Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Dec 10.
Published in final edited form as: J Control Release. 2012 May 18;164(2):192–204. doi: 10.1016/j.jconrel.2012.04.045

Opportunities and Challenges for use of Tumor Spheroids as Models to Test Drug Delivery and Efficacy

Geeta Mehta 1,2,*, Amy Y Hsiao 1, Marylou Ingram 3, Gary D Luker 4,5, Shuichi Takayama 1,6,7,*
PMCID: PMC3436947  NIHMSID: NIHMS378814  PMID: 22613880

Abstract

Multicellular spheroids are three dimensional in vitro microscale tissue analogs. The current article examines the suitability of spheroids as an in vitro platform for testing drug delivery systems. Spheroids model critical physiologic parameters present in vivo, including complex multicellular architecture, barriers to mass transport, and extracellular matrix deposition. Relative to two-dimensional cultures, spheroids also provide better target cells for drug testing and are appropriate in vitro model for studies of drug penetration. Key challenges associated with creation of uniformly sized spheroids, spheroids with small number of cells and co-culture spheroids are emphasized in the article. Moreover, the assay techniques required for the characterization of drug delivery and efficacy in spheroids and the challenges associated with such studies are discussed. Examples for the use of spheroids in drug delivery and testing are also emphasized. With these challenges and the possible solutions, multicellular spheroids are becoming an increasingly useful in vitro tool for drug screening and delivery to pathological tissues and organs.

Keywords: Spheroids, Drug delivery, Drug screening, High throughput, Tissue engineering, Imaging

Introduction

Drug delivery

Drug delivery is the process of administering an active pharmaceutical ingredient (API) in vivo to achieve a therapeutic effect in the patient [13]. Inside the body, the API reaches the target of interest after crossing many biological barriers such as other organs, the extracellular matrix, cells and intracellular compartments. During this process, the API may become inactivated or trapped at non-target sites as well as produce undesirable effects on non-target organs and tissues. Thus, efficient delivery of API to target tissues is critical to reduce unintended toxicity to other organs, while delivering the drug in a cost-effective manner. The challenge is that this complex physiological process is affected by many factors such as the nature of the API itself, carrier material used, targeting moieties attached if any, and additional external mechanism if any (such as changes in temperature, pH, magnetic field, ultrasound etc.) [2, 4]. Two common approaches to specifically target drug-loaded carrier systems to required pathological sites in the body are 1) active, involving cell-specific delivery of the API based on attachment of specific ligands to the surface of the pharmaceutical carriers to recognize and bind pathological cells (for example: antibody mediated targeting); and 2) passive, mediated by the enhanced permeability and retention (EPR) effect). EPR effect is based on the longevity of the APIs and its carriers in the blood, due to their extravasation and their accumulation in pathological sites with compromised vasculature (such as tumors, which have leaky blood vessels). This delivery method works very well for tumors, inflammation, and infarcts [3].

There are several requirements for the safe and effective delivery of therapeutic agents to target tissues and organs for human use. Before clinical trials and FDA approval, the drugs and delivery mechanism need to be rigorously tested to determine effectiveness, toxicity and safety. The first step in this process begins with in vitro testing of the drugs and their delivery methods. Typically such testing of delivery mechanisms against target cells is undertaken in 2-D, multi-well plate-based cell culture formats [5]. Results from such platforms, however, often are very different from the scenario in the body. From the drug delivery perspective, this is readily understood from the viewpoint of lack of appropriate physiological barriers arranged in a reasonable geometry. Therefore, in vitro models of tissues of interest that are more physiological than such conventional 2-D culture are needed for better prediction of drug effects and delivery mechanisms. This review focuses on one such platform: 3-D multicellular spheroids.

Spheroids as appropriate models for drug delivery

Platforms for 3-D cell culture, such as scaffolds, hydrogels, and microfluidics can provide enhanced models for testing of drug delivery, toxicity, and metabolism compared to conventional 2-D cultures due to their more physiological cell-cell contact geometry, mass transport, and mechanical properties [6]. Spheroids, which are microscale, spherical cell clusters formed by self-assembly, are one of the most common and versatile methods of culturing cells in 3-D [7]. Spheroids with radii of 200 micrometers and larger will have zones of proliferating cells on the outside and quiescent cells on the inside due to nutrient and oxygen transport limitations. Significantly larger spheroids can also harbor necrotic cells at the center as can be observed in some cancers in vivo. Spheroids have been used as models for evaluations of drug sensitivity and resistance, and they are typically more resistant to chemo-and radiotherapies compared to cells cultured as 2-D monolayers [8]. These models also allow analyses of different growth constraints such as oxygen tension and nutrient consumption, radiation effects, and angiogenesis [9].

Spheroids present a more physiological platform for drug delivery testing due to the following reasons (Figure 1):

Figure 1. Spheroids are appropriate models for testing drug delivery systems in vitro.

Figure 1

A) In vivo, concentration of drug, oxygen and nutrients decreases on increasing distance from blood vessels. The effectiveness of drug and carrier system in vivo is a function of drug dose, potency, kinetics, molecular weight, charge, solubility in water and lipids, diffusion, barriers in the microenvironment, binding, metabolism, and sequestration. B) Spheroids are 3-D microscale tissues which exhibit an inherent gradient of nutrients, oxygen and metabolites within themselves, which leads to a central necrotic core region surrounded by quiescent viable cells and an outer layer of actively proliferating cells. Due to similarities between in vivo tissues and spheroids as well as mass transport limitations, spheroids can serve as high throughput screening platforms for drug and carrier effectiveness. Scale bar is 200 μm.

  1. Spheroids model the 3-D architecture of tissues, including multicellular arrangement and extracellular matrix deposition, found in vivo. Such arrangements, which are absent in conventional culture formats, reproduce how drug delivery might occur in vivo.

  2. Being closely packed 3-D structures; spheroids have sizeable cell-cell interactions, including tight junctions, comparable to those in in vivo tissues. These cell-cell contacts and ensuing communication have been found to influence response of cells to drugs [10, 11].

  3. Spheroids have diffusional limits to mass transport of drugs, nutrients and other factors, similar to in vivo tissues. Due to their mimicry of the physiological barriers to drug delivery in vivo, spheroid can serve as an improved assay format for testing efficacy.

  4. Co-culture spheroids are formed with two or more cell types in varying ratios representing intercellular signaling and architecture seen in vivo. These co-culture spheroids can help decipher how multiple cell types found in tissues in vivo might impact drug delivery.

  5. Rare cells such as cancer stem cells or primary stem cells may be incorporated and maintained in spheroids, which can facilitate targeting these specific cells with drugs. It is often difficult to maintain small numbers of such cells in conventional culture formats and to decode how these cells respond to the drug and delivery mechanism.

  6. Spheroids are 3D models of in vivo solid tumors. Larger spheroids develop central necrosis and regions of hypoxia present in many cancers, which is critical for testing anti-cancer therapeutics. Cellular microenvironments such as hypoxia, which have been identified as one cause of drug resistance [1214], can be modeled and created within spheroids for accurate testing of drug efficacy.

Spheroids are advantageous in vitro models for recreating some of the mass transport limitations that the API and carrier system are likely to encounter in vivo. One valuable characteristic of spheroids is their diffusion limit of about 150–200 μm to many molecules, particularly oxygen [15]. As a result of these mass transport limitations, spheroids generally display gradients of oxygen, nutrient distribution, metabolic waste accumulation, and proliferation profile inside them. Spheroids with diameters larger than 400–500 μm commonly follow a concentrically layered structure consisting of a necrotic core surrounded by a viable layer of quiescent cells and an outer rim of proliferating cells [15, 16]. The peripheral cells closely reflect the in vivo environment of actively proliferating tumor cells next to capillaries, while the distant inner cells stay quiescent or die through necrosis and apoptosis [17]. Many groups have characterized the oxygen profile inside various types of spheroids, and the typical oxygen tensions inside spheroids are of the characteristic profile shown in Figure 1 [1822]. Spheroids thus are excellent models for generating 3-D cultures with such inherent pathophysiological gradients inside. For example, spheroids have been especially valuable in the study of therapeutic problems related to 3-D metabolic and proliferative gradients. Such applications of spheroids have increased understanding of the response of hypoxic tumor cells to chemotherapy and radiation therapy, as well as intricate 3-D cell-cell and cell-matrix interactions [17]. The combined effect from the pathophysiological gradients and complex 3-D cell-cell and cell-matrix interactions often lead to altered RNA and protein expression in spheroids, thus affecting the treatment responses from therapeutic agents. Therefore, they can serve as excellent models for testing drug delivery systems.

Cancer stem cells (CSCs), which are deemed to be responsible for relapse of cancers after treatment, have been cultured successfully as spheroids. Compared with conventional culture, spheroid cultures maintain key properties of stem cells, including gene expression profiles, colony-forming and/or tumorigenic activity, differentiation potential, cytokine secretion, and resistance to chemotherapy [2336]. As an example, sphere-forming primary human colon tumor cells maintain CD133 expression (a marker of colon CSC), and can generate and expand spheroids under serum-free culture conditions, initiate xenograft tumors, and exhibit resistance to chemotherapy-induced apoptosis, whereas cells in 2D conditions fail to do so [25]. Similarly, along with enhanced resistance to standard anticancer agents, cells within ovarian cancer spheroids display increased self-renewal potential, higher invasiveness, and migration potential relative to parental cells [35]. These examples illustrate that spheroids represent a physiological model for these rare cells, and present a great opportunity for therapeutic intervention targeting these cells with different drug delivery approaches.

Mathematical models predicting metabolic, transport and cellular phenomena in spheroids

Mathematical models are invaluable tools for understanding cellular and transport phenomena within spheroids and predicting metabolic responses to drug delivery. Along with experimental studies, mathematical models can provide pertinent perspectives for understanding mechanisms of drug delivery and effectiveness in spheroids. Due to their symmetry, spheroids can be easily defined mathematically in radial dimensions. As shown in Figure 1B, gradients and concentrations of nutrients, oxygen, lactate and glucose, and overall growth rate of cells within a spheroid can be predicted using mass transport and reaction diffusion equations. Mathematical models can also help predict the extent and location of quiescent cells in multicellular spheroids, leading to the finding that the oxygen transport has a greater effect than glucose transport on the distribution of quiescent cells within spheroids [37].

Mathematical models have also been developed that can simulate how the ligand-targeted drugs are distributed within spheroids as well as predict biological responses of the cells to such therapeutic treatments, providing an important tool to understand the kinetics and the dynamics of mechanisms responsible for cellular phenotypes in the spheroids [3740]. For example, mathematical models have helped understand that immunotoxins are generally less effective against spheroids than monolayer cells at equivalent conditions due to heterogeneous receptor distribution and barrier to penetration into the spheroid [39]. They also help predict the poor drug delivery and low rates of cell proliferation that occur in spheroids in part due to hypoxia in necrotic core regions, consequently explaining the limited efficacy of therapies that target proliferating cells [41]. Mathematical models are also applicable to understanding the dynamics of nanoparticle penetration into multicellular spheroids [42].

The utility of mathematical models is demonstrated specifically in the example below, where treatment of a spheroid with a drug is described with the following equations [38]:

nt+1r2(r2vn)r=[km(c)-kd(c)-KG(km(c))f(w)]n, (1)
ct+1r2(r2vc)r=Dr2r(r2cr)-βkm(c)n, (2)
1r2(r2v)r=[VLkm(c)-(VL-VD){kd(c)+KG(km(c))f(w)}]n, (3)
wt+1r2(r2vw)r=Dwr2r(r2wr)-KωG(km(c))f(w)n, (4)

where the dependent variables n, c, v and w are the live cell density (cells/unit volume), nutrient concentration, velocity and drug concentration, respectively. Eq. (1) declares that the rate of change of n is dependent on the difference between the birth (km(c)) and death rates, where death is either natural, at a rate of kd(c), or due to the drug, at a rate KG(km(c)f(w); the constant K is the maximum possible rate of drug induced cell death and f(w) and G(km(c) are positive functions described below. Functions km and kd are based on Michaelis–Menten kinetics:

km(c)=A(cm1ccm1+cm1),kd(c)=B(1-σcm2cdm2+cm2),

where A, B, cc, cd, m1, m2 and σ are positive constants. Eq. (2) arises from Fick’s law describing the diffusion of the nutrients. Eq. (3) states that the spheroid is made up of living cells (volume fraction VLn) and necrotic material (volume fraction 1 -VLn), and the rate of volume change is given by the difference in volume generated via birth (VL km(c)n) from that lost by death ((VLVD)[kd(c) + KG(km(c)f(w)]n), where VL and VD are the inverse densities (volume/cell) of an alive and dead cell. Eq. (4) states that the drug may diffuse (by Fick’s law) and is degraded only when it acts upon a living cell, giving a maximum degradation rate of K/ω, while neglecting any other process of drug breakdown. The dimensionless constant, ω, is a measure of the drug effectiveness, with increasing ω implying that less drug is consumed during the cell killing process.

Numerical analysis of these equations led to the conclusion that drug penetration is the most crucial factor in determining drug effectiveness in spheroids, and one of the main reasons for the difference in drug effectiveness between 2-D cultures and multicellular spheroids. Further, it was observed that an earlier exposure to a drug could generate an outer necrotic rim of cells, which provides an additional barrier to drug access to the remaining cells in the spheroid. Thus, the cells at the core can still survive despite further treatment with high doses of drug. In the future, mathematical models will continue to serve as important tools for understanding drug penetration, transport processes within a spheroid, and responses of cells within the spheroid to drug treatment.

Challenges for using spheroids as drug delivery and efficacy testing models

Although the advantages of multicellular spheroids have been widely recognized [9], challenges involved in the tedious procedures required for spheroid culture are still holding back the biological community from adapting the well-validated spheroid tissue models for studying drug delivery more widely. Here, we focus on the following challenges associated with the use of spheroid models for drug delivery and in vivo efficacy: (1) forming and maintaining spheroids of uniform size; (2) forming spheroids with small numbers of cells; (3) making tissue-like spheroids with multiple cell types; and (4) analyzing tissue-like spheroids (including quantitative assays incompatible with spheroid platforms) and making them compatible with readouts associated with drug delivery and efficacy testing. Each of the challenges and their possible solutions are outlined in the following sections.

Challenge 1: Forming and maintaining spheroids of uniform size

Formation of spheroids occurs spontaneously, in environments where cell-cell interactions dominate over cell-substrate interactions (Figure 2). Conventional methods for spheroid generation include hanging drops, culture of cells on non-adherent surfaces, spinner flask cultures, and NASA rotary cell culture systems [9, 15, 4348]. The hanging drop technique typically employs cell suspension droplets hanging on the underside of the lid of a tissue culture dish due to surface tension. It utilizes gravity to induce aggregation of cells into a single cluster, which then form into a spheroid. The hanging drop method, although simple and allows for defined size spheroid control, is extremely labor intensive. Maintaining spheroids in hanging drops and regular change of medium are time consuming.

Figure 2. Current state-of-the-art methods of generating uniformly sized spheroids.

Figure 2

A) A spheroid co-culture system utilizing chitosan hydrogel microarray molded with PDMS. The uniformity of spheroids generated was controlled by the patterned chitosan. Reproduced from [52] by the kind permission of Elsevier Press. B) Schematic of arrays of v-bottom PDMS microwells to generate uniformly-sized spheroids in a high-throughput fashion. Reproduced from [45] under the Creative Commons Attribution License. C) A microfluidic device with micro-posts to trap single cell suspensions into uniform clusters of cells that will then self-aggregate into spheroids. Reproduced from [49] by the kind permission of Elsevier Press. D) A microfluidic device with micro-chambers to trap single cell suspensions and induce spheroid formation. The size of the spheroids is controlled by the chamber dimensions. Reproduced from [51] by the kind permission of Springer Science. E) Cartoon of a v-bottom microwell-based device to generate spheroids of the same size. Reproduced from [44] by the kind permission of Elsevier Press. F) Schematic showing a two-layer PDMS-based microfluidic device for the generation of uniformly-sized embryoid bodies. After introducing cells into the microchannel (with surfaces blocked to prevent cell adhesion) as a uniform monolayer, cells spontaneously aggregate and break up into uniformly-sized embryoid bodies. Reproduced from [50] by permission of The Royal Society of Chemistry. (G) 384 hanging drop array plate and a cartoon of the spheroid formation process. The size of the spheroid is controlled by the number of cells seeded into each hanging drop. Reproduced from [59] by permission of The Royal Society of Chemistry.

Another technique for making spheroids involves culture of cells on non-adherent surfaces, which essentially prevents cells from attaching to a substrate, thus inducing cells to aggregate with themselves to form spheroids. Culture of cells in large bioreactors such as spinner flasks and NASA rotary cell vessels produces spheroids via continuous spinning of cell suspension to keep cells from settling down. Such systems are suitable for mass production and long-term culture but are limited by non-uniform spheroids and the need for specialized equipment. Size uniformity is critical for achieving reproducible experimental results since size affects cell behavior and function as well as drug penetration and transport.

More advanced spheroid formation methods developed recently often utilize micro- and nano-technologies. Such techniques generally provide great improvements over spheroid size-control, cellular composition, and throughput. Various groups have developed microfluidic devices to trap clusters of cells into micro-chambers, or posts [4951] to control spheroid sizes and simplify liquid handling procedures (Figure 2). Others have created arrays of microwells to increase throughput and generate defined-size spheroids [44, 45, 52]. Through forced-aggregation techniques and micro-textured surfaces, scalable uniform size-specified embryoid bodies were generated from human embryonic stem cells, which can differentiate to clinically relevant cell types [45]. In addition, some groups have developed surface-modified substrates, porous 3D scaffolds, or nano-imprinted scaffolds to induce spheroid formation [27, 5357]. Using photolithography, microwell arrays have been generated from poly(ethylene glycol) (PEG)-based hydrogels. These gel arrays are effective in confining single cells and aggregates, and in guiding their reorganization into spheroids [58]. However, many of these sophisticated platforms are complicated and thus require specially trained users to fabricate and operate the devices. Or they may allow efficient spheroid formation but not allow for convenient prolonged culture and drug testing where media exchange or reagent additions are required. In addition, many of these devices suffer from material compatibility issues with many hydrophobic compounds such as anti-cancer therapeutic drugs, which limits the applicability of these high-throughput devices to drug screening and testing.

Recently, our group has developed a 384-format hanging drop array plate as a user-friendly solution to spheroid formation and culture [5961]. This versatile platform is based on the conventional hanging drop method, allows for high-throughput drug screening and testing, is amenable to high-throughput multiplexing (e.g. liquid handling robots and plate readers), but greatly simplifies the tedious liquid handling processes as well as increased robustness of droplet stability to allow long-term spheroid culture (Figure 2G). The hanging drop plate comprises of access holes that can accommodate a 15–20 μl drop of cell suspension in culture medium confined by the diameter of the plateau on the bottom surface, which results in consistent geometry of the hanging drop without spreading out, and leads to more robust and stable culturing conditions not possible on conventional flat hanging drop substrates. The cells slowly aggregate in the bottom center of the hanging droplet, and eventually form into a tight spheroid. The access holes on the top of the plate allow direct manipulation of the droplets, greatly simplifying the initial droplet formation and subsequent media exchange procedures by eliminating the tedious hanging drop culture dish inversion required in the conventional hanging drop method.

Using this platform we have successfully formed spheroids (mono-, co- and tri-cultures) with varying initial cell seeding density, ranging from 50 cells to 15,000 cells with different cell lines and primary cells [5961]. The spheroids created on this platform have been analyzed for their cell-based assay capabilities by testing their sensitivity to anticancer drugs with distinctly different activity profiles: a conventional anticancer drug 5-fluorouracil (5-FU) that inhibits cellular proliferation, and a hypoxia-triggered cytotoxin tirapazamine (TPZ) that causes DNA damage at low oxygen tensions [59]. Compared to 2D cultures, 3D spheroids of epithelial carcinoma cell line (A431.H9) were more resistant to 10 μM 5-FU treatment, suggesting the presence of quiescent cells in the spheroids. However, 3D spheroids were susceptible to 10 M TPZ treatment compared to 2D controls, possibly because active oxygen consumption by cells and limits in diffusive oxygen transport creates a hypoxic core similar to in vivo tumors. When treated with these drugs together, there was an additive effect on the cell death in 3D spheroids, since 5-FU targets proliferating cells in the peripheral layers of spheroids and TPZ kills cells in the hypoxic core of spheroids [59].

Unfortunately, the first generation of the 384 hanging drop array plate was susceptible to small mechanical shocks and could not reliably maintain hanging drops for longer than 14 days. Stabilization of liquid droplets was enhanced in the second generation of the 384 well plate, which was optimized with micro-rings to stabilize droplets against mechanical perturbations and prevent surface fouling. This system enables long-term cell spheroid culture for more than 22 days within the droplet array [60, 61]. With enhanced droplet stability, the second-generation hanging drop array plates provide new opportunities for high-throughput preparation of microscale 3D cell constructs for drug screening and cell analysis. A systematic analysis of the advantages and disadvantages of different spheroid forming techniques is listed in Table 1.

Table 1.

Comparison of different spheroid forming approaches

Method Advantages Disadvantages
Hanging Drops [9, 15, 22, 43, 46, 47] Simple
No specialized equipment required
Defined spheroid size control
Uniform spheroid size
Co-culture with defined cellular composition control
Tedious handling
Time consuming
Not stable
Long-term culture difficult
Low throughput
Not efficient
Non-adherent Surfaces [15, 43] Simple
No specialized equipment required
High-throughput (*Multiwell plate-format only)
Compatible with HTS instruments (*Multiwell plate-format only)
Not efficient
Non-uniform spheroid size
No control of spheroid size
Tedious media exchange
Defined co-culture cellular composition difficult
Spinner Flask [15, 43] Simple
Mass production of spheroids
Long-term culture
Special equipment required
No individual compartment for each sample
No spheroid uniformity control (composition and size)
High shear force for cells
NASA Rotary Cell Vessel [7, 15, 43] Simple
Mass production of spheroids
Long-term culture
Special equipment
No individual compartment for each sample
No spheroid uniformity control (composition and size)
Micro/Nano Technology: Microfluidic Devices and Spheroids on a Chip [4951, 55, 85] Improved efficiency
Uniform spheroid size control
Uniform cellular composition
High-throughput
Simplified handling procedures
Analysis may be integrated into the device
Complicated set-up and device fabrication
Customized devices may require trained users to operate
Devices not mass producible for research community
Material compatibility issue with drugs (hydrophobic compounds)
Compatibility with HTS instruments
Spheroid retrieval for further analysis difficult
Long-term culture issue
Lack of isolated compartment for each sample
Micro/Nano Technology: Microwell Arrays [44, 45, 52, 86] Improved efficiency
High-throughput production
Uniform spheroid size control
Simplified handling
Long-term culture issue
Sample retrieval difficult
Fabrication of device require trained users
Devices not mass producible for research community
Material compatibility issue with drugs (hydrophobic compounds)
Compatibility with HTS instruments
Long-term culture issue
Micro/Nano Technology: Scaffold Technology [57, 8792] High-throughput production
Spheroids immobilized on scaffolds so easy handling
Long term culture
Mass production of scaffolds possible
May be compatible with multiwell plates
May be compatible with HTS instruments
Uniform spheroid size
Samples retrieval for further analysis difficult
Scaffold material biocompatibility and biodegradability issues
Multiwell Hanging Drop Plates [5961, 93] Material compatibility with drugs
High-throughput
Compatible with HTS instruments
Individual compartment for each sample for HTS
Long term culture
Robust assay performance
User-friendly
Mass production of device possible
Defined spheroid size control
Uniform spheroid size
Co-culture with defined cellular composition control
Media exchange tedious without HTS instrument
Hard to work with small number of cells

Challenge 2: Forming spheroids with small number of cells

Efficient formation of spheroids requires segregated groups of cells maintained within close proximity to each other to aggregate together, which becomes especially challenging when the number of available cells is small, as in the case of primary stem cells and CSCs. Many groups focusing on cancer stem cell research have formed spheroids from single cells by monoclonal growth [23, 6268]. This method has been particularly successful for certain cancer cell lines as many transformed cells are anchorage independent, thus do not require a stiff substrate to survive. Such cultures have been shown to enrich the stem cell subpopulation within bulk tumor cells. Nevertheless, manipulating, forming, growing, and maintaining spheroid from single cells can be very challenging. First of all, it is very difficult to titrate and manipulate cell suspensions into single cell levels, especially from small numbers of CSCs that may have already undergone stress through isolation and sorting. In addition, for cells that are still anchorage-dependent, lack of cell-cell interaction often slows down proliferation and may even lead to cell death. Therefore, such experiments have not only been challenging, but also very time-consuming. To overcome these challenges, the addition of Matrigel or other extracellular matrix (ECM) proteins (such as collagen and fibronectin) or materials such as methylcellulose to cultures have been shown to enhance spheroid assembly by providing cell-matrix support [6974]. Alternatively, small numbers of CSCs may be co-cultured with other supporting cell types as spheroids to achieve monoclonal expansion of the tumor cell within a supportive 3-D microenvironment.

Challenge 3: Making tissue-like spheroids with multiple cell types

Tissues in vivo typically integrate interactions between the surrounding multiple cell types: such as, epithelial cells, fibroblasts, mesenchymal stem cells (MSCs), endothelial cells, and immune cells, and the abundant extracellular matrix proteins, immobilized protein factors, proteoglycans, mineralized tissue, soluble protein factors, and small molecule signals [75]. Response of cells to drug delivery systems is dependent on these interactions; indeed, drugs effective against specific cells in one site may not work against the “same” cells in a different organ/tissue [76, 77]. In order to faithfully develop effective drug delivery systems, some of the in vivo microenvironmental interactions need to be captured in vitro in 3-D models of tissue, such as spheroids. We and others have established that spheroids can integrate multiple cell types, ECM and other molecules. With our 384 hanging drop system, we have successfully formed spheroids (mono-, co- and tri-cultures) with a varying initial cell seeding density, ranging from 50 cells to 15,000 cells with different cell lines and primary cells [5961]. Additionally, the 384 hanging drop array plate is also capable of spheroid transfer and retrieval for Janus spheroid formation, sequential addition of cells for concentric layer patterning of different cell types, and culture of a wide variety of cell types [60, 61]. Mouse mammary tumors were recapitulated recently in spheroids with tumor and endothelial cells and it was found that the initial primary or micrometastatic stages of solid tumor progression could be more accurately modeled in co-culture spheroids than conventional 2-D cultures [78]. When implanted in vivo, the co-culture tumor-endothelial spheroids enhanced angiogenesis and metastasis, compared to tumor cell only spheroids, emphasizing the importance of crosstalk between the cells comprising the tumor microenvironment.

Microscale technologies have also been applied to spheroid co-cultures. Co-culture spheroid arrays have been developed using micropatterning techniques [52, 79]. Hepatocyte spheroids co-cultured with endothelial cells or fibroblasts maintain their liver-specific functions for at least 1 month in such formats. Growing spheroids inside PDMS microfluidic bioreactors exposes them to uniform and non-uniform flow. The culture of HepG2 cells was demonstrated in such microbioreactors, and cell functionality (albumin production, glucose consumption) was evaluated [80]. Similarly, breast tumor cells were encapsulated within alginate, and gelled in situ within the microchannels of a microfluidic device. The cell aggregates were treated with various concentrations of doxorubicin. Drug effects on cell viability and proliferation were measured [81].

Interactions of cancer cells with the cellular and acellular components of their stroma control the growth and malignant behavior of invasive cancer [82] and should be included in at least rudimentary form in tumor models used in screening anti-cancer drug candidates. This requirement has been addressed by generating tumor histoids (TH) which are formed through the spontaneous interaction of tumor and stromal cells to form spheroidal TH that have a microscopic architecture closely resembling that of true tumor microlesions [83, 84]. TH generated in low shear, rotating suspension culture (original method) varied in size and tumor cell content (Figure 3). Although they could be analyzed and sorted by COPAS flow cytometry (which can accommodate samples as large as 100 – 1,500 μm) to deliver relatively uniform TH into multiwell plates, the additional manipulation was a drawback. TH production is currently being adapted to the 384 hanging drop format referred to above to address this problem.

Figure 3. Tissue and tumor histoids are generated through the spontaneous interaction of multiple cell types.

Figure 3

A) Histological section of breast cancer histoid (BCH) containing BT20 cancer cells and, generated in low shear, rotating suspension culture and immunostained to demonstrate cytokeratin in cancer cells. Photographed using 20x objective; inset shows size variability. Suspension of intact, viable pancreatic tumor histoids containing fibroblasts and pancreatic cancer cells (MIAPaCa) were generated in rotating suspension cultures, and imaged under B) 10x and C) 20x objectives. Pancreatic cancer cells were transfected to express green fluorescent ZsGreen 1, a variant of the widely used GFP. These histoids can be sorted according to size and fluorescent intensity (using COPAS flow cytometer) to obtain a relatively uniform population.

Challenge 4: Characterization of drug testing and efficacy in spheroid platforms

One of the major reasons spheroids have not entered mainstream drug screening process is due to the lack of simple, controlled techniques and protocols for rapid, standardized assay of cellular responses in situ, prediction of in vivo activity, as well as poor integration with high throughput systems widespread in industrial labs. However, new developments in spheroid technology are challenging these dogmas. Analyses must integrate quantitative population based information (cell number, position), subcellular functional readouts (viability, cell cycle, and signaling), along with more qualitative information (microscopy and histology). As spheroid models become widespread in biological studies, the feasibility of conducting standard biological assays with cells grown on such platforms must be carefully considered. Harvesting the spheroids from the culture platform such as microfluidic devices and hanging drops poses a challenge. Spheroids are often lost during medium exchange. Compatibility with standard biological assays such as RT-PCR, western blots, flow cytometry, are also a challenge when working with spheroid models. Here we discuss amenability of current 3-D spheroid systems for drug delivery systems.

Screening techniques currently applied to monolayer cultures in multiwell plate formats, such as various assays for cytotoxicity, proliferation, drug binding, apoptosis, and ATP level, can be adapted to spheroids. Measurement of the growth or shrinkage of spheroids in response to the drug can be accomplished by standard phase-contrast microscopy and image analysis. Moreover, confocal microscopy techniques can be developed for measuring drug penetration into individual spheroids [9497], which could potentially be automated into a high throughput approach. Important assays such as drug binding, uptake and penetration can be accomplished using spheroids in a moderate to high throughput manner as well.

Standard biological assays such as fluorescence activated cell sorting (FACS) can also be applied to spheroids. For example, the differential cytotoxicity in the cells near the hypoxic core of the spheroid relative to the well-oxygenated peripheral cells, was measured by FACS after selectively recovering cells from various depths within the spheroids [98].

Due to the need for a large number of cells within a test sample, molecular biological assays such as western blotting and RT-PCR are difficult to perform on lysates obtained from spheroids. However, microscale versions of western blot have now been developed [99101] which are capable of handling samples with low cell numbers. Quantitative RT-PCR assay still remains a challenge for the RNA harvested from spheroids. However, if lysates can be collected from at least 2000 cells per sample (by pooling duplicates), RNA can be purified using RNA purification kits designed for rare cell samples, and subsequently be used to make cDNA for qRT-PCR gene expression analysis. Alternatively, for low cell number spheroids, kits that convert cells directly to cDNA bypassing the RNA purification step (Ambion Inc. Cells-to-cDNA II Kit), can be utilized for qRT-PCR analysis of mRNA levels. Although the minimum number of cells required for flow cytometric analysis (at least 10,000) is higher than what can be typically isolated from a single spheroid, samples can be pooled and aggregates disassociated for analysis of cells post culture. Microscale flow cytometers that do not require such a large number of cells in a sample can also be utilized [102]. Pooling of samples would necessitate a large number of samples, which are resource and time consuming to maintain. Moreover, all of these techniques add unwanted complexity to the spheroids culture and maintenance, and may not be easily integrated into cell biology labs that have little expertise in microscale technologies.

Quantitative measurements of cell type specific phenotypes are often difficult to assess in co-culture spheroids. For example, one of the methods for quantifying proliferation rates of each cell type would require labeling each cell type via cell tracker stains, which tend to bleed over time, or tagging the cells with fluorescently expressed proteins such as GFP, which require additional transfection steps before spheroid formation. Another method for quantifying proliferation rates of each cell type could involve flow cytometric analysis of harvested spheroids at a given time point, stained with markers that identify each cell type, and would require a large sample size of spheroids in order to obtain at least 10,000 cells that are needed for flow cytometric analysis.

Emerging micro- and nanotechnologies have enabled fabrication of many different spheroids-on-a-chip devices that substantially increase throughput for preparing spheroid cultures. However, hydrophobic materials, such as poly(dimethylsiloxane) (PDMS), used in many of these devices bind hydrophobic compounds, making the devices incompatible with library screening and drug testing applications. Even if material is compatible with anti-cancer therapeutic compounds, chip formats are usually not compatible with HTS instruments (robots/plate readers). Therefore, there is a need for incorporating assay readouts and analyses within the devices.

Many commercially available cellular assays are designed for 2-D monolayer cultures, so caution must be taken when applied to 3-D spheroids. Such assays need to be validated in spheroid platforms with appropriate controls and generation of standard curves. We recently evaluated robustness of the 384 hanging drop plate array in terms of assay performance by calculating Z- factors for fluorescence- and colorimetric-based assays, to allow for practical 3-D cell-based high-throughput screening and enable broader use of the 384 hanging drop plate [61]. Z-factor, an assay performance measurement that provides an easy and useful summary of assay quality and robustness [103105], was found to be well above 0.5 for both the fluorescence- and colorimetric-based assays, indicating that such assays performed in the 384 hanging drop plate would have excellent accuracy and robustness in terms of readout reliability [60]. When using assay readout reagents, care must be taken that there is sufficient reagent penetration to provide readouts from a representative number of cells within the spheroid. Reagent penetration is in and of itself a sort of “drug delivery” process that depends on reagent type, cell types, as well as spheroid size. We and others have demonstrated that high throughput assays such as alamarBlue can be standardized for high throughput spheroid assays in general and the 384-hanging drop array platform in particular.

Bioluminescence assays with luciferase enzymes have emerged as a powerful alternative to fluorescence and colorimetric readouts for high throughput screening in cell-based formats. Advantages of bioluminescence relative to fluorescence include enhanced sensitivity due to lower background, greater dynamic range of quantitative signal, and reduced susceptibility to interference by colored compounds [106]. Bioluminescent reporters have been used successfully to identify promising therapeutic agents, including compounds that perturb tumor-stromal signaling [107]. Benefits and proven efficacy of bioluminescence for compound screening and drug testing in two dimensional cell cultures support use of luciferase-based readouts for high throughput screening and validation studies in spheroids [108].

Bioluminescence reporters for drug testing in spheroids can be designed to quantify basic parameters such as cell viability, or more advanced assays can be developed to measure protein-protein interactions, cell signaling, and/or promoter activity. Beetle luciferases, including firefly and click beetle luciferases, are ATP-dependent enzymes that provide robust quantification of cell viability [109]. Bioluminescence from these enzymes is directly proportional to numbers of living cells over 4 orders of magnitude, enabling facile analysis of drug cytotoxicity in intact spheroids (Figure 4). To measure effects of drugs on specific molecular pathways, investigators have established reporters and biosensors based on split luciferase complementation [110]. While not yet used for drug testing in spheroids, split luciferase complementation assays for key drug targets, such as ligand binding to seven-transmembrane receptors and activation of the proto-oncogene AKT, have been validated in 2-D cultures [111, 112]. Integration of such reporter systems into spheroid models will advance high throughput screening for molecularly-targeted drugs. In addition, availability of luciferase enzymes with distinct substrates and/or emission spectra will allow readouts of multiple drug targets in individual spheroids, increasing efficiency of a single screening assay.

Figure 4. Bioluminescence quantification of doxorubicin cytotoxicity in MDA-MB-231 human breast cancer cells.

Figure 4

A) Image of luciferase activity in intact spheroids, and B) quantification of bioluminescence after 48 hours of treatment. Luminescence data were normalized to pre-treatment values for each spheroid. Loss of bioluminescence directly measures relative cytotoxicity. Error bars denote mean values + SEM.

Comparison of drug delivery and efficacy between multicellular spheroids and in vivo models

In vitro models using multicellular spheroids provide an important link between monolayer cell cultures and animal experiments for effective drug delivery testing. Key similarities between outcomes of drug targeting in spheroid and animal models are discussed below:

Drugs can be delivered to cells within spheroids by several techniques such as liposomes, nanoparticles, and nonviral gene delivery [1, 113, 114]. Encapsulation of certain drugs in lipid vesicles (liposomes) has been shown to result in reduced toxic effects [115]. Parameters such as liposomal surface charge, mean diameter, lipid bilayer fluidity, lipid bilayer composition, duration of liposome-spheroid interaction, mean liposome size, steric stabilization of liposomes and fusogenicity impact spheroid penetration [95, 116, 117]. Non-viral gene transfer to spheroids was optimized for 22 kDa linear and 25 kDa branched polyethyleneimine (PEI) molecules [118]. Confocal imaging of fluorescent PEI indicated that the cationic complexes could only penetrate the outer 3–5 proliferating cell layers of a spheroid, sparing the deeper quiescent cells, which could be targeted via electroporation. Although the efficacy of electroporation in quiescent tissue was improved, gene expression was still confined to the outer regions of the spheroid, suggesting that expression of transgenes is better in diving cells, and that limited access to central regions of the spheroid remain a significant barrier to gene delivery.

Extracellular matrix in tumors restrict nanoparticle penetration for cancer imaging and therapy. To overcome this barrier, effects of nanoparticle size and collagenase treatment on penetration of carboxylated polystyrene nanoparticles were systematically assessed in a multicellular spheroid model [119]. It was found that penetration of nanoparticles into the spheroid core was limited to nanoparticles smaller than 100 nm. However, collagenase-coated, 100 nm nanoparticles demonstrated a 4-fold increase in the number of particles delivered to the spheroid core compared with control nanoparticles, revealing that nanoparticle delivery to tumors may be substantially improved by incorporation of extracellular matrix-modulating enzymes in the delivery formulation.

The rate of drug diffusion through the tissue is determined by physicochemical properties such as molecular weight, shape, charge and aqueous solubility. Penetration of a drug into a tissue is also dependent on its consumption, which removes the amount of free drug, thus inhibiting further permeation. Spheroids can serve as excellent model systems to study permeation of drugs and their carriers through organs and tissues. Low permeability of delivery systems across the blood-brain barrier (BBB), drug resistance, and poor penetration into tumor tissue limit treatment strategies in glial tumors. Dhanikula et al. studied permeability of methotrexate (MTX)-loaded polyether-copolyester (PEPE) dendrimers across an in vitro BBB model and their distribution into avascular human glioma tumor spheroids to evaluate their potential as drug carriers for gliomas [120]. Glucosylated dendrimers were endocytosed in significantly higher amounts than nonglucosylated dendrimers. Moreover, glucosylation enhanced cumulative permeation of dendrimers across the BBB and avascular tumor spheroids and increased the amount of MTX delivered to tumor cells. These results established that glucosamine can be used as an effective carrier not only for targeting glial tumors but also for enhancing permeability across BBB. Thus, glucosylated PEPE dendrimers can serve as potential delivery system for the treatment of gliomas.

To assess in vivo tumor targeting and therapeutic efficacy, neuroblastoma spheroids and nude mice xenografted with human neuroblastoma cells were treated with radioisotope iodine-131-metaiodobenzylguanidine ([131] MIBG) without any added carrier [121]. The drug was uptaken successfully in the tumor in vivo, where it reduced tumor burden, and prevented regrowth of spheroid in vitro, suggesting that the spheroid results correlate well with those in the animal model. However, the large spheroids were more vulnerable to the treatment than smaller spheroids, indicating that the in vivo treatment may spare smaller micrometastases [122]. Similarly, cytotoxic activity of the galactoside-specific lectin from mistletoe towards anaplastic glioma was investigated in a spheroid model and in vivo in rats intracerebrally implanted with F98 glioma cells [123]. Dose dependent cytotoxicity was observed in spheroids as well as in vivo. These studies indicate that spheroids are a relevant model of the in vivo tumors, and screening of cancer therapeutics in spheroid models can lead to selection of promising drug candidates and decrease time between drug discovery and clinical trials.

Drugs (such as anthracyclines) accumulate preferentially in cells at the periphery of spheroids, resulting in reduced rates and concentrations of drug delivered to cells in the interior. However, drugs are retained more effectively in intact spheroids than in dispersed cells, resulting in more cytotoxicity in situ in spheroids [124]. Doses of drugs can also have varying effects on tumor spheroids. While low concentrations (<100 μM) of gamma-linolenic acid (GLA), an essential fatty acid, increased both apoptosis and proliferation with a net increase in glioma tumor growth and invasion, higher doses significantly impaired cell growth. The proliferative effects of low-dose GLA could be a hazard in the clinical treatment of malignant glioma; however, because of the low toxicity of GLA against normal cells, local delivery of millimolar doses of GLA could significantly reduce tumor size [125].

Similar to in vivo tissues, spheroids are generally more resistant than 2-D cultures to a given dose of many drugs [126131]. Reasons for this difference involve factors including drug penetration, cell–cell contact, cell-cycle distribution and varying microenvironment (such as pH, extracellular matrix) within spheroids [12, 132138]. For example, intercellular contacts promote cell survival through activation of signaling pathways such as PI3K/Akt, NF-κB and Stat3. Moreover, the reductions in rate of cell division within the spheroid impairs the effects of many cancer drugs since most anti-cancer drugs exert selective toxicity on dividing cells. Further, drug binding to ECM not only slows down the movement of drugs towards target cells, but also changes the number of available drug molecules. Drug penetration can also change due to the modifications in its charge.

Not surprisingly, similar to the in vivo tissues, penetration of a drug into a spheroid is the rate-limiting step for drug delivery, and it has been shown that some drugs may penetrate at non-negligible quantities only to a depth of a few cells despite prolonged exposure [94, 95, 97, 137139]. Some drug molecules (non-conjugated or conjugated with targeting agents such as radionuclides) can easily penetrate through spheroids while others cannot. Durand et al. treated V79 (lung fibroblast) spheroids with doxorubicin, bleomycin, 5-fluorouracil, carmustine, cisplatin, chlorambucil, and mitomycin. Drug penetration within a spheroid is also a function of the drug molecule structure. As an example, doxorubicin penetrates poorly into V79 spheroids as compared with other tested drugs tested [140]. Lack of drug penetration can be due to avid binding to surface cells [141] and/or barriers to drug transport due to extracellular matrix, cell-cell junctions, and cell membranes [94, 97, 116, 137]. Due to pH gradients within the spheroids, cells in the acidic core region can be protected from weak acid drugs such as mitoxantrones and anthracyclines, at the same time potentiating the effect of weak base drugs such as chlorambucil and mitomycin C [128, 142, 143]. Due to mass transport limitations and a 3-D structure similar to in vivo tissues, drug penetration limitations occurring in vivo can be studied in spheroids.

Testing in multicellular spheroids is increasingly regarded as an essential step in drug development [16, 43, 134, 144, 145]. As discussed in this article, a variety of biomedical researchers have applied multicellular spheroid models for assessing various therapeutic treatments, some of which include: chemotherapy including target-specific approaches, experimental radiotherapy followed by photodynamic treatment, cell- and antibody-based immunotherapy, gene therapy and combinatorial therapies [43, 126, 127, 132, 133, 145158]. Together, these experiments demonstrate that spheroid-based assays have improved predictive power for in vivo therapeutic efficacy. These examples also reveal that structure-function relationships from interactions of different delivery systems with 3-D spheroids can lead to construction of improved delivery systems for drug penetration and targeting into tissues. The spherical shape is also easy to analyze and model mathematically, providing benefits as a screening tool for a variety of therapeutics.

Conclusions

Development of uniform, high throughput, multi-cell-type spheroids present an opportunity to screen and select drugs and drug delivery systems in vitro in a format that more closely resembles conditions in patients. Spheroids have similarities to in vivo tissues in terms of complexity of cell types (spheroids can be made with many cell types), key cell-cell interactions, extracellular matrix deposition by cells, and chemical gradients. These features of spheroids limit drug diffusion to an extent comparable to organs and tissues in vivo. Spheroids can support growth of rare cell types such as cancer stem cells, providing a model for testing drugs targeted against specific cell populations and types that are otherwise difficult to perform conveniently. Despite challenges that have limited broader use for the past half century, advances including new technologies for spheroid formation, culture, and analysis that are highlighted in this review suggest that use of spheroids are ripe for becoming a mainstream in vitro tool for testing drug delivery and efficacy, providing readouts that better predict results in patients than standard 2D cultures.

Acknowledgments

We apologize to those authors whose work has not been included in this review due to space limitations. This work is supported by grants R01CA136553, R01CA136829, and P50CA093990 from the US National Institutes of Health, the Coulter Foundation, the NRF-WCU program funded by MEST (R322008000200540), and a generous gift from Jacque Passino.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Han C, Wang B. Factors that impact developability of drug candidates: an overview. In: Wang B, Siahaan T, Soltero R, editors. Drug Delivery: Priciples and Applications. John Wiley & Sons; Hoboken, New Jersey: 2005. pp. 1–14. [Google Scholar]
  • 2.Saltzman WM. Drug Delivery: Engineering Principles for Drug Therapy. Oxford University Press; New York: 2001. [Google Scholar]
  • 3.Torchilin VP. Passive and active drug targeting: drug delivery to tumors as an example. In: Schafer-Korting M, editor. Drug Delivery. Springer; 2010. pp. 3–54. [DOI] [PubMed] [Google Scholar]
  • 4.Jain KK. Drug Delivery Systems- An Overview. In: Jain KK, editor. Drug Delivery Systems. Humana Press; 2008. pp. 1–50. [Google Scholar]
  • 5.Karunaratne DN, Silverstein PS, Vasadani V, Young AM, Rytting E, Yops B, Audus KL. Cell culture models for drug transport studies. In: Wang B, Siahaan T, Soltero R, editors. Drug Delivery: Priciples and Applications. John Wiley & Sons; Hoboken, New Jersey: 2005. pp. 103–124. [Google Scholar]
  • 6.Haycock JW. 3D cell culture: a review of current approaches and techniques. Methods Mol Biol. 2011;695:1–15. doi: 10.1007/978-1-60761-984-0_1. [DOI] [PubMed] [Google Scholar]
  • 7.Ingram M, Techy GB, Saroufeem R, Yazan O, Narayan KS, Goodwin TJ, Spaulding GF. Three-dimensional growth patterns of various human tumor cell lines in simulated microgravity of a NASA bioreactor. In Vitro Cell Dev Biol Anim. 1997;33:459–466. doi: 10.1007/s11626-997-0064-8. [DOI] [PubMed] [Google Scholar]
  • 8.Torisawa YS, Takagi A, Shiku H, Yasukawa T, Matsue T. A multicellular spheroid-based drug sensitivity test by scanning electrochemical microscopy. Oncol Rep. 2005;13:1107–1112. [PubMed] [Google Scholar]
  • 9.Kelm JM, Fussenegger M. Microscale tissue engineering using gravity-enforced cell assembly. Trends Biotechnol. 2004;22:195–202. doi: 10.1016/j.tibtech.2004.02.002. [DOI] [PubMed] [Google Scholar]
  • 10.Olive PL, Durand RE. Detection of hypoxic cells in a murine tumor with the use of the comet assay. J Natl Cancer Inst. 1992;84:707–711. doi: 10.1093/jnci/84.9.707. [DOI] [PubMed] [Google Scholar]
  • 11.Oloumi A, Lam W, Banath JP, Olive PL. Identification of genes differentially expressed in V79 cells grown as multicell spheroids. Int J Radiat Biol. 2002;78:483–492. doi: 10.1080/09553000210122299. [DOI] [PubMed] [Google Scholar]
  • 12.Kim SH, Kuh HJ, Dass CR. The reciprocal interaction: chemotherapy and tumor microenvironment. Curr Drug Discov Technol. 2011;8:102–106. doi: 10.2174/157016311795563875. [DOI] [PubMed] [Google Scholar]
  • 13.Shekhar MP. Drug resistance: challenges to effective therapy. Curr Cancer Drug Targets. 2011;11:613–623. doi: 10.2174/156800911795655921. [DOI] [PubMed] [Google Scholar]
  • 14.Andre F, Berrada N, Desmedt C. Implication of tumor microenvironment in the resistance to chemotherapy in breast cancer patients. Curr Opin Oncol. 2010;22:547–551. doi: 10.1097/CCO.0b013e32833fb384. [DOI] [PubMed] [Google Scholar]
  • 15.Lin RZ, Chang HY. Recent advances in three-dimensional multicellular spheroid culture for biomedical research. Biotechnol J. 2008;3:1172–1184. doi: 10.1002/biot.200700228. [DOI] [PubMed] [Google Scholar]
  • 16.Kunz-Schughart LA, Freyer JP, Hofstaedter F, Ebner R. The use of 3-D cultures for high-throughput screening: the multicellular spheroid model. J Biomol Screen. 2004;9:273–285. doi: 10.1177/1087057104265040. [DOI] [PubMed] [Google Scholar]
  • 17.Hirschhaeuser F, Menne H, Dittfeld C, West J, Mueller-Klieser W, Kunz-Schughart LA. Multicellular tumor spheroids: an underestimated tool is catching up again. J Biotechnol. 2010;148:3–15. doi: 10.1016/j.jbiotec.2010.01.012. [DOI] [PubMed] [Google Scholar]
  • 18.Bredel-Geissler A, Karbach U, Walenta S, Vollrath L, Mueller-Klieser W. Proliferation-associated oxygen consumption and morphology of tumor cells in monolayer and spheroid culture. J Cell Physiol. 1992;153:44–52. doi: 10.1002/jcp.1041530108. [DOI] [PubMed] [Google Scholar]
  • 19.Gassmann M, Fandrey J, Bichet S, Wartenberg M, Marti HH, Bauer C, Wenger RH, Acker H. Oxygen supply and oxygen-dependent gene expression in differentiating embryonic stem cells. Proc Natl Acad Sci U S A. 1996;93:2867–2872. doi: 10.1073/pnas.93.7.2867. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Kunz-Schughart LA, Groebe K, Mueller-Klieser W. Three-dimensional cell culture induces novel proliferative and metabolic alterations associated with oncogenic transformation. Int J Cancer. 1996;66:578–586. doi: 10.1002/(SICI)1097-0215(19960516)66:4<578::AID-IJC25>3.0.CO;2-2. [DOI] [PubMed] [Google Scholar]
  • 21.Sutherland RM. Cell and environment interactions in tumor microregions: the multicell spheroid model. Science. 1988;240:177–184. doi: 10.1126/science.2451290. [DOI] [PubMed] [Google Scholar]
  • 22.Wartenberg M, Donmez F, Ling FC, Acker H, Hescheler J, Sauer H. Tumor-induced angiogenesis studied in confrontation cultures of multicellular tumor spheroids and embryoid bodies grown from pluripotent embryonic stem cells. Faseb J. 2001;15:995–1005. doi: 10.1096/fj.00-0350com. [DOI] [PubMed] [Google Scholar]
  • 23.Krohn A, Song YH, Muehlberg F, Droll L, Beckmann C, Alt E. CXCR4 receptor positive spheroid forming cells are responsible for tumor invasion in vitro. Cancer Lett. 2009;280:65–71. doi: 10.1016/j.canlet.2009.02.005. [DOI] [PubMed] [Google Scholar]
  • 24.Bartosh TJ, Ylostalo JH, Mohammadipoor A, Bazhanov N, Coble K, Claypool K, Lee RH, Choi H, Prockop DJ. Aggregation of human mesenchymal stromal cells (MSCs) into 3D spheroids enhances their antiinflammatory properties. Proc Natl Acad Sci U S A. 2010;107:13724–13729. doi: 10.1073/pnas.1008117107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Fang DD, Kim YJ, Lee CN, Aggarwal S, McKinnon K, Mesmer D, Norton J, Birse CE, He T, Ruben SM, Moore PA. Expansion of CD133(+) colon cancer cultures retaining stem cell properties to enable cancer stem cell target discovery. Br J Cancer. 2010;102:1265–1275. doi: 10.1038/sj.bjc.6605610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Bratt-Leal AM, Kepple KL, Carpenedo RL, Cooke MT, McDevitt TC. Magnetic manipulation and spatial patterning of multi-cellular stem cell aggregates. Integr Biol (Camb) 2011;3:1224–1232. doi: 10.1039/c1ib00064k. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Huang GS, Dai LG, Yen BL, Hsu SH. Spheroid formation of mesenchymal stem cells on chitosan and chitosan-hyaluronan membranes. Biomaterials. 2011;32:6929–6945. doi: 10.1016/j.biomaterials.2011.05.092. [DOI] [PubMed] [Google Scholar]
  • 28.Miyagawa Y, Okita H, Hiroyama M, Sakamoto R, Kobayashi M, Nakajima H, Katagiri YU, Fujimoto J, Hata J, Umezawa A, Kiyokawa N. A microfabricated scaffold induces the spheroid formation of human bone marrow-derived mesenchymal progenitor cells and promotes efficient adipogenic differentiation. Tissue Eng Part A. 2011;17:513–521. doi: 10.1089/ten.TEA.2009.0810. [DOI] [PubMed] [Google Scholar]
  • 29.Raof NA, Raja WK, Castracane J, Xie Y. Bioengineering embryonic stem cell microenvironments for exploring inhibitory effects on metastatic breast cancer cells. Biomaterials. 2011;32:4130–4139. doi: 10.1016/j.biomaterials.2011.02.035. [DOI] [PubMed] [Google Scholar]
  • 30.Cheng NC, Wang S, Young TH. The influence of spheroid formation of human adipose-derived stem cells on chitosan films on stemness and differentiation capabilities. Biomaterials. 2012;33:1748–1758. doi: 10.1016/j.biomaterials.2011.11.049. [DOI] [PubMed] [Google Scholar]
  • 31.Hsu SH, Ho TT, Tseng TC. Nanoparticle uptake and gene transfer efficiency for MSCs on chitosan and chitosan-hyaluronan substrates. Biomaterials. 2012;33:3639–3650. doi: 10.1016/j.biomaterials.2012.02.005. [DOI] [PubMed] [Google Scholar]
  • 32.Hsueh YY, Chiang YL, Wu CC, Lin SC. Spheroid Formation and Neural Induction in Human Adipose-Derived Stem Cells on a Chitosan-Coated Surface. Cells Tissues Organs. 2012 doi: 10.1159/000332045. [DOI] [PubMed] [Google Scholar]
  • 33.Lopez J, Poitevin A, Mendoza-Martinez V, Perez-Plasencia C, Garcia-Carranca A. Cancer-initiating cells derived from established cervical cell lines exhibit stem-cell markers and increased radioresistance. BMC Cancer. 2012;12:48. doi: 10.1186/1471-2407-12-48. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Montales MT, Rahal OM, Kang J, Rogers TJ, Prior RL, Wu X, Simmen RC. Repression of mammosphere formation of human breast cancer cells by soy isoflavone genistein and blueberry polyphenolic acids suggests diet-mediated targeting of cancer stem-like/progenitor cells. Carcinogenesis. 2012;33:652–660. doi: 10.1093/carcin/bgr317. [DOI] [PubMed] [Google Scholar]
  • 35.Wang L, Mezencev R, Bowen NJ, Matyunina LV, McDonald JF. Isolation and characterization of stem-like cells from a human ovarian cancer cell line. Mol Cell Biochem. 2012;363:257–268. doi: 10.1007/s11010-011-1178-6. [DOI] [PubMed] [Google Scholar]
  • 36.Zhang Q, Nguyen AL, Shi S, Hill C, Wilder-Smith P, Krasieva TB, Le AD. Three-dimensional spheroid culture of human gingiva-derived mesenchymal stem cells enhances mitigation of chemotherapy-induced oral mucositis. Stem Cells Dev. 2012;21:937–947. doi: 10.1089/scd.2011.0252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Venkatasubramanian R, Henson MA, Forbes NS. Incorporating energy metabolism into a growth model of multicellular tumor spheroids. J Theor Biol. 2006;242:440–453. doi: 10.1016/j.jtbi.2006.03.011. [DOI] [PubMed] [Google Scholar]
  • 38.Ward JP, King JR. Mathematical modelling of drug transport in tumour multicell spheroids and monolayer cultures. Math Biosci. 2003;181:177–207. doi: 10.1016/s0025-5564(02)00148-7. [DOI] [PubMed] [Google Scholar]
  • 39.Wenning LA, Murphy RM. Coupled cellular trafficking and diffusional limitations in delivery of immunotoxins to multicell tumor spheroids. Biotechnol Bioeng. 1999;62:562–575. [PubMed] [Google Scholar]
  • 40.Owen MR, Byrne HM, Lewis CE. Mathematical modelling of the use of macrophages as vehicles for drug delivery to hypoxic tumour sites. J Theor Biol. 2004;226:377–391. doi: 10.1016/j.jtbi.2003.09.004. [DOI] [PubMed] [Google Scholar]
  • 41.Webb SD, Owen MR, Byrne HM, Murdoch C, Lewis CE. Macrophage-based anti-cancer therapy: modelling different modes of tumour targeting. Bull Math Biol. 2007;69:1747–1776. doi: 10.1007/s11538-006-9189-2. [DOI] [PubMed] [Google Scholar]
  • 42.Goodman TT, Chen J, Matveev K, Pun SH. Spatio-temporal modeling of nanoparticle delivery to multicellular tumor spheroids. Biotechnol Bioeng. 2008;101:388–399. doi: 10.1002/bit.21910. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Friedrich J, Ebner R, Kunz-Schughart LA. Experimental anti-tumor therapy in 3-D: spheroids--old hat or new challenge? Int J Radiat Biol. 2007;83:849–871. doi: 10.1080/09553000701727531. [DOI] [PubMed] [Google Scholar]
  • 44.Torisawa YS, Takagi A, Nashimoto Y, Yasukawa T, Shiku H, Matsue T. A multicellular spheroid array to realize spheroid formation, culture, and viability assay on a chip. Biomaterials. 2007;28:559–566. doi: 10.1016/j.biomaterials.2006.08.054. [DOI] [PubMed] [Google Scholar]
  • 45.Ungrin MD, Joshi C, Nica A, Bauwens C, Zandstra PW. Reproducible, ultra high-throughput formation of multicellular organization from single cell suspension-derived human embryonic stem cell aggregates. PLoS One. 2008;3:e1565. doi: 10.1371/journal.pone.0001565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Kelm JM, Timmins NE, Brown CJ, Fussenegger M, Nielsen LK. Method for generation of homogeneous multicellular tumor spheroids applicable to a wide variety of cell types. Biotechnol Bioeng. 2003;83:173–180. doi: 10.1002/bit.10655. [DOI] [PubMed] [Google Scholar]
  • 47.Kelm JM, Djonov V, Ittner LM, Fluri D, Born W, Hoerstrup SP, Fussenegger M. Design of custom-shaped vascularized tissues using microtissue spheroids as minimal building units. Tissue Eng. 2006;12:2151–2160. doi: 10.1089/ten.2006.12.2151. [DOI] [PubMed] [Google Scholar]
  • 48.Kelm JM, Fussenegger M. Scaffold-free cell delivery for use in regenerative medicine. Adv Drug Deliv Rev. 2010;62:753–764. doi: 10.1016/j.addr.2010.02.003. [DOI] [PubMed] [Google Scholar]
  • 49.Ong SM, Zhang C, Toh YC, Kim SH, Foo HL, Tan CH, van Noort D, Park S, Yu H. A gel-free 3D microfluidic cell culture system. Biomaterials. 2008;29:3237–3244. doi: 10.1016/j.biomaterials.2008.04.022. [DOI] [PubMed] [Google Scholar]
  • 50.Torisawa YS, Chueh BH, Huh D, Ramamurthy P, Roth TM, Barald KF, Takayama S. Efficient formation of uniform-sized embryoid bodies using a compartmentalized microchannel device. Lab Chip. 2007;7:770–776. doi: 10.1039/b618439a. [DOI] [PubMed] [Google Scholar]
  • 51.Wu LY, Di Carlo D, Lee LP. Microfluidic self-assembly of tumor spheroids for anticancer drug discovery. Biomed Microdevices. 2008;10:197–202. doi: 10.1007/s10544-007-9125-8. [DOI] [PubMed] [Google Scholar]
  • 52.Fukuda J, Khademhosseini A, Yeo Y, Yang X, Yeh J, Eng G, Blumling J, Wang CF, Kohane DS, Langer R. Micromolding of photocrosslinkable chitosan hydrogel for spheroid microarray and co-cultures. Biomaterials. 2006;27:5259–5267. doi: 10.1016/j.biomaterials.2006.05.044. [DOI] [PubMed] [Google Scholar]
  • 53.Sodunke TR, Turner KK, Caldwell SA, McBride KW, Reginato MJ, Noh HM. Micropatterns of Matrigel for three-dimensional epithelial cultures. Biomaterials. 2007;28:4006–4016. doi: 10.1016/j.biomaterials.2007.05.021. [DOI] [PubMed] [Google Scholar]
  • 54.Gotoh Y, Ishizuka Y, Matsuura T, Niimi S. Spheroid formation and expression of liver-specific functions of human hepatocellular carcinoma-derived FLC-4 cells cultured in lactose-silk fibroin conjugate sponges. Biomacromolecules. 2011;12:1532–1539. doi: 10.1021/bm101495c. [DOI] [PubMed] [Google Scholar]
  • 55.Torisawa YS, Mosadegh B, Luker GD, Morell M, O’Shea KS, Takayama S. Microfluidic hydrodynamic cellular patterning for systematic formation of co-culture spheroids. Integr Biol (Camb) 2009;1:649–654. doi: 10.1039/b915965g. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Lee KH, No da Y, Kim SH, Ryoo JH, Wong SF, Lee SH. Diffusion-mediated in situ alginate encapsulation of cell spheroids using microscale concave well and nanoporous membrane. Lab Chip. 2011;11:1168–1173. doi: 10.1039/c0lc00540a. [DOI] [PubMed] [Google Scholar]
  • 57.Yoshii Y, Waki A, Yoshida K, Kakezuka A, Kobayashi M, Namiki H, Kuroda Y, Kiyono Y, Yoshii H, Furukawa T, Asai T, Okazawa H, Gelovani JG, Fujibayashi Y. The use of nanoimprinted scaffolds as 3D culture models to facilitate spontaneous tumor cell migration and well-regulated spheroid formation. Biomaterials. 2011;32:6052–6058. doi: 10.1016/j.biomaterials.2011.04.076. [DOI] [PubMed] [Google Scholar]
  • 58.Cordey M, Limacher M, Kobel S, Taylor V, Lutolf MP. Enhancing the reliability and throughput of neurosphere culture on hydrogel microwell arrays. Stem Cells. 2008;26:2586–2594. doi: 10.1634/stemcells.2008-0498. [DOI] [PubMed] [Google Scholar]
  • 59.Tung YC, Hsiao AY, Allen SG, Torisawa YS, Ho M, Takayama S. High-throughput 3D spheroid culture and drug testing using a 384 hanging drop array. Analyst. 2011;136:473–478. doi: 10.1039/c0an00609b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Hsiao AY, Tung YC, Qu X, Patel LR, Pienta KJ, Takayama S. 384 hanging drop arrays give excellent Z-factors and allow versatile formation of co-culture spheroids. Biotechnol Bioeng. 2011 doi: 10.1002/bit.24399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Hsiao AY, Tung YC, Kuo CH, Mosadegh B, Bedenis R, Pienta KJ, Takayama S. Micro-ring structures stabilize microdroplets to enable long term spheroid culture in 384 hanging drop array plates. Biomed Microdevices. 2011 doi: 10.1007/s10544-011-9608-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Cioce M, Gherardi S, Viglietto G, Strano S, Blandino G, Muti P, Ciliberto G. Mammosphere-forming cells from breast cancer cell lines as a tool for the identification of CSC-like- and early progenitor-targeting drugs. Cell Cycle. 2010;9:2878–2887. [PubMed] [Google Scholar]
  • 63.Ginestier C, Wicinski J, Cervera N, Monville F, Finetti P, Bertucci F, Wicha MS, Birnbaum D, Charafe-Jauffret E. Retinoid signaling regulates breast cancer stem cell differentiation. Cell Cycle. 2009;8:3297–3302. doi: 10.4161/cc.8.20.9761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Chen KL, Pan F, Jiang H, Chen JF, Pei L, Xie FW, Liang HJ. Highly enriched CD133(+)CD44(+) stem-like cells with CD133(+)CD44(high) metastatic subset in HCT116 colon cancer cells. Clin Exp Metastasis. 2011;28:751–763. doi: 10.1007/s10585-011-9407-7. [DOI] [PubMed] [Google Scholar]
  • 65.Oh PS, Patel VB, Sanders MA, Kanwar SS, Yu Y, Nautiyal J, Patel BB, Majumdar AP. Schlafen-3 decreases cancer stem cell marker expression and autocrine/juxtacrine signaling in FOLFOX-resistant colon cancer cells. Am J Physiol Gastrointest Liver Physiol. 2011;301:G347–355. doi: 10.1152/ajpgi.00403.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Wan F, Zhang S, Xie R, Gao B, Campos B, Herold-Mende C, Lei T. The utility and limitations of neurosphere assay, CD133 immunophenotyping and side population assay in glioma stem cell research. Brain Pathol. 2010;20:877–889. doi: 10.1111/j.1750-3639.2010.00379.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Deleyrolle LP, Ericksson G, Morrison BJ, Lopez JA, Burrage K, Burrage P, Vescovi A, Rietze RL, Reynolds BA. Determination of somatic and cancer stem cell self-renewing symmetric division rate using sphere assays. PLoS One. 2011;6:e15844. doi: 10.1371/journal.pone.0015844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Anaka M, Freyer C, Gedye C, Caballero O, Davis ID, Behren A, Cebon J. Stem cell media culture of melanoma results in the induction of a nonrepresentative neural expression profile. Stem Cells. 2012;30:336–343. doi: 10.1002/stem.786. [DOI] [PubMed] [Google Scholar]
  • 69.Meng Q, Wu D, Zhang G, Qiu H. Direct self-assembly of hepatocytes spheroids within hollow fibers in presence of collagen. Biotechnol Lett. 2006;28:279–284. doi: 10.1007/s10529-005-5531-2. [DOI] [PubMed] [Google Scholar]
  • 70.Haeger JD, Hambruch N, Dilly M, Froehlich R, Pfarrer C. Formation of bovine placental trophoblast spheroids. Cells Tissues Organs. 2011;193:274–284. doi: 10.1159/000320544. [DOI] [PubMed] [Google Scholar]
  • 71.Ivascu A, Kubbies M. Rapid generation of single-tumor spheroids for high-throughput cell function and toxicity analysis. J Biomol Screen. 2006;11:922–932. doi: 10.1177/1087057106292763. [DOI] [PubMed] [Google Scholar]
  • 72.Korff T, Krauss T, Augustin HG. Three-dimensional spheroidal culture of cytotrophoblast cells mimics the phenotype and differentiation of cytotrophoblasts from normal and preeclamptic pregnancies. Exp Cell Res. 2004;297:415–423. doi: 10.1016/j.yexcr.2004.03.043. [DOI] [PubMed] [Google Scholar]
  • 73.Korff T, Augustin HG. Integration of endothelial cells in multicellular spheroids prevents apoptosis and induces differentiation. J Cell Biol. 1998;143:1341–1352. doi: 10.1083/jcb.143.5.1341. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Xin L, Lukacs RU, Lawson DA, Cheng D, Witte ON. Self-renewal and multilineage differentiation in vitro from murine prostate stem cells. Stem Cells. 2007;25:2760–2769. doi: 10.1634/stemcells.2007-0355. [DOI] [PubMed] [Google Scholar]
  • 75.Griffith LG, Swartz MA. Capturing complex 3D tissue physiology in vitro. Nat Rev Mol Cell Biol. 2006;7:211–224. doi: 10.1038/nrm1858. [DOI] [PubMed] [Google Scholar]
  • 76.Steeg PS. Tumor metastasis: mechanistic insights and clinical challenges. Nat Med. 2006;12:895–904. doi: 10.1038/nm1469. [DOI] [PubMed] [Google Scholar]
  • 77.Gilbert LA, Hemann MT. DNA damage-mediated induction of a chemoresistant niche. Cell. 2010;143:355–366. doi: 10.1016/j.cell.2010.09.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Upreti M, Jamshidi-Parsian A, Koonce NA, Webber JS, Sharma SK, Asea AA, Mader MJ, Griffin RJ. Tumor-Endothelial Cell Three-dimensional Spheroids: New Aspects to Enhance Radiation and Drug Therapeutics. Transl Oncol. 2011;4:365–376. doi: 10.1593/tlo.11187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Otsuka H, Hirano A, Nagasaki Y, Okano T, Horiike Y, Kataoka K. Two-dimensional multiarray formation of hepatocyte spheroids on a microfabricated PEG-brush surface. Chembiochem. 2004;5:850–855. doi: 10.1002/cbic.200300822. [DOI] [PubMed] [Google Scholar]
  • 80.Leclerc E, Sakai Y, Fujii T. Microfluidic PDMS (polydimethylsiloxane) bioreactor for large-scale culture of hepatocytes. Biotechnol Prog. 2004;20:750–755. doi: 10.1021/bp0300568. [DOI] [PubMed] [Google Scholar]
  • 81.Chen MC, Gupta M, Cheung KC. Alginate-based microfluidic system for tumor spheroid formation and anticancer agent screening. Biomed Microdevices. 2010;12:647–654. doi: 10.1007/s10544-010-9417-2. [DOI] [PubMed] [Google Scholar]
  • 82.Friedl P, Alexander S. Cancer invasion and the microenvironment: plasticity and reciprocity. Cell. 2011;147:992–1009. doi: 10.1016/j.cell.2011.11.016. [DOI] [PubMed] [Google Scholar]
  • 83.Kaur P, Ward B, Saha B, Young L, Groshen S, Techy G, Lu Y, Atkinson R, Taylor CR, Ingram M, Imam SA. Human breast cancer histoid: an in vitro 3-dimensional co-culture model that mimics breast cancer tissue. J Histochem Cytochem. 2011;59:1087–1100. doi: 10.1369/0022155411423680. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Ingram M, Techy GB, Ward BR, Imam SA, Atkinson R, Ho H, Taylor CR. Tissue engineered tumor models. Biotech Histochem. 2010;85:213–229. doi: 10.3109/10520295.2010.483655. [DOI] [PubMed] [Google Scholar]
  • 85.Toh YC, Zhang C, Zhang J, Khong YM, Chang S, Samper VD, van Noort D, Hutmacher DW, Yu H. A novel 3D mammalian cell perfusion-culture system in microfluidic channels. Lab Chip. 2007;7:302–309. doi: 10.1039/b614872g. [DOI] [PubMed] [Google Scholar]
  • 86.Karp JM, Yeh J, Eng G, Fukuda J, Blumling J, Suh KY, Cheng J, Mahdavi A, Borenstein J, Langer R, Khademhosseini A. Controlling size, shape and homogeneity of embryoid bodies using poly(ethylene glycol) microwells. Lab Chip. 2007;7:786–794. doi: 10.1039/b705085m. [DOI] [PubMed] [Google Scholar]
  • 87.Chua KN, Lim WS, Zhang P, Lu H, Wen J, Ramakrishna S, Leong KW, Mao HQ. Stable immobilization of rat hepatocyte spheroids on galactosylated nanofiber scaffold. Biomaterials. 2005;26:2537–2547. doi: 10.1016/j.biomaterials.2004.07.040. [DOI] [PubMed] [Google Scholar]
  • 88.Elkayam T, Amitay-Shaprut S, Dvir-Ginzberg M, Harel T, Cohen S. Enhancing the drug metabolism activities of C3A--a human hepatocyte cell line--by tissue engineering within alginate scaffolds. Tissue Eng. 2006;12:1357–1368. doi: 10.1089/ten.2006.12.1357. [DOI] [PubMed] [Google Scholar]
  • 89.Glicklis R, Shapiro L, Agbaria R, Merchuk JC, Cohen S. Hepatocyte behavior within three-dimensional porous alginate scaffolds. Biotechnol Bioeng. 2000;67:344–353. doi: 10.1002/(sici)1097-0290(20000205)67:3<344::aid-bit11>3.0.co;2-2. [DOI] [PubMed] [Google Scholar]
  • 90.Kotov NA, Liu Y, Wang S, Cumming C, Eghtedari M, Vargas G, Motamedi M, Nichols J, Cortiella J. Inverted colloidal crystals as three-dimensional cell scaffolds. Langmuir. 2004;20:7887–7892. doi: 10.1021/la049958o. [DOI] [PubMed] [Google Scholar]
  • 91.Lee J, Shanbhag S, Kotov NA. Inverted colloidal crystals as three-dimensional microenvironments for cellular co-cultures. J Mater Chem. 2006;16:3558–3564. [Google Scholar]
  • 92.Torok E, Vogel C, Lutgehetmann M, Ma PX, Dandri M, Petersen J, Burda MR, Siebert K, Dullmann J, Rogiers X, Pollok JM. Morphological and functional analysis of rat hepatocyte spheroids generated on poly(L-lactic acid) polymer in a pulsatile flow bioreactor. Tissue Eng. 2006;12:1881–1890. doi: 10.1089/ten.2006.12.1881. [DOI] [PubMed] [Google Scholar]
  • 93.Drewitz M, Helbling M, Fried N, Bieri M, Moritz W, Lichtenberg J, Kelm JM. Towards automated production and drug sensitivity testing using scaffold-free spherical tumor microtissues. Biotechnol J. 2011;6:1488–1496. doi: 10.1002/biot.201100290. [DOI] [PubMed] [Google Scholar]
  • 94.Bryce NS, Zhang JZ, Whan RM, Yamamoto N, Hambley TW. Accumulation of an anthraquinone and its platinum complexes in cancer cell spheroids: the effect of charge on drug distribution in solid tumour models. Chem Commun (Camb) 2009:2673–2675. doi: 10.1039/b902415h. [DOI] [PubMed] [Google Scholar]
  • 95.Kostarelos K, Emfietzoglou D, Papakostas A, Yang WH, Ballangrud AM, Sgouros G. Engineering lipid vesicles of enhanced intratumoral transport capabilities: correlating liposome characteristics with penetration into human prostate tumor spheroids. J Liposome Res. 2005;15:15–27. doi: 10.1081/lpr-64953. [DOI] [PubMed] [Google Scholar]
  • 96.Oishi M, Nagasaki Y, Nishiyama N, Itaka K, Takagi M, Shimamoto A, Furuichi Y, Kataoka K. Enhanced growth inhibition of hepatic multicellular tumor spheroids by lactosylated poly(ethylene glycol)-siRNA conjugate formulated in PEGylated polyplexes. ChemMedChem. 2007;2:1290–1297. doi: 10.1002/cmdc.200700076. [DOI] [PubMed] [Google Scholar]
  • 97.Ying X, Wen H, Lu WL, Du J, Guo J, Tian W, Men Y, Zhang Y, Li RJ, Yang TY, Shang DW, Lou JN, Zhang LR, Zhang Q. Dual-targeting daunorubicin liposomes improve the therapeutic efficacy of brain glioma in animals. J Control Release. 2010;141:183–192. doi: 10.1016/j.jconrel.2009.09.020. [DOI] [PubMed] [Google Scholar]
  • 98.Durand RE, Olive PL. Evaluation of bioreductive drugs in multicell spheroids. Int J Radiat Oncol Biol Phys. 1992;22:689–692. doi: 10.1016/0360-3016(92)90504-b. [DOI] [PubMed] [Google Scholar]
  • 99.He M, Herr AE. Automated microfluidic protein immunoblotting. Nat Protoc. 2010;5:1844–1856. doi: 10.1038/nprot.2010.142. [DOI] [PubMed] [Google Scholar]
  • 100.Paguirigan AL, Puccinelli JP, Su X, Beebe DJ. Expanding the available assays: adapting and validating In-Cell Westerns in microfluidic devices for cell-based assays. Assay Drug Dev Technol. 2010;8:591–601. doi: 10.1089/adt.2010.0274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Pan W, Chen W, Jiang X. Microfluidic Western blot. Anal Chem. 2010;82:3974–3976. doi: 10.1021/ac1000493. [DOI] [PubMed] [Google Scholar]
  • 102.Wlodkowic D, Darzynkiewicz Z. Rise of the micromachines: microfluidics and the future of cytometry. Methods Cell Biol. 2011;102:105–125. doi: 10.1016/B978-0-12-374912-3.00005-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Birmingham A, Selfors LM, Forster T, Wrobel D, Kennedy CJ, Shanks E, Santoyo-Lopez J, Dunican DJ, Long A, Kelleher D, Smith Q, Beijersbergen RL, Ghazal P, Shamu CE. Statistical methods for analysis of high-throughput RNA interference screens. Nat Methods. 2009;6:569–575. doi: 10.1038/nmeth.1351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Sui Y, Wu Z. Alternative statistical parameter for high-throughput screening assay quality assessment. J Biomol Screen. 2007;12:229–234. doi: 10.1177/1087057106296498. [DOI] [PubMed] [Google Scholar]
  • 105.Zhang JH, Chung TD, Oldenburg KR. A Simple Statistical Parameter for Use in Evaluation and Validation of High Throughput Screening Assays. J Biomol Screen. 1999;4:67–73. doi: 10.1177/108705719900400206. [DOI] [PubMed] [Google Scholar]
  • 106.Scott D, Dikici E, Ensor M, Daunert S. Bioluminescence and its impact on bioanalysis. Ann Rev Anal Chem. 2011;4:297–319. doi: 10.1146/annurev-anchem-061010-113855. [DOI] [PubMed] [Google Scholar]
  • 107.McMillin D, Delmore J, Weisberg E, Negri J, Geer D, Klippel S, Mitsiades N, Schlossman R, Munshi N, Kung A, Griffin J, Richardson P, Anderson K, Mitsiades C. Tumor cell-specific bioluminescence platform to identify stroma-induced changes to anticancer drug activity. Nat Med. 2010;16:483–489. doi: 10.1038/nm.2112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Michelini E, Cevenini L, Mezzanotte L, Coppa A, Roda A. Cell-based assays: fuelling drug discovery. Anal Bioanal Chem. 2010;398:227–238. doi: 10.1007/s00216-010-3933-z. [DOI] [PubMed] [Google Scholar]
  • 109.Sweeney T, Mailander V, Tucker A, Olomu A, Zhang W, Cao Y, Negrin R, Contag C. Visualizing the kinetics of tumor-cell clearance in living animals. Proc Natl Acad Sci USA. 1999;96:12044–12049. doi: 10.1073/pnas.96.21.12044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Luker G, Luker K. Luciferase protein complementation assays for bioluminescence imaging of cells and mice. Methods Mol Biol. 2011;680:29–43. doi: 10.1007/978-1-60761-901-7_2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Zhang L, Lee K, Bhojani M, Khan A, Shilman A, Holland E, Ross B, Rehemtulla A. Molecular imaging of Akt kinase activity. Nat Med. 2007;13:1114–1119. doi: 10.1038/nm1608. [DOI] [PubMed] [Google Scholar]
  • 112.Luker K, Mihalko L, Schmidt B, Lewin S, Ray P, Shcherbo D, Chudakov D, Luker G. In vivo imaging of ligand receptor binding with Gaussia luciferase complementation. Nat Med. 2012;18:172–177. doi: 10.1038/nm.2590. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Molema G. Drug Targeting: Basic Concepts and Novel Advances. In: Molema G, Meijer DKF, editors. Drug Targeting: Organ Specific Strategies. Wiley VCH; Weinhiem: 2001. pp. 1–21. [Google Scholar]
  • 114.Saltzman WM. Drug Delivery: Engineering Principles for Drug Therapy. Oxford University Press; New York: 2001. [Google Scholar]
  • 115.Parthasarathy R, Sacks PG, Harris D, Brock H, Mehta K. Interaction of liposome-associated all-trans-retinoic acid with squamous carcinoma cells. Cancer Chemother Pharmacol. 1994;34:527–534. doi: 10.1007/BF00685666. [DOI] [PubMed] [Google Scholar]
  • 116.Kostarelos K, Emfietzoglou D, Papakostas A, Yang WH, Ballangrud A, Sgouros G. Binding and interstitial penetration of liposomes within avascular tumor spheroids. Int J Cancer. 2004;112:713–721. doi: 10.1002/ijc.20457. [DOI] [PubMed] [Google Scholar]
  • 117.Emfietzoglou D, Kostarelos K, Papakostas A, Yang WH, Ballangrud A, Song H, Sgouros G. Liposome-mediated radiotherapeutics within avascular tumor spheroids: comparative dosimetry study for various radionuclides, liposome systems, and a targeting antibody. J Nucl Med. 2005;46:89–97. [PubMed] [Google Scholar]
  • 118.Mellor HR, Davies LA, Caspar H, Pringle CR, Hyde SC, Gill DR, Callaghan R. Optimising non-viral gene delivery in a tumour spheroid model. J Gene Med. 2006;8:1160–1170. doi: 10.1002/jgm.947. [DOI] [PubMed] [Google Scholar]
  • 119.Goodman TT, Olive PL, Pun SH. Increased nanoparticle penetration in collagenase-treated multicellular spheroids. Int J Nanomedicine. 2007;2:265–274. [PMC free article] [PubMed] [Google Scholar]
  • 120.Dhanikula RS, Argaw A, Bouchard JF, Hildgen P. Methotrexate loaded polyether-copolyester dendrimers for the treatment of gliomas: enhanced efficacy and intratumoral transport capability. Mol Pharm. 2008;5:105–116. doi: 10.1021/mp700086j. [DOI] [PubMed] [Google Scholar]
  • 121.Mairs RJ, Russell J, Cunningham S, O’Donoghue JA, Gaze MN, Owens J, Vaidyanathan G, Zalutsky MR. Enhanced tumour uptake and in vitro radiotoxicity of no-carrier-added [131I]meta-iodobenzylguanidine: implications for the targeted radiotherapy of neuroblastoma. Eur J Cancer. 1995;31A:576–581. doi: 10.1016/0959-8049(95)00052-k. [DOI] [PubMed] [Google Scholar]
  • 122.Gaze MN, Mairs RJ, Boyack SM, Wheldon TE, Barrett A. 131I-meta-iodobenzylguanidine therapy in neuroblastoma spheroids of different sizes. Br J Cancer. 1992;66:1048–1052. doi: 10.1038/bjc.1992.408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Lenartz D, Andermahr J, Plum G, Menzel J, Beuth J. Efficiency of treatment with galactoside-specific lectin from mistletoe against rat glioma. Anticancer Res. 1998;18:1011–1014. [PubMed] [Google Scholar]
  • 124.Durand RE. Slow penetration of anthracyclines into spheroids and tumors: a therapeutic advantage? Cancer Chemother Pharmacol. 1990;26:198–204. doi: 10.1007/BF02897199. [DOI] [PubMed] [Google Scholar]
  • 125.Bell HS, Wharton SB, Leaver HA, Whittle IR. Effects of N-6 essential fatty acids on glioma invasion and growth: experimental studies with glioma spheroids in collagen gels. J Neurosurg. 1999;91:989–996. doi: 10.3171/jns.1999.91.6.0989. [DOI] [PubMed] [Google Scholar]
  • 126.Desoize B, Gimonet D, Jardiller JC. Cell culture as spheroids: an approach to multicellular resistance. Anticancer Res. 1998;18:4147–4158. [PubMed] [Google Scholar]
  • 127.Olive PL, Durand RE. Drug and radiation resistance in spheroids: cell contact and kinetics. Cancer Metastasis Rev. 1994;13:121–138. doi: 10.1007/BF00689632. [DOI] [PubMed] [Google Scholar]
  • 128.Knuchel R, Hofstadter F, Jenkins WE, Masters JR. Sensitivities of monolayers and spheroids of the human bladder cancer cell line MGH-U1 to the drugs used for intravesical chemotherapy. Cancer Res. 1989;49:1397–1401. [PubMed] [Google Scholar]
  • 129.Kwok TT, Twentyman PR. The response to cytotoxic drugs of EMT6 cells treated either as intact or disaggregated spheroids. Br J Cancer. 1985;51:211–218. doi: 10.1038/bjc.1985.31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Oloumi A, MacPhail SH, Johnston PJ, Banath JP, Olive PL. Changes in subcellular distribution of topoisomerase IIalpha correlate with etoposide resistance in multicell spheroids and xenograft tumors. Cancer Res. 2000;60:5747–5753. [PubMed] [Google Scholar]
  • 131.Wibe E. Resistance to vincristine of human cells grown as multicellular spheroids. Br J Cancer. 1980;42:937–941. doi: 10.1038/bjc.1980.344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Desoize B, Jardillier J. Multicellular resistance: a paradigm for clinical resistance? Crit Rev Oncol Hematol. 2000;36:193–207. doi: 10.1016/s1040-8428(00)00086-x. [DOI] [PubMed] [Google Scholar]
  • 133.Durand RE, Olive PL. Resistance of tumor cells to chemo- and radiotherapy modulated by the three-dimensional architecture of solid tumors and spheroids. Methods Cell Biol. 2001;64:211–233. doi: 10.1016/s0091-679x(01)64015-9. [DOI] [PubMed] [Google Scholar]
  • 134.Friedrich J, Seidel C, Ebner R, Kunz-Schughart LA. Spheroid-based drug screen: considerations and practical approach. Nat Protoc. 2009;4:309–324. doi: 10.1038/nprot.2008.226. [DOI] [PubMed] [Google Scholar]
  • 135.Kerr DJ, Wheldon TE, Hydns S, Kaye SB. Cytotoxic drug penetration studies in multicellular tumour spheroids. Xenobiotica. 1988;18:641–648. doi: 10.3109/00498258809041702. [DOI] [PubMed] [Google Scholar]
  • 136.Kunz-Schughart LA. Multicellular tumor spheroids: intermediates between monolayer culture and in vivo tumor. Cell Biol Int. 1999;23:157–161. doi: 10.1006/cbir.1999.0384. [DOI] [PubMed] [Google Scholar]
  • 137.Minchinton AI, Tannock IF. Drug penetration in solid tumours. Nat Rev Cancer. 2006;6:583–592. doi: 10.1038/nrc1893. [DOI] [PubMed] [Google Scholar]
  • 138.Kwok TT, Twentyman PR. The relationship between tumour geometry and the response of tumour cells to cytotoxic drugs--an in vitro study using EMT6 multicellular spheroids. Int J Cancer. 1985;35:675–682. doi: 10.1002/ijc.2910350517. [DOI] [PubMed] [Google Scholar]
  • 139.Erlanson M, Daniel-Szolgay E, Carlsson J. Relations between the penetration, binding and average concentration of cytostatic drugs in human tumour spheroids. Cancer Chemother Pharmacol. 1992;29:343–353. doi: 10.1007/BF00686002. [DOI] [PubMed] [Google Scholar]
  • 140.Durand RE. Chemosensitivity testing in V79 spheroids: drug delivery and cellular microenvironment. J Natl Cancer Inst. 1986;77:247–252. [PubMed] [Google Scholar]
  • 141.Erlichman C, Vidgen D. Cytotoxicity of adriamycin in MGH-U1 cells grown as monolayer cultures, spheroids, and xenografts in immune-deprived mice. Cancer Res. 1984;44:5369–5375. [PubMed] [Google Scholar]
  • 142.Kozin SV, Gerweck LE. Cytotoxicity of weak electrolytes after the adaptation of cells to low pH: role of the transmembrane pH gradient. Br J Cancer. 1998;77:1580–1585. doi: 10.1038/bjc.1998.260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Tannock IF, Rotin D. Acid pH in tumors and its potential for therapeutic exploitation. Cancer Res. 1989;49:4373–4384. [PubMed] [Google Scholar]
  • 144.Abbott A. Cell culture: biology’s new dimension. Nature. 2003;424:870–872. doi: 10.1038/424870a. [DOI] [PubMed] [Google Scholar]
  • 145.Mueller-Klieser W. Tumor biology and experimental therapeutics. Crit Rev Oncol Hematol. 2000;36:123–139. doi: 10.1016/s1040-8428(00)00082-2. [DOI] [PubMed] [Google Scholar]
  • 146.Dubessy C, Merlin JM, Marchal C, Guillemin F. Spheroids in radiobiology and photodynamic therapy. Crit Rev Oncol Hematol. 2000;36:179–192. doi: 10.1016/s1040-8428(00)00085-8. [DOI] [PubMed] [Google Scholar]
  • 147.Hirschberg H, Sun CH, Tromberg BJ, Yeh AT, Madsen SJ. Enhanced cytotoxic effects of 5-aminolevulinic acid-mediated photodynamic therapy by concurrent hyperthermia in glioma spheroids. J Neurooncol. 2004;70:289–299. doi: 10.1007/s11060-004-9161-7. [DOI] [PubMed] [Google Scholar]
  • 148.Kunz-Schughart LA, Kreutz M, Knuechel R. Multicellular spheroids: a three-dimensional in vitro culture system to study tumour biology. Int J Exp Pathol. 1998;79:1–23. doi: 10.1046/j.1365-2613.1998.00051.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Mueller-Klieser W. Multicellular spheroids. A review on cellular aggregates in cancer research. J Cancer Res Clin Oncol. 1987;113:101–122. doi: 10.1007/BF00391431. [DOI] [PubMed] [Google Scholar]
  • 150.Santini MT, Rainaldi G. Three-dimensional spheroid model in tumor biology. Pathobiology. 1999;67:148–157. doi: 10.1159/000028065. [DOI] [PubMed] [Google Scholar]
  • 151.Santini MT, Rainaldi G, Indovina PL. Multicellular tumour spheroids in radiation biology. Int J Radiat Biol. 1999;75:787–799. doi: 10.1080/095530099139845. [DOI] [PubMed] [Google Scholar]
  • 152.Sutherland RM, Durand RE. Growth and cellular characteristics of multicell spheroids. Recent Results Cancer Res. 1984;95:24–49. doi: 10.1007/978-3-642-82340-4_2. [DOI] [PubMed] [Google Scholar]
  • 153.Carlsson J, Nederman T. Tumour spheroid technology in cancer therapy research. Eur J Cancer Clin Oncol. 1989;25:1127–1133. doi: 10.1016/0277-5379(89)90404-5. [DOI] [PubMed] [Google Scholar]
  • 154.Mueller-Klieser W. Three-dimensional cell cultures: from molecular mechanisms to clinical applications. Am J Physiol. 1997;273:C1109–1123. doi: 10.1152/ajpcell.1997.273.4.C1109. [DOI] [PubMed] [Google Scholar]
  • 155.Bates RC, Edwards NS, Yates JD. Spheroids and cell survival. Crit Rev Oncol Hematol. 2000;36:61–74. doi: 10.1016/s1040-8428(00)00077-9. [DOI] [PubMed] [Google Scholar]
  • 156.de Ridder L, Cornelissen M, de Ridder D. Autologous spheroid culture: a screening tool for human brain tumour invasion. Crit Rev Oncol Hematol. 2000;36:107–122. doi: 10.1016/s1040-8428(00)00081-0. [DOI] [PubMed] [Google Scholar]
  • 157.Shield K, Ackland ML, Ahmed N, Rice GE. Multicellular spheroids in ovarian cancer metastases: Biology and pathology. Gynecol Oncol. 2009;113:143–148. doi: 10.1016/j.ygyno.2008.11.032. [DOI] [PubMed] [Google Scholar]
  • 158.Meng Q. Three-dimensional culture of hepatocytes for prediction of drug-induced hepatotoxicity. Expert Opin Drug Metab Toxicol. 2010;6:733–746. doi: 10.1517/17425251003674356. [DOI] [PubMed] [Google Scholar]

RESOURCES