depending on the specific objectives, the most relevant 3D models must be carefully selected.

Author Contributions

Conceptualisation, O.H. and L.D.; Writing-original draft preparation, O.H.; writing and editing, L.D., M.D.-A., and F.C.-C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All pertinent data are presented within this manuscript.

Conflicts of Interest

The authors have no financial disclosures to make or conflict to report.

Abbreviations

2D: two-dimensional; 3D: three-dimensional; ECM: extracellular matrix; GAG: glycosaminoglycan; HA: hyaluronic acid; MMP: metalloproteinases; PCL: poly-ε-caprolactone; PEG: polyethylene glycol; PG: proteoglycans; PLA: poly-l-lactic acid; PLGA: lactic-co-glycolic acid.

References

  1. Fischbach, C.; Kong, H.J.; Hsiong, S.X.; Evangelista, M.B.; Yuen, W.; Mooney, D.J. Cancer Cell Angiogenic Capability Is Regulated by 3D Culture and Integrin Engagement. Proc. Natl. Acad. Sci. USA 2009, 106, 399–404. [Google Scholar] [CrossRef] [Green Version]
  2. Kleinman, H.K.; Philp, D.; Hoffman, M.P. Role of the Extracellular Matrix in Morphogenesis. Curr. Opin. Biotechnol. 2003, 14, 526–532. [Google Scholar] [CrossRef] [PubMed]
  3. Kramer, N.; Walzl, A.; Unger, C.; Rosner, M.; Krupitza, G.; Hengstschläger, M.; Dolznig, H. In Vitro Cell Migration and Invasion Assays. Mutat. Res./Rev. Mutat. Res. 2013, 752, 10–24. [Google Scholar] [CrossRef] [PubMed]
  4. Pampaloni, F.; Reynaud, E.G.; Stelzer, E.H.K. The Third Dimension Bridges the Gap between Cell Culture and Live Tissue. Nat. Rev. Mol. Cell Biol. 2007, 8, 839–845. [Google Scholar] [CrossRef] [PubMed]
  5. Becker, J.L.; Blanchard, D.K. Characterization of Primary Breast Carcinomas Grown in Three-Dimensional Cultures. J. Surg. Res. 2007, 142, 256–262. [Google Scholar] [CrossRef] [PubMed]
  6. Ferreira, L.P.; Gaspar, V.M.; Mano, J.F. Design of Spherically Structured 3D in Vitro Tumor Models-Advances and Prospects. Acta Biomater. 2018, 75, 11–34. [Google Scholar] [CrossRef] [PubMed]
  7. Breslin, S.; O’Driscoll, L. Three-Dimensional Cell Culture: The Missing Link in Drug Discovery. Drug Discov. Today 2013, 18, 240–249. [Google Scholar] [CrossRef]
  8. Ricci, C.; Moroni, L.; Danti, S. Cancer Tissue Engineering—New Perspectives in Understanding the Biology of Solid Tumours—A Critical Review. OA Tissue Eng. 2013, 1, 1–7. [Google Scholar] [CrossRef] [Green Version]
  9. Costa, E.C.; Moreira, A.F.; de Melo-Diogo, D.; Gaspar, V.M.; Carvalho, M.P.; Correia, I.J. 3D Tumor Spheroids: An Overview on the Tools and Techniques Used for Their Analysis. Biotechnol. Adv. 2016, 34, 1427–1441. [Google Scholar] [CrossRef]
  10. LaBarbera, D.V.; Reid, B.G.; Yoo, B.H. The Multicellular Tumor Spheroid Model for High-Throughput Cancer Drug Discovery. Expert Opin. Drug Discov. 2012, 7, 819–830. [Google Scholar] [CrossRef]
  11. Szot, C.S.; Buchanan, C.F.; Freeman, J.W.; Rylander, M.N. 3D in Vitro Bioengineered Tumors Based on Collagen I Hydrogels. Biomaterials 2011, 32, 7905–7912. [Google Scholar] [CrossRef] [Green Version]
  12. Hutchinson, L.; Kirk, R. High Drug Attrition Rates--Where Are We Going Wrong? Nat. Rev. Clin. Oncol. 2011, 8, 189–190. [Google Scholar] [CrossRef] [PubMed]
  13. Yamada, K.M.; Cukierman, E. Modeling Tissue Morphogenesis and Cancer in 3D. Cell 2007, 130, 601–610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Xu, F.; Burg, K.J.L. Three-Dimensional Polymeric Systems for Cancer Cell Studies. Cytotechnology 2007, 54, 135–143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Amer, L.D.; Holtzinger, A.; Keller, G.; Mahoney, M.J.; Bryant, S.J. Enzymatically Degradable Poly(Ethylene Glycol) Hydrogels for the 3D Culture and Release of Human Embryonic Stem Cell Derived Pancreatic Precursor Cell Aggregates. Acta Biomater. 2015, 22, 103–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Bray, L.J.; Binner, M.; Holzheu, A.; Friedrichs, J.; Freudenberg, U.; Hutmacher, D.W.; Werner, C. Multi-Parametric Hydrogels Support 3D in Vitro Bioengineered Microenvironment Models of Tumour Angiogenesis. Biomaterials 2015, 53, 609–620. [Google Scholar] [CrossRef] [Green Version]
  17. Debnath, J.; Brugge, J.S. Modelling Glandular Epithelial Cancers in Three-Dimensional Cultures. Nat. Rev. Cancer 2005, 5, 675–688. [Google Scholar] [CrossRef]
  18. Liu, C.; Liu, Y.; Xu, X.; Wu, H.; Xie, H.; Chen, L.; Lu, T.; Yang, L.; Guo, X.; Sun, G.; et al. Potential Effect of Matrix Stiffness on the Enrichment of Tumor Initiating Cells under Three-Dimensional Culture Conditions. Exp. Cell Res. 2015, 330, 123–134. [Google Scholar] [CrossRef]
  19. Kievit, F.M.; Florczyk, S.J.; Leung, M.C.; Wang, K.; Wu, J.D.; Silber, J.R.; Ellenbogen, R.G.; Lee, J.S.H.; Zhang, M. Proliferation and Enrichment of CD133+ Glioblastoma Cancer Stem Cells on 3D Chitosan-Alginate Scaffolds. Biomaterials 2014, 35, 9137–9143. [Google Scholar] [CrossRef] [Green Version]
  20. Feng, S.; Duan, X.; Lo, P.-K.; Liu, S.; Liu, X.; Chen, H.; Wang, Q. Expansion of Breast Cancer Stem Cells with Fibrous Scaffolds. Integr. Biol. 2013, 5, 768–777. [Google Scholar] [CrossRef] [Green Version]
  21. Moysidou, C.-M.; Barberio, C.; Owens, R.M. Advances in Engineering Human Tissue Models. Front. Bioeng. Biotechnol. 2021, 8, 1566. [Google Scholar] [CrossRef]
  22. Petersen, O.W.; Rønnov-Jessen, L.; Howlett, A.R.; Bissell, M.J. Interaction with Basement Membrane Serves to Rapidly Distinguish Growth and Differentiation Pattern of Normal and Malignant Human Breast Epithelial Cells. Proc. Natl. Acad. Sci. USA 1992, 89, 9064–9068. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Antoni, D.; Burckel, H.; Josset, E.; Noel, G. Three-Dimensional Cell Culture: A Breakthrough in Vivo. Int. J. Mol. Sci. 2015, 16, 5517–5527. [Google Scholar] [CrossRef]
  24. Redmond, J.; Mccarthy, H.; Buchanan, P.; Levingstone, T.J.; Dunne, N.J. Advances in Biofabrication Techniques for Collagen-Based 3D in Vitro Culture Models for Breast Cancer Research. Mater. Sci. Eng. C 2021, 122, 111944. [Google Scholar] [CrossRef] [PubMed]
  25. Horning, J.L.; Sahoo, S.K.; Vijayaraghavalu, S.; Dimitrijevic, S.; Vasir, J.K.; Jain, T.K.; Panda, A.K.; Labhasetwar, V. 3-D Tumor Model for in Vitro Evaluation of Anticancer Drugs. Mol. Pharm. 2008, 5, 849–862. [Google Scholar] [CrossRef]
  26. Kim, J.B. Three-Dimensional Tissue Culture Models in Cancer Biology. In Seminars in Cancer Biology; Academic Press: Cambridge, MA, USA, 2005; Volume 15, pp. 365–377. [Google Scholar] [CrossRef]
  27. Nyga, A.; Cheema, U.; Loizidou, M. 3D Tumour Models: Novel in Vitro Approaches to Cancer Studies. J. Cell Commun. Signal. 2011, 5, 239–248. [Google Scholar] [CrossRef] [Green Version]
  28. Debnath, J.; Muthuswamy, S.K.; Brugge, J.S. Morphogenesis and Oncogenesis of MCF-10A Mammary Epithelial Acini Grown in Three-Dimensional Basement Membrane Cultures. Methods 2003, 30, 256–268. [Google Scholar] [CrossRef]
  29. Debnath, J.; Mills, K.R.; Collins, N.L.; Reginato, M.J.; Muthuswamy, S.K.; Brugge, J.S. The Role of Apoptosis in Creating and Maintaining Luminal Space within Normal and Oncogene-Expressing Mammary Acini. Cell 2002, 111, 29–40. [Google Scholar] [CrossRef] [Green Version]
  30. Muthuswamy, S.K.; Li, D.; Lelievre, S.; Bissell, M.J.; Brugge, J.S. ErbB2, but Not ErbB1, Reinitiates Proliferation and Induces Luminal Repopulation in Epithelial Acini. Nat. Cell Biol. 2001, 3, 785–792. [Google Scholar] [CrossRef]
  31. Kumar, S.; Weaver, V.M. Mechanics, Malignancy, and Metastasis: The Force Journey of a Tumor Cell. Cancer Metastasis Rev. 2009, 28, 113–127. [Google Scholar] [CrossRef] [Green Version]
  32. Geiger, B.; Yamada, K.M. Molecular Architecture and Function of Matrix Adhesions. Cold Spring Harbor Perspect. Biol. 2011, 3, a005033. [Google Scholar] [CrossRef] [Green Version]
  33. Sawicki, L.A.; Choe, L.H.; Wiley, K.L.; Lee, K.H.; Kloxin, A.M. Isolation and Identification of Proteins Secreted by Cells Cultured within Synthetic Hydrogel-Based Matrices. ACS Biomater. Sci. Eng. 2018, 4, 836–845. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Malik, R.; Lelkes, P.I.; Cukierman, E. Biomechanical and Biochemical Remodeling of Stromal Extracellular Matrix in Cancer. Trends Biotechnol. 2015, 33, 230–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Tekle Abegaz, S. We Are IntechOpen, the World ’s Leading Publisher of Open Access Books Built by Scientists, for Scientists TOP 1%; University of Gondar Institutional Repository: Gondar, Ethiopia, 2021. [Google Scholar]
  36. Brown, N.H. Extracellular Matrix in Development: Insights from Mechanisms Conserved between Invertebrates and Vertebrates. Cold Spring Harbor Perspect. Biol. 2011, 3, a005082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Poltavets, V.; Kochetkova, M.; Pitson, S.M.; Samuel, M.S. The Role of the Extracellular Matrix and Its Molecular and Cellular Regulators in Cancer Cell Plasticity. Front. Oncol. 2018, 8, 1–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Ferreres, X.R.; Font, A.R.; Ibrahim, A.; Maximilien, N.; Lumbroso, D.; Hurford, A.; Winperry, J.; Wade, S.; Sataloff, R.T.; Johns, M.M.; et al. “We are IntechOpen, the world’s leading publisher of Open Access books Built by scientists, for scientists TOP 1 %,” Intech, vol. 32, no. July, pp. 137–144, 2013. Available online: http://www.intechopen.com/books/trends-in-telecommunications-technologies/gps-total-electron-content-tec-prediction-at-ionosphere-layer-over-the-equatorial-region%0AInTec%0Ahttp://www.asociatiamhc.ro/wp-content/uploads/2013/11/Guide-to-Hydropower.pdf (accessed on 27 October 2021).
  39. Frantz, C.; Stewart, K.M.; Weaver, V.M. The Extracellular Matrix at a Glance. J. Cell Sci. 2010, 123, 4195–4200. [Google Scholar] [CrossRef] [Green Version]
  40. Insua-rodríguez, J.; Oskarsson, T. The Extracellular Matrix in Breast Cancer. Adv. Drug Deliv. Rev. 2016, 97, 41–55. [Google Scholar] [CrossRef]
  41. Lu, P.; Takai, K.; Weaver, V.M.; Werb, Z. Extracellular Matrix Degradation and Remodeling in Development and Disease. Cold Spring Harbor Perspect. Biol. 2011, 3, 1–24. [Google Scholar] [CrossRef]
  42. Rozario, T.; DeSimone, D.W. The Extracellular Matrix in Development and Morphogenesis: A Dynamic View. Dev. Biol. 2010, 341, 126–140. [Google Scholar] [CrossRef] [Green Version]
  43. Soares, C.P.; Midlej, V.; de Oliveira, M.E.W.; Benchimol, M.; Costa, M.L.; Mermelstein, C. 2D and 3D-Organized Cardiac Cells Shows Differences in Cellular Morphology, Adhesion Junctions, Presence of Myofibrils and Protein Expression. PLoS ONE 2012, 7, e38147. [Google Scholar] [CrossRef] [Green Version]
  44. Lei, Y.; Schaffer, D.V. A Fully Defined and Scalable 3D Culture System for Human Pluripotent Stem Cell Expansion and Differentiation. Proc. Natl. Acad. Sci. USA 2013, 110, E5039–E5048. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Chopra, V.; Dinh, T.V.; Hannigan, E.V. Three-Dimensional Endothelial-Tumor Epithelial Cell Interactions in Human Cervical Cancers. In Vitro Cell. Dev. Biol.-Anim. 1997, 33, 432–442. [Google Scholar] [CrossRef] [PubMed]
  46. Lv, D.; Hu, Z.; Lu, L.; Lu, H.; Xu, X. Three-dimensional Cell Culture: A Powerful Tool in Tumor Research and Drug Discovery (Review). Oncol. Lett. 2017, 14, 6999–7010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Langhans, S.A. Three-Dimensional in Vitro Cell Culture Models in Drug Discovery and Drug Repositioning. Front. Pharmacol. 2018, 9, 6. [Google Scholar] [CrossRef] [PubMed]
  48. Li, Q.; Chow, A.B.; Mattingly, R.R. Three-Dimensional Overlay Culture Models of Human Breast Cancer Reveal a Critical Sensitivity to Mitogen-Activated Protein Kinase Kinase Inhibitors. J. Pharmacol. Exp. Ther. 2010, 332, 821–828. [Google Scholar] [CrossRef] [Green Version]
  49. Lee, G.Y.; Kenny, P.A.; Lee, E.H.; Bissell, M.J. Three-Dimensional Culture Models of Normal and Malignant Breast Epithelial Cells. Nat. Methods 2007, 4, 359–365. [Google Scholar] [CrossRef] [Green Version]
  50. Torisawa, Y.-S.; Shiku, H.; Yasukawa, T.; Nishizawa, M.; Matsue, T. Multi-Channel 3-D Cell Culture Device Integrated on a Silicon Chip for Anticancer Drug Sensitivity Test. Biomaterials 2005, 26, 2165–2172. [Google Scholar] [CrossRef]
  51. Li, Y.; Huang, G.; Li, M.; Wang, L.; Elson, E.L.; Lu, T.J.; Genin, G.M.; Xu, F. An Approach to Quantifying 3D Responses of Cells to Extreme Strain. Sci. Rep. 2016, 6, 19550. [Google Scholar] [CrossRef]
  52. Do Amaral, R.J.F.C.; Zayed, N.M.A.; Pascu, E.I.; Murphy, C.M.; Sridharan, R.; González-vázquez, A.; Sullivan, B.O. Functionalising Collagen-Based Scaffolds with Platelet-Rich Plasma for Enhanced Skin Wound Healing Potential. Front. Bioeng. Biotechnol. 2019, 7, 371. [Google Scholar] [CrossRef] [Green Version]
  53. Rosso, F.; Giordano, A.; Barbarisi, M.; Barbarisi, A. From Cell-ECM Interactions to Tissue Engineering. J. Cell. Physiol. 2004, 199, 174–180. [Google Scholar] [CrossRef]
  54. Yip, D.; Cho, C.H. A Multicellular 3D Heterospheroid Model of Liver Tumor and Stromal Cells in Collagen Gel for Anti-Cancer Drug Testing. Biochem. Biophys. Res. Commun. 2013, 433, 327–332. [Google Scholar] [CrossRef]
  55. Bokhari, M.; Carnachan, R.J.; Cameron, N.R.; Przyborski, S.A. Culture of HepG2 Liver Cells on Three Dimensional Polystyrene Scaffolds Enhances Cell Structure and Function during Toxicological Challenge. J. Anat. 2007, 211, 567–576. [Google Scholar] [CrossRef] [PubMed]
  56. Wang, F.; Weaver, V.M.; Petersen, O.W.; Larabell, C.A.; Dedhar, S.; Briand, P.; Lupu, R.; Bissell, M.J. Reciprocal Interactions between Beta1-Integrin and Epidermal Growth Factor Receptor in Three-Dimensional Basement Membrane Breast Cultures: A Different Perspective in Epithelial Biology. Proc. Natl. Acad. Sci. USA 1998, 95, 14821–14826. [Google Scholar] [CrossRef]
  57. Sokol, E.S.; Miller, D.H.; Breggia, A.; Spencer, K.C.; Arendt, L.M.; Gupta, P.B. Growth of Human Breast Tissues from Patient Cells in 3D Hydrogel Scaffolds. Breast Cancer Res. 2016, 18, 19. [Google Scholar] [CrossRef] [Green Version]
  58. Weaver, V.M.; Fischer, A.H.; Peterson, O.W.; Bissell, M.J. The Importance of the Microenvironment in Breast Cancer Progression: Recapitulation of Mammary Tumorigenesis Using a Unique Human Mammary Epithelial Cell Model and a Three-Dimensional Culture Assay. Biochem. Cell Biol. 1996, 74, 833–851. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Kenny, P.A.; Lee, G.Y.; Myers, C.A.; Neve, R.M.; Semeiks, J.R.; Spellman, P.T.; Lorenz, K.; Lee, E.H.; Barcellos-Hoff, M.H.; Petersen, O.W.; et al. The Morphologies of Breast Cancer Cell Lines in Three-Dimensional Assays Correlate with Their Profiles of Gene Expression. Mol. Oncol. 2007, 1, 84–96. [Google Scholar] [CrossRef] [PubMed]
  60. Järveläinen, H.; Sainio, A.; Koulu, M.; Wight, T.N.; Penttinen, R. Extracellular Matrix Molecules: Potential Targets in Pharmacotherapy. Pharmacol. Rev. 2009, 61, 198–223. [Google Scholar] [CrossRef] [Green Version]
  61. Bonnans, C.; Chou, J.; Werb, Z.; Bonnans, C.; Chou, J.; Werb, Z. Remodelling the extracellular matrix in development and disease. Nat. Rev. Mol. Cell Biol. 2014, 15, 786–801. [Google Scholar] [CrossRef]
  62. Lu, P.; Weaver, V.M.; Werb, Z. The Extracellular Matrix: A Dynamic Niche in Cancer Progression. J. Cell Biol. 2012, 196, 395–406. [Google Scholar] [CrossRef]
  63. Ricard-Blum, S. The Collagen Family. Cold Spring Harbor Perspect. Biol. 2011, 3, a004978. [Google Scholar] [CrossRef] [Green Version]
  64. Hashmi, S.; Marinkovich, M.P. Molecular Organization of the Basement Membrane Zone. Clin. Dermatol. 2011, 29, 398–411. [Google Scholar] [CrossRef]
  65. Hohenester, E.; Yurchenco, P.D. Laminins in Basement Membrane Assembly. Cell Adhes. Migr. 2013, 7, 56–63. [Google Scholar] [CrossRef] [Green Version]
  66. Mouw, J.K.; Ou, G.; Weaver, V.M. Extracellular Matrix Assembly: A Multiscale Deconstruction. Nat. Rev. Mol. Cell Biol. 2014, 15, 771–785. [Google Scholar] [CrossRef] [PubMed]
  67. Hynes, R.O.; Naba, A. Overview of the Matrisome-An Inventory of Extracellular Matrix Constituents and Functions. Cold Spring Harbor Perspect. Biol. 2012, 4, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Iozzo, R.V.; Schaefer, L. Proteoglycan Form and Function: A Comprehensive Nomenclature of Proteoglycans. Matrix Biol. 2015, 42, 11–55. [Google Scholar] [CrossRef]
  69. Schaefer, L.; Schaefer, R.M. Proteoglycans: From Structural Compounds to Signaling Molecules. Cell Tissue Res. 2010, 339, 237–246. [Google Scholar] [CrossRef] [PubMed]
  70. Roberts, K.; Alberts, B.; Johnson, A.; Lewis, J.; Morgan, D.; Raff, M.; Walter, P. Molecular Biology of the Cell, 6th ed.; Wilson, J., Hunt, T., Eds.; W.W. Norton & Company: New York, NY, USA, 2015; ISBN 0815340834. [Google Scholar] [CrossRef]
  71. Barsky, S.H.; Karlin, N.J. Myoepithelial Cells: Autocrine and Paracrine Suppressors of Breast Cancer Progression. J. Mammary Gland Biol. Neoplasia 2005, 10, 249–260. [Google Scholar] [CrossRef]
  72. Tsang, K.Y.; Cheung, M.C.H.; Chan, D.; Cheah, K.S.E. The Developmental Roles of the Extracellular Matrix: Beyond Structure to Regulation. Cell Tissue Res. 2010, 339, 93–110. [Google Scholar] [CrossRef] [Green Version]
  73. Huveneers, S.; Danen, E.H.J. Adhesion Signaling—Crosstalk between Integrins, Src and Rho. J. Cell Sci. 2009, 122, 1059–1069. [Google Scholar] [CrossRef] [Green Version]
  74. Discher, D.E.; Mooney, D.J.; Zandstra, P.W. Growth Factors, Matrices, and Forces Combine and Control Stem Cells. Science 2009, 324, 1673–1677. [Google Scholar] [CrossRef] [Green Version]
  75. Calvo, F.; Ege, N.; Grande-Garcia, A.; Hooper, S.; Jenkins, R.P.; Chaudhry, S.I.; Harrington, K.; Williamson, P.; Moeendarbary, E.; Charras, G.; et al. Mechanotransduction and YAP-Dependent Matrix Remodelling Is Required for the Generation and Maintenance of Cancer-Associated Fibroblasts. Nat. Cell Biol. 2013, 15, 637–646. [Google Scholar] [CrossRef] [PubMed]
  76. Varelas, X. The Hippo Pathway Effectors TAZ and YAP in Development, Homeostasis and Disease. Development 2014, 141, 1614–1626. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Bissell, M.J.; Radisky, D.C.; Rizki, A.; Weaver, V.M.; Petersen, O.W. The Organizing Principle: Microenvironmental Influences in the Normal and Malignant Breast. Differ. Res. Biol. Divers. 2002, 70, 537–546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Myllyharju, J.; Kivirikko, K.I. Collagens, Modifying Enzymes and Their Mutations in Humans, Flies and Worms. TIG 2004, 20, 33–43. [Google Scholar] [CrossRef] [PubMed]
  79. Ricard-Blum, S.; Ruggiero, F. The Collagen Superfamily: From the Extracellular Matrix to the Cell Membrane. Pathologie-Biologie 2005, 53, 430–442. [Google Scholar] [CrossRef]
  80. Camenisch, T.D.; Spicer, A.P.; Brehm-Gibson, T.; Biesterfeldt, J.; Augustine, M.L.; Calabro, A.J.; Kubalak, S.; Klewer, S.E.; McDonald, J.A. Disruption of Hyaluronan Synthase-2 Abrogates Normal Cardiac Morphogenesis and Hyaluronan-Mediated Transformation of Epithelium to Mesenchyme. J. Clin. Investig. 2000, 106, 349–360. [Google Scholar] [CrossRef] [Green Version]
  81. Chen, X.; Thibeault, S.L. Response of Fibroblasts to Transforming Growth Factor-Β1 on Two-Dimensional and in Three-Dimensional Hyaluronan Hydrogels. Tissue Eng. Part A 2012, 18, 2528–2538. [Google Scholar] [CrossRef] [Green Version]
  82. De Wever, O.; Demetter, P.; Mareel, M.; Bracke, M. Stromal Myofibroblasts Are Drivers of Invasive Cancer Growth. Int. J. Cancer 2008, 123, 2229–2238. [Google Scholar] [CrossRef]
  83. Hynes, R.O. The Extracellu. Science 2009, 326, 1216–1219. [Google Scholar] [CrossRef] [Green Version]
  84. Schaefer, L.; Iozzo, R.V. Biological Functions of the Small Leucine-Rich Proteoglycans: From Genetics to Signal Transduction. J. Biol. Chem. 2008, 283, 21305–21309. [Google Scholar] [CrossRef] [Green Version]
  85. Josefsson, A.; Adamo, H.; Hammarsten, P.; Granfors, T.; Stattin, P.; Egevad, L.; Laurent, A.E.; Wikström, P.; Bergh, A. Prostate Cancer Increases Hyaluronan in Surrounding Nonmalignant Stroma, and This Response Is Associated with Tumor Growth and an Unfavorable Outcome. Am. J. Pathol. 2011, 179, 1961–1968. [Google Scholar] [CrossRef] [PubMed]
  86. McAtee, C.O.; Barycki, J.J.; Simpson, M.A. Emerging Roles for Hyaluronidase in Cancer Metastasis and Therapy. Adv. Cancer Res. 2014, 123, 1–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Singh, P.; Carraher, C.; Schwarzbauer, J.E. Assembly of Fibronectin Extracellular Matrix. Annu. Rev. Cell Dev. Biol. 2010, 26, 397–419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Dallas, S.L.; Sivakumar, P.; Jones, C.J.P.; Chen, Q.; Peters, D.M.; Mosher, D.F.; Humphries, M.J.; Kielty, C.M. Fibronectin Regulates Latent Transforming Growth Factor-Beta (TGF Beta) by Controlling Matrix Assembly of Latent TGF Beta-Binding Protein-1. J. Biol. Chem. 2005, 280, 18871–18880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Sottile, J.; Hocking, D.C. Fibronectin Polymerization Regulates the Composition and Stability of Extracellular Matrix Fibrils and Cell-Matrix Adhesions. Mol. Biol. Cell 2002, 13, 3546–3559. [Google Scholar] [CrossRef] [Green Version]
  90. Dzamba, B.J.; Wu, H.; Jaenisch, R.; Peters, D.M. Fibronectin Binding Site in Type I Collagen Regulates Fibronectin Fibril Formation. J. Cell Biol. 1993, 121, 1165–1172. [Google Scholar] [CrossRef] [Green Version]
  91. Colombi, M.; Zoppi, N.; De Petro, G.; Marchina, E.; Gardella, R.; Tavian, D.; Ferraboli, S.; Barlati, S. Matrix Assembly Induction and Cell Migration and Invasion Inhibition by a 13-Amino Acid Fibronectin Peptide. J. Biol. Chem. 2003, 278, 14346–14355. [Google Scholar] [CrossRef] [Green Version]
  92. Domogatskaya, A.; Rodin, S.; Tryggvason, K. Functional Diversity of Laminins. Annu. Rev. Cell Dev. Biol. 2012, 28, 523–553. [Google Scholar] [CrossRef]
  93. Misra, S.; Hascall, V.C.; Markwald, R.R.; Ghatak, S. Interactions between Hyaluronan and Its Receptors (CD44, RHAMM) Regulate the Activities of Inflammation and Cancer. Front. Immunol. 2015, 6, 201. [Google Scholar] [CrossRef] [Green Version]
  94. Banerji, S.; Wright, A.J.; Noble, M.; Mahoney, D.J.; Campbell, I.D.; Day, A.J.; Jackson, D.G. Structures of the Cd44-Hyaluronan Complex Provide Insight into a Fundamental Carbohydrate-Protein Interaction. Nat. Struct. Mol. Biol. 2007, 14, 234–239. [Google Scholar] [CrossRef]
  95. Nelson, J.; McFerran, N.V.; Pivato, G.; Chambers, E.; Doherty, C.; Steele, D.; Timson, D.J. The 67 KDa Laminin Receptor: Structure, Function and Role in Disease. Biosci. Rep. 2008, 28, 33–48. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Harburger, D.S.; Calderwood, D.A. Integrin Signalling at a Glance. J. Cell Sci. 2009, 122, 159–163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Holen, I.; Nutter, F.; Wilkinson, J.M.; Evans, C.A.; Avgoustou, P.; Ottewell, P.D. Human Breast Cancer Bone Metastasis in Vitro and in Vivo: A Novel 3D Model System for Studies of Tumour Cell-Bone Cell Interactions. Clin. Exp. Metastasis 2015, 32, 689–702. [Google Scholar] [CrossRef]
  98. Egeblad, M.; Nakasone, E.S.; Werb, Z. Tumors as Organs: Complex Tissues That Interface with the Entire Organism. Dev. Cell 2010, 18, 884–901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Aumailley, M.; Gayraud, B. Structure and Biological Activity of the Extracellular Matrix. J. Mol. Med. 1998, 76, 253–265. [Google Scholar] [CrossRef] [PubMed]
  100. Cukierman, E.; Pankov, R.; Stevens, D.R.; Yamada, K.M. Taking Cell-Matrix Adhesions to the Third Dimension. Science 2001, 294, 1708–1712. [Google Scholar] [CrossRef]
  101. Sethi, T.; Rintoul, R.C.; Moore, S.M.; MacKinnon, A.C.; Salter, D.; Choo, C.; Chilvers, E.R.; Dransfield, I.; Donnelly, S.C.; Strieter, R.; et al. Extracellular Matrix Proteins Protect Small Cell Lung Cancer Cells against Apoptosis: A Mechanism for Small Cell Lung Cancer Growth and Drug Resistance in Vivo. Nat. Med. 1999, 5, 662–668. [Google Scholar] [CrossRef]
  102. Pal, A.; Kleer, C.G. Three Dimensional Cultures: A Tool to Study Normal Acinar Architecture vs. Malignant Transformation of Breast Cells. JoVE 2014, 86, 51311. [Google Scholar] [CrossRef]
  103. Espinoza-Sánchez, N.A.; Chimal-Ramírez, G.K.; Fuentes-Pananá, E.M. Analyzing the Communication Between Monocytes and Primary Breast Cancer Cells in an Extracellular Matrix Extract (ECME)-Based Three-Dimensional System. JoVE 2018, 131, 56589. [Google Scholar] [CrossRef]
  104. Knight, E.; Przyborski, S. Advances in 3D cell culture technologies enabling tissue-like structures to be created in vitro. J. Anat. 2015, 227, 746–756. [Google Scholar] [CrossRef] [Green Version]
  105. Moroni, L.; Burdick, J.A.; Highley, C.; Lee, S.J.; Morimoto, Y.; Takeuchi, S.; Yoo, J.J. Biofabrication Strategies for 3D in Vitro Models and Regenerative Medicine. Nat. Rev. Mater. 2021, 3, 21–37. [Google Scholar] [CrossRef]
  106. El-Sherbiny, I.M.; Yacoub, M.H. Hydrogel Scaffolds for Tissue Engineering: Progress and Challenges. Glob. Cardiol. Sci. Pract. 2013, 2013, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Benton, G.; Arnaoutova, I.; George, J.; Kleinman, H.K.; Koblinski, J. Matrigel: From Discovery and ECM Mimicry to Assays and Models for Cancer Research. Adv. Drug Deliv. Rev. 2014, 79–80, 3–18. [Google Scholar] [CrossRef] [PubMed]
  108. Baharvand, H.; Hashemi, S.M.; Ashtiani, S.K.; Farrokhi, A. Differentiation of Human Embryonic Stem Cells into Hepatocytes in 2D and 3D Culture Systems in Vitro. Int. J. Dev. Biol. 2006, 50, 645–652. [Google Scholar] [CrossRef] [Green Version]
  109. Rajakylä, K.; Krishnan, R.; Tojkander, S. Analysis of Contractility and Invasion Potential of Two Canine Mammary Tumor Cell Lines. Front. Vet. Sci. 2017, 4, 149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Mulfaul, K.; Giacalone, J.C.; Voigt, A.P.; Riker, M.J.; Ochoa, D.; Han, I.C.; Stone, E.M.; Mullins, R.F.; Tucker, B.A. Stepwise Differentiation and Functional Characterization of Human Induced Pluripotent Stem Cell-Derived Choroidal Endothelial Cells. Stem Cell Res. Ther. 2020, 11, 409. [Google Scholar] [CrossRef]
  111. Chen, D.; Qu, Y.; Hua, X.; Zhang, L.; Liu, Z.; Pflugfelder, S.C.; Li, D.-Q. A Hyaluronan Hydrogel Scaffold-Based Xeno-Free Culture System for Ex Vivo Expansion of Human Corneal Epithelial Stem Cells. Eye 2017, 31, 962–971. [Google Scholar] [CrossRef]
  112. Devarasetty, M.; Wang, E.; Soker, S.; Skardal, A. Mesenchymal Stem Cells Support Growth and Organization of Host-Liver Colorectal-Tumor Organoids and Possibly Resistance to Chemotherapy. Biofabrication 2017, 9, 21002. [Google Scholar] [CrossRef]
  113. Mannino, R.G.; Santiago-Miranda, A.N.; Pradhan, P.; Qiu, Y.; Mejias, J.C.; Neelapu, S.S.; Roy, K.; Lam, W.A. 3D Microvascular Model Recapitulates the Diffuse Large B-Cell Lymphoma Tumor Microenvironment in Vitro. Lab. Chip. 2017, 17, 407–414. [Google Scholar] [CrossRef]
  114. Chen, X.; Thibeault, S.L. Cell-Cell Interaction between Vocal Fold Fibroblasts and Bone Marrow Mesenchymal Stromal Cells in Three-Dimensional Hyaluronan Hydrogel. J. Tissue Eng. Regen. Med. 2016, 10, 437–446. [Google Scholar] [CrossRef] [Green Version]
  115. Engel, B.J.; Constantinou, P.E.; Sablatura, L.K.; Doty, N.J.; Carson, D.D.; Farach-Carson, M.C.; Harrington, D.A.; Zarembinski, T.I. Multilayered, Hyaluronic Acid-Based Hydrogel Formulations Suitable for Automated 3D High Throughput Drug Screening of Cancer-Stromal Cell Cocultures. Adv. Healthc. Mater. 2015, 4, 1664–1674. [Google Scholar] [CrossRef] [PubMed]
  116. Zhang, N.; Milleret, V.; Thompson-Steckel, G.; Huang, N.-P.; Vörös, J.; Simona, B.R.; Ehrbar, M. Soft Hydrogels Featuring In-Depth Surface Density Gradients for the Simple Establishment of 3D Tissue Models for Screening Applications. SLAS Discov. Adv. Life Sci. R&D 2017, 22, 635–644. [Google Scholar] [CrossRef] [Green Version]
  117. Saha, K.; Pollock, J.F.; Schaffer, D.V.; Healy, K.E. Designing synthetic materials to control stem cell phenotype. Curr. Opin. Chem. Biol. 2008, 11, 381–387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Seal, B.; Otero, T.C.; Panitch, A. Polymeric Biomaterials for Tissue and Organ Regeneration. Mater. Sci. Eng. R Rep. 2001, 34, 147–230. [Google Scholar] [CrossRef]
  119. Haycock, J.W. 3D Cell Culture: A Review of Current Approaches and Techniques. Methods Mol. Biol. 2011, 695, 1–15. [Google Scholar] [CrossRef]
  120. Sun, T.; Norton, D.; McKean, R.J.; Haycock, J.W.; Ryan, A.J.; MacNeil, S. Development of a 3D Cell Culture System for Investigating Cell Interactions with Electrospun Fibers. Biotechnol. Bioeng. 2007, 97, 1318–1328. [Google Scholar] [CrossRef]
  121. Aydin, H.M.; El Haj, A.J.; Pişkin, E.; Yang, Y. Improving Pore Interconnectivity in Polymeric Scaffolds for Tissue Engineering. J. Tissue Eng. Regen. Med. 2009, 3, 470–476. [Google Scholar] [CrossRef]
  122. Bokhari, M.; Carnachan, R.J.; Przyborski, A.; Cameron, N.R. Emulsion-Templated Porous Polymers as Scaffolds for Three Dimensional Cell Culture: Effect of Synthesis Parameters on Scaffold Formation and Homogeneity. J. Mater. Chem. 2007, 17, 4088–4094. [Google Scholar] [CrossRef]
  123. Salerno, A.; Oliviero, M.; Di Maio, E.; Iannace, S.; Netti, P.A. Design of Porous Polymeric Scaffolds by Gas Foaming of Heterogeneous Blends. J. Mater. Sci. Mater. Med. 2009, 20, 2043–2051. [Google Scholar] [CrossRef]
  124. Bai, C.; Yang, M.; Fan, Z.; Li, S.; Gao, T.; Fang, Z. Associations of Chemo- and Radio-Resistant Phenotypes with the Gap Junction, Adhesion and Extracellular Matrix in a Three-Dimensional Culture Model of Soft Sarcoma. J. Exp. Clin. Cancer Res. CR 2015, 34, 58. [Google Scholar] [CrossRef] [Green Version]
  125. Zhang, M.; Rose, B.; Lee, C.S.; Hong, A.M. In Vitro 3-Dimensional Tumor Model for Radiosensitivity of HPV Positive OSCC Cell Lines. Cancer Biol. Ther. 2015, 16, 1231–1240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Elisseeff, J. Hydrogels: Structure Starts to Gel. Nat. Mater. 2008, 7, 271–273. [Google Scholar] [CrossRef]
  127. Slaughter, B.V.; Khurshid, S.S.; Fisher, O.Z.; Khademhosseini, A.; Peppas, N.A. Hydrogels in Regenerative Medicine. Adv. Mater. 2009, 21, 3307–3329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Martínez-Ramos, C.; Lebourg, M. Three-Dimensional Constructs Using Hyaluronan Cell Carrier as a Tool for the Study of Cancer Stem Cells. J. Biomed. Mater. Research. Part B Appl. Biomater. 2015, 103, 1249–1257. [Google Scholar] [CrossRef] [PubMed]
  129. Hamdi, D.H.; Barbieri, S.; Chevalier, F.; Groetz, J.-E.; Legendre, F.; Demoor, M.; Galera, P.; Lefaix, J.-L.; Saintigny, Y. In Vitro Engineering of Human 3D Chondrosarcoma: A Preclinical Model Relevant for Investigations of Radiation Quality Impact. BMC Cancer 2015, 15, 579. [Google Scholar] [CrossRef] [Green Version]
  130. Tibbitt, M.W.; Anseth, K.S. Hydrogels as Extracellular Matrix Mimics for 3D Cell Culture. Biotechnol. Bioeng. 2009, 103, 655–663. [Google Scholar] [CrossRef] [Green Version]
  131. Fischbach, C.; Chen, R.; Matsumoto, T.; Schmelzle, T.; Brugge, J.S.; Polverini, P.J.; Mooney, D.J. Engineering Tumors with 3D Scaffolds. Nat. Methods 2007, 4, 855–860. [Google Scholar] [CrossRef]
  132. Worthington, P.; Pochan, D.J.; Langhans, S.A. Peptide Hydrogels—Versatile Matrices for 3D Cell Culture in Cancer Medicine. Front. Oncol. 2015, 5, 92. [Google Scholar] [CrossRef] [Green Version]
  133. Dhaliwal, A. 3D Cell Culture: A Review. Mater. Methods 2012, 2, 162. [Google Scholar] [CrossRef]
  134. Le, V.-M.; Lang, M.-D.; Shi, W.-B.; Liu, J.-W. A Collagen-Based Multicellular Tumor Spheroid Model for Evaluation of the Efficiency of Nanoparticle Drug Delivery. Artif. Cells Nanomed. Biotechnol. 2016, 44, 540–544. [Google Scholar] [CrossRef]
  135. Haisler, W.L.; Timm, D.M.; Gage, J.A.; Tseng, H.; Killian, T.C.; Souza, G.R. Three-Dimensional Cell Culturing by Magnetic Levitation. Nat. Protoc. 2013, 8, 1940–1949. [Google Scholar] [CrossRef] [PubMed]
  136. Lee, K.Y.; Mooney, D.J. Hydrogels for Tissue Engineering. Chem. Rev. 2001, 101, 1869–1879. [Google Scholar] [CrossRef]
  137. Hoffman, A.S. Hydrogels for Biomedical Applications. Adv. Drug Deliv. Rev. 2012, 64, 18–23. [Google Scholar] [CrossRef]
  138. Caliari, S.R.; Burdick, J.A. A Practical Guide to Hydrogels for Cell Culture. Nat. Methods 2016, 13, 405–414. [Google Scholar] [CrossRef] [Green Version]
  139. Domb, A.; Khan, W. Biodegradable Polymers as Drug Carrier Systems. In Encyclopedia of PHARMACEUTICAL TECHNOLOGY, 3rd ed.; Marcel Dekker: New York, NY, USA, 2013; Volume 1, pp. 135–176. ISBN 0824725387. [Google Scholar]
  140. Zhang, J.; Yun, S.; Du, Y.; Zannettino, A.C.W.; Zhang, H. Fabrication of a Cartilage Patch by Fusing Hydrogel-Derived Cell Aggregates onto Electrospun Film. Tissue Eng. Part A 2020, 26, 863–871. [Google Scholar] [CrossRef] [PubMed]
  141. Jeon, O.; Marks, R.; Wolfson, D.; Alsberg, E. Dual-Crosslinked Hydrogel Microwell System for Formation and Culture of Multicellular Human Adipose Tissue-Derived Stem Cell Spheroids. J. Mater. Chem. B 2016, 4, 3526–3533. [Google Scholar] [CrossRef] [PubMed]
  142. Pradhan, S.; Clary, J.M.; Seliktar, D.; Lipke, E.A. A Three-Dimensional Spheroidal Cancer Model Based on PEG-Fibrinogen Hydrogel Microspheres. Biomaterials 2017, 115, 141–154. [Google Scholar] [CrossRef]
  143. Wu, Y.; Zhao, Z.; Guan, Y.; Zhang, Y. Galactosylated Reversible Hydrogels as Scaffold for HepG2 Spheroid Generation. Acta Biomater. 2014, 10, 1965–1974. [Google Scholar] [CrossRef]
  144. Tseng, T.-C.; Wong, C.-W.; Hsieh, F.-Y.; Hsu, S. Biomaterial Substrate-Mediated Multicellular Spheroid Formation and Their Applications in Tissue Engineering. Biotechnol. J. 2017, 12, 1700064. [Google Scholar] [CrossRef]
  145. Lee, B.H.; Kim, M.H.; Lee, J.H.; Seliktar, D.; Cho, N.-J.; Tan, L.P. Modulation of Huh7.5 Spheroid Formation and Functionality Using Modified PEG-Based Hydrogels of Different Stiffness. PLoS ONE 2015, 10, e0118123. [Google Scholar] [CrossRef] [Green Version]
  146. Carvalho, M.P.; Costa, E.C.; Miguel, S.P.; Correia, I.J. Tumor Spheroid Assembly on Hyaluronic Acid-Based Structures: A Review. Carbohydr. Polym. 2016, 150, 139–148. [Google Scholar] [CrossRef] [PubMed]
  147. Yamada, Y.; Yoshida, C.; Hamada, K.; Kikkawa, Y.; Nomizu, M. Development of Three-Dimensional Cell Culture Scaffolds Using Laminin Peptide-Conjugated Agarose Microgels. Biomacromolecules 2020, 21, 3765–3771. [Google Scholar] [CrossRef]
  148. Barros, A.S.; Costa, E.C.; Nunes, A.S.; de Melo-Diogo, D.; Correia, I.J. Comparative Study of the Therapeutic Effect of Doxorubicin and Resveratrol Combination on 2D and 3D (Spheroids) Cell Culture Models. Int. J. Pharm. 2018, 551, 76–83. [Google Scholar] [CrossRef] [PubMed]
  149. Härmä, V.; Virtanen, J.; Mäkelä, R.; Happonen, A.; Mpindi, J.-P.; Knuuttila, M.; Kohonen, P.; Lötjönen, J.; Kallioniemi, O.; Nees, M. A Comprehensive Panel of Three-Dimensional Models for Studies of Prostate Cancer Growth, Invasion and Drug Responses. PLoS ONE 2010, 5, e10431. [Google Scholar] [CrossRef]
  150. Poincloux, R.; Collin, O.; Lizárraga, F.; Romao, M.; Debray, M.; Piel, M.; Chavrier, P. Contractility of the Cell Rear Drives Invasion of Breast Tumor Cells in 3D Matrigel. Proc. Natl. Acad. Sci. USA 2011, 108, 1943–1948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Dolega, M.E.; Abeille, F.; Picollet-D’hahan, N.; Gidrol, X. Controlled 3D Culture in Matrigel Microbeads to Analyze Clonal Acinar Development. Biomaterials 2015, 52, 347–357. [Google Scholar] [CrossRef]
  152. Chen, L.; Xiao, Z.; Meng, Y.; Zhao, Y.; Han, J.; Su, G.; Chen, B.; Dai, J. Biomaterials The Enhancement of Cancer Stem Cell Properties of MCF-7 Cells in 3D Collagen Scaffolds for Modeling of Cancer and Anti-Cancer Drugs. Biomaterials 2012, 33, 1437–1444. [Google Scholar] [CrossRef] [PubMed]
  153. Campbell, J.J.; Husmann, A.; Hume, R.D.; Watson, C.J.; Cameron, R.E. Biomaterials Development of Three-Dimensional Collagen Scaffolds with Controlled Architecture for Cell Migration Studies Using Breast Cancer Cell Lines. Biomaterials 2017, 114, 34–43. [Google Scholar] [CrossRef] [Green Version]
  154. Liverani, C.; De Vita, A.; Minardi, S.; Kang, Y.; Mercatali, L.; Bongiovanni, A.; La Manna, F.; Ibrahim, T.; Tasciotti, E. Open a Biomimetic 3D Model of Hypoxia-Driven Cancer Progression. Sci. Rep. 2019, 9, 12263. [Google Scholar] [CrossRef] [Green Version]
  155. Liverani, C.; Mercatali, L.; Cristofolini, L.; Giordano, E.; Minardi, S.; Della Porta, G.; De Vita, A.; Miserocchi, G.; Spadazzi, C.; Tasciotti, E.; et al. Investigating the Mechanobiology of Cancer Cell–ECM Interaction Through Collagen-Based 3D Scaffolds. Cell. Mol. Bioeng. 2017, 10, 223–234. [Google Scholar] [CrossRef]
  156. Egeblad, M.; Rasch, M.G.; Weaver, V.M. Dynamic Interplay between the Collagen Scaffold and Tumor Evolution. Curr. Opin. Cell Biol. 2010, 22, 697–706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. O’Connor, S.M.; Andreadis, J.D.; Shaffer, K.M.; Ma, W.; Pancrazio, J.J.; Stenger, D.A. Immobilization of Neural Cells in Three-Dimensional Matrices for Biosensor Applications. Biosens. Bioelectron. 2000, 14, 871–881. [Google Scholar] [CrossRef]
  158. Cross, V.L.; Zheng, Y.; Won Choi, N.; Verbridge, S.S.; Sutermaster, B.A.; Bonassar, L.J.; Fischbach, C.; Stroock, A.D. Dense Type I Collagen Matrices That Support Cellular Remodeling and Microfabrication for Studies of Tumor Angiogenesis and Vasculogenesis in Vitro. Biomaterials 2010, 31, 8596–8607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Suggs, L.J.; Mikos, A.G. Development of Poly(Propylene Fumarate-Co-Ethylene Glycol) as an Injectable Carrier for Endothelial Cells. Cell Transplant. 1999, 8, 345–350. [Google Scholar] [CrossRef] [PubMed]
  160. Model, D.; Cancer, B.; Subia, B.; Dey, T.; Sharma, S.; Kundu, S.C. Target Specific Delivery of Anticancer Drug in Silk Fibroin Based 3D Target Specific Delivery of Anticancer Drug in Silk Fibroin Based 3D Distribution Model of Bone−Breast Cancer Cells. ACS Appl. Mater. Interfaces 2015, 7, 2269–2279. [Google Scholar] [CrossRef]
  161. Talukdar, S.; Kundu, S.C. A Non-Mulberry Silk Fibroin Protein Based 3D In Vitro Tumor Model for Evaluation of Anticancer Drug Activity. Adv. Funct. Mater. 2012, 22, 4778–4788. [Google Scholar] [CrossRef]
  162. Talukdar, S.; Mandal, M.; Hutmacher, D.W.; Russell, P.J.; Soekmadji, C.; Kundu, S.C. Biomaterials Engineered Silk Fi Broin Protein 3D Matrices for in vitro Tumor Model. Biomaterials 2011, 32, 2149–2159. [Google Scholar] [CrossRef]
  163. Gurski, L.A.; Jha, A.K.; Zhang, C.; Jia, X.; Farach-Carson, M.C. Hyaluronic Acid-Based Hydrogels as 3D Matrices for in Vitro Evaluation of Chemotherapeutic Drugs Using Poorly Adherent Prostate Cancer Cells. Biomaterials 2009, 30, 6076–6085. [Google Scholar] [CrossRef] [Green Version]
  164. Wang, Y.; Mirza, S.; Wu, S.; Zeng, J.; Shi, W. 3D Hydrogel Breast Cancer Models for Studying the Effects of Hypoxia on Epithelial to Mesenchymal Transition. Oncotarget 2018, 9, 32191–32203. [Google Scholar] [CrossRef] [Green Version]
  165. Xu, W.; Qian, J.; Zhang, Y.; Suo, A.; Cui, N.; Wang, J.; Yao, Y.; Wang, H. Acta Biomaterialia A Double-Network Poly (Ne-Acryloyl L-Lysine)/Hyaluronic Acid Hydrogel as a Mimic of the Breast Tumor Microenvironment. Acta Biomater. 2016, 33, 131–141. [Google Scholar] [CrossRef]
  166. Fisher, S.A.; Anandakumaran, P.N.; Owen, S.C.; Shoichet, M.S. Tuning the Microenvironment: Click-Crosslinked Hyaluronic Acid-Based Hydrogels Provide a Platform for Studying Breast Cancer Cell Invasion. Adv. Funct. Mater. 2015, 25, 7163–7172. [Google Scholar] [CrossRef]
  167. Cavo, M.; Caria, M.; Pulsoni, I.; Beltrame, F.; Fato, M.; Scaglione, S. A New Cell-Laden 3D Alginate-Matrigel Hydrogel Resembles Human Breast Cancer Cell Malignant Morphology, Spread and Invasion Capability Observed “ in vivo”. Sci. Rep. 2018, 8, 5333. [Google Scholar] [CrossRef] [Green Version]
  168. Qiao, S.; Zhao, Y.; Li, C.; Yin, Y.; Meng, Q.; Lin, F.; Liu, Y.; Hou, X.; Guo, K.; Chen, X.; et al. Acta Biomaterialia An Alginate-Based Platform for Cancer Stem Cell Research. Acta Biomater. 2016, 37, 83–92. [Google Scholar] [CrossRef] [PubMed]
  169. Jiang, T.; Munguia-lopez, J.G.; Flores-torres, S.; Grant, J.; De Leon-rodriguez, A.; Kinsella, J.M. Directing the Self-Assembly of Tumour Spheroids by Bioprinting Cellular Heterogeneous Models within Alginate/Gelatin Hydrogels. Sci. Rep. 2017, 7, 4575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Arya, A.D.; Hallur, P.M.; Karkisaval, A.G.; Gudipati, A.; Rajendiran, S.; Dhavale, V.; Ramachandran, B.; Jayaprakash, A.; Gundiah, N.; Chaubey, A. Gelatin Methacrylate Hydrogels as Biomimetic Three-Dimensional Matrixes for Modeling Breast Cancer Invasion and Chemoresponse in Vitro. ACS Appl. Mater. Interfaces 2016, 8, 22005–22017. [Google Scholar] [CrossRef] [PubMed]
  171. Charmsaz, S.; Hughes, É.; Bane, F.T.; Tibbitts, P.; Mcilroy, M.; Byrne, C.; Cocchiglia, S.; Mcbryan, J.; Hennessy, B.T.; Dwyer, R.M.; et al. S100 β as a Serum Marker in Endocrine Resistant Breast Cancer. BMC Med. 2017, 15, 79. [Google Scholar] [CrossRef] [Green Version]
  172. Benton, G.; Kleinman, H.K.; George, J.; Arnaoutova, I. Multiple Uses of Basement Membrane-like Matrix (BME/Matrigel) In Vitro and In Vivo with Cancer Cells. Int. J. Cancer 2011, 128, 1751–1757. [Google Scholar] [CrossRef]
  173. Ghajar, C.M.; Bissell, M.J. Extracellular Matrix Control of Mammary Gland Morphogenesis and Tumorigenesis: Insights from Imaging. Histochem. Cell Biol. 2008, 130, 1105–1118. [Google Scholar] [CrossRef] [Green Version]
  174. Raub, C.B.; Unruh, J.; Suresh, V.; Krasieva, T.; Lindmo, T.; Gratton, E.; Tromberg, B.J.; George, S.C. Image Correlation Spectroscopy of Multiphoton Images Correlates with Collagen Mechanical Properties. Biophys. J. 2008, 94, 2361–2373. [Google Scholar] [CrossRef] [Green Version]
  175. Fridman, R.; Giaccone, G.; Kanemoto, T.; Martin, G.R.; Gazdar, A.F.; Mulshine, J.L. Reconstituted Basement Membrane (Matrigel) and Laminin Can Enhance the Tumorigenicity and the Drug Resistance of Small Cell Lung Cancer Cell Lines. Proc. Natl. Acad. Sci. 1990, 87, 6698–6702. [Google Scholar] [CrossRef] [Green Version]
  176. Xu, S.; Jia, Z.-F.; Kang, C.; Huang, Q.; Wang, G.; Liu, X.; Zhou, X.; Xu, P.; Pu, P. Upregulation of SEPT7 Gene Inhibits Invasion of Human Glioma Cells. Cancer Investig. 2010, 28, 248–258. [Google Scholar] [CrossRef]
  177. Arnaoutova, I.; George, J.; Kleinman, H.K.; Benton, G. The Endothelial Cell Tube Formation Assay on Basement Membrane Turns 20: State of the Science and the Art. Angiogenesis 2009, 12, 267–274. [Google Scholar] [CrossRef]
  178. Manuscript, A. Modular Extracellular Matrices: Solutions for the Puzzle. Methods 2009, 45, 93–98. [Google Scholar]
  179. Jafari, M.; Paknejad, Z.; Rad, M.R.; Motamedian, S.R.; Eghbal, M.J.; Nadjmi, N.; Khojasteh, A. Polymeric Scaffolds in Tissue Engineering: A Literature Review. J. Biomed. Mater. Res. Part B Appl. Biomater. 2017, 105, 431–459. [Google Scholar] [CrossRef] [PubMed]
  180. Nagaoka, M.; Jiang, H.L.; Hoshiba, T.; Akaike, T.; Cho, C.S. Application of Recombinant Fusion Proteins for Tissue Engineering. Ann. Biomed. Eng. 2010, 38, 683–693. [Google Scholar] [CrossRef] [PubMed]
  181. Kapoor, S.; Kundu, S.C. Acta Biomaterialia Silk Protein-Based Hydrogels: Promising Advanced Materials for Biomedical Applications. Acta Biomater. 2016, 31, 17–32. [Google Scholar] [CrossRef]
  182. Conklin, M.W.; Keely, P.J. Why the stroma matters in breast cancer. Cell Adhes. Migr. 2012, 6, 249–260. [Google Scholar] [CrossRef] [Green Version]
  183. Acerbi Cassereau, I.; Weaver, V.M. Human breast cancer invasion and aggression correlates with ECM stiffening and immune cell infiltration. Integrative Biology 2015, 7, 1120–1134. [Google Scholar] [CrossRef] [Green Version]
  184. Singh, D.; Singh, D.; Han, S.S. 3D Printing of Scaffold for Cells Delivery: Advances in Skin Tissue Engineering. Polymers 2016, 8, 19. [Google Scholar] [CrossRef]
  185. Levental, K.R.; Yu, H.; Kass, L.; Lakins, J.N.; Egeblad, M.; Erler, J.T.; Fong, S.F.T.; Csiszar, K.; Giaccia, A.; Weninger, W.; et al. Matrix Crosslinking Forces Tumor Progression by Enhancing Integrin Signaling. Cell 2009, 139, 891–906. [Google Scholar] [CrossRef] [Green Version]
  186. Fang, M.; Yuan, J.; Peng, C.; Li, Y. Collagen as a Double-Edged Sword in Tumor Progression. Tumor Biol. 2014, 35, 2871–2882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Gelse, K.; Pöschl, E.; Aigner, T. Collagens-Structure, Function, and Biosynthesis. Adv. Drug Deliv. Rev. 2003, 55, 1531–1546. [Google Scholar] [CrossRef] [Green Version]
  188. Gordon, M.K.; Hahn, R.A. Collagens. Cell Tissue Res. 2009, 339, 247. [Google Scholar] [CrossRef] [PubMed]
  189. Shoulders, M.D.; Raines, R.T. Collagen Structure and Stability. Annu. Rev. Biochem. 2009, 78, 929–958. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Ramtani, S.; Helary, C. Mechanical Behavior under Unconfined Compression Loading of Dense Fibrillar Collagen Matrices Mimetic Living Tissues Compression Loadings of Dense. J. Mech. Med. Biol. 2010, 10, 35–55. [Google Scholar] [CrossRef]
  191. Tierney, C.M.; Haugh, M.G.; Liedl, J.; Mulcahy, F.; Hayes, B.; O’Brien, F.J. The Effects of Collagen Concentration and Crosslink Density on the Biological, Structural and Mechanical Properties of Collagen-GAG Scaffolds for Bone Tissue Engineering. J. Mech. Behav. Biomed. Mater. 2009, 2, 202–209. [Google Scholar] [CrossRef]
  192. Ahmed, E.M. Hydrogel: Preparation, Characterization, and Applications: A Review. J. Adv. Res. 2015, 6, 105–121. [Google Scholar] [CrossRef] [Green Version]
  193. Antoine, E.E.; Vlachos, P.P.; Rylander, M.N. Review of Collagen I Hydrogels for Bioengineered Tissue Microenvironments: Characterization of Mechanics, Structure, and Transport. Tissue Eng. Part B Rev. 2014, 20, 683–696. [Google Scholar] [CrossRef] [Green Version]
  194. Martin, R.; Farjanel, J.; Eichenberger, D.; Colige, A.; Kessler, E.; Hulmes, D.J.S.; Giraud-Guille, M.M. Liquid Crystalline Ordering of Procollagen as a Determinant of Three-Dimensional Extracellular Matrix Architecture. J. Mol. Biol. 2000, 301, 11–17. [Google Scholar] [CrossRef] [Green Version]
  195. Achilli, M.; Mantovani, D. Tailoring Mechanical Properties of Collagen-Based Scaffolds for Vascular Tissue Engineering: The Effects of PH, Temperature and Ionic Strength on Gelation. Polymers 2010, 2, 664–680. [Google Scholar] [CrossRef] [Green Version]
  196. Hyllested, J.L.; Veje, K.; Ostergaard, K. Histochemical Studies of the Extracellular Matrix of Human Articular Cartilage—A Review. Osteoarthr. Cartil. 2002, 10, 333–343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Cox, R.F.; Jenkinson, A.; Pohl, K.; O’Brien, F.J.; Morgan, M.P. Osteomimicry of Mammary Adenocarcinoma Cells In Vitro; Increased Expression of Bone Matrix Proteins and Proliferation within a 3D Collagen Environment. PLoS ONE 2012, 7, e41679. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. James-Bhasin, M.; Siegel, P.M.; Nazhat, S.N. A Three-Dimensional Dense Collagen Hydrogel to Model Cancer Cell/Osteoblast Interactions. J. Funct. Biomater. 2018, 9, 72. [Google Scholar] [CrossRef] [Green Version]
  199. Tom, E.; Jonathan, B. Collagen substrata for studies on cell behavior. J. Cell Biol. 1972, 54, 626–637. [Google Scholar] [CrossRef]
  200. Folkman, J.; Haudenschild, C. Angiogenesis in Vitro. Nature 1980, 288, 551–556. [Google Scholar] [CrossRef] [PubMed]
  201. Krause, S.; Maffini, M.V.; Soto, A.M.; Sonnenschein, C. The Microenvironment Determines the Breast Cancer Cells’ Phenotype: Organization of MCF7 Cells in 3D Cultures. BMC Cancer 2010, 10, 263. [Google Scholar] [CrossRef] [Green Version]
  202. Millerot-Serrurot, E.; Guilbert, M.; Fourré, N.; Witkowski, W.; Said, G.; Van Gulick, L.; Terryn, C.; Zahm, J.-M.; Garnotel, R.; Jeannesson, P. 3D Collagen Type I Matrix Inhibits the Antimigratory Effect of Doxorubicin. Cancer Cell Int. 2010, 10, 26. [Google Scholar] [CrossRef] [Green Version]
  203. Patterson, D.M.; Nazarova, L.A.; Prescher, J.A. Finding the Right (Bioorthogonal) Chemistry. ACS Chem. Biol. 2014, 9, 592–605. [Google Scholar] [CrossRef]
  204. Lutolf, M.P.; Hubbell, J.A. Synthetic Biomaterials as Instructive Extracellular Microenvironments for Morphogenesis in Tissue Engineering. Nat. Biotechnol. 2005, 23, 47–55. [Google Scholar] [CrossRef]
  205. Lokeshwar, V.B.; Cerwinka, W.H.; Isoyama, T.; Lokeshwar, B.L. HYAL1 Hyaluronidase in Prostate Cancer: A Tumor Promoter and Suppressor. Cancer Res. 2005, 65, 7782–7789. [Google Scholar] [CrossRef] [Green Version]
  206. Murphy, A.R.; Laslett, A.; O’Brien, C.M.; Cameron, N.R. Scaffolds for 3D in Vitro Culture of Neural Lineage Cells. Acta Biomater. 2017, 54, 1–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Burdick, J.A.; Anseth, K.S. Photoencapsulation of Osteoblasts in Injectable RGD-Modified PEG Hydrogels for Bone Tissue Engineering. Biomaterials 2002, 23, 4315–4323. [Google Scholar] [CrossRef]
  208. Burdick, J.A.; Mason, M.N.; Hinman, A.D.; Thorne, K.; Anseth, K.S. Delivery of Osteoinductive Growth Factors from Degradable PEG Hydrogels Influences Osteoblast Differentiation and Mineralization. J. Control. Release 2002, 83, 53–63. [Google Scholar] [CrossRef]
  209. Bryant, S.J.; Durand, K.L.; Anseth, K.S. Manipulations in Hydrogel Chemistry Control Photoencapsulated Chondrocyte Behavior and Their Extracellular Matrix Production. J. Biomed. Mater. Res. Part A 2003, 67, 1430–1436. [Google Scholar] [CrossRef]
  210. Bryant, S.J.; Anseth, K.S. Hydrogel Properties Influence ECM Production by Chondrocytes Photoencapsulated in Poly(Ethylene Glycol) Hydrogels. J. Biomed. Mater. Res. 2002, 59, 63–72. [Google Scholar] [CrossRef] [PubMed]
  211. Adelöw, C.; Segura, T.; Hubbell, J.A.; Frey, P. The Effect of Enzymatically Degradable Poly(Ethylene Glycol) Hydrogels on Smooth Muscle Cell Phenotype. Biomaterials 2008, 29, 314–326. [Google Scholar] [CrossRef]
  212. Strutz, F.; Zeisberg, M.; Renziehausen, A.; Raschke, B.; Becker, V.; van Kooten, C.; Müller, G. TGF-Beta 1 Induces Proliferation in Human Renal Fibroblasts via Induction of Basic Fibroblast Growth Factor (FGF-2). Kidney Int. 2001, 59, 579–592. [Google Scholar] [CrossRef] [Green Version]
  213. Engler, A.J.; Sen, S.; Sweeney, H.L.; Discher, D.E. Matrix Elasticity Directs Stem Cell Lineage Specification. Cell 2006, 126, 677–689. [Google Scholar] [CrossRef] [Green Version]
  214. Hayward, A.S.; Sano, N.; Przyborski, S.A.; Cameron, N.R. Acrylic-Acid-Functionalized PolyHIPE Scaffolds for Use in 3D Cell Culture. Macromol. Rapid Commun. 2013, 34, 1844–1849. [Google Scholar] [CrossRef]
  215. Brown, T.E.; Anseth, K.S. Spatiotemporal Hydrogel Biomaterials for Regenerative Medicine. Chem. Soc. Rev. 2017, 46, 6532–6552. [Google Scholar] [CrossRef]
  216. Sell, S.A.; Wolfe, P.S.; Garg, K.; McCool, J.M.; Rodriguez, I.A.; Bowlin, G.L. The Use of Natural Polymers in Tissue Engineering: A Focus on Electrospun Extracellular Matrix Analogues. Polymers 2010, 2, 522–553. [Google Scholar] [CrossRef]
  217. Liu, J.; He, X.; Corbett, S.A.; Lowry, S.F.; Graham, A.M.; Fässler, R.; Li, S. Integrins Are Required for the Differentiation of Visceral Endoderm. J. Cell Sci. 2009, 122, 233–242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. El-Sherbiny, I.M.; Smyth, H.D.C. Poly(Ethylene Glycol)-Carboxymethyl Chitosan-Based PH-Responsive Hydrogels: Photo-Induced Synthesis, Characterization, Swelling, and in Vitro Evaluation as Potential Drug Carriers. Carbohydr. Res. 2010, 345, 2004–2012. [Google Scholar] [CrossRef]
  219. Davidenko, N.; Campbell, J.J.; Thian, E.S.; Watson, C.J.; Cameron, R.E. Acta Biomaterialia Collagen–Hyaluronic Acid Scaffolds for Adipose Tissue Engineering. Acta Biomater. 2010, 6, 3957–3968. [Google Scholar] [CrossRef] [PubMed]
  220. Pluen, A.; Boucher, Y.; Ramanujan, S.; McKee, T.D.; Gohongi, T.; di Tomaso, E.; Brown, E.B.; Izumi, Y.; Campbell, R.B.; Berk, D.A.; et al. Role of Tumor-Host Interactions in Interstitial Diffusion of Macromolecules: Cranial vs. Subcutaneous Tumors. Proc. Natl. Acad. Sci. USA 2001, 98, 4628–4633. [Google Scholar] [CrossRef] [Green Version]
  221. Toole, B.P. Hyaluronan Promotes the Malignant Phenotype. Glycobiology 2002, 12, 37R–42R. [Google Scholar] [CrossRef] [Green Version]
  222. Gallagher, J.T. Heparan Sulfate: Growth Control with a Restricted Sequence Menu. J. Clin. Investig. 2001, 108, 357–361. [Google Scholar] [CrossRef]
  223. Guimond, S.; Maccarana, M.; Olwin, B.B.; Lindahl, U.; Rapraeger, A.C. Activating and Inhibitory Heparin Sequences for FGF-2 (Basic FGF). Distinct Requirements for FGF-1, FGF-2, and FGF-4. J. Biol. Chem. 1993, 268, 23906–23914. [Google Scholar] [CrossRef]
  224. Nikitovic, D.; Assouti, M.; Sifaki, M.; Katonis, P.; Krasagakis, K.; Karamanos, N.K.; Tzanakakis, G.N. Chondroitin Sulfate and Heparan Sulfate-Containing Proteoglycans Are Both Partners and Targets of Basic Fibroblast Growth Factor-Mediated Proliferation in Human Metastatic Melanoma Cell Lines. Int. J. Biochem. Cell Biol. 2008, 40, 72–83. [Google Scholar] [CrossRef]
  225. Toole, B.P. Hyaluronan: From Extracellular Glue to Pericellular Cue. Nat. Rev. Cancer 2004, 4, 528–539. [Google Scholar] [CrossRef]
  226. Venning, F.A.; Wullkopf, L.; Erler, J.T. Targeting ECM Disrupts Cancer Progression. Front. Oncol. 2015, 5, 1–15. [Google Scholar] [CrossRef] [Green Version]
  227. Campbell, J.C.; Botos, L.A.; Sargeant, T.J.; Davidenko, N.; Cameron, R.E.; Watson, C.J. A 3-D in vitro co-culture model of mammary gland involution. Integr. Biol (Camb). 2014, 6, 618–626. [Google Scholar] [CrossRef] [PubMed]
  228. Laboratory for Synthetic Biology, RIKEN Quantitative Biology Center, Department of Systems Pharmacology, UTokyo Graduate School of Medicine. The CUBIC Clearing Protocol with Reagent-1A (for Whole Mouse Brain). 2015. Available online: http://cubic.riken.jp/data/CUBIC_clearing_protocol_with_Reagent-1A.pdf (accessed on 27 October 2021).
  229. Monteiro, J.; Gaspar, C.; Richer, W.; Franken, P.F.; Sacchetti, A.; Joosten, R.; Idali, A.; Brandao, J.; Decraene, C.; Fodde, R. Cancer Stemness in Wnt-Driven Mammary Tumorigenesis. Carcinogenesis 2014, 35, 2–13. [Google Scholar] [CrossRef] [Green Version]
  230. Hemmrich, K.; Von Heimburg, D.; Rendchen, R.; Di, C.; Milella, E.; Pallua, N. Implantation of Preadipocyte-Loaded Hyaluronic Acid-Based Scaffolds into Nude Mice to Evaluate Potential for Soft Tissue Engineering. Biomaterials 2005, 26, 7025–7037. [Google Scholar] [CrossRef] [PubMed]
  231. Flynn, L.; Prestwich, G.D.; Semple, J.L.; Woodhouse, K.A. Adipose Tissue Engineering in Vivo with Adipose-Derived Stem Cells on Naturally Derived Scaffolds. J. Biomed. Mater. Res. Part A 2009, 89, 929–941. [Google Scholar] [CrossRef] [PubMed]
  232. Flynn, L.; Prestwich, G.D.; Semple, J.L.; Woodhouse, K.A. Adipose Tissue Engineering with Naturally Derived Scaffolds and Adipose-Derived Stem Cells. Biomaterials 2007, 28, 3834–3842. [Google Scholar] [CrossRef] [PubMed]
  233. Halbleib, M.; Skurk, T.; De Luca, C.; Von Heimburg, D.; Hauner, H. Tissue Engineering of White Adipose Tissue Using Hyaluronic Acid-Based Scaffolds. I: In Vitro Differentiation of Human Adipocyte Precursor Cells on Scaffolds. Biomaterials 2003, 24, 3125–3132. [Google Scholar] [CrossRef]
  234. Offeddu, G.S.; Ashworth, J.C.; Cameron, R.E.; Oyen, M.L. Multi-Scale Mechanical Response of Freeze-Dried Collagen Scaffolds for Tissue Engineering Applications. J. Mech. Behav. Biomed. Mater. 2015, 42, 19–25. [Google Scholar] [CrossRef] [Green Version]
  235. Hume, R.D.; Pensa, S.; Brown, E.J.; Kreuzaler, P.A.; Hitchcock, J.; Husmann, A.; Campbell, J.J.; Lloyd-thomas, A.O.; Cameron, R.E.; Watson, C.J. Tumour Cell Invasiveness and Response to Chemotherapeutics in Adipocyte Invested 3D Engineered Anisotropic Collagen Scaffolds. Sci. Rep. 2018, 1–15. [Google Scholar] [CrossRef] [Green Version]
  236. Wondraczek, H.; Elschner, T.; Heinze, T. Synthesis of Highly Functionalized Dextran Alkyl Carbonates Showing Nanosphere Formation. Carbohydr. Polym. 2011, 83, 1112–1118. [Google Scholar] [CrossRef]
  237. Nowakowska, M.; Zapotoczny, S.; Sterzel, M.; Kot, E. Novel Water-Soluble Photosensitizers from Dextrans. Biomacromolecules 2004, 5, 1009–1014. [Google Scholar] [CrossRef] [PubMed]
  238. Cao, B.; Li, L.; Wu, H.; Tang, Q.; Sun, B.; Dong, H.; Zhe, J.; Cheng, G. Zwitteration of Dextran: A Facile Route to Integrate Antifouling, Switchability and Optical Transparency into Natural Polymers. Chem. Commun. 2014, 50, 3234–3237. [Google Scholar] [CrossRef] [PubMed]
  239. Perrino, C.; Lee, S.; Choi, S.W.; Maruyama, A.; Spencer, N.D. A Biomimetic Alternative to Poly(Ethylene Glycol) as an Antifouling Coating: Resistance to Nonspecific Protein Adsorption of Poly(L-Lysine)-Graft-Dextran. Langmuir 2008, 24, 8850–8856. [Google Scholar] [CrossRef]
  240. Zhang, X.; Yang, Y.; Yao, J.; Shao, Z.; Chen, X. Strong Collagen Hydrogels by Oxidized Dextran Modification. ACS Sustain. Chem. Eng. 2014, 2, 1318–1324. [Google Scholar] [CrossRef]
  241. Berillo, D.; Elowsson, L.; Kirsebom, H. Oxidized Dextran as Crosslinker for Chitosan Cryogel Scaffolds and Formation of Polyelectrolyte Complexes between Chitosan and Gelatin. Macromol. Biosci. 2012, 12, 1090–1099. [Google Scholar] [CrossRef]
  242. Wang, T.; Nie, J.; Yang, D. Dextran and Gelatin Based Photocrosslinkable Tissue Adhesive. Carbohydr. Polym. 2012, 90, 1428–1436. [Google Scholar] [CrossRef] [PubMed]
  243. Hoffmann, B.; Seitz, D.; Mencke, A.; Kokott, A.; Ziegler, G. Glutaraldehyde and Oxidised Dextran as Crosslinker Reagents for Chitosan-Based Scaffolds for Cartilage Tissue Engineering. J. Mater. Sci. Mater. Med. 2009, 20, 1495–1503. [Google Scholar] [CrossRef]
  244. Khor, E.; Lim, L.Y. Implantable Applications of Chitin and Chitosan. Biomaterials 2003, 24, 2339–2349. [Google Scholar] [CrossRef]
  245. Dash, M.; Federica, C.; Ottenbrite, R.; Chiellini, E. Chitosan—A Versatile Semi-Synthetic Polymer in Biomedical Applications. Prog. Polym. Sci. 2011, 36, 981–1014. [Google Scholar] [CrossRef]
  246. Madihally, S.V.; Matthew, H.W. Porous Chitosan Scaffolds for Tissue Engineering. Biomaterials 1999, 20, 1133–1142. [Google Scholar] [CrossRef]
  247. Hornof, M.D.; Kast, C.E.; Bernkop-Schnürch, A. In Vitro Evaluation of the Viscoelastic Properties of Chitosan-Thioglycolic Acid Conjugates. Eur. J. Pharm. Biopharm. 2003, 55, 185–190. [Google Scholar] [CrossRef]
  248. Loscalzo, D.E.H.R.C.J. NIH Public Access. Bone 2011, 23, 1–7. [Google Scholar] [CrossRef]
  249. Rabionet, M.; Yeste, M. Electrospinning PCL Scaffolds Manufacture for Three-Dimensional Breast Cancer Cell Culture. Polymers 2017, 9, 328. [Google Scholar] [CrossRef] [PubMed]
  250. Sundaresh, S.; Chu, B.; Hadjiargyrou, M. Cdk2 Silencing via a DNA/PCL Electrospun Scaffold Suppresses Proliferation and Increases Death of Breast Cancer Cells. PLoS ONE 2012, 7, e52356. [Google Scholar] [CrossRef] [Green Version]
  251. Sims-Mourtada, J.; Niamat, R.A.; Samuel, S.; Eskridge, C.; Kmiec, E.B. Enrichment of Breast Cancer Stem-like Cells by Growth on Electrospun Polycaprolactone-Chitosan Nanofiber Scaffolds. Int. J. Nanomed. 2014, 9, 995–1003. [Google Scholar] [CrossRef] [Green Version]
  252. Ampuja, M.; Jokimäki, R.; Juuti-uusitalo, K.; Rodriguez-martinez, A.; Alarmo, E.; Kallioniemi, A. BMP4 Inhibits the Proliferation of Breast Cancer Cells and Induces an MMP-Dependent Migratory Phenotype in MDA-MB-231 Cells in 3D Environment. BMC Cancer 2013, 13, 429. [Google Scholar] [CrossRef] [Green Version]
  253. Pradhan, S.; Hassani, I.; Seeto, W.J.; Lipke, E.A. PEG-Fibrinogen Hydrogels for Three-Dimensional Breast Cancer Cell Culture. J. Biomed. Mater. Res. Part A 2016, 105, 236–252. [Google Scholar] [CrossRef]
  254. Chang, F.-C.; Tsao, C.-T.; Lin, A.; Zhang, M.; Levengood, S.L.; Zhang, M. PEG-Chitosan Hydrogel with Tunable Stiffness for Study of Drug Response of Breast Cancer Cells. Polymers 2016, 8, 112. [Google Scholar] [CrossRef] [Green Version]
  255. Pathi, S.P.; Lin, D.D.W.; Dorvee, J.R.; Estroff, L.A.; Fischbach, C. Biomaterials Hydroxyapatite Nanoparticle-Containing Scaffolds for the Study of Breast Cancer Bone Metastasis. Biomaterials 2011, 32, 5112–5122. [Google Scholar] [CrossRef] [Green Version]
  256. Bryant, S.J.; Anseth, K.S.; Lee, D.A.; Bader, D.L. Crosslinking Density Influences the Morphology of Chondrocytes Photoencapsulated in PEG Hydrogels during the Application of Compressive Strain. J. Orthop. Res. 2004, 22, 1143–1149. [Google Scholar] [CrossRef]
  257. Bryant, S.J.; Chowdhury, T.T.; Lee, D.A.; Bader, D.L.; Anseth, K.S. Crosslinking Density Influences Chondrocyte Metabolism in Dynamically Loaded Photocrosslinked Poly(Ethylene Glycol) Hydrogels. Ann. Biomed. Eng. 2004, 32, 407–417. [Google Scholar] [CrossRef]
  258. Levenberg, S.; Huang, N.F.; Lavik, E.; Rogers, A.B.; Itskovitz-Eldor, J.; Langer, R. Differentiation of Human Embryonic Stem Cells on Three-Dimensional Polymer Scaffolds. Proc. Natl. Acad. Sci. USA 2003, 100, 12741–12746. [Google Scholar] [CrossRef] [Green Version]
  259. Mikos, A.G.; Sarakinos, G.; Leite, S.M.; Vacant, J.P.; Langer, R. Laminated Three-Dimensional Biodegradable Foams for Use in Tissue Engineering. Biomaterials 1993, 14, 323–330. [Google Scholar] [CrossRef]
  260. Ouyang, A.; Ng, R.; Yang, S.-T. Long-Term Culturing of Undifferentiated Embryonic Stem Cells in Conditioned Media and Three-Dimensional Fibrous Matrices Without Extracellular Matrix Coating. Stem Cells 2007, 25, 447–454. [Google Scholar] [CrossRef]
  261. Girard, Y.K.; Wang, C.; Ravi, S.; Howell, M.C.; Mallela, J.; Alibrahim, M.; Green, R.; Hellermann, G.; Mohapatra, S.S.; Mohapatra, S. A 3D Fibrous Scaffold Inducing Tumoroids: A Platform for Anticancer Drug Development. PLoS ONE 2013, 8, e75345. [Google Scholar] [CrossRef] [Green Version]
  262. Vasita, R.; Katti, D.S. Nanofibers and Their Applications in Tissue Engineering. Int. J. Nanomed. 2006, 1, 15–30. [Google Scholar] [CrossRef]
  263. Glass-Brudzinski, J.; Perizzolo, D.; Brunette, D.M. Effects of Substratum Surface Topography on the Organization of Cells and Collagen Fibers in Collagen Gel Cultures. J. Biomed. Mater. Res. 2002, 61, 608–618. [Google Scholar] [CrossRef]
  264. Karuri, N.W.; Liliensiek, S.; Teixeira, A.I.; Abrams, G.; Campbell, S.; Nealey, P.F.; Murphy, C.J. Biological Length Scale Topography Enhances Cell-Substratum Adhesion of Human Corneal Epithelial Cells. J. Cell Sci. 2004, 117, 3153–3164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Dalby, M.J. Cellular Response to Low Adhesion Nanotopographies. Int. J. Nanomed. 2007, 2, 373–381. [Google Scholar]
  266. Maltman, D.J.; Przyborski, S.A. Developments in Three-Dimensional Cell Culture Technology Aimed at Improving the Accuracy of in Vitro Analyses. Biochem. Soc. Trans. 2010, 38, 1072–1075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Knight, E.; Murray, B.; Carnachan, R.; Przyborski, S. Alvetex®: Polystyrene Scaffold Technology for Routine Three Dimensional Cell Culture. Methods Mol. Biol. 2011, 695, 323–340. [Google Scholar] [CrossRef] [PubMed]
  268. Hayman, M.W.; Smith, K.H.; Cameron, N.R.; Przyborski, S.A. Growth of Human Stem Cell-Derived Neurons on Solid Three-Dimensional Polymers. J. Biochem. Biophys. Methods 2005, 62, 231–240. [Google Scholar] [CrossRef] [PubMed]
  269. Hayman, M.W.; Smith, K.H.; Cameron, N.R.; Przyborski, S.A. Enhanced Neurite Outgrowth by Human Neurons Grown on Solid Three-Dimensional Scaffolds. Biochem. Biophys. Res. Commun. 2004, 314, 483–488. [Google Scholar] [CrossRef] [PubMed]
  270. Alayoubi, A.; Alqahtani, S.; Kaddoumi, A.; Nazzal, S. Effect of PEG Surface Conformation on Anticancer Activity and Blood Circulation of Nanoemulsions Loaded with Tocotrienol-Rich Fraction of Palm Oil. AAPS J. 2013, 15, 1168–1179. [Google Scholar] [CrossRef] [Green Version]
  271. Han, S.; Nie, K.; Li, J.; Sun, Q.; Wang, X.; Li, X.; Li, Q. 3D Electrospun Nanofiber-Based Scaffolds: From Preparations and Properties to Tissue Regeneration Applications. Stem Cells Int. 2021, 2021, 8790143. [Google Scholar] [CrossRef]
  272. Sancho, A.; Vazquez, L.; De Juan-Pardo, E. Effect of Cold Storage on Collagen-Based Hydrogels for the Three-Dimensional Culture of Adipose-Derived Stem Cells. Biofabrication 2014, 6, 35017. [Google Scholar] [CrossRef]
  273. Ghaedi, M.; Soleimani, M.; Shabani, I.; Duan, Y.; Lotfi, A. Hepatic Differentiation from Human Mesenchymal Stem Cells on a Novel Nanofiber Scaffold. Cell. Mol. Biol. Lett. 2012, 17, 89–106. [Google Scholar] [CrossRef]
  274. Puppi, D.; Zhang, X.; Yang, L.; Chiellini, F.; Sun, X.; Chiellini, E. Nano/Microfibrous Polymeric Constructs Loaded with Bioactive Agents and Designed for Tissue Engineering Applications: A Review. J. Biomed. Mater. Res. Part B Appl. Biomater. 2014, 102, 1562–1579. [Google Scholar] [CrossRef]
  275. Kundu, B.; Rajkhowa, R.; Kundu, S.C.; Wang, X. Silk Fibroin Biomaterials for Tissue Regenerations. Adv. Drug Deliv. Rev. 2013, 65, 457–470. [Google Scholar] [CrossRef]
  276. O’Brien, F.J.; Harley, B.A.; Yannas, I.V.; Gibson, L.J. The Effect of Pore Size on Cell Adhesion in Collagen-GAG Scaffolds. Biomaterials 2005, 26, 433–441. [Google Scholar] [CrossRef]
  277. Zhu, W.; Holmes, B.; Glazer, R.I.; Zhang, L.G. 3D Printed Nanocomposite Matrix for the Study of Breast Cancer Bone Metastasis. Nanomed. Nanotechnol. Biol. Med. 2016, 12, 69–79. [Google Scholar] [CrossRef]
  278. Yang, Y.; Yang, X.; Zou, J.; Jia, C.; Hu, Y.; Du, H.; Wang, H. Evaluation of Photodynamic Therapy Efficiency Using an In Vitro Three-Dimensional Microfluidic Breast Cancer Tissue Model. Lab Chip 2014, 15, 735–744. [Google Scholar] [CrossRef] [PubMed]
  279. Rao, S.S.; DeJesus, J.; Short, A.R.; Otero, J.J.; Sarkar, A.; Winter, J.O. Glioblastoma Behaviors in Three-Dimensional Collagen-Hyaluronan Composite Hydrogels. ACS Appl. Mater. Interfaces 2013, 5, 9276–9284. [Google Scholar] [CrossRef] [PubMed]
  280. Friedl, P.; Maaser, K.; Klein, C.E.; Niggemann, B.; Krohne, G.; Zã, K.S. Migration of Highly Aggressive MV3 Melanoma Cells in 3-Dimensional Collagen Lattices Results in Local Matrix Reorganization and Shedding of A2 and 131Integrins and CD441. Cancer Res. 1997, 57, 2061–2070. [Google Scholar] [PubMed]
  281. Nyga, A.; Loizidou, M.; Emberton, M.; Cheema, U. A Novel Tissue Engineered Three-Dimensional in Vitro Colorectal Cancer Model. Acta Biomater. 2013, 9, 7917–7926. [Google Scholar] [CrossRef] [Green Version]
  282. Heywood, H.K.; Sembi, P.K.; Lee, D.A.; Bader, D.L. Cellular Utilization Determines Viability and Matrix Distribution Profiles in Chondrocyte-Seeded Alginate Constructs. Tissue Eng. 2004, 10, 1467–1479. [Google Scholar] [CrossRef]
  283. Jongpaiboonkit, L.; King, W.J.; Lyons, G.E.; Paguirigan, A.L.; Warrick, J.W.; Beebe, D.J.; Murphy, W.L. An Adaptable Hydrogel Array Format for 3-Dimensional Cell Culture and Analysis. Biomaterials 2008, 29, 3346–3356. [Google Scholar] [CrossRef] [Green Version]
  284. Gutowska, A.; Bae, Y.H.; Feijen, J.; Kim, S.W. Heparin Release from Thermosensitive Hydrogels. J. Control. Release 1992, 22, 95–104. [Google Scholar] [CrossRef] [Green Version]
  285. Ferreira, L.; Vidal, M.M.; Gil, M.H. Evaluation of Poly(2-Hydroxyethyl Methacrylate) Gels as Drug Delivery Systems at Different PH Values. Int. J. Pharm. 2000, 194, 169–180. [Google Scholar] [CrossRef] [Green Version]
  286. D’Emanuele, A.; Staniforth, J.N. An Electrically Modulated Drug Delivery Device: I. Pharm. Res. 1991, 8, 913–918. [Google Scholar] [CrossRef]
  287. Barbosa, M.A.; Chiesa, R.; Cigada, A.; Ginebra, M.P.; Leali, P.T.; Lewis, A.L.; Lloyd, A.W.; Merolli, A.; Montufar, E.B.; Pérez, R.A. Biomimetic, bioresponsive, and bioactive materials: An Introduction to Integrating Materials with Tissues; Santin, M., Phillips, G., Eds.; John Wiley & Sons: Hoboken, NJ, USA, 2012; ISBN 9780470056714. [Google Scholar]
  288. Léon, Y. León, C.A. New Perspectives in Mercury Porosimetry. Adv. Colloid Interface Sci. 1998, 76–77, 341–372. [Google Scholar] [CrossRef]
  289. Yang, S.; Leong, K.F.; Du, Z.; Chua, C.K. The Design of Scaffolds for Use in Tissue Engineering. Part I. Traditional Factors. Tissue Eng. 2001, 7, 679–689. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  290. Brauker, J.H.; Carr-Brendel, V.E.; Martinson, L.A.; Crudele, J.; Johnston, W.D.; Johnson, R.C. Neovascularization of Synthetic Membranes Directed by Membrane Microarchitecture. J. Biomed. Mater. Res. 1995, 29, 1517–1524. [Google Scholar] [CrossRef] [PubMed]
  291. Klawitter, J.J.; Hulbert, S.F. Application of Porous Ceramics for the Attachment of Load Bearing Internal Orthopedic Applications. J. Biomed. Mater. Res. 1971, 5, 161–229. [Google Scholar] [CrossRef]
  292. Levingstone, T.J.; Matsiko, A.; Dickson, G.R.; Brien, F.J.O.; Gleeson, J.P. Acta Biomaterialia A Biomimetic Multi-Layered Collagen-Based Scaffold for Osteochondral Repair. Acta Biomater. 2014, 10, 1996–2004. [Google Scholar] [CrossRef]
  293. Haugh, M.G.; Murphy, C.M.; Mckiernan, R.C.; Altenbuchner, C.; Brien, F.J.O. Cell Attachment, Proliferation, and Migration Within Collagen Glycosaminoglycan Scaffolds. Tissue Eng. Part A 2011, 17, 1201–1208. [Google Scholar] [CrossRef]
  294. Curtin, C.M.; Cunniffe, G.M.; Lyons, F.G.; Bessho, K.; Dickson, G.R.; Duffy, G.P.; Brien, F.J.O. Innovative Collagen Nano-Hydroxyapatite Scaffolds Offer a Highly Efficient Non-Viral Gene Delivery Platform for Stem Cell-Mediated Bone Formation. Adv. Mater. 2012, 24, 749–754. [Google Scholar] [CrossRef]
  295. Ueda, H.; Hong, L.; Yamamoto, M.; Shigeno, K.; Inoue, M.; Toba, T.; Yoshitani, M.; Nakamura, T.; Tabata, Y.; Shimizu, Y. Use of Collagen Sponge Incorporating Transforming Growth Factor-Β1 to Promote Bone Repair in Skull Defects in Rabbits. Biomaterials 2002, 23, 1003–1010. [Google Scholar] [CrossRef] [Green Version]
  296. Pawelec, K.M.; Husmann, A.; Best, S.M.; Cameron, R.E. Understanding Anisotropy and Architecture in Ice-Templated Biopolymer Scaffolds. Mater. Sci. Eng. C 2014, 37, 141–147. [Google Scholar] [CrossRef]
  297. Haugh, M.G.; Murphy, C.M.; O’Brien, F.J. Novel Freeze-Drying Methods to Produce a Range of Collagen–Glycosaminoglycan Scaffolds with Tailored Mean Pore Sizes. Tissue Eng. Part C Methods 2009, 16, 887–894. [Google Scholar] [CrossRef]
  298. O’Brien, F.J.; Harley, B.A.; Yannas, I.V.; Gibson, L. Influence of Freezing Rate on Pore Structure in Freeze-Dried Collagen-GAG Scaffolds. Biomaterials 2004, 25, 1077–1086. [Google Scholar] [CrossRef]
  299. Szot, C.S.; Buchanan, C.F.; Gatenholm, P.; Nichole, M.; Freeman, J.W. Investigation of Cancer Cell Behavior on Nano Fi Brous Scaffolds. Mater. Sci. Eng. C 2011, 31, 37–42. [Google Scholar] [CrossRef]
  300. Tran, C.; Kalra, V. Fabrication of Porous Carbon Nanofibers with Adjustable Pore Sizes as Electrodes for Supercapacitors. J. Power Sources 2013, 235, 289–296. [Google Scholar] [CrossRef]
  301. Phipps, M.C.; Clem, W.C.; Grunda, J.M.; Clines, G.A.; Bellis, S.L. Increasing the Pore Sizes of Bone-Mimetic Electrospun Scaffolds Comprised of Polycaprolactone, Collagen I and Hydroxyapatite to Enhance Cell Infiltration. Biomaterials 2012, 33, 524–534. [Google Scholar] [CrossRef] [Green Version]
  302. Blakeney, B.A.; Tambralli, A.; Anderson, J.M.; Andukuri, A.; Lim, D.; Dean, D.R.; Jun, H. Biomaterials Cell in Fi Ltration and Growth in a Low Density, Uncompressed Three-Dimensional Electrospun Nano Fi Brous Scaffold. Biomaterials 2011, 32, 1583–1590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Zhang, Y.Z.; Feng, Y.; Huang, Z.-M.; Ramakrishna, S.; Lim, C.T. Fabrication of Porous Electrospun Nanofibres. Nanotechnology 2006, 17, 901–908. [Google Scholar] [CrossRef]
  304. Kim, T.G.; Chung, H.J.; Park, T.G. Macroporous and Nanofibrous Hyaluronic Acid/Collagen Hybrid Scaffold Fabricated by Concurrent Electrospinning and Deposition/Leaching of Salt Particles. Acta Biomater. 2008, 4, 1611–1619. [Google Scholar] [CrossRef] [PubMed]
  305. Li, W.-J.; Laurencin, C.T.; Caterson, E.J.; Tuan, R.S.; Ko, F.K. Electrospun Nanofibrous Structure: A Novel Scaffold for Tissue Engineering. J. Biomed. Mater. Res. 2002, 60, 613–621. [Google Scholar] [CrossRef]
  306. Ma, P.X. Biomimetic Materials for Tissue Engineering. Adv. Drug Deliv. Rev. 2008, 60, 184–198. [Google Scholar] [CrossRef] [Green Version]
  307. Chen, Z.G.; Wang, P.W.; Wei, B.; Mo, X.M.; Cui, F.Z. Electrospun Collagen–Chitosan Nanofiber: A Biomimetic Extracellular Matrix for Endothelial Cell and Smooth Muscle Cell. Acta Biomater. 2010, 6, 372–382. [Google Scholar] [CrossRef]
  308. Aras, O.; Kazanci, M. Production of Collagen Micro- and Nanofibers for Potential Drug-Carrier Systems. J. Enzym. Inhib. Med. Chem. 2015, 30, 1013–1016. [Google Scholar] [CrossRef] [PubMed]
  309. Fiorani, A.; Gualandi, C.; Panseri, S.; Montesi, M.; Marcacci, M.; Focarete, M.L.; Bigi, A. Comparative Performance of Collagen Nanofibers Electrospun from Different Solvents and Stabilized by Different Crosslinkers. J. Mater. Sci. Mater. Med. 2014, 25, 2313–2321. [Google Scholar] [CrossRef]
  310. Dong, B.; Arnoult, O.; Smith, M.E.; Wnek, G.E. Electrospinning of Collagen Nanofiber Scaffolds from Benign Solvents. Macromol. Rapid Commun. 2009, 30, 539–542. [Google Scholar] [CrossRef] [PubMed]
  311. Zhong, S.; Teo, W.E.; Zhu, X.; Beuerman, R.W.; Ramakrishna, S.; Yung, L.Y.L. An Aligned Nanofibrous Collagen Scaffold by Electrospinning and Its Effects on in Vitro Fibroblast Culture. J. Biomed. Mater. Res. Part A 2006, 79, 456–463. [Google Scholar] [CrossRef] [PubMed]
  312. Hartman, O.; Zhang, C.; Adams, E.L.; Farach-Carson, M.C.; Petrelli, N.J.; Chase, B.D.; Rabolt, J.F. Microfabricated Electrospun Collagen Membranes for 3-D Cancer Models and Drug Screening Applications. Biomacromolecules 2009, 10, 2019–2032. [Google Scholar] [CrossRef]
  313. Eltom, A.; Zhong, G.; Muhammad, A. Scaffold Techniques and Designs in Tissue Engineering Functions and Purposes: A Review. Adv. Mater. Sci. Eng. 2019, 2019, 3429527. [Google Scholar] [CrossRef] [Green Version]
  314. Creff, J.; Courson, R.; Mangeat, T.; Foncy, J.; Souleille, S.; Thibault, C.; Besson, A.; Malaquin, L. Fabrication of 3D Scaffolds Reproducing Intestinal Epithelium Topography by High-Resolution 3D Stereolithography. Biomaterials 2019, 221, 119404. [Google Scholar] [CrossRef]
  315. Lee, S.-J.; Nowicki, M.; Harris, B.; Zhang, L.G. Fabrication of a Highly Aligned Neural Scaffold via a Table Top Stereolithography 3D Printing and Electrospinning. Tissue Eng. Part A 2016, 23, 491–502. [Google Scholar] [CrossRef]
  316. Accardo, A.; Courson, R.; Riesco, R.; Raimbault, V.; Malaquin, L. Direct Laser Fabrication of Meso-Scale 2D and 3D Architectures with Micrometric Feature Resolution. Addit. Manuf. 2018, 22, 440–446. [Google Scholar] [CrossRef]
  317. Wang, Y.; Gunasekara, D.B.; Reed, M.I.; DiSalvo, M.; Bultman, S.J.; Sims, C.E.; Magness, S.T.; Allbritton, N.L. A Microengineered Collagen Scaffold for Generating a Polarized Crypt-Villus Architecture of Human Small Intestinal Epithelium. Biomaterials 2017, 128, 44–55. [Google Scholar] [CrossRef]
  318. Yu, J.; Peng, S.; Luo, D.; March, J.C. In Vitro 3D Human Small Intestinal Villous Model for Drug Permeability Determination. Biotechnol. Bioeng. 2012, 109, 2173–2178. [Google Scholar] [CrossRef]
  319. Guvendiren, M.; Molde, J.; Soares, R.M.D.; Kohn, J. Designing Biomaterials for 3D Printing. ACS Biomater. Sci. Eng. 2016, 2, 1679–1693. [Google Scholar] [CrossRef] [PubMed]
  320. Zhang, Z.; Nagrath, S. Microfluidics and Cancer: Are We There Yet? Biomed. Microdevices 2013, 15, 595–609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  321. Shimizu, A.; Goh, W.H.; Hashimoto, M.; Miura, S.; Onoe, H. ECM-Based Stretchable Microfluidic System for in Vitro 3D Tissue Culture. In Proceedings of the 2019 20th International Conference on Solid-State Sensors, Actuators and Microsystems & Eurosensors XXXIII (TRANSDUCERS & EUROSENSORS XXXIII), Berlin, Germany, 23–27 June 2019; pp. 752–755. [Google Scholar]
  322. Hutmacher, D.W.; Loessner, D.; Rizzi, S.; Kaplan, D.L.; Mooney, D.J.; Clements, J.A. Can Tissue Engineering Concepts Advance Tumor Biology Research? Trends Biotechnol. 2010, 28, 125–133. [Google Scholar] [CrossRef] [PubMed]
  323. Ghajar, C.M.; Bissell, M.J. Tumor Engineering: The Other Face of Tissue Engineering. Tissue Eng. Part A 2010, 16, 2153–2156. [Google Scholar] [CrossRef] [Green Version]
  324. Whitesides, G.M. The Origins and the Future of Microfluidics. Nature 2006, 442, 368–373. [Google Scholar] [CrossRef]
  325. El-Ali, J.; Sorger, P.K.; Jensen, K.F. Cells on Chips. Nature 2006, 442, 403–411. [Google Scholar] [CrossRef]
  326. Pisano, M.; Triacca, V.; Barbee, K.A.; Swartz, M.A. An in Vitro Model of the Tumor-Lymphatic Microenvironment with Simultaneous Transendothelial and Luminal Flows Reveals Mechanisms of Flow Enhanced Invasion. Integr. Biol. 2015, 7, 525–533. [Google Scholar] [CrossRef]
  327. Jeon, J.S.; Bersini, S.; Gilardi, M.; Dubini, G.; Charest, J.L.; Moretti, M.; Kamm, R.D. Human 3D Vascularized Organotypic Microfluidic Assays to Study Breast Cancer Cell Extravasation. Proc. Natl. Acad. Sci. USA 2015, 112, 214–219. [Google Scholar] [CrossRef] [Green Version]
  328. Chen, Y.; Gao, D.; Liu, H.; Lin, S.; Jiang, Y. Drug Cytotoxicity and Signaling Pathway Analysis with Three-Dimensional Tumor Spheroids in a Microwell-Based Microfluidic Chip for Drug Screening. Anal. Chim. Acta 2015, 898, 85–92. [Google Scholar] [CrossRef]
  329. Kim, J.-Y.; Fluri, D.A.; Marchan, R.; Boonen, K.; Mohanty, S.; Singh, P.; Hammad, S.; Landuyt, B.; Hengstler, J.G.; Kelm, J.M.; et al. 3D Spherical Microtissues and Microfluidic Technology for Multi-Tissue Experiments and Analysis. J. Biotechnol. 2015, 205, 24–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  330. Ruppen, J.; Wildhaber, F.D.; Strub, C.; Hall, S.R.R.; Schmid, R.A.; Geiser, T.; Guenat, O.T. Towards Personalized Medicine: Chemosensitivity Assays of Patient Lung Cancer Cell Spheroids in a Perfused Microfluidic Platform. Lab Chip 2015, 15, 3076–3085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  331. McMillan, K.S.; Boyd, M.; Zagnoni, M. Transitioning from Multi-Phase to Single-Phase Microfluidics for Long-Term Culture and Treatment of Multicellular Spheroids. Lab Chip 2016, 16, 3548–3557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  332. Fu, C.-Y.; Tseng, S.-Y.; Yang, S.-M.; Hsu, L.; Liu, C.-H.; Chang, H.-Y. A Microfluidic Chip with a U-Shaped Microstructure Array for Multicellular Spheroid Formation, Culturing and Analysis. Biofabrication 2014, 6, 15009. [Google Scholar] [CrossRef] [Green Version]
  333. Patra, B.; Peng, C.-C.; Liao, W.-H.; Lee, C.-H.; Tung, Y.-C. Drug Testing and Flow Cytometry Analysis on a Large Number of Uniform Sized Tumor Spheroids Using a Microfluidic Device. Sci. Rep. 2016, 6, 21061. [Google Scholar] [CrossRef] [Green Version]
  334. Ramanujan, S.; Pluen, A.; McKee, T.D.; Brown, E.B.; Boucher, Y.; Jain, R.K. Diffusion and Convection in Collagen Gels: Implications for Transport in the Tumor Interstitium. Biophys. J. 2002, 83, 1650–1660. [Google Scholar] [CrossRef] [Green Version]
  335. Lintz, M.; Muñoz, A.; Reinhart-King, C.A. The Mechanics of Single Cell and Collective Migration of Tumor Cells. J. Biomech. Eng. 2017, 139, 210051–210059. [Google Scholar] [CrossRef] [Green Version]
  336. Hollister, S.J.; Maddox, R.D.; Taboas, J.M. Optimal Design and Fabrication of Scaffolds to Mimic Tissue Properties and Satisfy Biological Constraints. Biomaterials 2002, 23, 4095–4103. [Google Scholar] [CrossRef]
  337. Conklin, M.W.; Eickhoff, J.C.; Riching, K.M.; Pehlke, C.A.; Eliceiri, K.W.; Provenzano, P.P.; Friedl, A.; Keely, P.J. Aligned Collagen Is a Prognostic Signature for Survival in Human Breast Carcinoma. Am. J. Pathol. 2011, 178, 1221–1232. [Google Scholar] [CrossRef]
  338. Provenzano, P.P.; Eliceiri, K.W.; Campbell, J.M.; Inman, D.R.; White, J.G.; Keely, P.J. Collagen Reorganization at the Tumor-Stromal Interface Facilitates Local Invasion. BMC Med. 2006, 4, 38. [Google Scholar] [CrossRef] [Green Version]
  339. Brabrand, A.; Kariuki, I.I.; Engstrøm, M.J.; Haugen, O.A.; Dyrnes, L.A.; Åsvold, B.O.; Lilledahl, M.B.; Bofin, A.M. Alterations in Collagen Fibre Patterns in Breast Cancer. A Premise for Tumour Invasiveness? APMIS 2015, 123, 1–8. [Google Scholar] [CrossRef] [PubMed]
  340. Lee, C.H.; Singla, A.; Lee, Y. Biomedical Applications of Collagen. Int. J. Pharm. 2001, 221, 1–22. [Google Scholar] [CrossRef]
  341. Schindelin, J.; Arganda-carreras, I.; Frise, E.; Kaynig, V.; Pietzsch, T.; Preibisch, S.; Rueden, C.; Saalfeld, S.; Schmid, B.; Tinevez, J.; et al. Fiji—An Open Source Platform for Biological Image Analysis. Nat. Methods 2019, 9, 10–38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  342. Hilliou, F.; Pairault, J.; Dominice, J.; Redziniak, G. Growth and Differentiation of 3T3-F442A Preadipocytes in Three-Dimensional Gels of Native Collagen. Exp. Cell Res. 1988, 177, 372–381. [Google Scholar] [CrossRef]
  343. Huss, F.R.M.; Kratz, G. Mammary Epithelial Cell and Adipocyte Co-Culture in a 3-D Matrix: The First Step towards Tissue-Engineered Human Breast Tissue. Cells Tissues Organs 2001, 169, 361–367. [Google Scholar] [CrossRef] [PubMed]
  344. Jha, B.S.; Ayres, C.E.; Bowman, J.R.; Telemeco, T.A.; Sell, S.A.; Bowlin, G.L.; Simpson, D.G. Electrospun Collagen: A Tissue Engineering Scaffold with Unique Functional Properties in a Wide Variety of Applications. J. Nanomater. 2011, 2011, 348268. [Google Scholar] [CrossRef] [Green Version]
  345. Zeng, J.; Xu, X.; Chen, X.; Liang, Q.; Bian, X.; Yang, L.; Jing, X. Biodegradable Electrospun Fibers for Drug Delivery. J. Control. Release 2003, 92, 227–231. [Google Scholar] [CrossRef]
  346. Chaurey, V.; Chiang, P.-C.; Polanco, C.; Su, Y.-H.; Chou, C.-F.; Swami, N.S. Interplay of Electrical Forces for Alignment of Sub-100 Nm Electrospun Nanofibers on Insulator Gap Collectors. Langmuir 2010, 26, 19022–19026. [Google Scholar] [CrossRef]
  347. Beachley, V.; Wen, X. Effect of Electrospinning Parameters on the Nanofiber Diameter and Length. Mater. Sci. Eng. C 2009, 29, 663–668. [Google Scholar] [CrossRef] [Green Version]
  348. Zong, X.; Kim, K.; Fang, D.; Ran, S.; Hsiao, B.S.; Chu, B. Structure and Process Relationship of Electrospun Bioabsorbable Nanofiber Membranes. Polymer 2002, 43, 4403–4412. [Google Scholar] [CrossRef]
  349. Deitzel, J.M.; Kleinmeyer, J.; Harris, D.; Beck Tan, N.C. The Effect of Processing Variables on the Morphology of Electrospun Nanofibers and Textiles. Polymer 2001, 42, 261–272. [Google Scholar] [CrossRef]
  350. Nezarati, R.M.; Eifert, M.B.; Cosgriff-Hernandez, E. Effects of Humidity and Solution Viscosity on Electrospun Fiber Morphology. Tissue Eng. Part C: Methods 2013, 19, 810–819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  351. Tarus, B.; Fadel, N.; Al-Oufy, A.; El-Messiry, M. Effect of Polymer Concentration on the Morphology and Mechanical Characteristics of Electrospun Cellulose Acetate and Poly (Vinyl Chloride) Nanofiber Mats. Alex. Eng. J. 2016, 55, 2975–2984. [Google Scholar] [CrossRef] [Green Version]
  352. Rnjak-Kovacina, J.; Weiss, A.S. Increasing the Pore Size of Electrospun Scaffolds. Tissue Eng. Part B Rev. 2011, 17, 365–372. [Google Scholar] [CrossRef]
  353. Matthews, J.A.; Wnek, G.E.; Simpson, D.G.; Bowlin, G.L. Electrospinning of Collagen Nanofibers. Biomacromolecules 2002, 3, 232. [Google Scholar] [CrossRef] [PubMed]
  354. Shields, K.J.; Beckman, M.J.; Bowlin, G.L.; Wayne, J.S. Mechanical Properties and Cellular Proliferation of Electrospun Collagen Type II. Tissue Eng. 2004, 10, 1510–1517. [Google Scholar] [CrossRef]
  355. Malakpour-Permlid, A.; Buzzi, I.; Hegardt, C.; Johansson, F.; Oredsson, S. Identification of Extracellular Matrix Proteins Secreted by Human Dermal Fibroblasts Cultured in 3D Electrospun Scaffolds. Sci. Rep. 2021, 11, 6655. [Google Scholar] [CrossRef]
  356. Yamada, D.; Kobayashi, S.; Wada, H.; Kawamoto, K.; Marubashi, S.; Eguchi, H.; Ishii, H.; Nagano, H.; Doki, Y.; Mori, M. Role of Crosstalk between Interleukin-6 and Transforming Growth Factor-Beta 1 in Epithelial–Mesenchymal Transition and Chemoresistance in Biliary Tract Cancer. Eur. J. Cancer 2013, 49, 1725–1740. [Google Scholar] [CrossRef]
  357. Pankov, R.; Yamada, K.M. Fibronectin at a Glance. J. Cell Sci. 2002, 115, 3861–3863. [Google Scholar] [CrossRef] [Green Version]
  358. Lee, M.P.; Cooper, G.J.T.; Hinkley, T.; Gibson, G.M.; Padgett, M.J.; Cronin, L. Development of a 3D Printer Using Scanning Projection Stereolithography. Sci. Rep. 2015, 5, 9875. [Google Scholar] [CrossRef] [Green Version]
  359. Achilli, T.-M.; Meyer, J.; Morgan, J.R. Advances in the Formation, Use and Understanding of Multi-Cellular Spheroids. Expert Opin. Biol. Ther. 2012, 12, 1347–1360. [Google Scholar] [CrossRef] [Green Version]
  360. Duguay, D.; Foty, R.A.; Steinberg, M.S. Cadherin-Mediated Cell Adhesion and Tissue Segregation: Qualitative and Quantitative Determinants. Dev. Biol. 2003, 253, 309–323. [Google Scholar] [CrossRef] [Green Version]
  361. Ryu, N.-E.; Lee, S.-H.; Park, H. Spheroid Culture System Methods and Applications for Mesenchymal Stem Cells. Cells 2019, 8, 1620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  362. Moscona, A.; Moscona, H. The Dissociation and Aggregation of Cells from Organ Rudiments of the Early Chick Embryo. J. Anat. 1952, 86, 287–301. [Google Scholar] [PubMed]
  363. Lovitt, C.J.; Shelper, T.B.; Avery, V.M. Advanced Cell Culture Techniques for Cancer Drug Discovery. Biology 2014, 3, 345–367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  364. Kang, A.; Seo, H.I.; Chung, B.G.; Lee, S.-H. Concave Microwell Array-Mediated Three-Dimensional Tumor Model for Screening Anticancer Drug-Loaded Nanoparticles. Nanomed. Nanotechnol. Biol. Med. 2015, 11, 1153–1161. [Google Scholar] [CrossRef] [PubMed]
  365. Amann, A.; Zwierzina, M.; Gamerith, G.; Bitsche, M.; Huber, J.M.; Vogel, G.F.; Blumer, M.; Koeck, S.; Pechriggl, E.J.; Kelm, J.M.; et al. Development of an Innovative 3D Cell Culture System to Study Tumour-Stroma Interactions in Non-Small Cell Lung Cancer Cells. PLoS ONE 2014, 9, 1–13. [Google Scholar] [CrossRef] [Green Version]
  366. Mueller-Klieser, W. Three-Dimensional Cell Cultures: From Molecular Mechanisms to Clinical Applications. Am. J. Physiol. Physiol. 1997, 273, C1109–C1123. [Google Scholar] [CrossRef]
  367. Dhiman, H.K.; Ray, A.R.; Panda, A.K. Three-Dimensional Chitosan Scaffold-Based MCF-7 Cell Culture for the Determination of the Cytotoxicity of Tamoxifen. Biomaterials 2005, 26, 979–986. [Google Scholar] [CrossRef]
  368. Ong, S.-M.; Zhao, Z.; Arooz, T.; Zhao, D.; Zhang, S.; Du, T.; Wasser, M.; van Noort, D.; Yu, H. Engineering a Scaffold-Free 3D Tumor Model for in Vitro Drug Penetration Studies. Biomaterials 2010, 31, 1180–1190. [Google Scholar] [CrossRef]
  369. Godugu, C.; Patel, A.R.; Desai, U.; Andey, T.; Sams, A.; Singh, M. AlgiMatrixTM Based 3D Cell Culture System as an In-Vitro Tumor Model for Anticancer Studies. PLoS One 2013, 8, e53708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  370. Holtfreter, J. A Study of the Mechanics of Gastrulation. J. Exp. Zool. 1944, 95, 171–212. [Google Scholar] [CrossRef]
  371. Laschke, M.W.; Menger, M.D. Life Is 3D: Boosting Spheroid Function for Tissue Engineering. Trends Biotechnol. 2017, 35, 133–144. [Google Scholar] [CrossRef] [PubMed]
  372. Li, L.; Neaves, W.B. Normal Stem Cells and Cancer Stem Cells: The Niche Matters. Cancer Res. 2006, 66, 4553–4557. [Google Scholar] [CrossRef] [Green Version]
  373. Murphy, K.C.; Hoch, A.I.; Harvestine, J.N.; Zhou, D.; Leach, J.K. Mesenchymal Stem Cell Spheroids Retain Osteogenic Phenotype Through A2β1 Signaling. Stem Cells Transl. Med. 2016, 5, 1229–1237. [Google Scholar] [CrossRef]
  374. Fennema, E.; Rivron, N.; Rouwkema, J.; van Blitterswijk, C.; de Boer, J. Spheroid Culture as a Tool for Creating 3D Complex Tissues. Trends Biotechnol. 2013, 31, 108–115. [Google Scholar] [CrossRef]
  375. Costa, E.C.; Gaspar, V.M.; Coutinho, P.; Correia, I.J. Optimization of Liquid Overlay Technique to Formulate Heterogenic 3D Co-Cultures Models. Biotechnol. Bioeng. 2014, 111, 1672–1685. [Google Scholar] [CrossRef]
  376. Metzger, W.; Sossong, D.; Bächle, A.; Pütz, N.; Wennemuth, G.; Pohlemann, T.; Oberringer, M. The Liquid Overlay Technique Is the Key to Formation of Co-Culture Spheroids Consisting of Primary Osteoblasts, Fibroblasts and Endothelial Cells. Cytotherapy 2011, 13, 1000–1012. [Google Scholar] [CrossRef]
  377. Wang, X.; Zhen, X.; Wang, J.; Zhang, J.; Wu, W.; Jiang, X. Doxorubicin Delivery to 3D Multicellular Spheroids and Tumors Based on Boronic Acid-Rich Chitosan Nanoparticles. Biomaterials 2013, 34, 4667–4679. [Google Scholar] [CrossRef]
  378. Gunay, G.; Kirit, H.A.; Kamatar, A.; Baghdasaryan, O.; Hamsici, S.; Acar, H. The Effects of Size and Shape of the Ovarian Cancer Spheroids on the Drug Resistance and Migration. Gynecol. Oncol. 2020, 159, 563–572. [Google Scholar] [CrossRef]
  379. Torisawa, Y.; Takagi, A.; Nashimoto, Y.; Yasukawa, T.; Shiku, H.; Matsue, T. A Multicellular Spheroid Array to Realize Spheroid Formation, Culture, and Viability Assay on a Chip. Biomaterials 2007, 28, 559–566. [Google Scholar] [CrossRef]
  380. Sarisozen, C.; Dhokai, S.; Tsikudo, E.G.; Luther, E.; Rachman, I.M.; Torchilin, V.P. Nanomedicine Based Curcumin and Doxorubicin Combination Treatment of Glioblastoma with ScFv-Targeted Micelles: In Vitro Evaluation on 2D and 3D Tumor Models. Eur. J. Pharm. Biopharm. 2016, 108, 54–67. [Google Scholar] [CrossRef] [PubMed]
  381. Perche, F.; Patel, N.R.; Torchilin, V.P. Accumulation and Toxicity of Antibody-Targeted Doxorubicin-Loaded PEG-PE Micelles in Ovarian Cancer Cell Spheroid Model. J. Control. Release 2012, 164, 95–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  382. Zhang, Y.S.; Duchamp, M.; Oklu, R.; Ellisen, L.W.; Langer, R.; Khademhosseini, A. Bioprinting the Cancer Microenvironment. ACS Biomater. Sci. Eng. 2016, 2, 1710–1721. [Google Scholar] [CrossRef] [Green Version]
  383. Foty, R. A Simple Hanging Drop Cell Culture Protocol for Generation of 3D Spheroids. JoVE 2011, 51, e2720. [Google Scholar] [CrossRef] [PubMed]
  384. Hongisto, V.; Jernström, S.; Fey, V.; Mpindi, J.-P.; Kleivi Sahlberg, K.; Kallioniemi, O.; Perälä, M. High-Throughput 3D Screening Reveals Differences in Drug Sensitivities between Culture Models of JIMT1 Breast Cancer Cells. PLoS One 2013, 8, e77232. [Google Scholar] [CrossRef] [Green Version]
  385. Fukuda, J.; Nakazawa, K. Orderly Arrangement of Hepatocyte Spheroids on a Microfabricated Chip. Tissue Eng. 2005, 11, 1254–1262. [Google Scholar] [CrossRef]
  386. Desroches, B.R.; Zhang, P.; Choi, B.-R.; King, M.E.; Maldonado, A.E.; Li, W.; Rago, A.; Liu, G.; Nath, N.; Hartmann, K.M.; et al. Functional Scaffold-Free 3-D Cardiac Microtissues: A Novel Model for the Investigation of Heart Cells. Am. J. Physiol. Circ. Physiol. 2012, 302, H2031–H2042. [Google Scholar] [CrossRef] [Green Version]
  387. Ekert, J.E.; Johnson, K.; Strake, B.; Pardinas, J.; Jarantow, S.; Perkinson, R.; Colter, D.C. Three-Dimensional Lung Tumor Microenvironment Modulates Therapeutic Compound Responsiveness In Vitro—Implication for Drug Development. PLoS One 2014, 9, e92248. [Google Scholar] [CrossRef]
  388. Perets, A.; Baruch, Y.; Weisbuch, F.; Shoshany, G.; Neufeld, G.; Cohen, S. Enhancing the Vascularization of Three-Dimensional Porous Alginate Scaffolds by Incorporating Controlled Release Basic Fibroblast Growth Factor Microspheres. J. Biomed. Mater. Res. 2003, 65, 489–497. [Google Scholar] [CrossRef]
  389. Huang, G.-S.; Dai, L.-G.; Yen, B.L.; Hsu, S. Spheroid Formation of Mesenchymal Stem Cells on Chitosan and Chitosan-Hyaluronan Membranes. Biomaterials 2011, 32, 6929–6945. [Google Scholar] [CrossRef] [PubMed]
  390. Cheng, N.-C.; Wang, S.; Young, T.-H. The Influence of Spheroid Formation of Human Adipose-Derived Stem Cells on Chitosan Films on Stemness and Differentiation Capabilities. Biomaterials 2012, 33, 1748–1758. [Google Scholar] [CrossRef]
  391. Maritan, S.M.; Lian, E.Y.; Mulligan, L.M. An Efficient and Flexible Cell Aggregation Method for 3D Spheroid Production. JoVE 2017, 121, 55544. [Google Scholar] [CrossRef] [PubMed]
  392. Li, J.; He, F.; Pei, M. Creation of an in Vitro Microenvironment to Enhance Human Fetal Synovium-Derived Stem Cell Chondrogenesis. Cell Tissue Res. 2011, 345, 357–365. [Google Scholar] [CrossRef]
  393. Carpenedo, R.L.; Sargent, C.Y.; McDevitt, T.C. Rotary Suspension Culture Enhances the Efficiency, Yield, and Homogeneity of Embryoid Body Differentiation. Stem Cells 2007, 25, 2224–2234. [Google Scholar] [CrossRef]
  394. Bosnakovski, D.; Mizuno, M.; Kim, G.; Ishiguro, T.; Okumura, M.; Iwanaga, T.; Kadosawa, T.; Fujinaga, T. Chondrogenic Differentiation of Bovine Bone Marrow Mesenchymal Stem Cells in Pellet Cultural System. Exp. Hematol. 2004, 32, 502–509. [Google Scholar] [CrossRef] [PubMed]
  395. Giovannini, S.; Diaz-Romero, J.; Aigner, T.; Heini, P.; Mainil-Varlet, P.; Nesic, D. Micromass Co-Culture of Human Articular Chondrocytes and Human Bone Marrow Mesenchymal Stem Cells to Investigate Stable Neocartilage Tissue Formation in Vitro. Eur. Cells Mater. 2010, 20, 245–259. [Google Scholar] [CrossRef] [PubMed]
  396. Qihao, Z.; Xigu, C.; Guanghui, C.; Weiwei, Z. Spheroid Formation and Differentiation into Hepatocyte-like Cells of Rat Mesenchymal Stem Cell Induced by Co-Culture with Liver Cells. DNA Cell Biol. 2007, 26, 497–503. [Google Scholar] [CrossRef] [PubMed]
  397. Bartosh, T.J.; Ylostalo, J.H. Preparation of Anti-Inflammatory Mesenchymal Stem/Precursor Cells (MSCs) through Sphere Formation Using Hanging-Drop Culture Technique. Curr. Protoc. Stem Cell Biol. 2014, 28, 2B.6.1–2B.6.23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  398. Tung, Y.-C.; Hsiao, A.Y.; Allen, S.G.; Torisawa, Y.; Ho, M.; Takayama, S. High-Throughput 3D Spheroid Culture and Drug Testing Using a 384 Hanging Drop Array. Analyst 2011, 136, 473–478. [Google Scholar] [CrossRef] [PubMed]
  399. Kelm, J.M.; Timmins, N.E.; Brown, C.J.; Fussenegger, M.; Nielsen, L.K. Method for Generation of Homogeneous Multicellular Tumor Spheroids Applicable to a Wide Variety of Cell Types. Biotechnol. Bioeng. 2003, 83, 173–180. [Google Scholar] [CrossRef]
  400. Timmins, N.E.; Nielsen, L.K. Generation of Multicellular Tumor Spheroids by the Hanging-Drop Method. Methods Mol. Med. 2007, 140, 141–151. [Google Scholar] [CrossRef] [PubMed]
  401. Gong, X.; Lin, C.; Cheng, J.; Su, J.; Zhao, H.; Liu, T.; Wen, X.; Zhao, P. Generation of Multicellular Tumor Spheroids with Microwell-Based Agarose Scaffolds for Drug Testing. PLoS ONE 2015, 10, e0130348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  402. Dahlmann, J.; Kensah, G.; Kempf, H.; Skvorc, D.; Gawol, A.; Elliott, D.A.; Dräger, G.; Zweigerdt, R.; Martin, U.; Gruh, I. The Use of Agarose Microwells for Scalable Embryoid Body Formation and Cardiac Differentiation of Human and Murine Pluripotent Stem Cells. Biomaterials 2013, 34, 2463–2471. [Google Scholar] [CrossRef] [PubMed]
  403. Carlsson, J.; Yuhas, J.M. Liquid-Overlay Culture of Cellular Spheroids. Recent Results Cancer Res. 1984, 95, 1–23. [Google Scholar] [CrossRef]
  404. Costa, E.C.; de Melo-Diogo, D.; Moreira, A.F.; Carvalho, M.P.; Correia, I.J. Spheroids Formation on Non-Adhesive Surfaces by Liquid Overlay Technique: Considerations and Practical Approaches. Biotechnol. J. 2018, 13, 1700417. [Google Scholar] [CrossRef]
  405. Enmon, R.M.J.; O’Connor, K.C.; Lacks, D.J.; Schwartz, D.K.; Dotson, R.S. Dynamics of Spheroid Self-Assembly in Liquid-Overlay Culture of DU 145 Human Prostate Cancer Cells. Biotechnol. Bioeng. 2001, 72, 579–591. [Google Scholar] [CrossRef]
  406. Landry, J.; Bernier, D.; Ouellet, C.; Goyette, R.; Marceau, N. Spheroidal Aggregate Culture of Rat Liver Cells: Histotypic Reorganization, Biomatrix Deposition, and Maintenance of Functional Activities. J. Cell Biol. 1985, 101, 914–923. [Google Scholar] [CrossRef] [Green Version]
  407. Castañeda, F.; Kinne, R.K. Short Exposure to Millimolar Concentrations of Ethanol Induces Apoptotic Cell Death in Multicellular HepG2 Spheroids. J. Cancer Res. Clin. Oncol. 2000, 126, 305–310. [Google Scholar] [CrossRef]
  408. Foty, R.A.; Steinberg, M.S. The Differential Adhesion Hypothesis: A Direct Evaluation. Dev. Biol. 2005, 278, 255–263. [Google Scholar] [CrossRef] [Green Version]
  409. Lin, R.-Z.; Chang, H.-Y. Recent Advances in Three-Dimensional Multicellular Spheroid Culture for Biomedical Research. Biotechnol. J. 2008, 3, 1172–1184. [Google Scholar] [CrossRef] [PubMed]
  410. Kelm, J.M.; Fussenegger, M. Microscale Tissue Engineering Using Gravity-Enforced Cell Assembly. Trends Biotechnol. 2004, 22, 195–202. [Google Scholar] [CrossRef]
  411. Nyberg, S.L.; Hardin, J.; Amiot, B.; Argikar, U.A.; Remmel, R.P.; Rinaldo, P. Rapid, Large-Scale Formation of Porcine Hepatocyte Spheroids in a Novel Spheroid Reservoir Bioartificial Liver. Liver Transplant. 2005, 11, 901–910. [Google Scholar] [CrossRef]
  412. Friedrich, J.; Seidel, C.; Ebner, R.; Kunz-Schughart, L.A. Spheroid-Based Drug Screen: Considerations and Practical Approach. Nat. Protoc. 2009, 4, 309–324. [Google Scholar] [CrossRef]
  413. Friedrich, J.; Ebner, R.; Kunz-Schughart, L.A. Experimental Anti-Tumor Therapy in 3-D: Spheroids—Old Hat or New Challenge? Int. J. Radiat. Biol. 2007, 83, 849–871. [Google Scholar] [CrossRef] [PubMed]
  414. Uchida, N.; Buck, D.W.; He, D.; Reitsma, M.J.; Masek, M.; Phan, T.V.; Tsukamoto, A.S.; Gage, F.H.; Weissman, I.L. Direct Isolation of Human Central Nervous System Stem Cells. Proc. Natl. Acad. Sci. USA 2000, 97, 14720–14725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  415. Ponti, D.; Costa, A.; Zaffaroni, N.; Pratesi, G.; Petrangolini, G.; Coradini, D.; Pilotti, S.; Pierotti, M.A.; Daidone, M.G. Isolation and in Vitro Propagation of Tumorigenic Breast Cancer Cells with Stem/Progenitor Cell Properties. Cancer Res. 2005, 65, 5506–5511. [Google Scholar] [CrossRef] [Green Version]
  416. Smart, C.E.; Morrison, B.J.; Saunus, J.M.; Vargas, A.C.; Keith, P.; Reid, L.; Wockner, L.; Askarian-Amiri, M.; Sarkar, D.; Simpson, P.T.; et al. In Vitro Analysis of Breast Cancer Cell Line Tumourspheres and Primary Human Breast Epithelia Mammospheres Demonstrates Inter- and Intrasphere Heterogeneity. PLoS ONE 2013, 8, e64388. [Google Scholar] [CrossRef]
  417. Yu, M.; Bardia, A.; Aceto, N.; Bersani, F.; Madden, M.W.; Donaldson, M.C.; Desai, R.; Zhu, H.; Comaills, V.; Zheng, Z.; et al. Cancer Therapy. Ex Vivo Culture of Circulating Breast Tumor Cells for Individualized Testing of Drug Susceptibility. Science 2014, 345, 216–220. [Google Scholar] [CrossRef] [Green Version]
  418. Sato, T.; Stange, D.E.; Ferrante, M.; Vries, R.G.J.; Van Es, J.H.; Van den Brink, S.; Van Houdt, W.J.; Pronk, A.; Van Gorp, J.; Siersema, P.D.; et al. Long-Term Expansion of Epithelial Organoids from Human Colon, Adenoma, Adenocarcinoma, and Barrett’s Epithelium. Gastroenterology 2011, 141, 1762–1772. [Google Scholar] [CrossRef]
  419. Kondo, J.; Endo, H.; Okuyama, H.; Ishikawa, O.; Iishi, H.; Tsujii, M.; Ohue, M.; Inoue, M. Retaining Cell-Cell Contact Enables Preparation and Culture of Spheroids Composed of Pure Primary Cancer Cells from Colorectal Cancer. Proc. Natl. Acad. Sci. USA 2011, 108, 6235–6240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  420. Rajcevic, U.; Knol, J.C.; Piersma, S.; Bougnaud, S.; Fack, F.; Sundlisaeter, E.; Søndenaa, K.; Myklebust, R.; Pham, T.V.; Niclou, S.P.; et al. Colorectal Cancer Derived Organotypic Spheroids Maintain Essential Tissue Characteristics but Adapt Their Metabolism in Culture. Proteome Sci. 2014, 12, 39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  421. Young, S.R.; Saar, M.; Santos, J.; Nguyen, H.M.; Vessella, R.L.; Peehl, D.M. Establishment and Serial Passage of Cell Cultures Derived from LuCaP Xenografts. Prostate 2013, 73, 1251–1262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  422. Weiswald, L.-B.; Richon, S.; Massonnet, G.; Guinebretière, J.-M.; Vacher, S.; Laurendeau, I.; Cottu, P.; Marangoni, E.; Nemati, F.; Validire, P.; et al. A Short-Term Colorectal Cancer Sphere Culture as a Relevant Tool for Human Cancer Biology Investigation. Br. J. Cancer 2013, 108, 1720–1731. [Google Scholar] [CrossRef] [Green Version]
  423. Sundlisæter, E.; Wang, J.; Sakariassen, P.Ø.; Marie, M.; Mathisen, J.R.; Karlsen, B.O.; Prestegarden, L.; Skaftnesmo, K.O.; Bjerkvig, R.; Enger, P.Ø. Primary Glioma Spheroids Maintain Tumourogenicity and Essential Phenotypic Traits after Cryopreservation. Neuropathol. Appl. Neurobiol. 2006, 32, 419–427. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Scanning Electron Microscopy micrographs of the longitudinal sections of freeze-dried scaffolds of (a) collagen-based and (b) collagen-HA based (adapted from [219]).
Figure 1. Scanning Electron Microscopy micrographs of the longitudinal sections of freeze-dried scaffolds of (a) collagen-based and (b) collagen-HA based (adapted from [219]).
Ijms 22 12200 g001
Figure 2. Polystyrene well insert holder for 3D culture Alvetex Scaffold (alvetex®/www.interchim.com).
Figure 2. Polystyrene well insert holder for 3D culture Alvetex Scaffold (alvetex®/www.interchim.com).
Ijms 22 12200 g002
Figure 3. Different synthetic strategies of 3D matrix-based: (a) collagen; (b) Lyophilization; (c) Electrospiding; (d) Stereolithography; (e) Micro fluid [271,272,273,277,278].
Figure 3. Different synthetic strategies of 3D matrix-based: (a) collagen; (b) Lyophilization; (c) Electrospiding; (d) Stereolithography; (e) Micro fluid [271,272,273,277,278].
Ijms 22 12200 g003
Figure 4. Technical methods of spheroid formation: (a) Pellet culture; (b) Hanging drop; (c) Liquid overlay; (d) Spinner culture.
Figure 4. Technical methods of spheroid formation: (a) Pellet culture; (b) Hanging drop; (c) Liquid overlay; (d) Spinner culture.
Ijms 22 12200 g004
Table 1. Comparison between 2D and 3D culture models.
Table 1. Comparison between 2D and 3D culture models.
Characteristic2D3DReferences
Support for cell fixationUtensils (plastic, polycarbonate)Extracellular matrix in vitro[2]
Instructions for useTraditional cultureImitating the natural microenvironment[11,12,13,25,26,52]
Interaction and communicationCell-cell (co-culture)Cell-cell and cell-matrix 3D interactions[53]
Cell formsFlat and extensibleNatural cellular structure preserved[23,24,43,45]
Media cell interfaceHomogeneous exposure of all cells to the mediaHeterogeneous exposure (the upper layer is more exposed than the lower layer)[26]
Cell junctionsLess commonMore common (cell-cell communication)[3,10,12,54]
Cell differentiationModerately and poorly differentiatedWell-differentiated[43,44]
Cell proliferationHigher proliferation rate than in the natural environmentMedium or high proliferation rate depending on cell type and 3D culture technique[45,50,51]
Treatment sensitivityCells more sensitive to treatmentCells less sensitive to treatment[46,47]
ViabilitySensitive to cytotoxinsHigh viability and less sensitivity to external factors[55]
CostCheapExpensive[7,8]
Table 2. Scaffold-based 3D techniques overview with attribute.
Table 2. Scaffold-based 3D techniques overview with attribute.
TechniqueProtein-Based EMC Natural Hydrogels Synthetic HydrogelsHard Polymer Scaffold
Product descriptionMatrigel®Collagen, hyaluronic acidTrueGel3D (polymers with crosslinkers)Polystyrene-polycaprolactone Alvetex
Biological relevanceEffective +++Effective ++++/−+/−
Consistency/reproducibilityLow −High ++Very high+++Very high +++
Risk of contaminationLow −High++Very high+++Very high +++
Modularity/customizationLow −Moderate +High ++low −
Cell recovery+/−++++++
Downstream analysis (imaging, molecular analysis)+++++++
References[107,108,109,110][111,112,113,114,115][116,117,118][119,120,121,122,123]
Table 3. Advantages and disadvantages of hydrogel-based.
Table 3. Advantages and disadvantages of hydrogel-based.
Hydrogel AdvantageDisadvantages
Matrigel®-Widely available
-Frequently used in cancer research [151,173,174,175,176]
-Unknown and uncontrollable amount growth factors [24]
-Lack of control over its exact composition
-Variable from batch to batch [178]
-Difficulties of handling when it is in a refrigerated liquid state [104]
Based on Collagen-Good adhesion and cell migration support [145,146,147,148,156]
-Biocompatibility, mechanical strength, degradability, and limited immunogenicity [157,158]
-The most widely used tissue engineering and in tumor culture [11,152,158,199,200,201,202]
-Cell signaling patterns [140,141,142,143,144,203]
-Animal origin can potentially transmit pathogens [204]
-Biodegradable [159]
Hyaluronic acid-Provide hydration and resistance for cellular mechanisms [33,35,39]
-Biodegradable, non-immunogenic, non-inflammatory [205]
-Hydrodynamic and swelling [33,35]
-Animal origin can potentially transmit pathogens[204].
-Mechanically poor [38]
-Biodegradable [159]
Synthetic (PEG), (PCL), (PLA) (PGA)-Most used in 3D neural culture, bones, cartilaginous, tissue, and kidney tissue [206,207,208,209,210,211,212]
-A defined chemical composition and adjustable mechanical properties for cultivation [213,214,215]
-Available [119]
-Easily modified and formulated with different rigidity only the type of fabric [119]

-Physiologically irrelevant and may release toxic degradation products to cells [216]
-Limited applications in in vitro tumor engineering [216]
-Contains active chemical groups sensitive to chemical reactions [217]
-Irrelevant and may release toxic degradation products to cells [216]
-Biophysical parameters (mechanical properties and permeability, stiffness) must be considered [119,206,212]
-Loss of cell signaling patterns [203]
-Sensitive to pH (PEG) [218]
Table 4. Comparison of technical characteristics between scaffolds made with pure collagen vs. collagen-HA-based ones [219,227,228,229,234,235].
Table 4. Comparison of technical characteristics between scaffolds made with pure collagen vs. collagen-HA-based ones [219,227,228,229,234,235].
Pure CollagenCollagen-HA
TechniqueBy lyophilization 1%By lyophilization 1%
Pore size100 et 220 μm100 et 220 μm
PorositySimilarSimilar
DenaturationAbsentAbsent
Efficacity+++++
Resistance of dissolution+++
Dissolution hydrolyte19.2% in 7 days11.4% to 13.3% in 7 days
Cellular proliferation+++++
Table 5. Comparison of different synthetic strategies of 3D matrix-based.
Table 5. Comparison of different synthetic strategies of 3D matrix-based.
Fabrication Method Method OverviewScaffolding MorphologyAdvantagesDisadvantages
Hydrogels [11,54,193,279,280,281]-Collagen gel solution (usually type 1 collagen and acetic acid) mixed on ice and usually neutralized (NaOH) and then gelled
-Physical parameters: collagen, pH, the temperature of desired gelling
-Dense gel network of string-like fibers. The thickness of the fiber depends on the manufacturing parameters-Easy to apply
-Matrigel is widely used in cancer research, so many user guides are available
-High level of cell viability
-The least porous
-Risk of poor distribution of cells and nutrients.
-An architecture is more difficult to control, therefore, has less reproducibility of the exact architectures desired
-Poor mechanical properties before cross-linking
Lyophilization
[153,276,292,293,294,295,296,297,298]
-Creation of a homogeneous suspension of collagen with acid (usually acetic acid) at high speed
-Heat treatment (controlled or quenched) for the sublimation of ice crystals under vacuum to the defined freezing point before returning to ~0 °C
The dried scaffolding must reach room temperature to complete the process
-Interconnected network
-Highly porous
-A well-defined pore shape and sizes
-Good control of scaffolding architecture
-A wide production range in terms of pore sizes and orientation
-High porosity levels.
-Inexpensive-High level of cell viability
-Problems in the freezing process affect the final scaffolding architecture from one batch to another
-Poor mechanical properties before cross-linking
Electrospiding [299,300,301,302,303,304,305,306,307,308,309,310,311,312]-Collagen solubilized (usually HFIP or TFE) and added to the syringe/injection system
-A high-voltage electric field is applied, causing the charge of the solution, the eruption of the polymer fiber of the tip of the needle, and the whip of the liquid jet
-The solvent evaporates during the process, leaving a network of dried fibers deposited on the collection plate (non-woven or aligned)
-Dense and tight fiber array (chain-shaped) of nanometric or micro size-Fibrous network that closely resembles native collagen fibers.
-Wide range of size/diameter/achievable fiber pattern
-High level of reported cell viability
-Use of harmful solvents (collagen scaffolding)
-Solvents are expensive
-Dense fiber networks can reduce the level of cellular infiltration.
Stereolithography [277,313,314,315,316,317,318,319]-prints layer by layer a UV-curable material in thin sheets
-Installation of a multiresolution 3D printer (Dilase 3D, Kloe France)
-Each layer is superimposed after drying the next layer
-Use of different light sources (visible, UV, IR) capable of polymerizing photosensitive materials.
-Hard layer set (UV)-Capable of producing scaffolding of size mm to cm
-Can be combined with different components to hydrogels or electro spinning (PCL fibers, PCL /gelatin)
-high differentiation rates and adhesion
-Imitates complex structures in vitro: as villi of the intestine
-Specific equipment
-Expensive
-Manufactured scaffolding is usually limited to a few tens of microns of resolution
Micro fluid [278,320,321,322,323,324,325,326,327,328,329,330,331,332,333]Support consisting of silicon/elastomer-based devices having microchannels with proportions from 1 to 1000 μm that exploit a small volume of fluids (10-9 to 10-18 L). These fluids are continuous flows of nutrients and therapeutic agents, establish a physiological profile such as that of blood circulation and intravenous injections-Matrix that has micro channels- which can be either strictly laminar (in parallel layers) or turbulent (parallel and strong numbers)-Labor-saving
-Microenvironment dynamics (fluid flow)
-Generate aggregates of different forms Co-culture of several cells
-Simulates cell-cell contacts and biological signals controlled by spatial and temporal gradients of soluble biological factors
-Study tumor progression, invasion, angiogenesis as well as treatment tests
-Low reagent consumption and low cell utilization
-Requiring professional equipment and special design
-Complexity.
-High cost
Table 6. Advantages and disadvantages of spheroid culture.
Table 6. Advantages and disadvantages of spheroid culture.
Advantage [371,372,373,376,377,378,379,380,381]Disadvantage [359,366,376,377,378,379,380,381]
  • Inexpensive
  • High efficiency
  • Improves cell viability and proliferation
  • Retains intrinsic phenotypic property
  • Keeps physical interactions that more closely reflect behavior in the three-dimensional native tissue (3D)
Variable diameter and size
Intense work
Diffusion gradient depends on the size (oxygen nutrient, paracrine factor) that decreases inwards
Self-disassembly is affected by the rate of production and consumption of factors
Table 7. Technical methods of spheroid formation.
Table 7. Technical methods of spheroid formation.
Technical Methods Means of ApplicationMode of OperationAdvantages and Disadvantages References
Pellet CultureConcentrate the cells at the conical bottom of a tube by centrifugal force (500 g/5 min)-Remove the supernatants to collect the cell cap
-Capus resuspended in a culture medium to form the spheroids
-To optimize: the suspension can be incubated on an agitator for one hour before centrifugation
-Maximized cell-to-cell adhesions
-Suitable for the differentiation of mesenchymal cells, chondrogenesis, and bone formation
-Disadvantage: Shear stress due to centrifugation can damage cells
[359,361,391,392,393,394,395,396]
Hanging dropUse of surface tension and gravitational force to form spheroids in the form of droplets that rely on gravity self-disassembly-Preparation of a cell suspension at desired density distribution in the wells of a mini-plateau
-Placed a lid on the mini-tray, and the entire mini-tray is overturned upside down
-The drop remains fixed on the mini-tray on the inverted surface (surface tension)
-Most commonly used
-Defined and controlled size of the spheroid (drop volume and suspension density)
-Coefficient of variation narrow size distribution from 10 to 15%
-Inexpensive equipment
-A large amount can be produced-Heterotypic spheroids (’to 384 spheroids in a single network)
[359,361,383,397,398,399,400]
The cultivation of molded lozengesNon-adhesive gel (agarose) usually prepared in molds-Cells are forced to aggregate by continuous agitation
-Can be accelerated by centrifugation
-Removes restrictions on spheroid size
-Increases production rate
-High centrifugation can disrupt spheroids (function)
[361,401,402]
Liquid overlay (static suspensionMaterials that do not adhere to cells that inhibit cell attachment, such as agarose (agar) gel or pHEMACell bindings to the support are inhibited; cells spontaneously form spheroids-Coefficient of variation narrow size distribution from 40% to 60%
-Easy to monitor the formation and growth of spheroids in a plate 96 wells
-Simple Method
-Heterogeneous spheroids in size and shape
[359,361,376,399,403,404,405,406]
Spinner CultureUse of convection force by stirring the bar in centrifugal flask bioreactor containers generated by a magnetic stirring wheel or barAdd the uniform and well-mixed single-celled suspension with constant continuous stirring-The spheroid depends on the size of the bioreactor container
-Speed must be constant
-A high stirring speed affects the spheroids and a slow speed makes the cells sink to the bottom of the container (blocks the spheroids)
-Forms heterotypic spheroids
-May not be useful for cells with low cohesion (risk of apoptosis)
-it is difficult to follow the spheroids during formation
[26,359,361,407,408,409,410,411]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Habanjar, O.; Diab-Assaf, M.; Caldefie-Chezet, F.; Delort, L. 3D Cell Culture Systems: Tumor Application, Advantages, and Disadvantages. Int. J. Mol. Sci. 2021, 22, 12200. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms222212200

AMA Style

Habanjar O, Diab-Assaf M, Caldefie-Chezet F, Delort L. 3D Cell Culture Systems: Tumor Application, Advantages, and Disadvantages. International Journal of Molecular Sciences. 2021; 22(22):12200. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms222212200

Chicago/Turabian Style

Habanjar, Ola, Mona Diab-Assaf, Florence Caldefie-Chezet, and Laetitia Delort. 2021. "3D Cell Culture Systems: Tumor Application, Advantages, and Disadvantages" International Journal of Molecular Sciences 22, no. 22: 12200. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms222212200

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop