Next Article in Journal
Spectral Algal Fingerprinting and Long Sequencing in Synthetic Algal–Microbial Communities
Previous Article in Journal
A Novel Non-Invasive Murine Model of Neonatal Hypoxic-Ischemic Encephalopathy Demonstrates Developmental Delay and Motor Deficits with Activation of Inflammatory Pathways in Monocytes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of Single Cell Type-Derived Spheroids Generated by Using a Hanging Drop Culture Technique in Various In Vitro Disease Models: A Narrow Review

1
Departments of Ophthalmology, Sapporo Medical University, S1W17, Chuo-ku, Sapporo 060-8556, Japan
2
Departments of Cardiovascular, Renal and Metabolic Medicine, Sapporo Medical University, S1W17, Chuo-ku, Sapporo 060-8556, Japan
3
Departments of Cellular Physiology and Signal Transduction, Sapporo Medical University, S1W17, Chuo-ku, Sapporo 060-8556, Japan
4
Departments of Molecular Biology, Sapporo Medical University, S1W17, Chuo-ku, Sapporo 060-8556, Japan
5
Departments of Oral Surgery, Sapporo Medical University, S1W17, Chuo-ku, Sapporo 060-8556, Japan
6
Departments of Medical Oncology, Sapporo Medical University, S1W17, Chuo-ku, Sapporo 060-8556, Japan
7
Departments of Dermatology, Sapporo Medical University, S1W17, Chuo-ku, Sapporo 060-8556, Japan
*
Authors to whom correspondence should be addressed.
Cells 2024, 13(18), 1549; https://doi.org/10.3390/cells13181549 (registering DOI)
Submission received: 28 June 2024 / Revised: 21 August 2024 / Accepted: 23 August 2024 / Published: 14 September 2024

Abstract

:
Cell culture methods are indispensable strategies for studies in biological sciences and for drug discovery and testing. Most cell cultures have been developed using two-dimensional (2D) culture methods, but three-dimensional (3D) culture techniques enable the establishment of in vitro models that replicate various pathogenic conditions and they provide valuable insights into the pathophysiology of various diseases as well as more precise results in tests for drug efficacy. However, one difficulty in the use of 3D cultures is selection of the appropriate 3D cell culture technique for the study purpose among the various techniques ranging from the simplest single cell type-derived spheroid culture to the more sophisticated organoid cultures. In the simplest single cell type-derived spheroid cultures, there are also various scaffold-assisted methods such as hydrogel-assisted cultures, biofilm-assisted cultures, particle-assisted cultures, and magnet particle-assisted cultures, as well as non-assisted methods, such as static suspension cultures, floating cultures, and hanging drop cultures. Since each method can be differently influenced by various factors such as gravity force, buoyant force, centrifugal force, and magnetic force, in addition to non-physiological scaffolds, each method has its own advantages and disadvantages, and the methods have different suitable applications. We have been focusing on the use of a hanging drop culture method for modeling various non-cancerous and cancerous diseases because this technique is affected only by gravity force and buoyant force and is thus the simplest method among the various single cell type-derived spheroid culture methods. We have found that the biological natures of spheroids generated even by the simplest method of hanging drop cultures are completely different from those of 2D cultured cells. In this review, we focus on the biological aspects of single cell type-derived spheroid culture and its applications in in vitro models for various diseases.

1. Introduction

In addition to conventional two-dimensional (2D) culture methods, much attention has recently been given to three-dimensional (3D) cell cultures that can replicate microenvironments of cells under physiological and pathological conditions, and applications of 3D culture methods in various biological research fields have been rapidly increasing [1]. There are various 3D cell culture methods ranging from the simplest single cell type-derived spheroid cultures to more sophisticated methods including spheroid co-cultures, spheroid on-chip cultures, 3D organoid cultures, tissue-like organoid cultures, and 3D organoid on-chip cultures [2,3,4]. Furthermore, for the simplest method, the single cell type-derived spheroid method, there are various methods using biomaterial scaffolds for assisting the formation of spheroids including hydrogels, biofilms, particles, and magnet particles and methods that do not require scaffolds including static suspension cultures, floating cultures, and hanging drop cultures [3]. Recently, we have been carrying out studies on the pathophysiology of various non-cancerous and cancerous tissues using single cell type-derived spheroids generated by using the hanging drop culture method [5,6,7,8,9]. Based on our experience, we mainly focus on single cell type-derived spheroid methods in this review and discuss the advantages and disadvantages as well as their contributions to elucidation of the pathophysiological aspects of various non-cancerous and cancerous diseases.

1.1. Various Methods for Single-Cell 3D Spheroid Generation

1.1.1. Scaffold-Assisted Methods

Hydrogel-Assisted Cultures

Hydrogels are crosslinked polymer chains with 3D spatial structures that can contain a large amount of water due to their hydrophilic groups such as -NH2, -COOH, -OH, -CONH2, -CONH, and -SO3H [10]. In practical use, these scaffolds are placed in a well [11], in a microfluidic channel [11], or on a micropillar [12]. An ad-vantage of this culture method is that the physical properties, including the size and stiffness, of spheroids can be optimized by selecting biomaterials such as alginate [13], fibrin [14], collagen [15], hyaluronic acid [16], and ECM-based hydrogels, such as Mat-rigel® [17,18,19] and Cultrex® [20], and their appropriate concentrations for fabrication of 3D spatial scaffolds [21].

Biofilm EPS-Assisted Cultures

Biofilm extracellular polymeric substances (Biofilm EPSs), which consist of a mix-ture of polysaccharides, extracellular DNA, and proteins [22], are used for spheroid generation [23]. It was shown that the size of spheroids can be controlled by the thick-ness of the film and composition of the EPS [24,25].

Gelatin Microparticle-Assisted Cultures

In this culture method, cells are cultured with gelatin microparticles containing biological factors on a cell repellent-coated culture plate [26,27]. By using desired bio-logical factors in the particles, the biological nature of the spheroids can be spatially controlled [28,29].

Magnetic Particle-Assisted Cultures

In this method, cells are mixed with magnetic particles and cultured on a mag-net-equipped cell culture plate. Because of the magnetic force, cells aggregate with floating magnetic particles in the culture medium [30]. However, it has been suggested that artificially manipulated gravity may induce changes in the cellular biological na-ture that result in apoptosis [31,32].

1.1.2. Biomaterial Non-Assisted Methods

Static Suspension Cultures

In this culture method, cells are concentrated to form a spheroid on culture plates with non-adhesive surfaces, such as ultra-low attachment (ULA) plates with U-shaped bottomed wells or surfaces coated with hydrophilic substances such as agar or poly Hema, to minimize cell–substrate adhesion so as to allow cells to remain suspended. Subsequent centrifugation is required to support the aggregation of cells and the formation of a single spheroid per well [33]. This culture method has been shown to generate reproducible spheroids with uniform sizes and shapes that are suitable for high-throughput drug screening [34]. To overcome the problem of limited generation of spheroids, a new 96-well plate microcavity array-based platform, which can generate spheroids in the microcavities, has recently been developed for drug testing with higher parallelization capacity than that of previous devices [35,36].

Floating Cultures

In this culture method, cells are cultured in a spinner or rotation chamber in which the cell suspension is continuously mixed by stirring to avoid cell adhesion to the bottom of the chamber [37]. According to the difference in intercellular affinity, the conditions of the stirring forces should be adjusted for proper spheroid formation [38].

Hanging Drop Cultures

This technique allows cells to spontaneously aggregate for the formation of a spheroid in a culture droplet. It is possible to control the spheroid size by adjusting the volume of the droplet or the concentration of cells in the suspension [39].

1.1.3. Comparison of the Various Influencing Factors among the 3D Spheroid Culture Methods

Table 1 shows a comparison of the (1) sites of spheroid generation, (2) effects of gravity force, (3) effects of buoyant force, and (4) contributions of additive factors among the above-described culture methods. In the magnetic particle-assisted cultures, floating cultures, and hanging drop cultures, spheroids are not generated on the ground and are thus more influenced by gravity force and buoyant force compared to the other culture methods. In addition, except for static suspension cultures, other methods are influenced by additive factors including some scaffolds, magnet force, centrifugal force, or methyl cellulose as a spheroid stabilizer. Therefore, collectively, hanging drop culture is the simplest method for generation of spheroids as it is only influenced by gravity force, buoyant force, and methyl cellulose. In fact, a previous study showed that the hanging drop culture method generated circular spheroids that had a narrow distribution of sizes with variation coefficients of 10% to 15%, which are smaller than the variation coefficients of 40% to 60% in spheroid preparation using non-adherent surface culture methods [40]. Furthermore, by using a 384-hanging drop array culture plate, very efficient and reproducible formation of spheroids with controlled sizes is possible [41].

2. Preparation and Characterization of 3D Spheroids Formed by Using a 384-Hanging Drop Array Culture Plate

2.1. Representative Protocol of Preparation (Figure 1)

  • Step 1. Cells to be used for spheroid generation are prepared by conventional 2D planar culture in 100 mm or 150 mm dishes until the cells reach approximately 90% confluence at 37 °C using a CO2 cell incubator, as described elsewhere.
  • Step 2. The cell pellet is collected from the 2D cell culture plate using 0.25% Tryp sin/EDTA and centrifugation after washing with phosphate-buffered saline (PBS) on a clean bench.
  • Step 3. The cells are resuspended in the 2D culture medium supplemented with methylcellulose (Methocel A4M, Sigma-Aldrich Co., St. Louis, MO, USA) to facilitate stable morphology at a concentration of 20,000 cells in 28 μL of the culture medium on a clean bench.
  • Step 4. A 28 μL aliquot of the cell suspension is placed in each well of a hanging drop culture plate (# HDP1385, Sigma-Aldrich Co., St. Louis, MO, USA) on a clean bench, and then the culture plate is placed in a CO2 cell incubator.
  • Step 5. Half of the medium (14 μL) in each well is replaced daily by 14 μL of a fresh culture medium on a clean bench, and then the culture plate is placed in a CO2 cell incubator.
An important point is that careful attention should be paid to performing the procedures during steps 4 and 5 very gently to prevent the culture medium containing spheroids from dropping off. Another important point is that spheroid culture should be carried out during a period of approximately 4 to 12 days to accomplish its maturation, which is checked by a phase contrast microscopic examination. A longer culture period may lead to apoptosis of cells located near the core of the spheroid.
Figure 1. Method for preparation of 3D spheroids by a hanging drop culture plate. Cells obtained by conventional 2D planar culture (A) formed spheroid cultures by using a 384-hanging drop array culture plate ((B): design drawing). As shown in the upward view (C), a drop of culture medium containing cells hangs down from each well. Representative images of a spheroid of mouse orbital fibroblasts with adipogenesis stained by BODIPY (red), phalloidin (green), and DAPI (blue) (D) and a spheroid of human orbital fibroblasts with adipogenesis stained by anti-hyaluronic acid (pink), anti-COL6 (green), and DAPI (blue) (E).
Figure 1. Method for preparation of 3D spheroids by a hanging drop culture plate. Cells obtained by conventional 2D planar culture (A) formed spheroid cultures by using a 384-hanging drop array culture plate ((B): design drawing). As shown in the upward view (C), a drop of culture medium containing cells hangs down from each well. Representative images of a spheroid of mouse orbital fibroblasts with adipogenesis stained by BODIPY (red), phalloidin (green), and DAPI (blue) (D) and a spheroid of human orbital fibroblasts with adipogenesis stained by anti-hyaluronic acid (pink), anti-COL6 (green), and DAPI (blue) (E).
Cells 13 01549 g001

2.2. Analyses of the Physical Properties, Including Size and Stiffness, and the Morphology of 3D Spheroids

2.2.1. Measurement of the Sizes of 3D Spheroids

For measurement of the sizes of 3D spheroids, a phase contrast (PC) microscopy image is taken and then the cross-sectional area of the PC image is drawn and the area is calculated by using Image-J software version 1.51n (National Institutes of Health, Bethesda, MD, USA).

2.2.2. Measurement of the Stiffness of 3D Spheroids

Measurement of the stiffness of a single living spheroid is carried out by using a MicroSquisher (CellScale, Waterloo, ON, Canada) consisting of a microscale cantilever-type pressure sensor with 3 mm × 3 mm basal and compression plates in a measurement chamber and a microscope with a camera outside the chamber. After the measurement chamber is filled with fresh PBS at 37 °C, a living single spheroid is transferred onto the basal plate and then gradually compressed by the compression plate with a compression bar until the vertical length of the spheroid becomes half of the length during a period of 20 s (Figure 2). The maximum force required (μN) is measured. The index for the stiffness of a spheroid is determined by maximum force/half of the length (μN/μm).

2.2.3. Analysis of Morphology by Scanning Electron Microscopy and Immunocytochemistry

To study the morphology of a 3D spheroid in detail, the 3D spheroid is fixed with 2.5% glutaraldehyde for evaluation by scanning electron microscopy and is fixed with 4% paraformaldehyde for evaluation by immunocytochemistry and is processed as described elsewhere.

2.3. Representative Process of 3D Spheroid Maturation

As for the maturation process of a spheroid using a hanging drop culture method, our group showed that 3T3-L1 preadipocytes started to coalesce at 1 h after cells were added into wells of the culture plate and had formed premature spheroids by 6 h. The formation of spheroids became more apparent with the formation of smaller, round-shaped, and stiffened mature spheroids from day 1 through day 7 of culture [42]. Similar to 3T3-L1 spheroids, as shown in PC images (Figure 3), after placing 20,000 H9c2 cardiomyoblasts in 28 μL of culture medium into each well of a spheroid drop culture plate, the cells start to aggregate and then gradually form an immature spheroid within 7.5 h. On day 1, a matured spheroid formed.

2.4. Estimation of the Total Number of Cells in a Single Spheroid

As shown in trypsin digestion of 2D H9c2 cells and a H9c2 spheroid (Figure 4), once a spheroid has formed, the time needed to break the spheroid into single cells is much longer than that in the case of 2D cultured cells. Most of the cells in a spheroid are arranged concentrically from the center to the surface, as shown in the DAPI nuclei staining image (Figure 5). Therefore, to estimate the total number of cells in a spheroid, the tentative diameter of a spheroid is measured by its cross-section in a Phalloidin image and the diameter of a single cell in the spheroid is estimated by the distance between two adjacent nuclei stained by DAPI. Using these diameters, the number of cells in a spheroid is calculated by dividing the volume of the overall spheroid by the volume of a cell.

2.5. Cellular Metabolic Analysis of 3D Spheroid

The use of Seahorse extracellular flux bioanalyzers enables real-time measurements of oxygen consumption rate (OCR) as a function of mitochondrial respiration and extracellular acidification rate (ECAR) as a glycolysis function occurring mainly in 2D cultured cells [43]. In addition, several past studies have applied OCR and ECAR measurements in a 3D culture system using single-well assays [44,45,46,47]. This system was later optimized to measure OCR and ECAR in specimens of multiple spheroids using 96-multi-well plates [48]. This system has been used mainly in cancer research [48,49,50,51]. Our group has also performed measurements of OCR and ECAR of single cell type-derived spheroids using cancerous cells, including A549 cells [52], Mia PaCa2 cells [53], malignant melanoma (MM) cells [7,54], oral squamous cell carcinoma (OSCC) cells [55], and cancer-associated fibroblasts (CAFs) [9], and non-cancerous cells, including 3T3-L1 cells [56], human trabecular meshwork (HTM) cells [57,58], and human conjunctival fibroblasts (HconFs) [59]. Based on these experiences, we confirmed that the essential biological functions of mitochondrial respiration and glycolysis of spheroids are indeed substantially different from those of 2D cultured cells even though the same cells were used with only a difference in their culture system: 2D vs. 3D.

3. Application of 3D Spheroids for Various Fields in Biological Science

3.1. Study of Adipogenesis Using 3T3-L1-1 Cells (Figure 6)

In the field of adipogenesis-related research, mouse preadipocyte 3T3-L1 cells are most frequently used [60,61]. Although conventional 2D planar cell culture methods have been used in most of the studies in this field, for a better understanding of the in vivo nature of adipogenesis of adipose tissues in spatial environments in vivo, various 3D cell culture methods including polymeric scaffold-assisted cultures [62,63,64,65,66,67], magnetic particle-assisted cultures [68,69], and static suspension cultures using an elastin-like polypeptide-polyethyleneimine (ELP-PEI)-coated surface [70,71], in addition to hanging drop cultures using a 384-hanging drop array culture plate, have been used [42,72,73,74,75,76]. Among these different 3T3-L1 spheroid culture methods, our hanging drop culture is much faster and reproducibly produces round-shaped 3T3-L1 spheroids, although most of the studies in which spheroid culture methods were used have shown that biological aspects, such as the adipogenesis efficacy of 3T3-L1 spheroids, are significantly different from those of 2D 3T3-L1 cells. Recently, our group [6] and another study group [62] have shown spontaneous adipogenesis before adipogenic differentiation in 3T3-L1 spheroids but not in 2D cultured 3T3-L1 cells. In fact, proteomics [51] and RNA sequencing analysis [6,56] have revealed significant changes in various adipogenesis-related factors, inflammatory cytokines, growth factors, transcription factors, mitochondrial and glycolytic-related factors, and other factors during spheroid generation [44]. Furthermore, ingenuity pathway analysis (IPA) suggested that STAT3 is the master upstream regulator for inducing 3T3-L1 spheroid generation by comparing 2D and 3D cultured 3T3-L1 preadipocytes [6]. Since STAT3 has been shown to be involved in some gravity-induced biological effects Since STAT3 has been shown to be involved in some gravity-induced biological effects [77,78,79], this suggests that gravity force pivotally affects spheroid generation. In addition, our recent RNA sequencing analysis before and after adipogenic differentiation showed that genes related to adipogenesis and lipid metabolism were altered in 3T3L-spheroids compared to those in 2D cultured 3T3-L1 cells, suggesting that a spheroid environment may be more suitable for adipogenesis of 3T3-L1 cells [56].
Figure 6. Representative images of spheroids obtained from various non-cancerous cells. HTM: human trabecular meshwork cell, HconF: human conjunctival fibroblast, G-OF: Graves-related human orbital fibroblast, N-OF: non-Graves-related human orbital fibroblast, 3T3-L1: 3T3-L1 mouse preadipocyte, HSSF: human scleral stromal fibroblast, H9c2: H9c2 rat cardiomyoblast. Scale bar: 100 μm.
Figure 6. Representative images of spheroids obtained from various non-cancerous cells. HTM: human trabecular meshwork cell, HconF: human conjunctival fibroblast, G-OF: Graves-related human orbital fibroblast, N-OF: non-Graves-related human orbital fibroblast, 3T3-L1: 3T3-L1 mouse preadipocyte, HSSF: human scleral stromal fibroblast, H9c2: H9c2 rat cardiomyoblast. Scale bar: 100 μm.
Cells 13 01549 g006

3.2. Study for Cardiology Using H9c2 Cells (Figure 6)

The metabolic aspects of cardiomyocytes are known to be different from those of other myocytes as well as non-myocytes [80]. In fact, it has been shown that matured cardiomyocytes require most of their energy from fatty acid oxidation rather than glycolysis [81,82] and that cellular metabolic functions are deteriorated by aging and/or cardiac functional defects [83,84]. For instance, insufficient oxygen (O2) perfusion into cardiac muscle tissues in the case of myocardial infarction or other etiology induces heart failure [85], suggesting that oxygen (O2) supply is an indispensable key factor for maintenance of the physiological functions of the heart. For studying the pathophysiology of the heart and for drug screening of the heart, much attention has been paid to in vitro 3D cell culture models using cardiomyocytes [86,87] for replicating in vivo spatial environments in addition to the standardized two-dimensional (2D) cell culture method.
H9c2 cardiomyoblasts originating from embryonic rat ventricular tissue are the most commonly used cells for cardiomyocyte research, and these cells have been applied to studies using different 3D culture methods including various scaffold-assisted methods [88,89,90,91,92,93,94,95,96,97,98,99], magnetic particles-assisted method [100] and floating culture using a rotating flask [101] in addition to hanging drop culture [102]. In those studies, in vitro 3D culture models were used cardiotoxicity screening and for investigation of drug-induced effects [89,100,101,102], apoptosis and cell survival [92,94], physical properties [91,102], cardiac homeostasis [88], myogenesis, differentiation and proliferation [93,96,97,99], cardiac disease modeling [98,102] and regeneration [95,96]. Great advantages of the 3D culture models compared to conventional 2D planar cell culture were shown in most of those studies. In fact, in our recent study, spontaneous expression of gap junction molecules and hypoxia-inducible factors (HIFs) were observed in H9c2 spheroids, but not in 2D H9c2 cells, and mitochondrial and glycolytic metabolic functions in H9c2 cells were significantly different in 2D and 3D culture conditions, suggesting that H9c2 spheroids indeed replicate essential physiological elements of cell-to-cell interaction in cardiomyocytes and O2 gradients in cardiac wall tissue [102]. Collectively, the results of studies indicated that an in vitro model of H9c2 spheroids is a useful in vitro model for understanding cardiac pathophysiology.

4. Studies for Ocular Pathophysiology

Three-dimensional cell cultures have been carried out using various parts of ocular tissues, and remarkable progress has recently been made in 3D organoid culture, but not spheroid culture, using human induced pluripotent stem cells (hiPSCs) for future retinal regenerative medicine [103]. Alternatively, spheroid culture methods have contributed to a better understanding of pathophysiology as well as various drug-induced effects in the ophthalmology research field as described in the following.

4.1. Ocular Surface and Sclera

In vitro spheroid models have been established using lacrimal gland (LG) cells and conjunctival cells by static suspension culture [104,105] and magnetic particle-assisted culture [106] to study tear-film dynamics and dry eye pathogenesis. Alternatively, our group established an in vitro subconjunctival fibrosis model using human conjunctival fibroblasts (HconF, Figure 6) by a hanging drop culture method to study the effects of various drugs including transforming growth factor-β (TGF-β isoforms [107], fibroblast growth factor-2 (FGF-2) [108], Rho-associated coiled-coil forming kinase (ROCK) inhibitors [109], all-trans retinoic acid (ATRA) [110], prostaglandin EP2 and FP2 agonists [111] and rosiglitazone [109]. For corneal disease modeling, spheroids were generated from bovine corneal endothelial cells by using a hydrogel-assisted method [112,113]. In addition, our group generated spheroids using human corneal stromal fibroblasts by a hanging drop culture method and we used the spheroids to study drug-induced effects of an EP2 agonist [114] and ROCK inhibitors [115]. We also established in vitro myopia models using spheroids that were generated from human scleral stroma fibroblasts (Figure 6) [116].

4.2. Intraocular Segments

To mimic in vivo microenvironments of the ocular lens, a particle-assisted method was used for generation of spheroids using lens epithelial cells [117]. To study aqueous humor (AH) dynamics related to glaucoma pathogenesis, various in vitro 3D models that replicate the trabecular meshwork (TM), which is the critical segment to produce AH outflow resistance during the AH conventional outflow route, were used (Figure 6) [118,119]. Scaffold-assisted 3D culture of TGF-β2 or dexamethasone (DEX)-treated human TM (HTM) cells was used in most of those studies, and those studies showed overexpressed ECM, defective phagocytosis of HTM cells resulting in increased resistance to AH outflow [120,121,122] and increase in the production of reactive oxidative species (ROS) [123]. Our group also obtained similar results using a hanging drop culture method without any scaffolds, and we successfully established in vitro primary open glaucoma (POAG) TM and steroid-induced glaucoma (SG) TM in the presence of TGF-β2 and DEX, respectively [8]. These in vitro HTM spheroid models were also used for studies to evaluate effects on the HTM of various drugs and factors including prostaglandin FP and EP2 agonist [124,125,126,127], autotaxin [128], ATRA [58], benzalkonium chloride (BAC) [129], ROCK inhibitors [130,131,132,133], α2-adrenergic agonist [134] and TGF-β isoforms [135,136,137]. The retinal pigment epithelium (RPE) is a highly organized and polarized single-cell sheet adjacent to the retinal choroid and photoreceptor outer segment (OS), and it functionally plays critical roles in the visual retinoid cycle, phagocytosis of the OS, formation of the outer blood-retinal barrier (oBRB) with Bruch’s membrane and retinal choroid, and transportation of various nutrients [138]. In turn, dysfunctions in the RPE, which can lead to disruption of the oBRB and photoreceptor cell death, are implicated in serious retinal diseases such as age-related macular degeneration (ARMD) and retinitis pigmentosa [139,140]. Therefore, for targeting the RPE, pilot therapies using hiPSCs or human embryonic stem cells (hESCs) have already been started [141,142,143]. To support this issue, various studies using stem cell-based 3D culture of the RPE have been carried out [144,145,146,147,148,149]. In addition, spheroid culture has also been used to replicate RPE-induced angiogenesis [150], proliferative vitreoretinopathy [151] and drug-induced toxicity to the RPE [152].

4.3. Orbit

Among various orbital tissues, the lacrimal gland (LG) and orbital fatty tissues have been frequently used for generation of in vitro spheroid models.

4.3.1. Lacrimal Gland (LG)

LG dysfunction is one of major causes of dry eye disease and is therefore a possible promising therapeutic target for dry eye disease [153]. For this purpose, 3D organoid cultures have mainly been performed using stem cells [154,155,156,157,158] or digested cells from human LG tissue [159]. Spheroids of LG-related cells including human epithelial cells, endothelial cells and mesenchymal stem cells [160] or rabbit conjunctival epithelium and lacrimal gland cells [104] have also been generated by using a static suspension culture method.

4.3.2. Orbital Fatty Tissue

A major disease of orbital fatty tissue is Graves’ orbitopathy (GO), for which no animal model has been developed. A suitable in vitro model is therefore necessary for obtaining a better understanding of the disease etiology and for developing a new therapy. For this purpose, by using a static suspension method, spheroids were generated from human orbital fibroblasts (OFs) obtained from patients without GO (N-OFs) and patients with GO (G-OFs) (Figure 6) [161]. Our group has also generated spheroids from N-OFs and G-OFs spheroids by a hanging drop culture method and we found several new biological differences between N-OFs and G-OFs [5,162,163,164,165,166,167,168]. As for the molecular pathology of GO, autoimmune-based overstimulation of the thyroid-stimulating hormone (TSH) receptor (TSHR) and insulin-like growth factor 1 (IGF1) receptor (IGF1R), both of which are expressed in OFs derived from the monocyte lineage and form a complex by a crosstalk mechanism [169], induces inflammation and differentiation into adipocytes and myofibroblasts as well as hyaluronan production [170,171,172].
Deepening of the upper eyelid sulcus (DUES) is known as a notable adverse effect of prostaglandin-derived anti-glaucoma medication with cosmetically annoying problems. Our group first established an in vitro spheroid model of DUES using N-OFs and 3T3-L1 cells and we provided valuable information for selection of anti-glaucoma medications to avoid DUES [42,72,73,74,75,162,163,164,165,166,167,168,173,174,175].

5. Three-Dimensional Spheroids Obtained from Various Cancerous Cells

As in the case of non-cancerous tissues, 3D culture systems are also being increasingly used in tumor research for establishing essential strategies [176,177]. As shown in Table 1, various methods have been used for generating single cell type-derived spheroids from cancerous cells as in the case of non-cancerous cells, and the spheroids obtained have been mostly used for drug screening to evaluate toxicity and to discover new anti-cancer drugs [176,177]. Among the various methods for generating spheroids from tumor cells, our group used a hanging drop culture method as described above, and we found that 3D configurations of non-cancerous cell-derived spheroids (globe shape, Figure 6) were significantly different from those of cancerous cell-derived spheroids (non-globe shape, Figure 7) [5,7,55,72,109,125]. A non-globe shape appearance was also observed in some tumor cells-derived spheroids generated by using a hanging drop culture method by other research groups [178,179,180,181,182]. Since, as stated above, the hanging drop culture method is the simplest method for generating spheroids that requires only gravity force and buoyant force, the non-spherical shape of spheroids may be caused by deviations from sphericity of the aggregation of cancerous cells. If this speculation is correct, such deviations from sphericity in the tumor cell-derived spheroids could be a key index that reflects some important biological aspects of tumorigenesis.

5.1. Malignant Melanoma (MM)

An MM originating from melanocytes is a very aggressive malignant tumor and occurs in various organs with melanin pigment including the skin, eye choroid, gastrointestinal tract, genitalia, sinuses, and meninges [183]. As recent therapies for MM, targeted inhibitors of BRAFV600E, MEK kinases and immune-based monoclonal antibody drugs have been effectively used for treating patients with MM [184,185,186,187]. However, a relatively large percentage of patients with MM show resistance to these treatments, and the development of an additional therapeutic strategy and the development of a method for control of drug resistance by using a reliable in vitro model such as a spheroid model are therefore necessary [188,189,190]. Recently, our group also prepared spheroids originating from 5 different MM cell lines including SK-mel-24, MM418, A375, WM266-4 and SM2-1 by using a hanging drop culture method [7]. Interestingly, the spheroid configurations were deformed and the degree of the deformity was significantly diverse among the 5 MM cell lines. In addition, cellular metabolic functions assessed by using a seahorse bioanalyzer were also different among the MM cell lines. RNA sequencing analysis of two distinct cell lines, WM266-4 and SK-mel-24, in the 3D appearances suggested that KRAS and SOX2 were potential master regulatory genes causing these diverse 3D configurations. Furthermore, knockdown of both factors reduced the morphological deformity and improved functional characteristics. In addition, ML329, a small molecule of a specific inhibitor of microphthalmia-associated transcription factor (MITF) that regulates the expression of various pigmentation-related genes including genes encoding tyrosinase [191] and is critically involved in the pathogenesis of MM [192,193], significantly and diversely modulated the spheroid appearance and cellular metabolic functions of various MM cell lines [54]. Collectively, the findings suggest that spheroid configuration may become a potential key indicator for pathophysiological activities associated with MM.

5.2. Oral Squamous Cell Carcinoma (OSCC)

Oral squamous cell carcinoma (OSCC) is a mucosal malignancy of the lips and oral cavity and is one of main causes of mortality worldwide [194,195]. The 5-year survival rate of patients with OSCC is less than 50% regardless of the stage [196,197,198]. In this research field, 3-D cell cultures have been increasingly used with the aim of establishing physiologically relevant in vitro models for studying various aspects of oral cancer [199]. As in the case of MM, deformed appearance of spheroids generated by a hanging drop culture method was also observed in pathologically different squamous cell carcinoma (OSCC) cell lines and one oral adeno squamous carcinoma (OAdSCC) cell line, and the degree of deformity correlated with the efficacy of the CDDP-induced cytotoxicity among the cell lines. Therefore, we suggested that spheroid architectures may be useful indicators for estimating drug sensitivity of OSCC and OAdSCC malignancies [55]. Collectively, the findings suggested that in vitro spheroid models may be more sensitive and suitable than in vitro conventional 2D culture models for achieving more precise screening, estimating appropriate dosage and estimating efficacies of new candidate antitumor agents [200,201,202,203,204].

5.3. Cancer-Associated Fibroblasts (CAFs)

Tumor surrounding microenvironments (TSM) have attracted much interest due to their crucial roles in tumorigenesis, tumor progression, and metastasis [205,206,207,208]. In the research field of TSM, it is thought the use of 3D cell culture methods can replicate the tumor microenvironment landscape and are therefore suitable for screening immunomodulatory drugs [209]. Among TSM, specific fibroblasts called “cancer-associated fibroblasts (CAFs)” are known to be critical participants that are mainly responsible for producing various extracellular matrix (ECM) proteins in the TSM [210,211,212,213,214]. Several recent studies have shown that fibroblasts not related to cancer (non-CAFs) may be transformed into CAFs in the TSM [215], and, in fact, recent transcriptome analyses have shown different gene expression profiles in non-CAFs and CAFs [206,216,217,218,219]. However, despite great pathological contributions of CAFs to cancer biology, the only method for identifying CAFs is to use certain markers with low specificities such as α-smooth muscle actin (αSMA) and fibroblast-activating protein-α (FAPα) [220]. In our recent study using a hanging drop culture method, we found that, unlike globe-shaped non-CAF spheroids, non-globe-shaped CAF spheroids were generated from four lines of CAFs obtained from the neighborhood of OSCCs and that the deformity and stiffness of the spheroids were different [9]. Therefore, the findings suggest that a spheroid culture method provides a new strategy for characterizing and specifying CAFs.

6. Summary of Current Concepts of the Biological Significance of Hanging Drop Cultures of Single Cell Type-Derived Spheroids and Their Future Prospects

In addition to conventional 2D cell culture, various 3D cell culture techniques provide new possibilities for unidentified biological aspects for advancing biological research. In fact, the use of various 3D culture methods has resulted in the establishment of better strategies for testing drug efficacy, the development of disease models that are closer to their in vivo conditions to facilitate a deeper understanding of disease etiology, and a better way to deal with stem cells for future regeneration medicine. However, the benefits and disadvantages of 3D cell cultures, even the simplest single cell type-derived spheroid culture, have not been fully characterized. As an identified disadvantage of spheroids, the possible inability to follow hanging-drop cultures over time should be considered compared to in vitro 2D culture models, and proper timing should be therefore adjusted for the most suitable experimental purpose. As shown in this review, our experience in the use of a hanging drop culture method to generate spheroids from various cells in proper timing has shown that the biological natures of spheroids are totally different from those of 2D cultured cells. The benefits of spheroids will lead to further improvements for tissue engineering, modeling of non-cancerous and cancerous diseases, and testing biological functions and drug-induced effects.

Author Contributions

H.O. and F.H. designed the structure of this review and wrote the manuscript. M.W., T.S., N.N., A.U., M.H., T.Y., H.S., A.M., K.T., H.U. and M.F. prepared the figures and table, and provided critical comments on the manuscript. 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

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Sumbalova Koledova, Z. 3D Cell Culture: Techniques For and Beyond Organoid Applications. Methods Mol. Biol. 2024, 2764, 1–12. [Google Scholar] [PubMed]
  2. Boucherit, N.; Gorvel, L.; Olive, D. 3D Tumor Models and Their Use for the Testing of Immunotherapies. Front. Immunol. 2020, 11, 603640. [Google Scholar] [CrossRef] [PubMed]
  3. 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]
  4. Lee, S.Y.; Koo, I.S.; Hwang, H.J.; Lee, D.W. In Vitro three-dimensional (3D) cell culture tools for spheroid and organoid models. SLAS Discov. Adv. Life Sci. R. D 2023, 28, 119–137. [Google Scholar] [CrossRef]
  5. Hikage, F.; Atkins, S.; Kahana, A.; Smith, T.J.; Chun, T.H. HIF2A-LOX Pathway Promotes Fibrotic Tissue Remodeling in Thyroid-Associated Orbitopathy. Endocrinology 2019, 160, 20–35. [Google Scholar] [CrossRef]
  6. Ohguro, H.; Ida, Y.; Hikage, F.; Umetsu, A.; Ichioka, H.; Watanabe, M.; Furuhashi, M. STAT3 Is the Master Regulator for the Forming of 3D Spheroids of 3T3-L1 Preadipocytes. Cells 2022, 11, 300. [Google Scholar] [CrossRef]
  7. Ohguro, H.; Watanabe, M.; Sato, T.; Hikage, F.; Furuhashi, M.; Okura, M.; Hida, T.; Uhara, H. 3D Spheroid Configurations Are Possible Indictors for Evaluating the Pathophysiology of Melanoma Cell Lines. Cells 2023, 12, 759. [Google Scholar] [CrossRef]
  8. Watanabe, M.; Ida, Y.; Ohguro, H.; Ota, C.; Hikage, F. Establishment of appropriate glaucoma models using dexamethasone or TGFβ2 treated three-dimension (3D) cultured human trabecular meshwork (HTM) cells. Sci. Rep. 2021, 11, 19369. [Google Scholar] [CrossRef]
  9. Nishikiori, N.; Takada, K.; Sato, T.; Miyamoto, S.; Watanabe, M.; Hirakawa, Y.; Sekiguchi, S.; Furuhashi, M.; Yorozu, A.; Takano, K.; et al. Physical Properties and Cellular Metabolic Characteristics of 3D Spheroids Are Possible Definitive Indices for the Biological Nature of Cancer-Associated Fibroblasts. Cells 2023, 12, 2160. [Google Scholar] [CrossRef]
  10. Ho, T.C.; Chang, C.C.; Chan, H.P.; Chung, T.W.; Shu, C.W.; Chuang, K.P.; Duh, T.H.; Yang, M.H.; Tyan, Y.C. Hydrogels: Properties and Applications in Biomedicine. Molecules 2022, 27, 2902. [Google Scholar] [CrossRef]
  11. Lv, D.; Hu, Z.; Lu, L.; Lu, H.; Xu, X. Three-dimensional cell culture: A powerful tool in tumor research and drug discovery. Oncol. Lett. 2017, 14, 6999–7010. [Google Scholar] [CrossRef] [PubMed]
  12. Lee, D.W.; Choi, Y.S.; Seo, Y.J.; Lee, M.Y.; Jeon, S.Y.; Ku, B.; Kim, S.; Yi, S.H.; Nam, D.H. High-throughput screening (HTS) of anticancer drug efficacy on a micropillar/microwell chip platform. Anal. Chem. 2014, 86, 535–542. [Google Scholar] [CrossRef] [PubMed]
  13. Ho, S.S.; Keown, A.T.; Addison, B.; Leach, J.K. Cell Migration and Bone Formation from Mesenchymal Stem Cell Spheroids in Alginate Hydrogels Are Regulated by Adhesive Ligand Density. Biomacromolecules 2017, 18, 4331–4340. [Google Scholar] [CrossRef] [PubMed]
  14. Murphy, K.C.; Whitehead, J.; Zhou, D.; Ho, S.S.; Leach, J.K. Engineering fibrin hydrogels to promote the wound healing potential of mesenchymal stem cell spheroids. Acta Biomater. 2017, 64, 176–186. [Google Scholar] [CrossRef] [PubMed]
  15. Lewis, N.S.; Lewis, E.E.; Mullin, M.; Wheadon, H.; Dalby, M.J.; Berry, C.C. Magnetically levitated mesenchymal stem cell spheroids cultured with a collagen gel maintain phenotype and quiescence. J. Tissue Eng. 2017, 8, 2041731417704428. [Google Scholar] [CrossRef]
  16. Gwon, K.; Kim, E.; Tae, G. Heparin-hyaluronic acid hydrogel in support of cellular activities of 3D encapsulated adipose derived stem cells. Acta Biomater. 2017, 49, 284–295. [Google Scholar] [CrossRef]
  17. 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, 3–18. [Google Scholar] [CrossRef]
  18. 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]
  19. 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]
  20. Chen, Y.H.; Ku, Y.H.; Wang, K.C.; Chiang, H.C.; Hsu, Y.P.; Cheng, M.T.; Wang, C.S.; Wee, Y. Bioinspired Sandcastle Worm-Derived Peptide-Based Hybrid Hydrogel for Promoting the Formation of Liver Spheroids. Gels 2022, 8, 149. [Google Scholar] [CrossRef]
  21. Li, Y.; Kumacheva, E. Hydrogel microenvironments for cancer spheroid growth and drug screening. Sci. Adv. 2018, 4, eaas8998. [Google Scholar] [CrossRef] [PubMed]
  22. Flemming, H.C.; van Hullebusch, E.D.; Neu, T.R.; Nielsen, P.H.; Seviour, T.; Stoodley, P.; Wingender, J.; Wuertz, S. The biofilm matrix: Multitasking in a shared space. Nat. Rev. Microbiol. 2023, 21, 70–86. [Google Scholar] [CrossRef]
  23. Cometta, S.; Hutmacher, D.W.; Chai, L. In vitro models for studying implant-associated biofilms—A review from the perspective of bioengineering 3D microenvironments. Biomaterials 2024, 309, 122578. [Google Scholar] [CrossRef]
  24. Huang, G.S.; Dai, L.G.; Yen, B.L.; Hsu, S.H. Spheroid formation of mesenchymal stem cells on chitosan and chitosan-hyaluronan membranes. Biomaterials 2011, 32, 6929–6945. [Google Scholar] [CrossRef] [PubMed]
  25. Yeh, H.Y.; Liu, B.H.; Hsu, S.H. The calcium-dependent regulation of spheroid formation and cardiomyogenic differentiation for MSCs on chitosan membranes. Biomaterials 2012, 33, 8943–8954. [Google Scholar] [CrossRef] [PubMed]
  26. Kim, Y.; Baipaywad, P.; Jeong, Y.; Park, H. Incorporation of gelatin microparticles on the formation of adipose-derived stem cell spheroids. Int. J. Biol. Macromol. 2018, 110, 472–478. [Google Scholar] [CrossRef]
  27. Hayashi, K.; Tabata, Y. Preparation of stem cell aggregates with gelatin microspheres to enhance biological functions. Acta Biomater. 2011, 7, 2797–2803. [Google Scholar] [CrossRef]
  28. Baraniak, P.R.; Cooke, M.T.; Saeed, R.; Kinney, M.A.; Fridley, K.M.; McDevitt, T.C. Stiffening of human mesenchymal stem cell spheroid microenvironments induced by incorporation of gelatin microparticles. J. Mech. Behav. Biomed. Mater. 2012, 11, 63–71. [Google Scholar] [CrossRef]
  29. Bratt-Leal, A.M.; Nguyen, A.H.; Hammersmith, K.A.; Singh, A.; McDevitt, T.C. A microparticle approach to morphogen delivery within pluripotent stem cell aggregates. Biomaterials 2013, 34, 7227–7235. [Google Scholar] [CrossRef]
  30. Anil-Inevi, M.; Yaman, S.; Yildiz, A.A.; Mese, G.; Yalcin-Ozuysal, O.; Tekin, H.C.; Ozcivici, E. Biofabrication of in situ Self Assembled 3D Cell Cultures in a Weightlessness Environment Generated using Magnetic Levitation. Sci. Rep. 2018, 8, 7239. [Google Scholar] [CrossRef]
  31. Meng, R.; Xu, H.Y.; Di, S.M.; Shi, D.Y.; Qian, A.R.; Wang, J.F.; Shang, P. Human mesenchymal stem cells are sensitive to abnormal gravity and exhibit classic apoptotic features. Acta Biochim. Biophys. Sin. 2011, 43, 133–142. [Google Scholar] [CrossRef] [PubMed]
  32. Sytkowski, A.J.; Davis, K.L. Erythroid cell growth and differentiation in vitro in the simulated microgravity environment of the NASA rotating wall vessel bioreactor. Vitr. Cell. Dev. Biol. Anim. 2001, 37, 79–83. [Google Scholar] [CrossRef]
  33. Zanoni, M.; Piccinini, F.; Arienti, C.; Zamagni, A.; Santi, S.; Polico, R.; Bevilacqua, A.; Tesei, A. 3D tumor spheroid models for in vitro therapeutic screening: A systematic approach to enhance the biological relevance of data obtained. Sci. Rep. 2016, 6, 19103. [Google Scholar] [CrossRef] [PubMed]
  34. 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]
  35. Zitzmann, F.D.; Schmidt, S.; Frank, R.; Weigel, W.; Meier, M.; Jahnke, H.G. Microcavity well-plate for automated parallel bioelectronic analysis of 3D cell cultures. Biosens. Bioelectron. 2024, 250, 116042. [Google Scholar] [CrossRef]
  36. Kukla, D.A.; Belair, D.G.; Stresser, D.M. Evaluation and Optimization of a Microcavity Plate-Based Human Hepatocyte Spheroid Model for Predicting Clearance of Slowly Metabolized Drug Candidates. Drug Metab. Dispos. Biol. Fate Chem. 2024, 52, 797–812. [Google Scholar] [CrossRef]
  37. Kim, J.B. Three-dimensional tissue culture models in cancer biology. Semin. Cancer Biol. 2005, 15, 365–377. [Google Scholar] [CrossRef]
  38. 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] [PubMed]
  39. 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]
  40. 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]
  41. Tung, Y.C.; Hsiao, A.Y.; Allen, S.G.; Torisawa, Y.S.; 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]
  42. Ida, Y.; Hikage, F.; Ohguro, H. ROCK inhibitors enhance the production of large lipid-enriched 3D organoids of 3T3-L1 cells. Sci. Rep. 2021, 11, 5479. [Google Scholar] [CrossRef]
  43. Duraj, T.; Carrión-Navarro, J.; Seyfried, T.N.; García-Romero, N.; Ayuso-Sacido, A. Metabolic therapy and bioenergetic analysis: The missing piece of the puzzle. Mol. Metab. 2021, 54, 101389. [Google Scholar] [CrossRef]
  44. Rodríguez-Enríquez, S.; Gallardo-Pérez, J.C.; Avilés-Salas, A.; Marín-Hernández, A.; Carreño-Fuentes, L.; Maldonado-Lagunas, V.; Moreno-Sánchez, R. Energy metabolism transition in multi-cellular human tumor spheroids. J. Cell. Physiol. 2008, 216, 189–197. [Google Scholar] [CrossRef]
  45. Mandujano-Tinoco, E.A.; Gallardo-Pérez, J.C.; Marín-Hernández, A.; Moreno-Sánchez, R.; Rodríguez-Enríquez, S. Anti-mitochondrial therapy in human breast cancer multi-cellular spheroids. Biochim. Biophys. Acta 2013, 1833, 541–551. [Google Scholar] [CrossRef]
  46. Marín-Hernández, Á.; Gallardo-Pérez, J.C.; Hernández-Reséndiz, I.; Del Mazo-Monsalvo, I.; Robledo-Cadena, D.X.; Moreno-Sánchez, R.; Rodríguez-Enríquez, S. Hypoglycemia Enhances Epithelial-Mesenchymal Transition and Invasiveness, and Restrains the Warburg Phenotype, in Hypoxic HeLa Cell Cultures and Microspheroids. J. Cell. Physiol. 2017, 232, 1346–1359. [Google Scholar] [CrossRef] [PubMed]
  47. Bloch, K.; Smith, H.; van Hamel Parsons, V.; Gavaghan, D.; Kelly, C.; Fletcher, A.; Maini, P.; Callaghan, R. Metabolic alterations during the growth of tumour spheroids. Cell Biochem. Biophys. 2014, 68, 615–628. [Google Scholar] [CrossRef]
  48. Javed, Z.; Worley, B.L.; Stump, C.; Shimko, S.S.; Crawford, L.C.; Mythreye, K.; Hempel, N. Optimization of Extracellular Flux Assay to Measure Respiration of Anchorage-independent Tumor Cell Spheroids. Bio-Protoc. 2022, 12, e4321. [Google Scholar] [CrossRef] [PubMed]
  49. Wang, C.; Yang, Z.; Xu, E.; Shen, X.; Wang, X.; Li, Z.; Yu, H.; Chen, K.; Hu, Q.; Xia, X.; et al. Apolipoprotein C-II induces EMT to promote gastric cancer peritoneal metastasis via PI3K/AKT/mTOR pathway. Clin. Transl. Med. 2021, 11, e522. [Google Scholar] [CrossRef]
  50. Bonuccelli, G.; Sotgia, F.; Lisanti, M.P. Identification of natural products and FDA-approved drugs for targeting cancer stem cell (CSC) propagation. Aging 2022, 14, 9466–9483. [Google Scholar] [CrossRef]
  51. Campioni, G.; Pasquale, V.; Busti, S.; Ducci, G.; Sacco, E.; Vanoni, M. An Optimized Workflow for the Analysis of Metabolic Fluxes in Cancer Spheroids Using Seahorse Technology. Cells 2022, 11, 866. [Google Scholar] [CrossRef] [PubMed]
  52. Ichioka, H.; Hirohashi, Y.; Sato, T.; Furuhashi, M.; Watanabe, M.; Ida, Y.; Hikage, F.; Torigoe, T.; Ohguro, H. G-Protein-Coupled Receptors Mediate Modulations of Cell Viability and Drug Sensitivity by Aberrantly Expressed Recoverin 3 within A549 Cells. Int. J. Mol. Sci. 2023, 24, 771. [Google Scholar] [CrossRef] [PubMed]
  53. Nakamura, H.; Watanabe, M.; Takada, K.; Sato, T.; Hikage, F.; Umetsu, A.; Muramatsu, J.; Furuhashi, M.; Ohguro, H. Modulation of Epithelial-Mesenchymal Transition Is a Possible Underlying Mechanism for Inducing Chemoresistance in MIA PaCa-2 Cells against Gemcitabine and Paclitaxel. Biomedicines 2024, 12, 1011. [Google Scholar] [CrossRef]
  54. Nishikiori, N.; Watanabe, M.; Sato, T.; Furuhashi, M.; Okura, M.; Hida, T.; Uhara, H.; Ohguro, H. Significant and Various Effects of ML329-Induced MITF Suppression in the Melanoma Cell Line. Cancers 2024, 16, 263. [Google Scholar] [CrossRef]
  55. Miyamoto, S.; Nishikiori, N.; Sato, T.; Watanabe, M.; Umetsu, A.; Tsugeno, Y.; Hikage, F.; Sasaya, T.; Kato, H.; Ogi, K.; et al. Three-Dimensional Spheroid Configurations and Cellular Metabolic Properties of Oral Squamous Carcinomas Are Possible Pharmacological and Pathological Indicators. Cancers 2023, 15, 2793. [Google Scholar] [CrossRef] [PubMed]
  56. Endo, K.; Sato, T.; Umetsu, A.; Watanabe, M.; Hikage, F.; Ida, Y.; Ohguro, H.; Furuhashi, M. 3D culture induction of adipogenic differentiation in 3T3-L1 preadipocytes exhibits adipocyte-specific molecular expression patterns and metabolic functions. Heliyon 2023, 9, e20713. [Google Scholar] [CrossRef]
  57. Watanabe, M.; Sato, T.; Tsugeno, Y.; Umetsu, A.; Suzuki, S.; Furuhashi, M.; Ida, Y.; Hikage, F.; Ohguro, H. Human Trabecular Meshwork (HTM) Cells Treated with TGF-β2 or Dexamethasone Respond to Compression Stress in Different Manners. Biomedicines 2022, 10, 1338. [Google Scholar] [CrossRef]
  58. Watanabe, M.; Sato, T.; Tsugeno, Y.; Higashide, M.; Furuhashi, M.; Umetsu, A.; Suzuki, S.; Ida, Y.; Hikage, F.; Ohguro, H. All-trans Retinoic Acids Synergistically and Beneficially Affect In Vitro Glaucomatous Trabecular Meshwork (TM) Models Using 2D and 3D Cell Cultures of Human TM Cells. Int. J. Mol. Sci. 2022, 23, 9912. [Google Scholar] [CrossRef]
  59. Tsugeno, Y.; Sato, T.; Watanabe, M.; Furuhashi, M.; Umetsu, A.; Ida, Y.; Hikage, F.; Ohguro, H. Benzalkonium Chloride, Even at Low Concentrations, Deteriorates Intracellular Metabolic Capacity in Human Conjunctival Fibroblasts. Biomedicines 2022, 10, 2315. [Google Scholar] [CrossRef]
  60. Ambele, M.A.; Dessels, C.; Durandt, C.; Pepper, M.S. Genome-wide analysis of gene expression during adipogenesis in human adipose-derived stromal cells reveals novel patterns of gene expression during adipocyte differentiation. Stem Cell Res. 2016, 16, 725–734. [Google Scholar] [CrossRef]
  61. Sadowski, H.B.; Wheeler, T.T.; Young, D.A. Gene expression during 3T3-L1 adipocyte differentiation. Characterization of initial responses to the inducing agents and changes during commitment to differentiation. J. Biol. Chem. 1992, 267, 4722–4731. [Google Scholar] [CrossRef] [PubMed]
  62. Aulthouse, A.L.; Freeh, E.; Newstead, S.; Stockert, A.L. Part 1: A Novel Model for Three-Dimensional Culture of 3T3-L1 Preadipocytes Stimulates Spontaneous Cell Differentiation Independent of Chemical Induction Typically Required in Monolayer. Nutr. Metab. Insights 2019, 12, 1178638819841399. [Google Scholar] [CrossRef] [PubMed]
  63. Chun, T.H.; Inoue, M. 3-D adipocyte differentiation and peri-adipocyte collagen turnover. Methods Enzymol. 2014, 538, 15–34. [Google Scholar] [PubMed]
  64. Davidenko, N.; Campbell, J.J.; Thian, E.S.; Watson, C.J.; Cameron, R.E. Collagen-hyaluronic acid scaffolds for adipose tissue engineering. Acta Biomater. 2010, 6, 3957–3968. [Google Scholar] [CrossRef]
  65. Fischbach, C.; Seufert, J.; Staiger, H.; Hacker, M.; Neubauer, M.; Göpferich, A.; Blunk, T. Three-dimensional in vitro model of adipogenesis: Comparison of culture conditions. Tissue Eng. 2004, 10, 215–229. [Google Scholar] [CrossRef]
  66. Fischbach, C.; Spruss, T.; Weiser, B.; Neubauer, M.; Becker, C.; Hacker, M.; Göpferich, A.; Blunk, T. Generation of mature fat pads in vitro and in vivo utilizing 3-D long-term culture of 3T3-L1 preadipocytes. Exp. Cell Res. 2004, 300, 54–64. [Google Scholar] [CrossRef]
  67. Josan, C.; Kakar, S.; Raha, S. Matrigel® enhances 3T3-L1 cell differentiation. Adipocyte 2021, 10, 361–377. [Google Scholar] [CrossRef]
  68. Daquinag, A.C.; Souza, G.R.; Kolonin, M.G. Adipose tissue engineering in three-dimensional levitation tissue culture system based on magnetic nanoparticles. Tissue Eng. Part. C Methods 2013, 19, 336–344. [Google Scholar] [CrossRef]
  69. Tseng, H.; Daquinag, A.C.; Souza, G.R.; Kolonin, M.G. Three-Dimensional Magnetic Levitation Culture System Simulating White Adipose Tissue. Methods Mol. Biol. 2018, 1773, 147–154. [Google Scholar]
  70. Turner, P.A.; Garrett, M.R.; Didion, S.P.; Janorkar, A.V. Spheroid Culture System Confers Differentiated Transcriptome Profile and Functional Advantage to 3T3-L1 Adipocytes. Ann. Biomed. Eng. 2018, 46, 772–787. [Google Scholar] [CrossRef]
  71. Turner, P.A.; Gurumurthy, B.; Bailey, J.L.; Elks, C.M.; Janorkar, A.V. Adipogenic Differentiation of Human Adipose-Derived Stem Cells Grown as Spheroids. Process Biochem. 2017, 59, 312–320. [Google Scholar] [CrossRef]
  72. Ida, Y.; Hikage, F.; Itoh, K.; Ida, H.; Ohguro, H. Prostaglandin F2α agonist-induced suppression of 3T3-L1 cell adipogenesis affects spatial formation of extra-cellular matrix. Sci. Rep. 2020, 10, 7958. [Google Scholar] [CrossRef]
  73. Ida, Y.; Hikage, F.; Umetsu, A.; Ida, H.; Ohguro, H. Omidenepag, a non-prostanoid EP2 receptor agonist, induces enlargement of the 3D organoid of 3T3-L1 cells. Sci. Rep. 2020, 10, 16018. [Google Scholar] [CrossRef]
  74. Ida, Y.; Watanabe, M.; Ohguro, H.; Hikage, F. Simultaneous Use of ROCK Inhibitors and EP2 Agonists Induces Unexpected Effects on Adipogenesis and the Physical Properties of 3T3-L1 Preadipocytes. Int. J. Mol. Sci. 2021, 22, 4648. [Google Scholar] [CrossRef]
  75. Ida, Y.; Watanabe, M.; Umetsu, A.; Ohguro, H.; Hikage, F. Addition of EP2 agonists to an FP agonist additively and synergistically modulates adipogenesis and the physical properties of 3D 3T3-L1 sphenoids. Prostaglandins Leukot. Essent. Fat. Acids 2021, 171, 102315. [Google Scholar] [CrossRef]
  76. Mery, B.; Vallard, A.; Rowinski, E.; Magne, N. High-throughput sequencing in clinical oncology: From past to present. Swiss Med. Wkly. 2019, 149, w20057. [Google Scholar] [CrossRef]
  77. Albi, E.; Curcio, F.; Lazzarini, A.; Floridi, A.; Cataldi, S.; Lazzarini, R.; Loreti, E.; Ferri, I.; Ambesi-Impiombato, F.S. A firmer understanding of the effect of hypergravity on thyroid tissue: Cholesterol and thyrotropin receptor. PLoS ONE 2014, 9, e98250. [Google Scholar] [CrossRef]
  78. Lee, S.G.; Lee, C.G.; Wu, H.M.; Oh, C.S.; Chung, S.W.; Kim, S.G. A load of mice to hypergravity causes AMPKα repression with liver injury, which is overcome by preconditioning loads via Nrf2. Sci. Rep. 2015, 5, 15643. [Google Scholar] [CrossRef]
  79. Cavey, T.; Pierre, N.; Nay, K.; Allain, C.; Ropert, M.; Loréal, O.; Derbré, F. Simulated microgravity decreases circulating iron in rats: Role of inflammation-induced hepcidin upregulation. Exp. Physiol. 2017, 102, 291–298. [Google Scholar] [CrossRef]
  80. Picca, A.; Mankowski, R.T.; Burman, J.L.; Donisi, L.; Kim, J.S.; Marzetti, E.; Leeuwenburgh, C. Mitochondrial quality control mechanisms as molecular targets in cardiac ageing. Nat. Rev. Cardiol. 2018, 15, 543–554. [Google Scholar] [CrossRef]
  81. Kolwicz, S.C., Jr.; Purohit, S.; Tian, R. Cardiac metabolism and its interactions with contraction, growth, and survival of cardiomyocytes. Circ. Res. 2013, 113, 603–616. [Google Scholar] [CrossRef]
  82. Chen, X.F.; Chen, X.; Tang, X. Short-chain fatty acid, acylation and cardiovascular diseases. Clin. Sci. 2020, 134, 657–676. [Google Scholar] [CrossRef]
  83. Heusch, G.; Libby, P.; Gersh, B.; Yellon, D.; Böhm, M.; Lopaschuk, G.; Opie, L. Cardiovascular remodelling in coronary artery disease and heart failure. Lancet 2014, 383, 1933–1943. [Google Scholar] [CrossRef]
  84. Patel, K.V.; Pandey, A.; de Lemos, J.A. Conceptual Framework for Addressing Residual Atherosclerotic Cardiovascular Disease Risk in the Era of Precision Medicine. Circulation 2018, 137, 2551–2553. [Google Scholar] [CrossRef]
  85. Poole, D.C.; Hirai, D.M.; Copp, S.W.; Musch, T.I. Muscle oxygen transport and utilization in heart failure: Implications for exercise (in)tolerance. Am. J. Physiol. Heart Circ. Physiol. 2012, 302, H1050–H1063. [Google Scholar] [CrossRef]
  86. Zuppinger, C. 3D Cardiac Cell Culture: A Critical Review of Current Technologies and Applications. Front. Cardiovasc. Med. 2019, 6, 87. [Google Scholar] [CrossRef]
  87. Arai, K.; Kitsuka, T.; Nakayama, K. Scaffold-based and scaffold-free cardiac constructs for drug testing. Biofabrication 2021, 13, 042001. [Google Scholar] [CrossRef]
  88. Navaee, F.; Renaud, P.; Kleger, A.; Braschler, T. Highly Efficient Cardiac Differentiation and Maintenance by Thrombin-Coagulated Fibrin Hydrogels Enriched with Decellularized Porcine Heart Extracellular Matrix. Int. J. Mol. Sci. 2023, 24, 2842. [Google Scholar] [CrossRef]
  89. Navaee, F.; Khornian, N.; Longet, D.; Heub, S.; Boder-Pasche, S.; Weder, G.; Kleger, A.; Renaud, P.; Braschler, T. A Three-Dimensional Engineered Cardiac In Vitro Model: Controlled Alignment of Cardiomyocytes in 3D Microphysiological Systems. Cells 2023, 12, 576. [Google Scholar] [CrossRef]
  90. Finosh, G.T.; Jayabalan, M.; Vandana, S.; Raghu, K.G. Hybrid alginate-polyester bimodal network hydrogel for tissue engineering—Influence of structured water on long-term cellular growth. Colloids Surf. B Biointerfaces 2015, 135, 855–864. [Google Scholar] [CrossRef]
  91. Chin, I.L.; Amos, S.E.; Jeong, J.H.; Hool, L.; Hwang, Y.; Choi, Y.S. Mechanosensation mediates volume adaptation of cardiac cells and spheroids in 3D. Mater. Today. Bio 2022, 16, 100391. [Google Scholar] [CrossRef]
  92. Hu, Y.; Jia, Y.; Wang, H.; Cao, Q.; Yang, Y.; Zhou, Y.; Tan, T.; Huang, X.; Zhou, Q. Low-intensity pulsed ultrasound promotes cell viability and inhibits apoptosis of H9C2 cardiomyocytes in 3D bioprinting scaffolds via PI3K-Akt and ERK1/2 pathways. J. Biomater. Appl. 2022, 37, 402–414. [Google Scholar] [CrossRef]
  93. Mei, C.; Chao, C.W.; Lin, C.W.; Li, S.T.; Wu, K.H.; Yang, K.C.; Yu, J. Three-dimensional spherical gelatin bubble-based scaffold improves the myotube formation of H9c2 myoblasts. Biotechnol. Bioeng. 2019, 116, 1190–1200. [Google Scholar] [CrossRef]
  94. Punia, K.; Bucaro, M.; Mancuso, A.; Cuttitta, C.; Marsillo, A.; Bykov, A.; L’Amoreaux, W.; Raja, K.S. Rediscovering Chemical Gardens: Self-Assembling Cytocompatible Protein-Intercalated Silicate-Phosphate Sponge-Mimetic Tubules. Langmuir ACS J. Surf. Colloids 2016, 32, 8748–8758. [Google Scholar] [CrossRef]
  95. Zhang, Y.; Zhang, Z.; Wang, Y.; Su, Y.; Chen, M. 3D myotube guidance on hierarchically organized anisotropic and conductive fibers for skeletal muscle tissue engineering. Mater. Sci. Eng. C Mater. Biol. Appl. 2020, 116, 111070. [Google Scholar] [CrossRef]
  96. Li, D.; Tao, L.; Shen, Y.; Sun, B.; Xie, X.; Ke, Q.; Mo, X.; Deng, B. Fabrication of Multilayered Nanofiber Scaffolds with a Highly Aligned Nanofiber Yarn for Anisotropic Tissue Regeneration. ACS Omega 2020, 5, 24340–24350. [Google Scholar] [CrossRef]
  97. Zhang, Y.; Le Friec, A.; Chen, M. 3D anisotropic conductive fibers electrically stimulated myogenesis. Int. J. Pharm. 2021, 606, 120841. [Google Scholar] [CrossRef]
  98. Rogers, A.J.; Miller, J.M.; Kannappan, R.; Sethu, P. Cardiac Tissue Chips (CTCs) for Modeling Cardiovascular Disease. IEEE Trans. Bio-Med. Eng. 2019, 66, 3436–3443. [Google Scholar] [CrossRef]
  99. Ciofani, G.; Ricotti, L.; Mattoli, V. Preparation, characterization and in vitro testing of poly(lactic-co-glycolic) acid/barium titanate nanoparticle composites for enhanced cellular proliferation. Biomed. Microdevices 2011, 13, 255–266. [Google Scholar] [CrossRef]
  100. Onbas, R.; Arslan Yildiz, A. Biopatterning of 3D Cellular Model by Contactless Magnetic Manipulation for Cardiotoxicity Screening. Tissue Eng. Part. A 2023, 30, 367–376. [Google Scholar] [CrossRef]
  101. Di, Y.; Zhao, S.; Fan, H.; Li, W.; Jiang, G.; Wang, Y.; Li, C.; Wang, W.; Wang, J. Mass Production of Rg1-Loaded Small Extracellular Vesicles Using a 3D Bioreactor System for Enhanced Cardioprotective Efficacy of Doxorubicin-Induced Cardiotoxicity. Pharmaceutics 2024, 16, 593. [Google Scholar] [CrossRef] [PubMed]
  102. Watanabe, M.; Yano, T.; Sato, T.; Umetsu, A.; Higashide, M.; Furuhashi, M.; Ohguro, H. mTOR Inhibitors Modulate the Physical Properties of 3D Spheroids Derived from H9c2 Cells. Int. J. Mol. Sci. 2023, 24, 11459. [Google Scholar] [CrossRef] [PubMed]
  103. Layer, P.G. In a century from agitated cells to human organoids. J. Neurosci. Methods 2024, 405, 110083. [Google Scholar] [CrossRef]
  104. Lu, Q.; Yin, H.; Grant, M.P.; Elisseeff, J.H. An In Vitro Model for the Ocular Surface and Tear Film System. Sci. Rep. 2017, 7, 6163. [Google Scholar] [CrossRef]
  105. Gleixner, S.; Zahn, I.; Dietrich, J.; Singh, S.; Drobny, A.; Schneider, Y.; Schwendner, R.; Socher, E.; Blavet, N.; Bräuer, L.; et al. A New Immortalized Human Lacrimal Gland Cell Line. Cells 2024, 13, 622. [Google Scholar] [CrossRef] [PubMed]
  106. Rodboon, T.; Yodmuang, S.; Chaisuparat, R.; Ferreira, J.N. Development of high-throughput lacrimal gland organoid platforms for drug discovery in dry eye disease. SLAS Discov. Adv. Life Sci. R. D 2022, 27, 151–158. [Google Scholar] [CrossRef] [PubMed]
  107. Watanabe, M.; Tsugeno, Y.; Sato, T.; Umetsu, A.; Nishikiori, N.; Furuhashi, M.; Ohguro, H. TGF-β Isoforms Affect the Planar and Subepithelial Fibrogenesis of Human Conjunctival Fibroblasts in Different Manners. Biomedicines 2023, 11, 2005. [Google Scholar] [CrossRef]
  108. Tsugeno, Y.; Furuhashi, M.; Sato, T.; Watanabe, M.; Umetsu, A.; Suzuki, S.; Ida, Y.; Hikage, F.; Ohguro, H. FGF-2 enhances fibrogenetic changes in TGF-β2 treated human conjunctival fibroblasts. Sci. Rep. 2022, 12, 16006. [Google Scholar] [CrossRef]
  109. Oouchi, Y.; Watanabe, M.; Ida, Y.; Ohguro, H.; Hikage, F. Rosiglitasone and ROCK Inhibitors Modulate Fibrogenetic Changes in TGF-β2 Treated Human Conjunctival Fibroblasts (HconF) in Different Manners. Int. J. Mol. Sci. 2021, 22, 7335. [Google Scholar] [CrossRef]
  110. Tsugeno, Y.; Sato, T.; Watanabe, M.; Higashide, M.; Furuhashi, M.; Umetsu, A.; Suzuki, S.; Ida, Y.; Hikage, F.; Ohguro, H. All Trans-Retinoic Acids Facilitate the Remodeling of 2D and 3D Cultured Human Conjunctival Fibroblasts. Bioengineering 2022, 9, 463. [Google Scholar] [CrossRef]
  111. Tsugeno, Y.; Sato, T.; Watanabe, M.; Furuhashi, M.; Ohguro, H. Prostanoid FP and EP2 Receptor Agonists Induce Epithelial and Subepithelial Fibrogenetic Changes in Human Conjunctival Fibroblasts in Different Manners. J. Ocul. Pharmacol. Ther. 2023, 39, 404–414. [Google Scholar] [CrossRef] [PubMed]
  112. Guo, Y.; Liu, Q.; Yang, Y.; Guo, X.; Lian, R.; Li, S.; Wang, C.; Zhang, S.; Chen, J. The effects of ROCK inhibitor Y-27632 on injectable spheroids of bovine corneal endothelial cells. Cell. Reprogram. 2015, 17, 77–87. [Google Scholar] [CrossRef] [PubMed]
  113. Li, S.; Han, Y.; Lei, H.; Zeng, Y.; Cui, Z.; Zeng, Q.; Zhu, D.; Lian, R.; Zhang, J.; Chen, Z.; et al. In vitro biomimetic platforms featuring a perfusion system and 3D spheroid culture promote the construction of tissue-engineered corneal endothelial layers. Sci. Rep. 2017, 7, 777. [Google Scholar] [CrossRef]
  114. Ida, Y.; Umetsu, A.; Furuhashi, M.; Watanabe, M.; Hikage, F.; Ohguro, H. The EP2 agonist, omidenepag, alters the physical stiffness of 3D spheroids prepared from human corneal stroma fibroblasts differently depending on the osmotic pressure. Faseb J. 2022, 36, e22067. [Google Scholar] [CrossRef]
  115. Ida, Y.; Umetsu, A.; Furuhashi, M.; Watanabe, M.; Tsugeno, Y.; Suzuki, S.; Hikage, F.; Ohguro, H. ROCK 1 and 2 affect the spatial architecture of 3D spheroids derived from human corneal stromal fibroblasts in different manners. Sci. Rep. 2022, 12, 7419. [Google Scholar] [CrossRef] [PubMed]
  116. Katayama, H.; Furuhashi, M.; Umetsu, A.; Hikage, F.; Watanabe, M.; Ohguro, H.; Ida, Y. Modulation of the Physical Properties of 3D Spheroids Derived from Human Scleral Stroma Fibroblasts (HSSFs) with Different Axial Lengths Obtained from Surgical Patients. Curr. Issues Mol. Biol. 2021, 43, 1715–1725. [Google Scholar] [CrossRef] [PubMed]
  117. Wang, E.; Wang, D.; Geng, A.; Seo, R.; Gong, X. Growth of hollow cell spheroids in microbead templated chambers. Biomaterials 2017, 143, 57–64. [Google Scholar] [CrossRef]
  118. Tirendi, S.; Saccà, S.C.; Vernazza, S.; Traverso, C.; Bassi, A.M.; Izzotti, A. A 3D Model of Human Trabecular Meshwork for the Research Study of Glaucoma. Front. Neurol. 2020, 11, 591776. [Google Scholar] [CrossRef]
  119. Buffault, J.; Brignole-Baudouin, F.; Labbé, A.; Baudouin, C. An Overview of Current Glaucomatous Trabecular Meshwork Models. Curr. Eye Res. 2023, 48, 1089–1099. [Google Scholar] [CrossRef]
  120. Torrejon, K.Y.; Papke, E.L.; Halman, J.R.; Bergkvist, M.; Danias, J.; Sharfstein, S.T.; Xie, Y. TGFβ2-induced outflow alterations in a bioengineered trabecular meshwork are offset by a rho-associated kinase inhibitor. Sci. Rep. 2016, 6, 38319. [Google Scholar] [CrossRef]
  121. Adhikari, B.; Osmond, M.J.; Pantcheva, M.B.; Krebs, M.D. Glycosaminoglycans Influence Extracellular Matrix of Human Trabecular Meshwork Cells Cultured on 3D Scaffolds. ACS Biomater. Sci. Eng. 2022, 8, 5221–5232. [Google Scholar] [CrossRef] [PubMed]
  122. Adhikari, B.; Stinson, B.S.; Osmond, M.J.; Pantcheva, M.B.; Krebs, M.D. Photoinduced Gelatin-Methacrylate Scaffolds to Examine the Impact of Extracellular Environment on Trabecular Meshwork Cells. Ind. Eng. Chem. Res. 2021, 60, 17417–17428. [Google Scholar] [CrossRef] [PubMed]
  123. Vernazza, S.; Tirendi, S.; Scarfì, S.; Passalacqua, M.; Oddone, F.; Traverso, C.E.; Rizzato, I.; Bassi, A.M.; Saccà, S.C. 2D- and 3D-cultures of human trabecular meshwork cells: A preliminary assessment of an in vitro model for glaucoma study. PLoS ONE 2019, 14, e0221942. [Google Scholar] [CrossRef]
  124. Kumon, M.; Fuwa, M.; Shimazaki, A.; Odani-Kawabata, N.; Iwamura, R.; Yoneda, K.; Kato, M. Downregulation of COL12A1 and COL13A1 by a selective EP2 receptor agonist, omidenepag, in human trabecular meshwork cells. PLoS ONE 2023, 18, e0280331. [Google Scholar] [CrossRef]
  125. Watanabe, M.; Ida, Y.; Furuhashi, M.; Tsugeno, Y.; Ohguro, H.; Hikage, F. Screening of the Drug-Induced Effects of Prostaglandin EP2 and FP Agonists on 3D Cultures of Dexamethasone-Treated Human Trabecular Meshwork Cells. Biomedicines 2021, 9, 930. [Google Scholar] [CrossRef] [PubMed]
  126. Suzuki, S.; Furuhashi, M.; Tsugeno, Y.; Umetsu, A.; Ida, Y.; Hikage, F.; Ohguro, H.; Watanabe, M. Comparison of the Drug-Induced Efficacies between Omidenepag Isopropyl, an EP2 Agonist and PGF2α toward TGF-β2-Modulated Human Trabecular Meshwork (HTM) Cells. J. Clin. Med. 2022, 11, 1652. [Google Scholar] [CrossRef]
  127. Kalouche, G.; Beguier, F.; Bakria, M.; Melik-Parsadaniantz, S.; Leriche, C.; Debeir, T.; Rostène, W.; Baudouin, C.; Vigé, X. Activation of Prostaglandin FP and EP2 Receptors Differently Modulates Myofibroblast Transition in a Model of Adult Primary Human Trabecular Meshwork Cells. Investig. Ophthalmol. Vis. Sci. 2016, 57, 1816–1825. [Google Scholar] [CrossRef]
  128. Watanabe, M.; Furuhashi, M.; Tsugeno, Y.; Ida, Y.; Hikage, F.; Ohguro, H. Autotaxin May Have Lysophosphatidic Acid-Unrelated Effects on Three-Dimension (3D) Cultured Human Trabecular Meshwork (HTM) Cells. Int. J. Mol. Sci. 2021, 22, 12039. [Google Scholar] [CrossRef]
  129. Bouchemi, M.; Roubeix, C.; Kessal, K.; Riancho, L.; Raveu, A.L.; Soualmia, H.; Baudouin, C.; Brignole-Baudouin, F. Effect of benzalkonium chloride on trabecular meshwork cells in a new in vitro 3D trabecular meshwork model for glaucoma. Toxicol. Vitr. Int. J. Publ. Assoc. BIBRA 2017, 41, 21–29. [Google Scholar] [CrossRef]
  130. Ota, C.; Ida, Y.; Ohguro, H.; Hikage, F. ROCK inhibitors beneficially alter the spatial configuration of TGFβ2-treated 3D organoids from a human trabecular meshwork (HTM). Sci. Rep. 2020, 10, 20292. [Google Scholar] [CrossRef]
  131. Watanabe, M.; Ida, Y.; Ohguro, H.; Ota, C.; Hikage, F. Diverse effects of pan-ROCK and ROCK2 inhibitors on 2 D and 3D cultured human trabecular meshwork (HTM) cells treated with TGFβ2. Sci. Rep. 2021, 11, 15286. [Google Scholar] [CrossRef]
  132. Buffault, J.; Brignole-Baudouin, F.; Reboussin, É.; Kessal, K.; Labbé, A.; Mélik Parsadaniantz, S.; Baudouin, C. The Dual Effect of Rho-Kinase Inhibition on Trabecular Meshwork Cells Cytoskeleton and Extracellular Matrix in an In Vitro Model of Glaucoma. J. Clin. Med. 2022, 11, 1001. [Google Scholar] [CrossRef]
  133. Watanabe, M.; Ida, Y.; Furuhashi, M.; Tsugeno, Y.; Hikage, F.; Ohguro, H. Pan-ROCK and ROCK2 Inhibitors Affect Dexamethasone-Treated 2D- and 3D-Cultured Human Trabecular Meshwork (HTM) Cells in Opposite Manners. Molecules 2021, 26, 6382. [Google Scholar] [CrossRef]
  134. Watanabe, M.; Sato, T.; Tsugeno, Y.; Higashide, M.; Furuhashi, M.; Umetsu, A.; Suzuki, S.; Ida, Y.; Hikage, F.; Ohguro, H. An α2-Adrenergic Agonist, Brimonidine, Beneficially Affects the TGF-β2-Treated Cellular Properties in an In Vitro Culture Model. Bioengineering 2022, 9, 310. [Google Scholar] [CrossRef]
  135. Zhang, Y.; Tseng, S.C.G.; Zhu, Y.T. Suppression of TGF-β1 signaling by Matrigel via FAK signaling in cultured human trabecular meshwork cells. Sci. Rep. 2021, 11, 7319. [Google Scholar] [CrossRef]
  136. Han, H.; Kampik, D.; Grehn, F.; Schlunck, G. TGF-β2-induced invadosomes in human trabecular meshwork cells. PLoS ONE 2013, 8, e70595. [Google Scholar] [CrossRef]
  137. Watanabe, M.; Sato, T.; Tsugeno, Y.; Higashide, M.; Furuhashi, M.; Ohguro, H. TGF-β-3 Induces Different Effects from TGF-β-1 and -2 on Cellular Metabolism and the Spatial Properties of the Human Trabecular Meshwork Cells. Int. J. Mol. Sci. 2023, 24, 4181. [Google Scholar] [CrossRef]
  138. Zhou, M.; Geathers, J.S.; Grillo, S.L.; Weber, S.R.; Wang, W.; Zhao, Y.; Sundstrom, J.M. Role of Epithelial-Mesenchymal Transition in Retinal Pigment Epithelium Dysfunction. Front. Cell Dev. Biol. 2020, 8, 501. [Google Scholar] [CrossRef]
  139. Raimondi, R.; Zollet, P.; De Rosa, F.P.; Tsoutsanis, P.; Stravalaci, M.; Paulis, M.; Inforzato, A.; Romano, M.R. Where Are We with RPE Replacement Therapy? A Translational Review from the Ophthalmologist Perspective. Int. J. Mol. Sci. 2022, 23, 682. [Google Scholar] [CrossRef]
  140. Blenkinsop, T.A.; Saini, J.S.; Maminishkis, A.; Bharti, K.; Wan, Q.; Banzon, T.; Lotfi, M.; Davis, J.; Singh, D.; Rizzolo, L.J.; et al. Human Adult Retinal Pigment Epithelial Stem Cell-Derived RPE Monolayers Exhibit Key Physiological Characteristics of Native Tissue. Investig. Ophthalmol. Vis. Sci. 2015, 56, 7085–7099. [Google Scholar] [CrossRef]
  141. Mandai, M.; Watanabe, A.; Kurimoto, Y.; Hirami, Y.; Morinaga, C.; Daimon, T.; Fujihara, M.; Akimaru, H.; Sakai, N.; Shibata, Y.; et al. Autologous Induced Stem-Cell-Derived Retinal Cells for Macular Degeneration. N. Engl. J. Med. 2017, 376, 1038–1046. [Google Scholar] [CrossRef]
  142. Kashani, A.H.; Lebkowski, J.S.; Rahhal, F.M.; Avery, R.L.; Salehi-Had, H.; Dang, W.; Lin, C.M.; Mitra, D.; Zhu, D.; Thomas, B.B.; et al. A bioengineered retinal pigment epithelial monolayer for advanced, dry age-related macular degeneration. Sci. Transl. Med. 2018, 10, eaao4097. [Google Scholar] [CrossRef] [PubMed]
  143. da Cruz, L.; Fynes, K.; Georgiadis, O.; Kerby, J.; Luo, Y.H.; Ahmado, A.; Vernon, A.; Daniels, J.T.; Nommiste, B.; Hasan, S.M.; et al. Phase 1 clinical study of an embryonic stem cell-derived retinal pigment epithelium patch in age-related macular degeneration. Nat. Biotechnol. 2018, 36, 328–337. [Google Scholar] [CrossRef]
  144. Hunt, N.C.; Hallam, D.; Karimi, A.; Mellough, C.B.; Chen, J.; Steel, D.H.W.; Lako, M. 3D culture of human pluripotent stem cells in RGD-alginate hydrogel improves retinal tissue development. Acta Biomater. 2017, 49, 329–343. [Google Scholar] [CrossRef] [PubMed]
  145. Zhu, D.; Xie, M.; Gademann, F.; Cao, J.; Wang, P.; Guo, Y.; Zhang, L.; Su, T.; Zhang, J.; Chen, J. Protective effects of human iPS-derived retinal pigmented epithelial cells on retinal degenerative disease. Stem Cell Res. Ther. 2020, 11, 98. [Google Scholar] [CrossRef]
  146. Kuwahara, A.; Nakano, T.; Eiraku, M. Generation of a Three-Dimensional Retinal Tissue from Self-Organizing Human ESC Culture. Methods Mol. Biol. 2017, 1597, 17–29. [Google Scholar]
  147. Shokoohmand, A.; Jeon, J.E.; Theodoropoulos, C.; Baldwin, J.G.; Hutmacher, D.W.; Feigl, B. A Novel 3D Cultured Model for Studying Early Changes in Age-Related Macular Degeneration. Macromol. Biosci. 2017, 17, 1700221. [Google Scholar] [CrossRef]
  148. Tian, Y.; Zonca, M.R., Jr.; Imbrogno, J.; Unser, A.M.; Sfakis, L.; Temple, S.; Belfort, G.; Xie, Y. Polarized, Cobblestone, Human Retinal Pigment Epithelial Cell Maturation on a Synthetic PEG Matrix. ACS Biomater. Sci. Eng. 2017, 3, 890–902. [Google Scholar] [CrossRef]
  149. Isla-Magrané, H.; Veiga, A.; García-Arumí, J.; Duarri, A. Multiocular organoids from human induced pluripotent stem cells displayed retinal, corneal, and retinal pigment epithelium lineages. Stem Cell Res. Ther. 2021, 12, 581. [Google Scholar] [CrossRef]
  150. Stahl, A.; Paschek, L.; Martin, G.; Feltgen, N.; Hansen, L.L.; Agostini, H.T. Combinatory inhibition of VEGF and FGF2 is superior to solitary VEGF inhibition in an in vitro model of RPE-induced angiogenesis. Graefe’s Arch. Clin. Exp. Ophthalmol. Albrecht Von. Graefes Arch. Fur Klin. Und Exp. Ophthalmol. 2009, 247, 767–773. [Google Scholar] [CrossRef]
  151. Suzuki, S.; Sato, T.; Watanabe, M.; Higashide, M.; Tsugeno, Y.; Umetsu, A.; Furuhashi, M.; Ida, Y.; Hikage, F.; Ohguro, H. Hypoxia Differently Affects TGF-β2-Induced Epithelial Mesenchymal Transitions in the 2D and 3D Culture of the Human Retinal Pigment Epithelium Cells. Int. J. Mol. Sci. 2022, 23, 5473. [Google Scholar] [CrossRef]
  152. Ida, Y.; Sato, T.; Watanabe, M.; Umetsu, A.; Tsugeno, Y.; Furuhashi, M.; Hikage, F.; Ohguro, H. The Selective α1 Antagonist Tamsulosin Alters ECM Distributions and Cellular Metabolic Functions of ARPE 19 Cells in a Concentration-Dependent Manner. Bioengineering 2022, 9, 556. [Google Scholar] [CrossRef]
  153. Veernala, I.; Jaffet, J.; Fried, J.; Mertsch, S.; Schrader, S.; Basu, S.; Vemuganti, G.K.; Singh, V. Lacrimal gland regeneration: The unmet challenges and promise for dry eye therapy. Ocul. Surf. 2022, 25, 129–141. [Google Scholar] [CrossRef]
  154. Bannier-Hélaouët, M.; Post, Y.; Korving, J.; Trani Bustos, M.; Gehart, H.; Begthel, H.; Bar-Ephraim, Y.E.; van der Vaart, J.; Kalmann, R.; Imhoff, S.M.; et al. Exploring the human lacrimal gland using organoids and single-cell sequencing. Cell Stem Cell 2021, 28, 1221–1232.e7. [Google Scholar] [CrossRef]
  155. Hayashi, R.; Okubo, T.; Kudo, Y.; Ishikawa, Y.; Imaizumi, T.; Suzuki, K.; Shibata, S.; Katayama, T.; Park, S.J.; Young, R.D.; et al. Generation of 3D lacrimal gland organoids from human pluripotent stem cells. Nature 2022, 605, 126–131. [Google Scholar] [CrossRef]
  156. Chen, H.; Huang, P.; Zhang, Y. Three-Dimensional, Serum-Free Culture System for Lacrimal Gland Stem Cells. J. Vis. Exp. 2022, 184, e63585. [Google Scholar]
  157. Jaffet, J.; Mohanty, A.; Veernala, I.; Singh, S.; Ali, M.J.; Basu, S.; Vemuganti, G.K.; Singh, V. Human Lacrimal Gland Derived Mesenchymal Stem Cells—Isolation, Propagation, and Characterization. Invest. Ophthalmol. Vis. Sci. 2023, 64, 12. [Google Scholar] [CrossRef]
  158. Xiao, S.; Zhang, Y. Establishment of long-term serum-free culture for lacrimal gland stem cells aiming at lacrimal gland repair. Stem Cell Res. Ther. 2020, 11, 20. [Google Scholar] [CrossRef]
  159. Jeong, S.Y.; Choi, W.H.; Jeon, S.G.; Lee, S.; Park, J.M.; Park, M.; Lee, H.; Lew, H.; Yoo, J. Establishment of functional epithelial organoids from human lacrimal glands. Stem Cell Res. Ther. 2021, 12, 247. [Google Scholar] [CrossRef]
  160. Massie, I.; Spaniol, K.; Barbian, A.; Geerling, G.; Metzger, M.; Schrader, S. Development of lacrimal gland spheroids for lacrimal gland tissue regeneration. J. Tissue Eng. Regen. Med. 2018, 12, e2001–e2009. [Google Scholar] [CrossRef]
  161. Li, H.; Fitchett, C.; Kozdon, K.; Jayaram, H.; Rose, G.E.; Bailly, M.; Ezra, D.G. Independent adipogenic and contractile properties of fibroblasts in Graves’ orbitopathy: An in vitro model for the evaluation of treatments. PLoS ONE 2014, 9, e95586. [Google Scholar] [CrossRef]
  162. Hikage, F.; Ichioka, H.; Watanabe, M.; Umetsu, A.; Ohguro, H.; Ida, Y. Addition of ROCK inhibitors to prostaglandin derivative (PG) synergistically affects adipogenesis of the 3D spheroids of human orbital fibroblasts (HOFs). Hum. Cell 2022, 35, 125–132. [Google Scholar] [CrossRef] [PubMed]
  163. Ichioka, H.; Ida, Y.; Watanabe, M.; Ohguro, H.; Hikage, F. Prostaglandin F2α and EP2 agonists, and a ROCK inhibitor modulate the formation of 3D organoids of Grave’s orbitopathy related human orbital fibroblasts. Exp. Eye Res. 2021, 205, 108489. [Google Scholar] [CrossRef] [PubMed]
  164. Ida, Y.; Ichioka, H.; Furuhashi, M.; Hikage, F.; Watanabe, M.; Umetsu, A.; Ohguro, H. Reactivities of a Prostanoid EP2 Agonist, Omidenepag, Are Useful for Distinguishing between 3D Spheroids of Human Orbital Fibroblasts without or with Graves’ Orbitopathy. Cells 2021, 10, 3196. [Google Scholar] [CrossRef]
  165. Hikage, F.; Ichioka, H.; Watanabe, M.; Umetsu, A.; Ohguro, H.; Ida, Y. ROCK inhibitors modulate the physical properties and adipogenesis of 3D spheroids of human orbital fibroblasts in different manners. FASEB Bioadv. 2021, 3, 866–872. [Google Scholar] [CrossRef] [PubMed]
  166. Itoh, K.; Ida, Y.; Ohguro, H.; Hikage, F. Prostaglandin F2α agonists induced enhancement in collagen1 expression is involved in the pathogenesis of the deepening of upper eyelid sulcus. Sci. Rep. 2021, 11, 9002. [Google Scholar] [CrossRef] [PubMed]
  167. Hikage, F.; Watanabe, M.; Sato, T.; Umetsu, A.; Tsugeno, Y.; Furuhashi, M.; Ohguro, H. Simultaneous Effects of a Selective EP2 Agonist, Omidenepag, and a Rho-Associated Coiled-Coil Containing Protein Kinase Inhibitor, Ripasudil, on Human Orbital Fibroblasts. J. Ocul. Pharmacol. Ther. 2023, 39, 439–448. [Google Scholar] [CrossRef]
  168. Hikage, F.; Ida, Y.; Ouchi, Y.; Watanabe, M.; Ohguro, H. Omidenepag, a Selective, Prostanoid EP2 Agonist, Does Not Suppress Adipogenesis in 3D Organoids of Human Orbital Fibroblasts. Transl. Vis. Sci. Technol. 2021, 10, 6. [Google Scholar] [CrossRef]
  169. Krieger, C.C.; Perry, J.D.; Morgan, S.J.; Kahaly, G.J.; Gershengorn, M.C. TSH/IGF-1 Receptor Cross-Talk Rapidly Activates Extracellular Signal-Regulated Kinases in Multiple Cell Types. Endocrinology 2017, 158, 3676–3683. [Google Scholar] [CrossRef]
  170. Dik, W.A.; Virakul, S.; van Steensel, L. Current perspectives on the role of orbital fibroblasts in the pathogenesis of Graves’ ophthalmopathy. Exp. Eye Res. 2016, 142, 83–91. [Google Scholar] [CrossRef]
  171. Turcu, A.F.; Kumar, S.; Neumann, S.; Coenen, M.; Iyer, S.; Chiriboga, P.; Gershengorn, M.C.; Bahn, R.S. A small molecule antagonist inhibits thyrotropin receptor antibody-induced orbital fibroblast functions involved in the pathogenesis of Graves ophthalmopathy. J. Clin. Endocrinol. Metab. 2013, 98, 2153–2159. [Google Scholar] [CrossRef] [PubMed]
  172. Smith, T.J. TSH-receptor-expressing fibrocytes and thyroid-associated ophthalmopathy. Nat. Rev. Endocrinol. 2015, 11, 171–181. [Google Scholar] [CrossRef] [PubMed]
  173. Ida, Y.; Sato, T.; Umetsu, A.; Watanabe, M.; Furuhashi, M.; Hikage, F.; Ohguro, H. Addition of ROCK Inhibitors Alleviates Prostaglandin-Induced Inhibition of Adipogenesis in 3T3L-1 Spheroids. Bioengineering 2022, 9, 702. [Google Scholar] [CrossRef]
  174. Ida, Y.; Furuhashi, M.; Watanabe, M.; Umetsu, A.; Hikage, F.; Ohguro, H. Prostaglandin F2 and EP2 Agonists Exert Different Effects on 3D 3T3-L1 Spheroids during Their Culture Phase. Biomedicines 2021, 9, 1821. [Google Scholar] [CrossRef]
  175. Umetsu, A.; Ida, Y.; Sato, T.; Watanabe, M.; Tsugeno, Y.; Furuhashi, M.; Hikage, F.; Ohguro, H. Brimonidine Modulates the ROCK1 Signaling Effects on Adipogenic Differentiation in 2D and 3D 3T3-L1 Cells. Biomedicines 2022, 9, 327. [Google Scholar] [CrossRef]
  176. 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. [Google Scholar] [CrossRef]
  177. Nath, S.; Devi, G.R. Three-dimensional culture systems in cancer research: Focus on tumor spheroid model. Pharmacol. Ther. 2016, 163, 94–108. [Google Scholar] [CrossRef]
  178. Butelmann, T.; Gu, Y.; Li, A.; Tribukait-Riemenschneider, F.; Hoffmann, J.; Molazem, A.; Jaeger, E.; Pellegrini, D.; Forget, A.; Shastri, V.P. 3D Printed Solutions for Spheroid Engineering and Cancer Research. Int. J. Mol. Sci. 2022, 23, 8188. [Google Scholar] [CrossRef]
  179. Zhao, L.; Xiu, J.; Liu, Y.; Zhang, T.; Pan, W.; Zheng, X.; Zhang, X. A 3D Printed Hanging Drop Dripper for Tumor Spheroids Analysis Without Recovery. Sci. Rep. 2019, 9, 19717. [Google Scholar] [CrossRef] [PubMed]
  180. Amaral, R.L.F.; Miranda, M.; Marcato, P.D.; Swiech, K. Comparative Analysis of 3D Bladder Tumor Spheroids Obtained by Forced Floating and Hanging Drop Methods for Drug Screening. Front. Physiol. 2017, 8, 605. [Google Scholar] [CrossRef]
  181. Shi, H.; Rath, E.M.; Lin, R.C.Y.; Sarun, K.H.; Clarke, C.J.; McCaughan, B.C.; Ke, H.; Linton, A.; Lee, K.; Klebe, S.; et al. 3-Dimensional mesothelioma spheroids provide closer to natural pathophysiological tumor microenvironment for drug response studies. Front. Oncol. 2022, 12, 973576. [Google Scholar] [CrossRef] [PubMed]
  182. Ganguli, A.; Mostafa, A.; Saavedra, C.; Kim, Y.; Le, P.; Faramarzi, V.; Feathers, R.W.; Berger, J.; Ramos-Cruz, K.P.; Adeniba, O.; et al. Three-dimensional microscale hanging drop arrays with geometric control for drug screening and live tissue imaging. Sci. Adv. 2021, 7, eabc1323. [Google Scholar] [CrossRef] [PubMed]
  183. Lo, J.A.; Fisher, D.E. The melanoma revolution: From UV carcinogenesis to a new era in therapeutics. Science 2014, 346, 945–949. [Google Scholar] [CrossRef] [PubMed]
  184. Luke, J.J.; Flaherty, K.T.; Ribas, A.; Long, G.V. Targeted agents and immunotherapies: Optimizing outcomes in melanoma. Nat. Rev. Clin. Oncol. 2017, 14, 463–482. [Google Scholar] [CrossRef]
  185. Genomic Classification of Cutaneous Melanoma. Cell 2015, 161, 1681–1696. [CrossRef]
  186. Davies, H.; Bignell, G.R.; Cox, C.; Stephens, P.; Edkins, S.; Clegg, S.; Teague, J.; Woffendin, H.; Garnett, M.J.; Bottomley, W.; et al. Mutations of the BRAF gene in human cancer. Nature 2002, 417, 949–954. [Google Scholar] [CrossRef] [PubMed]
  187. Carvajal, R.D.; Antonescu, C.R.; Wolchok, J.D.; Chapman, P.B.; Roman, R.A.; Teitcher, J.; Panageas, K.S.; Busam, K.J.; Chmielowski, B.; Lutzky, J.; et al. KIT as a therapeutic target in metastatic melanoma. Jama 2011, 305, 2327–2334. [Google Scholar] [CrossRef]
  188. Huh, D.; Hamilton, G.A.; Ingber, D.E. From 3D cell culture to organs-on-chips. Trends Cell Biol. 2011, 21, 745–754. [Google Scholar] [CrossRef]
  189. 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]
  190. Beaumont, K.A.; Mohana-Kumaran, N.; Haass, N.K. Modeling Melanoma In Vitro and In Vivo. Healthcare 2013, 2, 27–46. [Google Scholar] [CrossRef]
  191. D’Mello, S.A.; Finlay, G.J.; Baguley, B.C.; Askarian-Amiri, M.E. Signaling Pathways in Melanogenesis. Int. J. Mol. Sci. 2016, 17, 1144. [Google Scholar] [CrossRef] [PubMed]
  192. Yeh, I.; Busam, K.J. Spitz melanocytic tumours—A review. Histopathology 2022, 80, 122–134. [Google Scholar] [CrossRef] [PubMed]
  193. Ahmed, F.; Haass, N.K. Microenvironment-Driven Dynamic Heterogeneity and Phenotypic Plasticity as a Mechanism of Melanoma Therapy Resistance. Front. Oncol. 2018, 8, 173. [Google Scholar] [CrossRef]
  194. Jemal, A.; Bray, F.; Center, M.M.; Ferlay, J.; Ward, E.; Forman, D. Global cancer statistics. CA A Cancer J. Clin. 2011, 61, 69–90. [Google Scholar] [CrossRef]
  195. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA A Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [PubMed]
  196. McCullough, M.J.; Prasad, G.; Farah, C.S. Oral mucosal malignancy and potentially malignant lesions: An update on the epidemiology, risk factors, diagnosis and management. Aust. Dent. J. 2010, 55 (Suppl. S1), 61–65. [Google Scholar] [CrossRef]
  197. Farah, C.S.; Woo, S.B.; Zain, R.B.; Sklavounou, A.; McCullough, M.J.; Lingen, M. Oral cancer and oral potentially malignant disorders. Int. J. Dent. 2014, 2014, 853479. [Google Scholar] [CrossRef]
  198. Abrahão, R.; Anantharaman, D.; Gaborieau, V.; Abedi-Ardekani, B.; Lagiou, P.; Lagiou, A.; Ahrens, W.; Holcatova, I.; Betka, J.; Merletti, F.; et al. The influence of smoking, age and stage at diagnosis on the survival after larynx, hypopharynx and oral cavity cancers in Europe: The ARCAGE study. Int. J. Cancer 2018, 143, 32–44. [Google Scholar] [CrossRef]
  199. Chitturi Suryaprakash, R.T.; Kujan, O.; Shearston, K.; Farah, C.S. Three-Dimensional Cell Culture Models to Investigate Oral Carcinogenesis: A Scoping Review. Int. J. Mol. Sci. 2020, 21, 9520. [Google Scholar] [CrossRef]
  200. Jensen, C.; Teng, Y. Is It Time to Start Transitioning From 2D to 3D Cell Culture? Front. Mol. Biosci. 2020, 7, 33. [Google Scholar] [CrossRef]
  201. Miki, Y.; Ono, K.; Hata, S.; Suzuki, T.; Kumamoto, H.; Sasano, H. The advantages of co-culture over mono cell culture in simulating in vivo environment. J. Steroid Biochem. Mol. Biol. 2012, 131, 68–75. [Google Scholar] [CrossRef] [PubMed]
  202. Shang, M.; Soon, R.H.; Lim, C.T.; Khoo, B.L.; Han, J. Microfluidic modelling of the tumor microenvironment for anti-cancer drug development. Lab. A Chip 2019, 19, 369–386. [Google Scholar] [CrossRef] [PubMed]
  203. Yamada, K.M.; Cukierman, E. Modeling tissue morphogenesis and cancer in 3D. Cell 2007, 130, 601–610. [Google Scholar] [CrossRef]
  204. Zanoni, M.; Pignatta, S.; Arienti, C.; Bonafè, M.; Tesei, A. Anticancer drug discovery using multicellular tumor spheroid models. Expert Opin. Drug Discov. 2019, 14, 289–301. [Google Scholar] [CrossRef]
  205. Biffi, G.; Tuveson, D.A. Diversity and Biology of Cancer-Associated Fibroblasts. Physiol. Rev. 2021, 101, 147–176. [Google Scholar] [CrossRef] [PubMed]
  206. Chen, Y.; McAndrews, K.M.; Kalluri, R. Clinical and therapeutic relevance of cancer-associated fibroblasts. Nat. Rev. Clin. Oncol. 2021, 18, 792–804. [Google Scholar] [CrossRef]
  207. Kalluri, R. The biology and function of fibroblasts in cancer. Nat. Rev. Cancer 2016, 16, 582–598. [Google Scholar] [CrossRef]
  208. Choi, S.Y.; Kang, S.H.; Oh, S.Y.; Lee, K.Y.; Lee, H.J.; Gum, S.; Kwon, T.G.; Kim, J.W.; Lee, S.T.; Hong, Y.J.; et al. Differential Angiogenic Potential of 3-Dimension Spheroid of HNSCC Cells in Mouse Xenograft. Int. J. Mol. Sci. 2021, 22, 8245. [Google Scholar] [CrossRef]
  209. Di Modugno, F.; Colosi, C.; Trono, P.; Antonacci, G.; Ruocco, G.; Nisticò, P. 3D models in the new era of immune oncology: Focus on T cells, CAF and ECM. J. Exp. Clin. Cancer Res. CR 2019, 38, 117. [Google Scholar] [CrossRef]
  210. Kalluri, R.; Zeisberg, M. Fibroblasts in cancer. Nat. Rev. Cancer 2006, 6, 392–401. [Google Scholar] [CrossRef]
  211. Parsonage, G.; Filer, A.D.; Haworth, O.; Nash, G.B.; Rainger, G.E.; Salmon, M.; Buckley, C.D. A stromal address code defined by fibroblasts. Trends Immunol. 2005, 26, 150–156. [Google Scholar] [CrossRef] [PubMed]
  212. Tomasek, J.J.; Gabbiani, G.; Hinz, B.; Chaponnier, C.; Brown, R.A. Myofibroblasts and mechano-regulation of connective tissue remodelling. Nat. Rev. Mol. Cell Biol. 2002, 3, 349–363. [Google Scholar] [CrossRef] [PubMed]
  213. Watt, F.M.; Fujiwara, H. Cell-extracellular matrix interactions in normal and diseased skin. Cold Spring Harb. Perspect. Biol. 2011, 3, a005124. [Google Scholar] [CrossRef]
  214. Lynch, M.D.; Watt, F.M. Fibroblast heterogeneity: Implications for human disease. J. Clin. Investig. 2018, 128, 26–35. [Google Scholar] [CrossRef]
  215. Sahai, E.; Astsaturov, I.; Cukierman, E.; DeNardo, D.G.; Egeblad, M.; Evans, R.M.; Fearon, D.; Greten, F.R.; Hingorani, S.R.; Hunter, T.; et al. A framework for advancing our understanding of cancer-associated fibroblasts. Nat. Rev. Cancer 2020, 20, 174–186. [Google Scholar] [CrossRef]
  216. Lavie, D.; Ben-Shmuel, A.; Erez, N.; Scherz-Shouval, R. Cancer-associated fibroblasts in the single-cell era. Nat. Cancer 2022, 3, 793–807. [Google Scholar] [CrossRef] [PubMed]
  217. Yamamoto, Y.; Kasashima, H.; Fukui, Y.; Tsujio, G.; Yashiro, M.; Maeda, K. The heterogeneity of cancer-associated fibroblast subpopulations: Their origins, biomarkers, and roles in the tumor microenvironment. Cancer Sci. 2023, 114, 16–24. [Google Scholar] [CrossRef]
  218. Kanzaki, R.; Pietras, K. Heterogeneity of cancer-associated fibroblasts: Opportunities for precision medicine. Cancer Sci. 2020, 111, 2708–2717. [Google Scholar] [CrossRef]
  219. Kehrberg, R.J.; Bhyravbhatla, N.; Batra, S.K.; Kumar, S. Epigenetic regulation of cancer-associated fibroblast heterogeneity. Biochim. Biophys. Acta Rev. Cancer 2023, 1878, 188901. [Google Scholar] [CrossRef]
  220. Nurmik, M.; Ullmann, P.; Rodriguez, F.; Haan, S.; Letellier, E. In search of definitions: Cancer-associated fibroblasts and their markers. Int. J. Cancer 2020, 146, 895–905. [Google Scholar] [CrossRef]
Figure 2. Measurement of stiffness of spheroids. A living spheroid (3D sph) was initially held between a pedestal and a pressing plate ((A), D: diameter) and was gradually pressed by a press bar until the initial diameter was halved (D/2) over a period of 20 s (B).
Figure 2. Measurement of stiffness of spheroids. A living spheroid (3D sph) was initially held between a pedestal and a pressing plate ((A), D: diameter) and was gradually pressed by a press bar until the initial diameter was halved (D/2) over a period of 20 s (B).
Cells 13 01549 g002
Figure 3. Representative maturation process of H9c2 spheroids. Representative phase contrast (PC) images of a H9c2 spheroid in a well of a hanging drop culture plate during different culture periods at 0.5 h (A), 5 h (B), 7.5 h (C), 14.5 h (D), and 24 h (E). Panel F shows a representative PC image of a H9c2 spheroid taken out from the culture plate after 24 h culture period (F). Scale bar: 100 μm.
Figure 3. Representative maturation process of H9c2 spheroids. Representative phase contrast (PC) images of a H9c2 spheroid in a well of a hanging drop culture plate during different culture periods at 0.5 h (A), 5 h (B), 7.5 h (C), 14.5 h (D), and 24 h (E). Panel F shows a representative PC image of a H9c2 spheroid taken out from the culture plate after 24 h culture period (F). Scale bar: 100 μm.
Cells 13 01549 g003
Figure 4. Trypsin digestion of 2D H9c2 cells and 3D H9c2 spheroid. Representative phase contrast images of 2D and 3D cultured H9c2 cells treated by 0.05% trypsin for different time periods are shown. Scale bar: 100 μm.
Figure 4. Trypsin digestion of 2D H9c2 cells and 3D H9c2 spheroid. Representative phase contrast images of 2D and 3D cultured H9c2 cells treated by 0.05% trypsin for different time periods are shown. Scale bar: 100 μm.
Cells 13 01549 g004
Figure 5. Estimation of total number of cells in a single 3D spheroid.
Figure 5. Estimation of total number of cells in a single 3D spheroid.
Cells 13 01549 g005
Figure 7. Representative images of 3D spheroid obtained from various non-cancerous cells. MCF7: human breast cancer cell line, HTB26: breast adenocarcinoma cell line, LHK2: lung adenocarcinoma cell line, SK-mel-24: malignant melanoma cell line, MP-1000: oral squamous cell carcinoma cell line, CAFS1: cancer-associated fibroblast cell line. Scale bar: 100 μm.
Figure 7. Representative images of 3D spheroid obtained from various non-cancerous cells. MCF7: human breast cancer cell line, HTB26: breast adenocarcinoma cell line, LHK2: lung adenocarcinoma cell line, SK-mel-24: malignant melanoma cell line, MP-1000: oral squamous cell carcinoma cell line, CAFS1: cancer-associated fibroblast cell line. Scale bar: 100 μm.
Cells 13 01549 g007
Table 1. Various methods for 3D spheroid generation.
Table 1. Various methods for 3D spheroid generation.
MethodsPlace to Generate 3D SpheroidEffects of Gravity ForceEffects of Buoyant ForceContribution of Additive Factors
scaffold-assisted
hydrogel-assisted cultureon the groundlesslowhydrogels
biofilm-assisted cultureon the groundlesslowbiofilms
particle-assisted cultureon the groundlesslowparticles
magnet particle-assisted cultureabove the groundmoderatestrongmagnetic particles, magnetic force
scaffold non-assisted
static suspension cultureon the groundlesslownone
floating cultureabove the groundmoderatestrongcentrifugal force
hanging drop cultureunder the groundstrongmoderatenone
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ohguro, H.; Watanabe, M.; Sato, T.; Nishikiori, N.; Umetsu, A.; Higashide, M.; Yano, T.; Suzuki, H.; Miyazaki, A.; Takada, K.; et al. Application of Single Cell Type-Derived Spheroids Generated by Using a Hanging Drop Culture Technique in Various In Vitro Disease Models: A Narrow Review. Cells 2024, 13, 1549. https://doi.org/10.3390/cells13181549

AMA Style

Ohguro H, Watanabe M, Sato T, Nishikiori N, Umetsu A, Higashide M, Yano T, Suzuki H, Miyazaki A, Takada K, et al. Application of Single Cell Type-Derived Spheroids Generated by Using a Hanging Drop Culture Technique in Various In Vitro Disease Models: A Narrow Review. Cells. 2024; 13(18):1549. https://doi.org/10.3390/cells13181549

Chicago/Turabian Style

Ohguro, Hiroshi, Megumi Watanabe, Tatsuya Sato, Nami Nishikiori, Araya Umetsu, Megumi Higashide, Toshiyuki Yano, Hiromu Suzuki, Akihiro Miyazaki, Kohichi Takada, and et al. 2024. "Application of Single Cell Type-Derived Spheroids Generated by Using a Hanging Drop Culture Technique in Various In Vitro Disease Models: A Narrow Review" Cells 13, no. 18: 1549. https://doi.org/10.3390/cells13181549

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop